Next Article in Journal
Female Germ Cell Development in Chickens and Humans: The Chicken Oocyte Enriched Genes Convergent and Divergent with the Human Oocyte
Next Article in Special Issue
Aux/IAA11 Is Required for UV-AB Tolerance and Auxin Sensing in Arabidopsis thaliana
Previous Article in Journal
Comparison of Homologous and Heterologous Booster SARS-CoV-2 Vaccination in Autoimmune Rheumatic and Musculoskeletal Patients
Previous Article in Special Issue
Autophagy-Mediated Regulation of Different Meristems in Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Autophagy in the Lifetime of Plants: From Seed to Seed

Lushan Botanical Garden, Chinese Academy of Sciences, Jiujiang 332900, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(19), 11410; https://doi.org/10.3390/ijms231911410
Submission received: 3 September 2022 / Revised: 22 September 2022 / Accepted: 23 September 2022 / Published: 27 September 2022
(This article belongs to the Special Issue New Insight into Signaling and Autophagy in Plants 2.0)

Abstract

:
Autophagy is a highly conserved self-degradation mechanism in eukaryotes. Excess or harmful intracellular content can be encapsulated by double-membrane autophagic vacuoles and transferred to vacuoles for degradation in plants. Current research shows three types of autophagy in plants, with macroautophagy being the most important autophagic degradation pathway. Until now, more than 40 autophagy-related (ATG) proteins have been identified in plants that are involved in macroautophagy, and these proteins play an important role in plant growth regulation and stress responses. In this review, we mainly introduce the research progress of autophagy in plant vegetative growth (roots and leaves), reproductive growth (pollen), and resistance to biotic (viruses, bacteria, and fungi) and abiotic stresses (nutrients, drought, salt, cold, and heat stress), and we discuss the application direction of plant autophagy in the future.

1. Introduction

Autophagy, also known as self-eating, is an evolutionarily conserved process that occurs in eukaryotic cells and involves the degradation of organelles, protein complexes, and macromolecules [1]. Generally, the degraded material is sequestered into autophagic vesicles that are transported to the vacuole for breakdown. Under normal conditions, autophagy is a housekeeping process that degrades unwanted cytoplasmic content and maintains cellular homeostasis [2]. Under stress conditions (starvation, oxidative and abiotic stress, and pathogen infection), autophagy proteins are up-regulated and help in recycling damaged or non-essential cellular material [3].
Until now, three different types of autophagy in plants have been discovered, including microautophagy, macroautophagy, and mega-autophagy [4]. Microautophagy is the direct packaging of cargo into the vacuole for degradation through the invagination or protrusion of the vacuolar membrane [5]. Although the concept of microautophagy has been present for many years, little is known about the mechanism by which it occurs. In plants, microautophagy is found to play an important role in anthocyanin aggregates [6]. In addition, microautophagy is also involved in the degradation progress of damaged chloroplast, namely, chlorophagy [7]. The most well-studied type of autophagy in plants is macroautophagy, in which autophagosomes form and then fuse with vacuoles to degrade cargoes [8]. Until now, more than 40 ATG proteins have been found to be involved in the biological process of macroautophagy [1]. There is also a more direct type of autophagy, namely, mega-autophagy. Here, the tonoplast membrane ruptures to release the vacuolar hydrolases directly into the cytoplasm, where it degrades cytoplasmic materials [9,10]. Mega-autophagy often occurs during programmed cell death (PCD) including plant development or in response to pathogens [1].
In plants, autophagy is a feature in diverse biological processes such as development, nutrient recycling, and biotic and abiotic stresses [2] (Figure 1). It can be seen that autophagy plays a pivotal role in the life of plants. In this review, we summarize the research progress regarding how autophagy affects plant lifespan, especially concerning growth and development, resistance to abiotic stress, and interaction with microorganisms.

2. Autophagy in Vegetative Growth

2.1. Seed Development

For flowering plants, the cycle of life begins with a seed. Generally, atg mutations in plants produce fewer seeds compared to WT plants, suggesting that autophagy may function during plant seed development [11]. In Arabidopsis, several atg mutations show decreased seed production [2], and some ATG genes are up-regulated during seed maturation [12]. Similar results are shown in maize, such as ATG1a, Atg18e, Atg18e, Atg18f, and Atg18h, which are expressed in endosperm instead of other tissues [13]. However, the role of autophagy in seed development has not been explained at the mechanism level. Some studies show that autophagy may contribute to the transport of seed storage proteins [14,15]. Both total protein and 12S globulins are accumulated in atg5 seeds, indicating that autophagy affects the seed protein content [15]. In addition to seed protein accumulation, autophagy also contributes to seed germination. In Arabidopsis, the overexpression of Atg8-interacting proteins (ATI1 and ATI2) can stimulate seed germination under ABA conditions [16].
Some research shows that autophagy is also involved in plant oil production. In Brassicaceae, oil accumulates in embryonic tissues and endosperm during seed development and seed germination [17]. In addition, lipid droplets (LDs) need to be degraded under specific metabolic or physiological conditions, and this process occurs mainly via autophagy [18]. In Arabidopsis, the overexpression of ATG5 or ATG7 increases both the seed yield and the fatty acid (FA) content [11]. This process may not directly affect oil metabolism in seeds, but does affect the re-mobilization of resources from leaves [19].

2.2. Root Development

Structurally, plant roots are divided into three zones: the meristematic zone, elongation zone, and maturation zone [20]. A cross-section of roots includes three levels: dermal, cortex (ground tissue), and vascular tissues [21]. During root development, autophagy plays an important role in its establishment and functional differentiation.
ATG8 genes are mainly located in the root caps and maturation zone, which correspond to relevant protein degradation [22]. The role of autophagy in the root tips may be related to programmed cell death (PCD), but there is not enough evidence to prove this speculation [21].
Some evidence suggests that autophagy is involved in the formation of cortical tissue. Cortical parenchyma cells contain a large vacuole. In Arabidopsis roots, autophagy is required for ground tissue differentiation, in order to degrade cytoplasmic material and for vacuole formation from the meristem to the elongation zone [23]. In ground tissue cells, partial cytoplasmic accumulation is observed in central vacuoles, suggesting that autophagy occurs in these regions [23]. In Arabidopsis roots, the ATG8f protein localizes to autophagy-like structures in the central vacuole, suggesting that autophagy also determines vacuolar generation in cortical parenchymal cells [22].
Autophagy also plays a significant role in the differentiation of vascular tissue (xylem and phloem). In roots, xylem formation occurs primarily via nitric oxide (NO) signaling, through the synthesis of secondary cell walls and degradation of protoplast [24]. Autophagy-related processes appear to play a role in central vacuole formation and the degradation of cytoplasmic material at the onset of xylem partialization [21]. Several ATG genes (ATG8C, ATG8D, ATG11, and ATG18) are certified in the differentiation of the xylem [25]. Compared to the xylogenesis, there is a lack of discussion in the literature reporting on the mechanism of autophagy during phloemogenesis. There is a report indicating that ATG8 protein localization can be observed in differentiating the primary phloem [25].
In addition to root development, autophagy also plays a vital role in root senescence [26]. The up-regulation of ATG genes (ATG8C, ATG8D, and ATG8G) is characterized by the senescence of absorptive roots [26]. During the first stage of senescence, autophagy counteracts transient cell death and maintains cellular homeostasis [21]. Additionally, autophagy is also involved in the remobilization process, which is a key step in the senescence process [21].

2.3. Leaf Senescence

Leaf senescence is a late stage of plant vegetative growth. In Arabidopsis, premature leaf senescence is one of the common phenotypes in autophagy mutants. Most ATG genes are up-regulated under leaf senescence, while other autophagy genes are mainly expressed in leaf development. In barley, the transcript levels of both ATG7 and ATG18f are up-regulated during leaf senescence [27]. In rice, compared to the wild-type, the atg7 mutant not only reduces the plant height, root length, tiller number, and leaf area, but also has obvious premature leaf senescence [28]. In Arabidopsis, SAG12 (senescence-associated gene 12) is a marker gene for the onset of senescence, which is abundantly induced in atg2 and atg5 mutants [29]. Premature leaf senescence of these mutants can be alleviated by blocking SA (salicylic acid) biosynthesis or signal transduction. For example, the overexpression of SA hydroxylase NahG (salicylate hydroxylase) can inhibit the premature senescence phenotype of atg2 and atg5, and the administration of the SA analog BTH (benzothiadiazole) restores the normal phenotype of these mutants [29]. The regulation of autophagy in premature plant senescence may be attributed to the effect on the redistribution of plant nutrients, especially nitrogen. For example, rice atg7–1 mutant leaves prematurely senesce, and the nitrogen content of senescent leaves is higher than that of wild-type leaves, which reduces the nitrogen reuse efficiency [28]. In apples, the overexpression of ATG18a greatly improves resistance to low nitrogen stress and up-regulates the expression of nitrogen uptake and assimilation-related genes NIA2, NRT2.1, NRT2.4, and NRT2.5 [30]. Some new evidence suggests that SAG12 regulates plant senescence through involvement in protein degradation and N remobilization [31,32]. In future studies, it will be interesting to determine whether autophagy cooperates with senescence-associated proteases to cycle cellular components.

3. Autophagy in Reproductive Growth

Autophagy is not only involved in regulating the vegetative growth, but also regulates the reproductive growth in plants. Members of the PI3K complex (atg6, vps15, and vps34) are reported to regulate Arabidopsis pollen maturation, and none of them can produce mature pollen after mutation [33,34,35]. In the rice atg7 mutant, lipid and starch components in pollen grains cannot be accumulated normally during flowering, resulting in reduced pollen viability and sporophytic male sterility [36]. Autophagy also plays a key role in tobacco pollen germination. Autophagy flux is significantly increased in the early stage of pollen germination to degrade the cytoplasm in the germinal pores [37]. Cytoplasmic degradation of germinal pores during pollen germination is also inhibited after the silencing of ATG2, ATG5, and ATG7 in tobacco [38]. New research shows that autophagy is also involved in pollen tube elongation. In this process, a core protein SH3-domain-containing protein 2 (SH3P2) colocalizes with ATG proteins and participates in regulating mitophagosomes [38]. Down-regulation of SH3P2 expression significantly impairs pollen germination and pollen tube growth [38].

4. Autophagy in Abiotic Stress

Plants are exposed to various abiotic stresses during growth, such as salt, heat, cold, drought, and nutrition stress. Autophagy, a process that maintains cellular homeostasis, plays an important role in the defense of abiotic stresses.

4.1. Autophagy under Nutrient Starvation

Nutrient starvation triggers a strong induction of autophagy, and ATG mutants exhibit premature senescence upon carbon/nitrogen starvation. Using sucrose starvation in suspension-cultured cells shows that 30% to 50% of the total protein is degraded over a two day period [39], and the decrease in total proteins stems from non-selective degradation, rather than degradation of specific proteins [40]. Fusion between autophagosomes and central vacuoles is observed in cells treated with E-64c (cysteine protease inhibitors). Thus, during classical autophagy, the partially degraded cytoplasm in the autophagosome is likely to be released into the central vacuole for further degradation [40]. Using bafilomycin A1 and concanamycin A, which inhibits the activity of vacuolar hydrolase, it is observed that even non-degradable autophagosomes are still expelled into the central vacuole [41]. Nutrient stress also affects target of rapamycin (TOR) signaling and ultimately activates autophagy production. For example, autophagy induced by nutrient stress is inhibited in TOR-overexpressing plants, while the application of the TOR inhibitor AZD8055 in wild-type plants and raptor1a/b mutants leads to the production of structural autophagy [42]. Studies found that, although the ATG1 complex is involved in autophagy induced by nitrogen starvation and short-term carbon starvation, there is an ATG1-independent autophagy initiation pathway under long-term carbon starvation in Arabidopsis, in which the SnRK1 catalytic subunit KIN10 can directly phosphorylate the ATG6 to initiate autophagy [43].
In addition to N and C starvation, studies show that autophagy also plays an important role in other nutrient stresses. A previous study shows that autophagy can balance zinc (Zn) and iron (Fe) uptake in plants [44]. Previous studies also show that autophagy can increase zinc bioavailability in plants when zinc levels in the environment are low [45]. atg mutants exhibit more severe chlorosis compared to the wild-type under zinc-deficient conditions [45]. Zn deficiency not only induces autophagy to degrade various targets, but also observes that the amount of mobile Zn2+ in atg mutants is much lower than in the WT under Zn starvation conditions [45]. Interestingly, autophagy is also induced when zinc is in excess, and provides mobile iron ions from a non-mobile bound form to balance zinc–iron homeostasis in plants [46]. In this Zn–Fe balance process, bZIP19 and bZIP23 may be the switches to initiate/inhibit autophagy in plants under the condition of Zn deficiency/excess, and BRUTUS (BTS) may be involved in the initiation of autophagy under Fe deficiency [44]. A new study found that autophagy is also involved in the phosphate (Pi) response. However, the effect of Pi starvation on the degradation process of autophagy-regulated proteins is slight, and only a few proteins can be degraded, which can be used as characteristic target proteins under phosphorus starvation conditions for phosphorus deficiency-related studies [47].
A new study shows that autophagy is also involved in the regulation of sulfur starvation in plants [48]. Sulfur (S) remigration from the rosette to the seed is impaired in atg5 mutants compared to the wild-type [48]. These studies demonstrate that autophagy plays an indispensable role in maintaining cell homeostasis in plants under nutrient starvation.

4.2. Drought Stress

Drought stress increases the expression of many ATG genes in crops, such as ATG2 in peppers [49], ATG8a in millet [50], ATG6 in barley [51], and ATG3 and ATG18a in apples [28,52]. In tomatoes overexpressing HsfA1a, the silencing of ATG10 and ATG18f reduces HsfA1a-induced drought tolerance and autophagosome formation [53]. Conversely, the overexpression of MdATG18a in tomatoes degrades protein aggregation, limits oxidative damage, and ultimately improves drought tolerance [54]. Under drought stress, MtCAS31 promotes degradation of the MtPIP2;7 protein by autophagy, a negative regulator of drought, leading to a decrease in root hydraulic conductivity, thereby reducing water loss and improving drought tolerance [55]. In Arabidopsis, a plant-specific gene COST1, which is a negative regulator of drought, negatively regulates drought resistance by influencing the autophagy pathway [56]. COST1 co-localizes with ATG8e and the autophagy linker NBR1 on autophagosomes, suggesting a critical role in the direct regulation of autophagy [56]. A previous study found that mitochondrial alternative oxidase (AOX) may regulate autophagy through mitochondrial ROS during drought stress in tomatoes [57].

4.3. Heat and Cold Stress

Plants accumulate a large amount of oxidized and insoluble proteins at unsuitable temperatures. In this case, plants can eliminate these toxic proteins by inducing autophagy to improve plant resistance. ATG gene expression is up-regulated in various plants, and more autophagosomes are accumulated under heat stress [49,58,59]. On the contrary, silencing ATG5 or ATG7 in Arabidopsis and tomatoes under heat stress leads to heat sensitivity [58,59,60]. Plants degrade related proteins through NBR1-mediated selective autophagy under heat stress. Under heat stress, the expression of NBR1 is up-regulated and more puncta accumulate in the cytoplasm compared to the wild-type [61]. Furthermore, more NBR1 puncta accumulate in WT plants during the heat stress recovery stage, and the accumulation of the NBR1 protein is significantly higher in atg mutants than in WT plants [61,62]. Furthermore, the NBR1-mediated selective autophagy pathway degrades HSP90.1 and ROF1 to reduce plant resistance to heat stress memory [63].
Compared to heat stress, there are few studies on the regulation of cold stress by autophagy in plants. In rice, OsATG6b is down-regulated under cold stress, while OsATG6c expression is up-regulated [51]. In barley, the expression of HvATG6 is up-regulated under low temperatures [64]. This may suggest that ATG6 plays an important role in response to low plant temperature. In addition, NBR1-mediated selective autophagy also appears to be involved in plant responses to cold stress. In tomatoes, BRs (Brassinosteroids) and the positive regulator BZR1 induce autophagy and accumulation of the selective autophagy receptor NBR1 under cold stress [65].

4.4. Salt Stress

High concentrations of NaCl lead to a reduced photosynthetic rate, as well as excessive energy consumption and accumulation of excess reactive oxygen species (ROS) [66]. As an important regulator of cellular homeostasis, autophagy is also involved in the pathway of plant salt tolerance. Several autophagy genes are up-regulated under salt treatment in wheat seedlings [67]. Silencing metacaspase TaMCA-Id can reduce the tolerance of wheat seedlings to NaCl by promoting ROS production, which further participates in the regulation of autophagy and PCD triggered by NaCl treatment [68]. Within 3 h of salt treatment, accumulation of oxidized proteins in atg2 and atg7 is higher than that in the WT, and the mutants are highly sensitive to salt stress [69]. The Arabidopsis PI3K complex positively regulates salt tolerance by promoting the internalization of PIP2;1 from the plasma membrane into the vacuole under salt stress, thereby reducing root water permeability [70]. In addition, the overexpression of MdATG10 leads to increased autophagy activity in roots and enhances salt tolerance in apples [71]. Another study demonstrates spermidine (Spd), a kind of polyamine, activation of ATG gene expression and autophagosome formation under salt stress in cucumbers [72]. All of the above results indicate the role of autophagy in plants under salt stress.

5. Autophagy in Biotic Stress

Besides abiotic stresses, biotic stresses can influence autophagy. In plants, depending on the lifestyle of the pathogen, infecting the plant and activating autophagy is shown to lead to different outcomes [73]. The plant immune system is a complex mechanism, the most well-known of which is the hypersensitive response (HR)-related programmed cell death (PCD) [74]. It is reported that autophagy is involved in plant immunity by negatively regulating PCD [75].

5.1. Autophagy in Plant Viral Infection

Autophagy is a double-edged sword regarding defense against plant viruses. Plant autophagy is both antiviral and can be manipulated or inhibited by plant viruses to facilitate viral infection [76]. Meanwhile, autophagy can maintain a dynamic balance between viral infection and host survival during pathogen infection [76].
Some studies show that autophagy can defend DNA viruses from infecting plants [77]. The virulence factor βC1 from CLCuMuV is degraded by autophagy through interaction with ATG8. Thus, viral infection is enhanced in ATG7 and ATG5 mutants [78]. βC1 acts as the first plant viral activator of autophagy to activate autophagy by disrupting the interaction between ATG3 and glyceraldehyde-3-phosphate dehydrogenase, a negative regulator of autophagy [78]. In tomatoes, leaf curl Yunnan virus (TLCYnV) is degraded by autophagy through interaction with ATG8h [79]. In addition, infection by the double-stranded DNA pararetrovirus cauliflower mosaic virus (CaMV) also mediates its degradation through NBR1-regulated selective autophagy [80].
In addition to plant DNA viruses, autophagy has antiviral effects during infection by positive-strand RNA viruses. For example, the overexpression of Beclin1/ATG6 inhibits TuMV viral RNA accumulation, while the knockout of Beclin1 or ATG8a promotes its infection [81]. However, the molecular mechanism of autophagy inhibiting TuMV infection is still to be investigated further. A new study indicates that TuMV activates NBR1–ATG8f autophagy to target virus replication in the tonoplast for viral replication and accumulation [82].
Furthermore, autophagy also plays an antiviral role in negative-strand RNA virus infection. The viral suppressors of the RNAi (VSR) protein P3 from rice stripe virus (RSV) can interact with NbPI3P and can be degraded by autophagy, thereby inhibiting RSV infection [83]. In this process, eukaryotic translation initiation factor 4A (eIF4A) acts as a repressor by interacting with ATG5 to leave the ATG5–ATG12 interaction to inhibit autophagy [84]. From these current studies, autophagy can inhibit viral movement, not replication.
The above studies show that autophagy can suppress viral infection, but autophagy can also promote viral infection. For example, the overexpression of ATG5 and ATG8f promotes the virus bamboo mosaic virus (BaMV) infection [85]. This may be because ATG8f-rich virus-induced vesicles can provide sites for viral RNA replication or evade host-silencing mechanisms [85].
Other studies found that autophagy can be manipulated by certain viruses. The γb protein from the barley stripe mosaic virus (BSMV) directly binds to ATG7 and blocks the interaction of ATG7 with ATG8 to inhibit autophagosome formation [86]. In addition, the turnip crinkle virus (TCV) uses the viral silencing repressor protein P38 to inhibit antiviral autophagy, possibly by directly sequestering the ATG8 protein [87].

5.2. Autophagy and Fungi

Fungal infection is one of the main hazards of plants and poses a major threat to food security. Some fungi, such as the rice blast pathogen, infect leaves by forming a dome-shaped cellular structure called an appressorium to break through the cuticle and cell wall of the host [88]. The formation of appressorium requires the accumulation of glycerol to establish cell expansion, and the energy required for the accumulation of glycerol needs to be transported from adjacent conidial cells that undergo autophagy-related cell death [89]. Therefore, autophagy mutants cannot produce appressorium to penetrate the host. For example, the necrotizing plant pathogen Botrytis cinerea cannot form an appressorium to infect plant cells after mutating the atg1 gene [90]. In another study, a knockout mutant of a small Rab GTPases called MoYPT7 is shown to be impaired in autophagy and appressorium development in Magnaporthe oryzae [91]. In Magnaporthe oryzae, ATG1, ATG2, ATG3, ATG17, and ATG18 are shown to increase phosphorylation during appressorium formation, suggesting that post-translational modifications of ATG are involved in infecting the host [92]. Some reports found that host autophagy is also important for beneficial fungi. For example, the mycorrhizal fungus Glomus intraradices up-regulates the expression of ATG8f and ATG4a in the cortical cells and arbuscular cells of mycorrhizal colonized roots [93].
Autophagy can improve plant resistance to necrotrophic pathogens by limiting the hypersensitive response (HR) of host cells. Compared to the wild-type, multiple autophagy mutants in Arabidopsis show reduced resistance to Botrytis cinerea, specifically, severe leaf yellowing, larger lesion area, and more dead cells [94]. AtATG18a can positively regulate Arabidopsis necrotrophic pathogen immunity by interacting with WRKY33; at the same time, autophagy can induce the expression of PDF1.2 in jasmonic acid (JA) defense signals, which synergistically increases plant resistance to Botrytis cinerea [94]. Studies show that autophagy also has a positive regulatory role in response to biotrophic pathogens [95]. Silencing of TaATG8j results in the suppression of the wheat HR response after infection with Puccinia striiformis f. sp. tritici, resulting in reduced resistance [96]. Interestingly, knockout of ATG6 in wheat increases the resistance of sensitive lines to Blumeria graminis f. sp. tritici, while knockout of ATG6 in lines carrying the resistance gene Pm21 reduces resistance, indicating that ATG6 plays a complex role in wheat powdery mildew resistance [97].

5.3. Autophagy and Bacteria

Bacteria have evolved in many ways to manipulate host cells for successful infection. Numerous studies show that intracellular bacteria can manipulate autophagy as a pro-survival strategy [98].
Pseudomonas syringae is a Gram-negative bacterium with strong aerobic and saprotrophic properties, and the incidence of plant diseases caused by Pseudomonas syringae ranks first in the top 10 bacterial plant diseases [99]. Some studies show that autophagy has reverse resistance to Pseudomonas syringae. Arabidopsis atg5, atg18a mutants significantly reduce host cell HR after infection with the bacterium Pseudomonas syringae pathovar (pv) tomato (Pst DC3000) with the effector proteins AvrRps4 or AvrRpm1 [100]. Furthermore, when infected with Pseudomonas syringae, Arabidopsis atg5, atg10, and atg18a mutant leaves do not spread cell necrosis, and the plants show obvious resistance [101].
NBR1-mediated selective autophagy is also associated with immunity against bacteria [102,103]. While Pst-induced autophagy promotes bacterial proliferation, NBR1-mediated selective autophagy counteracts its induction of water-soaked lesions by inhibiting the formation of the Hrp outer protein M1(HopM1) and enhances Pst resistance in Arabidopsis [102]. While the way in which NBR1 targets pathogens and promotes local immunity remains unknown, these findings point to a more complex function of selective autophagy proteins in cell immunity [104]. A recent study uncovers the complex ways in which pathogens interact with their hosts. The type-III effector XopL can interact with the autophagy component SH3P2 through E3 ligase activity and degrade it to promote infection, while XopL is degraded by NBR1-mediated selective autophagy [105].

6. Conclusions and Future Perspectives

Plant autophagy research currently focuses on fundamental research, and how to apply it to crop improvement is an issue that needs to be considered in the future. The most direct approach is to use plant growth regulators to induce autophagy. For example, exogenous spraying of melatonin can improve the heat tolerance of tomatoes; it may be that melatonin increases the expression of ATGs and the formation of autophagic vesicles at high temperatures to degrade the denatured proteins produced under heat stress [106]. Exogenous BR treatment can also enhance tomato resistance to nitrogen starvation and cold stress through brassinazole resistance 1 (BZR1)-mediated autophagy [65,107]. Ethylene (ETH) is also reported to induce ATG expression and ROS levels to promote the survival of soybeans and tomatoes under flooding and hypoxia stress [108]. In addition to the use of hormone regulation, the development of biopesticides can also directly regulate the autophagy activity of pathogens and pests to reduce disease transmission, which is also an effective means to improve crop resistance. For example, since autophagy promotes the replication of the rice gall dwarf virus (RGDV) in leafhoppers, spraying the autophagy inhibitor 3-MA can reduce the spread of the RGDV virus [109]. With the development of nanotechnology, it is also reported that titanium dioxide (TiO2) and zinc oxide (ZnO) nanoparticles are involved in the regulation of plant autophagy [110,111]. The above three methods, using external spraying reagents, provide feasibility for the application of autophagy in agricultural production. Moreover, the use of gene editing or other means of gene manipulation to regulate the ATG genes is also a direction that can be developed in the future.
In recent years, the role of autophagy in plant growth and development, abiotic stress, and plant–microbe interactions has been clarified in a variety of plants (Table 1). The core process of autophagy is highly conserved in eukaryotes, and the functions and regulatory networks of autophagy are specific in different species. Although many ATG proteins have been identified, there are few studies on whether these proteins have other functions. In addition, although autophagy is involved in many life processes, the way to coordinate hormonal signals to regulate plant growth and stress resistance needs further research. Furthermore, the mechanism of autophagy regulation should be continuously studied, as it not only has theoretical significance, but also has very important application value for future crop breeding.

Author Contributions

The authors confirm their contributions to this work: F.L. conceived the project, S.W. wrote the manuscript, and W.H. and F.L. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the Lushan Botanical Garden, Chinese Academy of Sciences [No. 2021ZWZX09 to Song Wang; No. 2021ZWZX01 to Fen Liu], by grants from the National Natural Science Foundation of China [32100297 to Fen Liu], and by grants from Jiangxi high-level and urgently needed overseas talent introduction plan [20212BCJ25024 to Fen Liu].

Informed Consent Statement

Not applicable.

Acknowledgments

Due to space limitations, we apologize to the authors whose works are not cited.

Conflicts of Interest

I declare that I have no conflict of interest.

References

  1. 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] [PubMed]
  2. Liu, Y.; Bassham, D.C. Autophagy: Pathways for Self-Eating in Plant Cells. Annu. Rev. Plant Biol. 2012, 63, 215–237. [Google Scholar] [CrossRef]
  3. Su, W.; Bao, Y.; Yu, X.; Xia, X.; Liu, C.; Yin, W. Autophagy and Its Regulators in Response to Stress in Plants. Int. J. Mol. Sci. 2020, 21, 8889. [Google Scholar] [CrossRef] [PubMed]
  4. Van Doorn, W.G.; Papini, A. The ultrastructure of autophagy in plant cells: A review. Autophagy 2013, 9, 1922–1936. [Google Scholar] [CrossRef]
  5. Sie’nko, K.; Poormassalehgoo, A.; Yamada, K.; Goto-Yamada, S. Microautophagy in Plants: Consideration of Its Molecular Mechanism. Cells 2020, 9, 887. [Google Scholar] [CrossRef] [PubMed]
  6. Chanoca, A.; Kovinich, N.; Burkel, B.; Stecha, S.; Bohorquez-Restrepo, A.; Ueda, T.; Eliceiri, K.W.; Grotewold, E.; Otegui, M.S. Anthocyanin vacuolar inclusions form by a microautophagy mechanism. Plant Cell 2015, 27, 2545–2559. [Google Scholar] [CrossRef] [PubMed]
  7. Nakamura, S.; Hidema, J.; Sakamoto, W.; Ishida, H.; Izumi, M. Selective elimination of membrane-damaged chloroplasts via microautophagy. Plant Physiol. 2018, 177, 1007–1026. [Google Scholar] [CrossRef]
  8. Feng, Y.C.; He, D.; Yao, Z.Y.; Klionsky, D.J. The machinery of macroautophagy. Cell Res. 2014, 24, 24–41. [Google Scholar] [CrossRef]
  9. Hatsugai, N.; Kuroyanagi, M.; Yamada, K.; Meshi, T.; Tsuda, S.; Kondo, M.; Nishimura, M.; Hara-Nishimura, I. A plant vacuolar protease, VPE, mediates virus-induced hypersensitive cell death. Science 2004, 305, 855–858. [Google Scholar] [CrossRef]
  10. Nakatogawa, H. Mechanisms governing autophagosome biogenesis. Nat. Rev. Mol. Cell. Biol. 2020, 21, 439–458. [Google Scholar] [CrossRef]
  11. Minina, E.A.; Moschou, P.N.; Vetukuri, R.R.; Sanchez-Vera, V.; Cardoso, C.; Liu, Q.S.; Elander, P.H.; Dalman, K.; Beganovic, M.; Yilmaz, J.L.; et al. Transcriptional stimulation of rate-limiting components of the autophagic pathway improves plant fitness. J. Exp. Bot. 2018, 69, 1415–1432. [Google Scholar] [CrossRef]
  12. Angelovici, R.; Fait, A.; Zhu, X.; Szymanski, J.; Feldmesser, E.; Fernie, A.R.; Galili, G. Deciphering transcriptional and metabolic networks associated with lysine metabolism during Arabidopsis seed development. Plant Physiol. 2009, 151, 2058–2072. [Google Scholar] [CrossRef] [PubMed]
  13. Li, F.; Chung, T.; Pennington, J.G.; Federico, M.L.; Kaeppler, H.F.; Kaeppler, S.M.; Otegui, M.S.; Vierstra, R.D. Autophagic recycling plays a central role in maize nitrogen remobilization. Plant Cell 2015, 27, 1389–1408. [Google Scholar] [CrossRef] [PubMed]
  14. Levanony, H.; Rubin, R.; Altschuler, Y.; Galili, G. Evidence for a novel route of wheat storage proteins to vacuoles. J. Cell Biol. 1992, 119, 1117–1128. [Google Scholar] [CrossRef] [PubMed]
  15. Berardino, D.J.; Marmagne, A.; Berger, A.; Yoshimoto, K.; Cueff, G.; Chardon, F.; Masclaux-Daubresse, C.; Reisdorf-Cren, M. Autophagy controls resource allocation and protein storage accumulation in Arabidopsis seeds. J. Exp. Bot. 2018, 69, 1403–1414. [Google Scholar] [CrossRef] [PubMed]
  16. Honig, A.; Avin-Wittenberg, T.; Ufaz, S.; Galili, G. A new type of compartment, defined by plant-specific Atg8-interacting proteins, is induced upon exposure of Arabidopsis plants to carbon starvation. Plant Cell 2012, 24, 288–303. [Google Scholar] [CrossRef]
  17. Ortiz, R.; Geleta, M.; Gustafsson, C.; Lager, I.; Hofvander, P.; Löfstedt, C.; Cahoon, E.B.; Minina, E.; Bozhkov, P.; Stymne, S. Oil crops for the future. Curr. Opin. Plant Biol. 2020, 56, 181–189. [Google Scholar] [CrossRef]
  18. Huang, A.H.C. Plant lipid droplets and their associated proteins: Potential for rapid advances. Plant Physiol. 2018, 176, 1894–1918. [Google Scholar] [CrossRef]
  19. Fan, J.; Yu, L.; Xu, C. Dual role for autophagy in lipid metabolism in Arabidopsis. Plant Cell 2019, 31, 1598–1613. [Google Scholar] [CrossRef]
  20. Evert, R.F. Esau’s plant anatomy, meristems, cells, and tissues of the plant body: Their structure, function, and development. 3rd edn. Ann. Bot. 2007, 99, 785–786. [Google Scholar]
  21. Wojciechowska, N.; Michalak, K.M.; Bagniewska-Zadworna, A. Autophagy-an underestimated coordinator of construction and destruction during plant root ontogeny. Planta 2021, 254, 15. [Google Scholar] [CrossRef] [PubMed]
  22. Sláviková, S.; Shy, G.; Yao, Y.; Glozman, R.; Levanony, H.; Pietrokovski, S.; Elazar, Z.; Galili, G. The autophagy-associated ATG8 gene family operates both under favourable growth conditions and under starvation stresses in Arabidopsis plants. J. Exp. Bot. 2005, 56, 2839–2849. [Google Scholar] [CrossRef] [PubMed]
  23. Inoue, Y.; Suzuki, T.; Hattori, M.; Yoshimoto, K.; Ohsumi, Y.; Moriyasu, Y. AtATG genes, homologs of yeast autophagy genes, are involved in constitutive autophagy in Arabidopsis root tip cells. Plant Cell Physiol. 2006, 47, 1641–1652. [Google Scholar] [CrossRef] [PubMed]
  24. Bagniewska-Zadworna, A.; Arasimowicz-Jelonek, M.; Smolinski, D.J.; Stelmasik, A. New insights into pioneer root xylem development: Evidence obtained from Populus trichocarpa plants grown under field conditions. Ann. Bot. 2014, 113, 1235–1247. [Google Scholar] [CrossRef]
  25. Wojciechowska, N.; Smugarzewska, I.; Marzec-Schmidt, K.; ZarzyńskaNowak, A.; Bagniewska-Zadworna, A. Occurrence of autophagy during pioneer root and stem development in Populus trichocarpa. Planta 2019, 250, 1789–1801. [Google Scholar] [CrossRef]
  26. Wojciechowska, N.; Marzec-Schmidt, K.; Kalemba, E.M.; Zarzyńska-Nowak, A.; Jagodziński, A.M.; Bagniewska-Zadworna, A. Autophagy counteracts instantaneous cell death during seasonal senescence of the fine roots and leaves in Populus trichocarpa. BMC Plant Biol. 2018, 18, 260. [Google Scholar] [CrossRef]
  27. Hollmann, J.; Gregersen, P.L.; Krupinska, K. Identification of predominant genes involved in regulation and execution of senescence-associated nitrogen remobilization in flag leaves of field grown barley. J. Exp. Bot. 2014, 65, 3963–3973. [Google Scholar] [CrossRef]
  28. Wada, S.; Hayashida, Y.; Izumi, M.; Kurusu, T.; Hanamata, S.; Kanno, K.; Kojima, S.; Yamaya, T.; Kuchitsu, K.; Makino, A.; et al. Autophagy supports biomass production and nitrogen use efficiency at the vegetative stage in rice. Plant Physiol. 2015, 168, 60–73. [Google Scholar] [CrossRef]
  29. Yoshimoto, K.; Jikumaru, Y.; Kamiya, Y.; Kusano, M.; Consonni, C.; Panstruga, R.; Ohsumi, Y.; Shirasu, K. Autophagy negatively regulates cell death by controlling NPR1-dependent salicylic acid signaling during senescence and the innate immune response in Arabidopsis. Plant Cell 2009, 21, 2914–2927. [Google Scholar] [CrossRef]
  30. Sun, X.; Jia, X.; Huo, L.; Che, R.; Gong, X.; Wang, P.; Ma, F. MdATG18a overexpression improves tolerance to nitrogen deficiency and regulates anthocyanin accumulation through increased autophagy in transgenic apple. Plant Cell Environ. 2018, 41, 469–480. [Google Scholar] [CrossRef]
  31. Poret, M.; Chandrasekar, B.; van der Hoorn, R.A.; Avice, J.C. Characterization of senescence-associated protease activities involved in the efficient protein remobilization during leaf senescence of winter oilseed rape. Plant Sci. 2016, 246, 139–153. [Google Scholar] [CrossRef]
  32. James, M.; Poret, M.; Masclaux-Daubresse, C.; Marmagne, A.; Coquet, L.; Jouenne, T.; Chan, P.; Trouverie, J.; Etienne, P. SAG12, a major cysteine protease involved in nitrogen allocation during senescence for seed production in Arabidopsis thaliana. Plant Cell Physiol. 2018, 59, 2052–2063. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, W.; Zhang, L.; Xing, S.; Ma, Z.; Liu, J.; Gu, H.; Qin, G.; Qu, L. Arabidopsis AtVPS15 plays essential roles in pollen germination possibly by interacting with AtVPS34. J. Genet. Genom. 2012, 39, 81–92. [Google Scholar] [CrossRef] [PubMed]
  34. Fujiki, Y.; Yoshimoto, K.; Ohsumi, Y. An Arabidopsis homolog of yeast ATG6/VPS30 is essential for pollen germination. Plant Physiol. 2007, 143, 1132–1139. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, Y.; Kim, E.; Choi, Y.; Hwang, I.; Staiger, C.J.; Chung, Y.; Lee, Y. The Arabidopsis phosphatidylinositol 3-kinase is important for pollen development. Plant Physiol. 2008, 147, 1886–1897. [Google Scholar] [CrossRef]
  36. Kurusu, T.; Koyano, T.; Hanamata, S.; Kubo, T.; Noguchi, Y.; Yagi, C.; Nagata, N.; Yamamoto, T.; Ohnishi, T.; Okazaki, Y.; et al. OsATG7 is required for autophagy-dependent lipid metabolism in rice postmeiotic anther development. Autophagy 2014, 10, 878–888. [Google Scholar] [CrossRef]
  37. Zhao, P.; Zhou, X.; Zhao, L.; Cheung, A.Y.; Sun, M. Autophagy-mediated compartmental cytoplasmic deletion is essential for tobacco pollen germination and male fertility. Autophagy 2020, 16, 2180–2192. [Google Scholar] [CrossRef]
  38. Yan, H.; Zhuang, M.; Xu, X.; Li, S.; Yang, M.; Li, N.; Du, X.; Hu, K.; Peng, X.; Huang, W.; et al. Autophagy and its mediated mitochondrial quality control maintain pollen tube growth and male fertility in Arabidopsis. Autophagy 2022, 18, 1–16. [Google Scholar] [CrossRef]
  39. Takatsuka, C.; Inoue, Y.; Matsuoka, K.; Moriyasu, Y. 3-methyladenine inhibits autophagy in tobacco culture cells under sucrose starvation conditions. Plant Cell Physiol. 2004, 45, 265–274. [Google Scholar] [CrossRef]
  40. Moriyasu, Y.; Ohsumi, Y. Autophagy in tobacco suspension-cultured cells in response to sucrose starvation. Plant Physiol. 1996, 111, 1233–1241. [Google Scholar] [CrossRef]
  41. 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] [PubMed]
  42. Pu, Y.; Luo, X.; Bassham, D.C. TOR-dependent and -independent pathways regulate autophagy in Arabidopsis thaliana. Front. Plant Sci. 2017, 8, 1204. [Google Scholar] [CrossRef] [PubMed]
  43. Huang, X.; Zheng, C.; Liu, F.; Yang, C.; Zheng, P.; Lu, X.; Tian, J.; Chung, T.; Otegui, M.S.; Xiao, S.; et al. Genetic analyses of the Arabidopsis ATG1 kinase complex reveal both kinase-dependent and independent autophagic routes during fixed-carbon starvation. Plant Cell 2019, 31, 2973–2995. [Google Scholar] [CrossRef] [PubMed]
  44. Shinozaki, D.; Yoshimoto, K. Autophagy balances the zinc-iron seesaw caused by Zn-stress. Trends Plant Sci. 2021, 26, 882–884. [Google Scholar] [CrossRef]
  45. Shinozaki, D.; Merkulova, E.A.; Naya, L.; Horie, T.; Kanno, Y.; Seo, M.; Ohsumi, Y.; Masclaux-Daubresse, C.; Yoshimoto, K. Autophagy increases zinc bioavailability to avoid light-mediated reactive oxygen species production under zinc deficiency. Plant Physiol. 2020, 182, 1284–1296. [Google Scholar] [CrossRef]
  46. Shinozaki, D.; Tanoi, K.; Yoshimoto, K. Optimal distribution of iron to sink organs via autophagy is important for tolerance to excess zinc in Arabidopsis. Plant Cell Physiol. 2021, 62, 515–527. [Google Scholar] [CrossRef]
  47. Li, L.; Lee, C.P.; Ding, X.; Qin, Y.; Wijerathna-Yapa, A.; Broda, M.; Otegui, M.S.; Millar, A.H. Defects in autophagy lead to selective in vivo changes in turnover of cytosolic and organelle proteins in Arabidopsis. Plant Cell 2022, 34, koac185. [Google Scholar] [CrossRef]
  48. Lornac, A.; Havé, M.; Chardon, F.; Soulay, F.; Clément, G.; Avice, J.; Masclaux-Daubresse, C. Autophagy Controls Sulphur Metabolism in the Rosette Leaves of Arabidopsis and Facilitates S Remobilization to the Seeds. Cells 2020, 9, 332. [Google Scholar] [CrossRef]
  49. Zhai, Y.; Guo, M.; Wang, H.; Lu, J.; Liu, J.; Zhang, C.; Gong, Z.; Lu, M. Autophagy, a conserved mechanism for protein degradation, responds to heat, and other abiotic stresses in Capsicum annuum L. Front. Plant Sci. 2016, 7, 131. [Google Scholar] [CrossRef]
  50. 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]
  51. Zeng, X.; Zeng, Z.; Liu, C.; Yuan, W.; Hou, N.; Bian, H.; Zhu, M.; Han, N. A barley homolog of yeast ATG6 is involved in multiple abiotic stress responses and stress resistance regulation. Plant Physiol. Biochem. 2017, 115, 97–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Wang, P.; Sun, X.; Jia, X.; Ma, F. Apple autophagy-related protein MdATG3s afford tolerance to multiple abiotic stresses. Plant Sci. 2017, 256, 53–64. [Google Scholar] [CrossRef] [PubMed]
  53. 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]
  54. 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]
  55. Li, X.; Liu, Q.; Feng, H.; Deng, J.; Zhang, R.; Wen, J.; Dong, J.; Wang, T. Dehydrin MtCAS31 promotes autophagic degradation under drought stress. Autophagy 2020, 16, 862–877. [Google Scholar] [CrossRef] [PubMed]
  56. Bao, Y.; Song, W.; Wang, P.; Yu, X.; Li, B.; Jiang, C.; Shiu, S.; Zhang, H.; Bassham, D.C. COST1 regulates autophagy to control plant drought tolerance. Proc. Natl. Acad. Sci. USA 2020, 117, 7482–7493. [Google Scholar] [CrossRef]
  57. Zhu, T.; Zou, L.; Li, Y.; Yao, X.; Xu, F.; Deng, X.; Zhang, D.; Lin, H. Mitochondrial alternative oxidasedependent autophagy involved in ethylenemediated drought tolerance in Solanum lycopersicum. Plant Biotechnol. J. 2018, 16, 2063–2076. [Google Scholar] [CrossRef]
  58. Zhou, J.; Wang, J.; Yu, J.; Chen, Z. Role and regulation of autophagy in heat stress responses of tomato plants. Front. Plant Sci. 2014, 5, 174. [Google Scholar] [CrossRef]
  59. Cheng, F.; Yin, L.; Zhou, J.; Xia, X.; Shi, K.; Yu, J.; Zhou, Y.; Foyer, C. Interactions between 2-Cys peroxiredoxins and ascorbate in autophagosome formation during the heat stress response in Solanum lycopersicum. J. Exp. Bot. 2016, 67, 1919–1933. [Google Scholar] [CrossRef]
  60. Zhou, J.; Wang, J.; Cheng, Y.; Chi, Y.; Fan, B.; Yu, J.; Chen, Z. Correction: NBR1-mediated selective autophagy targets insoluble ubiquitinated protein aggregates in plant stress responses. PLoS Genet. 2013, 9, e1004477. [Google Scholar] [CrossRef]
  61. Jung, H.; Lee, H.N.; Marshall, R.S.; Lomax, A.W.; Yoon, M.J.; Kim, J.; Kim, J.H.; Vierstra, R.D.; Chung, T. Arabidopsis cargo receptor NBR1 mediates selective autophagy of defective proteins. J. Exp. Bot. 2020, 71, 73–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Thirumalaikumar, V.P.; Gorka, M.; Schulz, K.; Masclaux-Daubresse, C.; Sampathkumar, A.; Skirycz, A.; Vierstra, R.D.; Balazadeh, S. Selective autophagy regulates heat stress memory in Arabidopsis by NBR1-mediated targeting of HSP90 and ROF1. Autophagy 2020, 16, 1–16. [Google Scholar]
  63. Rana, R.M.; Dong, S.; Ali, Z.; Huang, J.; Zhang, H.S. Regulation of ATG6/Beclin-1 homologs by abiotic stresses and hormones in rice (Oryza sativa L.). Genet. Mol. Res. 2012, 11, 3676–3687. [Google Scholar] [CrossRef] [PubMed]
  64. Chi, C.; Li, X.; Fang, P.; Xia, X.; Shi, K.; Zhou, Y.; Zhou, J.; Yu, J. Brassinosteroids act as a positive regulator of NBR1-dependent selective autophagy in response to chilling stress in tomato. J. Exp. Bot. 2020, 71, 1092–1106. [Google Scholar] [CrossRef] [PubMed]
  65. Julkowska, M.M.; Testerink, C. Tuning plant signaling and growth to survive salt. Trends Plant Sci. 2015, 20, 586–594. [Google Scholar] [CrossRef]
  66. Yue, W.; Nie, X.; Cui, L.; Zhi, Y.; Zhang, T.; Du, X.; Song, W. Genome-wide sequence and expressional analysis of autophagy gene family in bread wheat (Triticum aestivum L.). J. Plant Physiol. 2018, 229, 7–21. [Google Scholar] [CrossRef]
  67. Yue, J.; Wang, Y.; Jiao, J.; Wang, W.; Wang, H. The Metacaspase TaMCA-Id Negatively Regulates Salt-Induced Programmed Cell Death and Functionally Links with Autophagy in Wheat. Front. Plant Sci. 2022, 13, 904933. [Google Scholar] [CrossRef]
  68. Luo, L.; Zhang, P.; Zhu, R.; Fu, J.; Su, J.; Zheng, J.; Wang, Z.; Wang, D.; Gong, Q. Autophagy Is Rapidly Induced by Salt Stress and Is Required for Salt Tolerance in Arabidopsis. Front. Plant Sci. 2017, 8, 1459. [Google Scholar] [CrossRef]
  69. Ueda, M.; Tsutsumi, N.; Fujimoto, M. Salt stress induces internalization of plasma membrane aquaporin into the vacuole in Arabidopsis thaliana. Biochem. Biophys. Res. Commun. 2016, 474, 742–746. [Google Scholar] [CrossRef] [PubMed]
  70. Huo, L.; Guo, Z.; Jia, X.; Sun, X.; Wang, P.; Gong, X.; Ma, F. Increased autophagic activity in roots caused by overexpression of the autophagy-related gene MdATG10 in apple enhances salt tolerance. Plant Sci. 2020, 294, 110444. [Google Scholar] [CrossRef] [PubMed]
  71. Zhang, Y.; Wang, Y.; Wen, W.; Shi, Z.; Gu, Q.; Ahammed, G.J.; Cao, K.; Shah Jahan, M.; Shu, S.; Wang, J.; et al. Hydrogen peroxide mediates spermidine-induced autophagy to alleviate salt stress in cucumber. Autophagy 2021, 17, 2876–2890. [Google Scholar] [CrossRef] [PubMed]
  72. Leary, A.Y.; Sanguankiattichai, N.; Duggan, C.; Tumtas, Y.; Pandey, P.; Segretin, M.E.; Linares, J.S.; Savage, Z.D.; Yow, R.J.; Bozkurt, T.O. Modulation of plant autophagy during pathogen attack. J. Exp. Bot. 2018, 69, 1325–1333. [Google Scholar] [CrossRef] [Green Version]
  73. Coll, N.S.; Epple, P.; Dangl, J.L. Programmed cell death in the plant immune system. Cell Death Differ. 2011, 18, 1247–1256. [Google Scholar] [CrossRef]
  74. Liu, Y.; Schiff, M.; Czymmek, K.; Tallóczy, Z.; Levine, B.; Dinesh-Kumar, S.P. Autophagy regulates programmed cell death during the plant innate immune response. Cell 2005, 121, 567–577. [Google Scholar] [CrossRef]
  75. Yang, M.; Liu, Y. Autophagy in plant viral infection. FEBS Lett. 2022, 596, 2152–2162. [Google Scholar] [CrossRef] [PubMed]
  76. Haxim, Y.; Ismayil, A.; Jia, Q.; Wang, Y.; Zheng, X.; Chen, T.; Qian, L.; Liu, N.; Wang, N.; Han, S.; et al. Autophagy functions as an antiviral mechanism against geminiviruses in plants. eLife 2017, 6, e23897. [Google Scholar] [CrossRef] [PubMed]
  77. Han, S.; Wang, Y.; Zheng, X.; Jia, Q.; Zhao, J.; Bai, F.; Hong, Y.; Liu, Y. Cytoplastic glyceraldehyde-3-phosphate dehydrogenases interact with ATG3 to negatively regulate autophagy and immunity in Nicotiana benthamiana. Plant Cell 2015, 27, 1316–1331. [Google Scholar] [CrossRef]
  78. Li, F.; Zhang, M.; Zhang, C.; Zhou, X. Nuclear autophagy degrades a geminivirus nuclear protein to restrict viral infection in solanaceous plants. New Phytol. 2020, 225, 1746–1761. [Google Scholar] [CrossRef]
  79. Hafren, 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]
  80. Li, F.; Zhang, C.; Li, Y.; Wu, G.; Hou, X.; Zhou, X.; Wang, A. Beclin1 restricts RNA virus infection in plants through suppression and degradation of the viral polymerase. Nat. Commun. 2018, 9, 1268. [Google Scholar] [CrossRef]
  81. Li, F.; Zhang, C.; Tang, Z.; Zhang, L.; Dai, Z.; Lyu, S.; Li, Y.; Hou, X.; Bernards, M.; Wang, A. A plant RNA virus activates selective autophagy in a UPR-dependent manner to promote virus infection. New Phytol. 2020, 228, 622–639. [Google Scholar] [CrossRef] [PubMed]
  82. Jiang, L.; Lu, Y.; Zheng, X.; Yang, X.; Chen, Y.; Zhang, T.; Zhao, X.; Wang, S.; Zhao, X.; Song, X.; et al. The plant protein NbP3IP directs degradation of rice stripe virus p3 silencing suppressor protein to limit virus infection through interaction with the autophagy related protein NbATG8. New Phytol. 2021, 229, 1036–1051. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, X.; Yin, Y.; Su, Y.; Jia, Z.; Jiang, L.; Lu, Y.; Zheng, H.; Peng, J.; Rao, S.; Wu, G.; et al. eIF4A, a target of siRNA derived from rice stripe virus, negatively regulates antiviral autophagy by interacting with ATG5 in Nicotiana benthamiana. PLoS Pathog. 2021, 17, e1009963. [Google Scholar] [CrossRef]
  84. Huang, Y.; Huang, Y.; Hsiao, Y.; Li, S.; Hsu, Y.; Tsai, C. Autophagy is involved in assisting the replication of Bamboo mosaic virus in Nicotiana benthamiana. J. Exp. Bot. 2019, 70, 4657–4669. [Google Scholar] [CrossRef] [PubMed]
  85. Yang, M.; Zhang, Y.; Xie, X.; Yue, N.; Li, J.; Wang, X.; Han, C.; Yu, J.; Liu, Y.; Li, D. Barley stripe mosaic virus γb protein subverts autophagy to promote viral infection by disrupting the ATG7-ATG8 interaction. Plant Cell 2018, 30, 1582–1595. [Google Scholar] [CrossRef] [PubMed]
  86. Shukla, A.; Hoffmann, G.; Hofius, D.; Hafren, A. Turnip crinkle virus targets host ATG8 proteins to attenuate antiviral autophagy. bioRxiv 2021. [CrossRef]
  87. Talbot, N.J. On the trail of a cereal killer: Exploring the biology of Magnaporthe grisea. Annu. Rev. Microbiol. 2003, 57, 177–202. [Google Scholar] [CrossRef] [PubMed]
  88. Wilson, R.A.; Talbot, N.J. Under pressure: Investigating the biology of plant infection by Magnaporthe oryzae. Nat. Rev. Microbiol. 2009, 7, 185–195. [Google Scholar] [CrossRef]
  89. Ren, W.; Zhang, Z.; Shao, W.; Yang, Y.; Zhou, M.; Chen, C. The autophagy-related gene BcATG1 is involved in fungal development and pathogenesis in Botrytis cinerea. Mol. Plant Pathol. 2017, 18, 238–248. [Google Scholar] [CrossRef]
  90. Liu, X.; Chen, S.; Gao, H.; Ning, G.; Shi, H.; Wang, Y.; Dong, B.; Qi, Y.; Zhang, D.; Lu, G.; et al. The small GTPase MoYpt7 is required for membrane fusion in autophagy and pathogenicity of Magnaporthe oryzae. Environ. Microbiol. 2015, 17, 4495–4510. [Google Scholar] [CrossRef]
  91. Franck, W.L.; Gokce, E.; Randall, S.M.; Oh, Y.; Eyre, A.; Muddiman, D.C.; Dean, R.A. Phosphoproteome analysis links protein phosphorylation to cellular remodeling and metabolic adaptation during Magnaporthe oryzae appressorium development. J. Proteome Res. 2015, 14, 2408–2424. [Google Scholar] [CrossRef] [PubMed]
  92. Gaude, N.; Bortfeld, S.; Duensing, N.; Lohse, M.; Krajinski, F. Arbuscule-containing and non-colonized cortical cells of mycorrhizal roots undergo extensive and specific reprogramming during arbuscular mycorrhizal development. Plant J. 2012, 69, 510–528. [Google Scholar] [CrossRef]
  93. Lai, Z.; Wang, F.; Zheng, Z.; Fan, B.; Chen, Z. A critical role of autophagy in plant resistance to necrotrophic fungal pathogens. Plant J. 2011, 66, 953–968. [Google Scholar] [CrossRef] [PubMed]
  94. Signorelli, S.; Tarkowski, Ł.P.; Ende Wim, V.E.; Bassham, D.V. Linking autophagy to abiotic and biotic stress responses. Trends Plant Sci. 2019, 24, 413–430. [Google Scholar] [CrossRef]
  95. Mamun, M.A.A.; Tang, C.; Sun, Y.; Islam, M.N.; Liu, P.; Wang, X.; Kang, Z. Wheat gene TaATG8j contributes to stripe rust resistance. Int. J. Mol. Sci. 2018, 19, 1666. [Google Scholar] [CrossRef] [PubMed]
  96. Yue, J.; Sun, H.; Zhang, W.; Pei, D.; He, Y.; Wang, H. Wheat homologs of yeast ATG6 function in autophagy and are implicated in powdery mildew immunity. BMC Plant Biol. 2015, 15, 95. [Google Scholar] [CrossRef]
  97. Winchell, C.G.; Steele, S.; Kawula, T.; Voth, D.E. Dining in: Intracellular bacterial pathogen interplay with autophagy. Curr. Opin. Microbiol. 2016, 29, 9–14. [Google Scholar] [CrossRef]
  98. Wang, D.; Wang, Q. Research advancement of molecular biology in Pseudomonas syringae. Acta Agric. Boreali-Occident. Sin. 2017, 26, 487–496. [Google Scholar]
  99. Hofius, D.; Li, L.; Hafrén, A.; Coll, N.S. Autophagy as an emerging arena for plant-pathogen interactions. Curr. Opin. Plant Biol. 2017, 38, 117–123. [Google Scholar] [CrossRef]
  100. Lenz, H.D.; Haller, E.; Melzer, E.; Kober, K.; Wurster, K.; Stahl, M.; Bassham, D.C.; Vierstra, R.D.; Parker, J.E.; Bautor, J.; et al. Autophagy differentially controls plant basal immunity to biotrophic and necrotrophic pathogens. Plant J. 2011, 66, 818–830. [Google Scholar] [CrossRef]
  101. Üstün, S.; Hafre’n, A.; Liu, Q.; Marshall, R.S.; Minina, E.A.; Bozhkov, P.V.; Vierstra, R.D.; Hofius, D. Bacteria exploit autophagy for proteasome degradation and enhanced virulence in plants. Plant Cell 2018, 30, 668–685. [Google Scholar] [CrossRef] [PubMed]
  102. Dagdas, Y.F.; Pandey, P.; Tumtas, Y.; Sanguankiattichai, N.; Belhaj, K.; Duggan, C.; Leary, A.Y.; Segretin, M.E.; Contreras, M.P.; Savage, Z.; et al. Host autophagy machinery is diverted to the pathogen interface to mediate focal defense responses against the Irish potato famine pathogen. eLife 2018, 7, e37476. [Google Scholar] [CrossRef] [PubMed]
  103. Leary, A.Y.; Savage, Z.; Tumtas, Y.; Bozkurt, T.O. Contrasting and emerging roles of autophagy in plant immunity. Curr. Opin. Plant Biol. 2019, 52, 46–53. [Google Scholar] [CrossRef]
  104. Xu, W.; Cai, S.Y.; Zhang, Y.; Wang, Y.; Ahammed, G.J.; Xia, X.; Shi, K.; Zhou, Y.; Yu, J.; Reiter, R.J.; et al. Melatonin enhances thermotolerance by promoting cellular protein protection in tomato plants. J. Pineal Res. 2016, 61, 457–469. [Google Scholar] [CrossRef]
  105. Leong, J.; Raffeiner, M.; Spinti, D.; Langin, G.; Franz-Wachtel, M.; Guzman, A.; Kim, J.; Pandey, P.; Minina, A.; Macek, B.; et al. A bacterial effector counteracts host autophagy by promoting degradation of an autophagy component. EMBO J. 2022, 41, e110352. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, Y.; Cao, J.; Wang, K.; Xia, X.J.; Shi, K.; Zhou, Y.; Yu, J.; Zhou, J. 4BZR1 mediates brassinosteroid-induced autophagy and nitrogen starvation in tomato. Plant Physiol. 2019, 179, 671–685. [Google Scholar] [CrossRef] [PubMed]
  107. Hartman, S.; Sasidharan, R.; Voesenek, L.A.C.J. The role of ethylene in metabolic acclimations to low oxygen. New Phytol. 2021, 229, 64–70. [Google Scholar] [CrossRef]
  108. Chen, Y.; Chen, Q.; Li, M.M.; Li, M.; Mao, Q.Z.; Chen, H.Y.; Wu, W.; Jia, D.S.; Wei, T.Y. Autophagy pathway induced by a plant virus facilitates viral spread and transmission by its insect vector. PLoS Pathog. 2017, 13, e1006727. [Google Scholar] [CrossRef]
  109. Shull, T.E.; Kurepa, J.; Smalle, J.A. Anatase TiO2 nanoparticles induce autophagy and chloroplast degradation in thale cress (Arabidopsis thaliana). Environ. Sci. Technol. 2019, 53, 9522–9532. [Google Scholar] [CrossRef]
  110. Balážová, Ľ.; Baláž, M.; Babula, P. Zinc oxide nanoparticles damage tobacco BY-2 cells by oxidative stress followed by processes of autophagy and programmed cell death. Nanomaterials 2020, 10, 1066. [Google Scholar] [CrossRef]
  111. Islam, M.M.; Ishibashi, Y.; Nakagawa, A.C.S.; Tomita, Y.; Iwaya-Inoue, M.; Arima, S.; Zheng, S.H. Nitrogenre distribution and its relationship with the expression of GmATG8c during seed filling in soybean. J. Plant Physiol. 2016, 192, 71–74. [Google Scholar] [CrossRef] [PubMed]
  112. Avila-Ospina, L.; Marmagne, A.; Soulay, F.; Masclaux-Daubresse, C. Identification of barley (Hordeum vulgare L.) autophagy genes and their expression levels during leaf senescence, chronic nitrogen limitation and in response to dark exposure. Agronomy 2016, 6, 15. [Google Scholar] [CrossRef]
  113. Barany, I.; Berenguer, E.; Solis, M.T.; Perez-Perez, Y.; Santamaria, M.E.; Crespo, J.L.; Risueno, M.C.; Diaz, I.; Testillano, P.S. Autophagy is activated and involved in cell death with participation of cathepsins during stress-induced microspore embryogenesis in barley. J. Exp. Bot. 2018, 69, 1387–1402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Hof, A.; Zechmann, B.; Schwammbach, D.; Huckelhoven, R.; Doehlemann, G. Alternative cell death mechanisms determine epidermal resistance in incompatible barley-Ustilago interactions. Mol. Plant Microbe Interact. 2014, 27, 403–414. [Google Scholar] [CrossRef]
  115. Wang, P.; Sun, X.; Wang, N.; Jia, X.; Ma, F. Ectopic expression of an autophagy-associated MdATG7b gene from apple alters growth and tolerance to nutrient stress in Arabidopsis thaliana. Plant Cell Tissue Organ 2017, 128, 9–23. [Google Scholar] [CrossRef]
  116. Yan, Y.; Wang, P.; He, C.; Shi, H. MeWRKY20 and its interacting and activating autophagy-related protein 8 (MeATG8) regulate plant disease resistance in cassava. Biochem. Biophys. Res. Commun. 2017, 494, 20–26. [Google Scholar] [CrossRef]
  117. Zeng, H.; Xie, Y.; Liu, G.; Lin, D.; He, C.; Shi, H. Molecular identification of GAPDHs in cassava highlights the antagonism of MeGAPCs and MeATG8s in plant disease resistance against cassava bacterial blight. Plant Mol. Biol. 2018, 97, 201–214. [Google Scholar] [CrossRef]
  118. 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 MaATG8 s in cell death and autophagy in immune response to Fusarium wilt. Plant Cell Rep. 2017, 36, 1237–1250. [Google Scholar] [CrossRef]
  119. 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]
  120. Pei, D.; Zhang, W.; Sun, H.; Wei, X.; Yue, J.; Wang, H. Identification of autophagy-related genes ATG4 and ATG8 from wheat (Triticum aestivum L.) and profiling of their expression patterns responding to biotic and abiotic stresses. Plant Cell Rep. 2014, 33, 1697–1710. [Google Scholar] [CrossRef]
  121. McLoughlin, F.; Augustine, R.; Marshall, R.; Li, F.; Kirkpatrick, L.; Otegui, M.; Vierstra, R. Maize multi-omics reveal roles for autophagic recycling in proteome remodelling and lipid turnover. Nat. Plants 2018, 4, 1056–1070. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Autophagy in plant life.
Figure 1. Autophagy in plant life.
Ijms 23 11410 g001
Table 1. Plants with identified ATG genes and potential processes that require autophagy.
Table 1. Plants with identified ATG genes and potential processes that require autophagy.
SpeciesBiological Process
Arabidopsis thalianaSeed development [15], root development [23], pollen development [33,34,35], abiotic stress [63,70], and biotic stress [94,100,101,102]
Capsicum annuumAbiotic stress [49]
Hordeum vulgareLeaf senescence [112], nutrient remobilization [27], microspore embryogenesis [113], abiotic stress [51], and biotic stress [114]
Manihot esculenta Biotic stress [115,116,117]
Musa acuminataCell death and immune response [118]
Nicotiana tabacumPollen maturation [37] and biotic stress [74,77,85].
Oryza sativaPollen maturation [36], nutrition stress [119], and leaf senescence [28]
Triticum aestivumNutrition stress [120] and biotic stress [96]
Zea maysNutrition stress [13] and lipid metabolism [121]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, S.; Hu, W.; Liu, F. Autophagy in the Lifetime of Plants: From Seed to Seed. Int. J. Mol. Sci. 2022, 23, 11410. https://doi.org/10.3390/ijms231911410

AMA Style

Wang S, Hu W, Liu F. Autophagy in the Lifetime of Plants: From Seed to Seed. International Journal of Molecular Sciences. 2022; 23(19):11410. https://doi.org/10.3390/ijms231911410

Chicago/Turabian Style

Wang, Song, Weiming Hu, and Fen Liu. 2022. "Autophagy in the Lifetime of Plants: From Seed to Seed" International Journal of Molecular Sciences 23, no. 19: 11410. https://doi.org/10.3390/ijms231911410

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