Next Article in Journal
Essential Role of Potassium in Apple and Its Implications for Management of Orchard Fertilization
Next Article in Special Issue
Role of Autophagy in Haematococcus lacustris Cell Growth under Salinity
Previous Article in Journal
Problems and Prospects of Improving Abiotic Stress Tolerance and Pathogen Resistance of Oil Palm
Previous Article in Special Issue
Electrical Signals, Plant Tolerance to Actions of Stressors, and Programmed Cell Death: Is Interaction Possible?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploration of Autophagy Families in Legumes and Dissection of the ATG18 Family with a Special Focus on Phaseolus vulgaris

by
Elsa-Herminia Quezada-Rodríguez
1,
Homero Gómez-Velasco
2,
Manoj-Kumar Arthikala
1,
Miguel Lara
3,
Antonio Hernández-López
1 and
Kalpana Nanjareddy
1,*
1
Ciencias Agrogenómicas, Escuela Nacional de Estudios Superiores Unidad León, Universidad Nacional Autónoma de México (UNAM), León C.P. 37684, Mexico
2
Instituto de Química, Universidad Nacional Autónoma de México (UNAM), Cuidad Universitaria, Cuidad de Mexico C.P. 04510, Mexico
3
Departamento de Biología Molecular de Plantas, Instituto de Biotecnología, Universidad Nacional Autónoma de México (UNAM), Cuernavaca C.P. 62271, Mexico
*
Author to whom correspondence should be addressed.
Plants 2021, 10(12), 2619; https://doi.org/10.3390/plants10122619
Submission received: 7 September 2021 / Revised: 3 November 2021 / Accepted: 3 November 2021 / Published: 29 November 2021
(This article belongs to the Special Issue Degradation of Plant Organelles and Cell Remodeling during Autophagy)

Abstract

:
Macroautophagy/autophagy is a fundamental catabolic pathway that maintains cellular homeostasis in eukaryotic cells by forming double-membrane-bound vesicles named autophagosomes. The autophagy family genes remain largely unexplored except in some model organisms. Legumes are a large family of economically important crops, and knowledge of their important cellular processes is essential. Here, to first address the knowledge gaps, we identified 17 ATG families in Phaseolus vulgaris, Medicago truncatula and Glycine max based on Arabidopsis sequences and elucidated their phylogenetic relationships. Second, we dissected ATG18 in subfamilies from early plant lineages, chlorophytes to higher plants, legumes, which included a total of 27 photosynthetic organisms. Third, we focused on the ATG18 family in P. vulgaris to understand the protein structure and developed a 3D model for PvATG18b. Our results identified ATG homologs in the chosen legumes and differential expression data revealed the nitrate-responsive nature of ATG genes. A multidimensional scaling analysis of 280 protein sequences from 27 photosynthetic organisms classified ATG18 homologs into three subfamilies that were not based on the BCAS3 domain alone. The domain structure, protein motifs (FRRG) and the stable folding conformation structure of PvATG18b revealing the possible lipid-binding sites and transmembrane helices led us to propose PvATG18b as the functional homolog of AtATG18b. The findings of this study contribute to an in-depth understanding of the autophagy process in legumes and improve our knowledge of ATG18 subfamilies.

1. Introduction

Autophagy is a degradation process essential in the maintenance of homeostasis in eukaryotic cells and is related to a wide variety of physiological and pathophysiological roles, such as host defense, development, infection, and tumorigenesis [1,2]. Autophagy/macroautophagy is a process in which cytosolic components are sequestered within double-membrane vesicles called autophagosomes, which fuse with lysosomes or vacuoles for degradation/recycling [3]. This process is mediated by evolutionarily conserved genes known as autophagy genes (ATGs) [4], which were originally discovered in and isolated from Saccharomyces cerevisiae [5,6,7,8]. Three major intracellular autophagy pathways, namely, macroautophagy, microautophagy and chaperone-mediated autophagy (CMA), have been elucidated, and these differ in the mode of cargo delivery to the lysosome or vacuole [9,10]. Macroautophagy can be nonselective or selective: Nonselective autophagy is a cellular response to nutrient deprivation that involves the random uptake of cytoplasm into phagophores (precursors to autophagosomes) [11], and selective autophagy is responsible for the specific removal of certain components, such as protein aggregates and damaged or superfluous organelles [12,13]. Selective autophagic degradation has been observed with several organelles, such as mitochondria [14], peroxisomes [15], lysosomes [16], endoplasmic reticulum and nucleus [17]. In contrast, microautophagy is the least characterized type of autophagy; during this nonselective process, smaller molecules acting as substrates and the cargo for degradation are transferred into vacuole via invagination of the tonoplast membrane. CMA involves molecular chaperones in the cytosol that unfold proteins and translocate them through the lysosomal membrane [18].
Research on plant autophagy has improved enormously since the first genetic analysis of plant autophagy was performed [19,20,21,22,23,24]. During the process of autophagy, ATG genes play a key role and are classified into several functional groups: The ATG1 kinase complex, the ATG9 recycling complex, the phosphatidylinositol 3-kinase (PI3K) complex and the ATG8 and ATG12 conjugation systems [12].
Autophagy/macroautophagy can be activated under nutrient-depletion conditions via the inhibition of mammalian target of rapamycin (mTOR) or the activation of AMPK. Under TOR-inhibiting conditions, ATG13 is rapidly dephosphorylated, which results in its association with ATG1 and the additional proteins ATG11 and ATG101 and thus stimulation of the autophagy process [25,26]. Phagophore expansion is driven by the transmembrane protein ATG9 along with its cycling factors ATG2 and ATG18 [27,28]. Furthermore, assembly of the phagophore is completed with phosphatidylinositol-3-phosphate (PI3P) by a complex containing class III phosphatidylinositol-3-kinase (PI3K), vacuolar protein sorting 34 (VPS34), ATG/VPS30/beclin-1, VPS38, ATG14 and VPS15 [28]. Phagophore expansion and maturation are completed by ATG8, which is cleaved by cysteine proteinase ATG4 to expose the C-terminal glycine residue [29]. Subsequently, the exposed glycine of ATG8 is conjugated to the membrane lipid phosphatidylethanolamine (PE) via a ubiquitin-like conjugation reaction catalyzed by ATG7 (E1-like enzyme), ATG3 (E2-like enzyme) and the ATG12-ATG5 complex (E3-like enzyme) [30,31,32]. The ATG8-PE adduct can be deconjugated from the membrane by ATG4 proteinase; hence, ATG8 is recycled to participate in new conjugation events [29,33].
ATG18 is an autophagy-related molecule that regulates the vacuolar shape and is conserved from yeast to higher organisms, including the human proteins WIPI1–WIPI4 [34]. While yeast has only one ATG18 gene and two other genes with WD40 repeats, the plant ATG18 family diversifies from two genes in algae to multiple genes in higher plants. The Atg18 protein is characterized by the presence of several WD-40 domains and has been predicted to form a β-propeller structure that binds to phosphatidylinositol 3-phosphate (PtdIns(3)P) and phosphatidylinositol 3,5-bisphosphate (PtdIns(3,5)P2) [35,36,37]. The binding of PtdIns(3)P and Atg18 is needed for the efficient recruitment of Atg8 and Atg16 during phagophore formation at the phagophore assembly site (PAS) [38]. A previous study showed that phagophore formation could also be affected in the absence of the Atg2-Atg18 complex, although other Atg proteins accumulate at the PAS [39]. The Atg2-Atg18 complex has also been shown to localize to a few specific spots on the opening edge of the isolation membrane that lie close to sites for COPII vesicle formation in the endoplasmic reticulum (ER) or ER exit sites [40,41].
Among plants, Arabidopsis contains eight ATG18 homologs, which are classified as AtATG18a–h, and multiple splice variants [42,43], and rice has six ATG18 homologs. AtATG18a is involved in oxidative, drought and salt stress [42,43,44,45]. Recent studies have also suggested the regulation of autophagy by the reversible persulfidation of AtATG18a under ER stress [46]. Similarly, ATG18 is reportedly involved in autophagy regulation under abiotic stress conditions in sweet orange (Citrus sinensis) [47], tomato (Solanum lycopersicum) [48] and apple (Malus domestica) [49,50]. To date, AtATG18a is the only member of the ATG18 family that has been established as an essential component of autophagy in A. thaliana.
Recent studies on ATG genes conducted by Norizuki and colleagues (2019) [51] have shown the diversification of ATGs from early plant lineages to higher plants. However, legumes are a large and economically important family of flowering plants, and few studies have investigated autophagy-related aspects. The aim of the present study was to expand the previous studies to higher clades, specifically to fabaceous plants, and thus understand the current diversity and complexity of ATGs. Furthermore, we focused on the ATG18 family to understand its evolutionary relationships, diversification, expression patterns and cis-regulatory elements in many plants ranging from early plant lineages to fabaceous members. We also performed a comprehensive study of various functional and structural aspects of ATG18b in P. vulgaris.

2. Results

2.1. Identification of ATG families in P. vulgaris, M. truncatula and G. max

In A. thaliana, a total of 39 ATG sequences divided into 17 families have been reported. In the present study, we identified a total of 32 genes in P. vulgaris (2n), 39 genes in M. truncatula (2n) and 61 genes in G. max (4x) (Table 1). A BLAST analysis of Arabidopsis sequences returned 19 (59.37%) homologs in P. vulgaris, 28 (77.77%) homologs in M. truncatula and 30 (48.38%) homologs in G. max with a query coverage of 93–94% and 66–77% identity (Supplementary Information SI1). For this reason, other ortholog analysis databases were used to identify any missing ATG members. The KEGG orthology table for the autophagy pathway was the second main tool because it contains a wide variety of species, and we used this table to obtain more than 70% of genes in P. vulgaris and M. truncatula and 58% in G. max. An analysis of legumes using Ensembl Plants provided more than 70% of ATGs in the legumes under study. Other studies were performed through a HMMER analysis using Ensembl databases and the InParanoid tools in Phytozome. The obtained sequences were verified using Pfam to acquire the positions of the families, domains and repeats, and the protein motifs were determined with MEME. Additional studies were performed using EggNOG, which provided a list of orthologs, particularly in P. vulgaris (Supplementary Figure S1). We also identified 21, 17 and 15 orthologs and 10, 17 and 21 paralogs in P. vulgaris, M. truncatula and G. max, respectively. The genes identified in P. vulgaris, M. truncatula and G. max are listed in Table S1.

2.2. Phylogenetic Relationships, Chromosome Localization, Synteny and Ka/Ks Ratio of ATG Families in Legumes

To understand the evolutionary relationships among ATGs, we generated 17 phylogenetic trees, one for each ATG family in A. thaliana, P. vulgaris, M. truncatula and G. max as per the classification in A. thaliana. The primary protein sequences of A. thaliana, P. vulgaris, M. truncatula and G. max were aligned using Clustal Omega with the default parameters, and phylogenetic trees were obtained with the neighbor-joining method. Each of the ATG sequences was also subjected to a motif analysis, which revealed that the sequences and motifs in all the studied legumes showed high identity to their homologs in Arabidopsis. The phylogenetic tree also revealed that the majority of the ATG family distributions was predominantly composed of Medicago sequences that were more closely related to those in Arabidopsis. Among all the phylogenetic trees of ATGs developed, 11 contained only one clade (ATG2, ATG3, ATG4, ATG5, ATG6, ATG7, ATG10, ATG11, ATG12, ATG14 and ATG101), even if there was more than one isoform, and most of the motif P-values were greater than 1e-100. ATG8 and ATG18 were the families with the highest number of members: ATG18, eight each in Arabidopsis, Medicago and Phaseolus and 19 in G. max; ATG8, nine in Arabidopsis, eight in Medicago, six in P. vulgaris and 10 in G. max. The phylogenetic analysis of ATG8 and ATG18 was divided into three clades with motif P-values between 1 × 10−13 and 1 × 10−90 (Figure 1). The close association of the homologs in all the species studied depicts the conservation of sequences and hence implies biological function conservation.
The chromosome localization of ATGs in the A. thaliana and legume genomes was mapped using Circos (Figure 2). The distribution of ATG homologs among the chromosomes was uneven in all the species compared. Among all 17 families, the maximal number of homologs was located on chromosome 3 in A. thaliana (8) and P. vulgaris (6), chromosome 4 in M. truncatula (6) and chromosomes 4 and 17 in G. max (6).
The Ka/Ks ratio among most of the ATG sequences was lower than 1 (average 0.17), which indicates purifying selection; in contrast, the sequences of ATG8 (1.24) and two sequences (GmATG18e and GmATG18b. I) of ATG18 (1.09 and 1.04) in G. max had values higher than 1, which indicated accelerated evolution and positive selection (Figure 3). The Ka/Ks ratios suggest the conservation of ATG homologs in terms of both sequence and biological function.

2.3. Promoter Analysis and Expression Profiling of ATG Families

Promoter analysis is an important method for understanding the regulatory mechanisms governing ATGs in response to growth and developmental issues and to environmental cues. The analysis of cis-acting elements in the promoters of all 17 ATG families resulted in 44 different transcription factors. The most abundant transcription factors identified were B-Proto-Oncogene-MYB involved in the ABA response and C-Proto-Oncogene-MYC related to jasmonate signaling, and the transcription factors with the motifs ethylene response elements (ERE), TATA box, CAATT-box and G-box were found for all ATGs in A. thaliana, P. vulgaris, M. truncatula and G. max (Supplementary Figure S5). Our results also showed that the ATG8 and ATG18 families contained the highest numbers of MYB, MYC, ERE and Box 4 (ATTAAT) transcription factor-binding sites. Most of the promoters contained MeJA-, SA-, GA- and ABA-responsive elements. Furthermore, light-responsive transcription factors such as BOX-4, G-box, GT1 motif, MRE and ACE were also detected abundantly in most of the families (Figure 4).
Interestingly, we elucidated the influence of nitrogen sources on ATG expression in the legume members P. vulgaris, M. truncatula and G. max due to their ability to establish symbiotic associations with nitrogen-fixing Rhizobia. Gene expression data from the Phytozome database were retrieved for leaf and root tissues under urea as the organic source and nitrate and ammonia as inorganic sources, as depicted in Figure 2. The highest expression of ATGs was recorded in roots treated with ammonia and leaves treated with urea. ATG8i and ATG3 showed the highest abundance in all the treatments, and the lowest expression levels were recorded for ATG18b, e, c and h, ATG2 and ATG2.II in G. max and ATG3 and ATG8c in M. truncatula. The ATG18 family homologs ATG18a.II, ATG18g and ATG18h showed induced expression in all tissues under all treatments (Figure 2 and Figure 5a).
Furthermore, the differential expression analysis of ATGs in P. vulgaris tissues showed very low expression in young pods collected 1 to 4 days post floral senescence, whereas the fix-(inefficient) nodules collected at 21 days showed the most abundant expression of all ATGs. Interestingly, inefficient fixation increased the expression levels compared with those found with efficient fixation. Among all PvATGs, the ATG1, ATG10, ATG13b, ATG18c and ATG18g.I genes showed the lowest expression in all the analyzed tissues, and a total of 16 ATGs were found to be expressed in most of the tissues (Figure 5a; Supplementary Information SI2). Following the interesting observation of ATG expression in nodules, we analyzed the expression of ATGs using our previously generated RNA-seq data of Rhizobium/mycorrhiza-inoculated P. vulgaris roots. The results were interesting: Six ATGs were upregulated and 16 ATGs were downregulated in mycorrhized roots, and nine ATGs were upregulated and 12 ATGs were downregulated in nodulated roots (Figure 5b; Supplementary Information SI2). The expression of ATG10 was found to be specifically induced in mycorrhized roots, ATG12 was highly induced and ATG18g.l was highly suppressed under both symbiotic conditions. The RNA-seq data was validated using RT-qPCR for PvATG2, PvATG8i, PvATG9 and PvATG10.

2.4. Identification of ATG18 Families in Plants

Through an extensive study aiming to identify and analyze the ATG18 family, we selected 27 plant species starting from the early plant lineage Chlorophyta, Charophyta, liverworts, mosses and higher plants such as monocots and dicots. As with other ATGs, the ATG18 family is also well conserved in all the studied plant species; herein, a total of 280 genes and amino acid sequences were identified and retrieved from various databases. Initially, we identified the ATG18 homologs through a BLAST search of NCBI, and we then used the Pfam database to ensure the presence of WD40 repeats in the characteristic ATG18 members. The identified members were named using the aliases registered in the legume information system, NCBI, Phytozome, InParanoid, EGGNOG and Ensembl (Supplementary Information SI3). The genes with the same names were distinguished by adding a Roman numeral: The number I indicated the closest sequence to that in NCBI. For the primitive plants Physcomitrella patens, Chara braunii, Chlamydomonas reinhardtii, Dunaliella salina, Volvox carteri, Klebsormidium nitens, Micromonas pusilla, Ostreococcus lucimarinus, Ostreococcus tauri and Coccomyxa subellipsoidea, we retained the same names that were reported by Norizuki and colleagues [51].
Starting from the most primitive photosynthetic organisms of Chlorophyta, all the members studied had two ATG18 homologs except C. subellipsoidea, which had three ATG18 genes. Charophyta (C. braunii), liverworts (Marchantia polymorpha) and mosses (P. patens) had two, four and eight genes, respectively. Among monocots, we found that Oryza sativa had the lower number of genes (8), and Z. mays had the highest number of genes (31). Arabidopsis had a total of eight ATG18 members, and the 12 legumes considered here together had a total of 180 genes belonging to the ATG18 family. P. sativum had a minimum of six, and a maximum of 27 genes were found in L. angustifolius. The details of the ATG18 homologs in every species are listed in Table 2 and Table 3.

2.5. Principal Component Analysis of the ATG18 Family

Multidimensional scaling analysis using Bios2mds demonstrates the similarity between 280 ATG18 protein sequences from 27 different species. The plot clearly shows that orthologs (genes with closely related sequences and having the same function in different species) are more similar than paralogs (genes that have similar sequences but have different functions in the same species). The plots show that all ATG18 sequences were grouped into three clusters (Figure 6 and Supplementary Figure S3A). The principal components (PCs) allowed us to construct graphs with PC1, PC2 and PC3, and we then applied the K-means method. Cluster I formed a subfamily with ATG18a, c, d and e members from all the higher plant species studied. Cluster II contained only ATG18b homologs, and cluster III contained ATG18f, g and h members. Cluster III consisted of 3 groups: Lower plants formed a distant group, the second group contained the monocot-derived proteins, and the third group harbored all dicots except Arabidopsis, which was more similar to monocots than dicots. Lower plant species were found to be distributed mostly in clusters I and II with the exception of K. nitens, C. subellipsoidea, M. polymorpha and P. patens, which were also grouped in cluster III but exhibited more similarities among themselves than with higher plants. These clusters were named subfamilies I, II and III for convenience.

2.6. Phylogenetic Relationships of the ATG18 Family in Plants

To understand the evolutionary relationship among primitive and advanced dicot plant species, a multiple sequence alignment of 280 ATG18 amino acid sequences was performed. The aligned sequences were used to generate phylogenetic trees based on the maximum likelihood and Bayes methods using MEGA and Phangorn software (Figure 7 and Supplementary Figure S3B). The largest clade was subfamily III followed by subfamily I, which was mainly composed of ATG18 a, c, d and e. Subfamily II harbored ATG18b. Subfamilies II and III consisted of the Bryopsida, Charophyceae, Klebsormidiophyceae, Mamiellophyceae and Trebouxiophyceae plants, which is important for understanding the divergence of ATG18 homologs.

2.7. Analysis of the Primary Structure and the Secondary Structure Predictions of the ATG18 Family in Plants

For the detection of motifs in 280 aa sequences, we identified four main motifs using MEME software. Motif 1 (SGVHLYKLRRGATNAVIQDIAFSHDSQWJAISSSKGTVHIF) contained 41 aa, and the motif sequence matched that of the WD40 family (PF00400) and β propeller clan 186 (CL0186) in the Pfam database. The InterProScan results also showed that motif 1 belongs to the superfamily WD40 (IPR036322), WD40 repeat-like (SSF50978) and breast carcinoma amplified sequence 3 (PTHR13268). Motif 2 (VIAQFRAHTSPISALCFDPSGTLLVTASVHGHNINVFRIMP) contained 41 aa and was similar to motif 1 but contained an additional domain (WD40/YVTN repeat-like domain, IPR015943). Moreover, motifs 3 (VRCSRDRVAVVLATQIYCYBA) and 4 (GYGPMAVGPRWLAYASNPPLLSNTGRLSPQN) did not belong to any protein family (Figure 8).
The motif sequences were further analyzed with PfamScan to identify the repeats, domains and families. Subfamily I was characterized by motifs 1 and 4, which consisted of WD40 and ANAPC4_WD40 repeats. These motifs also had two domains and eight families, although these Pfam family results are not representative of the subfamily. Subfamily II had motifs 1, 2 and 4, and we detected WD40 and ANAPC4_WD40 repeats in all the members. Only the green alga O. tauri contained leucine-rich repeats (LRR9 and LRR4). A total of four domains were identified: Gel_WD40, which was the largest, a defensin domain and PQQ and SecA preprotein crosslinking domains. Subfamily II also consisted of three families in six plants (Figure 9; Supplementary Information SI4).
Subfamily III had all four motifs, and we found PD40 repeats along with WD40 and ANAPC4_WD40 repeats. Among the 27 plant species analyzed, nine of them had 12 domains and ATP synthase was specific Z. mays. Breast carcinoma amplified sequence 3 (BCAS3) is a characteristic domain found in most members (Figure 9; Supplementary Information SI4).
The secondary structure of ATG18 was determined by protein alignment using JPred software. Here, we found that the sequence of ATG18h in A. thaliana was the largest sequence in the alignment with 927 aa. The protein contains seven blades with four beta blades commonly found in the WD40 family (Supplementary Figure S4). This sequence composition was 1% alpha-helix (H), 29% beta-sheet and 68% coil. ATG18 sequences have four antiparallel β-strands, which are named blades [52]. The beta-sheets in ATG18 proteins contain flexible loops that facilitate molecule binding.
AtATG18h has an LHRG sequence in the same place where the alignments have the FRRG sequence, and we found the BCAS3 domain with Phe17 (Figure 10). The sequence alignment performed to identify the FRRG motif revealed that FRRGs appeared in subfamily II, which consists of ATG18b. In addition, subfamily I contained the LRRG or VRRG sequences, whereas subfamily III contained LQRG, LHRG or LYRG sequences. The sequences that appear in ATG18 contain the same pattern of two polar and neutral amino acids in the center of the sequence between two neutral and nonpolar amino acids. ATG18b in subfamily II has the conserved sequence for PtdInsP binding, and other subfamilies likely also show PtdInsP binding (Figure 10, Supplementary Figure S4).

2.8. Microsynteny of ATG18 in P. vulgaris

To explore the origins and evolutionary processes of the P. vulgaris ATG18 family genes, a comparative synteny map between the eight PvATG18 homologs and 15 other genomes was constructed. The species compared in this study were based on their availability in the GCV database. The classification of the ATG18 family was based on the subfamilies obtained by multidimensional scaling (Figure 6).

2.8.1. Subfamily I

ATG18a was highly conserved in all species with the exception of A. ipaensis. SPATA 20 (legfed_v1_0.L_1H5ZXB) is tandemly duplicated in P. vulgaris. In contrast, the lyase dihydroneopterin aldolase (legfed_v1_0.L_2MWVJ4) was only found in P. vulgaris in the syntenic block. Other genes conserved in the syntenic block were related to cell cycle regulation, transcriptional regulation, transcription factors, zinc finger proteins and other structural motifs involved in peroxisomal and mitochondrial import (Supplementary Figure S5A).
ATG18c was not located in the syntenic block in L. albus, M. truncatula, P. sativum or V. angularis. Genes related to ABC transport, vacuolar iron transport, proteins with WD40 repeats involved in protein–protein interactions, cytochrome P450, oxidoreductases and zinc-binding dehydrogenase were highly conserved in the syntenic block. T. pratense and P. lunatus show duplication of oxidoreductases and zinc-binding dehydrogenase family proteins (Supplementary Figure S5B).
ATG18c II was not located in the syntenic block in L. japonicus. Transcriptional regulator SUPERMAN-like (legfed_v1_0.L_Tx802x and legfed_v1_0.L_NLQvfk) were specific to P. vulgaris. Furthermore, an uncharacterized protein (legfed_v1_0.L_2ffJFT) was found to have undergone duplications in G. max, indicating a putative functional role. Pre-mRNA-splicing factor (legfed_v1_0.L_1Bt8v9) was specifically found in milletioid members of legumes, such as P. vulgaris, G. max, G. soja and V. unguiculata (Supplementary Figure S5C).

2.8.2. Subfamily II

ATG18b was not located in L. japonicus or V. angularis. L. japonicus exhibited inversions in the syntenic block involving the synthesis of pectic cell wall components, ATPases and DUF788 proteins, which have been proven to be involved in autophagy regulation. ATG11 was also found in the same syntenic block (Supplementary Figure S5D; Supplementary Information SI5).

2.8.3. Subfamily III

ATG18f.I was identified in most of the species compared, and most of the flanking genes were conserved. An important observation from this syntenic block is the tandem duplication of Histone H2A (legfed_v1_0.L_0mwghf) in all species except Arachis and Lotus. Fe(II)-dependent dioxygenase-like (legfed_v1_0.L_81S90D) was missing in L. albus and L. angustifolia (Supplementary Figure S6A; Supplementary Information SI5).
ATG18g.I was only found in P. vulgaris, C. cajan, G. max, L. japonicus and V. angularis, and in the other species, the circadian clock-regulated growth regulator Zinc knuckle family protein (legfed_v1_0.L_001qtq) was found in the same syntenic block. The most significant feature of this block was the repeated duplication of disease resistance-responsive dirigent-like protein family protein (legfed_v1_0.L_08frmp) in all the species except V. angularis. In Arachis species, the clustering of vacuolar protein-sorting protein (legfed_v1_0.L_0c0sd2) and breast carcinoma amplified sequence 3 protein (legfed_v1_0.L_cdgcy6) with other genes was an important observation (Supplementary Figure S6B; Supplementary Information SI5).
ATG18g.II was missing in L. albus and was well conserved in other species. In Arachis, gene clusters involving FANTASTIC FOUR 3-like (legfed_v1_0.L_xmq5fm) protein were found associated with shoot meristem growth (Supplementary Figure S6C; Supplementary Information SI5).

2.9. ATG18 Protein Characterization

As mentioned previously, ATG18 homologs in P. vulgaris were also divided into three subfamilies with the characteristic motifs FRRG in PvATG18b, VRRG in PvATG18a and PvATG18c, LQRG in PvATG18f and LHRG in PvATG18g. Characterization of the PvATG18 homologs revealed that PvATG18b had the lowest molecular weight, was stable with an isoelectric point of 8.86 and had a high aliphatic index. High-molecular-weight proteins were specifically found in subfamily III (Supplementary Table S2).
Prediction of the subcellular localization of ATG18 homologs showed that ATG18a, c.I, c.II, g.I and g.II were localized in the cytoplasm, and ATG18f.I and f.II were located in the ER membrane and plasma membrane. Only ATG18c homologs were localized to the lumen of lysosomes. ATG18b was unique because it was found in the mitochondrial inner membrane, inner membrane space and ER membrane (Supplementary Table S2). Furthermore, only three of the PvATG18 proteins had a transmembrane helix spanning the aa 44–67 in PvATG18b and located between the aa 12 and 34 in PvATG18f.I and the aa 7 and 26 in PvATG18f.II (Supplementary Figure S7). Furthermore, we predicted the putative phosphorylation sites in PvATG18 homologs and found that these were located on the amino acids threonine and serine in all sequence alignments (Supplementary Figure S8).

2.10. Protein Structure Prediction and Molecular Dynamics Simulation of ATG18b in P. vulgaris

The above-described analysis implies that PvATG18b is the functional ortholog of AtATG18b; hence, we attempted to understand the structure of this protein using the Robetta Server. This model was submitted to 2.1-µs-long unbiased MD to evaluate the predicted protein model (Figure 11a). In the simulation, we monitored the root mean square deviation (RMSD) of the model protein. The graph clearly indicates a change in the RMSD during the first 1.8 µs of simulation, but the RMSD then reached a plateau. This finding indicates that after 1.8 µs of simulation, the 3D structural model of PvATG18b represents a stable folding conformation (Figure 11b). The model shows the seven-bladed β-propeller architecture conserved among the ATG18 family of proteins [52]. The PvATG18 protein structure consists of seven blades formed by antiparallel β-stands connected by short loop regions. The blades are listed with the numbers 1 to 7 beginning at the C-terminus, whereas the β-stands are named with letters from an inner to outer location as A to D. These structures were similar to those observed with the biophysical characterization of PROPPIN ATG18 in Pichia angusta [52]. We also found a CD loop (S269 to T288) located between the two phosphoinositide-binding sites and the FRRG motif at positions F218, R219, R220 and G221 between blades 5 and 6 (Figure 10d). PROPPINs are WD-40 family propeller proteins that act as scaffolds for protein–protein interactions. The binding of PvATG18b to PtdIns(3,5)P2 and PtdIns3P might be mediated by additional protein–protein interactions, as observed in Kluyveromyces lactis [37]. Earlier models of PROPPINS predicted the insertion of two loops into the membrane in a perpendicular orientation in the phagophore membrane through nonspecific electrostatic interactions [53,54]. Our results for PvATG18 reveal the previously reported nonspecific electrostatic interaction in the protein structure and the presence of one transmembrane motif (Figure 10b, c.)

3. Discussion

Autophagy is recognized as a highly selective cellular clearance pathway that helps maintain homeostasis in eukaryotic cells. The genes involved in autophagy are highly conserved from yeast to humans, and the process is the result of the interaction of these ATGs and other associated genes. The number of identified ATGs shows a marked variation among different species. In yeast, a total of 41 genes have been identified to date, and several studies on plant ATGs have also identified a varied number of genes. In the present investigation, we attempted to perform a comprehensive study for identifying ATG families in three important legume species, namely, P. vulgaris, M. truncatula and G. max. Furthermore, we focused on the ATG18 gene family, the largest of all the families, to identify and phylogenetically compare 27 plant species starting from early plant lineages, chlorophytes to higher plants including legumes.

3.1. Autophagy Genes in Legumes Are Highly Conserved

Using Arabidopsis ATGs as a reference, we retrieved ATG homologs in all the species listed in various databases, including Phytozome, and the sequences were confirmed to be affiliated with ATG-like homologs by analyzing their Pfam matches in the Pfam database. We identified a total of 32, 28 and 61 ATG homologs in P. vulgaris, M. truncatula and G. max, respectively. The identified homologs could be classified into 17 families based on their phylogenetic relationships and motifs. The phylogenetic analysis revealed that homologs in Medicago were located closer to Arabidopsis than those in other species. Unlike in yeast, which contains a single copy of each family, many of the gene families have multiple copies. ATG1 has 4, 3, 2 and 6 homologs in Arabidopsis, Medicago, Phaseolus and Glycine, respectively, ATG13 has 2 homologs in Arabidopsis, Medicago and Phaseolus (2 in each) and 4 homologs in G. max, ATG9 has 2 or 4 homologs in Medicago, Phaseolus and G. max and ATG14 and ATG4 have 2 homologs in Arabidopsis and 2 homologs in G. max. The analysis of larger families revealed that ATG8 has 9, 6, 7 and 10 homologs in Arabidopsis, Medicago, Phaseolus and G. max, respectively, and that ATG18 has 8 homologs in Arabidopsis, Medicago and Phaseolus (8 in each) and a maximum of 19 homologs in G. max. Similar results were also obtained with O. sativa [55], Nicotiana tabacum [56], Vitis vinifera [57], Musa acuminate [58] and Setaria italic [59]. However, in most of the families, the homologs were placed in one clade, which clearly showed sequence similarity and the derivation of statistically reliable pairs of possible orthologous proteins sharing similar functions from a common ancestor, consistent with the results from a previous study conducted by Kellogg (2001) [60]. Furthermore, the ATG families identified constituted a relatively complete autophagic machinery in forming the complexes, namely, the ATG1 kinase complex, class III PI3K complex, ATG9 recycling complex, Atg8-lipidation system and Atg12-conjugation system.
ATG17 is an important accessory protein along with ATG31-ATG29, which acts as a scaffold/modulator in linking the ATG1-ATG13 complex to the phagophore assembly site in yeast. Homologs of the ATG17-ATG31-ATG29 subcomplex were not detected in Arabidopsis. However, single orthologs of ATG11 and ATG101 were identified, and ATG11 reportedly contains a short cryptic ATG17-like domain with weak identity to yeast ATG17 [61]. The identification of ATG homologs in the present study revealed one homolog of ATG11 and one homolog of ATG101 in all the legumes analyzed.
For further exploration of the origin and evolutionary process of ATGs, a comparative synteny map that depicted the presence of 160 genes in Arabidopsis and three legumes compared was constructed. The results suggested that the majority of ATGs had a common ancestor. The Ka/Ks ratio is an important genetic parameter for determining whether positive Darwinian selection is related to gene differentiation [62]. Positive Darwinian selection will retain the advantages of nonsynonymous mutations, and purification selection will gradually remove deleterious nonsynonymous mutations. Herein, the Ka/Ks ratio among most of the ATG sequences was lower than 1 (average of 0.17), indicating purifying selection; in contrast, the sequences of ATG8 (1.24) and two ATG18s (1.09 and 1.04) in G. max had higher values, indicating accelerated evolution and positive selection.
Plant macroautophagy is a process in which macromolecules and cellular components are recycled in lytic vacuoles to be reused. Recycling is crucial for the maintenance of cellular homeostasis by acting as a quality control mechanism under nonstressful conditions and is stimulated under stress conditions [63]. Stress-induced autophagy is well documented in some plant species. Our study of the transcription factors binding to the ATGs revealed that several light-responsive transcription factors, such as BOX-4, G-box, GT1-motif, MRE and ACE, were abundant in most of the ATGs. Furthermore, cis-acting elements related to circadian control were also identified. Phytohormones play key roles in different plant processes, including stress responses. The ATGs analyzed exhibited TF-binding sites for EREs, ABA-responsive ABREs, MeJA-responsive CGTCA motifs, auxin-responsive TGA elements and gibberellin-responsive GARE motifs. Ethylene is considered a key regulator of autophagy in petal senescence in petunia, and ERF5 is also shown to induce autophagy by binding to ATG8 and ATG18h under drought stress in tomato. Upregulation of autophagy by low concentrations of salicylic acid is found to delay methyl jasmonate-induced leaf senescence in Arabidopsis [64,65,66]. In addition, several wound-responsive, pathogen-responsive, flavonoid biosynthetic gene regulation-related and meristem-specific elements were also detected. Based on all the results, the involvement of autophagy in the regulation of plant responses to biotic and abiotic stresses is undeniable.

3.2. Autophagy Genes Are Responsive to Nitrate

To assess the differential expression pattern and responsive nature of ATGs to the presence of different nitrate sources, we developed heatmaps using the data retrieved from databases and from a previous RNA-seq analysis performed by our research group. The differential expression pattern in Phaseolus tissues showed that most of the ATGs were expressed in all tested tissues. Nitrogen is an essential component of life that is needed for building proteins and DNA, and despite its abundance in the atmosphere, only limited reserves of soil inorganic nitrogen are accessible to plants, and this nitrogen is primarily in the forms of nitrate and ammonium. Legumes have a unique ability to establish a symbiotic association with nitrogen-fixing rhizobia. Due to our understanding of the evolution of ATGs in legumes, we opted to understand the response of both arial and root tissues of these legumes to different nitrate sources. The expression patterns showed that the highest expression was found in roots treated with ammonia and leaves treated with urea. ATG18 homologs a, g and h were specifically induced in all tissues and by all treatments, indicating the nitrate-responsive nature of these genes.
Furthermore, an analysis of the differential expression patterns of ATGs in Phaseolus tissues revealed that the highest expression level was noted in 21-day fix (-) nodules, which could be due to the involvement of the autophagic process in providing the necessary amino acids for the synthesis of nitrogen in the absence of the symbiont. In yeast and other eukaryotes, it has been proven that nitrogen deficiency induces autophagy. A recent study using yeast cells also suggested that autophagy sustains glutamate and aspartate synthesis during nitrogen starvation [67]. RNA-seq data from early symbiosis with rhizobia and mycorrhizae showed differential ATG expression, and more ATGs were upregulated in rhizobia-inoculated roots than in mycorrhizae-inoculated roots. This analysis provided candidate genes that could play pivotal roles in symbiosis. The involvement of ATG6/beclin has previously been reported in P. vulgaris during rhizobial infection progression and arbuscule maturation [68].

3.3. The ATG18 Family Is Highly Conserved and Has a Broader Sequence-Based Classification

Atg18 is one of the autophagy-related molecules responsible for autophagic processes and is conserved from yeast to higher organisms [34]. ATG18 proteins belong to the PROPPINs (β-propellers that bind polyphosphoinositides) family and work as PI3P effectors. Earlier studies that focused on the identification of ATG genes in primitive and higher plants showed that each family is represented by only one gene for each component of the core autophagy machinery. ATG8 and ATG18 are exceptions and have multiple homologs with lower redundancy in Arabidopsis and P. patens [51].
ATG18 was the family with the highest number of homologs; hence, we chose this family for a comprehensive analysis of the family from the early plant lineage to legumes. The multiple sequence alignment and phylogeny of ATG18 homologs resulted in separation of the homologs into three clades. Each of the clades had subfamily members, as determined by the multidimensional scaling projection of 280 ATG18 homologs in 27 photosynthetic organisms. Unlike previous studies by Norizuki and colleagues [51], the classification of the ATG18 family was not based on the BCAS3 domain alone. Knockout of the BCAS3 gene in Dictyostelium resulted in a reduction in early autophagosomes compared with that found in wild-type cells [69]. In the present study, due to the multidimensional scaling projection of the retrieved sequences, we classified the ATG18 sequences into three subfamilies. Subfamily I contained ATG18a, ATG18c, ATG18d and ATG18e homologs, subfamily II had only ATG18b and subfamily III had ATG18f, ATG18g and ATG18h members. All homologs with BCAS3 were found to be clustered within subfamily III.
Subfamily II, which contained only ATG18b homologs, had few members but was detected in all the plant species investigated in this study, which suggested the sequence and functional conservation of these proteins. Among the early photosynthetic organisms, we identified at least one homolog in subfamilies I and II, but significant divergence was detected, particularly within subfamily III. Among monocots, O. sativa had 8 homologs, whereas 32 and 21 homologs were found in Z. mays and T. aestivum, respectively. The analysis of dicots revealed 8 homologs in each of Arabidopsis, L. japonicus, M. truncatula and P. vulgaris, whereas Arachis sp. had 9 and 10. The maximum number of homologs was recorded in C. cajan (18), G. max (18), C. arietinum (20), Vigna sp. and L. angustifolius (27).
The legume family includes one of the most agroeconomically important plant crops after Poaceae [70]. Of the three subfamilies within Fabaceae, Papilionoideae is the largest, the most recently evolved and monophyletic. Because Papilionoideae includes the most important cultivated legumes, we sought to determine the members of this subfamily in different clades. In the present study, the maximum number of homologs (27) was identified in L. angustifolius, which belongs to the genistoid clade and exhibited an early divergence at approximately 56.4 ± 2 mya. Furthermore, in Arachis species, we found less than half of the ATG18 homologs, indicating possible deletions. Among the members of the next recent (45 mya) clade, which consisted of milletoids, an increase in the number of homologs (18) was detected, which might be due to whole-genome duplication in G. max. However, P. vulgaris had only eight members of ATG18, indicating possible divergence prior to whole-genome duplications, whereas Vigna sp. was found to have high numbers of homologs. Furthermore, more recent robinioid (48.3 ± 1.0 mya) and IRLC (39.0 ± 2.4 mya) clade members had fewer members with the exception of the tribe Vicieae, whose gene numbers were due to genome expansion and related genomic events. In contrast, syntenic relations were not disrupted due to differences in genome sizes [71,72]. A phylogenetic analysis revealed that the ATG18 homologs of Chlorophyta, Charophyta, Marchantiophyta and Bryophyta were always grouped together, and similar results were obtained for monocots and dicots. However, in a comparison of a broad class of species, it is often not simple to precisely define orthologous genes or genomic loci in a straightforward manner, and this analysis is complicated due to gene duplication, recurring polyploidy and extensive genome rearrangement [73].

3.4. The ATG18 Protein Structure Predicts Possible Functional Diversification

In addition, the prediction of the primary and secondary structures of the proteins strengthens the classification of ATG18 proteins into subfamilies. The protein size, motif structure and changes in FRRG motifs among the ATG18 homologs were identified as the fundamental features that contribute to the classification. The changes in the FRRG motifs found in members of subfamily II comprising ATG18b to LRRG, VRRG in subfamily I, LQRG, LHRG or LYRG in subfamily III indicate functional diversification. The WD40 domain is among the top ten most abundant domains in eukaryotic genomes and is also ranked as the top interacting domain in S. cerevisae [74] (Stirnimann et al., 2010). Based on the SMART database, the human genome contains approximately 349 WD40 domain-containing proteins [75]. The presence of the WD40 domain in ATG18 homologs could indicate their involvement in cellular functions. Proteins containing WD40 domains are known to be involved in signal transduction, vesicular trafficking, cytoskeletal assembly, cell cycle control, apoptosis, chromatin dynamics and transcription regulation due to their ability to bind and thus function as interchangeable substrate receptors to target different substrates and recruit different substrates in distinct modes [76]. In C. elegans, ATG18 and WIPI 1/2 (WD-repeat protein interacting with phosphoinositides) in mammals have FRRGs and EPG-6 and WIPI 3/4 have LRRGs. The substitution of the FRRG motif by FTTG and FKKG does not allow PtdInsP binding; however, the changes in LKKG and LTTG still allow PtdInsP binding [77], implying a possible functional diversification of ATG18 homologs. The studies conducted thus far also demonstrate the involvement of ATG18 homologs in abiotic stress responses in plants [42,43,44,45,46,47,48,49,50].

3.5. ATG18 Family in P. vulgaris

In P. vulgaris, a total of eight ATG18 homologs were identified in the current study and were also classified into three subfamilies. While the functional roles of these subfamilies were not determined in this study, the involvement of these proteins in diversified cellular functions cannot be ruled out. All the subfamilies showed conserved phosphorylation sites but different subcellular localizations.
The conserved nature of serine/threonine sites could indicate the functional roles corresponding to several cellular responses in P. vulgaris. In yeast, Pichia pastoris, Atg18 phosphorylation in the loops in the propeller structure of blades 6 and 7 decreases its binding affinity to phosphatidylinositol 3,5-bisphosphate. The association of ATG18 with the vacuolar membrane is inhibited until dephosphorylation [78]. A recent study in Arabidopsis showed that the phosphorylation of ATG18a by brassinosteroid insensitive 1-associated receptor kinase 1 (BAK1) suppresses autophagy and attenuates plant resistance against necrotrophic pathogens [79].
The microsynteny of P. vulgaris ATG18 homologs showed that subfamily I members were highly conserved across the compared species and were flanked by genes involved in cell cycle regulation, transcriptional regulation, cellular transport and metal ion binding. Furthermore, subfamily II was flanked by the ATPase and DUF788 proteins, which have been proven to be involved in autophagy regulation. ATG11, which is a part of the ATG13-ATG1 complex in autophagy initiation, was also found in the same syntenic block. The subfamily III syntenic block contained conserved genes related to histones, circadian clock, growth and vacuolar transport.

3.6. PvATG18b Could Be the Homolog of AtATG18b

In accordance with a well-established fact, the most important feature of ATG18 proteins is the presence of the FRRG motif and its ability to bind to phosphoinositide. Among P. vulgaris ATG18 homologs, the FRRG motif was found only in ATG18b belonging to subfamily II. Hence, we propose PvATG18b as the functional homolog of A. thaliana ATG18b. We also hypothesize that other ATG18 homologs might be involved in other molecular recognition events through binding to surface molecules that play a distinctive role in autophagy, and similar findings have been observed with human ATG18 homologs, e.g., WIPI 1/WIPI 2 with FRRG repeats and WIPI 3/WIPI 4 with LRRG repeats bind to various PtdIns and thus play distinct roles in autophagy [76,80].
We then performed a molecular dynamic simulation of PvATG18b that is unique to ATG models in legumes. Our model shows the stable folding conformation of the seven-bladed β-propeller architecture. PvATG18b is composed of 359 amino acids, and we found the CD loop (S269 to T288) in blade 6. While this loop sequence differs among species, it forms an amphipathic alpha-helix and might insert into a membrane to allow two lipid-binding sites (PtdIns3P and PtdIns(3,5)P2) [81]. Additionally, PvATG18b contains the FRRG repeat and helps form the site for binding to lipids. The FRRG repeat is in F218 to G221 and is conserved in ATG18b to form the PROPPIN family. The FRRG motif (Phe-Arg-Arg-Gly) in ATG18 proteins has been studied in mammals, yeast and C. elegance [79,82]. In Kluyveromyces lactis, the mutation of the blade 6 β3-β4 loop affects the loss of liposome binding, and the flexible loop coordinates two distinct lipid-binding sites [83]. Previous studies with S. cerevisiae have demonstrated that loops A and B of blade 7 are the locations where ATG2 interacts with ATG18. Further research should be performed to understand the interaction of ATG18 with ATG2 and thus ensure the binding site and vacuole scission function of PvATG18b.

4. Materials and Methods

4.1. Identification of ATG Families in Legumes

Arabidopsis (taxid: 3702) ATG family gene sequences were retrieved from the Araport (https://www.araport.org; accessed on 13 May 2020) and TAIR (https://www.arabidopsis.org; accessed on 15 May 2020) databases through Phytozome v.13. Using these sequences, a BLAST [84] (http://www.ncbi.nlm.nih.gov; Stephen et al., 1997; accessed on 19 May 2020) search was conducted to identify the homologs of ATG genes in Phaseolus vulgaris v 2.1 (taxid: 3885), Medicago truncatula Mt4.0v1 (taxid: 3880) and Glycine max Wm82.a2.v1 (taxid: 3847). The stringency of the search was maintained by keeping the mean BLAST results within a query coverage of 93.85% and 67.78% identity.
The detection of homologs was further optimized using other programs, such as KEGG (www.genome.jp/kegg/; accessed on 2 June 2020) [85], Ensembl Plants (https://plants.ensembl.org; accessed on 4 June 2020) [86], HMMER suite server (http://hmmer.org; accessed on 4 June 2020) [87] and InParanoid 4.1 [88]. Additionally, we examined the ontology IDs for all ATG families using KOG (EuKaryotic Orthologous subfamilies) in the EggNOG v5.0 database [89] (http://eggnog.embl.de; accessed on 7 June 2020) and Protein ANalysis THrough Evolutionary Relationships (PANTHER v14.0, http://www.pantherdb.org; accessed on 10 June 2020) and Pfam IDs were identified in Portal v33.1 (http://pfam.xfam.org/about accessed on 30 October 2020).
The ATG18 protein family was studied in 27 photosynthetic organisms, 13 dicots (legumes), 3 monocots and 10 plants through the evolution of land plants from an algal ancestor. We obtained the ATG18 protein sequences of monocotyledonous crops such as Zea mays (taxid: 4577), Triticum aestivum (taxid: 4565) and Oryza sativa (rice, taxid: 4530) and legumes such as Arachis duranensis (peanut, taxid: 130453), Arachis ipaensis (taxid: 130454), Cajanus cajan (taxid: 3821), Lotus japonicus (taxid: 34305), Cicer arietinum (taxid: 3827), Lupinus angustifolius (taxid: 3871), Pisum sativum (pea, taxid: 3888), Vigna angularis (taxid: 3914), Vigna radiata (taxid: 157791) and Trifolium pratense (red clover, taxid: 57577) through a BLAST analysis of the NCBI, Phytozome, LegumeInfo (https://legumeinfo.org; accessed on 18 June 2020), KEGG, InParanoid, Ensembl, EggNOG and Pfam databases. Additionally, we used the Norizuki report of early-divergent plant lineages to extract the ATG18 protein sequences of Bryopsida (Physcomitrella patens, taxid: 3218), Charophyceae (Chara braunii, taxid: 69332), Chlorophyceae (Chlamydomonas reinhardtii, taxid: 3055, Dunaliella salina, taxid: 3046), (Volvox carteri, taxid: 3067), Klebsormidiophyceae (Klebsormidium nitens, taxid: 105231), Mamiellophyceae (Micromonas pusilla, taxid: 38833; Ostreococcus lucimarinus, taxid: 242159; Ostreococcus tauri, taxid: 70448) and Trebouxiophyceae (Coccomyxa subellipsoidea, taxid: 248742) [51].

4.2. Alignment and Phylogenetic Tree Analyses

The protein sequences of ATG families were aligned using Clustal Omega (1.2.4) [90] (www.clustal.org and www.ebi.ac.uk; accessed on 5 July 2020) with the default parameters. The phylogenetic tree was a neighbor-joining tree without distance corrections, and we extracted the outputs from the tree and generated circular phylogram and cladogram tree images using EvolView. The different phylogenetic trees were combined with the MEME results for all sequences, and the final details were obtained using Inkscape software [91] (https://www.evolgenius.info/evolview/; accessed on 6 July 2020).
Multiple sequence alignment of 280 intraspecies protein sequences of ATG18 family members was performed using Clustal Omega. The phylogenetic analysis was performed using MEGA X with the maximum likelihood method and Bayes analyses with 1000 bootstrap replicates and the default parameters [92]. Phangorn and APE packages in R were used to build the phylogenetic trees [93,94]. In Phangorn, we used the Akaike information criterion and the Whelan and Goldman matrix (WAG) as the substitution model.

4.3. Chromosome Localization, Synteny and Ka/Ks Calculation

The chromosomal localization of ATG family homologs in A. thaliana, P. vulgaris, M. truncatula and G. max was verified using NCBI. Furthermore, Ensembl Plants was used to compare and explore the gene alignments and generate a segment to link the genomes. The synteny relation of ATG genes was drawn using OmicCircos in R36 [95]. The macro- and microsynteny of the ATG18 family was developed using the Genome Context Viewer (GCV) in the Legume information system [96] (https://legumeinfo.org/lis_context_viewer/instructions; accessed on 16 July 2020).
The CDSs and protein sequences were obtained from Phytozome and used to calculate the synonymous substitutions (Ks) and nonsynonymous substitutions (Ka) with TBtools software (https://github.com/CJ-Chen/TBtools; accessed on 26 July 2020). Using the data table, we developed a graph of the Ka and Ks values for all ATG families in P. vulgaris, M. truncatula and G. max using the ggplot2 R packages (https://ggplot2.tidyverse.org/; accessed on 28 July 2020).

4.4. Promoter Analysis, Expression Profiling and Transcriptome of ATG Families

The 2000-bp upstream sequences of ATG genes were retrieved from Phytozome, and these sequences were used as query sequences in PlantCARE software (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/; accessed on 2 August 2020). The results were analyzed, and the most abundant transcription factors were identified using ggplot2 in R.
ATG gene expression data for A. thaliana, M. truncatula and G. max were extracted from Phytozome to determine the differential expression of the genes under different nitrogen treatments [97]. Data on the differential expression of genes in P. vulgaris under nitrogen treatments and after fixation and inoculation with Rhizobium tropici (CIAT899) were obtained from the PvGEA website (https://plantgrn.noble.org/PvGEA/; accessed on 2 July 2020).
We calculated the log2 values of the RPKM values for the comparison. To show the data for A. thaliana, M. truncatula and G. max, we used the OmicCircos package and constructed subfamilies using the synteny graph. However, for P. vulgaris, we constructed an independent heatmap of ggplot2 because the amounts of treatments and tissues were higher. The expression data for ATGs under rhizobial and mycorrhizal symbiotic conditions were obtained from our previous global transcriptomic analysis [98]. A heatmap of the fold change values was constructed using the ggplot2 package.

4.5. Quantitative Real-Time PCR Analysis

Four genes were selected for RT-qPCR analysis, which was performed to validate the RNA-seq data. High-quality total RNA was isolated from frozen root tissues using TRIzol reagent (Sigma) according to the manufacturer’s instructions. RNA integrity was verified by gel electrophoresis and RNA concentration was assessed using a NanoDrop spectrophotometer (Thermo Scientific). RNA was treated with DNase to eliminate DNA contamination (1 U/μL; Roche, USA) according to the manufacturer’s instructions. Reverse-transcription quantitative PCR (RT-qPCR) analysis was performed using a DNA-free RNA and iScriptTM One-Step RT-PCR Kit with SYBR® Green (Bio-Rad) according to the manufacturer’s instructions. To confirm the absence of DNA contamination, a sample lacking reverse transcriptase was included. Relative expression values were calculated using the 2-ΔCt method, where the quantification cycle (Cq) value equals the Cq value of the gene of interest minus the Cq value of the reference gene [99]. Gene-specific primers were used for RT-qPCR analysis (Table S3). PvEF1α and PvIDE were used as reference as described previously by Arthikala et al. [100]. The relative expression values were normalized with respect to two reference genes EF1α and IDE as described previously by Vandesompele et al. [101]. The values presented are averages of three biological replicates, and each data set was recorded using triplicate samples.

4.6. Principal Components Analysis of the ATG18 Family

Based on multiple alignments of ATG18 protein sequences, we converted the information into a distance matrix calculated using the bios2mds packages (https://CRAN.R-project.org/package=bios2mds; accessed on 3 July 2020) in R. The matrix used was BLOSUM62 (BLOcks of Amino Acid SUbstitution Matrix), and sequences with 62% identity were obtained. Using the same packages, we obtain the K-means and principal components to generate the multidimensional scaling projection and thus define the subfamilies within the protein family.

4.7. Detection of Motifs, Domains, Repeats, Families and Secondary Protein Structure of the ATG18 Family

ATG sequences were analyzed for a repeated sequence motif pattern using Multiple Expectation Maximization for Motif Elicitation [102] (http://meme-suite.org/tools/meme; accessed on 18 July 2020) in the classical motif discovery mode and using a limit of three motifs. The secondary structures of the proteins were developed after alignment with Clustal Omega using the online tool JPred in FASTA format. To obtain the repeats, domains and families, a Pfam scan of EMBL-EBI was performed (https://www.ebi.ac.uk/Tools/pfa/pfamscan/; accessed on 26 August 2020).

4.8. Microsynteny and Protein Sequence Parameters of ATG18 in P. vulgaris

The computed parameters for PvATG18, including the molecular weight, theoretical pI, amino acid composition, atomic composition, extinction coefficient, estimated half-life, instability index, aliphatic index, grand average of hydropathicity (GRAVY), phosphorylation sites, predicted transmembrane helixes and subcellular localization, were obtained using ProtParam, PSORT, THMHMM and NetPhos 51 (https://web.expasy.org; accessed on 5 July 2020). The ATG positions were extracted from Phytozome, and microsynteny calculations were generated using GCV v1.2.0 [103] (https://legumeinfo.org/lis_context_viewer/; accessed on 6 August 2020).

4.9. ATG18b Protein in P. vulgaris

The 3D structure of the PvATG18b protein was determined using the Robetta server [102]. Comparative models were built from structures detected and aligned using HHSEARCH, SPARKS and Raptor [104,105,106,107]. The loop regions were assembled from fragments and optimized to fit the aligned template structures. The final structure prediction was selected using the lowest-energy model as determined by a low-resolution Rosetta energy function. The final 3D image was colored with Quimera [108].

5. Conclusions

The present study was carried out to understand the diversification of ATG genes during plant evolution with special emphasis on legumes and P. vulgaris. In the present study, we identified 32, 39 and 61 core ATG genes in P. vulgaris, M. truncatula and G. max, respectively. The ATG genes were conserved across the species, but the higher plants revealed great redundancy. Most of the ATGs in Phaseolus were found to be nitrate responsive and were differentially expressed under rhizobial and mycorrhizal symbiosis, implying their possible role during symbiosis. Further, analysis ATG18 of the family in 27 photosynthetic organisms showed their classification into three subfamilies based on the sequence. In Phaseolus, ATG18 members belonging to all the three subfamilies were conserved. Comparison of Phaseolus ATG18b structure to the crystal structure in Arabidopsis showed conserved FRRG sequence.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/plants10122619/s1, Figure S1: Percentage of legume ATG homologs in different software programs. Figure S2: Validation of expression patterns of ATGs of symbiont-colonized P. vulgaris roots by RT-qPCR analysis. Figure S3A: Representation of 280 ATG18 proteins from different plant species analyzed by multidimensional scaling using Bios2mds. Figure S3B: Phylogenetic tree of ATG18 in plants. Figure S4: Secondary structure prediction using JPred. Figure S5: Microsynteny analysis of ATG18 (Subfamily I & II) in P. vulgaris. Figure S6: Microsynteny analysis of ATG18 (Subfamily III) in P. vulgaris. Figure S7: Phosphorylation sites of ATG18 in P. vulgaris identified using NetPhos. Figure S8: Prediction of transmembrane helices in PvATG18 proteins using TMHMM. Table S1: List of identifiers of the genes, transcripts, and proteins of each ATG in P. vulgaris, Table S2: ATG18 protein characterization in P. vulgaris. Table S3: List of Oligos for RT-Qpcr. Supplementary information: Supplementary Information SI1: Analysis of ATG genes homologs in P. vulgaris, M. truncatula, G. max in different databases; Supplementary Information SI2: Expression profiles of ATGs in P. vulgaris; Supplementary Information SI3: Analysis of ATG18 homologs; Supplementary Information SI4: Family, repeats, motifs and domain positions in legumes; Supplementary Information SI5: Alignment and synteny of ATG genes between A. thaliana and legumes using the comparative genomics in Ensembl.

Author Contributions

Conceptualization, K.N. and E.-H.Q.-R.; methodology, E.-H.Q.-R., H.G.-V. and K.N.; software, E.-H.Q.-R.; validation, K.N., M.-K.A. and A.H.-L.; formal analysis, E.-H.Q.-R. and H.G.-V.; investigation, K.N. and E.-H.Q.-R.; resources, E.-H.Q.-R. and H.G.-V.; writing—original draft preparation, E.-H.Q.-R.; writing—review and editing, K.N., E.-H.Q.-R., M.-K.A. and M.L.; visualization, K.N., E.-H.Q.-R. and M.-K.A.; supervision, K.N.; project administration, K.N.; funding acquisition, K.N., M.-K.A. and M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Dirección General de Asuntos del Personal Académico, DGAPA/PAPIIT-UNAM grant no. IN211218 to K.N. Partially supported by CONACyT project CF-MI-20191017134234199/316538 to M.-K.A., DGAPA/PAPIIT-UNAM grant no. IN216321 to K.N. and DGAPA/PAPIIT-UNAM grant no. IN205619 to M.L.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data reported in this study are available in the supplementary information provided in the supplementary data.

Acknowledgments

Elsa Herminia Quezada Rodríguez is a student from the Programa de Doctorado en Ciencias Biomédicas, Universidad Nacional Autónoma de México (UNAM) and has received CONACyT fellowship 409344. We greatly acknowledge technical support by Martín Munguía Ortiz.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, L.; Ye, X.; Zhao, T. The Physiological Roles of Autophagy in the Mammalian Life Cycle. Biol. Rev. Camb. Philos. Soc. 2019, 94, 503–516. [Google Scholar] [CrossRef] [Green Version]
  2. Gou, W.; Li, X.; Guo, S.; Liu, Y.; Li, F.; Xie, Q. Autophagy in Plant: A New Orchestrator in the Regulation of the Phytohormones Homeostasis. Int. J. Mol. Sci. 2019, 20, 2900. [Google Scholar] [CrossRef] [Green Version]
  3. Klionsky, D.J.; Emr, S.D. Autophagy as a Regulated Pathway of Cellular Degradation. Science 2000, 290, 1717–1721. [Google Scholar] [CrossRef]
  4. Klionsky, D.J.; Cregg, J.M.; Dunn, W.A.; Emr, S.D.; Sakai, Y.; Sandoval, I.V.; Sibirny, A.; Subramani, S.; Thumm, M.; Veenhuis, M.; et al. A Unified Nomenclature for Yeast Autophagy-Related Genes. Dev. Cell 2003, 5, 539–545. [Google Scholar] [CrossRef] [Green Version]
  5. Takeshige, K.; Baba, M.; Tsuboi, S.; Noda, T.; Ohsumi, Y. Autophagy in Yeast Demonstrated with Proteinase-Deficient Mutants and Conditions for Its Induction. J. Cell Biol. 1992, 119, 301–311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Tsukada, M.; Ohsumi, Y. Isolation and Characterization of Autophagy-Defective Mutants of Saccharomyces Cerevisiae. FEBS Lett. 1993, 333, 169–174. [Google Scholar] [CrossRef] [Green Version]
  7. Harding, T.M.; Morano, K.A.; Scott, S.V.; Klionsky, D.J. Isolation and Characterization of Yeast Mutants in the Cytoplasm to Vacuole Protein Targeting Pathway. J. Cell Biol. 1995, 131, 591–602. [Google Scholar] [CrossRef] [Green Version]
  8. Thumm, M.; Egner, R.; Koch, B.; Schlumpberger, M.; Straub, M.; Veenhuis, M.; Wolf, D.H. Isolation of Autophagocytosis Mutants of Saccharomyces Cerevisiae. FEBS Lett. 1994, 349, 275–280. [Google Scholar] [CrossRef] [Green Version]
  9. Xie, Z.; Klionsky, D.J. Autophagosome Formation: Core Machinery and Adaptations. Nat. Cell Biol. 2007, 9, 1102–1109. [Google Scholar] [CrossRef]
  10. González-Polo, R.A.; Pizarro-Estrella, E.; Yakhine-Diop, S.M.S.; Rodríguez-Arribas, M.; Gómez-Sánchez, R.; Casado-Naranjo, I.; Bravo-San Pedro, J.M.; Fuentes, J.M. The Basics of Autophagy. In Autophagy Networks in Inflammation; Maiuri, M.C., De Stefano, D., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 3–20. [Google Scholar] [CrossRef]
  11. Thompson, A.R.; Vierstra, R.D. Autophagic Recycling: Lessons from Yeast Help Define the Process in Plants. Curr. Opin. Plant Biol. 2005, 8, 165–173. [Google Scholar] [CrossRef]
  12. Li, F.; Vierstra, R.D. Autophagy: A Multifaceted Intracellular System for Bulk and Selective Recycling. Trends Plant Sci. 2012, 17, 526–537. [Google Scholar] [CrossRef]
  13. Marshall, R.S.; Vierstra, R.D. Autophagy: The Master of Bulk and Selective Recycling. Annu. Rev. Plant Biol. 2018, 69, 173–208. [Google Scholar] [CrossRef]
  14. Ashrafi, G.; Schwarz, T.L. The Pathways of Mitophagy for Quality Control and Clearance of Mitochondria. Cell Death Differ. 2013, 20, 31–42. [Google Scholar] [CrossRef] [Green Version]
  15. Hutchins, M.U.; Veenhuis, M.; Klionsky, D.J. Peroxisome Degradation in Saccharomyces Cerevisiae Is Dependent on Machinery of Macroautophagy and the Cvt Pathway. J. Cell Sci. 1999, 112, 4079–4087. [Google Scholar] [CrossRef]
  16. Hung, Y.-H.; Chen, L.M.-W.; Yang, J.-Y.; Yuan Yang, W. Spatiotemporally Controlled Induction of Autophagy-Mediated Lysosome Turnover. Nat. Commun. 2013, 4, 2111. [Google Scholar] [CrossRef] [Green Version]
  17. Nakatogawa, H.; Mochida, K. Reticulophagy and Nucleophagy: New Findings and Unsolved Issues. Autophagy 2015, 11, 2377–2378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Dice, J.F. Chaperone-Mediated Autophagy. Autophagy 2007, 3, 295–299. [Google Scholar] [CrossRef] [Green Version]
  19. Doelling, J.H.; Walker, J.M.; Friedman, E.M.; Thompson, A.R.; Vierstra, R.D. The APG8/12-Activating Enzyme APG7 Is Required for Proper Nutrient Recycling and Senescence in Arabidopsis Thaliana. J. Biol. Chem. 2002, 277, 33105–33114. [Google Scholar] [CrossRef] [Green Version]
  20. Hanaoka, H.; Noda, T.; Shirano, Y.; Kato, T.; Hayashi, H.; Shibata, D.; Tabata, S.; Ohsumi, Y. Leaf Senescence and Starvation-Induced Chlorosis Are Accelerated by the Disruption of an Arabidopsis Autophagy Gene. Plant Physiol. 2002, 129, 1181–1193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Minina, E.A.; Filonova, L.H.; Fukada, K.; Savenkov, E.I.; Gogvadze, V.; Clapham, D.; Sanchez-Vera, V.; Suarez, M.F.; Zhivotovsky, B.; Daniel, G.; et al. Autophagy and Metacaspase Determine the Mode of Cell Death in Plants. J. Cell Biol. 2013, 203, 917–927. [Google Scholar] [CrossRef] [Green Version]
  22. Gao, C.; Zhuang, X.; Cui, Y.; Fu, X.; He, Y.; Zhao, Q.; Zeng, Y.; Shen, J.; Luo, M.; Jiang, L. Dual Roles of an Arabidopsis ESCRT Component FREE1 in Regulating Vacuolar Protein Transport and Autophagic Degradation. Proc. Natl. Acad. Sci. USA 2015, 112, 1886–1891. [Google Scholar] [CrossRef] [Green Version]
  23. Marshall, R.S.; Vierstra, R.D. Eat or Be Eaten: The Autophagic Plight of Inactive 26S Proteasomes. Autophagy 2015, 11, 1927–1928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hafrén, A.; Macia, J.-L.; Love, A.J.; Milner, J.J.; Drucker, M.; Hofius, D. Selective Autophagy Limits Cauliflower Mosaic Virus Infection by NBR1-Mediated Targeting of Viral Capsid Protein and Particles. Proc. Natl. Acad. Sci. USA 2017, 114, E2026–E2035. [Google Scholar] [CrossRef] [Green Version]
  25. Suttangkakul, A.; Li, F.; Chung, T.; Vierstra, R.D. The ATG1/ATG13 Protein Kinase Complex Is Both a Regulator and a Target of Autophagic Recycling in Arabidopsis. Plant Cell 2011, 23, 3761–3779. [Google Scholar] [CrossRef] [Green Version]
  26. Li, F.; Chung, T.; Vierstra, R.D. Autophagy-Related11 Plays a Critical Role in General Autophagy- and Senescence-Induced Mitophagy in Arabidopsis. Plant Cell 2014, 26, 788–807. [Google Scholar] [CrossRef] [Green Version]
  27. Zhuang, X.; Chung, K.P.; Cui, Y.; Lin, W.; Gao, C.; Kang, B.-H.; Jiang, L. ATG9 Regulates Autophagosome Progression from the Endoplasmic Reticulum in Arabidopsis. Proc. Natl. Acad. Sci. USA 2017, 114, E426–E435. [Google Scholar] [CrossRef] [Green Version]
  28. Zhuang, X.; Chung, K.P.; Luo, M.; Jiang, L. Autophagosome Biogenesis and the Endoplasmic Reticulum: A Plant Perspective. Trends Plant Sci. 2018, 23, 677–692. [Google Scholar] [CrossRef]
  29. Yoshimoto, K.; Hanaoka, H.; Sato, S.; Kato, T.; Tabata, S.; Noda, T.; Ohsumi, Y. Processing of ATG8s, Ubiquitin-Like Proteins, and Their Deconjugation by ATG4s Are Essential for Plant Autophagy. Plant Cell 2004, 16, 2967–2983. [Google Scholar] [CrossRef]
  30. Thompson, A.R.; Doelling, J.H.; Suttangkakul, A.; Vierstra, R.D. Autophagic Nutrient Recycling in Arabidopsis Directed by the ATG8 and ATG12 Conjugation Pathways. Plant Physiol. 2005, 138, 2097–2110. [Google Scholar] [CrossRef] [Green Version]
  31. Phillips, A.R.; Suttangkakul, A.; Vierstra, R.D. The ATG12-Conjugating Enzyme ATG10 Is Essential for Autophagic Vesicle Formation in Arabidopsis Thaliana. Genetics 2008, 178, 1339–1353. [Google Scholar] [CrossRef] [Green Version]
  32. Chung, T.; Phillips, A.R.; Vierstra, R.D. ATG8 Lipidation and ATG8-Mediated Autophagy in Arabidopsis Require ATG12 Expressed from the Differentially Controlled ATG12A AND ATG12B Loci. Plant J. 2010, 62, 483–493. [Google Scholar] [CrossRef]
  33. Woo, J.; Park, E.; Dinesh-Kumar, S.P. Differential Processing of Arabidopsis Ubiquitin-like Atg8 Autophagy Proteins by Atg4 Cysteine Proteases. Proc. Natl. Acad. Sci. USA 2014, 111, 863–868. [Google Scholar] [CrossRef] [Green Version]
  34. Mizushima, N.; Yoshimori, T.; Ohsumi, Y. The Role of Atg Proteins in Autophagosome Formation. Annu. Rev. Cell Dev. Biol. 2011, 27, 107–132. [Google Scholar] [CrossRef]
  35. Dove, S.K.; Piper, R.C.; McEwen, R.K.; Yu, J.W.; King, M.C.; Hughes, D.C.; Thuring, J.; Holmes, A.B.; Cooke, F.T.; Michell, R.H.; et al. Svp1p Defines a Family of Phosphatidylinositol 3,5-Bisphosphate Effectors. EMBO J. 2004, 23, 1922–1933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Strømhaug, P.E.; Reggiori, F.; Guan, J.; Wang, C.-W.; Klionsky, D.J. Atg21 Is a Phosphoinositide Binding Protein Required for Efficient Lipidation and Localization of Atg8 during Uptake of Aminopeptidase I by Selective Autophagy. Mol. Biol. Cell 2004, 15, 3553–3566. [Google Scholar] [CrossRef] [Green Version]
  37. Krick, R.; Busse, R.A.; Scacioc, A.; Stephan, M.; Janshoff, A.; Thumm, M.; Kühnel, K. Structural and Functional Characterization of the Two Phosphoinositide Binding Sites of PROPPINs, a β-Propeller Protein Family. Proc. Natl. Acad. Sci. USA 2012, 109, E2042–E2049. [Google Scholar] [CrossRef] [Green Version]
  38. Nair, U.; Yen, W.-L.; Mari, M.; Cao, Y.; Xie, Z.; Baba, M.; Reggiori, F.; Klionsky, D.J. A Role for Atg8–PE Deconjugation in Autophagosome Biogenesis. Autophagy 2012, 8, 780–793. [Google Scholar] [CrossRef] [Green Version]
  39. Suzuki, K.; Kubota, Y.; Sekito, T.; Ohsumi, Y. Hierarchy of Atg Proteins in Pre-Autophagosomal Structure Organization. Genes Cells 2007, 12, 209–218. [Google Scholar] [CrossRef]
  40. Graef, M.; Friedman, J.R.; Graham, C.; Babu, M.; Nunnari, J. ER Exit Sites Are Physical and Functional Core Autophagosome Biogenesis Components. Mol. Biol. Cell 2013, 24, 2918–2931. [Google Scholar] [CrossRef]
  41. Suzuki, K.; Akioka, M.; Kondo-Kakuta, C.; Yamamoto, H.; Ohsumi, Y. Fine Mapping of Autophagy-Related Proteins during Autophagosome Formation in Saccharomyces Cerevisiae. J. Cell Sci. 2013, 126, 2534–2544. [Google Scholar] [CrossRef] [Green Version]
  42. Xiong, Y.; Contento, A.L.; Bassham, D.C. AtATG18a Is Required for the Formation of Autophagosomes during Nutrient Stress and Senescence in Arabidopsis Thaliana. Plant J. 2005, 42, 535–546. [Google Scholar] [CrossRef]
  43. Bassham, D.C.; Laporte, M.; Marty, F.; Moriyasu, Y.; Ohsumi, Y.; Olsen, L.J.; Yoshimoto, K. Autophagy in Development and Stress Responses of Plants. Autophagy 2006, 2, 2–11. [Google Scholar] [CrossRef] [PubMed]
  44. Xiong, Y.; Contento, A.L.; Nguyen, P.Q.; Bassham, D.C. Degradation of Oxidized Proteins by Autophagy during Oxidative Stress in Arabidopsis. Plant Physiol. 2007, 143, 291–299. [Google Scholar] [CrossRef] [Green Version]
  45. Liu, Y.; Xiong, Y.; Bassham, D.C. Autophagy Is Required for Tolerance of Drought and Salt Stress in Plants. Autophagy 2009, 5, 954–963. [Google Scholar] [CrossRef] [Green Version]
  46. Aroca, A.; Yruela, I.; Gotor, C.; Bassham, D.C. Persulfidation of ATG18a Regulates Autophagy under ER Stress in Arabidopsis. Proc. Natl. Acad. Sci. USA 2021, 118, e2023604118. [Google Scholar] [CrossRef] [PubMed]
  47. Fu, X.-Z.; Zhou, X.; Xu, Y.-Y.; Hui, Q.-L.; Chun, C.-P.; Ling, L.-L.; Peng, L.-Z. Comprehensive Analysis of Autophagy-Related Genes in Sweet Orange (Citrus Sinensis) Highlights Their Roles in Response to Abiotic Stresses. Int. J. Mol. Sci. 2020, 21, 2699. [Google Scholar] [CrossRef] [Green Version]
  48. Wang, Y.; Cai, S.; Yin, L.; Shi, K.; Xia, X.; Zhou, Y.; Yu, J.; Zhou, J. Tomato HsfA1a Plays a Critical Role in Plant Drought Tolerance by Activating ATG Genes and Inducing Autophagy. Autophagy 2015, 11, 2033–2047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Sun, X.; Wang, P.; Jia, X.; Huo, L.; Che, R.; Ma, F. Improvement of Drought Tolerance by Overexpressing MdATG18a Is Mediated by Modified Antioxidant System and Activated Autophagy in Transgenic Apple. Plant Biotechnol. J. 2018, 16, 545–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Huo, L.; Sun, X.; Guo, Z.; Jia, X.; Che, R.; Sun, Y.; Zhu, Y.; Wang, P.; Gong, X.; Ma, F. MdATG18a Overexpression Improves Basal Thermotolerance in Transgenic Apple by Decreasing Damage to Chloroplasts. Hortic. Res. 2020, 7, 1–15. [Google Scholar] [CrossRef] [Green Version]
  51. Norizuki, T.; Kanazawa, T.; Minamino, N.; Tsukaya, H.; Ueda, T. Marchantia Polymorpha, a New Model Plant for Autophagy Studies. Front. Plant Sci. 2019, 10, 935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Scacioc, A.; Schmidt, C.; Hofmann, T.; Urlaub, H.; Kühnel, K.; Pérez-Lara, Á. Structure Based Biophysical Characterization of the PROPPIN Atg18 Shows Atg18 Oligomerization upon Membrane Binding. Sci. Rep. 2017, 7, 14008. [Google Scholar] [CrossRef] [Green Version]
  53. Lemmon, M.A. Membrane Recognition by Phospholipid-Binding Domains. Nat. Rev. Mol. Cell Biol. 2008, 9, 99–111. [Google Scholar] [CrossRef]
  54. Dove, S.K.; Dong, K.; Kobayashi, T.; Williams, F.K.; Michell, R.H. Phosphatidylinositol 3,5-Bisphosphate and Fab1p/PIKfyve UnderPPIn Endo-Lysosome Function. Biochem. J. 2009, 419, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Xia, K.; Liu, T.; Ouyang, J.; Wang, R.; Fan, T.; Zhang, M. Genome-Wide Identification, Classification, and Expression Analysis of Autophagy-Associated Gene Homologues in Rice (Oryza Sativa, L.). DNA Res. 2011, 18, 363–377. [Google Scholar] [CrossRef] [Green Version]
  56. Zhou, X.; Zhao, P.; Wang, W.; Zou, J.; Cheng, T.; Peng, X.; Sun, M. A Comprehensive, Genome-Wide Analysis of Autophagy-Related Genes Identified in Tobacco Suggests a Central Role of Autophagy in Plant Response to Various Environmental Cues. DNA Res. 2015, 22, 245–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Shangguan, L.; Fang, X.; Chen, L.; Cui, L.; Fang, J. Genome-Wide Analysis of Autophagy-Related Genes (ARGs) in Grapevine and Plant Tolerance to Copper Stress. Planta 2018, 247, 1449–1463. [Google Scholar] [CrossRef]
  58. Wei, Y.; Liu, W.; Hu, W.; Liu, G.; Wu, C.; Liu, W.; Zeng, H.; He, C.; Shi, H. Genome-Wide Analysis of Autophagy-Related Genes in Banana Highlights MaATG8s in Cell Death and Autophagy in Immune Response to Fusarium Wilt. Plant Cell Rep. 2017, 36, 1237–1250. [Google Scholar] [CrossRef] [PubMed]
  59. Li, W.; Chen, M.; Wang, E.; Hu, L.; Hawkesford, M.J.; Zhong, L.; Chen, Z.; Xu, Z.; Li, L.; Zhou, Y.; et al. Genome-Wide Analysis of Autophagy-Associated Genes in Foxtail Millet (Setaria Italica, L.) and Characterization of the Function of SiATG8a in Conferring Tolerance to Nitrogen Starvation in Rice. BMC Genom. 2016, 17, 797. [Google Scholar] [CrossRef] [Green Version]
  60. Kellogg, E.A. Evolutionary History of the Grasses1. Plant Physiol. 2001, 125, 1198–1205. [Google Scholar] [CrossRef] [Green Version]
  61. Li, F.; Vierstra, R.D. Arabidopsis ATG11, a Scaffold That Links the ATG1-ATG13 Kinase Complex to General Autophagy and Selective Mitophagy. Autophagy 2014, 10, 1466–1467. [Google Scholar] [CrossRef] [Green Version]
  62. Juretic, N.; Hoen, D.R.; Huynh, M.L.; Harrison, P.M.; Bureau, T.E. The Evolutionary Fate of MULE-Mediated Duplications of Host Gene Fragments in Rice. Genome Res. 2005, 15, 1292–1297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Avin-Wittenberg, T.; Baluška, F.; Bozhkov, P.V.; Elander, P.H.; Fernie, A.R.; Galili, G.; Hassan, A.; Hofius, D.; Isono, E.; Le Bars, R.; et al. Autophagy-Related Approaches for Improving Nutrient Use Efficiency and Crop Yield Protection. J. Exp. Bot. 2018, 69, 1335–1353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Yin, R.; Liu, X.; Yu, J.; Ji, Y.; Liu, J.; Cheng, L.; Zhou, J. Up-Regulation of Autophagy by Low Concentration of Salicylic Acid Delays Methyl Jasmonate-Induced Leaf Senescence. Sci. Rep. 2020, 10, 11472. [Google Scholar] [CrossRef]
  65. Shibuya, K.; Niki, T.; Ichimura, K. Pollination Induces Autophagy in Petunia Petals via Ethylene. J. Exp. Bot. 2013, 64, 1111–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zhu, T.; Zou, L.; Li, Y.; Yao, X.; Xu, F.; Deng, X.; Zhang, D.; Lin, H. Mitochondrial alternative oxidase-dependent autophagy involved in ethylene-mediated drought tolerance in Solanum lycopersicum. Plant Biotechnol. J. 2018, 16, 2063–2076. [Google Scholar] [CrossRef] [Green Version]
  67. Liu, K.; Sutter, B.M.; Tu, B.P. Autophagy Sustains Glutamate and Aspartate Synthesis in Saccharomyces Cerevisiae during Nitrogen Starvation. Nat. Commun. 2021, 12, 57. [Google Scholar] [CrossRef] [PubMed]
  68. Estrada-Navarrete, G.; Cruz-Mireles, N.; Lascano, R.; Alvarado-Affantranger, X.; Hernández-Barrera, A.; Barraza, A.; Olivares, J.E.; Arthikala, M.-K.; Cárdenas, L.; Quinto, C.; et al. An Autophagy-Related Kinase Is Essential for the Symbiotic Relationship between Phaseolus Vulgaris and Both Rhizobia and Arbuscular Mycorrhizal Fungi. Plant Cell 2016, 28, 2326–2341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Yamada, Y.; Schaap, P. The Proppin Bcas3 and Its Interactor Kinky A Localize to the Early Phagophore and Regulate Autophagy. Autophagy 2021, 17, 640–655. [Google Scholar] [CrossRef] [Green Version]
  70. Lewis, G.P.; Schrire, B.; Mackinder, B.; Lock, M. Legumes of the World; Royal Botanic Gardens: Kew, UK, 2005. [Google Scholar]
  71. Choi, H.-K.; Mun, J.-H.; Kim, D.-J.; Zhu, H.; Baek, J.-M.; Mudge, J.; Roe, B.; Ellis, N.; Doyle, J.; Kiss, G.B.; et al. Estimating Genome Conservation between Crop and Model Legume Species. Proc. Natl. Acad. Sci. USA 2004, 101, 15289–15294. [Google Scholar] [CrossRef] [Green Version]
  72. Lee, C.; Yu, D.; Choi, H.-K.; Kim, R.W. Reconstruction of a Composite Comparative Map Composed of Ten Legume Genomes. Genes Genom. 2017, 39, 111–119. [Google Scholar] [CrossRef] [Green Version]
  73. Tang, H.; Bowers, J.E.; Wang, X.; Ming, R.; Alam, M.; Paterson, A.H. Synteny and Collinearity in Plant Genomes. Science 2008, 320, 486–488. [Google Scholar] [CrossRef] [Green Version]
  74. Stirnimann, C.U.; Petsalaki, E.; Russell, R.B.; Müller, C.W. WD40 Proteins Propel Cellular Networks. Trends Biochem. Sci. 2010, 35, 565–574. [Google Scholar] [CrossRef]
  75. Letunic, I.; Doerks, T.; Bork, P. SMART 7: Recent Updates to the Protein Domain Annotation Resource. Nucleic Acids Res. 2012, 40, D302–D305. [Google Scholar] [CrossRef]
  76. Xu, C.; Min, J. Structure and Function of WD40 Domain Proteins. Protein Cell 2011, 2, 202–214. [Google Scholar] [CrossRef]
  77. Lu, Q.; Yang, P.; Huang, X.; Hu, W.; Guo, B.; Wu, F.; Lin, L.; Kovács, A.L.; Yu, L.; Zhang, H. The WD40 Repeat PtdIns(3)P-Binding Protein EPG-6 Regulates Progression of Omegasomes to Autophagosomes. Dev. Cell 2011, 21, 343–357. [Google Scholar] [CrossRef] [Green Version]
  78. Tamura, N.; Oku, M.; Ito, M.; Noda, N.N.; Inagaki, F.; Sakai, Y. Atg18 Phosphoregulation Controls Organellar Dynamics by Modulating Its Phosphoinositide-Binding Activity. J. Cell Biol. 2013, 202, 685–698. [Google Scholar] [CrossRef] [Green Version]
  79. Zhang, B.; Shao, L.; Wang, J.; Zhang, Y.; Guo, X.; Peng, Y.; Cao, Y.; Lai, Z. Phosphorylation of ATG18a by BAK1 Suppresses Autophagy and Attenuates Plant Resistance against Necrotrophic Pathogens. Autophagy 2021, 17, 2093–2110. [Google Scholar] [CrossRef]
  80. Proikas-Cezanne, T.; Ruckerbauer, S.; Stierhof, Y.-D.; Berg, C.; Nordheim, A. Human WIPI-1 Puncta-Formation: A Novel Assay to Assess Mammalian Autophagy. FEBS Lett. 2007, 581, 3396–3404. [Google Scholar] [CrossRef] [Green Version]
  81. Gopaldass, N.; Fauvet, B.; Lashuel, H.; Roux, A.; Mayer, A. Membrane Scission Driven by the PROPPIN Atg18. EMBO J. 2017, 36, 3274–3291. [Google Scholar] [CrossRef]
  82. Polson, H.E.J.; de Lartigue, J.; Rigden, D.J.; Reedijk, M.; Urbé, S.; Clague, M.J.; Tooze, S.A. Mammalian Atg18 (WIPI2) Localizes to Omegasome-Anchored Phagophores and Positively Regulates LC3 Lipidation. Autophagy 2010, 6, 506–522. [Google Scholar] [CrossRef] [Green Version]
  83. Baskaran, S.; Ragusa, M.J.; Hurley, J.H. How Atg18 and the WIPIs Sense Phosphatidylinositol 3-Phosphate. Autophagy 2012, 8, 1851–1852. [Google Scholar] [CrossRef] [Green Version]
  84. Drozdetskiy, A.; Cole, C.; Procter, J.; Barton, G.J. JPred4: A Protein Secondary Structure Prediction Server. Nucleic Acids Res. 2015, 43, W389–W394. [Google Scholar] [CrossRef]
  85. 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] [Green Version]
  86. Feng, X.; Xu, Y.; Chen, Y.; Tang, Y.J. MicrobesFlux: A Web Platform for Drafting Metabolic Models from the KEGG Database. BMC Syst. Biol. 2012, 6, 94. [Google Scholar] [CrossRef] [Green Version]
  87. Bolser, D.M.; Staines, D.M.; Perry, E.; Kersey, P.J. Ensemble Plants: Integrating Tools for Visualizing, Mining, and Analyzing Plant Genomic Data. In Plant Genomics Databases: Methods and Protocols; Van Dijk, A.D.J., Ed.; Springer: New York, NY, USA, 2017; pp. 1–31. [Google Scholar] [CrossRef]
  88. Potter, S.C.; Luciani, A.; Eddy, S.R.; Park, Y.; Lopez, R.; Finn, R.D. HMMER Web Server: 2018 Update. Nucleic Acids Res. 2018, 46, W200–W204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Remm, M.; Storm, C.E.; Sonnhammer, E.L. Automatic Clustering of Orthologs and In-Paralogs from Pairwise Species Comparisons. J. Mol. Biol. 2001, 314, 1041–1052. [Google Scholar] [CrossRef] [Green Version]
  90. Huerta-Cepas, J.; Szklarczyk, D.; Heller, D.; Hernández-Plaza, A.; Forslund, S.K.; Cook, H.; Mende, D.R.; Letunic, I.; Rattei, T.; Jensen, L.J.; et al. EggNOG 5.0: A Hierarchical, Functionally and Phylogenetically Annotated Orthology Resource Based on 5090 Organisms and 2502 Viruses. Nucleic Acids Res. 2019, 47, D309–D314. [Google Scholar] [CrossRef] [Green Version]
  91. Sievers, F.; Higgins, D.G. Clustal Omega for Making Accurate Alignments of Many Protein Sequences. Protein Sci. 2018, 27, 135–145. [Google Scholar] [CrossRef] [Green Version]
  92. Subramanian, B.; Gao, S.; Lercher, M.J.; Hu, S.; Chen, W.-H. Evolview v3: A Webserver for Visualization, Annotation, and Management of Phylogenetic Trees. Nucleic Acids Res. 2019, 47, W270–W275. [Google Scholar] [CrossRef]
  93. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  94. Akaike, H. A New Look at the Statistical Model Identification. IEEE Trans. Autom. Control 1974, 19, 716–723. [Google Scholar] [CrossRef]
  95. Paradis, E.; Schliep, K. Ape 5.0: An Environment for Modern Phylogenetics and Evolutionary Analyses in R. Bioinformatics 2019, 35, 526–528. [Google Scholar] [CrossRef]
  96. Hu, Y.; Yan, C.; Hsu, C.-H.; Chen, Q.-R.; Niu, K.; Komatsoulis, G.A.; Meerzaman, D. OmicCircos: A Simple-to-Use R Package for the Circular Visualization of Multidimensional Omics Data. Cancer Inform. 2014, 13, CIN-S13495. [Google Scholar] [CrossRef]
  97. Cleary, A.; Farmer, A. Genome Context Viewer: Visual Exploration of Multiple Annotated Genomes Using Microsynteny. Bioinformatics 2018, 34, 1562–1564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Wang, J.; Hossain, M.S.; Lyu, Z.; Schmutz, J.; Stacey, G.; Xu, D.; Joshi, T. SoyCSN: Soybean Context-specific Network Analysis and Prediction Based on Tissue-specific Transcriptome Data. Plant Direct 2019, 3, e00167. [Google Scholar] [CrossRef] [Green Version]
  99. Nanjareddy, K.; Arthikala, M.-K.; Gómez, B.-M.; Blanco, L.; Lara, M. Differentially Expressed Genes in Mycorrhized and Nodulated Roots of Common Bean Are Associated with Defense, Cell Wall Architecture, N Metabolism, and P Metabolism. PLoS ONE 2017, 12, e0182328. [Google Scholar] [CrossRef] [Green Version]
  100. Bustin, S.A.; Benes, V.; Garson, J.A.; Hellemans, J.; Huggett, J.; Kubista, M.; Mueller, R.; Nolan, T.; Pfaffl, M.W.; Shipley, G.L. The MIQE guidelines: Minimum information for publication of quantitative real-time PCR experiments. Clin. Chem. 2009, 55, 611–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Arthikala, M.K.; Montiel, J.; Nava, N.; Santana, O.; Sánchez-López, R.; Cárdenas, L.; Quinto, S. PvRbohB negatively regulates Rhizophagus irregularis colonization in Phaseolus vulgaris. Plant Cell Physiol. 2013, 54, 1391–13402. [Google Scholar] [CrossRef] [Green Version]
  102. Vandesompele, J.; De Preter, K.; Pattyn, F.; Poppe, B.; Van Roy, N.; De Paepe, A.; Speleman, F. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal reference genes. Genome Biol. 2002, 3, 1–12. [Google Scholar] [CrossRef] [Green Version]
  103. Bailey, T.L.; Johnson, J.; Grant, C.E.; Noble, W.S. The MEME Suite. Nucleic Acids Res. 2015, 43, W39–W49. [Google Scholar] [CrossRef] [Green Version]
  104. Kim, D.E.; Chivian, D.; Baker, D. Protein Structure Prediction and Analysis Using the Robetta Server. Nucleic Acids Res. 2004, 32, W526–W531. [Google Scholar] [CrossRef] [Green Version]
  105. Söding, J. Protein Homology Detection by HMM-HMM Comparison. Bioinform. Oxf. Engl. 2005, 21, 951–960. [Google Scholar] [CrossRef] [Green Version]
  106. Yang, Y.; Faraggi, E.; Zhao, H.; Zhou, Y. Improving Protein Fold Recognition and Template-Based Modeling by Employing Probabilistic-Based Matching between Predicted One-Dimensional Structural Properties of Query and Corresponding Native Properties of Templates. Bioinformatics 2011, 27, 2076–2082. [Google Scholar] [CrossRef] [Green Version]
  107. Källberg, M.; Wang, H.; Wang, S.; Peng, J.; Wang, Z.; Lu, H.; Xu, J. Template-Based Protein Structure Modeling Using the RaptorX Web Server. Nat. Protoc. 2012, 7, 1511–1522. [Google Scholar] [CrossRef] [Green Version]
  108. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A Visualization System for Exploratory Research and Analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Phylogenetic analysis and protein motifs of 17 ATG families in A. thaliana, P. vulgaris, M. truncatula and G. max. The phylogenetic tree was constructed with the neighbor-joining method with 1000 repeated bootstrap tests, p-distance and pairwise deletion in MEGA X software and visualized using EvolView. MEME was used to identify motifs of the ATG homologs in A. thaliana, P. vulgaris, M. truncatula and G. max.
Figure 1. Phylogenetic analysis and protein motifs of 17 ATG families in A. thaliana, P. vulgaris, M. truncatula and G. max. The phylogenetic tree was constructed with the neighbor-joining method with 1000 repeated bootstrap tests, p-distance and pairwise deletion in MEGA X software and visualized using EvolView. MEME was used to identify motifs of the ATG homologs in A. thaliana, P. vulgaris, M. truncatula and G. max.
Plants 10 02619 g001
Figure 2. The chromosomal localization, synteny relationship and gene expression of autophagy genes were integrated into the Circos plot designed using OmicCircos. The outermost circle shows the A. thaliana (blue), P. vulgaris (green), M. truncatula (pink) and G. max (brown) chromosomes. The inner circle is a heatmap that shows the log2 RPKM values of gene expression in leaves and roots under ammonia, nitrate and urea treatments. The innermost line is the synteny of autophagy genes, but the yellow, purple and red lines represent ATG18b subfamilies I, II and III, respectively.
Figure 2. The chromosomal localization, synteny relationship and gene expression of autophagy genes were integrated into the Circos plot designed using OmicCircos. The outermost circle shows the A. thaliana (blue), P. vulgaris (green), M. truncatula (pink) and G. max (brown) chromosomes. The inner circle is a heatmap that shows the log2 RPKM values of gene expression in leaves and roots under ammonia, nitrate and urea treatments. The innermost line is the synteny of autophagy genes, but the yellow, purple and red lines represent ATG18b subfamilies I, II and III, respectively.
Plants 10 02619 g002
Figure 3. Ka/Ks ratios of 17 families of ATGs in A. thaliana, P. vulagris, G. max and M. truncatula. The distribution of Ka and Ks values are obtained using TBtools. The dark blue line divides the Ka/Ks ratios lower and higher than 1 (dots in the highlighted area Ka/Ks > 1).
Figure 3. Ka/Ks ratios of 17 families of ATGs in A. thaliana, P. vulagris, G. max and M. truncatula. The distribution of Ka and Ks values are obtained using TBtools. The dark blue line divides the Ka/Ks ratios lower and higher than 1 (dots in the highlighted area Ka/Ks > 1).
Plants 10 02619 g003
Figure 4. Transcription factor-binding sites in the promoter regions of ATGs (2000 bp) identified using PlantCARE.
Figure 4. Transcription factor-binding sites in the promoter regions of ATGs (2000 bp) identified using PlantCARE.
Plants 10 02619 g004
Figure 5. Expression profiles of ATGs in P. vulgaris tissues. (a) The transcription abundances of P. vulgaris ATGs in different tissues and organs during different stages of development and during rhizobial infections obtained from the PvGEA database. (b) Expression data from nodulated roots (R. tropici) and mycorrhized roots (R. irregularis) obtained from RNA-seq analysis. A violin plot shows total number of up/dowregulated ATGs under nodulated/mycorrhized conditions. The highlighted box represents higher number of downregulated genes in mycorrhized condition.
Figure 5. Expression profiles of ATGs in P. vulgaris tissues. (a) The transcription abundances of P. vulgaris ATGs in different tissues and organs during different stages of development and during rhizobial infections obtained from the PvGEA database. (b) Expression data from nodulated roots (R. tropici) and mycorrhized roots (R. irregularis) obtained from RNA-seq analysis. A violin plot shows total number of up/dowregulated ATGs under nodulated/mycorrhized conditions. The highlighted box represents higher number of downregulated genes in mycorrhized condition.
Plants 10 02619 g005
Figure 6. Three-dimensional representation of 280 ATG18 proteins from different plant species analyzed by multidimensional scaling using Bios2mds. The ATG18 subfamilies are color-coded as follows: Subfamily I (yellow), subfamily II (purple) and subfamily III (red). PC, principal component. The axes are the principal components (PC): x-axis (PC1), y-axis (PC2) and z-axis (PC3).
Figure 6. Three-dimensional representation of 280 ATG18 proteins from different plant species analyzed by multidimensional scaling using Bios2mds. The ATG18 subfamilies are color-coded as follows: Subfamily I (yellow), subfamily II (purple) and subfamily III (red). PC, principal component. The axes are the principal components (PC): x-axis (PC1), y-axis (PC2) and z-axis (PC3).
Plants 10 02619 g006
Figure 7. Phylogenetic tree of ATG18 proteins in plants. The protein sequences were aligned using Clustal Omega, and the phylogenetic tree was constructed using the ML method in MEGA X software with 1000 bootstrap replications. A total of 280 sequences of ATG18 are differentiated into subfamilies: Subfamily I (yellow), subfamily II (purple) and subfamily III (red). The plant species are differentiated by letters: A. thaliana (At), M. polymorpha (Mpo), O. sativa (Os), Triticum aestivum (Ta), Zea mays (Zm), Arachis duranensis (Ad), Arachis ipaensis (Ai), Cajanus cajan (Cc), Lotus japonicus (Lj), Cicer arietinum (Ca), Lupinus angustifolius (La), Pisum sativum (Ps), Vigna angularis (Va), Vigna radiata (Vr) and Trifolium pratense (Tp), P. patens, C. braunii (Cb), C. reinhardtii (Cr), D. salina (Ds), V. carteri (Vc), K. nitens (Kn), M. pusilla (Mpu), O. lucimarinus (Ol), O. tauri (Ot) and C. subellipsoidea (Cs). The branch lengths are labeled.
Figure 7. Phylogenetic tree of ATG18 proteins in plants. The protein sequences were aligned using Clustal Omega, and the phylogenetic tree was constructed using the ML method in MEGA X software with 1000 bootstrap replications. A total of 280 sequences of ATG18 are differentiated into subfamilies: Subfamily I (yellow), subfamily II (purple) and subfamily III (red). The plant species are differentiated by letters: A. thaliana (At), M. polymorpha (Mpo), O. sativa (Os), Triticum aestivum (Ta), Zea mays (Zm), Arachis duranensis (Ad), Arachis ipaensis (Ai), Cajanus cajan (Cc), Lotus japonicus (Lj), Cicer arietinum (Ca), Lupinus angustifolius (La), Pisum sativum (Ps), Vigna angularis (Va), Vigna radiata (Vr) and Trifolium pratense (Tp), P. patens, C. braunii (Cb), C. reinhardtii (Cr), D. salina (Ds), V. carteri (Vc), K. nitens (Kn), M. pusilla (Mpu), O. lucimarinus (Ol), O. tauri (Ot) and C. subellipsoidea (Cs). The branch lengths are labeled.
Plants 10 02619 g007
Figure 8. Protein motifs of the ATG18b family from different plant species. The conserved motifs were identified with MEME. The amino acid sequence of the ATG18 family is represented by lines, and the motifs identified using TBtools are shown with boxes: Motif 1 (green), motif 2 (yellow), motif 3 (dark green) and motif 4 (pink).
Figure 8. Protein motifs of the ATG18b family from different plant species. The conserved motifs were identified with MEME. The amino acid sequence of the ATG18 family is represented by lines, and the motifs identified using TBtools are shown with boxes: Motif 1 (green), motif 2 (yellow), motif 3 (dark green) and motif 4 (pink).
Plants 10 02619 g008aPlants 10 02619 g008b
Figure 9. Repeats, domains and families of ATG18b sub-families. (a) The ATG18 protein functions were determined using Pfam, and the proteins were divided into subfamilies: Subfamily I (yellow), subfamily II (purple) and subfamily III (red). (b) Pfam identifiers and their annotations.
Figure 9. Repeats, domains and families of ATG18b sub-families. (a) The ATG18 protein functions were determined using Pfam, and the proteins were divided into subfamilies: Subfamily I (yellow), subfamily II (purple) and subfamily III (red). (b) Pfam identifiers and their annotations.
Plants 10 02619 g009
Figure 10. Three-dimensional structural model of PvATG18b determined by molecular dynamics simulation and alignment of ATG18 protein sequences of P. vulgaris. (a) The PvATG18b protein structure preserves seven blades of four β-strands. (ad) In the colored rainbow, the N-terminus is shown in blue, the C-terminal is shown in red, the FRRG repeat (F218-G221) is colored pink, the conserved T131 residue is shown in orange and S246 Ser is presented in blue. The region consisting of site I PI(3)P and site II PI(3,5)P2 are shown in the gray circle. (b) PvATG18b protein structure rotated 180° and showing the CD loop (S269-T288) in yellow. (c) PvATG18b protein structure surfaces (positive and negative charges are shown in blue and red, respectively) showing a nonspecific electrostatic charge. (d) The FRRG repeat position is highlighted with the following colors: Subfamily I (yellow), subfamily II (purple) and subfamily III (red).
Figure 10. Three-dimensional structural model of PvATG18b determined by molecular dynamics simulation and alignment of ATG18 protein sequences of P. vulgaris. (a) The PvATG18b protein structure preserves seven blades of four β-strands. (ad) In the colored rainbow, the N-terminus is shown in blue, the C-terminal is shown in red, the FRRG repeat (F218-G221) is colored pink, the conserved T131 residue is shown in orange and S246 Ser is presented in blue. The region consisting of site I PI(3)P and site II PI(3,5)P2 are shown in the gray circle. (b) PvATG18b protein structure rotated 180° and showing the CD loop (S269-T288) in yellow. (c) PvATG18b protein structure surfaces (positive and negative charges are shown in blue and red, respectively) showing a nonspecific electrostatic charge. (d) The FRRG repeat position is highlighted with the following colors: Subfamily I (yellow), subfamily II (purple) and subfamily III (red).
Plants 10 02619 g010
Figure 11. ATG18 structure. (a) Three-dimensional structural model of Atg18b before (gray) and after (purple) running the molecular dynamics simulation. (b) RMSD of the modeled ATG18b protein over a time period of 2.1 µs.
Figure 11. ATG18 structure. (a) Three-dimensional structural model of Atg18b before (gray) and after (purple) running the molecular dynamics simulation. (b) RMSD of the modeled ATG18b protein over a time period of 2.1 µs.
Plants 10 02619 g011
Table 1. List of 17 autophagy gene families in A. thaliana, P. vulgaris, M. truncatula and G. max.
Table 1. List of 17 autophagy gene families in A. thaliana, P. vulgaris, M. truncatula and G. max.
Arabidopsis thalianaPhaseolus vulgarisMedicago truncatulaGlycine max
ComplexFamilyNameIDNameIDNameIDNameID
Initiation of autophagyATG1 complexATG1AtATG1aAt3g61960 MtATG1aMedtr8g024100GmATG1a.IGlyma.07g048400
GmATG1a.IIGlyma.16g017300
AtATG1bAt3g53930PvATG1bPhvul.010g015100MtATG1bMedtr4g019410GmATG1b.IGlyma.03g069800
AtATG1cAt2g37840 GmATG1b.IIGlyma.01g099600
AtATG1tAt1g49180PvATG1tPhvul.010g120500MtATG1tMedtr3g095620GmATG1t.IGlyma.06g150700
GmATG1t.IIGlyma.04g215500
ATG11AtATG11At4g30790PvATG11Phvul.003g153800MtATG11Medtr4g130370GmATG11Glyma.17g071400
ATG13AtATG13At3g49590PvATG13aPhvul.008g187800MtATG13aMedtr5g068710GmATG13a.IGlyma.02g220700
GmATG13a.IIGlyma.14g187000
AtATG13bAt3g18770PvATG13bPhvul.002g269600MtATG13bMedtr3g095570GmATG13b.IGlyma.05g189000
MtATG13cMedtr8g093050GmATG13b.IIGlyma.08g146700
ATG101AtATG101At5g66930PvATG101Phvul.003g248000MtATG101Medtr8g079240GmATG101Glyma.17g180900
Membrane recruitment to autophagosomezComplex ATG2-ATG18ATG9AtATG9At2g31260PvATG9aPhvul.001g159900MtATG9aMedtr7g096680GmATG9a.IGlyma.03g162100
GmATG9a.IIGlyma.19g163500
PvATG9bPhvul.007g194300MtATG9bMedtr1g070160GmATG9b.IIIGlyma.10g035800
GmATG9b.vIGlyma.13g122200
ATG2AtATG2At3g19190PvATG2Phvul.003g295800MtATG2Medtr4g086370GmATG2.IGlyma.02g133400
GmATG2.IIGlyma.07g211600
ATG18AtATG18aAt3g62770PvATG18aPhvul.001g205000MtATG18aMedtr1g083230GmATG18a.IGlyma.10g152500
GmATG18a.IIGlyma.20g235800
GmATG18a.IIIGlyma.03g212100
GmATG18a.IvGlyma.19g209200
AtATG18bAt4g30510PvATG18bPhvul.003g152800MtATG18bMedtr4g130190GmATG18b.IGlyma.17g070200
GmATG18b.IIGlyma.02g207500
GmATG18b.IIIGlyma.10g126200
AtATG18cAt2g40810PvATG18c.IPhvul.009g041700MtATG18cMedtr7g108520GmATG18c.IGlyma.04g224300
PvATG18c.IIPhvul.007g196400 GmATG18c.IIGlyma.06g140400
AtATG18dAt3g56440 MtATG18dMedtr1g088855
AtATG18eAt5g05150 MtATG18eMedtr3g093590GmATG18eGlyma.16g109400
AtATG18fAt5g54730PvATG18f.IPhvul.011g140900MtATG18fMedtr2g082770GmATG18f.IGlyma.12g214600
PvATG18f.IIPhvul.005g091300 GmATG18f.IIGlyma.12g136000
GmATG18f.IIIGlyma.13g287000
GmATG18f.IVGlyma.06g267000
AtATG18gAt1g03380PvATG18g.IPhvul.001g146700MtATG18gMedtr1g089110GmATG18g.IGlyma.03g148700
PvATG18g.IIPhvul.007g183100 GmATG18g.IIGlyma.19g152000
GmATG18g.IIIGlyma.20g230900
AtATG18hAt1g54710 MtATG18hMedtr1g082300GmATG18hGlyma.10g157700
Autophagosome formation ATG6AtATG6At3g61710PvATG6Phvul.005g029900MtATG6Medtr3g018770GmATG6.IGlyma.11g153900
GmATG6.IIGlyma.04g141000
PI3K complexATG14AtATG14aAt1g77890PvATG14Phvul.008g169200MtATG14Medtr5g061040GmATG14.IGlyma.13g085400
GmATG14.IIGlyma.14g167200
AtATG14bAt4g08540
Ubiquitin-like protein conjugation systemsUbiquitin-like conjugation (ATG8)ATG3AtATG3At5g61500PvATG3Phvul.011g006500MtATG3Medtr4g036265GmATG3.IGlyma.12g005700
GmATG3.IIGlyma.09g231000
AtATG4aAt2g44140PvATG4aPhvul.008g048900MtATG4aMedtr7g081230GmATG4a.IGlyma.18g248400
ATG4 GmATG4a.IIGlyma.09g244800
AtATG4bAt3g59950
ATG7AtATG7At5g45900PvATG7Phvul.011g010700MtATG7Medtr0003s0540GmATG7Glyma.12g010000
ATG8AtATG8aAt4g21980 MtATG8aMedtr2g023430
AtATG8bAt4g04620 MtATG8bMedtr4g037225GmATG8bGlyma.15g188600
AtATG8cAt1g62040PvATG8c.IPhvul.003g079300MtATG8cMedtr4g048510GmATG8c.IGlyma.12g098400
PvATG8c.IIPhvul.006g149640 GmATG8c.IIGlyma.06g306300
GmATG8c.IIIGlyma.09g003900
GmATG8c.IVGlyma.17g013000
GmATG8c.VGlyma.07g261000
GmATG8c.VIGlyma.15g108200
AtATG8dAt2g05630PvATG8dPhvul.011g103300MtATG8dMedtr2g088230
AtATG8eAt2g45170 MtATG8eMedtr4g101090
AtATG8fAt4g16520PvATG8f.IPhvul.003g219600MtATG8fMedtr1g086310GmATG8fGlyma.17g140700
PvATG8f.IIPhvul.002g062200
AtATG8gAt3g60640 MtATG8gMedtr4g123760
AtATG8hAt3g06420 MtATG8hMedtr7g096540
AtATG8iAt3g15580PvATG8iPhvul.007g210800 GmATG8iGlyma.02g008800
ATG5AtATG5At5g17290PvATG5Phvul.008g241000MtATG5Medtr5g076920GmATG5.IGlyma.14g210200
GmATG5.IIGlyma.02g240700
Ubiquitin-like conjugation (ATG12)ATG10AtATG10At3g07525PvATG10Phvul.010g036300MtATG10Medtr8g010140GmATG10Glyma.03g097000
ATG12AtATG12aAt1g54210
AtATG12bAt3g13970PvATG12bPhvul.010g130300MtATG12bMedtr8g020500GmATG12b.IGlyma.07g038100
GmATG12b.IIGlyma.16g007300
ATG16AtATG16At5g50230PvATG16Phvul.003g207100MtATG16aMedtr3g075400GmATG16.IGlyma.05g043700
MtATG16bMedtr4g104380GmATG16.IIGlyma.17g126200
MtATG16cMedtr4g007500
Table 2. List of ATG18 homologs in early plant lineages.
Table 2. List of ATG18 homologs in early plant lineages.
ChlorophytaCharophytaLiverwortsBryophyta Monocots Arabidopsis
Dunaliella salinaVolvox carteriOstreococcus tauriOOstreococcus lucimarinusMicromonas pusillaCoccomyxa subellipsoideaChlamydomonas reinhardtiiChara brauniiKlebsormidium nitensMarchantia polymorphaPhyscomitrella patensOryza sativaZea maysTriticum aestivumArabidopsis thaliana
Subfamily IADsATG18 (Dusal.0227s00002.1)VcATG18 (Vocar.0005s0363)OtATG18 (Ot06g00830)OlATG18 (OlATG18.3284.
fragment)
MpuATG18 (MpuATG1849616)CsubATG18 (CsATG18.65175)CrATG18 (Cre10.g425750.t1)CbATG18 (CHBRA95g00960)KnATG18 (kfl00229.0060)MpoATG18a.I (MARPO.0005s0065)PpATG18 (Phpat.005G022700)OsATG18a (XP.015621196)ZmATG18a (Zm00001d011920)TaATG18a.I (CDM86058)AtATG18a (AT3G62770)
MpoATG18a.II (MARPO.0001s0033)PpATG18 (Phpat.006G095100) TaATG18a.II (AGW81806)
PpATG18 (Phpat.017G015900) TaATG18a.III (Traes.3B.19AF6BFF0)
TaATG18a.IV (TRAES.3B.113DC4275)
ZmATG18b.IV (Zm00001d042215.T002)
ZmATG18b.V (GRMZM2G143211)
C ZmATG18c.I (AQK90439)TaATG18c.I (Traes.3DS.985ED34D7)AtATG18c (AT2G40810)
ZmATG18c.II (Zm00001d008691)TaATG18c.II (Traes.3AS.71D103050)
ZmATG18c.III (GRMZM2G069177)TaATG18c.III (TraesCS3B02G110900)
ZmATG18c.IV (AQK90440)TaATG18c.IV (CDM81498)
D OsATG18d.I (XP.015620970) TaATG18d (AGW81809)AtATG18d (AT3G56440)
E OsATG18eII (XP.015639564) AtATG18e (AT5G05150)
Subfamily IIBDsATG18 (Dusal.0460s00003)VcATG18 (Vocar.0020s0155)OtATG18 (Ot06g00720)OlATG18 (OlATG18.41442.fragment)MpuATg18 (MpuATG18.156491.
fragment)
CsubATG18 (CsATG18.3880.fragment)CrATG18
(Cre10.g457550)
KnATG18 (kfl00404.0130)MpoATG18b (MARPO.0027s0044)PpATG18 (Phpat.007G038400)OsATG18b (XP.015627655)ZmATG18b.I (NP_00114563.1)TaATG18b (Traes.6AL.DDF2EBF31)AtATG18b (AT4G30510)
ZmATG18b.II (XP.020408852)TaATG18e.I (Traes_6BL_B2A8BBB52)
ZmATG18b.III (Zm00001d018355)TaATG18e.II (Traes.6DL.9F29527A0)
Subfamily IIIF CsubATG18 (CsATG18.63899) CbATG18 (CHBRA141g00400)KnATG18 (kfl00046.0070)MpoATG18f (MARPO.0006s0048)PpATG18 (Phpat.008G022700)OsATG18f.I (XP.015621123)ZmATG18f.I (ONM37261)TaATG18f.I (Traes.3B.F4F2FC6FA)AtATG18f (AT5G54730)
PpATG18 (Phpat.020G070000)OsATG18f.II (XP.025877429)ZmATG18f.II (ONM37262)TaATG18f.II (Traes.3DL.E400E521A)
PpATG18 (Phpat.023G024100)OsATG18f.III (LOC.Os05g33610)ZmATG18f.III (Zm00001d043239)TaATG18f.III (TraesCS3D02G318200)
PpATG18 (Phpat.024G018700) ZmATG18f.IV (ONM37265)TaATG18f.IV (CDM84501)
ZmATG18f.V (PWZ31673)TaATG18f.V (Traes.3B.7A23DFB41)
TaATG18f.VI (Traes.3AL.B27F0D4FF)
G ZmATG18g.I (AQK85845) AtATG18g (AT1G03380)
ZmATG18g.II (AQK85860)
ZmATG18g.III (AQK93836)
ZmATG18g.IV (AQK93828)
ZmATG18g.V (AQK93834)
ZmATG18g.VI (AQK85849)
ZmATG18g.VII (GRMZM2G078468)
ZmATG18g.VIII (PWZ17532)
ZmATG18g.IX (AQK93830)
ZmATG18g.X (AQK93829)
ZmATG18g.XI (AQK93835)
ZmATG18g.XII (AQK85856)
ZmATG18g.XIII (AQK93827)
H ZmATG18h.I (XP.008649626)TaATG18h.I (Traes.1BL.45E2558BB.1)AtATG18h (AT1G54710)
OsATG18h (XP.015639663)ZmATG18h.II (PWZ11786)TaATG18h.II (TraesCS1A02G254200.1)
ZmATG18h.III (XP.008656294)TaATG18h.III (Traes.1DL.DB75BFD8A.1)
TaATG18h.IV (Traes.1AL.C4A651390.1)
Table 3. List of ATG18 homologs in legumes.
Table 3. List of ATG18 homologs in legumes.
GenestoidsDalbergioidsMilletioidsRobinioidsIRLC
Lupinus angustifoliusArachis duranensisArachis ipaensisGlycine maxVigna angularisVigna radiataPhaseolus vulgarisLotus JaponicaCicer arietinumCajanus
cajan
Medicago truncatulaPisum sativumTrifolium pratense
Subfamily IALaATG18a.I (XP.019421581.1) AdATG18a.I (XP.015939789.1)AiATG18a (XP.016174738.1)GmATG18a.I
(Glyma.10G152500.1)
VaATG18a.I (VIGAN03G286700)VrATG18a.I (VRADI08G12430)PvATG18a (Phvul.001G205000.1) CaATG18a.I (XP.004495714.1)CcATG18a.I (XP.020209984.1)MtATG18a (Medtr1G083230.1)PsATG18a (PSAT0S3233G0120.1)TpATG18a.I (TRIPR.GENE96259)
LaATG18a.II (XP.019452261.1)AdATG18a.II (XP.015967701.1) GmATG18a.II
(Glyma.20G235800.1)
VaATG18a.II (XP.017412432.1)VrATG18a.II
(VRADI03G05850)
CaATG18a.II (XP.004494924.1)CcATG18a.II (C.CAJAN.10296.1) TpATG18a.II (TRIPR.GENE33973)
LaATG18a.III (XP.019419463.1) GmATG18a.III (Glyma.03G212100.1)VaATG18a.III (VANG04G16030.1) CaATG18a.III (C.CA.05407.1)CcATG18a.III (XP.020212010.1) TpATG18a.III (PNX79795.1)
LaATG18a.IV (XP.019441771.1) GmATG18a.IV
(Glyma.19G209200.1)
VaATG18a.IV (VANG06G12920.1)
LaATG18a.V (XP.019441170.1)
LaATG18b.III (TanjilG.02747) LjATG18b.II (Lj5g3v1496760.1)CaATG18b.V (Ca.04089)
LjATG18b.III (Lj0g3v0083309.1)CaATG18b.VI (CC4958C.Ca14068.1)
LjATG18b.IV (Lj1g3v4912170.1)
CLaATG18c.I (XP.019430950.1)AdATG18c
(XP.015945005.1)
AiATG18c (XP.016181861.1)GmATG18c.I
(Glyma.04G224300.1)
PvATG18c.I (Phvul.009G041700.1) CaATG18c (C.CA.03673) PsATG18c (PSAT5G069920.1)TpATG18c.I (TRIPR.GENE13965)
LaATG18c.II (XP.019417508.1) GmATG18c.II
(Glyma.06G140400.1)
PvATG18c.II (Phvul.007G196400.1) MtATG18c (Medtr7G108520.1) TpATG18c.II (PNX92525.1)
LaATG18c.III (LUP000470) LjATG18c (Lj1G3V1112870.1)
DLaATG18d (XP.019430946.1) VaATG18d.I (VIGAN04G120000)VrATG18d.I (VRADI0239S00050) CaATG18d (XP.004502800.1)CcATG18d.I (XP.029129536.1)MtATG18d (Medtr1G088855.1)
VaATG18d.II (VANG0200S00330.1)VrATG18d.II (XP.022632145.1) CcATG18d.II (XP.020229011.1)
E VrATG18d.VI (XP.022632144.1) MtATG18e (Medtr3G093590.1)
Subfamily II LaATG18b.I (XP.019441874.1)AdATG18b
(XP.015933286.1)
AiATG18b (XP.016200540.1)GmATG18b.I
(Glyma.17G070200.1)
VaATG18b.I (VIGAN01G240600)VrATG18b.I (VRADI07G21660)PvATG18b (Phvul.003G152800.1)LjATG18b.I (Lj4G3V2018270.1)CaATG18b.I (XP.027192941.1) MtATG18b (Medtr4G130190.1)PsATG18b (PSAT0S2826G0080.1)TpATG18b.I (PNX94509)
B*LaATG18b.II (XP.019441865.1 GmATG18b.II
(Glyma.02G207500.2)
VaATG18b.II (XP.017411081.1)VrATG18b.II (XP.014510099.1) CaATG18b.II (XP.004507771.1) TpATG18b.II (PNY02700.1)
GmATG18b.III
(Glyma.10G126200.1)
VaATG18b.III (XP.017411091.1) CaATG18b.III (XP.027192940.1)
VaATG18b.IV (VANG11G12160.2) CaATG18b.IV (ICC4958.CA.21790.1)
VaATG18b.V (XP.017411074.1)
GmATG18e
(Glyma.16G109400.1)
* VrATG18d.III (XP.014522590.1)
LaATG18f.I (XP.019437124.1)AdATG18f.I (ARADU.XJ3JE.1)AiATG18f.I (XP.016170472.1)GmATG18f.I
(Glyma.12G214600.1)
VaATG18f.I (VIGAN05G145500)VrATG18f.I (XP.014522059.1)PvATG18f.I (Phvul.011G140900.1)LjATG18f (Lj3G3V1544540.1)CaATG18f.I (XP.004487613.1)CcATG18f.I (XP.020229318.1)MtATG18f
(Medtr2G082770.1)
PsATG18f (PSAT5G249880.1)TpATG18f (TRIPR.GENE36798)
Subfamily IIIFLaATG18f.II (XP.019453655.1)AdATG18f.II (XP.015936500.1)AiATG18f.II (ARAIP.FRI7H.1)GmATG18f.II
(Glyma.12G136000.1)
VaATG18f.II (XP.017425518.1)VrATG18f.II (XP.014494161.1)PvATG18f.II (Phvul.005G091300.1) CaATG18f.II (XP.027187641.1)CcATG18f.II (XP.020229320.1)
LaATG18f.III (OIW15456.1) GmATG18f.III
(Glyma.13G287000.1)
VaATG18f.III (VIGAN08G077000)VrATG18f.III (XP.022634400.1) CaATG18f.III (XP.004487612.1)CcATG18f.III (C.CAJAN32508.1)
LaATG18f.IV (XP.019453653.1) GmATG18f.IV.
(Glyma.06G267000.1)
VaATG18f.IV (VANG1095S00020.1)VrATG18f.IV (VRADI02G09460.1) CaATG18f.IV (CA.00864.1)CcATG18f.IV (XP.020235274.1)
CcATG18f.V (XP.020229319.1)
CcATG18f.VI (XP.020229316.1)
LaATG18g.I (XP.019441802.1)AdATG18g (XP.015951046.1)AiATG18g (XP.016184366.1)GmATG18g.I
(Glyma.03G148700.1)
VaATG18g.I (XP.017419622.1)VrATG18g (VRADI03G00450)PvATG18g.I (Phvul.001G146700.1)LjATG18g (Lj1G3V4404380.1)CaATG18g.I (CA.09934.1)CcATG18g.I (XP.020211839.1)MtATG18g.I (Medtr1G089110.1)PsATG18g (PSAT6G169560.1)TpATG18g.I (TRIPR.GENE16922)
GLaATG18g.II (XP.019441803.1) GmATG18g.II
(Glyma.19G152000.1)
VaATG18g.II
(KOM38883.1)
PvATG18g.II (Phvul.007G183100.1) CaATG18g.III (CA.08309)CcATG18g.II (C.CAJAN09614.1) TpATG18g.II (TRIPR.GENE2713)
GmATG18g.III
(Glyma.20G230900.1)
VaATG18g.III (VIGAN.VANG07G05180.1) CcATG18g.III (KYP70659.1)
LaATG18h.I (XP.019421306.1)AdATG18h.I (XP.015939933.1)AiATG18h.I (XP.016205481.1)GmATG18h
(Glyma.10G157700.1)
VaATG18h.I
(KOM55039.1)
VrATG18h (VRADI08G12840.1) LjATG18h (Lj5G3V1451080.1)CaATG18h.I (XP.027189075.1)CcATG18h.I (XP.020233978.1)MtATG18h (Medtr1G082300.1)PsATG18h (PSAT6G148560.1)TpATG18h.I (PNY09258.1)
HLaATG18h.II (XP.019421307.1)AdATG18h.II (XP.015939934.1)AiATG18h.II (XP.016176031.1) VaATG18h.II (VANG06G10190.1) CaATG18h.II (XP.027189076.1)CcATG18h.II (C.CAJAN06885.1) TpATG18h.II (PNY17060.1)
LaATG18h.III (XP.019421305.1)AdATG18h.III (XP.015968551.1)AiATG18h.III (XP.016176030.1) CaATG18h.III (CA.09238.1)CcATG18h.III (XP.020233954.1) TpATG18h.III (PNY12850.1)
LaATG18h.IV (XP.019452235.1) AiATG18h.IV (XP.016176032.1) CcATG18h.IV (XP.029125824.1)
LaATG18h.V (TANJILG.10103)
LaATG18h.VI (OIW07130.1)
LaATG18h.VII (XP.019452236.1)
LaATG18h.VIII (XP.019452234.1)
LaATG18h.IX (OIW12695.1)
* Sequence ID with assigned the letter but belongs to other ATG18 Subfamily.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Quezada-Rodríguez, E.-H.; Gómez-Velasco, H.; Arthikala, M.-K.; Lara, M.; Hernández-López, A.; Nanjareddy, K. Exploration of Autophagy Families in Legumes and Dissection of the ATG18 Family with a Special Focus on Phaseolus vulgaris. Plants 2021, 10, 2619. https://doi.org/10.3390/plants10122619

AMA Style

Quezada-Rodríguez E-H, Gómez-Velasco H, Arthikala M-K, Lara M, Hernández-López A, Nanjareddy K. Exploration of Autophagy Families in Legumes and Dissection of the ATG18 Family with a Special Focus on Phaseolus vulgaris. Plants. 2021; 10(12):2619. https://doi.org/10.3390/plants10122619

Chicago/Turabian Style

Quezada-Rodríguez, Elsa-Herminia, Homero Gómez-Velasco, Manoj-Kumar Arthikala, Miguel Lara, Antonio Hernández-López, and Kalpana Nanjareddy. 2021. "Exploration of Autophagy Families in Legumes and Dissection of the ATG18 Family with a Special Focus on Phaseolus vulgaris" Plants 10, no. 12: 2619. https://doi.org/10.3390/plants10122619

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