Next Article in Journal
Exogenous Nitric Oxide and Silicon Applications Alleviate Water Stress in Apricots
Next Article in Special Issue
Influence of Plant Growth Regulators and Artificial Light on the Growth and Accumulation of Inulin of Dedifferentiated Chicory (Cichorium intybus L.) Callus Cells
Previous Article in Journal
Study on Cytotoxic and Genotoxic Potential of Bulgarian Rosa damascena Mill. and Rosa alba L. Hydrosols—In Vivo and In Vitro
Previous Article in Special Issue
Role of WRKY Transcription Factors in Regulation of Abiotic Stress Responses in Cotton
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genomic Analysis of LEA Genes in Carica papaya and Insight into Lineage-Specific Family Evolution in Brassicales

Hainan Key Laboratory for Biosafety Monitoring and Molecular Breeding in Off-Season Reproduction Regions, Institute of Tropical Biosciences and Biotechnology, Sanya Research Institute of Chinese Academy of Tropical Agricultural Sciences, Haikou 571101, China
*
Authors to whom correspondence should be addressed.
Life 2022, 12(9), 1453; https://doi.org/10.3390/life12091453
Submission received: 1 August 2022 / Revised: 8 September 2022 / Accepted: 15 September 2022 / Published: 19 September 2022
(This article belongs to the Special Issue Plant Biotic and Abiotic Stresses)

Abstract

:
Late embryogenesis abundant (LEA) proteins comprise a diverse superfamily involved in plant development and stress responses. This study presents a first genome-wide analysis of LEA genes in papaya (Carica papaya L., Caricaceae), an economically important tree fruit crop widely cultivated in the tropics and subtropics. A total of 28 members were identified from the papaya genome, which belong to eight families with defined Pfam domains, i.e., LEA_1 (3), LEA_2 (4), LEA_3 (5), LEA_4 (5), LEA_5 (2), LEA_6 (2), DHN (4), and SMP (3). The family numbers are comparable to those present in Ricinus communis (Euphorbiaceae, 28) and Moringa oleifera (Moringaceae, 29), but relatively less than that found in Moringa oleifera (Cleomaceae, 39) and Arabidopsis thaliana (Brassicaceae, 51), implying lineage-specific evolution in Brassicales. Indeed, best-reciprocal-hit-based sequence comparison and synteny analysis revealed the presence of 29 orthogroups, and significant gene expansion in Tarenaya and Arabidopsis was mainly contributed by whole-genome duplications that occurred sometime after their split with the papaya. Though a role of transposed duplication was also observed, tandem duplication was shown to be a key contributor in gene expansion of most species examined. Further comparative analyses of exon-intron structures and protein motifs supported fast evolution of this special superfamily, especially in Arabidopsis. Transcriptional profiling revealed diverse expression patterns of CpLEA genes over various tissues and different stages of developmental fruit. Moreover, the transcript level of most genes appeared to be significantly regulated by drought, cold, and salt stresses, corresponding to the presence of cis-acting elements associated with stress response in their promoter regions. These findings not only improve our knowledge on lineage-specific family evolution in Brassicales, but also provide valuable information for further functional analysis of LEA genes in papaya.

1. Introduction

Late embryogenesis abundant (LEA) proteins comprise a large and diverse superfamily that is widely involved in plant development as well as stress responses [1,2,3]. Since their first discovery as accumulating late in cotton (Gossypium hirsutum) embryogenesis [4,5,6], over the past four decades, LEA proteins have been found in a wide range of plants as well as bacteria, fungi, and animals [1,7]. According to sequence similarity and particular Pfam domains present, LEAs can be classified into eight main families, i.e., LEA_1 (Pfam accession number PF03760), LEA_2 (PF03168), LEA_3 (PF03242), LEA_4 (PF02987), LEA_5 (PF00477), LEA_6 (PF10714), DHN (dehydrin, PF00257), and SMP (seed maturation protein, PF04927) [3,8,9]. In the model plant arabidopsis (Arabidopsis thaliana), the presence of 51 LEA-encoding genes was reported, whereby two members (i.e., AtEM10 and AtEM17) comprise one more family named AtM without significant protein domains [2,10]. Generally, LEA proteins are extremely hydrophilic; however, some members in the LEA_2 family were shown to be hydrophobic and even have a three-dimensional structure [11]. Increasing evidence shows that the accumulation of LEA proteins is not only found in seeds, but also different vegetative tissues especially under stress conditions, e.g., high temperature, low temperature, drought, and salt [2,3,12,13]. Moreover, improved stress tolerance was also observed after overexpressing LEA genes in Escherichia coli, yeast (Saccharomyces cerevisiae), and several model plants such as tobacco (Nicotiana tabacum), arabidopsis, and rice (Oryza sativa) [14,15,16,17]. Although the exact mechanism has not been clarified, LEA proteins are able to stabilize other proteins and membrane structures during water stress [16,18].
Papaya (Carica papaya L., 2n = 18) is an important tree fruit crop that belongs to the Caricaceae family within the order Brassicales, which also includes arabidopsis as a representative in Brassicaceae, spider flower (Tarenaya hassleriana) in Cleomaceae, and horseradish tree (Moringa oleifera) in Moringaceae. Compared with the occurrence of two recent whole-genome duplications (WGDs) in both spider flower and arabidopsis, papaya and horseradish tree did not experience any additional WGD after the ancient so-called γ WGD shared by all core eudicots [19,20,21,22]. Although originated in Central America, the high nutritional value with significant vitamins and minerals in papaya fruits has prompted its wide cultivation in tropics and subtropics, e.g., India, Nigeria, Brazil, Mexico, Indonesia, and China [23]. In contrast to the considerable drought tolerance of wild relatives, commercial papaya cultivars are highly susceptible to cold and drought stresses [24,25], which frequently occur in subtropical regions such as south China. Therefore, exploring genes involved in stress responses and breeding resistant varieties in these areas are of particular importance. By taking advantage of available genome and transcriptome datasets, in this study, we would like to report a genome-wide analysis of LEA genes in papaya, which includes gene locations, exon-intron structures, sequence characteristics, evolutionary relationships, and cis-acting elements in the promoter regions, as well as gene expression patterns with a focus on fruit development and stress responses. These findings provide a global view of CpLEA genes that can facilitate further functional studies, and the comparative analysis with arabidopsis, spider flower, horseradish tree, and castor bean (Ricinus communis) contributes to our knowledge on the lineage-specific evolution of this special superfamily in Brassicales.

2. Materials and Methods

2.1. Data Retrieval and Identification of LEA Genes in Papaya, Horseradish Tree, and Spider Flower

LEA genes reported in arabidopsis and castor bean (see Table S1) were retrieved from Araport11 (https://www.arabidopsis.org/, accessed on 18 June 2022) and Phytozome v13 (https://phytozome-next.jgi.doe.gov/, accessed on 18 June 2022), respectively. Their protein sequences were used to identify homologs from papaya, horseradish tree, and spider flower, whose genome sequences were accessed from Phytozome v13, NCBI (http://www.ncbi.nlm.nih.gov/, accessed on 18 June 2022), and NGDC (https://ngdc.cncb.ac.cn/, accessed on 18 June 2022). The E-value of the tBLASTn search [26] was set to 1 × 105, and gene models of candidates were curated with available mRNAs as described before [27]. The presence of certain Pfam domains was confirmed using MOTIF Search (https://www.genome.jp/tools/motif/, accessed on 18 June 2022). Systematic names were assigned with two italic letters denoting the source organism and family name followed by a progressive number of their locations on chromosomes (Chrs) or scaffolds (Scfs).

2.2. Synteny Analysis and Gene Expansion Patterns

Homolog pairs were identified using the all-to-all BLASTP method (E-value cutoff 1 × 1010) and syntenic blocks were inferred using MCScanX (BLAST hits ≥ 5) [26,28]. Tandem repeats were defined when two paralogs were consecutive in a genome; WGD repeats were considered when duplicated genes were located in syntenic blocks of duplicated chromosomes, and transposed repeats were identified using the DupGen_finder pipeline as previously described [29]. Orthologs between different species were determined using the Best Reciprocal Hit (BRH) method [30], as well as information from synteny analysis; and orthogroups (OGs) were assigned only when they were present in at least two species examined.

2.3. Exon-Intron Structure, Phylogenetic Analysis, and Structural Characterization

The exon-intron structure was analyzed using GSDS 2.0 [31] by aligning the coding sequence (CDS) to the corresponding genomic sequence. The molecular weight (MW), theoretical isoelectric point (pI), and grand average of hydropathy (GRAVY) were calculated using ProtParam (http://web.expasy.org/protparam/, accessed on 18 June 2022), and protein subcellular localization was predicted using WoLF PSORT (http://www.genscript.com/wolf-psort.html, accessed on 18 June 2022). Multiple sequence alignment and phylogenetic reconstruction were performed using MEGA6 [32] with MUSCLE and the maximum likelihood method (bootstrap: 1000 replicates), respectively. Conserved motifs in LEA proteins were identified using MEME (v 5.4.1) [33]: any number of repetitions; maximum number of motifs, 20; minimum sites, 2; and, the optimum width of each motif, between 6 and 100 residues.

2.4. Promoter Analysis

PLACE (http://www.dna.affrc.go.jp/PLACE/, accessed on 18 June 2022) was used to examine the presence of two stress-related cis-acting elements (i.e., abscisic acid response (ABRE, ACGTG) and low temperature response (LTRE, CCGAC)) in the 2000-bp promoter region of CpLEA genes.

2.5. Plant Materials, RNA-seq, and Gene Expression Analysis

Gene expression profiles were analyzed on the basis of RNA sequencing (RNA-seq) samples as shown in Table S2. Various tissues, i.e., root, apical bud, leaf, petiole, leaf vein, male flower, female flower, fruit, peel, and seed, were collected from one-year-old hermaphrodite plants of the cultivar Zhongbai that were planted in 2019 at the Wenchang experimental base, Institute of Tropical Biosciences and Biotechnology, Chinese Academy of Tropical Agricultural Sciences (Wenchang, Hainan, China: 19°32′15.39″ N, 110°45′47.26″ E). Routine management was performed, and three groups of more than five trees were used. As for cold and salt stresses, eight-week-old plantlets were used and treatments of 4 °C low temperature (i.e., 0, 7, 21, and 40 h) and 300 mmol/L NaCl (i.e., 0, 10, 15, and 20 d) were applied. To ensure the consistency of materials, only the second leaf from the top of a plantlet was collected and at least 10 leaves were pooled for total RNA isolation and subsequent Illumina RNA-seq as previously described [34,35]. As for drought stress, watering was withheld from three-month-old plants for 0, 10, and 20 d; and samples of roots, leaves, and phloem sap were sequenced as previously described [36]. Quality control and read mapping were carried out using Trimmomatic [37] and TopHat (v2.0.8) [38], respectively. The gene expression level was represented using FKPM (fragments per kilobase of exon per million fragments mapped) [39], and differentially expressed genes were determined using RSEM (v1.2.27) [40] with default parameters.

3. Results

3.1. Identification, Chromosome Localization, and Synteny Analysis of 28 LEA Genes in Papaya

Thus far, three genome assemblies have been reported in papaya, i.e., two for a virus-resistant transgenic variety SunUp, and one for its progenitor Sunset [20,41]. Whereas the ASGPBv0.4 assembly of SunUp is fragmented in 17,766 scaffolds [20], two recently available assemblies for SunUp and Sunset are chromosomal-level genomes [41], providing a good chance for comparative genomics analysis. Since the LEA genes identified in two chromosomal-level genomes are exactly the same, only results from the Sunset genome, as well as the ASGPBv0.4 assembly, were presented in Table 1, where an ortholog (i.e., sunset04G0003920/evm.TU.supercontig_6.122) of AtLEA13/-43 was not included due to the absence of a significant LEA_4 domain. Based on the presence of Pfam domains in deduced proteins, 28 identified CpLEA genes were assigned into eight out of nine families as described in arabidopsis (only excluding the AtM family), and each family contains two to five members, respectively, i.e., CpLEA1-1 to -3, CpLEA2-1 to -4, CpLEA3-1 to -5, CpLEA4-1 to -5, CpLEA5-1 to -2, CpLEA6-1 to -2, CpDHN1 to -4, and CpSMP1 to -3 (Table 1). Gene localization analysis indicated that they are not randomly distributed across eight out of nine chromosomes (excluding Chr9), varying from one (i.e., Chr7) to nine (i.e., Chr5) genes. Notably, several hotspots were observed, and a good example is the top of Chr5, which contains the maximum of seven genes (Figure 1). Correspondingly, eight duplicate pairs were identified, which include two tandem repeats (CpLEA2-4/-3/-2) and three transposed repeats (CpLEA2-1/-4, CpLEA3-4/-5, and CpSMP1/-3) (Table S1); on the contrary, synteny analysis revealed that the other three duplicate pairs are located in syntenic blocks and thus were defined as WGD repeats, i.e., CpLEA3-1/-3, CpLEA5-1/-2, and CpSMP1/-2. Among them, CpLEA2-4/-3/-2/-1 as well as CpLEA3-3 are located in the top region of Chr5, though CpLEA3-4 is located in the bottom region (Figure 1). Whereas the protein identity between tandem repeats CpLEA2-3 and CpLEA2-4 is relatively low (about 29.0%), CpLEA2-2 and CpLEA2-3 exhibit 51.1% and 47.2% sequence identity at the nucleotide or protein level, respectively. Moreover, the first 483-bp sequences (counting from the initiation codon) of these two genes even harbor a relatively high sequence identity of 88.4%, and the low sequence identity of the full CDS was shown to result from the divergence of 3′ sequences (Figure S1).

3.2. Identification of LEA Genes in Horseradish Tree and Spider Flower and Definition of Orthogroups

The finding of almost half the amount of LEA genes in papaya relative to those in arabidopsis impelled us to investigate the lineage-specific evolution of the LEA superfamily in different families of Brassicales, i.e., Caricaceae, Moringaceae, Cleomaceae, and Brassicaceae. For this purpose, LEA genes were also identified from horseradish tree and spider flower, whose genome sequences have recently been accessible [21,22]. As shown in Table S1, 29 LEA genes identified in the horseradish tree are comparable to 28 present in papaya, as well as castor bean (an Euphorbiaceae plant also not having experienced any recent WGD), relatively less than 39 found in spider flower, and considerably less than 51 reported in arabidopsis, implying lineage-specific gene contraction and expansion. The species-specific distribution of LEA genes in nine defined gene families is summarized in Figure 2. Notably, no AtM homolog was found beyond arabidopsis.
To gain insights into species-specific evolution patterns, we further conducted BRH-based homology analysis between different species, resulting in 29 orthogroups that are present in more than one species compared (Table 2). In total, 28 CpLEA genes belong to 27 orthogroups, and each orthogroup includes one, with the exception of LEA2b containing two. As for two other orthogroups, DHNe is only present in horseradish tree and spider flower, whereas LEA4f is widely found, though a papaya homolog (see above) has lost the LEA_4 domain. Among three species without a recent WGD, i.e., papaya, horseradish tree, and castor bean, nearly one-to-one orthologous relationships were observed, though no member was identified in castor bean for LEA2b, DHNe, or LEA4a. Notably, a LEA4a homolog is actually found in castor bean, i.e., 30074.t000080; however, no significant LEA_4 domain was identified, supporting species-specific divergence. Like papaya, orthogroups that include more than one member were also found in horseradish tree and castor bean, i.e., MoLEA5-2/-3 in LEA5b, RcDHN2/-3 in DHNb, and RcSMP1/-2 in SMPb, all of which were characterized as tandem repeats (Table S1). On the contrary, orthologous relationships between papaya and spider flower/arabidopsis are relatively complex, including one-to-one, one-to-two, one-to-three, and two-to-four. In spider flower, the majority (84.6%) of duplicate pairs within an orthogroup were characterized as WGD repeats, which is relatively more than the 69.2% found in arabidopsis. Moreover, the duplication mode of the remaining duplicate pairs is also different, i.e., dispersed duplication in spider flower and tandem duplication in arabidopsis, respectively (Table S1).
Compared with other species examined, 27.5% of AtLEA genes seem to be arabidopsis-specific. To uncover their evolution patterns in Brassicaceae, we further traced their orthologs in representative Brassicaceae plants whose genome sequences are available in Phytozome v13, i.e., A. lyrata, A. halleri, Capsella rubella, C. grandiflora, Eutrema salsugineum, Brassica oleracea, and B. rapa. As expected, all of them have orthologs in at least one out of seven species examined, though species-specific evolution was observed (Table S3).

3.3. Exon-Intron Structure, Phylogenetic Analysis, and Structural Characterization

To learn more about the divergence between papaya and arabidopsis, we performed phylogenetic analysis of LEA proteins according to families, and further compared their gene structures and protein motifs. As observed in arabidopsis, CpLEA genes feature few introns, varying from zero to two in the coding region, accounting for 14.3%, 75%, and 10.7% of total genes, respectively. Notably, an additional intron was also found in 5′ or 3′ untranslated regions (UTR) of CpLEA2-1 and CpDHN4, though no intron is present in the coding region of CpDHN4 (Figure 3). Moreover, 12 out of 25 intron-containing CpLEA genes appeared to have alternative splicing (AS) isoforms, and the proportion of 48% is relatively more than the 39.5% found in arabidopsis (Table S1). For convenience, the most expressed transcript was selected for further analyses. The deduced protein length of CpLEA genes varies from 78 to 590 amino acids (AA), and molecular weight (MW) and isoelectric point (pI) values range from 8.77 to 66.20 kDa, or from 4.56 to 10.07, respectively. Except for CpLEA2-4, the GRAVY value of other CpLEA proteins is less than 0, implying their hydrophilic feature. These proteins were predicted to target mitochondria, chloroplast, nuclear, cytoplasmic as well as extracellular genes (Table 1). A further MEME search resulted in 20 conserved motifs, which were shown to significantly distribute over different families (Figure 3 and Figure S2).

3.3.1. LEA_1

The LEA_1 family is also known as D-113 [42]. In papaya, this family includes three members, which is equal to that of arabidopsis (Figure 2). However, their gene origin is not exactly the same. In fact, these genes belong to three phylogenetic groups or orthogroups, i.e., LEA1a, LEA1b, and LEA1c (Figure 2 and Table 2). Among them, AtLEA18 was characterized as a paralog of AtLEA6 that were resulted from the α WGD [43]. Whereas the majority of members in this family contain one intron, CpLEA1-1 and AtLEA18 in LEA1a are intronless (Figure 3), gene-specific loss of an intron can be speculated. Most proteins in this family were shown to harbor Motif 20, which was characterized as the LEA_1 domain. By contrast, despite the presence of the LEA_1 domain in CpLEA1-2 and AtLEA6 as supported by a MOTIF Search, no motif was detected in CpLEA1-2 due to the parameter of 20 motifs set in this study, whereas AtLEA6 was shown to harbor Motif 1, which was characterized as a LEA_4-like domain, supporting their sequence divergence (Figure 3). The length of three CpLEA1s varies from 102 to 160 AA, and the average of 140 AA is relatively longer than the 130 AA observed in arabidopsis. Correspondingly, the MW value varies from 11.41 to 17.01 kDa, and the average of 14.83 kDa is relatively larger than 13.85 kDa in arabidopsis (Table 1). Nevertheless, the pI value in two species appeared to be greater than 7.0, implying their basic feature.

3.3.2. LEA_2

This family is also known as LEA14 or D-95 [42]. The four members found in papaya are relatively more than the three present in arabidopsis (Figure 2). Similar to LEA_1, the LEA_2 family also includes three orthogroups, i.e., LEA2a, LEA2b, and LEA2c (Table 2). In contrast to AtLEA1 and AtLEA27 that are repeats derived from the β WGD [43], CpLEA2-1 was characterized as a transposed repeat of CpLEA2-4, which also resulted in CpLEA2-3 via tandem duplication; and CpLEA2-2 is a more recent tandem repeat of CpLEA2-3 (Figure 1 and Table S1). Most genes in this family harbor a single intron in the coding region; however, CpLEA2-3 contains two instead and the gain of the second intron can be speculated. Moreover, one more intron was also observed in the 5′ UTR of both CpLEA2-1 and AtLEA26, implying their early origin. All members in this family include Motif 6 and Motif 5, which were characterized as the LEA_2 or LEA_3-like domain, respectively. Moreover, both CpLEA2-1 and AtLEA26 harbor two additional motifs, i.e., Motif 13 and Motif 10, where the latter was characterized as the LEA_2 domain; both CpLEA2-2 and CpLEA2-3 include Motif 16, while CpLEA2-2 also contains eight copies of Motif 13 (Figure 3). The length of CpLEA2s varies from 151 to 316 AA, and the average of 239 AA is relatively longer than 214 AA in arabidopsis. Correspondingly, the MW value varies from 16.16 to 35.12 kDa, and the average of 26.40 kDa is relatively larger than 23.48 kDa in arabidopsis. Nevertheless, the pI value in these two species varies from 4.53 to 5.65 (Table 1), suggesting that they are acidic.

3.3.3. LEA_3

This family is also known as LEA5 or D-73 [42], and the five members present in papaya are relatively more than the four present in arabidopsis (Figure 2), which can be assigned into five orthogroups, i.e., LEA3a, LEA3b, LEA3c, LEA3d, and LEA3e (Table 1). Among them, AtLEA38 and AtLEA41 are repeats of AtLEA2 and were derived from the α or γ WGD, respectively [43]; CpLEA3-1 may also be derived from CpLEA3-3 via the γ WGD, whereas CpLEA3-4 was characterized as a transposed repeat of CpLEA3-5, which only exhibit 33.3% sequence identity at the protein level. This family features one intron; however, AtLEA37 has gained an additional intron in the coding region. All members in this family harbor a single motif (i.e., Motif 7), which was characterized as the LEA_3 domain (Figure 3). The length of CpLEA3s varies from 95 to 104 AA, and the average of 100 AA is relatively shorter than 104 AA in arabidopsis. Correspondingly, the MW value varies from 10.61 to 11.78 kDa, and the average of 11.03 kDa is slightly smaller than 11.38 kDa in arabidopsis. The pI value in two species varies from 9.39 to 10.07 (Table 1), indicating that they are basic.

3.3.4. LEA_4

This family is also known as D-7 or D-29 [42], which contains the most number of 6 or 18 members in papaya and arabidopsis, respectively (Figure 2). This family was shown to be highly diverse, including six main orthogroups and six Brassicaceae-specific groups, i.e., LEA4a, LEA4b, LEA4c, LEA4d, LEA4e, LEA4f, AtLEA7/-29, AtLEA11/-12, AtLEA23/-24, AtLEA28, AtLEA39, and AtLEA40 (Table 2 and Table S3). Among them, AtLEA42/-48, AtLEA19/-36, AtLEA13/-43, and AtLEA7/-29 are duplicates that resulted from the α WGD [43], AtLEA11/-12 and AtLEA7/-40 are transposed repeats, and AtLEA23/-24 are tandem repeats (Table S1). The intron number also varies from zero to two, and the copy number of the widely distributed Motif 1, which was characterized as the LEA_4 domain, varies from one to eleven. Additionally, both CpLEA4-1 and AtLEA9 harbor two more motifs, i.e., Motif 12 and Motif 18, where the former was characterized as a domain of unknown function (DUF4149, PF13664) (Figure 3). The length of CpLEA4s varies from 193 to 590 AA, and the average of 358 AA is considerably longer than 280 AA in arabidopsis. Correspondingly, the MW value varies from 23.63 to 61.45 kDa, and the average of 39.33 kDa is relatively smaller than 30.37 kDa in arabidopsis. Unlike most families, the pI value in both species is highly diverse, varying from 4.82 to 9.71 (Table 1).

3.3.5. LEA_5

This family is also known as D-19 or EM [42], which includes two members in both papaya and arabidopsis, comprising two orthogroups, i.e., LEA5a and LEA5b (Figure 2 and Table 2). Whereas CpLEA5-1 and -2 were characterized as WGD repeats, AtLEA20 and -35 are dispersed repeats (Table S1), implying possible chromosome rearrangement after papaya-arabidopsis divergence. All members in this family feature a single intron and harbor Motif 4 that was characterized as the LEA_5 domain (Figure 3). Nevertheless, the sequence length of LEA5b is relatively longer than LEA5a (i.e., 89–92 vs. 111–152) due to fragment insertion. The MW value of CpLEA5-1 and CpLEA5-2 is 9.64 or 12.10 kDa, respectively, and the average of 10.87 kDa is relatively smaller than 13.27 kDa in arabidopsis. The pI value in two species varies from 5.51 to 6.75 (Table 1), suggesting that they are acidic.

3.3.6. LEA_6

This family is also known as PvLEA18 [44], which harbors two or three members in papaya and arabidopsis, respectively (Figure 2). It is composed of two orthogroups, i.e., LEA6a and LEA6b (Table 2), where AtLEA15 and AtLEA16 in LEA6b are tandem repeats (Table S1). Although most genes are intronless, AtLEA15 was shown to gain one intron in the 3′ UTR. The unique motif identified in this family (i.e., Motif 15) was characterized as the LEA_6 domain (Figure 3). CpLEA6-1 and CpLEA6-2 are 97 or 78 AA in length, respectively, and the average of 88 AA is slightly longer at 83 AA in arabidopsis, whereas the average MW value of 9.60 kDa in papaya is relatively larger than 8.71 kDa in arabidopsis. The pI value in these two species varies from 4.46 to 5.56 (Table 1), implying that they are acidic.

3.3.7. DHN

This family is also known as D-11 [42], and the 4 members found in papaya is considerably less than the 10 present in arabidopsis (Figure 2). These genes constitute five orthogroups and one Brassicaceae-specific group, i.e., DHNa, DHNb, DHNc, DHNd, DHNe, and AtLEA44 (Table 2 and Table S3). Among them, AtLEA4/-5 and AtLEA33/-34 are tandem repeats (Table S1), where AtLEA4/-10, AtLEA14/-45, and AtLEA33/-51 are duplicates that were derived from the α WGD [43]. Most members in this family harbor one intron in the coding region; however, AtLEA33 has lost the corresponding intron present in its paralogs (i.e., AtLEA34 and AtLEA51). By contrast, one conserved intron was found in the 3′ UTR of both CpDHN4 and AtLEA8, though the intron retention was observed in one alternative splicing isoform of CpDHN4, supporting species-specific evolution. All members in this family include Motif 3, which was characterized as the DHN domain (or more precisely as the K-segment), and the motif copies vary from one to six. One copy of Motif 9, which was also characterized as the DHN domain (or more precisely as the S-segment), is widely found with the exception of CpDHN4, AtLEA8, AtLEA33, and AtLEA45. Further sequence alignment revealed the presence of the S-segment at the C-terminal of both CpDHN4 and AtLEA8, and one to three copies of the Y-segment at the N-terminal of CpDHN2, CpDHN3, AtLEA14, AtLEA34, AtLEA45, and AtLEA51. Based on the presence and order of these conserved domains, all five architectures (i.e., Kn, SKn, KnS, YnKn, and YnSKn) were found in arabidopsis, while only SKn, KnS, and YnSKn were identified in papaya (Figure S3). Additionally, members in DHNa as well as AtLEA44 also harbor Motif 19 (Figure 3 and Figure S3), whose function has not been described yet. The length of CpDHNs varies from 93 to 211 AA, and the average of 152 AA is relatively shorter than 181 AA in arabidopsis. Correspondingly, the MW value varies from 10.50 to 24.10 kDa, and the average of 16.82 kDa is relatively smaller than 19.76 kDa in arabidopsis. Like the LEA_4 family, the pI value in both species is also diverse, varying from 4.74 to 9.38 (Table 1).

3.3.8. SMP

This family is also known as D-34 [42,45], and the three members identified in papaya is considerably less than the six present in arabidopsis (Figure 2). They comprise three orthogroups and one Brassicaceae-specific group, i.e., SMPa, SMPb, SMPc, and AtLEA49/-50 (Table 2 and Table S3). Among them, AtLEA31/-32 and AtLEA49/-50 are tandem repeats, AtLEA3/31 are transposed repeats, and CpSMP2/-3 were characterized as WGD and transposed repeats of CpSMP1, respectively. Members in SMPa and SMPc feature two introns, whereas other group members have no or a single one instead. Despite the close evolutionary relationship between AtLEA49 and AtLEA50, they include one intron in the coding region or 5′ UTR, respectively, implying fast evolution and sequence divergence. All members in this family include Motif 2, which was characterized as the SMP domain. Moreover, Motif 14 is also present in members of SMPa, SMPb, and SMPc, whereas two more motifs (i.e., Motif 8 and Motif 17) were also found in members of SMPa and SMPb (Figure 3). Noteworthy, Motif 8 was also characterized as the SMP domain, implying possible fragment duplication or gene fusion. The length of CpSMPs varies from 244 to 267 AA, and the average of 258 AA is relatively longer than 204 AA in arabidopsis. Correspondingly, the MW value varies from 25.13 to 27.97 kDa, and the average of 26.60 kDa is relatively larger than 21.15 kDa in arabidopsis. The pI value in the two species varies from 4.56 to 6.44 (Table 1), indicating that they are acidic.

3.4. ABRE and LTRE cis-Acting Elements Present in the Promoter Region of CpLEA Genes

LTRE, also known as DRE (drought responsive) or CRT (C-repeat), is a key cis-acting element for CBF/DREB1 transcription factors, whereas ABRE is a key element involved in ABA signaling [46,47]. Previous studies showed that these two elements are overrepresented in the promoter region of AtLEA genes and are associated with ABA, cold and/or drought responses [3]. To reveal possible response patterns of CpLEA genes to stresses, we examined the presence of ABRE and LTRE elements in the 2,000-bp promoter regions. Results showed that 89.3% of CpLEA genes contain 1 to 10 copies of the ABRE element, only excluding CpLEA2-3, CpLEA3-5, and CpLEA6-1, while 67.9% of them contain 1 to 4 copies of the LTRE element, excluding CpLEA1-2, CpLEA2-2, CpLEA3-4, CpLEA3-5, CpLEA5-2, CpLEA6-1, CpDHN3, CpSMP1, and CpSMP3 (Figure 4). The proportion is similar to the 82.0% and 69.0% reported for AtLEA genes, respectively [3].

3.5. Tissue-Specific Expression Profiles of CpLEA Genes

Although some LEA proteins have been reported to be regulated by posttranslational modifications (e.g., phosphorylation), cellular trafficking, homo- and heteromerization [18,48,49,50], and transcriptional regulation still represent a key mechanism to perform their functions. For this purpose, we first performed global expression profiling of CpLEA genes in various tissues.
As shown in Figure 5, our transcriptional profiling supported the expression of all CpLEA genes in at least one of 11 tissues examined in this study, i.e., root, apical bud, leaf, petiole, leaf vein, phloem sap, male flower, female flower, fruit, peel, and seed, though the transcript level was highly diverse. As expected, CpLEA genes were most expressed in the seed, but considerably less expressed in the leaf and root, which is consistent with the cluster analysis. In total, 22 out of 28 CpLEA genes (75.9%) possessed a FKPM value >1 in the seed, which is relatively more than the 15 in the petiole, 15 in the vein, 13 in the root, 12 in the bud, 11 in the fruit, 11 in the peel, 10 in the leaf, 10 in the female flower, 9 in the male flower, and 7 in the sap. Five genes, i.e., CpLEA3-3, CpDHN4, CpDHN1, CpLEA2-1, and CpLEA2-4, appeared to constitutively express in these tissues, whereas other genes were tissue-specific. As for a certain tissues, several key genes were also identified: CpLEA3-3 represents the most expressed gene in most tested tissues, whereas CpLEA1-3 and CpDHN1 represent the most expressed genes in the seed or bud/fruit, respectively; CpDHN4 represents the second most expressed gene in the male flower, female flower, petiole, vein, and peel, whereas CpLEA3-3, CpDHN1, CpDHN3, and CpLEA2-2 represent the second most expressed genes in the bud/fruit, root/leaf, seed, or sap, respectively. According to tissue-specific expression patterns, CpLEA genes can be divided into five main clusters: Cluster I includes the most of the 13 genes that are predominantly expressed in the seed; Cluster II includes CpLEA3-5 (preferentially expressed in fruit), CpLEA3-2 (preferentially expressed in vein), CpSMP2 (preferentially expressed in seed), and other four rarely expressed genes; Cluster III includes CpLEA3-3, CpDHN1, and CpDHN4, which are constitutively expressed; Cluster IV includes CpLEA2-2, CpLEA2-3, and CpLEA2-4, which are typically expressed in sap; and Cluster V includes the constitutively expressed CpLEA2-1, as well as CpLEA3-1, which is preferentially expressed in fruit (Figure 5).

3.6. Expression Patterns of CpLEA Genes during Fruit Development

To learn more about the expression pattern of CpLEA genes during fruit development, six typical stages were investigated, i.e., 30 days post-anthesis (30 DPA), 150 DPA, and stages 1–4 of fruit flesh from immature to ripe, i.e., S1, S2, S3, and S4, as previously defined [51]. Unlike rapid accumulation of LEA genes during the late stage of seed development as described in other species, CpLEA genes were shown to be the most expressed in the early stages of fruit development, but considerably less expressed in mature fresh fruit. Based on the expression patterns of 15 genes with the FKPM value >1 in at least one of the stages tested, these genes could be grouped into four clusters: Cluster I includes CpLEA3-1, CpLEA3-3, CpDHN1, and CpDHN4, which were highly abundant in all stages; Cluster II includes CpLEA1-3, CpLEA3-2, CpLEA3-4, and CpLEA5-1, which were rarely or lowly expressed in a few stages; Cluster III includes CpLEA4-2 and CpSMP1, which were lowly expressed in most stages; Cluster IV includes CpLEA2-1, CpLEA2-2, CpLEA2-3, CpLEA2-4, and CpLEA3-5, which were moderately expressed in most stages (Figure 6).

3.7. Expression Patterns of CpLEA Genes under Drought, Cold and Salt Stresses

The response of CpLEA genes to mild (10 d) and severe (20 d) drought was investigated based on transcriptomes of the roots, leaves, and phloem sap [36]. As shown in Figure 7, a total of 15 CpLEA genes were differentially expressed in at least one tissue per treatment, and the majority of them (86.7%) were shown to be significantly up-regulated. As for the root, six genes, i.e., CpLEA1-3, CpLEA4-2, CpLEA4-3, CpLEA4-5, CpLEA5-1, and CpSMP1, were up-regulated under both conditions; CpLEA3-2 was up-regulated by mild drought, whereas CpLEA4-1 and CpDHN4 were up-regulated by severe drought; by contrast, CpDHN1 was down-regulated by severe drought. As for the leaf, in contrast to the down-regulation of CpDHN1, four genes, i.e., CpLEA1-3, CpLEA4-2, CpLEA4-5, and CpDHN4, were up-regulated by both treatments; CpSMP1 was up-regulated only by mild drought, whereas CpLEA2-1, CpLEA3-3, CpLEA4-3, CpLEA5-1, and CpLEA6-2 were up-regulated only by severe drought; CpLEA2-4 and CpLEA4-1 were down-regulated by mild and severe drought, respectively. As for the sap, only one gene (i.e., CpLEA2-1) was up-regulated by severe drought (Figure 7).
To study the response of CpLEA genes to cold and salt stresses, eight-week-old plantlets were subjected to 4 °C chilling or 300 mM/L NaCl treatment, and the leaf transcriptome was characterized at 0–40 h or 0–20 d post treatment, respectively. Among 18 CpLEA genes with a FKPM value >1, 16 genes were shown to be significantly regulated: six (i.e., CpLEA1-3, CpLEA2-4, CpLEA4-1, CpLEA4-2, CpLEA6-2, and CpDHN4) are shared by cold and salt stresses, whereas five (i.e., CpLEA2-1, CpLEA2-2, CpLEA2-3, CpLEA3-3, and CpDHN1) and five (i.e., CpLEA1-1, CpLEA3-1, CpLEA3-4, CpLEA4-3, and CpLEA4-5) are cold- or salt-specific, respectively. Similar to drought stress, most genes were up-regulated, accounting for about 68.8% of total LEA genes, though some of them (i.e., CpLEA1-3, CpLEA2-4, and CpLEA4-1) were occasionally down-regulated at a certain time point. As for cold stress, five regulated genes are shared by three time points, including four up-regulated (i.e., CpLEA2-4, CpLEA3-3, CpLEA4-2, and CpDHN4) and one down-regulated (i.e., CpLEA2-3); CpLEA2-2 was down-regulated at two former time points, whereas CpLEA1-3 and CpDHN1 were up-regulated at the latter two time points; CpLEA2-1 and CpLEA6-2 were down-regulated at 7 or 40 h post-treatment, respectively; CpLEA4-1 was down-regulated at 7 h but up-regulated at 40 h post-treatment. As for salt stress, CpLEA4-2 and CpLEA4-5 were up-regulated at three time points, whereas CpLEA1-3 was down-regulated at 10 d but up-regulated at the latter two time points; CpLEA1-1, CpLEA4-1, and CpLEA4-3 were up-regulated at the latter two time points, whereas CpLEA3-4 was down-regulated at the same time points; CpDHN4 was up-regulated at 10 d post-treatment, whereas CpLEA2-4 and CpLEA6-2 were down-regulated at the same time point; CpLEA3-1 was up-regulated at 10 and 20 d post-treatment (Figure 8).

4. Discussion

4.1. Small Number but High Diversity of LEA Genes in Papaya

Although first identified for their accumulation in the later stages of seed development, LEA proteins have been found in a wide range of plant tissues, as well as different types of organisms [1,7,21,45]. In contrast to a single or few members present in algae, rapid expansion of the LEA superfamily was observed in terrestrial plants, which was shown to be essential for survival under water stress [9,52]. Rapid gene expansion is usually accompanied by WGDs, which are widespread and play an important role in the radiation of flowering plants [53]. In eudicots, studies established that the γ whole genome triplication event occurred at 117 million years ago (Mya), sometime before the diversification of core eudicots [54]. After that, arabidopsis, a Brassicaceae plant within the order Brassicales, was proven to experience two additional whole genome doubling events, i.e., β and α, occurred within a window of 61–65 and 23–50 Mya, respectively [19,55]. As a result, a high number of 51 LEA genes are present in arabidopsis, including seven dispersed repeats as well as 21 repeats that resulted from γ WGD (1), β WGD (1), α WGD (9), tandem duplication (7), and transposed duplication (4) (Table S1).
In this study, a first genome-wide identification of LEA genes was conducted in an important tropical fruit tree of the Caricaceae family, papaya, as well as another two Brassicales plants, i.e., horseradish tree and spider flower. Horseradish tree is an important multipurpose shrub with medicinal and nutritional properties and the ability to grow in the low water conditions of the Moringaceae family, whereas spider flower belongs to a phylogenetic outgroup of the Brassicaceae sister family Cleomaceae [21,22]. Like castor bean (Euphorbiaceae), the papaya and horseradish tree did not experience any additional WGD after the γ WGD. By contrast, the spider flower shared the β WGD but further experienced one genome triplication that is independent of the Brassicaceae-specific α WGD as described in arabidopsis [19,20,21,22,56]. As expected, a relatively small number of 28 or 29 LEA genes were found in the papaya and horseradish tree, respectively, which are comparable to 28 reported in castor bean, but relatively less than the 39 and 51 present in spider flower and arabidopsis, respectively, reflecting the occurrence of lineage-specific WGDs in the latter after their divergence [3,8,19,21].
LEA genes identified in this study belong to eight out of nine families as described in arabidopsis, i.e., LEA_1, LEA_2, LEA_3, LEA_4, LEA_5, LEA_6, DHN, and SMP [3]. As for the AtM family, which includes two tandem repeats in arabidopsis, it is more likely to be Brassicaceae-specific, because it is widely present in Brassicaceae plants (Table S3) but has not yet been identified in other species [3,8,9,12,13], including species examined in this study. Nevertheless, 28 CpLEA genes represent 27 out of 29 orthogroups based on sequence comparison of the above five species, though a LEA4f homolog has lost the corresponding LEA_4 domain. Moreover, no orthologs were identified for CpLEA1-2, CpLEA2-2, CpLEA2-3, CpLEA3-4, or CpLEA3-5 in arabidopsis, though their counterparts are present in at least one of three other species examined.

4.2. Comparative Genomics Analysis Reveals Lineage-Specific Evolution of the LEA Superfamily in Brassicales

Orthology defines genes in different organisms that evolved from a common ancestral gene via speciation, which may perform similar functions [57]. Characterization of 29 orthogroups in five representative species allows us to infer lineage-specific evolution in Brassicales. Notably, a nearly one-to-one orthologous relationship was observed between the papaya/horseradish tree and castor bean, though they belong to different plant families, implying that few LEA genes have been lost in either the papaya or horseradish tree after the split with the castor bean. By contrast, tandem duplication plays a predominant role in gene expansion within an orthogroup, i.e., RcDHN2/-3 in DHNb, and RcSMP1/-2 in SMPb, CpLEA2-2/-3 in LEA2b, and MoLEA5-2/-3 in LEA5b. As for the spider flower, which experienced two WGDs (including the β WGD shared by Brassicaceae plants) after the split with papaya at approximately 72 Mya [21,58], duplicate pairs are mainly contributed by WGD (12), followed by dispersed duplication (3) and transposed duplication (1) (Table S1). The transposed duplication is shared by all five species examined, whereas WGD repeats appear to be spider flower-specific. By contrast, AtLEA2/-41 and AtLEA1/-27 were characterized as γ and β WGD-derived repeats, respectively [22], supporting species-specific evolution following WGDs. Nevertheless, since the spider flower-specific WGD is a triplication event, theoretically, it should have given rise to three gene copies from a single ancestral gene. However, in most cases, only one or two copies are maintained. Unlike the spider flower, tandem duplication also plays a key role in gene expansion in arabidopsis.
Further comparative analysis of exon-intron structures and protein motifs revealed frequent gain and/or loss of certain introns/motifs, which includes the loss of the second intron in CpLEA2-3 relative to CpLEA2-2. In fact, compared with papaya, such an occurrence is relatively more prevalent in arabidopsis, which is consistent with a relatively faster evolution of annual than perennial shrubs [59]. Nevertheless, family-specific Pfam domains are highly conserved. It is worth noting that CpLEA2-1 and AtLEA26 contain two LEA_2 domains relative to a single one present in other LEA_2 family members, implying a possible fragment repetition. From an evolutionary perspective, further characterization of these species-specific genes is of particular interest.

4.3. Diverse Expression Patterns of CpLEA Genes and a Role in Fruit Development and Abiotic Stress Responses

As reported in other species, our transcriptional profiling revealed diverse expression patterns of CpLEA genes in 11 tissues, as well as six typical stages of fruit development examined in this study. In contrast to the constitutive expression of a few members, e.g., CpLEA2-1, CpLEA2-4, CpLEA3-3, CpDHN1, and CpDHN4, most CpLEA genes appeared to preferentially express in a few tissues, especially in seed. However, except for CpLEA1-3 and CpSMP1 that preferentially accumulated in mature fruits, the expression patterns of most CpLEA genes differ from that observed in seeds, which undergo a dehydration process [2,3,4,5,6,21]. The high abundance of CpDHN4, CpDHN1, CpLEA3-1, and CpLEA3-3 in fruits implies their possible important role in this special tissue.
Analyzing promoter sequences of CpLEA genes revealed the presence of a high number of ABRE and LTRE cis-acting elements, implying their possible involvement in stress responses. As expected, the transcript levels of most CpLEA were shown to be significantly regulated by the cold, drought, and high salt conditions examined in this study. Among three genes (i.e., CpLEA2-3, CpLEA3-5, and CpLEA6-1) without ABRE elements in their promoters, none of them were regulated by drought as well as salt, though CpLEA2-3 was down-regulated by cold, which is consistent with the presence of one copy of the LTRE element in its promoter. Among nine genes (i.e., CpLEA1-2, CpLEA2-2, CpLEA3-4, CpLEA3-5, CpLEA5-2, CpLEA6-1, CpDHN3, CpSMP1, and CpSMP3) without LTRE elements, only CpLEA2-2 was shown to be down-regulated by cold, while CpLEA3-4 and CpSMP1 were regulated by salt or drought, respectively. Among 20 genes containing both ABRE and LTRE cis-acting elements, most of them (85.0%) were regulated by at least one of the three stresses tested, only excluding CpLEA4-4, CpDHN2, and CpSMP2, which were preferentially expressed in seed but lowly expressed in the leaf, root and sap examined in this study. Among these 17 regulated genes, all of them were up-regulated by at least one treatment in at least one of three examined tissues: nine genes (i.e., CpLEA1-1, CpLEA3-1, CpLEA3-2, CpLEA3-3, CpLEA4-2, CpLEA4-3, CpLEA4-5, CpLEA5-1, and CpDHN4) exhibit a single up-regulated pattern; CpLEA2-4, the unique gene regulated in sap, was up-regulated by drought but down-regulated by cold in leaf; CpDHN1, a cold-induced gene, was down-regulated by drought in both the root and leaf; CpLEA2-4 was up-regulated by cold but down-regulated by both drought and NaCl in the leaf; CpLEA6-2 was down-regulated by both cold and NaCl but up-regulated by drought in the leaf; CpLEA4-5, a NaCl-induced gene, was down-regulated in leaf but up-regulated in root upon drought stress; by contrast, an initial decline followed by a steady increasing trend was observed. Regulation by stresses has been frequently reported in arabidopsis, rice, cassava (Manihot esculenta), and other species [2,3,12,13]. In arabidopsis, a study revealed that 54.5% of genes highly expressed in non-seed tissues were induced more than threefold by various stresses, mainly by cold, drought and salt [3]. For example, AtLEA18, the ortholog of CpLEA1-1, was also induced by salt; AtLEA41, the ortholog of CpLEA3-1, was induced by ABA, cold, and salt; AtLEA46, the ortholog of CpLEA1-3, was induced by ABA, cold, drought, and salt [2,3]. Thereby, similar functions could be speculated.

5. Conclusions

This study presents the first genome-wide identification of LEA genes in papaya as well another two Brassicales plants, horseradish tree and spider flower; resulting in 28, 29, and 39 members, respectively. These genes belong to eight out of nine families as described in arabidopsis, i.e., LEA_1, LEA_2, LEA_3, LEA_4, LEA_5, LEA_6, DHN, and SMP. Further comparison of LEA genes in papaya, horseradish tree, spider flower, castor bean, and arabidopsis reveals lineage-specific evolution in Brassicales, and significant expansion in spider flower and arabidopsis was mainly contributed by WGDs sometime after their split with papaya. Analysis of exon-intron structures and protein motifs supported the fast evolution of this special family, especially in arabidopsis. Moreover, global expression profiles of CpLEA genes were comprehensively analyzed, which revealed tissue-specific expression patterns and key roles in fruit development and stress responses. Taken together, these findings provide valuable information for further functional analysis of LEA genes in papaya and other species.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/life12091453/s1, Figure S1: Nucleotide and protein sequence alignments of CpLEA2-2 and CpLEA2-3; Figure S2: Sequence logos of 20 motifs identified in this study; Figure S3: Alignment of DHN proteins in papaya and arabidopsis; Table S1: Detailed information of LEA genes present in papaya, horseradish tree, spider flower, castor bean, and arabidopsis; Table S2: Detailed information of transcriptome data used in this study; Table S3: Orthologs in representative Brassicaceae plants for 14 arabidopsis-specific LEA genes identified in this study.

Author Contributions

Z.Z.: methodology, data curation and writing—original draft; Z.Z., J.G., Y.Z. and Y.X.: data curation; Z.Z., J.G., Y.Z. and Y.X.: conceptualization and methodology; Z.Z., Y.Z. and Y.X.: software; Z.Z. and A.G.: formal analysis and preparation of materials; Z.Z.: conceptualization, data curation and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Hainan province (320RC705), the National Natural Science Foundation of China (31971688), and the Central Public-interest Scientific Institution Basal Research Fund for Chinese Academy of Tropical Agricultural Sciences (1630052022001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Battaglia, M.; Olvera-Carrillo, Y.; Garciarrubio, A.; Campos, F.; Covarrubias, A.A. The enigmatic LEA proteins and other hydrophilins. Plant. Physiol. 2008, 148, 6–24. [Google Scholar] [CrossRef] [PubMed]
  2. Bies-Etheve, N.; Gaubier-Comella, P.; Debures, A.; Lasserre, E.; Jobet, E.; Raynal, M.; Cooke, R.; Delseny, M. Inventory, evolution and expression profiling diversity of the LEA (late embryogenesis abundant) protein gene family in Arabidopsis thaliana. Plant. Mol. Biol. 2008, 67, 107–124. [Google Scholar] [CrossRef] [PubMed]
  3. Hundertmark, M.; Hincha, D.K. LEA (late embryogenesis abundant) proteins and their encoding genes in Arabidopsis thaliana. BMC Genom. 2008, 9, 118. [Google Scholar] [CrossRef] [PubMed]
  4. Dure, L.; Chlan, C. Developmental biochemistry of cottonseed embryogenesis and germination. XII. Purification and properties of principal storage proteins. Plant. Physiol. 1981, 68, 180–186. [Google Scholar] [CrossRef] [PubMed]
  5. Dure, L.; Galau, G.A. Developmental biochemistry of cottonseed embryogenesis and germination. XIII. Regulation of biosynthesis of principal storage proteins. Plant. Physiol. 1981, 68, 187–194. [Google Scholar] [CrossRef]
  6. Dure, L.; Greenway, S.C.; Galau, G.A. Developmental biochemistry of cottonseed embryogenesis and germination: Changing messenger ribonucleic acid populations as shown by in vitro and in vivo protein synthesis. Biochemistry 1981, 20, 4162–4168. [Google Scholar] [CrossRef]
  7. Hand, S.C.; Menze, M.A.; Toner, M.; Boswell, L.; Moore, D. LEA proteins during water stress: Not just for plants anymore. Annu. Rev. Physiol. 2011, 73, 115–134. [Google Scholar] [CrossRef]
  8. Zou, Z.; Huang, Q.X.; An, F. Genome-wide identification, classification and phylogenetic analysis of LEA gene family in castor bean (Ricinus communis L.). Chin. J. Oil Crop. Sci. 2013, 35, 637–643. [Google Scholar]
  9. Artur, M.A.S.; Zhao, T.; Ligterink, W.; Schranz, E.; Hilhorst, H.W. Dissecting the genomic diversification of late embryogenesis abundant (LEA) protein gene families in plants. Genome Biol. Evol. 2019, 11, 459–471. [Google Scholar] [CrossRef]
  10. Raynal, M.; Guilleminot, J.; Gueguen, C.; Cooke, R.; Delseny, M.; Gruber, V. Structure, organization and expression of two closely related novel Lea (late-embryogenesis abundant) genes in Arabidopsis thaliana. Plant. Mol. Biol. 1999, 40, 153–165. [Google Scholar] [CrossRef]
  11. Singh, S.; Cornilescu, C.C.; Tyler, R.C.; Cornilescu, G.; Tonelli, M.; Lee, M.S.; Markley, J.L. Solution structure of a late embryogenesis abundant protein (LEA14) from Arabidopsis thaliana, a cellular stress-related protein. Protein Sci. 2005, 14, 2601–2609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Wang, X.S.; Zhu, H.B.; Jin, G.L.; Liu, H.L.; Wu, W.R.; Zhu, J. Genome-scale identification and analysis of LEA genes in rice (Oryza sativa L.). Plant. Sci. 2007, 172, 414–420. [Google Scholar] [CrossRef]
  13. Wu, C.; Hu, W.; Yan, Y.; Tie, W.; Ding, Z.; Guo, J.; He, G. The late embryogenesis abundant protein family in cassava (Manihot esculenta Crantz): Genome-wide characterization and expression during abiotic stress. Molecules 2018, 23, 1196. [Google Scholar] [CrossRef]
  14. Salleh, F.M.; Evans, K.; Goodall, B.; Machin, H.; Mowla, S.B.; Mur, L.A.; Runions, J.; Theodoulou, F.L.; Foyer, C.H.; Rogers, H.J. A novel function for a redox-related LEA protein (SAG21/AtLEA5) in root development and biotic stress responses. Plant. Cell Environ. 2012, 35, 418–429. [Google Scholar] [CrossRef] [PubMed]
  15. Dang, N.X.; Popova, A.V.; Hundertmark, M.; Hincha, D.K. Functional characterization of selected LEA proteins from Arabidopsis thaliana in yeast and in vitro. Planta 2014, 240, 325–336. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, X.; Lu, S.; Jiang, C.; Wang, Y.; Lv, B.; Shen, J.; Ming, F. RcLEA, a late embryogenesis abundant protein gene isolated from Rosa chinensis, confers tolerance to Escherichia coli and Arabidopsis thaliana and stabilizes enzyme activity under diverse stresses. Plant. Mol. Biol. 2014, 85, 333–347. [Google Scholar] [CrossRef]
  17. Xiang, D.J.; Man, L.L.; Zhang, C.L.; Li, Z.G.; Zheng, G.C. A new Em-like protein from Lactuca sativa, LsEm1, enhances drought and salt stress tolerance in Escherichia coli and rice. Protoplasma 2018, 255, 1089–1106. [Google Scholar] [CrossRef]
  18. Hernández-Sánchez, I.E.; Maruri-López, I.; Molphe-Balch, E.P.; Becerra-Flora, A.; Jaimes-Miranda, F.; Jiménez-Bremont, J.F. Evidence for in vivo interactions between dehydrins and the aquaporin AtPIP2B. Biochem. Bioph. Res. Commun. 2019, 510, 545–550. [Google Scholar] [CrossRef]
  19. Bowers, J.E.; Chapman, B.A.; Rong, J.; Paterson, A.H. Unravelling angiosperm genome evolution by phylogenetic analysis of chromosomal duplication events. Nature 2003, 422, 433–438. [Google Scholar] [CrossRef]
  20. Ming, R.; Hou, S.; Feng, Y.; Yu, Q.; Dionne-Laporte, A.; Saw, J.H.; Senin, P.; Wang, W.; Ly, B.V.; Lewis, K.L.; et al. The draft genome of the transgenic tropical fruit tree papaya (Carica papaya Linnaeus). Nature 2008, 452, 991–996. [Google Scholar] [CrossRef]
  21. Cheng, S.; van den Bergh, E.; Zeng, P.; Zhong, X.; Xu, J.; Liu, X.; Hofberger, J.; de Bruijn, S.; Bhide, A.S.; Kuelahoglu, C.; et al. The Tarenaya hassleriana genome provides insight into reproductive trait and genome evolution of crucifers. Plant. Cell 2013, 25, 2813–2830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Shyamli, P.S.; Pradhan, S.; Panda, M.; Parida, A. De novo whole-genome assembly of Moringa oleifera helps identify genes regulating drought stress tolerance. Front. Plant. Sci. 2021, 12, 766999. [Google Scholar] [CrossRef] [PubMed]
  23. Ming, R.; Moore, P. Genetics and Genomics of Papaya. Plant Genetics and Genomics: Crops and Models; Springer: Cham, Switzerland, 2014; Volume 10. [Google Scholar] [CrossRef]
  24. Allan, P. Carica papaya responses under cool subtropical growth conditions. Acta Hortic. 2002, 575, 757–763. [Google Scholar] [CrossRef]
  25. Mahouachi, J.; Socorro, A.; Talon, M. Responses of papaya seedlings (Carica papaya L.) to water stress and re-hydration: Growth, photosynthesis and mineral nutrient imbalance. Plant. Soil. 2006, 281, 137–146. [Google Scholar] [CrossRef]
  26. Altschul, S.F.; Madden, T.L.; Schäffer, A.A.; Zhang, J.; Zhang, Z.; Miller, W.; Lipman, D.J. Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 3389–3402. [Google Scholar] [CrossRef]
  27. Zou, Z.; Yang, L.; Gong, J.; Mo, Y.; Wang, J.; Cao, J.; An, F.; Xie, G. Genome-wide identification of Jatropha curcas aquaporin genes and the comparative analysis provides insights into the gene family expansion and evolution in Hevea brasiliensis. Front. Plant. Sci. 2016, 7, 395. [Google Scholar] [CrossRef]
  28. 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]
  29. Qiao, X.; Li, Q.; Yin, H.; Qi, K.; Li, L.; Wang, R.; Zhang, S.; Paterson, A.H. Gene duplication and evolution in recurring polyploidization-diploidization cycles in plants. Genome Biol. 2019, 20, 38. [Google Scholar] [CrossRef]
  30. Moreno-Hagelsieb, G.; Latimer, K. Choosing BLAST options for better detection of orthologs as reciprocal best hits. Bioinformatics 2008, 24, 319–324. [Google Scholar] [CrossRef]
  31. Hu, B.; Jin, J.; Guo, A.Y.; Zhang, H.; Luo, J.; Gao, G. GSDS 2.0: An upgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Tamura, K.; Stecher, G.; Peterson, D.; Filipski, A.; Kumar, S. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol. Biol. Evol. 2013, 30, 2725–2729. [Google Scholar] [CrossRef] [PubMed]
  33. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. MEME SUITE: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef] [PubMed]
  34. Zou, Z.; Gong, J.; An, F.; Xie, G.; Wang, J.; Mo, Y.; Yang, L. Genome-wide identification of rubber tree (Hevea brasiliensis Muell. Arg.) aquaporin genes and their response to ethephon stimulation in the laticifer, a rubber-producing tissue. BMC Genom. 2015, 16, 1001. [Google Scholar] [CrossRef] [PubMed]
  35. Zou, Z.; Yang, J.H.; Zhang, X.C. Insights into genes encoding respiratory burst oxidase homologs (RBOHs) in rubber tree (Hevea brasiliensis Muell. Arg.). Ind. Crop. Prod. 2019, 128, 126–139. [Google Scholar] [CrossRef]
  36. Gamboa-Tuz, S.D.; Pereira-Santana, A.; Zamora-Briseño, J.A.; Castano, E.; Espadas-Gil, F.; Ayala-Sumuano, J.T.; Keb-Llanes, M.Á.; Sanchez-Teyer, F.; Rodríguez-Zapata, L.C. Transcriptomics and co-expression networks reveal tissue-specific responses and regulatory hubs under mild and severe drought in papaya (Carica papaya L.). Sci. Rep. 2018, 8, 14539. [Google Scholar] [CrossRef] [PubMed]
  37. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef]
  38. Trapnell, C.; Williams, B.A.; Pertea, G.; Mortazavi, A.; Kwan, G.; Van Baren, M.J.; Salzberg, S.L.; Wold, B.J.; Pachter, L. Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat. Biotechnol. 2010, 28, 511–515. [Google Scholar] [CrossRef]
  39. Langmead, B.; Salzberg, S.L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 2012, 9, 357–359. [Google Scholar] [CrossRef]
  40. Li, B.; Dewey, C.N. RSEM: Accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinform. 2011, 12, 323. [Google Scholar] [CrossRef]
  41. Yue, J.; VanBuren, R.; Liu, J.; Fang, J.; Zhang, X.; Liao, Z.; Wai, C.M.; Xu, X.; Chen, S.; Zhang, S.; et al. SunUp and Sunset genomes revealed impact of particle bombardment mediated transformation and domestication history in papaya. Nat. Genet. 2022, 54, 715–724. [Google Scholar] [CrossRef]
  42. Dure, L. A repeating 11-mer amino acid motif and plant desiccation. Plant. J. 1993, 3, 363–369. [Google Scholar] [CrossRef]
  43. Wang, Y.; Tan, X.; Paterson, A.H. Different patterns of gene structure divergence following gene duplication in Arabidopsis. BMC Genom. 2013, 14, 652. [Google Scholar] [CrossRef] [PubMed]
  44. Colmenero-Flores, J.M.; Moreno, L.P.; Smith, C.E.; Covarrubias, A.A. Pvlea18, a member of a new late-embryogenesis-abundant protein family that accumulates during water stress and in the growing region of well-irrigated bean seedlings. Plant. Physiol. 1999, 120, 93–103. [Google Scholar] [CrossRef] [PubMed]
  45. Zou, Z.; Zhao, Y.; Zhang, L.; Xiao, Y.; Guo, A. Analysis of Cyperus esculentus SMP family genes reveals lineage-specific evolution and seed desiccation-like transcript accumulation during tuber maturation. Ind. Crop. Prod. 2022, 187, 115382. [Google Scholar] [CrossRef]
  46. Bartels, D.; Sunkar, R. Drought and salt tolerance in plants. Crit. Rev. Plant. Sci. 2005, 24, 23–58. [Google Scholar] [CrossRef]
  47. 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] [PubMed]
  48. Nylander, M.; Svensson, J.; Palva, E.T.; Welin, B.V. Stress-induced accumulation and tissue-specific localization of dehydrins in Arabidopsis thaliana. Plant. Mol. Biol. 2001, 45, 263–279. [Google Scholar] [CrossRef]
  49. Alsheikh, M.K.; Svensson, J.T.; Randall, S.K. Phosphorylation regulated ion-binding is a property shared by the acidic subclass dehydrins. Plant. Cell Environ. 2005, 28, 1114–1122. [Google Scholar] [CrossRef]
  50. Candat, A.; Paszkiewicz, G.; Neveu, M.; Gautier, R.; Logan, D.C.; Avelange-Macherel, M.H.; Macherel, D. The ubiquitous distribution of late embryogenesis abundant proteins across cell compartments in Arabidopsis offers tailored protection against abiotic stress. Plant. Cell 2014, 26, 3148–3166. [Google Scholar] [CrossRef]
  51. Lü, P.; Yu, S.; Zhu, N.; Chen, Y.R.; Zhou, B.; Pan, Y.; Tzeng, D.; Fabi, J.P.; Argyris, J.; Garcia-Mas, J.; et al. Genome encode analyses reveal the basis of convergent evolution of fleshy fruit ripening. Nat. Plants. 2018, 4, 784–791. [Google Scholar] [CrossRef] [Green Version]
  52. Rensing, S.A.; Lang, D.; Zimmer, A.D.; Terry, A.; Salamov, A.; Shapiro, H.; Nishiyama, T.; Perroud, P.F.; Lindquist, E.A.; Kamisugi, Y.; et al. The Physcomitrella genome reveals evolutionary insights into the conquest of land by plants. Science 2008, 319, 64–69. [Google Scholar] [CrossRef] [PubMed]
  53. Van de Peer, Y.; Fawcett, J.A.; Proost, S.; Sterck, L.; Vandepoele, K. The flowering world: A tale of duplications. Trends Plant. Sci. 2009, 14, 680–688. [Google Scholar] [CrossRef] [PubMed]
  54. Jiao, Y.; Leebens-Mack, J.; Ayyampalayam, S.; Bowers, J.E.; McKain, M.R.; McNeal, J.; Rolf, M.; Ruzicka, D.R.; Wafula, E.; Wickett, N.J.; et al. A genome triplication associated with early diversification of the core eudicots. Genome Biol. 2012, 13, R3. [Google Scholar] [CrossRef] [PubMed]
  55. Vanneste, K.; Baele, G.; Maere, S.; Van de Peer, Y. Analysis of 41 plant genomes supports a wave of successful genome duplications in association with the Cretaceous-Paleogene boundary. Genome Res. 2014, 24, 1334–1347. [Google Scholar] [CrossRef] [PubMed]
  56. Chan, A.P.; Crabtree, J.; Zhao, Q.; Lorenzi, H.; Orvis, J.; Puiu, D.; Melake-Berhan, A.; Jones, K.M.; Redman, J.; Chen, G. Draft genome sequence of the oilseed species Ricinus communis. Nat. Biotechnol. 2010, 28, 951–956. [Google Scholar] [CrossRef]
  57. Koonin, E.V. Orthologs, paralogs, and evolutionary genomics. Annu. Rev. Genet. 2005, 39, 309–338. [Google Scholar] [CrossRef] [PubMed]
  58. Carvalho, F.A.; Renner, S.S. The Phylogeny of the Caricaceae. In Genetics and Genomics of Papaya. Plant Genetics and Genomics: Crops and Models; Ming, R., Moore, P., Eds.; Springer: Cham, Switzerland, 2014. [Google Scholar]
  59. Luo, M.C.; You, F.M.; Li, P.; Wang, J.R.; Zhu, T.; Dandekar, A.M.; Leslie, C.A.; Aradhya, M.; McGuire, P.E.; Dvorak, J. Synteny analysis in Rosids with a walnut physical map reveals slow genome evolution in long-lived woody perennials. BMC Genom. 2015, 16, 707. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Chromosomal locations and duplication events of 28 CpLEA genes. Chromosome serial numbers are indicated at the top of each chromosome. CpLEA2-4/-3/-2 are clustered as tandem repeats (lines in green); CpLEA2-1/-4, CpLEA3-4/-5, and CpSMP1/3 are transposed repeats (lines in blue); and CpLEA3-1/-3, CpLEA5-1/-2, and CpSMP1/2 are WGD repeats (lines in red) that are located in syntenic blocks.
Figure 1. Chromosomal locations and duplication events of 28 CpLEA genes. Chromosome serial numbers are indicated at the top of each chromosome. CpLEA2-4/-3/-2 are clustered as tandem repeats (lines in green); CpLEA2-1/-4, CpLEA3-4/-5, and CpSMP1/3 are transposed repeats (lines in blue); and CpLEA3-1/-3, CpLEA5-1/-2, and CpSMP1/2 are WGD repeats (lines in red) that are located in syntenic blocks.
Life 12 01453 g001
Figure 2. Distribution of papaya, horseradish tree, spider flower, arabidopsis, and castor been LEA genes in nine defined gene families.
Figure 2. Distribution of papaya, horseradish tree, spider flower, arabidopsis, and castor been LEA genes in nine defined gene families.
Life 12 01453 g002
Figure 3. Phylogenetic analysis, gene structure, and motif distribution of papaya and arabidopsis LEA genes. (A) Phylogenetic analysis of eight families of Cp/AtLEA proteins; (B) Exon-intron structures of Cp/AtLEA genes; (C) Distribution of 20 conserved motifs. Multiple sequence alignments were conducted using MUSCLE and unrooted phylogenetic trees were constructed using MEGA6 (maximum likelihood method; bootstrap, 1000 replicates; shown are bootstrap values at nodes supported by a posterior probability of ≥50%). Motifs were identified using MEME.
Figure 3. Phylogenetic analysis, gene structure, and motif distribution of papaya and arabidopsis LEA genes. (A) Phylogenetic analysis of eight families of Cp/AtLEA proteins; (B) Exon-intron structures of Cp/AtLEA genes; (C) Distribution of 20 conserved motifs. Multiple sequence alignments were conducted using MUSCLE and unrooted phylogenetic trees were constructed using MEGA6 (maximum likelihood method; bootstrap, 1000 replicates; shown are bootstrap values at nodes supported by a posterior probability of ≥50%). Motifs were identified using MEME.
Life 12 01453 g003
Figure 4. ABRE and LTRE cis-acting elements present in 2000-bp promoter regions of CpLEA genes.
Figure 4. ABRE and LTRE cis-acting elements present in 2000-bp promoter regions of CpLEA genes.
Life 12 01453 g004
Figure 5. Tissue-specific expression profiles of CpLEA genes. Color scale represents FKPM normalized log10 transformed counts, where blue indicates low expression and red indicates high expression.
Figure 5. Tissue-specific expression profiles of CpLEA genes. Color scale represents FKPM normalized log10 transformed counts, where blue indicates low expression and red indicates high expression.
Life 12 01453 g005
Figure 6. Expression profiles of CpLEA genes during fruit development. Color scale represents FKPM normalized log10 transformed counts, where blue indicates low expression and red indicates high expression. (DPA, days post-anthesis; S, stage of developmental fruit).
Figure 6. Expression profiles of CpLEA genes during fruit development. Color scale represents FKPM normalized log10 transformed counts, where blue indicates low expression and red indicates high expression. (DPA, days post-anthesis; S, stage of developmental fruit).
Life 12 01453 g006
Figure 7. Expression profiles of CpLEA genes upon drought stress. The FKPM value of all genes in controls was normalized to one, and the color scale represents normalized log10 transformed fold changes, where blue indicates low expression and red indicates high expression.
Figure 7. Expression profiles of CpLEA genes upon drought stress. The FKPM value of all genes in controls was normalized to one, and the color scale represents normalized log10 transformed fold changes, where blue indicates low expression and red indicates high expression.
Life 12 01453 g007
Figure 8. Expression profiles of CpLEA genes upon cold and salt stresses. The FKPM value of all genes in the controls was normalized to one, and the color scale represents normalized log10 transformed fold changes, where blue indicates low expression and red indicates high expression.
Figure 8. Expression profiles of CpLEA genes upon cold and salt stresses. The FKPM value of all genes in the controls was normalized to one, and the color scale represents normalized log10 transformed fold changes, where blue indicates low expression and red indicates high expression.
Life 12 01453 g008
Table 1. LEA genes identified in papaya.
Table 1. LEA genes identified in papaya.
FamilyGene NameLocusASDeduced Protein
SunsetASGPBv0.4AAMW (kDa)pIGRAVYLoc
LEA_1CpLEA1-1sunset05G0006380evm.TU.supercontig_18.65-16017.019.65−0.755Nucl
CpLEA1-2sunset05G0013060evm.TU.supercontig_41.41-10211.417.03−0.908Mito
CpLEA1-3sunset08G0019430evm.TU.supercontig_85.72Yes15816.078.83−0.878Mito
LEA_2CpLEA2-1sunset05G0003590evm.TU.supercontig_9.242Yes31635.124.69−0.384Cyto
CpLEA2-2sunset05G0009060evm.TU.supercontig_11.66Yes30534.105.38−0.243Chlo
CpLEA2-3sunset05G0009070evm.TU.supercontig_11.68Yes18520.235.65−0.056Chlo
CpLEA2-4sunset05G0009080evm.TU.supercontig_11.69Yes15116.164.750.094Cyto
LEA_3CpLEA3-1sunset03G0023320evm.TU.supercontig_16.192-10311.2110.07−0.472Chlo
CpLEA3-2sunset04G0017920evm.TU.supercontig_25.184Yes9810.949.52−0.526Chlo
CpLEA3-3sunset05G0003680evm.TU.supercontig_9.251Yes9910.619.89−0.531Cyto
CpLEA3-4sunset05G0018090evm.TU.supercontig_2471.1Yes9510.629.66−0.997Mito
CpLEA3-5sunset06G0002130evm.TU.supercontig_200.7-10411.789.69−0.839Cyto
LEA_4CpLEA4-1sunset01G0016400evm.TU.supercontig_66.6-59066.208.91−0.515Extr
CpLEA4-2sunset03G0025310evm.TU.supercontig_209.19-58161.455.20−0.864Nucl
CpLEA4-3sunset05G0000220evm.TU.supercontig_146.20-19321.635.21−1.053Extr
CpLEA4-4sunset07G0004690evm.TU.supercontig_464.2-22224.578.95−1.333Chlo
CpLEA4-5sunset08G0016230evm.TU.supercontig_5.110Yes28030.346.17−1.360Nucl
LEA_5CpLEA5-1sunset02G0011780evm.TU.supercontig_19.160-899.645.51−1.319Cyto
CpLEA5-2sunset08G0009640evm.TU.supercontig_2485.2-11112.105.51−1.338Nucl
LEA_6CpLEA6-1sunset01G0017510evm.TU.supercontig_88.61-9710.425.56−0.705Nucl
CpLEA6-2sunset04G0003310evm.TU.supercontig_6.54-788.775.22−1.573Nucl
DHNCpDHN1sunset01G0014930evm.TU.supercontig_26.225Yes21124.105.05−1.584Nucl
CpDHN2sunset04G0004410evm.TU.supercontig_6.176-13714.769.45−1.222Nucl
CpDHN3sunset06G0003520evm.TU.supercontig_106.3Yes16717.935.94−1.265Nucl
CpDHN4sunset06G0021280evm.TU.supercontig_161.14Yes9310.506.62−1.984Nucl
SMPCpSMP1sunset03G0005590evm.TU.supercontig_58.99-26226.704.70−0.270Chlo
CpSMP2sunset03G0027120evm.TU.supercontig_487.3-26727.974.56−0.246Cyto
CpSMP3sunset06G0024460evm.TU.contig_34050.2-24425.136.44−0.359Nucl
AA, Amino acid; AS, Alternative splicing; Chlo, Chloroplast; Cyto, Cytoplasmic; Extr, Extracellular; GRAVY, Grand average of hydropathicity; Mito, Mitochondria; MW, Molecular weight; Nucl, Nuclear; pI, Isoelectric point; Loc, Subcellular localization.
Table 2. 29 Orthogroups identified in this study.
Table 2. 29 Orthogroups identified in this study.
FamilyOrthogroupPapayaHorseradish TreeSpider FlowerCastor BeenArabidopsis
LEA_1LEA1aCpLEA1-1MoLEA1-1ThLEA1-1
ThLEA1-2
RcLEA1-2AtLEA6
AtLEA18
LEA1bCpLEA1-2MoLEA1-2-RcLEA1-1-
LEA1cCpLEA1-3MoLEA1-3ThLEA1-3
ThLEA1-4
RcLEA1-3AtLEA46
LEA_2LEA2aCpLEA2-1MoLEA2-1ThLEA2-1
ThLEA2-2
RcLEA2-2AtLEA26
LEA2bCpLEA2-2
CpLEA2-3
-ThLEA2-3
ThLEA2-4
ThLEA2-5
ThLEA2-6
--
LEA2cCpLEA2-4MoLEA2-2ThLEA2-7RcLEA2-1AtLEA1
AtLEA27
LEA_3LEA3aCpLEA3-1MoLEA3-1ThLEA3-1RcLEA3-5AtLEA41
LEA3bCpLEA3-2MoLEA3-2ThLEA3-2RcLEA3-4AtLEA37
LEA3cCpLEA3-3MoLEA3-3ThLEA3-3
ThLEA3-4
RcLEA3-1AtLEA2
AtLEA38
LEA3dCpLEA3-4MoLEA3-4ThLEA3-5RcLEA3-2-
LEA3eCpLEA3-5MoLEA3-5-RcLEA3-3-
LEA_4LEA4aCpLEA4-1MoLEA4-1ThLEA4-1-AtLEA9
LEA4bCpLEA4-2MoLEA4-2ThLEA4-2RcLEA4-2AtLEA25
LEA4cCpLEA4-3MoLEA4-3-RcLEA4-4AtLEA30
LEA4dCpLEA4-4MoLEA4-4ThLEA4-3RcLEA4-3AtLEA42
AtLEA48
LEA4eCpLEA4-5MoLEA4-5ThLEA4-4RcLEA4-5AtLEA19
AtLEA36
LEA4f-MoLEA4-6ThLEA4-5
ThLEA4-6
RcLEA4-1AtLEA13
AtLEA43
LEA_5LEA5aCpLEA5-1MoLEA5-1ThLEA5-1RcLEA5-1AtLEA20
LEA5bCpLEA5-2MoLEA5-2
MoLEA5-3
ThLEA5-2RcLEA5-2AtLEA35
LEA_6LEA6aCpLEA6-1MoLEA6-1ThLEA6-1RcLEA6-1AtLEA17
LEA6bCpLEA6-2MoLEA6-2ThLEA6-2RcLEA6-2AtLEA15
AtLEA16
DHNDHNaCpDHN1MoDHN1ThDHN1
ThDHN2
RcDHN1AtLEA4
AtLEA5
AtLEA10
DHNbCpDHN2MoDHN2ThDHN3RcDHN2
RcDHN3
AtLEA33
AtLEA34
AtLEA51
DHNcCpDHN3MoDHN3ThDHN4
ThDHN5
RcDHN4AtLEA14
AtLEA45
DHNdCpDHN4MoDHN4ThDHN6
ThDHN7
ThDHN8
RcDHN5AtLEA8
DHNe-MoDHN5ThDHN9--
SMPSMPaCpSMP1MoSMP1ThSMP1RcSMP3AtLEA31
AtLEA32
SMPbCpSMP2MoSMP2ThSMP2RcSMP1
RcSMP2
AtLEA3
SMPcCpSMP3MoSMP3ThSMP3
ThSMP4
RcSMP4AtLEA47
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zou, Z.; Guo, J.; Zheng, Y.; Xiao, Y.; Guo, A. Genomic Analysis of LEA Genes in Carica papaya and Insight into Lineage-Specific Family Evolution in Brassicales. Life 2022, 12, 1453. https://doi.org/10.3390/life12091453

AMA Style

Zou Z, Guo J, Zheng Y, Xiao Y, Guo A. Genomic Analysis of LEA Genes in Carica papaya and Insight into Lineage-Specific Family Evolution in Brassicales. Life. 2022; 12(9):1453. https://doi.org/10.3390/life12091453

Chicago/Turabian Style

Zou, Zhi, Jingyuan Guo, Yujiao Zheng, Yanhua Xiao, and Anping Guo. 2022. "Genomic Analysis of LEA Genes in Carica papaya and Insight into Lineage-Specific Family Evolution in Brassicales" Life 12, no. 9: 1453. https://doi.org/10.3390/life12091453

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