Next Article in Journal
Dynamics of Greenhouse Gas Fluxes in Açaí Cultivation: Comparing Amazonian Upland and Floodplain Soils
Previous Article in Journal
Thermal Modulation of Leaf Nitrogen Forms in Chinese Fir Under Soil-Warming Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification and Expression Analysis of AT-Hook Motif Nuclear Localized Gene Family in Birch

1
Institute of Carbon Neutrality, Key Laboratory of Sustainable Forest Ecosystem Management-Ministry of Education, School of Ecology, Northeast Forestry University, Harbin 150040, China
2
Heilongjiang Maoershan Forest Ecosystem National Observation and Research Station, School of Ecology, Northeast Forestry University, Harbin 150040, China
3
College of Biology Resources and Environmental Sciences, Jishou University, Jishou 416000, China
4
College of Life Science, Northeast Forestry University, Harbin 150040, China
5
State Key Laboratory of Tree Genetics and Breeding, Northeast Forestry University, Harbin 150040, China
6
College of Agriculture, Guangxi University, Nanning 530004, China
7
College of Pharmacy, Hebei North University, Zhangjiakou 075031, China
8
College of Agriculture and Forestry Science and Technology, Hebei North University, Zhangjiakou 075031, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Forests 2025, 16(6), 943; https://doi.org/10.3390/f16060943
Submission received: 23 April 2025 / Revised: 27 May 2025 / Accepted: 1 June 2025 / Published: 4 June 2025
(This article belongs to the Section Genetics and Molecular Biology)

Abstract

The AT-hook motif nuclear localized (AHL) gene family encodes transcription factors pivotal in regulating plant growth, development, and responses to abiotic stimuli, including low temperature, salinity, darkness, and drought. In this study, we systematically identified 21 BpAHL genes in birch and characterized their sequence features, evolutionary relationships, and expression dynamics. Phylogenetic analysis classified BpAHLs into two clades (Clade-A and Clade-B) and three types (Type-I, -II, and -III), based on PPC domain and AT-hook motifs. Chromosomal mapping revealed an even distribution across nine chromosomes and one contig, with dispersed duplication events recognized as the major driver of BpAHL family expansion. Tissue-specific expression profiling uncovered striking divergence: Type-I BpAHLs displayed root-predominant expression, whereas Type-II/III BpAHLs were highly expressed in plant flowers and leaves. Notably, Type-II/III BpAHL genes in leaves showed distinct expression patterns in response to cold and heat stresses, while Type-I BpAHLs in roots were down-regulated under salt stress. This study provides a comprehensive phylogenomic and functional analysis of the AHLs in birch, providing insights into their roles in enhancing abiotic stress resilience in forest trees.

1. Introduction

With the increasing frequency of extreme weather events (e.g., droughts, floods) under climate change scenarios, understanding transcription factors (TFs)-mediated stress-responsive mechanisms in staple crops (e.g., rice, wheat) and economically important forest tree species (e.g., poplar, birch) has become crucial for ensuring global food security and sustainable forestry development [1]. Among these TFs, the AT-hook motif nuclear localized (AHL) gene family has garnered worldwide research attention because of its pivotal role in regulating plant responses to various biotic and abiotic stresses [2].
The AHL gene family was first discovered in mammalian high mobility group (HMG) proteins [3] and has subsequently been characterized in numerous terrestrial plant species with sequenced genomes, including Arabidopsis thaliana (29 members) [4], Oryza sativa (26) [5], Zea mays (37) [6], Glycine max (63) [7], Vitis vinifera (14) [8], Daucus carota (47) [9], Brassica rapa (42) [10], Populus trichocarpa (37) [11], Liriodendron chinense (21) [12], and Gossypium species (48, 51, 99) [13]. AHL proteins function as nuclear-localized transcription factors that modulate both chromosome structure assembly and target gene expression. For instance, overexpression of AHL15 can induce somatic embryogenesis without exogenous hormones and promote tetraploid formation in L. chinense, suggesting its ability to influence cell differentiation through ploidy alteration [12]. In A. thaliana, AHL13 protein interacts with poly(ADP-ribose) modification, where PARylation suppresses its DNA-binding activity, thereby fine-tuning the transcriptional regulation of some immune-related genes [14]. The DNA-binding capability of AHL protein is conferred by two conserved structural features: the AT-hook motifs and the Plants and Prokaryotes Conserved (PPC/DUF296) [4,15]. The AT-hook contains a core sequence ‘Arg-Gly-Arg-Pro’ (R-G-R-P) essential for DNA binding, while the PPC domain (~120 amino acids) features a conserved structure comprising seven β-sheets surrounding a central α-helix [4,16]. The hydrophobic C-terminal region facilitates nuclear localization and mediates protein–protein interaction, enabling AHL proteins to regulate transcription through both direct DNA binding and protein complex formation [17].
The AHL proteins in land plants are phylogenetically classified into two clades (Clade-A and Clade-B), distinguished by characteristic PPC domain signatures: ‘Leu-Arg-Ser-His’ (L-R-S-H) in Clade-A versus ‘Phe-Thr-Pro-His’ (F-T-P-H) in Clade-B [4]. These AHL proteins are further categorized into three subtypes (Type-I, -II, and -III) based on sequence variations in their AT-hook motifs. Type-I AHLs uniquely possess a C-terminal ‘Gly-Ser-Lys-Asn-Lys’ (G-S-K-N-K) extension, while both Type-II and Type-III subtypes feature a conserved ‘Arg-Lys-Tyr’ (R-K-Y) tail [4,16]. Their conserved binding capability to AT-rich DNA and expression patterns suggest AHLs have maintained critical functions throughout evolution across plant species.
The AHLs family governs pivotal biological processes in plants, encompassing three major aspects:
1.
Growth and Development Regulation.
AtAHL22 modulates flowering time and hypocotyl elongation in A. thaliana through direct regulation of FLOWERING LOCUS T (FT) and PHYTOCHROME-INTERACTING FACTOR 4 (PIF4). While its overexpression delays flowering, knockout lines exhibit enhanced hypocotyl elongation [18]. Heterotopic expression of AtAHL15 suppresses axillary meristem (AM) development while promoting apical meristem growth, thereby altering shoot architecture [19]. Additionally, AHL proteins can participate in reproductive development, including maize ear morphogenesis [20].
2.
Transcriptional Regulation.
In cotton, GhAT1 acts as a transcriptional repressor of the FSltp4 promoter in non-fiber tissues, thereby fine-tuning fiber development [21]. The soybean AT-1SNBP, which shares structural homology with HMG-A proteins, binds to the glutamine synthetase GSI5 promoter to stabilize chromatin architecture and mediate distalproximal regulatory element interactions [22].
3.
Stress-Responsive Regulation.
Overexpression of AtAHL20 enhances immunity against bacterial pathogens in A. thaliana [23], while transgenic pepper lines overexpressing CaATL1 demonstrate broad-spectrum biotic stress resistance [24]. In rice, drought-induced OsAHL1 up-regulates OsCDPK7 expression in vascular tissues, conferring improved water-deficit tolerance [25]. Transcriptomic analyses associate specific GmAHLs with drought and flooding responses in soybean, underscoring their stress-responsive regulatory functions [7].
Despite these advances, research on AHLs in perennial woody plants remains limited. Birch (Betula platyphylla Suk.), a keystone broad-leaved deciduous hardwood tree species, is widely distributed across northern Europe, Russia, and northern China, where it forms both pure stands and mixed forests [26]. Birch holds significant medicinal, commercial, and ecological values, e.g., betulin—a bioactive compound extracted from its bark—exhibits cytotoxic activity with potential anticancer applications [27,28]. Its high-density wood, prized for its attractive white grain, is extensively used in furniture manufacturing, construction, and paper production [29]. Ecologically, sustainable birch cultivation enhances vegetation coverage, mitigates soil erosion, improves water conservation, and maintains forest ecosystem balance. However, natural birch are highly vulnerable to abiotic stresses, including salinity [30], osmotic stress [31], and low-temperatures [32]. Moreover, the AHL gene family in birch has not yet been comprehensively characterized. Here, we systematically investigated the birch AHL gene family through bioinformatics identification, phylogenetic and evolution analysis, structural predictions, expression profiling, and abiotic stress response assays. Our findings deepen the understanding of AHL gene functions in forest trees and provide a foundation for their application in breeding and stress resilience engineering.

2. Materials and Methods

2.1. Plant Material and Treatments

The seeds of Betula platyphylla used in this study were derived from ‘DL-1’, a superior white birch strain characterized by high differentiation efficiency, rapid growth, and elevated biomass [33]. These seeds were obtained from the State Key Laboratory of Tree Genetics and Breeding (TGB) at Northeast Forestry University in 2020. For germination, surface-sterilized seeds were cultured on Hoagland’s solution in Petri dishes for one week. Germinated seedlings were then transferred to plastic pots containing a sterilized growth medium (peat moss: garden soil: river sand = 5:2:1, v/v/v) and maintained at 26 °C under a 14 h light/10 h dark photoperiod [28]. After three months, various tissues (leaves, stems, and roots) were collected in three independent biological replicates per sample, flash-frozen in liquid nitrogen, and stored at −80 °C for qRT-PCR experiments.
To induce drought stress, three-month-old seedlings of uniform growth status were irrigated with 0.2 mol/L mannitol solution. Whole plants were photographed at 9:00 a.m. on days 1, 3, and 5 post-treatment in a PERCIVAL growth chamber (26 °C, ~40% relative humidity). Simultaneously, root tissues were independently collected (n = 3 biological replicates per group), immediately flash-frozen in liquid nitrogen, and stored at −80 °C for qRT-PCR experiments.

2.2. Characterization of BpAHL Gene Family

Putative AHL proteins in the birch genome [34] were identified using the HMMER 3.0 program [35] with the PPC/DUF296 (PF03479) domain (https://pfam.xfam.org; accessed on 2 January 2024) [36,37], and further validated via NCBI-BLASTp (E-value ≤ 1 × 10−10). Domain architecture was confirmed using Conserved Domains Database (CDD) (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi; accessed on 4 January 2024) [38] and SMART (https://smart.embl.de; accessed on 5 January 2024) [39], with an E-value cutoff of 1 × 10−5. Only proteins containing both one PPC domain and at least one AT-hook motif were classified as BpAHL family members; these were systematically numbered according to genomic position.
Chromosomal localization, gene length, CDS length, protein length, and GO annotation of the BpAHL gene family were extracted from birch genome annotations (https://genomevolution.org/coge/GenomeInfo.pl?gid=35080; accessed on 1 July 2023) [34]. The molecular weight (MW) and theoretical isoelectric point (pI) of each BpAHL protein were computed using ExPASy Proteomics Server (https://www.expasy.org; accessed on 20 January 2024) [40], and subcellular localization was predicted via the Plant-PLoc website (http://www.csbio.sjtu.edu.cn/bioinf/plant; accessed on 25 January 2024) [41].
Tertiary structures of the BpAHL proteins were modeled via SWISS-MODEL (https://swissmodel.expasy.org/interactive; accessed on 3 February 2024) [42], with output PDB files visualized in Swiss PDB Viewer v4.0.1 (https://spdbv.unil.ch; accessed on 5 February 2024) [43]. For biological function, GO enrichment analysis was performed with goatools v0.8.2 (https://github.com/tanghaibao/goatools; accessed on 10 February 2024) [44], where significantly enriched terms (Fisher’s exact test, p-value ≤ 0.05) were identified.

2.3. Phylogenetic Analysis of BpAHL Protein Family

To elucidate the evolutionary relationships of AHL proteins across diverse plant species, we conducted a comprehensive phylogenetic analysis with maize (https://plants.ensembl.org/index.html; accessed on 10 July 2023, assembly: Zm-B73-REFERENCE-NAM-5.0), soybean (https://phytozome-next.jgi.doe.gov; accessed on 10 July 2023, assembly: Glycine max Wm82.a2.v1), sorghum (https://plants.ensembl.org/index.html; accessed on 10 July 2023, assembly: Sorghum_bicolor_NCBIv3), A. thaliana (https://www.arabidopsis.org; accessed on 10 July 2023, assembly: TAIR10), and birch. The phylogenetic tree was constructed with MEGA11 (https://www.megasoftware.net; accessed on 12 July 2023) [45], which involved three key steps: (1) multiple sequence alignment with MUSCLE algorithm, (2) tree construction using the neighbor-joining (NJ) method, and (3) evaluation of branch support with 1000 bootstrap replicates to ensure topological reliability. The phylogenetic tree was exported in Newick format and visualized using the Interactive Tree of Life (iTOL) platform [46], with all protein sequences and the corresponding Newick file provided in Tables S1 and S2.

2.4. Protein Domain, Gene Structure, and Motif Prediction of BpAHL Gene Family

The BpAHL protein sequences were initially analyzed using the SMART website to identify conserved amino acid residues. Following our established methodology [7], we recorded specific residues as domain identifiers: the first 4/7 and 8/10 residues of the N-terminal and C-terminal AT-hook motifs, respectively, and the first 4 residues of the PPC domain. Phylogenetic classification of BpAHL proteins was performed using MEGA11 software, adopting an approach consistent with our previous work [7]. Genomic features including coding sequences (CDSs), exons, and introns of BpAHL genes were extracted from genome annotation files. These structural elements were visualized using the ‘Gene Structure View’ module in TBtools v2.136 (https://github.com/CJ-Chen/TBtools-II/releases; accessed on 17 October 2023) [47], enabling clear representation of gene architectures. Conserved protein motifs were predicted using the MEME website (https://meme-suite.org/meme; accessed on 15 February 2024) [48], with a stringent E-value cutoff of 1 × 105. The ten most statistically significant motifs were identified and graphically represented using the ‘Visualize MEME Motif Pattern’ module in TBtools software, facilitating comparative analysis of motif distributions across BpAHL proteins.

2.5. Chromosomal Localization, Gene Duplication Events, Synteny and Ka/Ks Analysis of BpAHL Gene Family

The chromosomal locations of BpAHL genes were mapped and visualized using the ‘Gene Location Visualize’ module in TBtools. Gene duplication events were identified by analyzing BpAHL proteins with MCScanX (https://github.com/wyp1125/MCScanX; accessed on 27 February 2024) [49], using an E-value threshold of 1 × 105, with numerical identifiers representing duplication types: ‘0’ (singleton), ‘1’ (dispersed), ‘2’ (proximal), ‘3’ (tandem), and ‘4’ (WGD/segmental). Synteny analysis of AHL proteins between five species (A. thaliana, soybean, rice, maize, Populus) and birch was performed using the ‘One Step MCScanX’ module in TBtools, with comparative genomic data obtained from the Ensembl Plants, Phytozome and TAIR websites. The evolutionary constraints on duplicated gene pairs (type ‘4’) were evaluated by calculating synonymous (Ks) and non-synonymous (Ka) substitution rates and their ratios (Ka/Ks) using the ‘Simple Ka/Ks Calculator’ module in TBtools. Divergence times were estimated using the formula T = Ks/2r, where r represents the clock-like rate of divergence. For dicots, r is typically estimated at 1.5 × 108 substitutions per synonymous site per year [50].

2.6. Cis-Acting Regulatory Elements (CREs) in the Promoters of BpAHL Genes Family

The 2.0 kb upstream promoter regions of BpAHL genes were extracted from the birch genome using the ‘Gtf/Gff3 Sequences Extract’ module in TBtools, followed by comprehensive CREs prediction via PlantCARE (https://bioinformatics.psb.ugent.be/webtools/plantcare/html; accessed on 8 March 2024) [51]. The analysis focused on identifying and classifying the ten most abundant CREs, which were subsequently visualized using the ‘Simple BioSequence Viewer’ module in TBtools for comparative analysis of promoter architectures.

2.7. Tissue-Specific Transcriptome Analysis of BpAHL Gene Family

Illumina high-throughput sequencing data encompassing multiple tissue types (leaf, root, flower, and xylem) from two-year-old birch specimens were obtained from the Experimental Station of Northeast Forestry University [52] and downloaded from the NCBI Sequence Read Archive (PRJNA535361). For expression quantification, clean reads were aligned to the birch genome using HISAT2 (https://daehwankimlab.github.io/hisat2; accessed on 2 April 2024) [53], followed by quantification with StringTie (https://ccb.jhu.edu/software/stringtie; accessed on 6 April 2024) [54], with default parameters, to generate transcripts per million (TPM) values. This approach employs probabilistic read allocation to transcript models, preserving multi-mapping reads to more accurately represent transcript abundance and minimize quantification bias. Finally, BpAHL expression profiles were row-scaled and visualized as a heatmap using the ‘pheatmap’ function in R (https://www.r-project.org; accessed on 15 August 2023).

2.8. Abiotic Transcriptome Analysis of BpAHL Gene Family

Transcriptome datasets were analyzed to examine BpAHL gene expression profiling under diverse abiotic stress conditions in Betula platyphylla. Four distinct experimental datasets were obtained from the NCBI Sequence Read Archive (SRA):
1.
Low-Temperature Treatments in Short Term.
Transcriptome data (PRJNA532995) from leaves of two-month-old birch treated with low-temperature (6 °C) for six periods (0.5 h, 1.0 h, 1.5 h, 2.0 h, 2.5 h, and 3.0 h) [55].
2.
Low- and High-Temperature Treatments in Long Term.
Transcriptome data (PRJNA811313) from leaves of two-month-old birch treated with low-temperature (6 °C) and high-temperature (35 °C) for six periods (6.0 h, 1 d, 2 d, 4 d, 7 d, and 14 d) [56].
3.
Salt Stress Treatments.
Transcriptome data (PRJNA790472) from leaves and roots of 10-week-old birch treated with 24 h of 0.2 M NaCl condition [57].
4.
Light/Dark Treatments.
Transcriptome data (PRJNA759706) from leaves of one-month-old birch treated with four days of light and dark cycles [58].
All datasets were processed using the bioinformatics pipelines described in Section 2.7, with genes showing TPM values < 2.0 excluded from subsequent analyses.

2.9. RNA Extraction and Quantitative RT-PCR (qRT-PCR)

Total RNA was isolated from fresh leaves, stems and roots of three-month-old birch seedlings using the OminiPlant RNA Kit (DNase I). High-quality RNA was reverse-transcribed into cDNA using the PrimeScripTM RT Reagent Kit with gDNA Eraser in a two-step process:
1.
Genomic DNA elimination.
A total of 10 µL of reaction mixture containing 2 µL RNA (500 ng/μL), 2 µL 5× gDNA Eraser Buffer, 1 µL gDNA Eraser, and 5 µL RNase-free ddH2O was incubated at 42 °C for 2 min followed by 10 min on ice.
2.
cDNA synthesis.
The reaction mixture was brought to 20 µL by adding 1 µL PrimeScript™ RT Enzyme Mix I, 4 µL 5× PrimeScript™ Buffer2, 1 µL RT Primer Mix, and 4 µL RNase-free ddH2O, and then incubated at 37 °C for 15 min, 85 °C for 5 s, and 4 °C for 10 min. The synthesized cDNA was diluted 20-fold and stored at −40 °C.
For qRT-PCR, the reaction mixtures (15 µL) contained 6.3 µL diluted cDNA, 0.6 µL each of forward and reverse primers (Table S3), and 7.5 µL FastStart Universal SYBR Green Master. Amplification was performed on an ABI7500 instrument (Applied Biosystems) under the following conditions: 95 °C for 10 min; 40 cycles of 95 °C for 10 s, and 60 °C for 30 s. This was followed by melt curve analysis (95 °C for 1 min, 65 °C for 30 s, then 95 °C for 0.11 °C per second down to 5 °C). Technical triplicates were performed with each sample, with relative transcript levels calculated using the 2−ΔΔCt method [59] and normalized to tubulin as the internal control [60,61].

2.10. Statistical Analysis

Differences in relative gene expression levels resulting from qRT-PCR were evaluated using one-way ANOVA. For transcriptome-derived gene expression levels (TPMs), differences were assessed using Student’s t-test. Statistical significance was set at a threshold of p-value ≤ 0.05. All figures were generated using the R program (https://www.r-project.org; accessed on 15 August 2023).

3. Results

3.1. Identification of BpAHL Gene Family

In this study, we identified 21 putative AHL genes in birch using the Hidden Markov Model (HMM) engine, systematically designated as BpAHL01 to BpAHL21 (Table 1). Tertiary structure analysis confirmed that all candidate proteins contain conserved architectural features, including PPC/DUF296 domains and AT-hook motifs (Figure S1 and Table S4), validating their classification as AHL family members. The BpAHL proteins exhibited molecular weights ranging from 24,150 Da (BpAHL04) to 52,128 Da (BpAHL16), with isoelectric points (pI) varying between 4.73 (BpAHL07) and 10.02 (BpAHL18), averaging 7.7. Subcellular localization predictions indicated exclusive nuclear targeting for all BpAHL proteins (Table 1), consistent with their functional role as transcription factors in nuclear gene regulation.

3.2. Phylogenetic Analysis of BpAHL Protein Family

To infer the evolutionary relationships among the AHL proteins in birch, we conducted a phylogenetic analysis using full-length BpAHL protein sequences. Our results indicated that BpAHL proteins are grouped into two major clades: Clade-A (11 members, 52%), and Clade-B (10, 48%) (Figure 1). Further protein sequence alignments subdivided these clades into three types: Type-I (11, 52%), Type-II (6, 29%), and Type-III (4, 19%) (Figure 1 and Figures S2–S4). This mirrors the classification patterns observed in other land plants such as maize, rice, and soybean [4,5,6,7,13].
Comparative sequence analysis uncovered distinct conservation patterns in the functional domains. The PPC domain showed greater variability in Clade-A (‘L-R-S-H’ in 45% and ‘L-R-A-H’ in 27% of members) compared to the predominant ‘F-T-P-H’ motif (80%) in Clade-B (Figure 1 and Figures S2–S4). While N-terminal AT-hook motifs were highly conserved (‘R-G-R-P’ in Clade-A; ‘R-G-R-P-R-K-Y’ in Clade-B), C-terminal motifs displayed significant divergence, with Type-I members containing ‘G-S-K-N’ (73%) and Types II-III lacking this motif entirely. Taken together, the sequence diversity of AT-hook motifs and the PPC domain observed in the BpAHL proteins likely contributes to a varied array of biological functions.
Reconstruction of phylogenetic relationships across five species (A. thaliana, soybean, maize, sorghum, and birch) confirmed the universal conservation of Clade-A and Clade-B, though with species-specific expansion patterns (Figure 2). Specifically, there are 15, 34, 20, 14, and 11 AHL proteins in A. thaliana, soybean, maize, sorghum, and birch, respectively, belonging to Clade-A, while Clade-B includes 14, 29, 17, 11, and 10 proteins from these species. Notably, birch contained the fewest AHL members (11 in Clade-A, 10 in Clade-B) compared to the expanded families in soybean (34/29) and maize (20/17). This conserved clade structure supports previous findings and an ancient divergence event [4], while the limited paralog expansion in birch highlights potential functional optimization within this woody species.

3.3. Sequence Characterization of BpAHL Gene Family

Comparative analysis of BpAHL gene structures revealed distinct organizational patterns among BpAHL types (Figure 3A and Table 1). Type-I members predominantly (10/11) feature single-exon structures (723–939 bp), while BpAHL07 represents an exception with two exons (7507 bp). In contrast, Type-II/III members exhibit more complex architectures, containing 5–8 exons within substantially longer genes (5017–12,190 bp), suggesting evolutionary expansion from simpler Type-I ancestors—a pattern consistent with maize AHL evolution [6]. This structural complexity, particularly through intron acquisition, likely facilitates alternative splicing and functional diversification in higher plants.
Subsequently, we compared the motifs among BpAHL proteins (Figure 3B). The ten most frequently conserved motifs are shown, with the corresponding amino acid residues illustrated in Figure S5. The lengths of these motifs ranged from 8 to 41 amino acids, with site ranges varying from 3 to 21 (Table S5). As expected, the 4th to 6th amino acid residues of Motif3, ‘R-G-R-P’, and the 5th to 11th amino acid residues of Motif5, ‘R-G-R-P-R-K-Y’, both represent conserved residues in N-terminal AT-hook motifs, which are predominantly found in Type-I (11 members, 100%) and Type-II/III (10, 100%) BpAHLs, respectively. The 9th to 16th amino acid residues of Motif3, ‘G-S-K-N-K-P-K-P’, as conserved residues in C-terminal AT-hook motifs, were also exclusively present in Type-I (11, 100%). Notably, both conserved amino acid residues at C- and N-terminals of the AT-hook motifs share the same core, ‘R-G-R-P’, indicating their role in binding to the minor groove of AT-rich B-form DNA for regulatory functions [62]. While Motif1, Motif2, Motif4, and Motif6 were universally conserved, Motif7 appeared to be specific to Type-I, suggesting both functional conservation and type-specific specialization within the BpAHL protein family.

3.4. Chromosome Locations of BpAHL Gene Family

Genomic mapping revealed that 20 of the 21 identified BpAHL genes are distributed across nine birch chromosomes (1, 2, 3, 4, 5, 8, 9, 13, and 14), with chromosome 14 harboring the highest number (five genes) followed by chromosome 5 (four genes) (Figure 4A). The remaining chromosomes showed variable distributions: chromosomes 2, 4, 9, and 13 each contained two BpAHL genes, while chromosomes 1, 3, and 8 each possessed a single copy. The exceptional BpAHL21 was localized to an unanchored genomic fragment (Contig1465), suggesting it may reside in an unassembled chromosomal region.

3.5. Gene Functions of BpAHL Gene Family

To elucidate the biological functions of BpAHLs, we conducted gene ontology (GO) enrichment analysis across three aspects: biological process (BP), cellular component (CC), and molecular function (MF) (Figure 4B, Table S6). In terms of the MF, BpAHLs exhibited significant enrichment in DNA binding (GO:0003677, p-value = 3.5 × 10−11) and adenine-thymine-rich DNA binding (GO:0003680, p-value = 3.2 × 10−10), reinforcing their role in regulating AT-rich DNA sequences. Regarding BPs, BpAHLs were primarily engaged in metabolic and biosynthetic processes, including the brassinosteroid metabolic process (GO:0016132, p-value = 4.58 × 10−6). Considering CCs, BpAHLs were primarily localized in the nucleus (GO:0005730, p-value = 2.2 × 10−3) and nucleoplasm (GO:0005654, p-value = 1.6 × 10−4), suggesting their involvement in nuclear activities. Collectively, the systematic GO enrichment offers a comprehensive understanding of genes’ biological roles, facilitating further exploration of their mechanisms and potential significance in organisms.

3.6. Evolutionary Analysis of BpAHL Gene Family

Gene duplication events, arising from region-specific duplication or polyploidization, play a crucial role in gene family expansion and plant genome evolution [63,64]. Here, we investigated duplication patterns in the BpAHL gene family and found that ~81% (17 members) originated from dispersed duplication, while ~19% (4) resulted from segmental duplication (Figure 4C). Surprisingly, no tandem duplications were detected among the BpAHLs (Table S7), suggesting that BpAHLs have never undergone tandem duplication during birch long-term evolution.
Furthermore, we computed Ka/Ks ratios for those segmental duplication gene pairs, finding that BpAHL11/BpAHL14 and BpAHL09/BpAHL18 exhibited ratios of 0.33 and 0.27, respectively, both significantly below 1, indicating strong purifying selection during their evolutionary history. The estimated divergence times for these pairs were ~48.5 and ~36.4 million years ago (MYA), aligning with the evolutionary timeline of Populus trichocarpa AHLs (1.36–58.62 MYA) but substantially later than most AHLs in rice (13.82–173.89 MYA) [11,65].
To investigate evolutionary relationships, we assessed collinearity between BpAHLs and AHL proteins from two monocots (rice, maize) and three dicots (A. thaliana, soybean, Populus) (Figure 4D). Notably, 8 (~38%), 7 (~33%), 12 (~57%), 15 (~71%), and 16 (~76%) BpAHL members showed high homology with the AHLs from these species, respectively. These results support earlier findings in birch [60] and highlight the evolutionary divergence between monocot and dicot AHLs.

3.7. Cis-Acting Regulatory Elements in Promoters of BpAHL Gene Family

In organisms, promoter regions located upstream of the genes that bind to transcription factors (TFs) are called cis-acting regulatory elements (CREs) and encode blueprints for proper spatiotemporal patterning of gene expression and responses to environmental stimuli [66,67]. Here, we predicted ten classes of CREs in BpAHL promoters, associated with hormones (gibberellin, GA; methyl jasmonate, MeJA; abscisic acid, ABA; auxin; and salicylic acid, SA) and abiotic stress responses (light response; anaerobic induction; low-temperature response; and drought response) (Figure 5A, Table S8). The predicted CREs were present in 7 to 21 BpAHL genes, with each promoter containing three to eight distinct CREs (Figure 5B–D). The abiotic stress-related CREs, particularly light-responsive and anaerobic induction CREs, showed the broadest distribution, appearing in 21 and 19 BpAHLs, respectively. Hormone-responsive CREs (ABA, GA, and MeJA) were detected in 11 to 13 BpAHLs, while SA-responsive and low-temperature-responsive CREs exhibited more limited distributions, present in 7–8 BpAHLs. Notably, BpAHL05, BpAHL08, and BpAHL17 contained the most comprehensive set of CREs, with all eight types equally represented between hormone (four classes) and abiotic stress (four classes) responses (Figure 5B–D). These findings suggest that diverse CREs enable BpAHL genes to potentially respond to various environmental and hormonal signals.

3.8. Transcriptome Profiles of BpAHL Gene Family in Different Plant Tissues

To gain insights into potential functions of BpAHL genes in plant growth and development, we analyzed their gene expression patterns in four tissues of two-year-old birch using previously reported transcriptome data [52]. The 21 BpAHL genes showed predominant expression in flowers, leaves, and roots, displaying two distinct expression profiles (Figure 6A).
Specifically, the nine BpAHL genes (BpAHL11, BpAHL20, BpAHL03, BpAHL08, BpAHL13, BpAHL14, BpAHL18, BpAHL05, and BpAHL15) exhibited significantly higher expression in flowers and leaves compared to roots and xylem. Notably, ~78% (7/9) of these belonged to Type-II/III BpAHLs (Figure 6A, Table S9). In contrast, the remaining 11 genes (BpAHL02, BpAHL16, BpAHL12, BpAHL06, BpAHL19, BpAHL21, BpAHL07, BpAHL04, BpAHL10, BpAHL01, and BpAHL17) showed preferential root expression, with ~81% (9/11) classified as Type-I. In addition, the expression patterns were further validated by qRT-PCR experiments of randomly selected BpAHLs (Figure 6B). The distinct demarcation in expression patterns between BpAHL types (Type-I versus Type-II/III) across various tissues—the belowground versus the aboveground—implies that different types of BpAHLs may fulfill diverse biological functions in the tissues of birch.

3.9. Transcriptome Analysis of BpAHL Gene Family in Response to Abiotic Stress

To investigate the responses of the BpAHL gene family to various abiotic stresses, we performed a transcriptome analysis based on a series of artificial treatments guided by predicted CREs in promoters of BpAHLs (Figure 5). The treatments included temperature extremes (6 °C and 35 °C), salt stress (0.2 M NaCl for 24 h), and light/dark cycles (4 days). We systematically examined expression patterns of all 21 BpAHL genes under each condition (Figure 7).

3.9.1. Type-I BpAHLs Show a Limited Cold Response, While Type-II/III Show a Dynamic Response

Under low-temperature (6 °C) conditions, nearly all (10/11) Type-I BpAHLs (BpAHL01/03/04/06/10/12/15/17/19/21) exhibited low expression levels and no obvious responses in leaves (Table S10). Contrastingly, five Type-II (BpAHL09/11/13/18/20) and four Type-III (BpAHL02/05/08/16) BpAHLs, along with Type-I BpAHL07, displayed dynamic responses, revealing two distinct expression patterns (Figure 7A, Table S10). Specifically, six BpAHLs of Type-II/III (BpAHL02/08/11/13/18/20) showed progressive down-regulation during cold exposure (Figure 7A); conversely, four genes (BpAHL05/07/09/16) peaked at 1 h treatment. Extended cold exposure revealed further complexity: BpAHL18/20 showed early induction in the first 4 days, and BpAHL05/13 peaked on days 1–2, while BpAHL14/15 exhibited delayed responses, with BpAHL14 peaking on day 4 and BpAHL15 declining progressively (Figure 7B, Table S11).

3.9.2. Almost BpAHLs Show Down-Regulation Under Heat Stress on the First Day

Under high-temperature (35 °C) conditions, the expression patterns of BpAHLs were strikingly similar (Figure 7C, Table S12). With the exception of BpAHL12, BpAHL05, and BpAHL07, the expression levels of the other 10 BpAHLs dropped to their lowest values on the first day before gradually rebounding. Collectively, these results indicate that primarily Type-II/III BpAHLs respond to both low- and high-temperature stresses in leaves, with their varying expression patterns suggesting different biological functions in temperature stress responses.

3.9.3. Almost BpAHLs Down-Regulated Under Salt Stress in Root Tissue

Only 24% of BpAHLs (five members) responded to salt stress in leaves, predominantly Type-II/III (BpAHL05/08/09/11) and Type-I BpAHL07 (Figure 7D, Table S13). Two BpAHLs were up-regulated while three showed down-regulation, indicating specialized roles in leaf salt tolerance. In roots, 71% of BpAHLs (15 members) responded to salt stress, with 80% (12 members) showing significant down-regulation (Figure 7E, Table S14). This included six Type-I (BpAHL01/06/07/12/19/21) and six Type-II/III (BpAHL02/05/08/09/18) genes, revealing tissue-specific regulatory networks for salt adaptation.

3.9.4. Only Type-II/III BpAHLs Respond to Light/Dark Cycles

Approximately 24% of BpAHLs (all Type-II/III: BpAHL02/05/08/09/20) responded significantly to light/dark cycles in leaves (Figure 7F, Table S15), suggesting their specialized role in photoperiod-related processes.

3.10. Expression Analysis of BpAHL Genes Under Drought Treatment

Previous research has established that AHL family proteins (AT-hook motif nuclear localized proteins containing ankyrin repeats) play crucial regulatory roles in Oryza sativa under drought stress responses [25]. In our drought stress experiments with birch (Betula platyphylla), visible phenotypic alterations were observed (Figure 8A). Following 5–6 days of progressive drought treatment, stressed individuals exhibited marked growth retardation and wilting symptoms compared to well-watered controls, although the phenotypic differences remained relatively moderate during the initial period. Intriguingly, qRT-PCR experiments revealed a significant up-regulation of gene expression in root tissues under drought stresses (Figure 8B), suggesting their potential involvement in coordinating root-specific adaptive responses to these contrasting abiotic stresses.

4. Discussion

The AT-hook motif nuclear localized (AHL) gene family has been well documented to play pivotal roles in transcriptional regulation, plant growth and development, and stress responses [15,17,68]. Extensive systematic research has been conducted on this gene family across a variety of terrestrial plant species, encompassing dicotyledonous herbs such as A. thaliana (29 members) [4], Glycine max (63) [7], Daucus carota (47) [9], Brassica rapa (42) [10], and Gossypium species (48–99) [13]; monocotyledonous herbs like Oryza sativa (26) [5], and Zea mays (37) [6]; and woody plants including Vitis vinifera (14) [8], Populus trichocarpa (37) [11], and Liriodendron chinense (21) [12]. Here, we identified 21 AHL members in birch (Figure 1 and Table 1), representing the second-lowest number among characterized species, with only grape having a lower number. Notably, the number of AHL members is generally lower in woody plants compared to herbs, suggesting that the AHL gene family might fulfill indispensable functions throughout the entire lifecycle of trees.
Phylogenetic analysis classified the BpAHL proteins into two major clades (Clade-A and Clade-B) based on their PPC domain sequences—Clade-A harbors the conserved ‘L-R-S-H’ motif, while Clade-B contains the ‘F-T-P-H’ motif. Further subdivision based on AT-hook motifs categorized the members into three types: Type-I (with the ‘R-G-R-P’ motif at the N-terminus), Type-II (with ‘R-G-R-P’ at the C-terminus), and Type-III (lacking this conserved C-terminal residues) (Figure 1). These two functional domains are essential for nuclear localization and transcriptional regulation of AHL proteins [15,37,69,70]. Notably, mutations in the AT-hook motif could impair chromatin-binding activity [71].
Notably, distinct differences were present in sequence characteristics and physicochemical properties between Type-I and Type-II/III BpAHLs. Specifically, except for BpAHL07, all other Type-I BpAHLs are shorter than 1.0 kb, whereas Type-II/III BpAHLs exceed 5.0 kb in length (Figure 3A and Table 1). However, the variation in CDS lengths—particularly protein lengths—is relatively minor. This observation suggests that, although Type-II/III BpAHLs contain significantly more CDSs than Type-I BpAHLs, their translated proteins remain similar in size. This phenomenon may result from evolutionary processes such as genome duplication, transposon insertion, non-neutral selection, and co-evolving residues [72]. Additionally, these findings support the hypothesis that Type-II/III BpAHLs originated from Type-I BpAHLs and may employ a greater number of introns to generate novel splice variants, enabling precise regulation of gene function in response to specific developmental stages and tissue types [4,73].
Comparative analysis of molecular weight (MW), isoelectric point (pI), and subcellular localization among the BpAHL proteins revealed distinct patterns (Table 1). Type-I BpAHLs exhibited smaller molecular weights (24,150–36,878 Da; average ~29,845 Da) compared to Type-II (33,737–37,807 Da; average ~36,030 Da) and Type-III (25,202–52,128 Da; average ~37,173 Da). Similarly, Type-I proteins showed slightly lower pI values (4.73–8.97; average ~6.80) than Type-II (5.68–10.25; average ~9.00) and Type-III (5.64–9.73; average ~8.30), suggesting potential functional specialization in different micro-environments. Subcellular localization predictions demonstrated that all Type-I BpAHLs (11 proteins, 100%) were nuclear-localized, while half of Type-II/III BpAHLs (5 members) showed additional chloroplast localization. This finding raises the intriguing possibility that some BpAHL proteins may exert regulatory functions through chloroplast DNA binding [12]. Together, these results reinforce the relative conservation of Type-I BpAHL properties, in contrast to the greater variability observed in Type-II/III members.
Chromosomal distribution analysis revealed that these 21 BpAHL genes are relatively evenly distributed across birch’s nine chromosomes and one contig (Figure 4A), showing no significant correlation with chromosome length (Spearman correlation, p-value > 0.05). Gene duplication analysis indicated that most BpAHLs (17/21) arose from dispersed duplication events (Figure 4C), a pattern similar to maize but distinct from soybean [6,7]. This difference likely reflects soybean’s status as an ancient tetraploid [74], in contrast to birch, which lacks evidence of whole-genome duplication (WGD) in its evolutionary history [34]. This evolutionary distinction might also account for the relatively small BpAHL gene family in birch. Comparative synteny analysis demonstrated strong conservation between birch (a dicot) and other dicots (Glycine max, Populus trichocarpa, and A. thaliana), but weaker synteny with monocots (Oryza sativa and Zea mays) (Figure 4D). This divergence may result from mutated Type-I BpAHLs in monocots [6], highlighting an important evolutionary differentiation in the AHL gene family between monocotyledonous and dicotyledonous plants.
The AHL gene family displayed conserved tissue-specific expression patterns across various plant species, with soybean AHLs predominantly expressed in roots, meristems and epicotyls [7], cabbage AHLs in roots and buds [10], and maize AHLs in roots, embryos, endosperm and seeds [6]. Here, this study reveals that birch AHLs exhibited similar type-dependent tissue specificity: Type-I BpAHLs show root-predominant expression (below-ground tissues), while Type-II/III members exhibit preferential expression in flowers and leaves (above-ground tissues) (Figure 6A and Table S9), a pattern confirmed by qRT-PCR experiments of BpAHL06, BpAHL21 and BpAHL09 (Figure 6B). This expression dichotomy suggests functional specialization, with Type-I genes likely involved in root development and nutrient uptake (consistent with Arabidopsis AtAHL18/AtAHL29 [75,76]), whereas Type-II/III genes might participate in aerial tissue processes like photosynthesis, flowering or defense mechanisms [4,5,6].
The CREs within gene promoters play crucial roles in modulating transcriptional networks governing abiotic stimulus responses, hormone signaling, and developmental processes [77]. Previous research has demonstrated this regulatory function in AHL family members: AtAHL15 and other Clade-A proteins directly regulate downstream flowering-related genes like SOC1 (SUPPRESSOR OF OVEREXPRESSION OF CONSTANS 1) and FUL (FRUITFULL), and upstream GA signaling [19]; SlAHL1 and SlAHL7 in tomato show dramatic responses to ABA, salt, and drought stresses [78]; while BrAHL02 and BrAHL24 in Brassica rapa exhibit rapid up-regulation under cold and cadmium stress, peaking at 12 h [10]. Consistent with these results, our analysis revealed that BpAHL promoters are enriched in CREs associated with hormonal responses (MeJA, ABA, and SA) and environmental stresses (light, temperature, and drought) (Figure 5), suggesting conserved regulatory mechanisms across plant species.
To validate this hypothesis, we conducted transcriptomic analyses of birch subjected to various abiotic stresses (Figure 7). The results revealed distinct expression dynamics between BpAHL genes’ types: (a) Under short-term low-temperature stress, all Type-II/III BpAHLs showed significant down-regulation, either immediately or within 1.0 h (Figure 7A). (b) During prolonged high-temperature exposure, all Type-II/III members (except BpAHL05) exhibited marked suppression, reaching minimal expression at 24 h before partial recovery (Figure 7C). (c) Salt stress induced consistent gene expression patterns for BpAHL05/07/08/09/11 in both leaves and roots (Figure 7D,E), while most Type-I members (BpAHL01/06/07/12/19/21) were significantly down-regulated in roots. These type-specific and stress-dependent expression profiles suggest functional diversification of BpAHLs in environmental stress responses, with different members potentially participating in distinct regulatory pathways within the same tissue.

5. Conclusions

In summary, our study systematically identified and characterized 21 AT-hook motif nuclear localized (AHL) genes in birch (BpAHLs), all containing a PPC domain and at least one AT-hook motif. These genes are evenly spanned across birch chromosomes and primarily originated through dispersed duplication events. Phylogenetic analysis classified them into two clades (Clade-A/B) or three types (Type-I/II/III), with distinct structural features and gene expression patterns. Type-I BpAHLs (single CDSs) show root-predominant expression and evolutionary conservation, while Type-II/III members (multiple CDSs) exhibit aerial tissue-specific expression and pronounced responsiveness to diverse environmental stresses (temperature, salinity, and light), suggesting their involvement in more complex regulatory networks. These findings provide valuable genomic resources and molecular insights for birch tree improvement. Importantly, the stress-responsive BpAHLs, particularly members in Type-II/III, represent promising candidates for molecular breeding and genetic engineering applications. Their manipulation through overexpression or genome editing could potentially enhance birch adaptability to abiotic stresses, while their tissue-specific promoters may enable targeted modulation of stress tolerance mechanisms in different organs. This study thus establishes a foundation for developing stress-resilient birch varieties through biotechnological approaches.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/f16060943/s1. Figure S1: Tertiary structure of the 21 BpAHL proteins; Figure S2: Multiple sequence alignments of Type-I AHL proteins between birch and Arabidopsis thaliana; Figure S3: Multiple sequence alignments of Type-II AHL proteins between birch and Arabidopsis thaliana; Figure S4: Multiple sequence alignments of Type-III AHL proteins between birch and Arabidopsis thaliana; Figure S5: Annotation of MEME motifs for AHL proteins in birch; Table S1: The Newick file obtained from MEGA software for constructing the phylogenetic tree; Table S2: The AHL protein sequences for constructing the phylogenetic tree; Table S3: Primer sequences for qRT-PCR experiments; Table S4: Comprehensive tertiary structure information for the 21 BpAHL proteins; Table S5: E-values, sites, and widths of the conserved motifs in the BpAHL proteins; Table S6: GO annotation of the BpAHL gene family; Table S7: The classification of duplication events in the BpAHL gene family; Table S8: Plantcare predicted Cis-acting element results package; Table S9: Expression levels (TPM) of BpAHLs in different plant tissues of birch; Table S10: Expression levels (TPM) of BpAHLs under short-term (from 0 to 3 h) low-temperature (6 °C) stress in birch leaf; Table S11: Expression levels (TPM) of BpAHLs under long-term (from 0 to 14 days) low-temperature (6 °C) stress in birch leaf; Table S12: Expression levels (TPM) of BpAHLs under long-term (from 0 to 14 days) high-temperature (35 °C) stress in birch leaf; Table S13: Expression levels (TPM) of BpAHLs under salt stress (0.2 M NaCl for 24 h) in birch leaf; Table S14: Expression levels (TPM) of BpAHLs under salt stress (0.2 M NaCl for 24 h) in birch root; Table S15: Expression levels (TPM) of BpAHLs under dark and light conditions.

Author Contributions

Data curation, B.C. and M.W.; formal analysis, B.C.; funding acquisition, L.X.; investigation, H.C.; methodology, B.C.; project administration, B.C.; resources, H.C.; software, B.C., H.C., B.L., Z.Z., T.Z. and W.Z.; supervision, L.X.; validation, H.C., B.L., Y.G., Q.X. and M.H.; visualization, B.C., H.C., B.L., Z.Z. and T.Z.; writing—original draft, B.C., S.A., L.Z., Z.J. and M.W.; writing—review and editing, B.C. and M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Key R&D Program of China during the 14th Five-year Plan Period (2021YFD2200103 and 2021YFD2200304), the Heilongjiang Provincial Natural Science Foundation of China (LH2024C047), Fundamental Research Funds for the Central Universities (2572024DP24), and the Scientific Research Foundation of Hunan Provincial Education Department (24B0482).

Data Availability Statement

Data supporting the findings of this study are available from the corresponding author upon request. All sequencing data are deposited in NCBI Sequence Read Archive (SRA) and are accessible under the following accession numbers: PRJNA535361, PRJNA532995, PRJNA811313, PRJNA790472, and PRJNA759706.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yoon, Y.; Seo, D.H.; Shin, H.; Kim, H.J.; Kim, C.M.; Jang, G. The Role of Stress-Responsive Transcription Factors in Modulating Abiotic Stress Tolerance in Plants. Agronomy 2020, 10, 788. [Google Scholar] [CrossRef]
  2. Zhang, W.M.; Cheng, X.Z.; Fang, D.; Cao, J. AT-Hook Motif Nuclear Localized (AHL) Proteins of Ancient Origin Radiate New Functions. Int. J. Biol. Macromol. 2022, 214, 290–300. [Google Scholar] [CrossRef] [PubMed]
  3. Goodwin, G.H.; Sanders, C.; Johns, E.W. A New Group of Chromatin-Associated Proteins with a High Content of Acidic and Basic Amino Acids. Eur. J. Biochem. 1973, 38, 14–19. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, J.; Favero, D.S.; Qiu, J.; Roalson, E.H.; Neff, M.M. Insights into the Evolution and Diversification of the AT-Hook Motif Nuclear Localized Gene Family in Land Plants. BMC Plant Biol. 2014, 14, 266. [Google Scholar] [CrossRef]
  5. Kumar, A.; Singh, S.; Mishra, A. Genome-Wide Identification and Analyses of the AHL Gene Family in Rice (Oryza sativa). 3 Biotech 2023, 13, 248. [Google Scholar] [CrossRef]
  6. Bishop, E.H.; Kumar, R.; Luo, F.; Saski, C.; Sekhon, R.S. Genome-Wide Identification, Expression Profiling, and Network Analysis of AT-Hook Gene Family in Maize. Genomics 2020, 112, 1233–1244. [Google Scholar] [CrossRef]
  7. Wang, M.; Chen, B.; Zhou, W.; Xie, L.; Wang, L.; Zhang, Y.; Zhang, Q. Genome-Wide Identification and Expression Analysis of the AT-Hook Motif Nuclear Localized Gene Family in Soybean. BMC Genom. 2021, 22, 361. [Google Scholar] [CrossRef]
  8. Li, X.; He, H.; Wang, H.; Wu, X.; Wang, H.; Mao, J. Identification and Expression Analysis of the AHL Gene Family in Grape (Vitis vinifera). Plant Gene 2021, 26, 100285. [Google Scholar] [CrossRef]
  9. Machaj, G.; Grzebelus, D. Characteristics of the AT-Hook Motif Containing Nuclear Localized (AHL) Genes in Carrot Provides Insight into Their Role in Plant Growth and Storage Root Development. Genes 2021, 12, 764. [Google Scholar] [CrossRef]
  10. Zhang, X.; Li, J.; Cao, Y.; Huang, J.; Duan, Q. Genome-Wide Identification and Expression Analysis under Abiotic Stress of BrAHL Genes in Brassica rapa. Int. J. Mol. Sci. 2023, 24, 12447. [Google Scholar] [CrossRef]
  11. Wang, H.; Leng, X.; Yang, J.; Zhang, M.; Zeng, M.; Xu, X.; Wang, F.; Li, C. Comprehensive Analysis of AHL Gene Family and Their Expression under Drought Stress and ABA Treatment in Populus trichocarpa. PeerJ 2021, 9, e10932. [Google Scholar] [CrossRef] [PubMed]
  12. Tang, Y.; Wu, W.; Zheng, X.; Lu, L.; Chen, X.; Hao, Z.; Liu, S.; Chen, Y. AT-Hook Transcription Factors Show Functions in Liriodendron chinense under Drought Stress and Somatic Embryogenesis. Plants 2023, 12, 1353. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, L.; Lü, Y.; Chen, W.; Yao, J.; Li, Y.; Li, Q.; Pan, J.; Fang, S.; Sun, J.; Zhang, Y. Genome-Wide Identification and Analyses of the AHL Gene Family in Cotton (Gossypium). BMC Genom. 2020, 21, 69. [Google Scholar] [CrossRef] [PubMed]
  14. Rayapuram, N.; Jarad, M.; Alhoraibi, H.M.; Bigeard, J.; Abulfaraj, A.A.; Völz, R.; Mariappan, K.G.; Almeida, M.; Schlöffel, M.; Lastrucci, E.; et al. Chromatin Phosphoproteomics Unravels A Function for AT-Hook Motif Nuclear Localized Protein AHL13 in PAMP-Triggered Immunity. Proc. Natl. Acad. Sci. USA 2021, 118, e2004670118. [Google Scholar] [CrossRef]
  15. Huth, J.R.; Bewley, C.A.; Nissen, M.S.; Evans, J.; Reeves, R.; Gronenborn, A.M.; Clore, G.M. The Solution Structure of An HMG-I(Y)–DNA Complex Defines A New Architectural Minor Groove Binding Motif. Nat. Struct. Biol. 1997, 4, 657–665. [Google Scholar] [CrossRef]
  16. Aravind, L.; Landsman, D. AT-Hook Motifs Identified in A Wide Variety of DNA-Binding Proteins. Nucleic Acids Res. 1998, 26, 4413–4421. [Google Scholar] [CrossRef]
  17. Zhao, J.; Favero, D.S.; Peng, H.; Neff, M.M. Arabidopsis thaliana AHL Family Modulates Hypocotyl Growth Redundantly by Interacting with Each Other Via the PPC/DUF296 Domain. Proc. Natl. Acad. Sci. USA 2013, 110, 4688–4697. [Google Scholar] [CrossRef]
  18. Yun, J.; Kim, Y.S.; Jung, J.H.; Seo, P.J.; Park, C.M. The AT-Hook Motif-Containing Protein AHL22 Regulates Flowering Initiation by Modifying FLOWERING LOCUS T Chromatin in Arabidopsis. J. Biol. Chem. 2012, 287, 15307–15316. [Google Scholar] [CrossRef]
  19. Karami, O.; Rahimi, A.; Khan, M.; Bemer, M.; Hazarika, R.R.; Mak, P.; Compier, M.; Noort, V.; Offringa, R. A Suppressor of Axillary Meristem Maturation Promotes Longevity in Flowering Plants. Nat. Plants 2020, 6, 368–376. [Google Scholar] [CrossRef]
  20. Gallavotti, A.; Malcomber, S.; Gaines, C.; Stanfield, S.; Whipple, C.; Kellogg, E.; Schmidt, R.J. BARREN STALK FASTIGIATE1 Is An AT-Hook Protein Required for the Formation of Maize Ears. Plant Cell 2011, 23, 1756–1771. [Google Scholar] [CrossRef]
  21. Delaney, S.K.; Orford, S.J.; Martin-Harris, M.; Timmis, J.N. The Fiber Specificity of the Cotton FSltp4 Gene Promoter is Regulated by An AT-Rich Promoter Region and the AT-Hook Transcription Factor GhAT1. Plant Cell Physiol. 2007, 48, 1426–1437. [Google Scholar] [CrossRef] [PubMed]
  22. Reisdorf-Cren, M.; Carrayol, E.; Tercé-Laforgue, T.; Hirel, B. A Novel HMG A-Like Protein Binds Differentially to the AT-Rich Regions Located in the Far Distal and Proximal Parts of a Soybean Glutamine Synthetase Gene (GS15) Promoter. Plant Cell Physiol. 2002, 43, 1006–1016. [Google Scholar] [CrossRef] [PubMed]
  23. Lu, H.; Zou, Y.; Feng, N. Overexpression of AHL20 Negatively Regulates Defenses in Arabidopsis. J. Integr. Plant Biol. 2010, 52, 801–808. [Google Scholar] [CrossRef] [PubMed]
  24. Kim, S.Y.; Kim, Y.C.; Seong, E.S.; Lee, Y.H.; Park, J.M.; Choi, D. The Chili Pepper CaATL1: An AT-Hook Motif-Containing Transcription Factor Implicated in Defense Responses Against Pathogens. Mol. Plant Pathol. 2007, 8, 761–771. [Google Scholar] [CrossRef]
  25. Zhou, L.; Liu, Z.; Liu, Y.; Kong, D.; Li, T.; Yu, S.; Mei, H.; Xu, X.; Liu, H.; Chen, L.; et al. A Novel Gene OsAHL1 Improves Both Drought Avoidance and Drought Tolerance in Rice. Sci. Rep. 2016, 6, 30264. [Google Scholar] [CrossRef]
  26. Geng, W.; Li, Y.; Sun, D.; Li, B.; Zhang, P.; Chang, H.; Rong, T.; Liu, Y.; Shao, J.; Liu, Z.; et al. Prediction of the Potential Geographical Distribution of Betula platyphylla Suk. in China under Climate Change Scenarios. PLoS ONE 2022, 17, e0262540. [Google Scholar] [CrossRef]
  27. Isidorov, V.; Szoka, Ł.; Nazaruk, J. Cytotoxicity of White Birch Bud Extracts: Perspectives for Therapy of Tumours. PLoS ONE 2018, 13, e0201949. [Google Scholar] [CrossRef]
  28. Wang, J.; Chen, B.; Ali, S.; Zhang, T.; Wang, Y.; Zhang, H.; Wang, L.; Zhang, Y.; Xie, L.; Jiang, T.; et al. Epigenetic Modification Associated with Climate Regulates Betulin Biosynthesis in Birch. J. For. Res. 2023, 34, 21–35. [Google Scholar] [CrossRef]
  29. Liu, Y.; Xu, H.W.; Shang, F.Q.; Jiao, H.; Zhang, L.M.; Luo, J.X.; Teng, W.H.; Liu, G.F. Variation and Zoning of 16-Year-Old Betula platyphylla Provenance. Sci. Silvae Sin. 2016, 52, 48–56. [Google Scholar]
  30. Liu, Z.J.; Shi, X.X.; Wang, Z.B.; Qu, M.; Gao, C.Q.; Wang, C.; Wang, Y.C. Acetylation of Transcription Factor BpTCP20 by Acetyltransferase BpPDCE23 Modulates Salt Tolerance in Birch. Plant Physiol. 2024, 195, 2354–2371. [Google Scholar] [CrossRef]
  31. Wen, X.J.; Ding, Y.; Tan, Z.L.; Wang, J.X.; Zhang, D.Y.; Wang, Y.C. Identification and Characterization of Cadmium Stress-Related LncRNAs from Betula platyphylla. Plant Sci. 2020, 299, 110601. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, X.; Yu, J.; Qu, G.; Chen, S. The Cold-Responsive C-repeat Binding Factors in Betula platyphylla Suk. Positively Regulate Cold Tolerance. Plant Sci. 2024, 341, 112012. [Google Scholar] [CrossRef] [PubMed]
  33. Cheng, D.; Liu, Y.; Wang, Y.; Cao, L.; Wu, S.; Yu, S.; Xie, L.; Li, H.; Jiang, J.; Liu, G.; et al. Establishment of High-Efficiency Genome Editing in White Birch (Betula platyphylla Suk.). Plant Biotechnol. J. 2024, 22, 7–9. [Google Scholar] [CrossRef] [PubMed]
  34. Salojärvi, J.; Smolander, O.P.; Nieminen, K.; Rajaraman, S.; Safronov, O.; Safdari, P.; Lamminmäki, A.; Immanen, J.; Lan, T.; Tanskanen, J.; et al. Genome Sequencing and Population Genomic Analyses Provide Insights into the Adaptive Landscape of Silver Birch. Nat. Genet. 2017, 49, 904–912. [Google Scholar] [CrossRef]
  35. 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]
  36. Lin, L.; Nakano, H.; Nakamura, S.; Uchiyama, S.; Fujimoto, S.; Matsunaga, S.; Kobayashi, Y.; Ohkubo, T.; Fukui, K. Crystal Structure of Pyrococcus horikoshii PPC Protein at 1.60 A Resolution. Proteins 2007, 67, 505–507. [Google Scholar] [CrossRef]
  37. Fujimoto, S.; Matsunaga, S.; Yonemura, M.; Uchiyama, S.; Azuma, T.; Fukui, K. Identification of a Novel Plant MAR DNA Binding Protein Localized on Chromosomal Surfaces. Plant Mol. Biol. 2004, 56, 225–239. [Google Scholar] [CrossRef]
  38. Yang, M.; Derbyshire, M.K.; Yamashita, R.A.; Marchler-Bauer, A. NCBI’s Conserved Domain Database and Tools for Protein Domain Analysis. Curr. Protoc. Bioinform. 2020, 69, e90. [Google Scholar] [CrossRef]
  39. Schultz, J.; Milpetz, F.; Bork, P.; Ponting, C.P. SMART, A Simple Modular Architecture Research Tool: Identification of Signaling Domains. Proc. Natl. Acad. Sci. USA 1998, 95, 5857–5864. [Google Scholar] [CrossRef]
  40. Gasteiger, E.; Gattiker, A.; Hoogland, C.; Ivanyi, I.; Appel, R.D.; Bairoch, A. ExPASy: The Proteomics Server for In-Depth Protein Knowledge and Analysis. Nucleic Acids Res. 2003, 31, 3784–3788. [Google Scholar] [CrossRef]
  41. Chou, K.C.; Shen, H.B. Cell-PLoc: A Package of Web Servers for Predicting Subcellular Localization of Proteins in Various Organisms. Nat. Protoc. 2008, 3, 153–162. [Google Scholar] [CrossRef] [PubMed]
  42. Waterhouse, A.; Bertoni, M.; Bienert, S.; Studer, G.; Tauriello, G.; Gumienny, R.; Heer, F.T.; Beer, T.; Rempfer, C.; Bordoli, L.; et al. SWISS-MODEL: Homology Modelling of Protein Structures and Complexes. Nucleic Acids Res. 2018, 46, 296–303. [Google Scholar] [CrossRef] [PubMed]
  43. Johansson, M.U.; Zoete, V.; Michielin, O.; Guex, N. Defining and Searching for Structural Motifs Using DeepView/Swiss-PdbViewer. BMC Bioinf. 2012, 13, 173. [Google Scholar] [CrossRef] [PubMed]
  44. Klopfenstein, D.V.; Zhang, L.; Pedersen, B.S.; Ramírez, F.; Warwick, V.A.; Naldi, A.; Mungall, C.J.; Yunes, J.M.; Botvinnik, O.; Weigel, M.; et al. GOATOOLS: A Python library for Gene Ontology Analyses. Sci. Rep. 2018, 8, 10872. [Google Scholar] [CrossRef]
  45. Tamura, K.; Stecher, G.; Kumar, S. MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol. Biol. Evol. 2021, 38, 3022–3027. [Google Scholar] [CrossRef]
  46. Letunic, I.; Bork, P. Interactive Tree of Life (iTOL): An Online Tool for Phylogenetic Tree Display and Annotation. Bioinform. 2006, 23, 127–128. [Google Scholar] [CrossRef]
  47. 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]
  48. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. MEME Suite: Tools for Motif Discovery and Searching. Nucleic Acids Res. 2009, 37, 202–208. [Google Scholar] [CrossRef]
  49. Wang, Y.; Tang, H.; DeBarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.H.; Jin, H.; Marler, B.; Guo, H.; et al. MCScanX: A Toolkit for Detection and Evolutionary Analysis of Gene Synteny and Collinearity. Nucleic Acids Res. 2012, 40, e49. [Google Scholar] [CrossRef]
  50. Koch, M.A.; Haubold, B.; Mitchell, 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]
  51. Lescot, M.; Dehairs, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Peer, Y.; Rouzé, P.; Rombauts, S. PlantCARE, A Database of Plant Cis-Acting Regulatory Elements and A Portal to Tools for in Silico Analysis of Promoter Sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, S.; Lin, X.; Zhang, D.; Li, Q.; Zhao, X.; Chen, S. Genome-Wide Analysis of NAC Gene Family in Betula pendula. Forests 2019, 10, 741. [Google Scholar] [CrossRef]
  53. Kim, D.; Paggi, J.M.; Park, C.; Bennett, C.; Salzberg, S.L. Graph-Based Genome Alignment and Genotyping with HISAT2 and HISAT-Genotype. Nat. Biotechnol. 2019, 37, 907–915. [Google Scholar] [CrossRef] [PubMed]
  54. Pertea, M.; Pertea, G.M.; Antonescu, C.M.; Chang, T.C.; Mendell, J.T.; Salzberg, S.L. StringTie Enables Improved Reconstruction of A Transcriptome from RNA-Seq Reads. Nat. Biotechnol. 2015, 33, 290–295. [Google Scholar] [CrossRef]
  55. Yan, S.; Zhang, D.; Chen, S.; Chen, S. Transcriptome Sequencing Analysis of Birch (Betula platyphylla Sukaczev) under Low-Temperature Stress. Forests 2020, 11, 970. [Google Scholar] [CrossRef]
  56. Ritonga, F.N.; Chen, S.; Indriani, F.; Song, R.K.; Zhang, X.; Lan, X.; Chen, S. Comparative RNA-Seq Analysis of Betula platyphylla under Low and High Temperature Stresses. CERNE 2023, 29, e103147. [Google Scholar] [CrossRef]
  57. Jia, Y.; Zhao, H.; Niu, Y.; Wang, Y. Identification of Birch lncRNAs and mRNAs Responding to Salt Stress and Characterization of Functions of lncRNA. Hortic. Res. 2023, 10, uhac277. [Google Scholar] [CrossRef]
  58. Song, S.; Ge, M.; Wang, W.; Gu, C.; Chen, K.; Zhang, Q.; Yu, Q.; Liu, G.; Jiang, J. BpEIN3.1 Represses Leaf Senescence by Inhibiting Synthesis of Ethylene and Abscisic Acid in Betula platyphylla. Plant Sci. 2022, 321, 111330. [Google Scholar] [CrossRef]
  59. Skern, R.; Frost, P.; Nilsen, F. Relative Transcript Quantification by Quantitative PCR: Roughly Right or Precisely Wrong? BMC Mol. Biol. 2005, 6, 10. [Google Scholar] [CrossRef]
  60. Chen, B.; Ali, S.; Zhang, X.; Zhang, Y.; Wang, M.; Zhang, Q.; Xie, L. Genome-Wide Identification, Classification, and Expression Analysis of the JmjC Domain-Containing Histone Demethylase Gene Family in Birch. BMC Genom. 2021, 22, 772. [Google Scholar] [CrossRef]
  61. Chen, B.; Guo, Y.; Zhang, X.; Wang, L.; Cao, L.; Zhang, T.; Zhang, Z.; Zhou, W.; Xie, L.; Wang, J.; et al. Climate-Responsive DNA Methylation is Involved in the Biosynthesis of Lignin in Birch. Front. Plant Sci. 2022, 13, 1090967. [Google Scholar] [CrossRef] [PubMed]
  62. Lim, P.O.; Kim, Y.; Breeze, E.; Koo, J.C.; Woo, H.R.; Ryu, J.S.; Park, D.H.; Beynon, J.; Tabrett, A.; Buchanan, V.; et al. Overexpression of A Chromatin Architecture-Controlling AT-Hook Protein Extends Leaf Longevity and Increases the Post-Harvest Storage Life of Plants. Plant J. 2007, 52, 1140–1153. [Google Scholar] [CrossRef] [PubMed]
  63. Cannon, S.B.; Mitra, A.; Baumgarten, A.; Young, N.D.; May, G. The Roles of Segmental and Tandem Gene Duplication in the Evolution of Large Gene Families in Arabidopsis thaliana. BMC Plant Biol. 2004, 4, 10. [Google Scholar] [CrossRef] [PubMed]
  64. Mehan, M.R.; Freimer, N.B.; Ophoff, R.A. A Genome-Wide Survey of Segmental Duplications that Mediate Common Human Genetic Variation of Chromosomal Architecture. Hum. Genomics. 2004, 1, 335. [Google Scholar] [CrossRef]
  65. Ambadas, D.A.; Singh, A.; Jha, R.K.; Chauhan, D.; Sharma, V.K. Genome-Wide Dissection of AT-Hook Motif Nuclear-Localized Gene Family and Their Expression Profiling for Drought and Salt Stress in Rice (Oryza sativa). Front. Plant Sci. 2023, 14, 1283555. [Google Scholar] [CrossRef]
  66. Marand, A.P.; Eveland, A.L.; Kaufmann, K.; Springer, N.M. Cis-Regulatory Elements in Plant Development, Adaptation, and Evolution. Annu. Rev. Plant Biol. 2023, 74, 111–137. [Google Scholar] [CrossRef]
  67. Wittkopp, P.J.; Kalay, G. Cis-Regulatory Elements: Molecular Mechanisms and Evolutionary Processes Underlying Divergence. Nat. Rev. Genet. 2012, 13, 59–69. [Google Scholar] [CrossRef]
  68. Jin, Y.; Luo, Q.; Tong, H.; Wang, A.; Cheng, Z.; Tang, J.; Li, D.; Zhao, X.; Li, X.; Wan, J.; et al. An AT-Hook. Gene is Required for Palea Formation and Floral Organ Number Control in Rice. Dev. Biol. 2011, 359, 277–288. [Google Scholar] [CrossRef]
  69. Xiao, C.; Chen, F.; Yu, X.; Lin, C.; Fu, Y.F. Over-Expression of An AT-Hook Gene, AHL22, Delays Flowering and Inhibits the Elongation of the Hypocotyl in Arabidopsis thaliana. Plant Mol. Biol. 2009, 71, 39–50. [Google Scholar] [CrossRef]
  70. Sgarra, R.; Lee, J.; Tessari, M.A.; Altamura, S.; Spolaore, B.; Giancotti, V.; Bedford, M.T.; Manfioletti, G. The AT-Hook of the Chromatin Architectural Transcription Factor High Mobility Group A1a is Arginine-Methylated by Protein Arginine Methyltransferase 6. J. Biol. Chem. 2006, 281, 3764–3772. [Google Scholar] [CrossRef]
  71. Turlure, F.; Maertens, G.; Rahman, S.; Cherepanov, P.; Engelman, A. A Tripartite DNA-Binding Element, Comprised of the Nuclear Localization Signal and Two AT-Hook Motifs, Mediates the Association of LEDGF/p75 with Chromatin in vivo. Nucleic Acids Res. 2006, 34, 1653–1665. [Google Scholar] [CrossRef] [PubMed]
  72. Parakkunnel, R.; Bhojaraja, N.K.; Susmita, C.; Girimalla, V.; Bhaskar, K.U.; Sripathy, K.V.; Shantharaja, C.S.; Aravindan, S.; Kumar, S.; Lakhanpaul, S.; et al. Evolution and Co-Evolution: Insights into the Divergence of Plant Heat Shock Factor Genes. Physiol. Mol. Biol. Plants. 2022, 28, 1029–1047. [Google Scholar] [CrossRef] [PubMed]
  73. Gorlova, O.; Fedorov, A.; Logothetis, C.; Amos, C.; Gorlov, I. Genes With A Large Intronic Burden Show Greater Evolutionary Conservation on the Protein Level. BMC Evol. Biol. 2014, 14, 50. [Google Scholar] [CrossRef] [PubMed]
  74. Schmutz, J.; Cannon, S.B.; Schlueter, J.; Ma, J.; Mitros, T.; Nelson, W.; Hyten, D.L.; Song, Q.; Thelen, J.J.; Cheng, J.; et al. Genome Sequence of the Palaeopolyploid Soybean. Nature 2010, 463, 178–183. [Google Scholar] [CrossRef]
  75. Street, I.H.; Shah, P.K.; Smith, A.M.; Avery, N.; Neff, M.M. The AT-hook-Containing Proteins SOB3/AHL29 and ESC/AHL27 Are Negative Modulators of Hypocotyl Growth in Arabidopsis. Plant J. 2008, 54, 1–14. [Google Scholar] [CrossRef]
  76. Širl, M.; Šnajdrová, T.; Gutiérrez-Alanís, D.; Dubrovsky, J.G.; Vielle-Calzada, J.P.; Kulich, I.; Soukup, A. At-Hook Motif Nuclear Localized Protein 18 as A Novel Modulator of Root System Architecture. Int. J. Mol. Sci. 2020, 21, 1886. [Google Scholar] [CrossRef]
  77. Yamaguchi-Shinozaki, K.; Shinozaki, K. Organization of Cis-Acting Regulatory Elements in Osmotic- and Cold-Stress-Responsive Promoters. Trends Plant Sci. 2005, 10, 88–94. [Google Scholar] [CrossRef]
  78. Wang, L.; Li, T.; Liu, N.; Liu, X. Identification of Tomato AHL Gene Families and Functional Analysis Their Roles in Fruit Development and Abiotic Stress Response. Plant Physiol. Biochem. 2023, 202, 107931. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic analysis of the 21 BpAHL proteins. The neighbor-joining phylogenetic tree is depicted in the left panel, and the conserved amino acid residues of AT-hook motifs and the PPC domain are illustrated in the right panel.
Figure 1. Phylogenetic analysis of the 21 BpAHL proteins. The neighbor-joining phylogenetic tree is depicted in the left panel, and the conserved amino acid residues of AT-hook motifs and the PPC domain are illustrated in the right panel.
Forests 16 00943 g001
Figure 2. Phylogenetic analysis of AHL proteins across five species. The AHL proteins from Arabidopsis thaliana, Glycine max, Zea mays, and Sorghum bicolor are indicated in red, green, blue, and orange, respectively. The AHL proteins from birch are highlighted with a purple circle. Clade-A and Clade-B of the BpAHL proteins are annotated with green and brown curves, respectively, to denote their positions outside the phylogenetic tree.
Figure 2. Phylogenetic analysis of AHL proteins across five species. The AHL proteins from Arabidopsis thaliana, Glycine max, Zea mays, and Sorghum bicolor are indicated in red, green, blue, and orange, respectively. The AHL proteins from birch are highlighted with a purple circle. Clade-A and Clade-B of the BpAHL proteins are annotated with green and brown curves, respectively, to denote their positions outside the phylogenetic tree.
Forests 16 00943 g002
Figure 3. Schematic representations of the 21 BpAHLs. (A) Gene structures of BpAHLs. The x-axis represents the length of the DNA sequences (kb), with green squares indicating CDSs and black lines denoting introns. (B) Prediction of conserved motifs in BpAHL proteins. The x-axis reflects the length of the protein sequences, and different categories of conserved motifs are represented by squares in various colors. A total of ten of the most frequently occurring motifs are shown in this graph, with most proteins within the same clade exhibiting similar motif patterns.
Figure 3. Schematic representations of the 21 BpAHLs. (A) Gene structures of BpAHLs. The x-axis represents the length of the DNA sequences (kb), with green squares indicating CDSs and black lines denoting introns. (B) Prediction of conserved motifs in BpAHL proteins. The x-axis reflects the length of the protein sequences, and different categories of conserved motifs are represented by squares in various colors. A total of ten of the most frequently occurring motifs are shown in this graph, with most proteins within the same clade exhibiting similar motif patterns.
Forests 16 00943 g003
Figure 4. Chromosomal localization, gene ontology enrichments, gene duplication events, and collinearity analysis of the BpAHL gene family. (A) Genomic distribution of the BpAHLs across the 14 chromosomes of birch. (B) Gene ontology enrichment analysis of the BpAHLs. (C) Analysis of gene duplication events of the BpAHLs. (D) Collinearity analysis of AHLs among A. thaliana, soybean, rice, maize, Populus, and birch. The grey lines represent collinear blocks between birch and the other plants, while the red lines highlight syntenic AHL gene pairs.
Figure 4. Chromosomal localization, gene ontology enrichments, gene duplication events, and collinearity analysis of the BpAHL gene family. (A) Genomic distribution of the BpAHLs across the 14 chromosomes of birch. (B) Gene ontology enrichment analysis of the BpAHLs. (C) Analysis of gene duplication events of the BpAHLs. (D) Collinearity analysis of AHLs among A. thaliana, soybean, rice, maize, Populus, and birch. The grey lines represent collinear blocks between birch and the other plants, while the red lines highlight syntenic AHL gene pairs.
Forests 16 00943 g004
Figure 5. Cis-acting regulatory elements (CREs) of the 21 BpAHLs promoters. (A) Genomic DNA sequences of the 2.0 kb upstream promoters of the BpAHLs were submitted to the PlantCare website to identify CREs. These elements can be classified into two main categories: CREs associated with hormone responses, and CREs related to abiotic stress responses. The squares in various colors represent different classes of CREs. (B) Summary of the ten classes of CREs identified in the promoters of BpAHLs. The black dots with numbers indicate the specific BpAHLs that contain the corresponding number of CREs. (C) Number of CREs present in each BpAHL gene. The bars, from left to right, represent BpAHL01 to BpAHL21. (D) Number of BpAHLs in each class of CRE. The bars in different colors represent different CREs, with corresponding annotations at the bottom left.
Figure 5. Cis-acting regulatory elements (CREs) of the 21 BpAHLs promoters. (A) Genomic DNA sequences of the 2.0 kb upstream promoters of the BpAHLs were submitted to the PlantCare website to identify CREs. These elements can be classified into two main categories: CREs associated with hormone responses, and CREs related to abiotic stress responses. The squares in various colors represent different classes of CREs. (B) Summary of the ten classes of CREs identified in the promoters of BpAHLs. The black dots with numbers indicate the specific BpAHLs that contain the corresponding number of CREs. (C) Number of CREs present in each BpAHL gene. The bars, from left to right, represent BpAHL01 to BpAHL21. (D) Number of BpAHLs in each class of CRE. The bars in different colors represent different CREs, with corresponding annotations at the bottom left.
Forests 16 00943 g005
Figure 6. Tissue-specific expression profiles of BpAHL genes. (A) Hierarchical clustering of the expression profiles of 21 BpAHLs across four tissues: flower, leaf, root, and xylem. The color scale represents normalized transcripts per million (TPM), where yellow cells indicate low gene expression and blue cells indicate high gene expression. The yellow circles represent Type-I BpAHLs, while the green circles represent Type-II/III BpAHLs. The cluster dendrograms are shown in the left panel. (B) Relative expression levels of BpAHL06, BpAHL09, and BpAHL21 as validated by qRT-PCR. The qRT-PCR results were normalized using the tubulin gene and are presented relative to the leaf. The x-axis represents different tissues (stem, root, and leaf), while the y-axis illustrates the relative gene expression levels. Error bars indicate the standard deviations (SD) of three technical replicates, and the asterisk (‘***’) indicates that a p-value calculated using the one-way ANOVA is ≤0.001.
Figure 6. Tissue-specific expression profiles of BpAHL genes. (A) Hierarchical clustering of the expression profiles of 21 BpAHLs across four tissues: flower, leaf, root, and xylem. The color scale represents normalized transcripts per million (TPM), where yellow cells indicate low gene expression and blue cells indicate high gene expression. The yellow circles represent Type-I BpAHLs, while the green circles represent Type-II/III BpAHLs. The cluster dendrograms are shown in the left panel. (B) Relative expression levels of BpAHL06, BpAHL09, and BpAHL21 as validated by qRT-PCR. The qRT-PCR results were normalized using the tubulin gene and are presented relative to the leaf. The x-axis represents different tissues (stem, root, and leaf), while the y-axis illustrates the relative gene expression levels. Error bars indicate the standard deviations (SD) of three technical replicates, and the asterisk (‘***’) indicates that a p-value calculated using the one-way ANOVA is ≤0.001.
Forests 16 00943 g006
Figure 7. Expression profiles of BpAHL genes in response to abiotic stresses. (A) Expression of BpAHLs under short-term cold stress. The x-axis indicates the duration of cold treatments, with time points at 0.5 h, 1.0 h, 1.5 h, 2.0 h, 2.5 h, and 3.0 h. The y-axis indicates the expression levels of BpAHLs, clustered by their TPM. Expression of BpAHLs under (B) long-term cold stress and (C) heat stress. In these panels, the x-axis indicates the duration of treatments, including 0 h, 6 h, 24 h, 2 d, 4 d, 7 d, and 14 d. Similar annotations are as described for Figure 6. Expression of BpAHLs under salt stress (0.2 M NaCl for 24 h) is depicted in (D) leaf and (E) root. (F) Expression of BpAHLs treated with light/dark treatments over 4 days. Gene expression levels (TPM) below 2.0 are not shown. An asterisk (‘*’) indicates that a p-value calculated using the Student’s t-test is ≤0.05, while ‘n.s’ denotes that the differences are not significant.
Figure 7. Expression profiles of BpAHL genes in response to abiotic stresses. (A) Expression of BpAHLs under short-term cold stress. The x-axis indicates the duration of cold treatments, with time points at 0.5 h, 1.0 h, 1.5 h, 2.0 h, 2.5 h, and 3.0 h. The y-axis indicates the expression levels of BpAHLs, clustered by their TPM. Expression of BpAHLs under (B) long-term cold stress and (C) heat stress. In these panels, the x-axis indicates the duration of treatments, including 0 h, 6 h, 24 h, 2 d, 4 d, 7 d, and 14 d. Similar annotations are as described for Figure 6. Expression of BpAHLs under salt stress (0.2 M NaCl for 24 h) is depicted in (D) leaf and (E) root. (F) Expression of BpAHLs treated with light/dark treatments over 4 days. Gene expression levels (TPM) below 2.0 are not shown. An asterisk (‘*’) indicates that a p-value calculated using the Student’s t-test is ≤0.05, while ‘n.s’ denotes that the differences are not significant.
Forests 16 00943 g007
Figure 8. Expression patterns of BpAHL genes in birch under drought conditions. (A) The growth of birch under drought stress, with the treatment group on the left and the control group on the right. (B) Expression of BpAHL06, BpAHL13, and BpAHL20 in roots on the fifth day. The asterisk (‘*’, ‘**’, and ‘***’) indicates that a p-value calculated using the one-way ANOVA is ≤0.05, ≤0.01, and ≤0.001, respectively.
Figure 8. Expression patterns of BpAHL genes in birch under drought conditions. (A) The growth of birch under drought stress, with the treatment group on the left and the control group on the right. (B) Expression of BpAHL06, BpAHL13, and BpAHL20 in roots on the fifth day. The asterisk (‘*’, ‘**’, and ‘***’) indicates that a p-value calculated using the one-way ANOVA is ≤0.05, ≤0.01, and ≤0.001, respectively.
Forests 16 00943 g008
Table 1. Genomic and chemical characteristics of the 21 BpAHLs, including gene name, gene accession, genomic location, type, gene length, CDS length, protein length, molecular weight (MW), isoelectric point (pI), and subcellular localization. The designations ‘C’, ‘Cm’, and ‘N’ correspond to ‘Chloroplast’, ‘Cell membrane’, and ‘Nucleus’, respectively.
Table 1. Genomic and chemical characteristics of the 21 BpAHLs, including gene name, gene accession, genomic location, type, gene length, CDS length, protein length, molecular weight (MW), isoelectric point (pI), and subcellular localization. The designations ‘C’, ‘Cm’, and ‘N’ correspond to ‘Chloroplast’, ‘Cell membrane’, and ‘Nucleus’, respectively.
Gene NameGene AccessionGenomic LocationTypeGene
Length
(bp)
CDS Length
(bp)
Protein Length
(aa)
Molecular Weight
(Da)
Isoelectric PointSubcellular
Localization
BpAHL01Bpev01.c0774.g0008Chr1: 11,387,186–11,387,992I80780726827,4895.97N
BpAHL02Bpev01.c0616.g0003Chr2: 12,611,347–12,618,513III7167104134635,8008.97Cm/C/N
BpAHL03Bpev01.c0362.g0004Chr2: 19,921,747–19,922,535I78978926227,3218.38C/N
BpAHL04Bpev01.c1484.g0011Chr3: 16,791,132–16,791,854I72372324024,1508.97N
BpAHL05Bpev01.c1390.g0007Chr4: 10,015,299–10,026,351III11,053103834535,5609.00C/N
BpAHL06Bpev01.c1390.g0016Chr4:10,121,683–10,122,537I85585528429,3855.48N
BpAHL07Bpev01.c0365.g0010Chr5: 2,456,997–2,464,503I7507108636136,8784.73N
BpAHL08Bpev01.c0540.g0008Chr5: 6,984,368–6,996,557III12,19074124625,2029.73N
BpAHL09Bpev01.c0027.g0117Chr5: 23,300,208–23,306,294II6087102033934,88310.05N
BpAHL10Bpev01.c0874.g0035Chr5: 26,805,688–26,806,626I93993931232,9237.16N
BpAHL11Bpev01.c0169.g0027Chr8: 22,673,422–22,678,438II5017109536437,4218.51N
BpAHL12Bpev01.c0480.g0053Chr9: 391,896–392,816I92192130630,8387.93N
BpAHL13Bpev01.c0480.g0043Chr9: 485,086–490,203II5118112237337,10110.08N
BpAHL14Bpev01.c0902.g0012Chr13: 7,069,248–7,078,300II905397232333,7375.68C
BpAHL15Bpev01.c0223.g0044Chr13: 10,271,229–10,272,056I82882827528,0927.05N
BpAHL16Bpev01.c1544.g0004Chr14: 2,458,829–2,467,966III9138145248352,1285.64C/N
BpAHL17Bpev01.c1544.g0009Chr14: 2,534,436–2,535,362I92792730833,0026.53N
BpAHL18Bpev01.c0449.g0016Chr14: 4,176,010–4,181,875II5866103234335,22810.25Cm/C/N
BpAHL19Bpev01.c0964.g0005Chr14: 15,052,985–15,053,881I89789729830,7706.15N
BpAHL20Bpev01.c0964.g0003Chr14: 15,082,749–15,091,115II8367111937237,8079.57N
BpAHL21Bpev01.c1465.g0006Contig1465: 35,258–36,043I78678626127,4476.06N
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

Chen, B.; Chu, H.; Lv, B.; Guo, Y.; Zhang, Z.; Zhang, T.; Xie, Q.; Hao, M.; Ali, S.; Zhou, W.; et al. Genome-Wide Identification and Expression Analysis of AT-Hook Motif Nuclear Localized Gene Family in Birch. Forests 2025, 16, 943. https://doi.org/10.3390/f16060943

AMA Style

Chen B, Chu H, Lv B, Guo Y, Zhang Z, Zhang T, Xie Q, Hao M, Ali S, Zhou W, et al. Genome-Wide Identification and Expression Analysis of AT-Hook Motif Nuclear Localized Gene Family in Birch. Forests. 2025; 16(6):943. https://doi.org/10.3390/f16060943

Chicago/Turabian Style

Chen, Bowei, Huaixue Chu, Bin Lv, Yile Guo, Zihui Zhang, Tianxu Zhang, Qingyi Xie, Menghan Hao, Shahid Ali, Wei Zhou, and et al. 2025. "Genome-Wide Identification and Expression Analysis of AT-Hook Motif Nuclear Localized Gene Family in Birch" Forests 16, no. 6: 943. https://doi.org/10.3390/f16060943

APA Style

Chen, B., Chu, H., Lv, B., Guo, Y., Zhang, Z., Zhang, T., Xie, Q., Hao, M., Ali, S., Zhou, W., Zhao, L., Jiang, Z., Wang, M., & Xie, L. (2025). Genome-Wide Identification and Expression Analysis of AT-Hook Motif Nuclear Localized Gene Family in Birch. Forests, 16(6), 943. https://doi.org/10.3390/f16060943

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