Abstract

The genetic bases of halophytes for salinity tolerance are crucial for genetically breeding salt-tolerant crops. All natural Nitrariaceae species that exclusively occur in arid environments are highly tolerant to salt stress, but the underlying genomic bases to this adaptation remain unknown. Here we present a high-quality, chromosome-level genome sequence of Nitraria sibirica, with an assembled size of 456.66 Mb and 23,365 annotated genes. Phylogenomic analyses confirmed N. sibirica as the sister to all other sampled representatives from other families in Sapindales, and no lineage-specific whole-genome duplication was found except the gamma triplication event. Still, we found that the genes involved in K+ retention, energy supply, and Fe absorption expanded greatly in N. sibirica. Deep transcriptome analyses showed that leaf photosynthesis and cuticular wax formation in roots were enhanced under salt treatments. Furthermore, many transcription factors involved in salt tolerance changed their expressions significantly and displayed tissue- and concentration-dependent signalling in response to salt stress. Additionally, we found vacuolar Na+ compartmentalization is an ongoing process under salt treatment, while Na+ exclusion tends to function at high salt concentrations. These genomic and transcriptomic changes conferred salt tolerance in N. sibirica and pave the way for the future breeding of salt-tolerant crops.

1. Introduction

Soil salinization is a global problem that threatens crop growth and production and hampers modern agriculture’s sustainable development.1–3 High salt concentrations induce ionic imbalance and hyperosmotic stress that may lead to poor plant growth or death.4,5 However, many plants occurring in salty or arid environments have developed a series of genetic changes to prevent salt toxicity and maintain osmotic homeostasis.6–13 For example, in the desert poplar, the up-regulated expression of ion transport-related genes and the expansion of genes involved in maintaining intracellular homeostasis make this species salt-tolerant.7 In grasses, salt tolerance is enhanced by some duplicated genes resulting from whole-genome duplication (WGD).12 However, the genomic changes of many other salt-tolerant lineages remain largely unexplored.

The family Nitrariaceae in the broad sense (including Tetradiclidaceae and Peganaceae), in the order Sapindales, comprise around four genera and 20 species that mainly occur in arid and salty environments and show high salt tolerance.14,15 However, other families of this order, including Rutaceae, Anacardiaceae, and Aceraceae are distributed in subtropical and tropical environments, and most species are sensitive to salt stress. Therefore, Nitrariaceae comprise a specific lineage to examine genomic changes for salt tolerance compared with other Sapindales families. Furthermore, the family has attracted research interests in many other aspects, such as secondary chemistry, biogeography, and floral morphology.16–19 For example, Nitrariaceae plants are rich and diverse in metabolites, of which numerous chemical constituents isolated from the genera Nitraria and Peganum are shown to have anticancer and antioxidant effects18,19; In addition, although a small family but it has multiple intercontinental discontinuous distribution patterns, making it an ideal plant group to investigate biogeographical pattern in drylands.17 However, up to now, no genome sequence is available for Nitrariaceae, although de novo genomes have been reported for other species across several families in Sapindales.9,20,21 In this study, we aimed to assemble the genome sequence of the type species of the family, Nitraria sibirica Pall. This shrub, occurring in the desert and salt-alkaline regions like other species, can tolerate extremely saline and drought conditions (Fig. 1A).22–24 Both physiological and biochemical tests have also confirmed such traits.24,25 For example, mesophyll vacuoles of leaves act as the main Na+ sink to store excess sodium ions,26 while the roots have superior K+ retention ability.26,27

Genomic features of N. sibirica. A. Habitat, leaf, fruit, flower, and distribution of N. sibirica (a–e). The blue triangles in ‘e’ represents the distribution information of N. sibirica obtained from GBIF (Global Biodiversity Information Facility, https://www.gbif.org/). B. Synteny and gene distribution features of the N. sibirica genome within a window size of 200 kb. The tracks from outer to inner circles indicate the following: (I) the assembled 12 chromosomes (Chr1–Chr12); (II) GC content (0.3–0.5); (III) gene density; (IV) gypsy coverage; (V) copia coverage; (VI) LTR coverage; (VII) synteny blocks.
Figure 1.

Genomic features of N. sibirica. A. Habitat, leaf, fruit, flower, and distribution of N. sibirica (a–e). The blue triangles in ‘e’ represents the distribution information of N. sibirica obtained from GBIF (Global Biodiversity Information Facility, https://www.gbif.org/). B. Synteny and gene distribution features of the N. sibirica genome within a window size of 200 kb. The tracks from outer to inner circles indicate the following: (I) the assembled 12 chromosomes (Chr1–Chr12); (II) GC content (0.3–0.5); (III) gene density; (IV) gypsy coverage; (V) copia coverage; (VI) LTR coverage; (VII) synteny blocks.

Here, we report a high-quality chromosome-scale genome of N. sibirica from a combination of several sequencing platforms and strategies, including Illumina short reads sequencing, Oxford Nanopore Technologies (ONT) long reads sequencing, and chromosome conformation capture (Hi-C) technology. RNA-seq-based transcriptomics was used to optimize gene structural prediction and annotation and explore differentially expressed genes (DEGs), co-responsive expression patterns, and their regulatory networks under different salt concentrations across various tissues. As the first genome sequence of the family Nitrariaceae, we aimed to use it to address the following questions. (1) Did special genomic changes, for example, polyploidization that usually increases the salinity adaptation,12,28,29 occur in this lineage? (2) If not, did gene families expand their copies as found in the other salt-tolerant species?7 Especially what roles do transcription factors (TFs) play in response to salt stress? Moreover, (3) do TFs and their associated regulatory networks display tissue- or concentration-dependent signalling in response to salt stress?

2. Materials and methods

2.1. Plant materials

Plant samples for this study were collected from salinized land (E: 103°34ʹ17.9472″, N: 39°0ʹ30.5172″) near Minqin Desert Botanical Garden, Gansu Province, China. The collected tissues of roots, stems, leaves, and flowers of N. sibirica were immediately stored in liquid nitrogen and subsequently used for genome or transcriptome sequencing. Plant seeds collected from the same population were germinated and planted in the greenhouse for subsequent salt stress treatment and transcriptome sequencing.

2.2. Sequencing and genome assembly

Total DNA was extracted from the fresh young leaves using the sodium dodecyl sulphate method and purified with the QIAGEN® Genomic kit (Hilden, Germany) according to the standard operating procedure. The DNA library for Illumina short reads with 300 bp insert sizes were constructed using the Illumina DNA Prep Kit and sequenced by Illumina HiSeq 4000 platform in 150-bp paired-end mode. The genomic library of long-reads was constructed and sequenced utilizing PromethION, an Oxford Nanopore Technology sequencer, to obtain raw sequencing data with 20 Kb insert sizes. Then a common standard was used to filter raw reads (removal of sequencing adapter and trimming of low-quality bases with an ‘average quality score < 7’). The GC content distribution range of filtered long reads is shown in Supplementary Fig. 1. The Hi-C sequencing was carried out as follows. Interoperable DNA fragments were obtained by cross-linkage of 2% formaldehyde with sampled DNA, and Dpn II restriction enzymes were used to digest chromatin, then the Illumina Nova-seq platform was used for sequencing in 150-bp paired-end mode.30

The genome size of N. sibirica was estimated with a k-mer depth-frequency distribution (k-mer = 17) using KmerFreq_AR as implemented in SOAPdenovo231 from Illumina 150 bp paired-end read data. For de novo genome assembly, the NextDenovo v2.3.0 software (https://github.com/Nextomics/NextDenovo), an overlap layout-consensus (OLC) assembled method specially developed for ONT reads, was used to generate the primary assembly for N. sibirica as follows. First, the program NextCorrect was used to self-correct the original subreads, and then the consensus sequences (CS reads) were obtained. Second, CS reads were compared with the NextGraph module to capture each correlation between them. The preliminary genome was assembled based on the correlation relationships of CS reads. Finally, to improve the accuracy of the assembly, the contigs were refined with Racon v1.3.132 using ONT long reads and Nextpolish v1.2.433 using Illumina short reads, respectively. After, we assessed the potential contamination by comparing the assembled contigs with the National Center for Biotechnology Information (NCBI)—non-redundant (NR) nucleotide database (https://www.ncbi.nlm.nih.gov/nucleotide/) using BLAST with an E-value cutoff of 1e – 5. As shown in Supplementary Table 1, no external contaminations (endophytes and bacteria, etc.) were found and four contigs derived from mitochondrion and chloroplast was removed from subsequent chromosome-level assembly. For the Hi-C-assisted assembly, 370 million paired-end reads were generated. The raw Hi-C data were then quality controlled using Hi-C-Pro v2.8.1,34 and valid interaction paired reads were used for subsequent scaffolding. Subsequently, the assembled scaffolds were divided into bins of 100 Kb size, and the number of contacted Hi-C read pairs between each pair of bins was calculated. Based on the contact information, the scaffolds were further clustered, sorted, and oriented to chromosomes by LACHESIS.35 Finally, some errors in position and orientation were manually adjusted based on chromatin interaction frequency by using Juicebox (https://github.com/aidenlab/Juicebox). The completeness of genome assembly was assessed using Benchmarking Universal Single-Copy Orthologs (BUSCO) v4.0.536 and Core Eukaryotic Gene Mapping Approach (CEGMA) v2.37 We evaluated the accuracy of the assembly by mapping all the Illumina paired-end reads to the assembled genome using Burrows–Wheeler Aligner (BWA) v0.7.12.38 Mapping rate and genome coverage were assessed using SAMtools v0.1.1855.39 Besides, base accuracy of the assembly was calculated with BCFtools v1.8.0.40

2.3. Genome annotation

Tandem repeats were annotated using GMATA v2.241 and Tandem Repeats Finder (TRF) v4.07b.42 Transposable elements (TEs) were identified using a combination of ab inito and homology-based methods. Briefly, an ab inito repeat library was first predicted using MITE-hunter43 and RepeatModeler v1.0.1144 with default parameters. For further identification of the repeats throughout the genome, RepeatMasker was used to search for known and novel TEs by mapping sequences against the de novo repeat library and Repbase.45 Finally, overlapping transposable elements of the same repeat class were collated and combined. In addition, we used tRNAscan-SE v2.0,46 Infernal v1.1.2,47 and RNAmmer v1.248 to predict non-coding RNAs (including miRNAs, tRNAs, rRNAs, and snRNAs).

Three independent approaches were used for protein-coding gene prediction in a repeat-masked genome, including ab initio prediction, homology search, and RNA-seq-based. In detail, GeMoMa v1.6.149 was used to align the homologous peptides from four closely related species with high-quality published genomes (Acer yangbiense, Citrus maxima, Athaliana thaliana, and Vitis vinifera) to the repeat-masked assembly of N. sibirica to obtain the gene structure information. Meanwhile, Augustus v3.3.150 was run with default parameters for ab initio gene prediction with the training set. For RNA-seq-based gene prediction, filtered mRNA-seq reads were aligned to the N. sibirica reference genome using STAR v2.7.3a51 with default parameters. The transcripts were then assembled using Stringtie v1.3.4d,52 and open reading frames (ORFs) were predicted using PASA v2.3.3.53 Finally, comprehensive gene sets obtained by different strategies were integrated by EVidenceModeler (EVM) v1.1.1.53 For gene function annotation, BLASTP (E-value < 1 – e05) was used in two public databases, including SwissProt (https://web.expasy.org/docs/swiss-prot_guideline.html) and the NCBI-NR protein database (https://ftp.ncbi.nlm.nih.gov/blast/db/FASTA/) to obtain functional annotations of predicted protein-coding genes. Protein structural domains were predicted by InterProScan v5.23.54 Gene Ontology (GO) entries for each gene were searched using Blast2GO v3.0,55 and pathway information was assigned using the Kyoto Encyclopedia of Genes and Genomes (KEGG) database (https://www.genome.jp/kegg/).

2.4. Phylogenomic analyses

To explore the phylogenetic position of Nitrariaceae, we collected the protein-coding gene sequences of 11 other sequenced plant species in eudicots: including four Sapindales species (A. yangbiense, Citrus clementina, Poncirus trifoliata, and Pistacia vera) and seven additional eudicot species (A. thaliana, Corchorus capsularis, Caragana korshinskii, Cucumis sativus, Populus trichocarpa, Tripterygium wilfordii, and V. vinifera). The detailed genome information and sources are listed in Supplementary Table 2. Only the longest transcript of each gene was used for subsequent analysis. We used OrthoFinder v2.4.056 to cluster the protein sequences of N. sibirica and the other 11 species to putative gene families (or ortho groups) based on sequence similarity with default parameters. In total, 554 strict single-copy gene families were obtained. MAFFT v7.47157 was used to perform multiple sequence alignments of the protein sequences in each sing-copy gene family. Then, the protein sequence alignments were converted into corresponding DNA sequence alignments using PAL2NAL v14.58 We used maximum likelihood and coalescent-based methods for phylogenetic analyses. For maximum likelihood tree reconstruction, the DNA sequence alignments of single-copy gene families were concatenated into a super-gene matrix, and then model selection and tree inference were automatically performed with IQTREE v2.1.2.59 For coalescent-based phylogenetic analysis, the trees of each gene family were constructed using IQTREE and subsequently summarized using ASTRALL-III v5.7.5.60 In addition, the chloroplast genome of N. sibirica was assembled and annotated with GetOrganelle v1.7.561 and CPGAVAS2,62 respectively. The chloroplast genomes of other representative species were downloaded from the public database (Supplementary Table 3). Phylogenetic analysis of chloroplast genomes was performed using the maximum likelihood method as in the nuclear data. The divergence time estimation of the 12 species was done using the MCMCtree pipeline of the PAML program v4.9.63 We used three calibration points obtained from the TimeTree database (http://timetree.org/), including the crown group of Sapindales (56–99 Mya), the split of C. capsularisA. thaliana (84–95 Mya), and the crown group of V. vinifera and other species (106–125 Mya).

2.5. Comparative genomic analyses

To estimate the evolutionary dynamics of gene families, we employed the computational analysis of gene family evolution (CAFÉ) v4.2 program64 to explore gene family expansion and contraction events across the time-calibrated phylogeny based on a random birth and death model. GO enrichment analysis of the expanded and extracted gene families was performed with TBtools.65

We used Ks-based age distributions of paralogous genes and intra- and inter-genomic synteny to analyze the WGD history in N. sibirica. Three other Sapindales species (A. yangbiense, C. clementina, and P. trifoliata) with published chromosomal-level genomes were used for comparisons, and V. vinifera, a lineage without any further WGDs after the gamma event, was used an outgroup. Ks values of paralogous genes were calculated using the Nei–Gojobori method66 by the yn00 package of PAML v4.9.61 Intra- and inter-genomic synteny comparisons were performed using WGDI with default parameters.67

2.6. Identification of key gene members relevant to salinity adaptation

We used BLASTP and hidden Markov model (HMM) search68 to identify the gene members in ion transport, tricarboxylic acid cycle, nicotianmine biosynthesis, and salt overly sensitive pathways. Protein sequences of gene members of each gene family in A. thaliana were retrieved from the Arabidopsis Information Resource (TAIR) (https://www.arabidopsis.org) or relevant references. The retrieved A. thaliana proteins were then searched against the protein sequences in other four Sapindales species (A. yangbiense, C. clementina, P. trifoliata, and P. vera) and V. vinifera using BLASTP with E-value < 1e – 10 and sequence identity > 30%. Meanwhile, the HMM profiles with protein domains of each family were downloaded from the Pfam website (http://pfam.xfam.org), and then batch searched in each species with HMMER v3.3.2.68 All candidates of each gene family were used for maximum likelihood tree construction using IQTREE59 as previously described. Any outliers in the phylogeny were manually checked and removed, and the remaining genes were regarded as true members of a gene family.

2.7. Transcriptome sequencing and analyses

Two months N. sibirica seedlings of equal growth conditions were used for salt treatments. It has been suggested that the optimal salt concentration for N. sibirica growth is around 200 mM NaCl, and it could tolerant at least a 400 mM NaCl salt treatment.22,24,26 Therefore, we set two salt concentrations: moderate with 200 mM NaCl treatment and high with 400 mM NaCl treatment. Three major tissues (roots, stems, and leaves) were sampled to examine the gene expression profiles in response to salt stress. Salt treatments were performed by adding 50 mM NaCl every half day to avoid overstimulation until it reached the two treatment concentrations (T200 and T400) and a further week of stress. Seedlings without salt treatment served as control. A total of three biological replicates were set up. Roots, stems, and leaves of NaCl-treated and control seedlings were collected and sorted in liquid nitrogen for subsequent sequencing. Total RNA extraction, library construction, and sequencing were performed by Novogene Co, Ltd (Beijing, China) on the Illumina Nova-seq platform.

A total of 27 transcriptomes from three tissues with two salt concentrations and controls were generated and yielded 193.38 Gb (Supplementary Table 4). For each sample, the quality of raw reads was evaluated with FastQC (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/), and low-quality reads was trimmed with fastp v0.20.0.69 After quality control, clean reads were mapped to the N. sibirica genome using HISAT2 v 2.1.0.70 StringTie v1.3.4d53 was used to calculate the transcripts per million (TPM) of genes in each sample. Differentially expressed genes between treatments and controls in different tissues were identified using the DESeq2 v3.11.71 The candidate genes with at least 2-fold differential expression levels in comparisons and a false discovery rate (FDR) value < 0.05 were identified as DEGs. To check whether differentially expressed TFs evolved under positive selection, we analyzed the accumulation of non-synonymous changes by comparing N. sibirica with three salt-sensitive species (A. yangbiense, C. clementina, and P. trifoliata) in Sapindales following the procedures of Ma et al. (2013).7 The R package WGCNA v1.6972 was used to perform weighted gene co-expression network analysis and the resulting networks were visualized using Cytoscape v3.9.1.73 We analyzed the average expression level (TPM) tendency of genes in each tissue (root, stem, and leaf) under different salt concentrations (control, T200, and T400) on the Omicshare Tools website (https://www.omicshare.com/). The KOBAS website (http://kobas.cbi.pku.edu.cn/kobas3/) was used to perform a KEGG enrichment analysis of the genes in each cluster.

3. Results

3.1 Chromosomal-level genome assembly of N. sibirica

Nitraria sibirica is a diploid halophyte (2n = 2x = 24),74 with an estimated genome size of 482.77 Mb based on genome survey analysis (Supplementary Table 5; Supplementary Fig. 2). For de novo whole-genome sequencing, we produced 38 Gb (~79 × coverage) Nanopore long-read sequencing data, 32 Gb (~67 × coverage) short reads by Illumina sequencing, and 55 Gb (~116 × coverage) of Hi-C paired-end reads (Supplementary Tables 6–8). After self-correction of long reads, contig-level primary assembly, polishing with short reads, removing contaminations, and further scaffolding (see Materials and Methods), we obtained a correct genome assembly totalling 456.66 Mb. The assembled genome is slightly smaller than the estimated genome, consisting of 170 contigs with an N50 value of 9.07 Mb and a longest contig of 21.14 Mb (Table 1, Supplementary Table 9). We further used Hi-C contact information to construct a chromosomal-level genome assembly that resulted in 98.44% of the genomic sequences assigned unambiguously to discrete chromosome locations (Fig. 1B and Supplementary Fig. 3), including 12 pseudochromosomes with sizes ranging from 24.3 to 56.2 Mb (Supplementary Table 10).

Table 1.

Assembly and annotation features of N. sibirica.

FeatureStatistics
Genome assembly
 Estimated genome size (Mb)482.55
 Assembled genome size (Mb)456.66
 GC content (%)35.26
 Number of contigs170
 Contig N50 (Mb)9.07
 Scaffold N50 (Mb)44.88
 Longest contig (Mb)21.13
 Repeat region % of assembly62.85
 BUSCO score of assembly (%)98.14
Genome annotation
 Predicted gene models23,365
 Average gene length (bp)4154.54
 Average exons per gene5.91
 Average exon length (bp)215.58
 Average intron length (bp)586.61
 Average CDS length (bp)1274.13
 Number of long non-coding RNAs
 BUSCO score of annotation (%)
5,934
97.71
FeatureStatistics
Genome assembly
 Estimated genome size (Mb)482.55
 Assembled genome size (Mb)456.66
 GC content (%)35.26
 Number of contigs170
 Contig N50 (Mb)9.07
 Scaffold N50 (Mb)44.88
 Longest contig (Mb)21.13
 Repeat region % of assembly62.85
 BUSCO score of assembly (%)98.14
Genome annotation
 Predicted gene models23,365
 Average gene length (bp)4154.54
 Average exons per gene5.91
 Average exon length (bp)215.58
 Average intron length (bp)586.61
 Average CDS length (bp)1274.13
 Number of long non-coding RNAs
 BUSCO score of annotation (%)
5,934
97.71
Table 1.

Assembly and annotation features of N. sibirica.

FeatureStatistics
Genome assembly
 Estimated genome size (Mb)482.55
 Assembled genome size (Mb)456.66
 GC content (%)35.26
 Number of contigs170
 Contig N50 (Mb)9.07
 Scaffold N50 (Mb)44.88
 Longest contig (Mb)21.13
 Repeat region % of assembly62.85
 BUSCO score of assembly (%)98.14
Genome annotation
 Predicted gene models23,365
 Average gene length (bp)4154.54
 Average exons per gene5.91
 Average exon length (bp)215.58
 Average intron length (bp)586.61
 Average CDS length (bp)1274.13
 Number of long non-coding RNAs
 BUSCO score of annotation (%)
5,934
97.71
FeatureStatistics
Genome assembly
 Estimated genome size (Mb)482.55
 Assembled genome size (Mb)456.66
 GC content (%)35.26
 Number of contigs170
 Contig N50 (Mb)9.07
 Scaffold N50 (Mb)44.88
 Longest contig (Mb)21.13
 Repeat region % of assembly62.85
 BUSCO score of assembly (%)98.14
Genome annotation
 Predicted gene models23,365
 Average gene length (bp)4154.54
 Average exons per gene5.91
 Average exon length (bp)215.58
 Average intron length (bp)586.61
 Average CDS length (bp)1274.13
 Number of long non-coding RNAs
 BUSCO score of annotation (%)
5,934
97.71

To assess the accuracy of the final genome assembly, the genomic data from Illumina short reads and transcriptome reads (i.e. those from roots, stems, leaves, flowers, fruits, and various tissues under salt treatments) were mapped to the assembled genome, resulting in mapping rates of above 99.5% and 92.3%, respectively (Supplementary Table 4). Furthermore, 98.14% of the 1,614 BUSCO genes were detected in the assembled genome (Supplementary Table 11). Similarly, 97.98% of the ultra-conserved core eukaryotic genes, based on CEGMA analysis, were completely detected in the assembly (Supplementary Table 12). The genome assembly achieved ‘reference quality’ with an estimated long-terminal repeat (LTR) assembly index (LAI) value of 12.17 (Supplementary Table 13).75 Together, these results suggest that the continuity and integrity of the N. sibirica genome assembly are both of high quality.

3.2 Genome annotation of N. sibirica

We annotated 23,365 protein-coding gene models (Table 1) from the N. sibirica genome by integrating de novo, homology, and RNA-seq-based prediction methods (see Materials and Methods). Of these, 19,918 high-confidence genes (85.25% of the total genes) were supported by transcriptome gene expressive evidence (Supplementary Table 14). The average gene length, coding-sequence length, and exons per gene are 4,154.54 bp, 1,274.13 bp, and 5.91, respectively (Table 1), like in other eudicot species (Supplementary Fig. 4). About 87.17% of the total genes were annotated as functional by at least one of the following protein-related databases: GO database (54.71%), KEGG database (32.6%), NCBI-NR protein database (86.48%), the Swiss-Prot database (73.4%), and Clusters of Orthologous Groups for Eukaryotic complete Genomes (KOG) database (49.29%) (Supplementary Table 15). BUSCO assessment showed that our annotation recovered 97.71% of the gene set (Supplementary Table 16).

The N. sibirica genome is composed of 62.85% of repetitive elements (Table 1), which is slightly lower than in A. yangbiense (68.0%)21 and P. vera (70.7%),9 but much higher than in C. clementina (45%).20 With retrotransposons (Class I) and transposons (Class II) transposable elements (TEs) representing 45.61% and 12.79% of the genome, respectively (Supplementary Table 17). LTR retrotransposons are the most abundant TEs, representing 40.64% of the assembled genome (Supplementary Table 17). Moreover, most of these LTRs burst around one million years ago (Mya) (Supplementary Fig. 5). In addition to protein-coding genes and repetitive elements, we predicted 734 transfer RNAs (tRNA), 264 microRNAs (miRNA), 865 small nuclear RNAs (snRNA) and 3,789 ribosomal RNAs (rRNA; Supplementary Table 18).

3.3. Phylogenomic inference of the evolutionary position of Nitrariaceae

To resolve the phylogenetic position of Nitrariaceae, we extracted 554 strict single-copy nuclear genes from the N. sibirica genome and other representative species with published genomes, including four of Sapindales—C. clementina (Rutaceae), P. trifoliata (Rutaceae), P. vera (Anacardiaceae), and A. yangbiense (Aceraceae), six of other eudicots lineages—C. capsularis (Malvales), A. thaliana (Brassicales), P. trichocarpa (Malpighiales), T. wilfordii (Celastrales), C. korshinskii (Fabales), and C. sativus (Cucurbitales), and one outgroup species—V. vinifera (Vitales; Supplementary Table 2). We employed maximum likelihood and coalescent-based methods to reconstruct the phylogeny of Nitrariaceae and other representative lineages in eudicots (see Materials and Methods). The two methods retrieved the same topology, that is, N. sibirica of Nitrariaceae as a sister to all other representatives of the remaining families in Sapindales (bootstrap support (BS) = 100%; Fig. 2 and Supplementary Fig. 6A). In addition, we also assembled a complete chloroplast genome of N. sibirica and performed plastid phylogenomic analysis for N. sibirica and related species. The results showed that the topology was consistent with the nuclear phylogeny for the placement of N. sibirica, indicating the robustness of Nitrariaceae as a member of Sapindales (BS = 100%; Supplementary Fig. 6B). Based on this well-resolved phylogeny, the divergence time between N. sibirica and other Sapindales species was estimated at around 75 Mya (Supplementary Fig. 7).

Comparative genomic analysis of N. sibirica and eleven other plant species. A. Ks distributions for paralogues in N. sibirica and three other Sapindales species (A. yangbiense, C. clementina, and P. trifoliata), and V. vinifera. Whole-genome triplication (WGT) event was found in all species. B. Chromosomal synteny blocks among N. sibirica, three other Sapindales species, and V. vinifera. The numbers represent individual chromosomes. The green lines indicate one versus one syntenic depth ratio between these Sapindales species with V. vinifera as an outgroup. C. Time-calibrated phylogenetic tree of N. sibirica and eleven other species based on 554 strict single-copy nuclear genes. Bootstrap values for all branches are 100% from both maximum likelihood and coalescent-based analyses. The numbers of expanded (blue) and contracted (red) gene families are shown above the branches. The boxed number indicates the remaining gene-family size at each node.
Figure 2.

Comparative genomic analysis of N. sibirica and eleven other plant species. A. Ks distributions for paralogues in N. sibirica and three other Sapindales species (A. yangbiense, C. clementina, and P. trifoliata), and V. vinifera. Whole-genome triplication (WGT) event was found in all species. B. Chromosomal synteny blocks among N. sibirica, three other Sapindales species, and V. vinifera. The numbers represent individual chromosomes. The green lines indicate one versus one syntenic depth ratio between these Sapindales species with V. vinifera as an outgroup. C. Time-calibrated phylogenetic tree of N. sibirica and eleven other species based on 554 strict single-copy nuclear genes. Bootstrap values for all branches are 100% from both maximum likelihood and coalescent-based analyses. The numbers of expanded (blue) and contracted (red) gene families are shown above the branches. The boxed number indicates the remaining gene-family size at each node.

3.4. Comparative genomic analyses

The distributions of synonymous substitutions per site (Ks) and intragenomic syntenic analysis suggested that the genomes of Sapindales species have not undergone any additional genome duplication, apart from the common whole-genome triplication (gamma) event shared by all core-eudicots species (Fig. 2A and Supplementary Fig. 8).76 Intergenomic syntenic analysis between N. sibirica and three Sapindales species (A. yangbiense, C. clementina, and P. trifoliata) and V. vinifera revealed an apparent 1:1 syntenic depth ratio, which further confirmed that no lineage-specific WGD occurred in N. sibirica and also other species of the order (Fig. 2B).

Since no WGD exists, we examined the expansion and contraction of gene families in N. sibirica. We found that many gene families in the N. sibirica genome have contracted (i.e. 4,481) and while only a few gene families have expanded (i.e. 1,073) (Fig. 2C; Supplementary Table 19). The contracted gene families are mainly involved in essential molecular functions (i.e. binding) and biological processes (i.e. nucleic acid metabolism) (Supplementary Fig. 9). However, GO enrichment indicated that all of the expanded gene families are involved in essential biological processes related to salinity adaptation, including potassium ion transport (GO: 0006813), and potassium ion transmembrane transport (GO:0071805), nicotianamine metabolic processes (GO:0030417), nicotianamine biosynthetic process (GO: 0030418), tricarboxylic acid (TCA) cycle metabolic process (GO:0072350), TCA biosynthetic process (GO:0072350), amine biosynthetic process (GO: 0009308) and carbohydrate metabolic process (GO: 0005975) (corrected P-value ≤ 0.01; Supplementary Table 20 and Supplementary Fig. 10).

The roots of N. sibirica have superior K+ retention ability,26 which is carried out through K+ channels and transporters2,77 that comprise a series of gene families, such as K+ uptake permease/high-affinity K+ transporter/K+ transporter (KUP/HAK/KT), Two-pore K+ channel (TPK), High-affinity K+ transporter (HKT), Shaker-type potassium channels (KV-like), K+ efflux antiporters (KEA), and Na+/H+ exchangers (NXH). Copy numbers of these gene families revealed that in N. sibirica, only the HAK5 subfamily was significantly expanded, which belongs to the KUP/HAK/KT gene family (Fig. 3A; Supplementary Table 21). Phylogenetic analysis of the HUP/HAK/KT gene family showed that the orthologs of HAK5 increased from one or two copies in other species of the Sapindales to six copies in N. sibirica, five of which were generated by tandem duplications (Fig. 3B). Three copies (LG02.979, LG02.984, and LG04.306) displayed up-regulated expression pattern under salt stress (Supplementary Fig. 11).

Evolutionary analysis of gene families in N. sibirica and six other plant species. A. Overview of the copy number of major gene families associated with salt response in N. sibirica, four other Sapindales species, and V. vinifera and A. thaliana. Gene families involved in biological processes such as potassium ion transport, tricarboxylic acid cycle, and nicotianmine synthase were significantly expanded in N. sibirica. The length of the bar indicates the relative size of each gene family copy number, and the number on the bar indicates the copy number of this gene family; the colour of the bar represents different species. B. Maximum likelihood phylogenetic tree of the KUP/HAK/KT gene family in N. sibirica and A. thaliana. The four clades (I–IV) of the phylogenetic tree are shown in different colours. The blue and red bold fonts represent the HAK5 genes of N. sibirica (six genes) and A. thaliana (one gene), respectively. The chromosomal locations of five HAK5 genes generated by tandem duplications are indicated near the phylogenetic tree. C. Maximum likelihood phylogenetic tree of the nicotianmine synthase (NAS) gene family in seven plant genomes. Different species are represented with coloured boxes. The chromosomal locations of NAS genes generated by tandem duplications are indicated around the phylogenetic tree.
Figure 3.

Evolutionary analysis of gene families in N. sibirica and six other plant species. A. Overview of the copy number of major gene families associated with salt response in N. sibirica, four other Sapindales species, and V. vinifera and A. thaliana. Gene families involved in biological processes such as potassium ion transport, tricarboxylic acid cycle, and nicotianmine synthase were significantly expanded in N. sibirica. The length of the bar indicates the relative size of each gene family copy number, and the number on the bar indicates the copy number of this gene family; the colour of the bar represents different species. B. Maximum likelihood phylogenetic tree of the KUP/HAK/KT gene family in N. sibirica and A. thaliana. The four clades (I–IV) of the phylogenetic tree are shown in different colours. The blue and red bold fonts represent the HAK5 genes of N. sibirica (six genes) and A. thaliana (one gene), respectively. The chromosomal locations of five HAK5 genes generated by tandem duplications are indicated near the phylogenetic tree. C. Maximum likelihood phylogenetic tree of the nicotianmine synthase (NAS) gene family in seven plant genomes. Different species are represented with coloured boxes. The chromosomal locations of NAS genes generated by tandem duplications are indicated around the phylogenetic tree.

As an important Fe chelator in plants, nicotianamine (NA) is vital for Fe deficiency toleration in high pH soils, such as saline–alkali soils.78 In N. sibirica, the Nicotianmine Synthase (NAS) gene family was identified as significantly expanded to include 16 copies, about fourfold higher than in other species of the Sapindales (Fig. 3A and C). The phylogenetic analysis found that most of these genes were the product of tandem duplications (Fig. 3C). Notably, under salt stress, most of these NAS genes were highly expressed in roots but not in stems and leaves (Supplementary Fig. 12), suggesting that it is likely that enhanced nicotianmine biosynthesis allowed roots to absorb sufficient Fe from the soil.

The TCA cycle is one of the most fundamental processes for cell energy production.79 It consists of nine metabolic steps and eight enzymes, of which citrate synthase (CSY), isocitrate dehydrogenase (IDH), and α-ketoglutarate dehydrogenase (α-KGDHC) are three rate-limiting enzymes that regulate the cycling rate (Supplementary Fig. 13). A total of seven CSY genes were identified in N. sibirica, while only two or three genes were in the other Sapindales species (Fig. 3A; Supplementary Table 22). Transcriptomic analysis showed that the expression of three CSY genes was induced by salt treatment (Supplementary Fig. 13).

3.5. Salt-stressed gene expressions based on transcriptomes

We generated RNA-seq data sets across various tissues under different salt concentration treatments. We analyzed RNA from roots, stems, and leaves of N. sibirica seedlings, which were first grown in vermiculite medium (watered 1/2 Hoagland’s solution) for two months. Then seedlings were supplemented for one week with moderate (200 mM, termed as T200) and high (400 mM, termed as T400) concentrations of NaCl, respectively, to identify genes for salinity responses (see Materials and Methods). Seedlings grown continuously in vermiculite medium poured with 1/2 Hoagland’s solution were used as control. With three replicates per sample, 27 transcriptomes were generated to calculate expression profiles. Over 92.3% of quality-filtered reads were properly mapped to the N. sibirica genome with strong transcriptome correlation across tissues (Supplementary Table 4; Supplementary Fig. 14), indicating our data are viable for subsequent analyses.

Expression tendency analyses (see Materials and Methods) of genes in each tissue (root, stem, and leaf) under different salt concentrations (control, T200, and T400) yielded eight tightly grouped clusters, which display co-responsive expression patterns in each condition (Fig. 4A). Consistently up-regulated or down-regulated gene clusters in response to increased salt concentrations treatment are particularly interesting. For example, the expression of each gene in cluster_8 (including 596 genes in leaves, 554 genes in roots, and 107 genes in stems) consistently increased with increasing salt concentration (Fig. 4A). KEGG pathway enrichment analyses revealed that genes of cluster_8 in leaves were enriched in photosynthesis-related pathways (e.g. porphyrin and chlorophyll metabolism, photosynthesis—antenna proteins, and photosynthesis; Fig. 4B), indicating that photosynthesis was enhanced with the increase of salt concentration. In roots, 554 consistently up-regulated genes of cluster_8 were functionally enriched in cutin, suberine and wax biosynthesis pathways (Fig. 4B), which probably are involved in the cuticular wax formation and maintenance of cellular structural stability under high osmotic pressure.

Expression tendency and pathway enrichment analysis of N. sibirica under salt stress. A. Cluster analysis of DEGs displaying a log2-fold change (with absolute value > 2) of transcripts under moderate (200 mM) and high (400 mM) concentration NaCl stress in leaf, root, and stem. The comparisons include control (CK), 200 mM versus control (T200) and 400 mM versus control (T400). B. Bubble diagram of significant enrichment pathways during salt stress. The colour of the circles represents different clusters, and the size of the circles indicates the –log10 correct P-values (correct P-value ≤ 0.05) of significant enrichment pathways in each cluster. L: leaf, R: root, and S: stem.
Figure 4.

Expression tendency and pathway enrichment analysis of N. sibirica under salt stress. A. Cluster analysis of DEGs displaying a log2-fold change (with absolute value > 2) of transcripts under moderate (200 mM) and high (400 mM) concentration NaCl stress in leaf, root, and stem. The comparisons include control (CK), 200 mM versus control (T200) and 400 mM versus control (T400). B. Bubble diagram of significant enrichment pathways during salt stress. The colour of the circles represents different clusters, and the size of the circles indicates the –log10 correct P-values (correct P-value ≤ 0.05) of significant enrichment pathways in each cluster. L: leaf, R: root, and S: stem.

In contrast, genes involved in plant-pathogen interaction displayed down-regulated expression patterns in both Cluster_1 and Cluster_2 as salt concentration increased (Fig. 4), suggesting this type of biological process is weakened under salt stress. From Clusters 5–8 displaying increased expression under either T200 or T400 or both conditions, we observed co-enrichment of phenylpropanoid biosynthesis and phenylalanine metabolism pathways (Fig. 4B). It has been proposed that phenylpropanoids contributed to plant response to biotic and abiotic stresses from various aspects.80 More importantly, metabolic pathways and biosynthesis of secondary metabolites are the most enriched pathways in almost all clusters (Fig. 4B), suggesting active metabolic levels are crucial for salt stress tolerance.

The salt overly sensitive (SOS) pathway, including SOS3, SCaBP8, SOS2, and SOS1, is critical for maintaining cellular ion homeostasis through regulating Na+ exclusion.1,81 Therefore, we analyzed the expression changes of related gene members under salt stress in N. sibirica (Supplementary Fig. 15). SOS1 is a key member in the SOS pathway that squeezes excess Na+ out of the cell.81 The expression of SOS1 slightly decreased at a moderate salt concentration (T200) but significantly increased at a high salt concentration (T400) (Supplementary Fig. 15) in both roots and leaves, suggesting the enhanced Na+ exclusion under high salt conditions. In addition to Na+ exclusion, vacuolar Na+ compartmentalization, mainly mediated by NXH1, a vital Na+/H+ transporter, is another mechanism to reduce excess Na+ in the cytoplasm.1,2,79 We found that the expression of NXH1 increased under salt treatments and maintained at high levels in both T200 and T400 treatments (Supplementary Fig. 15).

3.6. Key transcription factors involved in response to salt stress and tissue- and concentration-dependent signalling

As TFs play central roles in triggering the expression of a series of salt-responsive genes, we analyzed the differentially expressed TFs (Fig. 5A) and their associated regulatory networks (Fig. 5B and C) in the above-identified DEGs. Among the 14 differentially expressed TF families identified, AP2/ERF-AP2 family has the highest number of TF members, all of which were significantly up-regulated, indicating the important role of salt tolerance of this family (Fig. 5A). For example, the expression level of NsiCBF4, a regulator of drought response in Arabidopsis,82 was remarkably increased in leaves (8.02-fold at T200) and roots (9.78-fold at T200 and 7.51-fold at T400). Elevated expressions were also found in NsiERF109, a factor involved in retarding programmed cell death under salt stress,83 and NsiERF040, a member involved in generating thickened cell walls.84 Several TFs related to abscisic acids (ABA), such as NsiERF030, NsiWRKY33, and NsiWRKY46, were up-regulated under salt stress.

Differentially expressed transcription factors and their associated regulatory networks of N. sibirica under salt stress. A. Statistics of differentially expressed transcription factors in leaves and roots under moderate (200 mM) and high (400 mM) concentrations of NaCl. Dark red: fold change ≥ 6, light red: 1 ≤ fold change < 6, blue: 0 ≤ fold change < 1. The seven positively selected TFs are marked with asterisks in the upper right of the gene names. B. Sub-network for black module transcription factor expression at moderate (200 mM) and high (400 mM) concentrations of NaCl. Transcription factors inside the circles are expressed in both roots and leaves. The three transcription factors above the circles are expressed only in leaves, and the two transcription factors below the circles are expressed only in roots. The colour of the heatmap indicates the log2-fold change value. C. Sub-network for red module transcription factor expression at moderate (200 mM) and high (400 mM) concentrations of NaCl. Transcription factors inside the circles are expressed in both roots and leaves. The four transcription factors above the circles are expressed only in leaves, and the two transcription factors below the circles are expressed only in roots.
Figure 5.

Differentially expressed transcription factors and their associated regulatory networks of N. sibirica under salt stress. A. Statistics of differentially expressed transcription factors in leaves and roots under moderate (200 mM) and high (400 mM) concentrations of NaCl. Dark red: fold change ≥ 6, light red: 1 ≤ fold change < 6, blue: 0 ≤ fold change < 1. The seven positively selected TFs are marked with asterisks in the upper right of the gene names. B. Sub-network for black module transcription factor expression at moderate (200 mM) and high (400 mM) concentrations of NaCl. Transcription factors inside the circles are expressed in both roots and leaves. The three transcription factors above the circles are expressed only in leaves, and the two transcription factors below the circles are expressed only in roots. The colour of the heatmap indicates the log2-fold change value. C. Sub-network for red module transcription factor expression at moderate (200 mM) and high (400 mM) concentrations of NaCl. Transcription factors inside the circles are expressed in both roots and leaves. The four transcription factors above the circles are expressed only in leaves, and the two transcription factors below the circles are expressed only in roots.

In addition, we found that seven differentially expressed TFs in N. sibirica bears the signatures of positive selection (Fig. 5B), compared with the other Sapindales species distributed in subtropical and tropical environments (see Materials and Methods). These include three functionally related to abiotic stress response (NsiSZF1, NsiSCL13, and NsiWRKY40) and four unrelated or unknown function TFs (NsiNAC036, NsiDof1, NsiHB16, and NsiMYB). By investigating the expression levels of these TFs’ orthologs in a salt-sensitive species of Sapindales, we found most of these TFs did not show differential expression under salt stress (Supplementary Table 23) except NsiSZF1 and NsiWRKY40, functionally involved in response to salt and wounding, respectively. These contrasting expression patterns provide critical evidence for the rapid evolution of TFs confer salinity adaptation of Nitrariaceae.

Moreover, we observed a significant difference in differentially expressed TFs between T200 and T400 treatments in leaves, with 21 significantly up-regulated and three TFs down-regulated under T200 treatment, while only seven TFs down-regulated and no TFs up-regulated under T400 treatment, suggesting concentration-dependent signalling in response to salt stress. In roots, the number of differentially expressed TFs was similar between T200 (13 TFs) and T400 (14 TFs), with some TFs especially expressed in T200 or T400 treatments (Fig. 5A). Weighted gene co-expression network analysis (WGCNA) further revealed that the most differentially expressed TFs were in the black and red modules (Supplementary Fig. 16). The expression of key transcription factors and their associated regulatory networks displayed tissue- and concentration-specific patterns. For example, HHO3, WRKY33, and ERF041 in the black module, bHLH35, PAT1, SCL13, and bZIP55 in the red module were only expressed in the leaves, while NAC036, ERF030 in the black module, and MYB77, ERF3 in the red module were only expressed in the roots (Fig. 5B and C). More interestingly, some TFs showed differences in expression patterns in different tissues in response to different salt treatments; for example, bHLH25, WEKY53, and WRKY46 were only expressed at a high salt concentration (T400) in roots, but at a low salt concentration (T200) in leaves, and ERF109, CBF4, SZF1, and ERF040 were expressed at both low and high salt concentrations in roots, but only at low salt concentrations in leaves (Fig. 5B and C). These results suggest that different tissues of N. sibirica may have different regulatory mechanisms in response to various salt concentrations.

4. Discussion

In the ‘published plant genome’ database (https://www.plabipd.de/plant_genomes_pa.ep, last accessed on 1 February 2023), 141 families of angiosperms have published genome data, which account for only about 1/3 of the total families of angiosperms. In this study, we presented the whole genome sequence of N. sibirica, which is the first chromosomal-level genome in the family Nitrariaceae. BUSCO, CEGMA, and LAI assessments all indicated that the N. sibirica genome reached ‘reference genome’ quality (Supplementary Tables 11–13). This genome provides a valuable resource for filling genomic gaps in flowering plants at the family level. Based on phylogenomic analyses, we confirmed that Nitrariaceae is part of Sapindales14,15,85–87 rather than Zygophyllales.88 Our results also suggested that N. sibirica is the sister to all other sampled representative species of Sapindales occurring in non-arid environments (Fig. 2), providing a basis for their genomic analyses.

All species representing different lineages of Sapindales did not show WGDs (Fig. 2A and B), suggesting that polyploidization could not account for genomic changes for salinity adaptation in N. sibirica. Interestingly, many gene families decrease their copies in N. sibirica. This is also found in salt-tolerant mangroves.89 However, many gene families involved in salt stress expanded in N. sibirica, including genes involved in potassium ion transport, TCA cycle, and nicotianamine biosynthesis, similar to those genomic changes in other salt-tolerant species.7,12,89 These significantly expanded genes functioned at multiple levels in response to salt stress. For example, the tandem duplications of HAK5 gene (six copies; Figs. 3B and 6) in N. sibirica (only 2–3 copies in other Sapindales species) may partly explain the superior K+ retention ability of N. sibirica roots as revealed by previous physiological examination.26,27 Furthermore, the tremendous expansion and synchronized expression of NAS genes may have enhanced nicotianmine biosynthesis for sufficient Fe uptake to maintain normal growth in saline-alkali soils (Fig. 3 and Supplementary Fig. 12).76 Our results highlight the importance of gene expansion in salt stress adaptations in this and other halophytes.

Adaptation to the saline-alkali environment—supportive mechanisms in N. sibirica. From the aboveground part (filled with light blue), increased photosynthesis and energy supply were identified under salt stress in N. sibirica (see ‘Results’). These enhancements could provide more organic matter and energy to regulate intracellular environmental homeostasis and maintain normal growth and development. Moreover, vacuoles of the succulent leaves acted as the main Na+ sink to store excess sodium ions through the expression of the NHX1 transporter. In the belowground part (filled with light brown), the increased copy number of HAK5 genes might confer a superior K+ retention ability of N. sibirica. The expansion of nicotianmine synthase (NAS) genes may have enhanced nicotianmine (NA) biosynthesis for sufficient Fe uptake in saline–alkali soils. The expression of SOS pathway members suggests vacuolar Na+ compartmentalization is an ongoing process under salt treatment while Na+ exclusion tends to function at high salt concentrations. In addition, the expression of vital transcription factors (e.g. ERF, NAC, and bHLH) also plays important roles for N. sibirica to cope with salt stress.
Figure 6.

Adaptation to the saline-alkali environment—supportive mechanisms in N. sibirica. From the aboveground part (filled with light blue), increased photosynthesis and energy supply were identified under salt stress in N. sibirica (see ‘Results’). These enhancements could provide more organic matter and energy to regulate intracellular environmental homeostasis and maintain normal growth and development. Moreover, vacuoles of the succulent leaves acted as the main Na+ sink to store excess sodium ions through the expression of the NHX1 transporter. In the belowground part (filled with light brown), the increased copy number of HAK5 genes might confer a superior K+ retention ability of N. sibirica. The expansion of nicotianmine synthase (NAS) genes may have enhanced nicotianmine (NA) biosynthesis for sufficient Fe uptake in saline–alkali soils. The expression of SOS pathway members suggests vacuolar Na+ compartmentalization is an ongoing process under salt treatment while Na+ exclusion tends to function at high salt concentrations. In addition, the expression of vital transcription factors (e.g. ERF, NAC, and bHLH) also plays important roles for N. sibirica to cope with salt stress.

We further found that many genes involved in salt stress changed their expressions. Analysis of a series of transcriptomes revealed that genes in photosynthesis-related pathways consistently increased expression with increased salt concentrations (Fig. 4B), suggesting the enhanced photosynthesis ability of N. sibirica under salt stress (Fig. 6). This phenomenon is contrasted with the observation in crops. For example, salt would lead to toxic effects on photosynthesis-related components and reduced growth rates in barley.90 In addition to increasing photosynthetic carbon fixation, the expansion and synchronized expression of CSY genes, a key rate-limiting enzyme in the TCA cycle, may also enhance the energy supply of N. sibirica under salt stress (Figs. 3A and 6). These enhancements are partly due to the enormous energy consumption required to regulate intracellular environmental homeostasis and to maintain normal growth and development. This finding is supported by previous metabolomic evidence.24

Furthermore, we found up-regulated genes related to cuticular wax formation in roots (Fig. 4B), which is favourable for cellular structural stability against high osmotic pressure caused by salt stress. Preventing the accumulation of excess Na+ in the cytoplasm through Na+ exclusion and vacuolar Na+ compartmentalization is critical for maintaining ion homeostasis under salt stress.1–3 Analyzing the expression patterns of core genes involved in these two processes indicated that enhanced Na+ exclusion occurred under high salt stress, in addition to vacuolar Na+ compartmentalization, a process likely play roles in both mild and high salt stresses (Fig. 6).

Moreover, many TFs related to salinity adaptations changed their expression significantly under salt stress (Figs. 5A and 6). These TF genes are mainly from the AP2/ERF-AP2, bHLH, bZIP, C2H2, C3H, GRAS, MYB, NAC, and WRKY families, of which AP2/ERF-AP2 is likely of particular importance for salt tolerance in terms of the number and degree of differentially expressed TFs. Because of the regulatory roles of TF, the expression changes of these genes could largely initiate the responses of a series of downstream genes.91 By analyzing the expression pattern of differentially expressed TFs and their associated regulatory networks, we found tissue- and concentration-dependent signalling in response to different salt concentrations across various tissues (Fig. 5B and C). Importantly, we identified that seven of these TFs experienced rapid evolution by positive selection compared with other salt-sensitive Sapindales species (see Results), potentially contributing to expression changes and salt tolerance in N. sibirica. Such adaptive divergence of sequences within genes participating in salt adaptation was also reported in a desert poplar.7

In summary, our results suggest that the integration of multiple genomic and transcriptomic changes might have together contributed to the high salt tolerance of N. sibirica (as summarized in Fig. 6). From the aboveground part, increased photosynthesis and energy supply could provide more organic matter and energy to regulate intracellular environmental homeostasis and maintain normal growth and development. In the belowground part, the increased ability for K+ retention and Fe uptake through the expansion of key genes are crucial for maintaining cellular ion homeostasis and chlorophyll synthesis, respectively. Moreover, vacuoles of the succulent leaves acted as the main Na+ sink to store excess sodium ions. In addition to vacuolar Na+ compartmentalization, under a high level of salt stress, Na+ exclusion also play a significant role (Fig. 6). In addition, the expression of vital transcription factors (e.g. ERF, NAC, and bHLH) and their associated regulatory networks also play essential roles for N. sibirica to cope with salt stress. Together, our findings shed new light on the genetic bases of salt tolerance in N. sibirica, enhanced our understanding of salinity adaptation in halophytes, and pave the way for breeding salt-tolerant crops in the future.

Acknowledgements

We received support for computational work from the Big Data Computing Platform for Western Ecological Environment and Regional Development and the Supercomputing Center of Lanzhou University. We would like to thank Dr Xinxing Fu for her helpful discussions of this work and Mr Weidong Tang for his suggestions on plant material collection.

Funding

This work was supported by the National Natural Science Foundation of China (grant No. 32100170) and Fundamental Research Funds for the Central Universities (lzujbky-2021-49).

Conflict of Interest

The authors declare that they have no competing interests.

Data availability

The genomic and transcriptomic sequencing data were deposited in the National Center for Biotechnology Information (NCBI; https://www.ncbi.nlm.nih.gov/) under the BioProject accession number PRJNA939936. The genome assembly and annotation files are available at the public figshare database under the link https://figshare.com/search?q=10.6084%2Fm9.figshare.22490920.

References

1.

Yang
,
Y.Q.
and
Guo
,
Y.
2018
,
Elucidating the molecular mechanisms mediating plant salt-stress responses
,
New Phytol.
,
217
,
523
39
.

2.

van Zelm
,
E.
,
Zhang
,
Y.X.
, and
Testerink
,
C.
2020
,
Salt tolerance mechanisms of plants
,
Annu. Rev. Plant Biol.
,
71
,
403
33
.

3.

Zhao
,
C.Z.
,
Zhang
,
H.
,
Song
,
C.P.
,
Zhu
,
J.K.
, and
Shabala
,
S.
2020
,
Mechanisms of plant responses and adaptation to soil salinity
,
Innovation
,
1
,
100017
.

4.

Zhu
,
J.K.
2001
,
Plant salt tolerance
,
Trends Plant Sci.
,
6
,
66
71
.

5.

Møller
,
I.S.
and
Tester
,
M.
2007
,
Salinity tolerance of Arabidopsis: a good model for cereals?
,
Trends Plant Sci.
,
12
,
534
40
.

6.

Dassanayake
,
M.
,
Oh
,
D.H.
,
Haas
,
J.S.
, et al.
2011
,
The genome of the extremophile crucifer Thellungiella parvula
,
Nat. Genet.
,
43
,
913
8
.

7.

Ma
,
T.
,
Wang
,
J.Y.
,
Zhou
,
G.K.
, et al.
2013
,
Genomic insights into salt adaptation in a desert poplar
,
Nat. Commun.
,
4
,
2797
.

8.

Zou
,
C.S.
,
Chen
,
A.
,
Xiao
,
L.H.
, et al.
2017
,
A high-quality genome assembly of quinoa provides insights into the molecular basis of salt bladder-based salinity tolerance and the exceptional nutritional value
,
Cell Res.
,
27
,
1327
40
.

9.

Zeng
,
L.
,
Tu
,
X.L.
,
Dai
,
H.
, et al.
2019
,
Whole genomes and transcriptomes reveal adaptation and domestication of pistachio
,
Genome Biol.
,
20
,
79
.

10.

Zhang
,
W.T.
,
Liu
,
J.
,
Zhang
,
Y.X.
, et al.
2020
,
A high-quality genome sequence of alkaligrass provides insights into halophyte stress tolerance
,
Sci. China Life Sci.
,
63
,
1269
82
.

11.

Feng
,
X.
,
Li
,
G.H.
,
Xu
,
S.H.
, et al.
2021
,
Genomic insights into molecular adaptation to intertidal environments in the mangrove Aegiceras corniculatum
,
New Phytol.
,
231
,
2346
58
.

12.

Ren
,
G.P.
,
Jiang
,
Y.Y.
,
Li
,
A.
, et al.
2022
,
The genome sequence provides insights into salt tolerance of Achnatherum splendens (Gramineae), a constructive species of alkaline grassland
,
Plant Biotechnol. J.
,
20
,
116
28
.

13.

Yuan
,
F.
,
Wang
,
X.
,
Zhao
,
B.Q.
, et al.
2022
,
The genome of the recretohalophyte Limonium bicolor provides insights into salt gland development and salinity adaptation during terrestrial evolution
,
Mol. Plant.
,
15
,
1024
44
.

14.

Temirbayeva
,
K.
and
Zhang
,
M.L.
2015
,
Molecular phylogenetic and biogeographical analysis of Nitraria based on nuclear and chloroplast DNA sequences
,
Plant Syst Evol.
,
301
,
1897
906
.

15.

APG, III.
2009
,
An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG III
,
Bot. J. Linn. Soc.
,
161
,
105
21
.

16.

Bachelier
J.
,
Endress
,
P.
, and
Craene
,
L.
2011
, In:
Wanntorp
,
L.
, and
De Craene
,
L.R.
(eds.)
Flowers on the tree of life
(Systematics Association Special Volume Series),
Cambridge:
Cambridge University Press
, p.
181
217
.

17.

Zhang
,
M.L.
,
Temirbayeva
,
K.
,
Sanderson
,
S.
, et al.
2015
,
Young dispersal of xerophil Nitraria lineages in intercontinental disjunctions of the Old World
,
Sci. Rep.
,
5
,
13840
.

18.

Du
,
Q.H.
,
Xin
,
H.L.
, and
Peng
,
C.
2015
,
Pharmacology and phytochemistry of the Nitraria genus (Review)
,
Mol. Med. Rep.
,
11
,
11
20
.

19.

Tölke
,
E.D.
,
Medina
,
M.C.
,
Souto
,
A.L.
, et al.
2022
,
Diversity and evolution of secretory structures in Sapindales
,
Braz. J. Bot.
,
45
,
251
79
.

20.

Wu
,
G.A.
,
Prochnik
,
S.
,
Jenkins
,
J.
, et al.
2014
,
Sequencing of diverse mandarin, pummelo and orange genomes reveals complex history of admixture during citrus domestication
,
Nat. Biotechnol.
,
32
,
656
62
.

21.

Yang
,
J.
,
Wariss
,
H.M.
,
Tao
,
L.
, et al.
2019
,
De novo genome assembly of the endangered Acer yangbiense, a plant species with extremely small populations endemic to Yunnan Province, China
,
GigaScience
,
104
,
1215
32
.

22.

Li
,
Q.H.
,
Wang
,
S.X.
,
Xu
,
J.
,
Ren
,
W.J.
, and
Zhao
,
Y.M.
2012
,
Comprehensive evaluation on salt tolerance of different desert shrubs in Ulan Buh desert region
,
Pratacultural Sci.
,
29
,
1132
6
.

23.

Ni
,
J.W.
,
Wu
,
X.
,
Zhang
,
H.X.
,
Liu
,
T.
, and
Zhang
,
L.
2012
,
Comparative analysis of salt tolerance of three Nitraria species
,
Forest Res.
,
25
,
48
53
.

24.

Li
,
H.Y.
,
Tang
,
X.Q.
,
Yang
,
X.Y.
, and
Zhang
,
H.X.
2021
,
Comprehensive transcriptome and metabolome profiling reveal metabolic mechanisms of Nitraria sibirica Pall. to salt stress
,
Sci. Rep.
,
11
,
12878
.

25.

Li
,
H.Y.
,
Tang
,
X.Q.
,
Zhu
,
J.F.
,
Yang
,
X.Y.
, and
Zhang
,
H.X.
2017
,
De novo transcriptome characterization, gene expression profiling and ionic responses of Nitraria sibirica Pall. under Salt Stress
,
Forests
,
8
,
211
.

26.

Tang
,
X.Q.
,
Zhang
,
H.L.
,
Shabala
,
S.
,
Li
,
H.Y.
,
Yang
,
X.Y.
, and
Zhang
,
H.X.
2020
,
Tissue tolerance mechanisms conferring salinity tolerance in a halophytic perennial species Nitraria sibirica Pall
,
Tree Physiol.
,
41
,
1264
77
.

27.

Cheng
,
T.L.
,
Li
,
H.Y.
,
Wu
,
H.L.
, et al.
2015
,
Comparison on osmotica accumulation of different salt-tolerant plants under Salt Stress
,
Forest Res.
,
28
,
826
32
.

28.

Wu
,
S.D.
,
Han
,
B.
, and
Jiao
,
Y.N.
2020
,
Genetic contribution of paleopolyploidy to adaptive evolution in angiosperms
,
Mol. Plant
,
13
,
59
71
.

29.

Van de Peer
,
Y.
,
Ashman
,
T.L.
,
Soltis
,
P.S.
, and
Soltis
,
D.E.
2021
,
Polyploidy: an evolutionary and ecological force in stressful times
,
Plant Cell
,
33
,
11
26
.

30.

Lieberman-Aiden
,
E.
,
van Berkum
,
N.L.
,
Williams
,
L.
, et al.
2009
,
Comprehensive mapping of long-range interactions reveals folding principles of the human genome
,
Science
,
326
,
289
93
.

31.

Luo
,
R.B.
,
Liu
,
B.H.
,
Xie
,
Y.L.
, et al.
2012
,
SOAPdenovo2: an empirically improved memory-efficient short-read de novo assembler
,
GigaScience
,
1
,
18
.

32.

Vaser
,
R.
,
Sović
,
I.
,
Nagarajan
,
N.
, and
Šikić
,
M.
2017
,
Fast and accurate de novo genome assembly from long uncorrected reads
,
Genome Res.
,
27
,
737
46
.

33.

Hu
,
J.
,
Fan
,
J.P.
,
Sun
,
Z.Y.
, and
Liu
,
S.L.
2020
,
NextPolish: a fast and efficient genome polishing tool for long-read assembly
,
Bioinformatics
,
36
,
2253
5
.

34.

Servant
,
N.
,
Varoquaux
,
N.
,
Lajoie
,
B.R.
, et al.
2015
,
HiC-Pro: an optimized and flexible pipeline for Hi-C data processing
,
Genome Biol.
,
16
,
259
.

35.

Burton
,
J.N.
,
Adey
,
A.
,
Patwardhan
,
R.P.
,
Qiu
,
R.
,
Kitzman
,
J.O.
, and
Shendure
,
J.
2013
,
Chromosome-scale scaffolding of de novo genome assemblies based on chromatin interactions
,
Nat. Biotechnol.
,
31
,
1119
25
.

36.

Simão
,
F.A.
,
Waterhouse
,
R.M.
,
Ioannidis
,
P.
,
Kriventseva
,
E.V.
, and
Zdobnov
,
E.M.
2015
,
BUSCO: assessing genome assembly and ­annotation completeness with single-copy orthologs
,
Bioinformatics
,
31
,
3210
2
.

37.

Parra
,
G.
,
Bradnam
,
K.
, and
Korf
,
I.
2007
,
CEGMA: a pipeline to accurately annotate core genes in eukaryotic genomes
,
Bioinformatics
,
23
,
1061
7
.

38.

Li
,
H.
and
Durbin
,
R.
2010
,
Fast and accurate long-read alignment with Burrows–Wheeler transform
,
Bioinformatics
,
26
,
589
95
.

39.

Li
,
H.
,
Handsaker
,
B.
,
Wysoker
,
A.
, et al. ;
1000 Genome Project Data Processing Subgroup.
2009
,
The sequence alignment/map format and SAMtools
,
Bioinformatics
,
25
,
2078
9
.

40.

Danecek
,
P.
and
McCarthy
,
S.A.
2017
,
BCFtools/csq: haplotype-aware variant consequences
,
Bioinformatics
,
33
,
2037
9
.

41.

Wang
,
X.W.
and
Wang
,
L.
2016
,
GMATA: an integrated software package for genome-scale SSR mining, marker development and viewing
,
Front. Plant Sci.
,
7
,
1350
.

42.

Benson
,
G.
1999
,
Tandem repeats finder: a program to analyze DNA sequences
,
Nucleic Acids Res.
,
27
,
573
80
.

43.

Han
,
Y.
and
Wessler
,
S.R.
2010
,
MITE-Hunter: a program for discovering miniature inverted-repeat transposable elements from genomic sequences
,
Nucleic Acids Res.
,
38
,
e199
.

44.

Bedell
,
J.A.
,
Korf
,
I.
, and
Gish
,
W.
2000
,
MaskerAid: a performance enhancement to RepeatMasker
,
Bioinformatics
,
16
,
1040
1
.

45.

Jurka
,
J.
,
Kapitonov
,
V.V.
,
Pavlicek
,
A.
,
Klonowski
,
P.
,
Kohany
,
O.
, and
Walichiewicz
,
J.
2005
,
Repbase Update, a database of eukaryotic repetitive elements
,
Cytogenet. Genome Res.
,
110
,
462
7
.

46.

Lowe
,
T.M.
and
Eddy
,
S.R.
1997
,
tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence
,
Nucleic Acids Res.
,
25
,
955
64
.

47.

Nawrocki
,
E.P.
and
Eddy
,
S.R.
2013
,
Infernal 1.1: 100-fold faster RNA homology searches
,
Bioinformatics
,
29
,
2933
5
.

48.

Lagesen
,
K.
,
Hallin
,
P.
,
Rødland
,
E.A.
,
Staerfeldt
,
H.H.
,
Rognes
,
T.
, and
Ussery
,
D.W.
2007
,
RNAmmer: consistent and rapid annotation of ribosomal RNA genes
,
Nucleic Acids Res.
,
35
,
3100
8
.

49.

Keilwagen
,
J.
,
Wenk
,
M.
,
Erickson
,
J.L.
,
Schattat
,
M.H.
,
Grau
,
J.
, and
Hartung
,
F.
2016
,
Using intron position conservation for homology-based gene prediction
,
Nucleic Acids Res.
,
44
,
e89
.

50.

Stanke
,
M.
,
Diekhans
,
M.
,
Baertsch
,
R.
, and
Haussler
,
D.
2008
,
Using native and syntenically mapped cDNA alignments to improve de novo gene finding
,
Bioinformatics
,
24
,
637
44
.

51.

Dobin
,
A.
,
Davis
,
C.A.
,
Schlesinger
,
F.
, et al.
2013
,
STAR: ultrafast universal RNA-seq aligner
,
Bioinformatics
,
29
,
15
21
.

52.

Kovaka
,
S.
,
Zimin
,
A.V.
,
Pertea
,
G.M.
,
Razaghi
,
R.
,
Salzberg
,
S.L.
, and
Pertea
,
M.
2019
,
Transcriptome assembly from long-read RNA-seq alignments with StringTie2
,
Genome Biol.
,
20
,
278
.

53.

Haas
,
B.J.
,
Salzberg
,
S.L.
,
Zhu
,
W.
, et al.
2008
,
Automated eukaryotic gene structure annotation using EVidenceModeler and the program to assemble spliced alignments
,
Genome Biol.
,
9
,
R7
.

54.

Zdobnov
,
E.M.
and
Apweiler
,
R.
2001
,
InterProScan—an integration platform for the signature-recognition methods in InterPro
,
Bioinformatics
,
17
,
847
8
.

55.

Conesa
,
A.
,
Götz
,
S.
,
García-Gómez
,
J.M.
,
Terol
,
J.
,
Talón
,
M.
, and
Robles
,
M.
2005
,
Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research
,
Bioinformatics
,
21
,
3674
6
.

56.

Emms
,
D.M.
and
Kelly
,
S.
2015
,
OrthoFinder: solving fundamental biases in whole genome comparisons dramatically improves orthogroup inference accuracy
,
Genome Biol.
,
16
,
157
.

57.

Katoh
,
K.
and
Standley
,
D.M.
2013
,
MAFFT multiple sequence alignment software version 7: improvements in performance and usability
,
Mol. Biol. Evol.
,
30
,
772
80
.

58.

Suyama
,
M.
,
Torrents
,
D.
, and
Bork
,
P.
2006
,
PAL2NAL: robust conversion of protein sequence alignments into the corresponding codon alignments
,
Nucleic Acids Res.
,
34
,
W609
12
.

59.

Nguyen
,
L.T.
,
Schmidt
,
H.A.
,
von Haeseler
,
A.
, and
Minh
,
B.Q.
2015
,
IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies
,
Mol. Biol. Evol.
,
32
,
268
74
.

60.

Mirarab
,
S.
and
Warnow
,
T.
2015
,
ASTRAL-II: coalescent-based species tree estimation with many hundreds of taxa and thousands of genes
,
Bioinformatics
,
31
,
i44
52
.

61.

Jin
,
J.J.
,
Yu
,
W.B.
,
Yang
,
J.B.
, et al.
2020
,
GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes
,
Genome Biol.
,
21
,
241
.

62.

Shi
,
L.C.
,
Chen
,
H.M.
,
Jiang
,
M.
, et al.
2019
,
CPGAVAS2, an integrated plastome sequence annotator and analyzer
,
Nucleic Acids Res.
,
47
,
W65
73
.

63.

Yang
,
Z.H.
2007
,
PAML 4: phylogenetic analysis by maximum likelihood
,
Mol. Biol. Evol.
,
24
,
1586
91
.

64.

Han
,
M.V.
,
Thomas
,
G.W.
,
Lugo-Martinez
,
J.
, and
Hahn
,
M.W.
2013
,
Estimating gene gain and loss rates in the presence of error in genome assembly and annotation using CAFE 3
,
Mol. Biol. Evol.
,
30
,
1987
97
.

65.

Chen
,
C.J.
,
Chen
,
H.
,
Zhang
,
Y.
, et al.
2020
,
TBtools: an integrative toolkit developed for interactive analyses of big biological data
,
Mol. Plant
,
13
,
1194
202
.

66.

Nei
,
M.
and
Gojobori
,
T.
1986
,
Simple methods for estimating the numbers of synonymous and nonsynonymous nucleotide substitutions
,
Mol. Biol. Evol.
,
3
,
418
26
.

67.

Sun
,
P.C.
,
Jiao
,
B.B.
,
Yang
,
Y.Z.
, et al.
2022
,
WGDI: A user-friendly toolkit for evolutionary analyses of whole-genome duplications and ancestral karyotypes
,
Mol. Plant
,
15
,
1841
51
.

68.

Johnson
,
L.S.
,
Eddy
,
S.R.
, and
Portugaly
,
E.
2010
,
Hidden Markov model speed heuristic and iterative HMM search procedure
,
BMC Bioinf.
,
11
,
431
.

69.

Chen
,
S.
,
Zhou
,
Y.Q.
,
Chen
,
Y.
, and
Gu
,
J.
2018
,
fastp: an ultra-fast all-in-one FASTQ preprocessor
,
Bioinformatics
,
34
,
i884
90
.

70.

Kim
,
D.
,
Langmead
,
B.
, and
Salzberg
,
S.L.
2015
,
HISAT: a fast spliced aligner with low memory requirements
,
Nat. Methods
,
12
,
357
60
.

71.

Love
,
M.I.
,
Huber
,
W.
, and
Anders
,
S.
2014
,
Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2
,
Genome Biol.
,
15
,
550
.

72.

Langfelder
,
P.
and
Horvath
,
S.
2008
,
WGCNA: an R package for weighted correlation network analysis
,
BMC Bioinf.
,
9
,
559
.

73.

Smoot
,
M.E.
,
Ono
,
K.
,
Ruscheinski
,
J.
,
Wang
,
P.L.
, and
Ideker
,
T.
2011
,
Cytoscape 2.8: new features for data integration and network visualization
,
Bioinformatics
,
27
,
431
2
.

74.

Pan
,
X.Y.
,
Wei
,
X.P.
,
Wei
,
Q.S.
,
Chen
,
J.K.
, and
Wang
,
G.X.
2003
,
Polyploidy: classification, evolution and applied perspective of the genus Nitraria L
,
Chin. Bull. Bot.
,
20
,
632
8
.

75.

Ou
,
S.J.
and
Jiang
,
N.
2018
,
LTR_retriever: a highly accurate and sensitive program for identification of long terminal repeat retrotransposons
,
Plant Physiol.
,
176
,
1410
22
.

76.

Jiao
,
Y.
,
Leebens-Mack
,
J.
,
Ayyampalayam
,
S.
, et al.
2012
,
A genome triplication associated with early diversification of the core eudicots
,
Genome Biol.
,
13
,
R3
.

77.

Wang
,
Y.
and
Wu
,
W.H.
2013
,
Potassium transport and signaling in higher plants
,
Annu. Rev. Plant Biol.
,
64
,
451
76
.

78.

Nozoye
,
T.
2018
,
The nicotianamine synthase gene is a useful candidate for improving the nutritional qualities and Fe-deficiency tolerance of various crops
,
Front. Plant Sci.
,
9
,
340
.

79.

Akram
,
M.
2014
,
Citric acid cycle and role of its intermediates in metabolism
,
Cell Biochem. Biophys.
,
68
,
475
8
.

80.

Dong
,
N.Q.
and
Lin
,
H.X.
2021
,
Contribution of phenylpropanoid metabolism to plant development and plants–environment interactions
,
J. Integr. Plant Biol.
,
63
,
180
209
.

81.

Ji
,
H.T.
,
Pardo
,
J.M.
,
Batelli
,
G.
,
Van Oosten
,
M.J.
,
Bressan
,
R.A.
, and
Li
,
X.
2013
,
The Salt Overly Sensitive (SOS) pathway: established and emerging roles
,
Mol. Plant
,
6
,
275
86
.

82.

Haake
,
V.
,
Cook
,
D.
,
Riechmann
,
J.L.
,
Pineda
,
O.
,
Thomashow
,
M.F.
, and
Zhang
,
J.Z.
2002
,
Transcription factor CBF4 is a regulator of drought adaptation in Arabidopsis
,
Plant Physiol.
,
130
,
639
48
.

83.

Bahieldin
,
A.
,
Atef
,
A.
,
Edris
,
S.
, et al.
2016
,
Ethylene responsive transcription factor ERF109 retards PCD and improves salt tolerance in plant
,
BMC Plant Biol.
,
16
,
216
.

84.

Sakamoto
,
S.
,
Somssich
,
M.
,
Nakata
,
M.T.
, et al.
2018
,
Complete substitution of a secondary cell wall with a primary cell wall in Arabidopsis
,
Nat. Plants
,
10
,
777
83
.

85.

Sheahan
,
M.C.
and
Chase
,
M.W.
1996
,
A phylogenetic analysis of Zygophyllaceae R.Br. based on morphological, anatomical and rbcL DNA sequence data
,
Bot. J. Linn. Soc.
,
122
,
279
300
.

86.

Muellner-Riehl
,
A.N.
,
Weeks
,
A.
,
Clayton
,
J.W.
, et al.
2016
,
Molecular phylogenetics and molecular clock dating of Sapindales based on plastid rbcL, atpB and trnL-trnF DNA sequences
,
Taxon
,
65
,
1019
36
.

87.

Lu
,
L.
,
Li
,
X.
,
Hao
,
Z.D.
, et al.
2018
,
Phylogenetic studies and comparative chloroplast genome analyses elucidate the basal position of halophyte Nitraria sibirica (Nitrariaceae) in the Sapindales
,
Mitochondrial DNA A DNA Mapp. Seq. Anal
,
29
,
745
55
.

88.

Engler
,
A.
1931
,
Zygophyllaceae, Rutaceae, Simaroubaceae, Burseraceae.
In:
Engler
,
A.
,
Prantl
,
K.
(eds.)
Die natürlichen Pflanzenfamilien
, 2nd edition, vol.
19a
.
Leipzig
:
Engelmann
, p.
144
184
,
187
456
.

89.

Xie
,
W.
,
Guo
,
Z.
,
Wang
,
J.
, et al.
2022
,
Evolution of woody plants to the land-sea interface—the atypical genomic features of mangroves with atypical phenotypic adaptation
,
Mol. Ecol.
,
32
,
1351
65
.

90.

Fricke
,
W.
,
Akhiyarova
,
G.
,
Veselov
,
D.
, and
Kudoyarova
,
G.
2004
,
Rapid and tissue-specific changes in ABA and in growth rate in response to salinity in barley leaves
,
J. Exp. Bot.
,
55
,
1115
23
.

91.

de Mendoza
,
A.
,
Sebé-Pedrós
,
A.
,
Šestak
,
M.S.
, et al.
2013
,
Transcription factor evolution in eukaryotes and the assembly of the regulatory toolkit in multicellular lineages
,
Proc. Natl. Acad. Sci. U.S.A.
,
110
,
E4858
66
.

Author notes

These authors contributed equally to this work.

This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial License (https://creativecommons.org/licenses/by-nc/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original work is properly cited. For commercial re-use, please contact [email protected]