Next Article in Journal
Galectin-9 in Gastroenterological Cancer
Next Article in Special Issue
Genome-Wide Analysis of the Expansin Gene Family in Populus and Characterization of Expression Changes in Response to Phytohormone (Abscisic Acid) and Abiotic (Low-Temperature) Stresses
Previous Article in Journal
Influence of MMR, MGMT Promotor Methylation and Protein Expression on Overall and Progression-Free Survival in Primary Glioblastoma Patients Treated with Temozolomide
Previous Article in Special Issue
Genome-Wide Identification of the CER Gene Family and Significant Features in Climate Adaptation of Castanea mollissima
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comprehensive Analysis of the INDETERMINATE DOMAIN (IDD) Gene Family and Their Response to Abiotic Stress in Zea mays

1
The Characteristic Laboratory of Crop Germplasm Innovation and Application, Provincial Department of Education, College of Agronomy, Qingdao Agricultural University, Qingdao 266109, China
2
Department of Agronomy, College of Agriculture and Biotechnology, Zijingang Campus, Zhejiang University, Hangzhou 310058, China
3
State Agricultural Biotechnology Centre, College of Science, Health, Engineering and Education, Murdoch University, Perth, WA 6150, Australia
4
The Key Laboratory of the Plant Development and Environmental Adaptation Biology, School of Life Sciences, Ministry of Education, Shandong University, Qingdao 266237, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6185; https://doi.org/10.3390/ijms24076185
Submission received: 15 February 2023 / Revised: 20 March 2023 / Accepted: 23 March 2023 / Published: 24 March 2023

Abstract

:
Transcription factors (TFs) are important regulators of numerous gene expressions due to their ability to recognize and combine cis-elements in the promoters of target genes. The INDETERMINATE DOMAIN (IDD) gene family belongs to a subfamily of C2H2 zinc finger proteins and has been identified only in terrestrial plants. Nevertheless, little study has been reported concerning the genome-wide analysis of the IDD gene family in maize. In total, 22 ZmIDD genes were identified, which can be distributed on 8 chromosomes in maize. On the basis of evolutionary relationships and conserved motif analysis, ZmIDDs were categorized into three clades (1, 2, and 3), each owning 4, 6, and 12 genes, respectively. We analyzed the characteristics of gene structure and found that 3 of the 22 ZmIDD genes do not contain an intron. Cis-element analysis of the ZmIDD promoter showed that most ZmIDD genes possessed at least one ABRE or MBS cis-element, and some ZmIDD genes owned the AuxRR-core, TCA-element, TC-rich repeats, and LTR cis-element. The Ka:Ks ratio of eight segmentally duplicated gene pairs demonstrated that the ZmIDD gene families had undergone a purifying selection. Then, the transcription levels of ZmIDDs were analyzed, and they showed great differences in diverse tissues as well as abiotic stresses. Furthermore, regulatory networks were constructed through the prediction of ZmIDD-targeted genes and miRNAs, which can inhibit the transcription of ZmIDDs. In total, 6 ZmIDDs and 22 miRNAs were discovered, which can target 180 genes and depress the expression of 9 ZmIDDs, respectively. Taken together, the results give us valuable information for studying the function of ZmIDDs involved in plant development and climate resilience in maize.

1. Introduction

To survive better, terrestrial plants have evolved many mechanisms to accommodate a variety of complex environments, including drought, salt, and low temperature stress. Transcription factors (TFs) play dominant roles in plant growth and development and response to abiotic stress [1,2]. TFs can recognize and combine cis-elements, thus activating or repressing the transcription of downstream genes [3]. A number of important TF families have been identified in plants, including C2H2 (Cys2His2 zinc finger) [4], WRKY [5], bHLH (basic helix-loop-helix) [6], MYB [7], bZIP (basic region-leucine zipper) [8], and MADS-box [9]. Of these, the INDETERMINATE DOMAIN (IDD) family, a class of C2H2 zinc finger proteins, has been found only in land plants [10,11].
The IDD gene family is characterized by the INDETERMINATE (ID) domain, which contains four zinc fingers (ZFs) [10,12]. The four ZFs include two C2H2-type ZFs involved in DNA binding and two C2HC-type ZFs responsible for protein interaction [13]. The first IDD genes, ZmID1, were identified in Zea mays through transposon insertion, which regulates flowering time via a leaf-generated signal [14]. The IDD family gene has been discovered in numerous plants, including Arabidopsis [15], rice [16], Moso bamboo [17], and so on. So far, 16 IDD genes have been found in Arabidopsis, and 12 of them have been functionally verified [18]. In Arabidopsis thaliana, IDDs function in various metabolic and development processes, including root and leaf development, flowering time, seed maturation, hormone signaling, and abiotic stress [19]. For example, six AtIDD genes (IDD2, IDD3, IDD4, IDD5, IDD9, and IDD10) are involved in gibberellin (GA) homeostasis, thus modulating florescence through interactions with DELLA proteins [20]. Coelho et al. [18] reported that AtIDD14 could be alternatively spliced and produce two transcripts. AtIDD14α exists in normal conditions, but AtIDD14β accumulates when plants are subjected to cold stress. Moreover, AtIDD14 can also interact with ABFs/AREBs and cooperatively mediate ABA-dependent drought tolerance [21]. This suggests that IDDs are responsible for responding to different environmental stimuli in plants, but research concerning IDDs remains rare.
IDDs also play vital roles in the grass family, such as rice and barley. Ghd10, the ortholog of ZmID1, participates in mediating yield component characters via increasing plant height and tillering under short-day (SD) conditions [22]. OsIDD2 is a negative regulator in second cell wall formation through repressing target genes, which are related to sucrose metabolism and lignin synthesis [23]. OsIDD3/ROC1 enhances cold tolerance by directly targeting cis-elements of dehydration-responsive element-binding protein (DREB)/CBF1 [24]. OsIDD10 was reported to regulate ammonium absorption and nitrogen metabolism in roots [25]. In barley, BROAD LEAF1 (BLF1) was found to regulate leaf size by restraining longitudinal cell proliferation [26].
Maize is the most common cereal crop in the world. Nevertheless, genome-wide analysis of IDDs has not been conducted in maize. Thus far, three IDDs have been characterized in maize. For example, ZmID1 can regulate flowering time [11], and its paralogs ZmIDD9 and ZmIDDveg9 control endosperm development [27,28]. With the rapid development of sequencing technology, whole-genome analysis of ZmIDD gene families becomes feasible. In this study, we identified 22 candidate ZmIDD genes through bioinformatics analysis. Then, a systematic analysis of the phylogenetic tree, chromosome localization, gene structures, conserved motifs, and promoter cis-elements was conducted. In addition, tissue-specific and abiotic stress-mediated expression profiles of ZmIDDs were analyzed. At last, miRNA-target ZmIDDs and ZmIDDs-target genes were predicted. The results will give us valuable information for better comprehending the function of ZmIDDs in plant development and response to environmental stress in maize.

2. Results

2.1. Identification and Evolution Analysis of ZmIDD Family Genes in Zea mays

In this study, we identified 22 ZmIDDs. The gene and protein characteristics, such as gene names, ID, chromosomal locations, amino acid numbers, molecular weights (MW), and isoelectric points (pI), were shown in Table S1. For instance, the amino acid length of 22 ZmIDD proteins ranges from 354 to 815. The molecular weight ranges from 38 to 90 kDa. The coding sequence and protein sequence can be seen in Table S2. Evolutionary relationships of IDD genes were shown via a phylogenetic tree using 15 rice IDDs, 16 Arabidopsis IDDs, and 22 maize IDDs using the maximum likelihood method, the minimum evolution method, and the Neighbor-Joining method (Figure 1; Figures S1 and S2; Table S3). Three major clades were identified. Clades 1, 2, and 3 contained 4, 6, and 12 ZmIDDs, respectively (Figure 1; Figures S1 and S2).

2.2. Structure Analysis of ZmIDD

The DNA structure analysis showed that ZmIDDs had 1–4 exons distributed unevenly (Figure 2B; Table S2). Meanwhile, eight conserved motifs were found in ZmIDDs and named motifs 1–8 (Table S4). In detail, all ZmIDDs have motif 1, 2, and 3. Motifs 4, 5, and 7 were presented in all subfamilies. Motif 8 only exists in the members of clade 1. Motif 6 exists exclusively in the members of clade 2 and clade 3 (Figure 2C). Multi-sequence alignment of ZmIDDs demonstrated that all ZmIDD proteins owned two C2H2-type (ZF1 and ZF2) and two C2HC-type (ZF3 and ZF4) ZFs (Figure 3). In addition, 3D structures of ZmIDD1, ZmIDD2, and ZmIDD4 proteins, belonging to clade 3, clade 1, and clade 2, respectively, were predicted, indicating each ZmIDD had different 3D structures in spite of having four ZF fingers in all ZmIDDs (Figure 4).

2.3. Location and Duplication of ZmIDDs

The chromosome localization showed that 22 ZmIDDs were distributed across the maize genome except for chromosome 4 and 6 (Figure 5). In detail, ZmIDD clade 1 genes were distributed on chromosomes 1, 7, and 9; ZmIDD clade 2 genes were found on chromosomes 1, 2, 5, 7, and 10; and ZmIDD clade 3 genes were localized on chromosomes 1, 2, 3, 5, 7, and 8. Then, we detected gene duplication and found 8 segmental duplication events in ZmIDDs (Figure 6A; Table S5). It indicated that segmental duplication was the primary reason for the enlargement of ZmIDD genes in maize. Furthermore, a synteny analysis between the maize and rice genomes was also performed. It found that three IDD gene pairs (ZmIDD13/OsIDD9, ZmIDD1/OsIDD9, and ZmIDD17/OsIDD11) were found in Zea mays and Oryza sativa (Figure 6B; Table S6).
To illustrate the evolutionary constraints acting, we calculated the Ks value, Ka value, Ka:Ks ratio, and divergence time of paralogous IDD genes. The majority of Ka:Ks ratios in segmental duplicated ZmIDD gene pairs were less than one, with the exception of ZmIDD15/ZmIDD3 with 1.1, and divergence time occurred between 6.788 Mya and 112.101 Mya ago (Table S5).

2.4. Cis-Elements Analysis in ZmIDDs Promoters

To figure out the function and regulatory pattern of ZmIDD genes, we scanned the promoter sequence of 22 ZmIDDs to analyze cis-elements, such as ABRE, AuxRR-core, MBS, TCA-element, TC-rich repeats, and LTR, related to ABA, auxin, drought-inducibility, salicylic acid, defense and stress, and low temperature responses (Figure 7; Table S7). Generally, 15 ZmIDDs (68%) had ABRE cis-elements, 7 ZmIDD genes (32%) owned AuxRR-core elements, 12 ZmIDD genes (46.7%) owned MBS cis-elements, and 3 (14%) ZmIDD genes carried LTR cis-elements. Six ZmIDD genes owned TCA elements, and four ZmIDD genes carried TC-rich repeats.

2.5. Tissue-Specific Expression Patterns of ZmIDDs

A number of reports have shown that IDD genes are expressed in many tissues of plants. For instance, some IDDs are mainly expressed in mature leaves [29], and some in immature leaves [10] or roots [30]. To determine the tissue-specific expression profiles of ZmIDDs, the transcription levels of 22 ZmIDDs in six tissues of B73, such as roots, leaves, stems, embryo, endosperm, and pericarp, were compared. Based on the difference in expression patterns, 22 ZmIDD genes were categorized into three groups (Figure 8A; Table S8). Group 1 consists of one gene (ZmIDD9), which is not expressed in six tissues. Group 2 contains 11 genes expressed only in certain tissues. For example, ZmIDD13 was expressed only in the root but not in other tissues. Group 3 had 10 genes expressed in all tissues. Moreover, group 3 could be divided into two subgroups. Subgroup 1 contains four ZmIDDs with a high transcription level (log2TPM + 1 > 1) in six tissues, including ZmIDD8, ZmIDD14, ZmIDD17, and ZmIDD22. The rest of the six genes belong to subgroup 2.

2.6. Stress-Induced Expression Patterns of ZmIDDs

We collected transcriptome data from ZmIDD genes in maizeGDB and compared gene expression patterns when plants were subjected to drought, salt, heat, cold, and ultraviolet light. Generally, the expression of ZmIDD genes displayed significant differences under diverse abiotic stresses (Figure 8B; Table S9). Interestingly, most ZmIDD genes showed down-regulation or no change in response to abiotic stress except for ZmIDD8 and ZmIDD21. For example, ZmIDD8 and ZmIDD21 exhibited up-regulation after salt and cold stress, respectively. Meanwhile, many ZmIDD genes were inhibited when plants were subjected to abiotic stress. For instance, the expression of ZmIDD6 and ZmIDD11 was decreased under cold and drought treatments, respectively. ZmIDD15 showed down-regulation under heat, salt, and cold stress. However, there are some ZmIDDs that were not in response to all the abiotic stresses analyzed in the current study, such as ZmIDD7, ZmIDD9, ZmIDD10, ZmIDD13, and ZmIDD19. In addition, ZmIDD8 displayed opposite expression patterns under heat and salt stress.

2.7. ZmIDDs-Regulated Genes and miRNA-Targeted ZmIDDs

There were 6 ZmIDDs (ZmIDD 1, 4, 5, 9, 10, and 13) that could bind the cis-elements and mediate the transcription of 180 downstream genes in maize. The detailed information is listed in Table S10. The GO enrichment analysis concerning downstream genes demonstrated that five ZmIDD-targeted genes were involved in protein dimerization activity (GO:0046983) and four ZmIDD-targeted genes were directed hydrolase activities, hydrolyzing O-glycosyl compounds (GO:0004553) (Figure 9; Tables S11 and S12).
To determine whether a miRNA-mRNA regulatory network exists in maize, 321 known miRNAs were scanned. Finally, 22 miRNAs were predicted to inhibit the transcription level of ZmIDDs (Figure 10; Table S13). In short, miRNA-targeted ZmIDDs were divided into 5 networks: group 1/4/5/6 involving ZmIDD 4/7/18/5, respectively; group 2 involving two ZmIDDs (ZmIDD8 and ZmIDD17); and group 3 involving three ZmIDDs (ZmIDD 2/3/21) (Figure 10; Table S13).

3. Discussion

Transcription factors have regulatory and functional roles in plant growth and development [31]. The indeterminate domain (IDD) genes exist universally in all plants. In Zea mays, only three IDDs have been functionally verified: INDETERMINATE1 (ID1), the dominant regulator in flowering [14,32], and Naked Endosperm 1 and 2 (NKD1 and NKD2) associated with seed development [26,28]. Hence, whole-genome analysis and function prediction of the ZmIDD genes were necessary in maize. In this study, we identified 22 ZmIDD genes, which were named ZmIDD1 to ZmIDD22 according to their chromosome localization in maize (Figure 5; Tables S1 and S2). Three IDDs (ID1, NKD1, and NKD2) correspond to ZmIDD5, ZmIDD7, and ZmIDD22 identified in this work, respectively. Gene duplication and diversification play an essential role in plant evolution [33]. Prochetto et al. [19] reported that IDD derived from Chalorophyta and experienced a gene duplication event about 470 million years ago (MYA). Several times, duplication led to the emergence and diversification of terrestrial plants. Hence, monocots and dicots owned the same IDD gene numbers (16–23) [10,34,35]. In the current study, eight segmental duplication events were found in ZmIDDs (Figure 6; Table S5). Meanwhile, divergence time occurred between 6.788 Mya and 112.101 Mya ago, and most Ka/Ks values were less than one (Table S5), implying that the ZmIDD gene family may have suffered robust purifying selective pressure in the course of evolution.
Plant hormones, as important signaling molecules, can control plant growth, development, and stress responses [36]. Previous research has reported that many IDDs are related to hormone homeostasis [12,37,38]. For example, IDD14, IDD15, and IDD16 can cooperatively function in lateral organ morphogenesis and gravitropism through accelerating auxin synthesis and transport in Arabidopsis [39]. As shown in Figure 1, Figures S1 and S2, AtIDD14 was a paralog of AtIDD15 and AtIDD16, and ZmIDD2, ZmIDD3, ZmIDD6, ZmIDD15, and ZmIDD21 were orthologous genes of IDD14, IDD15, and IDD16 in Arabidopsis. Meanwhile, the promoter of 7 ZmIDDs has an AuxRR-element involved in auxin biosynthesis, including ZmIDD1, ZmIDD2, ZmIDD6, ZmIDD7, ZmIDD9, ZmIDD13, and ZmIDD15 (Figure 7; Table S7). It was suggested that although ZmIDD3 and ZmIDD21 were orthologs of AtIDD14, the promoter of ZmIDD3 and ZmIDD21 does not have the AuxRR-element, thus developing different functions in plant growth. Wang et al. [40] reported that miRNA167 could directly regulate the auxin response factors GmARF8a and GmARF8b and participate in lateral root development in soybean. In this study, miR167 could target and lead to the degradation of ZmIDD7 (Figure 10; Table S13). In addition, xyloglucan endotransglucosylase/hydrolase (XTH) is regulated by auxin and functions in plant developmental plasticity [41]. Gullner et al. [42] reported that glutathione-S-transferases (GSTs) were involved in intracellular auxin transport. Here, ZmIDD1 and ZmIDD13 could target xyloglucan endotransglucosylase/hydrolase protein 32 and glutathione S-transferase GST 21 (Table S10). The results demonstrated that ZmIDD may also be responsible for auxin biosynthesis, thus regulating plant growth.
Except for the zinc finger domain, the IDD proteins owned two conserved amino acid residues (Ser73 and Ser182): Ser73 could be phosphorylated by MPK6 [38], and Ser182 has been reported to be modified by AKIN10 (the catalytic subunit of SnRK1) [37]. Here, all 22 ZmIDDs have Ser73 and Ser182 (Figure 3). Meanwhile, the amino acid length of 22 ZmIDD proteins ranged from 354 to 815 (Table S1). It indicated that although the ZmIDDs contained a highly similar ID domain in their N-terminal, the flank sequences varied greatly. Jeong et al. [37] reported that phosphorylation of AtIDD8 at Ser-182 obviously decreased its transcriptional activity, and atidd8 mutants or over-expression of AKIN10 led to a delay in flowering in Arabidopsis. In the current study, protein-protein interactions (PPI) between ZmIDDs and ZmSnRKs were predicted using the STRING web server. Only one ZmSnRK protein (ZmSnRK1.1) that could interact with the ZmIDD protein (ZmIDD5) was found (Figure S3). ZmIDD5 exhibited expression in most tissues except for the root and showed down-regulation under cold and heat stress (Figure 6; Tables S8 and S9). Meanwhile, miR156 might bind and cause the cleavage of ZmIDD5 (Figure 10; Table S13). It was reported that miR156 could enhance cold and heat stress tolerance by repressing the expression level of transcription factors in plants [43,44]. It suggests that ZmIDD5 may also respond to abiotic stresses and regulate flowering time. The starch degradation in guard cells can cause stomatal opening [45], which is likely to increase guard cell turgor pressure by supplying soluble sugars. Furthermore, Seo et al. [30] discovered that AtIDD14 directly activated qua-quine starch, thereby mediating starch metabolism in Arabidopsis. Hence, IDD14-regulated starch degradation is likely to be responsible for stomatal opening. In this study, ZmIIDD6 and ZmIDD11 were down-regulated under drought stress, but other ZmIDDs do not respond to drought responses (Figure 8; Table S9). The promoters of ZmIDD6 and ZmIDD11 have an ABRE element and a MBS element, respectively (Figure 7; Table S7). It indicated that ZmIDD could also respond to drought stress in maize. The Indeterminate Domain Protein ROC1 (the ortholog of ZmIDD21) enhanced chilling tolerance by activating DREB1B/CBF1 in rice [24]. Phytochrome A signal transduction 1 (PAT1) could interact with IDD3 to activate lipoxygenase 3 (LOX3) and increase JA-Ile accumulation in grape calli under low temperature stress [46]. In the current study, ZmIDD21 was up-regulated under cold stress (Figure 8; Table S9), implying ZmIDD21 was also a positive regulator of cold stress.
Generally, we conducted a genome-wide analysis of IDD genes in maize. It has a vital implication for further comprehending the biological functions of ZmIDDs. However, a lot of functional verification work needs to be conducted in future studies.

4. Materials and Methods

4.1. Identification of ZmIDDs in Zea mays

The genome and protein sequences of maize B73 were obtained from MaizeGDB (https://maizegdb.org/ (accessed on 12 October 2022)). With default parameters, the Hidden Markov Model (HMMER3.0) profile of the C2H2 protein domain (PF00096) was used to search the protein database in the maize genome [47]. Then, we used the maize genome database to conduct BLASP. Gene and protein sequences in Arabidopsis were downloaded from the Ensemble Plants database [48]. The longest transcripts were retained, and incomplete sequences were deleted. In addition, we used the SMART database to reconfirm sequences [49]. Putative ZmIDD genes were named ZmIDD1 to ZmIDD22 based on their chromosome localization. Moreover, protein characteristics were determined by the ExPASy tool (http://www.expasy.org/tools/ (accessed on 20 October 2022)) [50].

4.2. Phylogenetic Tree, Gene Structure, Conserved Motifs, and Cis-Elements Analysis

Multiple protein sequence alignments were conducted using ClustalW [51]. A phylogenetic tree was constructed by MEGA version 7.0 through the maximum likelihood method, Neighbor-Joining method, and Minimum Evolution method [52]. Firstly, the evolutionary history was inferred by using the Maximum Likelihood method based on the Jones-Taylor-Thornton (JTT) matrix-based model. Initial tree(s) for the heuristic search were obtained automatically by applying Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances estimated using a JTT model and then selecting the topology with the highest log likelihood value. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. Secondly, evolutionary history was inferred using the Minimum Evolution method. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) is shown next to the branches. The evolutionary distances were computed using the p-distance method and are in units of the number of amino acid differences per site. The ME tree was searched using the Close-Neighbor-Interchange (CNI) algorithm at a search level of 1. The tree is drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. Finally, evolutionary history was inferred using the Neighbor-Joining method. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) is shown next to the branches. The evolutionary distances were computed using the p-distance method and are in units of the number of amino acid differences per site. In addition, we used iTOL to beautify the evolutionary tree (https://itol.embl.de/ (accessed on 21 October 2022)) [53].
The DNA structure, such as exon and intron arrangement, was detected by the Gene Structure Display Server (GSDS) tool (http://gsds.cbi.pku.edu.cn/ (accessed on 21 October 2022)) [54]. The conserved motifs were analyzed by the MEME program with default parameters [55] and annotated using the InterProScan database (http://www.ebi.ac.uk/Tools/pfa/iprscan/ (accessed on 1 November 2022)) [56]. The gene structures and motifs were then imaged using the TBtools software (v1.09832) [57].
Promoter sequences (−1.5 kb) of each ZmIDD gene were used to scan any potential cis-elements by PlantCARE (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/ (accessed on 12 November 2022)) [58]. Then, we used TBtools to visualize the architecture [57].

4.3. Chromosome Distribution and Synteny Block of ZmIDDs

We used the Circos tool to locate ZmIDD genes on maize chromosome [59]. The collinearity of the orthologous IDD genes between maize and rice was analyzed using MCScanX software [60]. The nonsynonymous substitution rate (Ka) and synonymous substitution rate (Ks) of ZmIDDs were calculated by the ParaAT tool [61]. The Ka:Ks ratio was then calculated using Calculator 2.0 software [62]. In addition, Ks/2λ was used to estimate gene duplication time [63], where λ = 1.5 × 10 − 8.

4.4. Modeling of 3D Structures of ZmIDDs and Zinc Finger Alignment

We used Swiss-Model (https://swissmodel.expasy.org/interactive/ (accessed on 12 January 2023)) to predict the 3D structure of ZmIDD proteins and the SAVES server to assess the 3D structure (http://saves.mbi.ucla.edu/ (accessed on 12 January 2023)) [64]. Meanwhile, the amino acid alignment of the zinc finger was performed by DNAMAN.

4.5. Prediction of ZmIDDs-Regulated Genes

To better comprehend the regulatory mechanism of ZmIDDs, the target genes regulated by ZmIDDs were predicted through the online PlantRegMap tool [65], where the species and parameters were set to Zea mays, Organ:All, Method:FunTFBS, and Mode:TF (retrieve targets). Then, Gene Ontology (GO) enrichment of target genes was analyzed by an online bioinformatics server (http://www.bioinformatics.com.cn/?p=1 (accessed on 1 February 2023)) with a p-value < 0.05.

4.6. Prediction of miRNA-ZmIDD Regulatory Networks

The miRNAs, which can regulate the ZmIDD gene expression, were predicted using the psRNATarget server with the following parameters: the penalty for a G:U pair was one, and the number of mismatches allowed in the seed region was zero [66]. Then, we used Cytoscape V3.8.2 software (https://cytoscape.org/download.html (accessed on 1 February 2023)) to image the interaction networks.

4.7. Expression Profile Analysis of ZmIDDs

The tissue-specific and abiotic stress-induced transcriptional data of ZmIDDs in maize B73 were downloaded from qTeller in MaizeGDB [67,68]. We used the DSEeq2 R package to conduct differential expression analysis, and the heatmaps were imaged by TBtools software. The genes with a |log2 ratio|≥1 were considered differentially expressed genes (DEGs).

5. Conclusions

In the current study, we identified and characterized 22 ZmIDD proteins that contained a complete IDD domain. On the basis of their amino acid sequences, the 22 ZmIDDs were categorized into three clades. In the course of the evolution of ZmIDD genes, segmental duplication events played a dominant role. Moreover, the cis-acting elements, gene expression, and regulatory network of ZmIDD families were also analyzed. These findings add to our understanding of the ZmIDD gene family’s characteristics and provide useful information for further functional characterization of ZmIDDs in maize climate resilience.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24076185/s1.

Author Contributions

X.F. and W.L. conceived and designed the research. X.F., Q.Y., J.Z. and X.H. performed the experiments and data analyses. W.L., X.F. and Q.Y. wrote the article. L.G. and W.M. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Shandong Province (grant No. ZR2021QC052) and the National Natural Science Foundation of China (grant No. 32101660).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data analyzed during this study are included in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, T.; Chen, T.; Kan, J.; Yao, Y.; Guo, D.; Yang, Y.; Ling, X.; Wang, J.; Zhang, B. The GhMYB36 transcription factor confers resistance to biotic and abiotic stress by enhancing PR1 gene expression in plants. Plant Biotechnol. J. 2022, 20, 722–735. [Google Scholar] [CrossRef] [PubMed]
  2. Song, L.; Huang, S.C.; Wise, A.; Castanon, R.; Nery, J.R.; Chen, H.; Watanabe, M.; Thomas, J.; Bar-Joseph, Z.; Ecker, J.R. A transcription factor hierarchy defines an environmental stress response network. Science 2016, 354, aag1550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Franco-Zorrilla, J.M.; López-Vidriero, I.; Carrasco, J.L.; Godoy, M.; Vera, P.; Solano, R. DNA-binding specificities of plant transcription factors and their potential to define target genes. Proc. Natl. Acad. Sci. USA 2014, 111, 2367–2372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Han, G.; Lu, C.; Guo, J.; Qiao, Z.; Sui, N.; Qiu, N.; Wang, B. C2H2 Zinc finger proteins: Master tegulators of abiotic stress responses in plants. Front. Plant Sci. 2020, 20, 115. [Google Scholar] [CrossRef] [Green Version]
  5. Khoso, M.A.; Hussain, A.; Ritonga, F.N.; Ali, Q.; Channa, M.M.; Alshegaihi, R.M.; Meng, Q.; Ali, M.; Zaman, W.; Brohi, R.D.; et al. WRKY transcription factors (TFs): Molecular switches to regulate drought, temperature, and salinity stresses in plants. Front. Plant Sci. 2022, 13, 1039329. [Google Scholar] [CrossRef]
  6. Lu, R.; Li, Y.; Zhang, J.; Wang, Y.; Zhang, J.; Li, Y.; Zheng, Y.; Li, X.B. The bHLH/HLH transcription factors GhFP2 and GhACE1 antagonistically regulate fiber elongation in cotton. Plant Physiol. 2022, 189, 628–643. [Google Scholar] [CrossRef]
  7. Ma, D.; Constabel, C.P. MYB Repressors as regulators of phenylpropanoid metabolism in plants. Trends Plant Sci. 2019, 24, 275–289. [Google Scholar] [CrossRef]
  8. Joo, H.; Baek, W.; Lim, C.W.; Lee, S.C. Post-translational modifications of bZIP transcription factors in abscisic acid signaling and drought responses. Curr. Genom. 2021, 22, 4–15. [Google Scholar] [CrossRef]
  9. Wang, Y.; Zhang, J.; Hu, Z.; Guo, X.; Tian, S.; Chen, G. Genome-wide analysis of the MADS-Box transcription factor family in Solanum lycopersicum. Int. J. Mol. Sci. 2019, 20, 2961. [Google Scholar] [CrossRef] [Green Version]
  10. Colasanti, J.; Tremblay, R.; Wong, A.Y.; Coneva, V.; Kozaki, A.; Mable, B.K. The maize INDETERMINATE1 flowering time regulator defines a highly conserved zinc finger protein family in higher plants. BMC Genom. 2006, 7, 158. [Google Scholar] [CrossRef] [Green Version]
  11. Coneva, V.; Zhu, T.; Colasanti, J. Expression differences between normal and indeterminate1 maize suggest downstream targets of ID1, a floral transition regulator in maize. J. Exp. Bot. 2007, 58, 3679–3693. [Google Scholar] [CrossRef] [Green Version]
  12. Yoshida, H.; Hirano, K.; Sato, T.; Mitsuda, N.; Nomoto, M.; Maeo, K.; Koketsu, E.; Mitani, R.; Kawamura, M.; Ishiguro, S.; et al. DELLA protein functions as a transcriptional activator through the DNA binding of the indeterminate domain family proteins. Proc. Natl. Acad. Sci. USA 2014, 111, 7861–7866. [Google Scholar] [CrossRef] [Green Version]
  13. Hirano, Y.; Nakagawa, M.; Suyama, T.; Murase, K.; Shirakawa, M.; Takayama, S.; Sun, T.; Hakoshima, T. Structure of the SHR-SCR heterodimer bound to the BIRD/IDD transcriptional factor JKD. Nat. Plants 2017, 3, 17010. [Google Scholar] [CrossRef] [Green Version]
  14. Colasanti, J.; Yuan, Z.; Sundaresan, V. The indeterminate gene encodes a zinc finger protein and regulates a leaf-generated signal required for the transition to flowering in maize. Curr. Opin. Plant Biol. 1998, 1, 276. [Google Scholar] [CrossRef] [Green Version]
  15. Kumar, M.; Le, D.T.; Hwang, S.; Seo, P.J.; Kim, H.U. Role of the INDETERMINATE DOMAIN genes in Plants. Int. J. Mol. Sci. 2019, 20, 2286. [Google Scholar] [CrossRef] [Green Version]
  16. Zhang, T.; Tan, M.; Geng, L.; Li, J.; Xiang, Y.; Zhang, B.; Zhao, Y. New insight into comprehensive analysis of INDETERMINATE DOMAIN (IDD) gene family in rice. Plant Physiol. Biochem. 2020, 154, 547–556. [Google Scholar] [CrossRef]
  17. Guo, X.; Zhou, M.; Chen, J.; Shao, M.; Zou, L.; Ying, Y.; Liu, S. Genome-wide identification of the highly conserved INDETERMINATE DOMAIN (IDD) zinc finger gene family in Moso Bamboo (Phyllostachys edulis). Int. J. Mol. Sci. 2022, 23, 13952. [Google Scholar] [CrossRef]
  18. Coelho, C.P.; Huang, P.; Lee, D.Y.; Brutnell, T.P. Making roots, shoots, and seeds: IDD gene family diversification in plants. Trends Plant Sci. 2018, 23, 66–78. [Google Scholar] [CrossRef]
  19. Prochetto, S.; Reinheimer, R. Step by step evolution of Indeterminate Domain (IDD) transcriptional regulators: From algae to angiosperms. Ann. Bot. 2020, 126, 85–101. [Google Scholar] [CrossRef]
  20. Fukazawa, J.; Teramura, H.; Murakoshi, S.; Nasuno, K.; Nishida, N.; Ito, T.; Yoshida, M.; Kamiya, Y.; Yamaguchi, S.; Takahashi, Y. DELLAs function as coactivators of GAI-ASSOCIATED FACTOR1 in regulation of gibberellin homeostasis and signaling in Arabidopsis. Plant Cell 2014, 26, 2920–2938. [Google Scholar] [CrossRef] [Green Version]
  21. Liu, J.; Shu, D.; Tan, Z.; Ma, M.; Guo, N.; Gao, S.; Duan, G.; Kuai, B.; Hu, Y.; Li, S.; et al. The Arabidopsis IDD14 transcription factor interacts with bZIP-type ABFs/AREBs and cooperatively regulates ABA-mediated drought tolerance. New Phytol. 2022, 236, 929–942. [Google Scholar] [CrossRef] [PubMed]
  22. Hu, S.; Dong, G.; Xu, J.; Su, Y.; Shi, Z.; Ye, W.; Li, Y.; Li, G.; Zhang, B.; Hu, J.; et al. A point mutation in the zinc finger motif of RID1/EHD2/OsID1 protein leads to outstanding yield-related traits in japonica rice variety Wuyunjing 7. Rice 2013, 6, 24. [Google Scholar] [CrossRef] [Green Version]
  23. Takehara, S.; Huang, J.; Hirano, K.; Ordonio, R.L.; Matsuoka, M.; UeguchiTanaka, M. OsIDD2, a zinc finger and Indeterminate Domain protein, regulates secondary cell wall formation. J. Integr. Plant Biol. 2018, 60, 130–143. [Google Scholar]
  24. Dou, M.; Cheng, S.; Zhao, B.; Xuan, Y.; Shao, M. The Indeterminate Domain Protein ROC1 regulates chilling tolerance via activation of DREB1B/CBF1 in rice. Int. J. Mol. Sci. 2016, 17, 233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Xuan, Y.H.; Priatama, R.A.; Huang, J.; Je, B.I.; Liu, J.M.; Park, S.J.; Piao, H.L.; Son, D.Y.; Lee, J.J.; Park, S.H.; et al. INDETERMINATE DOMAIN 10 regulates ammonium-mediated gene expression in rice roots. New Phytol. 2013, 197, 791–804. [Google Scholar] [CrossRef]
  26. Jöst, M.; Hensel, G.; Kappel, C.; Druka, A.; Sicard, A.; Hohmann, U.; Beier, S.; Himmelbach, A.; Waugh, R.; Kumlehn, J.; et al. The INDETERMINATE DOMAIN protein BROAD LEAF1 limits barley leaf width by restricting lateral proliferation. Curr. Biol. 2016, 26, 903–909. [Google Scholar] [CrossRef] [Green Version]
  27. Yi, G.; Neelakandan, A.K.; Gontarek, B.C.; Vollbrecht, E.; Becraft, P.W. The naked endosperm genes encode duplicate INDETERMINATE domain transcription factors required for maize endosperm cell patterning and differentiation. Plant Physiol. 2015, 167, 443–456. [Google Scholar] [CrossRef] [Green Version]
  28. Gontarek, B.C.; Neelakandan, A.K.; Wu, H.; Becraft, P.W. NKD transcription factors are central regulators of maize endosperm development. Plant Cell 2016, 28, 2916–2936. [Google Scholar] [CrossRef] [Green Version]
  29. Deng, L.; Li, L.; Zhang, S.; Shen, J.; Li, S.; Hu, S.; Peng, Q.; Xiao, J.; Wu, C. Suppressor of rid1 (SID1) shares common targets with RID1 on florigen genes to initiate floral transition in rice. PLoS Genet. 2017, 13, e1006642. [Google Scholar] [CrossRef] [Green Version]
  30. Seo, P.J.; Kim, M.J.; Ryu, J.Y.; Jeong, E.Y.; Park, C.M. Two splice variants of the IDD14 transcription factor competitively form nonfunctional heterodimers which may regulate starch metabolism. Nat. Commun. 2011, 2, 303. [Google Scholar] [CrossRef] [Green Version]
  31. Lee, S.C.; Lim, M.H.; Kim, J.A.; Lee, S.I.; Kim, J.S.; Jin, M.; Kwon, S.J.; Mun, J.H.; Kim, Y.K.; Kim, H.U.; et al. Transcriptome analysis in Brassica rapa under the abiotic stresses using Brassica 24K oligo microarray. Mol. Cells 2008, 26, 595–605. [Google Scholar]
  32. Wong, A.Y.; Colasanti, J. Maize floral regulator protein INDETERMINATE1 is localized to developing leaves and is not altered by light or the sink/source transition. J. Exp. Bot. 2007, 5, 403–414. [Google Scholar] [CrossRef]
  33. 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] [Green Version]
  34. Su, X.; Meng, T.; Zhao, Y.; Li, G.; Cheng, X.; Abdullah, M.; Sun, X.; Cai, Y.; Lin, Y. Comparative genomic analysis of the IDD genes in five Rosaceae species and expression analysis in Chinese white pear (Pyrus bretschneideri). PeerJ 2019, 7, e6628. [Google Scholar] [CrossRef] [Green Version]
  35. Fan, S.; Zhang, D.; Xing, L.; Qi, S.; Du, L.; Wu, H.; Shao, H.; Li, Y.; Ma, J.; Han, M. Phylogenetic analysis of IDD gene family and characterization of its expression in response to flower induction in Malus. Mol. Genet. Genom. 2017, 292, 755–771. [Google Scholar] [CrossRef]
  36. Waadt, R.; Seller, C.A.; Hsu, P.K.; Takahashi, Y.; Munemasa, S.; Schroeder, J.I. Plant hormone regulation of abiotic stress responses. Nat. Rev. Mol. Cell Biol. 2022, 23, 680–694. [Google Scholar] [CrossRef]
  37. Jeong, E.Y.; Seo, P.J.; Woo, J.C.; Park, C.M. AKIN10 delays flowering by inactivating IDD8 transcription factor through protein phosphorylation in Arabidopsis. BMC Plant Biol. 2015, 15, 110. [Google Scholar] [CrossRef] [Green Version]
  38. Völz, R.; Rayapuram, N.; Hirt, H. Phosphorylation regulates the activity of INDETERMINATE-DOMAIN (IDD/BIRD) proteins in response to diverse environmental conditions. Plant Signal. Behav. 2019, 14, e1642037. [Google Scholar] [CrossRef]
  39. Cui, D.; Zhao, J.; Jing, Y.; Fan, M.; Liu, J.; Wang, Z.; Xin, W.; Hu, Y. The Arabidopsis IDD14, IDD15 and IDD16 cooperatively regulate lateral organ morphogenesis and gravitropism by promoting auxin biosynthesis and transport. PLoS Genet. 2013, 9, e1003759. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, Y.; Li, K.; Chen, L.; Zou, Y.; Liu, H.; Tian, Y.; Li, D.; Wang, R.; Zhao, F.; Ferguson, B.J.; et al. MicroRNA167-Directed Regulation of the Auxin Response Factors GmARF8a and GmARF8b Is Required for Soybean Nodulation and Lateral Root Development. Plant Physiol. 2015, 168, 984–999. [Google Scholar] [CrossRef] [Green Version]
  41. Sasidharan, R.; Keuskamp, D.H.; Kooke, R.; Voesenek, L.A.; Pierik, R. Interactions between auxin, microtubules and XTHs mediate green shade-induced petiole elongation in arabidopsis. PLoS ONE 2014, 9, e90587. [Google Scholar] [CrossRef] [PubMed]
  42. Gullner, G.; Komives, T.; Király, L.; Schröder, P. Glutathione S-Transferase Enzymes in Plant-Pathogen Interactions. Front. Plant Sci. 2018, 9, 1836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zhou, M.; Tang, W. MicroRNA156 amplifies transcription factor-associated cold stress tolerance in plant cells. Mol. Genet. Genom. 2019, 294, 379–393. [Google Scholar] [CrossRef] [PubMed]
  44. Arshad, M.; Hannoufa, A. Alfalfa transcriptome profiling provides insight into miR156-mediated molecular mechanisms of heat stress tolerance. Genome 2022, 65, 315–330. [Google Scholar] [CrossRef] [PubMed]
  45. Horrer, D.; Flutsch, S.; Pazmino, D.; Matthews, J.S.; Thalmann, M.; Nigro, A.; Leonhardt, N.; Lawson, T.; Santelia, D. Blue light induces a distinct starch degradation pathway in guard cells for stomatal opening. Curr. Biol. 2016, 26, 362–370. [Google Scholar] [CrossRef] [Green Version]
  46. Wang, Z.; Wong, D.C.J.; Wang, Y.; Xu, G.; Ren, C.; Liu, Y.; Kuang, Y.; Fan, P.; Li, S.; Xin, H.; et al. GRAS-domain transcription factor PAT1 regulates jasmonic acid biosynthesis in grape cold stress response. Plant Physiol. 2021, 186, 1660–1678. [Google Scholar] [CrossRef]
  47. Finn, R.D.; Clements, J.; Eddy, S.R. HMMER web server: Interactive sequence similarity searching. Nucleic Acids Res. 2011, 39, 29–37. [Google Scholar] [CrossRef] [Green Version]
  48. Kersey, P.J.; Allen, J.E.; Allot, A.; Barba, M.; Boddu, S.; Bolt, B.J.; Carvalho-Silva, D.; Christensen, M.; Davis, P.; Grabmueller, C.; et al. Ensemble Genomes 2018: An integrated omics infrastructure for non-vertebrate species. Nucleic Acids Res. 2018, 46, 802–808. [Google Scholar] [CrossRef]
  49. Letunic, I.; Doerks, T.; Bork, P. SMART 7: Recent updates to the protein domain annotation resource. Nucleic Acids Res. 2012, 40, 302–305. [Google Scholar] [CrossRef]
  50. Artimo, P.; Jonnalagedda, M.; Arnold, K.; Baratin, D.; Csardi, G.; Castro, E.; Duvaud, S.; Flegel, V.; Fortier, A.; Gasteiger, E.; et al. ExPASy: SIB bioinformatics resource portal. Nucleic Acids Res. 2012, 40, 597–603. [Google Scholar] [CrossRef]
  51. Larkin, M.A.; Blackshields, G.; Brown, N.P.; Chenna, R.; McGettigan, P.A.; McWilliam, H.; Valentin, F.; Wallace, I.M.; Wilm, A.; Lopez, R.; et al. Clustal W and clustal X version 2.0. Bioinformatics 2007, 23, 2947–2948. [Google Scholar] [CrossRef] [Green Version]
  52. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [Green Version]
  53. Letunic, I.; Bork, P. Interactive Tree of Life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 2019, 47, 256–259. [Google Scholar] [CrossRef] [Green Version]
  54. Hu, B.; Jin, J.; Guo, A.; Zhang, H.; Luo, J.; Gao, G. GSDS 2.0: An upgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef] [Green Version]
  55. Bailey, T.L.; Johnson, J.; Grant, C.E.; Noble, W.S. The MEME Suite. Nucleic Acids Res. 2015, 43, 39–49. [Google Scholar] [CrossRef] [Green Version]
  56. Jones, P.; Binns, D.; Chang, H.Y.; Fraser, M.; Li, W.; McAnulla, C.; McWilliam, H.; Maslen, J.; Mitchell, A.; Nuka, G.; et al. InterProScan 5: Genome-scale protein function classification. Bioinformatics 2014, 30, 1236–1240. [Google Scholar] [CrossRef] [Green Version]
  57. 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]
  58. Lescot, M.; Dehais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouze, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef]
  59. Gu, Z.; Gu, L.; Eils, R.; Schlesner, M.; Brors, B. Circlize Implements and enhances circular visualization in R. Bioinformatics 2014, 30, 2811–2812. [Google Scholar] [CrossRef] [Green Version]
  60. Wang, Y.; Tang, H.; DeBarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.H.; Jin, H.; Marler, B.; Guo, H.; et al. MCScanX: A toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012, 40, 49. [Google Scholar] [CrossRef] [Green Version]
  61. Zhang, Z.; Xiao, J.; Wu, J.; Zhang, H.; Liu, G.; Wang, X.; Dai, L. ParaAT: A parallel tool 665 for constructing multiple protein-coding DNA alignments. Biochem. Biophys. Res. Commun. 2012, 419, 779–781. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, D.; Zhang, Z.; Zhang, Z.; Zhu, J.; Yu, J. KaKs_Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef] [Green Version]
  63. Koch, M.A.; Haubold, B.; Mitchell-Olds, T. Comparative evolutionary analysis of chalcone synthase and alcohol dehydrogenase loci in Arabidopsis, Arabis, and related genera (Brassicaceae). Mol. Biol. Evol. 2000, 17, 1483–1498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Arnold, K.; Bordoli, L.; Kopp, J.; Schwede, T. The SWISS-MODEL workspace: A web-based environment for protein structure homology modelling. Bioinformatics 2006, 22, 195–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Tian, F.; Yang, D.C.; Meng, Y.Q.; Jin, J.; Gao, G. PlantRegMap: Charting functional regulatory maps in plants. Nucleic Acids Res. 2020, 48, 1104–1113. [Google Scholar] [CrossRef]
  66. Su, W.; Raza, A.; Zeng, L.; Gao, A.; Lv, Y.; Ding, X.; Cheng, Y.; Zou, X. Genome-wide analysis and expression patterns of lipid phospholipid phospholipase gene family in Brassica napus L. BMC Genom. 2021, 22, 548. [Google Scholar] [CrossRef]
  67. Opitz, N.; Paschold, A.; Marcon, C.; Malik, W.A.; Lanz, C.; Piepho, H.P.; Hochholdinger, F. Transcriptomic complexity in young maize primary roots in response to low water potentials. BMC Genom. 2014, 15, 741. [Google Scholar] [CrossRef] [Green Version]
  68. 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, plantgenome2015.04.0025. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic tree of full-length ZmIDD, AtIDD, and OsIDD proteins using the Maximum Likelihood method based on the JTT matrix-based model. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The analysis involved 54 amino acid sequences. All positions containing gaps and missing data were eliminated. There were a total of 94 positions in the final dataset. The different colored arcs indicate subfamilies of the IDD proteins. Different colored shapes represent IDDs from maize (☆), rice (√), and Arabidopsis (☐). IDD in Chlamydomonas reinhardtii was selected as an outgroup.
Figure 1. Phylogenetic tree of full-length ZmIDD, AtIDD, and OsIDD proteins using the Maximum Likelihood method based on the JTT matrix-based model. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The analysis involved 54 amino acid sequences. All positions containing gaps and missing data were eliminated. There were a total of 94 positions in the final dataset. The different colored arcs indicate subfamilies of the IDD proteins. Different colored shapes represent IDDs from maize (☆), rice (√), and Arabidopsis (☐). IDD in Chlamydomonas reinhardtii was selected as an outgroup.
Ijms 24 06185 g001
Figure 2. Phylogenetic relationships, architecture of conserved protein motifs, and gene structure in IDD genes from maize. (A) The phylogenetic tree was constructed based on the full-length sequences of maize IDD proteins using MEGA version 7.0 software. (B) The motif compositions of 22 ZmIDD proteins. The motifs were identified using the MEME program. Boxes of different colors represent motifs 1 to 10. The length of the amino acid sequences can be estimated by the scale at the bottom. (C) Gene structures of 22 ZmIDD genes. Yellow boxes represent exons, green boxes represent 5′ or 3′ untranslated regions (UTR), and black lines represent introns. The length of nucleotide sequences of exons/introns/UTRs can be estimated by the scale at the bottom.
Figure 2. Phylogenetic relationships, architecture of conserved protein motifs, and gene structure in IDD genes from maize. (A) The phylogenetic tree was constructed based on the full-length sequences of maize IDD proteins using MEGA version 7.0 software. (B) The motif compositions of 22 ZmIDD proteins. The motifs were identified using the MEME program. Boxes of different colors represent motifs 1 to 10. The length of the amino acid sequences can be estimated by the scale at the bottom. (C) Gene structures of 22 ZmIDD genes. Yellow boxes represent exons, green boxes represent 5′ or 3′ untranslated regions (UTR), and black lines represent introns. The length of nucleotide sequences of exons/introns/UTRs can be estimated by the scale at the bottom.
Ijms 24 06185 g002
Figure 3. Comparative amino acid sequence alignment of INDETERMINATE DOMAIN (IDD) genes that shows motifs or domains that are conserved in maize. Black boxes mark the positions of cysteines (C, in blue triangles) and histidines (H, in yellow triangles) characterized for each zinc finger.
Figure 3. Comparative amino acid sequence alignment of INDETERMINATE DOMAIN (IDD) genes that shows motifs or domains that are conserved in maize. Black boxes mark the positions of cysteines (C, in blue triangles) and histidines (H, in yellow triangles) characterized for each zinc finger.
Ijms 24 06185 g003
Figure 4. The 3D structure modeling of ZmIDD proteins. The Pymol software was used to create the structural image.
Figure 4. The 3D structure modeling of ZmIDD proteins. The Pymol software was used to create the structural image.
Ijms 24 06185 g004
Figure 5. Distribution of ZmIDD genes in maize chromosomes. The chromosome numbers are indicated at the top of each chromosome image.
Figure 5. Distribution of ZmIDD genes in maize chromosomes. The chromosome numbers are indicated at the top of each chromosome image.
Ijms 24 06185 g005
Figure 6. The synteny analysis of ZmIDD family genes. (A) The synteny analysis of the ZmIDD family in maize. Gray lines indicate all synteny blocks in the maize genome. The genes linked by red lines represent homologues. (B) Synteny analysis of IDD genes between maize and rice. Gray lines: all collinear blocks within maize and other plant genomes. Red lines: the synteny of IDD gene pairs. The species names with the prefixes Zm and Os indicate maize and rice, respectively.
Figure 6. The synteny analysis of ZmIDD family genes. (A) The synteny analysis of the ZmIDD family in maize. Gray lines indicate all synteny blocks in the maize genome. The genes linked by red lines represent homologues. (B) Synteny analysis of IDD genes between maize and rice. Gray lines: all collinear blocks within maize and other plant genomes. Red lines: the synteny of IDD gene pairs. The species names with the prefixes Zm and Os indicate maize and rice, respectively.
Ijms 24 06185 g006
Figure 7. Predicted cis-regulatory elements in ZmIDD promoters. Promoter sequences (about 1.5 kb) of 22 ZmIDD genes were analyzed by PlantCARE. The upstream length to the translation starting site can be inferred according to the scale at the bottom.
Figure 7. Predicted cis-regulatory elements in ZmIDD promoters. Promoter sequences (about 1.5 kb) of 22 ZmIDD genes were analyzed by PlantCARE. The upstream length to the translation starting site can be inferred according to the scale at the bottom.
Ijms 24 06185 g007
Figure 8. Expression profiles of the ZmIDD genes. (A) Expression profiles of the ZmIDD genes in different tissues. The color scale represents expression data with a row scale. Blue: low expression; red: high expression. (B) Expression profiles of the ZmIDD genes under different abiotic stresses. Expression data were the ratios to control values. The color scale represents expression levels from upregulation (red) to downregulation (blue).
Figure 8. Expression profiles of the ZmIDD genes. (A) Expression profiles of the ZmIDD genes in different tissues. The color scale represents expression data with a row scale. Blue: low expression; red: high expression. (B) Expression profiles of the ZmIDD genes under different abiotic stresses. Expression data were the ratios to control values. The color scale represents expression levels from upregulation (red) to downregulation (blue).
Ijms 24 06185 g008
Figure 9. Bubble map of GO enrichment of ZmIDD-target genes. Genes were listed in Table S12. The X-axis represents the Rich Ratio. Rich Ratio = Term Candidate Gene Number/Term Gene Number. The Y-axis represents the GO Term. The size of the bubble represents the number of different genes annotated to a GO Term, and the color represents the enriched p value. 0 < p value < 1. The smaller the p value, the more significant the GO enrichment.
Figure 9. Bubble map of GO enrichment of ZmIDD-target genes. Genes were listed in Table S12. The X-axis represents the Rich Ratio. Rich Ratio = Term Candidate Gene Number/Term Gene Number. The Y-axis represents the GO Term. The size of the bubble represents the number of different genes annotated to a GO Term, and the color represents the enriched p value. 0 < p value < 1. The smaller the p value, the more significant the GO enrichment.
Ijms 24 06185 g009
Figure 10. Interaction networks between miRNAs and miRNA-acted ZmIDDs. Information on miRNAs and miRNAs-acted ZmIDDs is shown in Table S13.
Figure 10. Interaction networks between miRNAs and miRNA-acted ZmIDDs. Information on miRNAs and miRNAs-acted ZmIDDs is shown in Table S13.
Ijms 24 06185 g010
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

Feng, X.; Yu, Q.; Zeng, J.; He, X.; Ma, W.; Ge, L.; Liu, W. Comprehensive Analysis of the INDETERMINATE DOMAIN (IDD) Gene Family and Their Response to Abiotic Stress in Zea mays. Int. J. Mol. Sci. 2023, 24, 6185. https://doi.org/10.3390/ijms24076185

AMA Style

Feng X, Yu Q, Zeng J, He X, Ma W, Ge L, Liu W. Comprehensive Analysis of the INDETERMINATE DOMAIN (IDD) Gene Family and Their Response to Abiotic Stress in Zea mays. International Journal of Molecular Sciences. 2023; 24(7):6185. https://doi.org/10.3390/ijms24076185

Chicago/Turabian Style

Feng, Xue, Qian Yu, Jianbin Zeng, Xiaoyan He, Wujun Ma, Lei Ge, and Wenxing Liu. 2023. "Comprehensive Analysis of the INDETERMINATE DOMAIN (IDD) Gene Family and Their Response to Abiotic Stress in Zea mays" International Journal of Molecular Sciences 24, no. 7: 6185. https://doi.org/10.3390/ijms24076185

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