Plant peroxisomes host critical metabolic reactions and insulate the rest of the cell from reactive byproducts. The specialization of peroxisomal reactions is rooted in how the organelle modulates its proteome to be suitable for the tissue, environment, and developmental stage of the organism. The story of plant peroxisomal proteostasis begins with transcriptional regulation of peroxisomal protein genes and the synthesis, trafficking, import, and folding of peroxisomal proteins. The saga continues with assembly and disaggregation by chaperones and degradation via proteases or the proteasome. The story concludes with organelle recycling via autophagy. Some of these processes as well as the proteins that facilitate them are peroxisome-specific, while others are shared among organelles. Our understanding of translational regulation of plant peroxisomal protein transcripts and proteins necessary for pexophagy remain based in findings from other models. Recent strides to elucidate transcriptional control, membrane dynamics, protein trafficking, and conditions that induce peroxisome turnover have expanded our knowledge of plant peroxisomal proteostasis. Here we review our current understanding of the processes and proteins necessary for plant peroxisome proteostasis—the emergence, maintenance, and clearance of the peroxisomal proteome.

Harvesting energy, orchestrating development, and adapting to environmental changes are critical to organismal success—peroxisomes contribute to all three processes. Peroxisomes isolate β-oxidation and other oxidative reactions that produce toxic byproducts, including reactive oxygen and nitrogen species (ROS/RNS), and detoxifying these compounds is a conserved peroxisomal function. Peroxisomes are found in most eukaryotes and encompass a unique combination of features found in other organelles. Like several organelles defined by delimiting membranes and lacking DNA, peroxisomes can originate from the endoplasmic reticulum (ER) [1]. Like the nucleus, peroxisomes can import proteins without unfolding [2,3]. Like mitochondria and chloroplasts, peroxisomes can divide by fission [4]. Like multivesicular endosomes, peroxisomes can contain intralumenal vesicles derived from their outer membrane [5]. These sub-peroxisomal structures may operate in fatty acid metabolism [5], a fundamental peroxisomal function. Plants offer distinctive insights into peroxisomal fat processing, as fatty acid β-oxidation is exclusively peroxisomal in plants [6]. In contrast, metazoans split this task between peroxisomes and mitochondria.

Specific peroxisomal reactions vary based on organism, tissue, and developmental stage, which necessitates modification of the peroxisomal proteome to prioritize specialized reactions. Several plant hormones, including jasmonate and auxin, are partially synthesized in peroxisomes. Jasmonates are needed for pathogen responses and fertility (reviewed in [7,8]), and the auxin output from peroxisomes contributes to cotyledon expansion [9], root hair elongation [9], and lateral root development [10]. Peroxisomal catabolism of stored lipids fuels early seedling growth (reviewed in [11]). After seedlings are established, leaf peroxisomes house key steps in photorespiration, which improves photosynthetic efficiency by recycling glycolate to glycerate, allowing fixed carbon to reenter the Calvin cycle (reviewed in [12]). Peroxisome functions return to lipid catabolism during senescence, another transitional stage characterized by coordinated shifts in protein levels [13]. During senescence, peroxisomes also serve as signaling hubs of reactive species that coordinate with other metabolic organelles (reviewed in [14]).

Because plant peroxisomes carry out critical metabolism, peroxisome proteostasis is vital. Peroxisomal biogenesis and maintenance is facilitated by proteins termed peroxins (PEX proteins), which orchestrate biogenesis, division, and protein delivery to the lumen and membranes of the organelle. Peroxisomes are essential in plants; nearly all viable pex mutants are partial loss-of-function alleles, as complete loss of most Arabidopsis PEX proteins results in embryonic or gametophytic lethality [15–22]. Even partial peroxin dysfunction can confer serious plant growth defects [23–28].

Peroxisomal proteostasis contributes to peroxisomal specialization in response to developmental and environmental cues and counters the damaging effects of the byproducts of oxidative peroxisomal reactions. Here, we summarize our current understanding of peroxisomal proteostasis in plants and supplement these findings with insights from other models to bridge the gaps in plant studies.

Like many stories in biology, the narrative of peroxisomal proteins begins in the nucleus. Peroxisomal proteins are encoded by nuclear DNA; thus, initial regulation takes place outside of the organelle. Transcripts encoding proteins that decompose peroxisomal ROS (e.g., catalase (CAT), glutathione S-transferase, superoxide dismutase, and ascorbate peroxidase) are dynamically expressed in conditions that stimulate ROS production and/or promote peroxisome turnover, including pathogen and wound response, high light stresses, nutrient limitations, metal imbalances, drought stress, and senescence [29–43]. For example, the Arabidopsis ABI5 and GBF1 transcription factors activate CAT1 during seed germination and inhibit CAT2 during senescence and pathogen infection, respectively (Figure 1A) [29,44]. Moreover, transcriptional profiling of autophagy mutants and senescing Arabidopsis and maize leaves reveals enrichment in peroxisome-associated gene ontology terms [30,33,36], suggesting that peroxisomal protein levels and reactions are modulated during cellular remodeling.

Establishing the peroxisomal proteome

Figure 1
Establishing the peroxisomal proteome

(A) Nuclear encoded peroxisomal protein genes are governed by various transcriptional activators (mauve) and repressors (teal) and are modulated during development and by abiotic and biotic challenges that elicit individual-protein and whole-organelle changes. (B) As part of the endomembrane system, peroxisomes arise from the ER as pre-peroxisomes lacking lumenal proteins. (C) mRNA of some peroxisomal membrane proteins (PMPs) are likely translated by ER-localized ribosomes and the protein is trafficked through the ER to bud with pre-peroxisomes. Other PMPs are translated by cytosolic and potentially peroxisome-localized ribosomes. PEX19 acts as a chaperone, binding to the mPTS (membrane peroxisome targeting signal) of PMPs and transferring mPTS-cargo to PEX3 for membrane insertion. (D) Mature peroxisomes contain PMPs, intralumenal vesicles (ILVs) derived from the outer membrane, and lumenal proteins. PTS1 and PTS2 (peroxisome targeting signal 1 and 2) cargo (enzymes, chaperones, proteases) are imported into the lumen via receptors (PEX5 and PEX7) and the membrane docking complex (PEX13 and PEX14). After delivery of lumenal cargo, PEX5 is monoubiquitinated by peroxisome-associated ubiquitination machinery (the PEX4-PEX22 ubiquitin conjugating enzyme complex and the PEX2-PEX10-PEX12 ubiquitin ligase enzyme complex) and retrotranslocated for reuse by an ATPase complex (PEX1, PEX6, and PEX26).

Figure 1
Establishing the peroxisomal proteome

(A) Nuclear encoded peroxisomal protein genes are governed by various transcriptional activators (mauve) and repressors (teal) and are modulated during development and by abiotic and biotic challenges that elicit individual-protein and whole-organelle changes. (B) As part of the endomembrane system, peroxisomes arise from the ER as pre-peroxisomes lacking lumenal proteins. (C) mRNA of some peroxisomal membrane proteins (PMPs) are likely translated by ER-localized ribosomes and the protein is trafficked through the ER to bud with pre-peroxisomes. Other PMPs are translated by cytosolic and potentially peroxisome-localized ribosomes. PEX19 acts as a chaperone, binding to the mPTS (membrane peroxisome targeting signal) of PMPs and transferring mPTS-cargo to PEX3 for membrane insertion. (D) Mature peroxisomes contain PMPs, intralumenal vesicles (ILVs) derived from the outer membrane, and lumenal proteins. PTS1 and PTS2 (peroxisome targeting signal 1 and 2) cargo (enzymes, chaperones, proteases) are imported into the lumen via receptors (PEX5 and PEX7) and the membrane docking complex (PEX13 and PEX14). After delivery of lumenal cargo, PEX5 is monoubiquitinated by peroxisome-associated ubiquitination machinery (the PEX4-PEX22 ubiquitin conjugating enzyme complex and the PEX2-PEX10-PEX12 ubiquitin ligase enzyme complex) and retrotranslocated for reuse by an ATPase complex (PEX1, PEX6, and PEX26).

Close modal

In addition to transcripts encoding peroxisomal enzymes, transcripts encoding peroxins, which build peroxisomes, are also regulated. Wounding and Pseudomonas syringae infection induce Arabidopsis PEX1, PEX5, PEX10, and PEX14 transcripts (Figure 1A) [35], but the transcriptional regulators of these genes are not identified. PEX11 is implicated in peroxisome proliferation, and Arabidopsis PEX11 mRNA levels are dynamic. For example, PEX11A and PEX11B are induced by cadmium and high light, respectively (Figure 1A) [31,39,40]. Phytochrome A (PHYA) and the HYH transcription factor up-regulate PEX11B transcription, and phya and hyh mutants have fewer peroxisomes [31]. HYH directly binds the PEX11B promoter to promote peroxisome proliferation during seedling photomorphogenesis [31], and FHA3 negatively regulates PEX11B via HYH binding (Figure 1A) [45]. Further elucidation of transcriptional regulators specific to peroxisomal protein genes is needed to fully understand plant peroxisome proteostasis.

In addition to assessing transcriptional regulation following environmental challenges, analysis of transcript changes in mutants can elucidate the importance of specific proteins in peroxisomal processes. Dysfunctional peroxisomes can trigger retrograde signaling from the peroxisome to the nucleus. For example, catalase disruption in Arabidopsis alters expression of genes involved in abiotic and biotic stress response, plant growth regulation, and MAPK cascades [42], implicating ROS regulation in these transcriptional responses. Moreover, impaired peroxisomal protein import in a Caenorhabditis elegans pex mutant triggers induction of peroxisomal catalase and Lon protease genes by a ligand-activated transcription factor and a mediator of RNA polymerase II transcription [46]. Additional exploration of transcriptional changes in peroxisome-specific mutants is needed to fully understand this signaling in plants.

Protein delivery to the peroxisome is accomplished by PEX proteins that recognize Peroxisome Targeting Signals (PTSs) on peroxisomal cargo. Peroxisomal membrane proteins (PMPs), including most peroxins, are targeted through an mPTS (membrane PTS) and can be directly inserted into the peroxisome membrane or inserted into the ER before moving to nascent peroxisomes via ER budding. For example, a fluorescent protein tagged with the PEX26 mPTS localizes to peroxisome membranes in Arabidopsis seedlings, whereas the PEX22 mPTS localizes a reporter to both ER and peroxisome membranes [5]. Cytosolic PEX19 binds the mPTS and acts as a chaperone that accompanies the PMP cargo to PEX3, a PMP that assists with membrane insertion [47,48] (Figure 1C).

In addition to PEX3 and PEX19, Arabidopsis PEX16 assists in targeting PMPs that first insert into the ER [49]. Viable pex16 alleles show enlarged peroxisomes [50], and peroxisomes are enlarged or misshapen when PEX3, PEX16, or PEX19 levels are reduced via RNAi [51]. Exploration of PEX3 and PEX19 functions is complicated by gene duplications resulting in two isoforms of both peroxins in Arabidopsis [20,52]. Whether Arabidopsis pex3, pex16, or pex19 mutants display PMP sorting defects has not been examined.

mRNAs from nuclear genes that encode proteins destined for the ER, chloroplasts, and mitochondria can be translated by organelle-associated ribosomes— streamlining protein synthesis and organellar targeting (reviewed in [53]). For example, yeast PEX3 transcripts are found in ER-associated ribosomes [54]. Although proteins inserted directly into peroxisomes are typically thought to originate from cytosolic ribosomes, peroxisome-specific ribosomal profiling and single-molecule RNA fluorescence in situ hybridization in yeast reveals 11 peroxisomal protein transcripts, mostly PMPs, translated by peroxisome-proximal ribosomes [55]. Similarly, transcripts of several yeast peroxisome lumenal proteins are peroxisome-associated [54]. It will be interesting to learn whether localized translation is also associated with plant peroxisomes.

Peroxisomes are unusual among organelles in that they can import fully folded proteins [2,3]. Lumenal protein import relies on one of two peroxisome targeting signals that are recognized by cytosolic receptors (reviewed in [12]). PEX5 binds PTS1 cargo and PEX7 binds PTS2 cargo (Figure 1D). Like sequences targeting proteins to the ER, mitochondria, and chloroplasts, the nonapeptide PTS2 is near the N-terminus and is usually removed after import [56]. In contrast, the C-terminal PTS1 tripeptide [57] is not removed. Cargo-bound PEX5 and PEX7 complex with docking machinery, PEX13 and PEX14, at the peroxisomal membrane, resulting in lumenal delivery (Figure 1D). After cargo delivery, PEX5 is mono-ubiquitinated by dedicated ubiquitination machinery in the peroxisomal membrane to allow retrotranslocation from the membrane by peroxisome-anchored ATPases for further import rounds (Figure 1D).

The presence of PTS1 and PTS2 signals has enabled robust bioinformatic prediction of the peroxisomal proteome (reviewed in [58]), which is supported by proteomic analysis of purified peroxisomes from a variety of tissues (reviewed in [59]). In addition, some peroxisomal proteins are targeted to multiple locations (reviewed in [60]). For example, catalase, although predominantly peroxisomal, may also localize in the nucleus under certain conditions in Arabidopsis [61].

After proteins are incorporated into the organelle, peroxisomal proteostasis is managed by both lumenal and cytosolic machinery. Peroxisomes employ the cytosolic ubiquitin (Ub) 26S proteasome system to degrade ubiquitinated proteins to maintain peroxisomal health. The sequential action of Ub-activating enzyme (E1), Ub-conjugating enzymes (E2s), and Ub-protein ligases (E3s) covalently attach one or more ubiquitin moieties to target proteins (reviewed in [62]). Like yeast, plant peroxisomes have a dedicated E2, PEX4, that is tethered to the outside of the organelle by the PEX22 PMP [63,64]. Arabidopsis PEX4 is implicated in both PEX5 recycling and degradation [64,65] as well as degradation of a mutated PEX12 [23]. E2s work with E3s that provide substrate specificity, and many E3s are characterized by a RING domain. PEX2, PEX10, and PEX12 (Figure 1D) are RING-domain peroxins that have E3 activity in vitro [66]. These peroxins are implicated in both monoubiquitinating PEX5 for recycling (Figure 1D) and polyubiquitinating PEX5 for degradation by the cytosolic proteasome system (Figure 2) [23,66,67]. Furthermore, SUPPRESSOR OF PPI1 LOCUS1 (SP1), a RING-type E3 with well-documented chloroplast functions [68] may moonlight at the peroxisome to ubiquitinate the PEX13 PMP for proteasomal degradation [69], although this function is disputed [70].

Renovating the peroxisomal proteome

Figure 2
Renovating the peroxisomal proteome

(A) Intralumenal proteostasis involves proteases and chaperones. Proteases cleave the PTS2 (peroxisome targeting signal 2) peptide and degrade damaged and obsolete proteins; chaperones assist protein folding, complex assembly, and aggregate dissociation. Peroxisomal proteins with cytosolic domains (e.g., PEX5) can be polyubiquitinated by a ubiquitin (Ub)-conjugating enzyme (PEX4) and the Ub-protein ligase complex (PEX2, PEX10, and PEX12) at the peroxisome membrane. (B) Polyubiquitinated proteins are degraded by the cytosolic 26S proteosome.

Figure 2
Renovating the peroxisomal proteome

(A) Intralumenal proteostasis involves proteases and chaperones. Proteases cleave the PTS2 (peroxisome targeting signal 2) peptide and degrade damaged and obsolete proteins; chaperones assist protein folding, complex assembly, and aggregate dissociation. Peroxisomal proteins with cytosolic domains (e.g., PEX5) can be polyubiquitinated by a ubiquitin (Ub)-conjugating enzyme (PEX4) and the Ub-protein ligase complex (PEX2, PEX10, and PEX12) at the peroxisome membrane. (B) Polyubiquitinated proteins are degraded by the cytosolic 26S proteosome.

Close modal

Beyond PEX5 and PMPs, there is indirect evidence of ubiquitin system involvement in degrading lumenal proteins, which would require retrotranslocation of lumenal substrates out of the organelle to access the cytosolic ubiquitination machinery. For example, the lumenal proteins isocitrate lyase (ICL) and malate synthase (MLS) are stabilized in a pex4 pex22 mutant [64,71], and a screen for mutants that stabilize a GFP-ICL fusion [72] yielded pex2 and pex10 alleles [73].

In addition to cytosolic quality control, the peroxisome lumen houses chaperones and proteases that have been identified through various proteomic and bioinformatic studies (Table 1). Heat shock proteins (HSPs) are molecular chaperones that regulate protein folding, assembly, and disaggregation. Chaperones also can contribute to protein degradation, and mammalian Hsp70 facilitates substrate degradation via the ubiquitin-proteasome system and autophagy pathways [74,75]. An Hsp70 with a consensus PTS2 in watermelon (Citrullus vulgaris) cotyledons is similar to pea (Pisum sativum) and cucumber (Cucumis sativus) plastidic Hsp70s [76]. Interestingly, cucumber peroxisomes house two Hsp70s (71 and 78 kDa) and an Hsp40 [77]. Cucumber Hsp40 is a PMP that in the ADP-bound conformation specifically binds Hsp70-1 (71 kDa), but not the heavier Hsp70-2, [77], perhaps assisting in folding and import of peroxisomal proteins.

Table 1
Predicted or confirmed plant peroxisomal proteases and chaperones
IdentifierNameAliasClassPTSLocalization1
At1g28320 DEGRADATION OF PERIPLASMIC PROTEINS15 DEG15 PTS2-processing protease PTS1 Microscopy [91
At4g12910 SERINE CARBOXYPEPTIDASE-LIKE20 SCPL20 Serine carboxypeptidase PTS1 Microscopy [81
At4g36880 RD21A-LIKE PROTEASE1 RDL1 Papain-like Cys proteinase PTS1 Microscopy [81
At5g47040 LON PROTEASE2 LON2 Lon protease homolog PTS1 Microscopy [82
At2g41790 Peroxisomal M16 protease PXM16 Zinc-metallopeptidase PTS1 Proteomics [83
At2g18080 EMBRYO SAC DEVELOPMENT ARREST2 EDA2 Serine carboxypeptidase PTS1 Predicted [57
At4g20310 SITE 2 PROTEASE S2P Peptidase M50 family protein PTS2 Predicted [57
At1g06460 ALPHA-CRYSTALLIN DOMAIN 31.2 ACD31.2 Hsp20/α-crystallin domain PTS2 Microscopy [80
At5g12020 17.6 kDa Class II heat shock protein HSP17.C6II Hsp20/α-crystallin domain non-canonical PTS1 Microscopy [79
At5g37670 15.7 kDa heat shock protein HSP15.7 Hsp20/α-crystallin domain PTS1 Microscopy [80
Cucumis sativus Q04960 DnaJ peptide-binding protein HSP40 J-domain-containing protein Proteomics [135
Citrullus lanatus U92815 Heat shock protein 70 HSP70 70-kDa heat shock protein PTS2 Proteomics [76
Cucumis sativus AJ249330 Heat shock protein 70 HSP70-1 70-kDa heat shock protein Proteomics [77
Cucumis sativus AJ249331 Heat shock protein 70 HSP70-2 70-kDa heat shock protein Proteomics [77
IdentifierNameAliasClassPTSLocalization1
At1g28320 DEGRADATION OF PERIPLASMIC PROTEINS15 DEG15 PTS2-processing protease PTS1 Microscopy [91
At4g12910 SERINE CARBOXYPEPTIDASE-LIKE20 SCPL20 Serine carboxypeptidase PTS1 Microscopy [81
At4g36880 RD21A-LIKE PROTEASE1 RDL1 Papain-like Cys proteinase PTS1 Microscopy [81
At5g47040 LON PROTEASE2 LON2 Lon protease homolog PTS1 Microscopy [82
At2g41790 Peroxisomal M16 protease PXM16 Zinc-metallopeptidase PTS1 Proteomics [83
At2g18080 EMBRYO SAC DEVELOPMENT ARREST2 EDA2 Serine carboxypeptidase PTS1 Predicted [57
At4g20310 SITE 2 PROTEASE S2P Peptidase M50 family protein PTS2 Predicted [57
At1g06460 ALPHA-CRYSTALLIN DOMAIN 31.2 ACD31.2 Hsp20/α-crystallin domain PTS2 Microscopy [80
At5g12020 17.6 kDa Class II heat shock protein HSP17.C6II Hsp20/α-crystallin domain non-canonical PTS1 Microscopy [79
At5g37670 15.7 kDa heat shock protein HSP15.7 Hsp20/α-crystallin domain PTS1 Microscopy [80
Cucumis sativus Q04960 DnaJ peptide-binding protein HSP40 J-domain-containing protein Proteomics [135
Citrullus lanatus U92815 Heat shock protein 70 HSP70 70-kDa heat shock protein PTS2 Proteomics [76
Cucumis sativus AJ249330 Heat shock protein 70 HSP70-1 70-kDa heat shock protein Proteomics [77
Cucumis sativus AJ249331 Heat shock protein 70 HSP70-2 70-kDa heat shock protein Proteomics [77
1

Localizations are based on the presence of a PTS (predicted), from peroxisome proteomic studies (proteomics), or confirmed using fusion to fluorescent reporters (microscopy).

Although apparently lacking the peroxisomal Hsp70s found in crop plants, Arabidopsis peroxisomes house several smaller heat shock proteins in the Hsp20/α-crystallin superfamily (Table 1). Hsp15.7 and Hsp17.6CII are PTS1-targeted [78–81], and ALPHA-CRYSTALLIN DOMAIN 31.2 (ACD31.2) is PTS2-targeted [80]. Both Hsp15.7 and ACD31.2 reduce non-specific protein aggregation when expressed in yeast [80]. Hsp15.7 transcripts are induced by heat stress and when catalase is inhibited [80]. Hsp17.6CII prevents CAT2 aggregation, increases CAT2 activity, and increases tolerance to abiotic stresses (Figure 2A) [79].

The nexus of protein quality control is the balance between repair and destruction—folding, assembly, or disaggregation versus breakdown for recycling. This balance is typified by the Lon family of ATP-dependent proteases, originally described in bacteria, which have both chaperone and protease functions. Arabidopsis encodes four LON isoforms; LON2 (Figure 2A) contains a PTS1 and is the only peroxisomal isoform [82,83]. LON proteases contain an N-terminal substrate-recognition domain, a central ATPase associated with various cellular activities (AAA) domain, and a C-terminal proteolytic domain (Figure 2A) [84,85]. LON N-terminal and ATPase domains together unfold misfolded proteins, contributing chaperone activity [85,86], and yeast and mammalian LONs exhibit chaperone-like activity independent of protease activity [87]. Arabidopsis lon2 mutants display sparse, enlarged peroxisomes [88–90]. Although peroxisomal proteins are not stabilized in a lon2 mutant [89], peroxisomal MLS and ICL (glyoxylate cycle enzymes) are synergistically stabilized in lon2 mutants when autophagy is disabled [88,90], implicating MLS and ICL as LON2 protease substrates. Though LON2 chaperone clients are not identified in Arabidopsis, LON2 chaperone activity is necessary to prevent excessive peroxisome turnover [90].

A second protein family implicated in both proteolytic and chaperone activities is the Deg peptidase subfamily S1B proteases. Citrullus vulgaris GLYOXYSOMAL PROCESSING PROTEASE (GPP) and the Arabidopsis ortholog DEGRADATION OF PERIPLASMIC PROTEINS 15 (DEG15) are ATP-independent serine endopeptidases that remove the N-terminal region of PTS2 proteins in vivo [91]. In vitro, DEG15 interacts with different substrates as a monomer and dimer [92]. In the calcium-depleted monomeric state, DEG15 degrades denatured proteins. However, when bound to the peroxisomal calcium-dependent Calmodulin-like protein CML3, DEG15 specifically cleaves PTS2 proteins (Figure 2A) [92,93]. DEG15 PTS2 substrates include malate dehydrogenase, citrate synthase, acyl-CoA oxidase, and 3-ketoacyl-CoA thiolase [91,92]. PTS2 cleavage is the only DEG15 activity that is validated in vivo; deg15 mutants fail to process PTS2 proteins, but reported lumenal protein levels are unaltered [91,92]. Despite failing to process PTS2 proteins, deg15 mutants have only mild physiological defects [89], indicating that PTS2-containing enzymes are at least partially functional when uncleaved. However, the deg15 lon2 double mutant displays synergistic defects [89], hinting that DEG15 might have roles beyond PTS2 processing. Related enzymes in bacteria also show in vivo chaperone activity [94–96], but whether DEG15 has proteolytic or chaperone substrates beyond PTS2 proteins remains unknown.

Regulating individual proteins allows peroxisomes to tune their contents and reactions; however, plants also use more drastic measures to orchestrate large-scale organellar remodeling. The peroxisome population can be decreased via pexophagy, a selective form of autophagy, which allows for comprehensive shifts in the peroxisomal proteome (reviewed in [97]).

During pexophagy, the entire organelle, including lumenal and membrane contents, is targeted by the autophagy machinery for disassembly in the vacuole (Figure 3), allowing the molecular components to be reused by the cell. Autophagy requires coordination of numerous autophagy-related (ATG) proteins [98]. Although the first Arabidopsis ATG genes were identified by homology in 2002 [99], pexophagy was not reported in plants until over a decade later [88,100,101]. Analysis of atg null mutants, which are unable to carry out either general or cargo-specific autophagy, underpins much of our knowledge of pexophagy in plants.

Dismantling the peroxisomal proteome

Figure 3
Dismantling the peroxisomal proteome

(A) Peroxisomes are targeted for pexophagy by recruitment of autophagy machinery, including a double-bilayer isolation membrane marked by ATG8 (violet circles). Recruitment could occur through direct interaction of peroxisomal proteins with ATG8 or via an intermediary receptor (purple ovals) that recognizes peroxisomal proteins or their ubiquitinated (Ub) derivates. Illustrated targeting mechanisms are based on receptor-mediated partners observed in non-plant systems or demonstrated ATG8 interactions in plants; none of the interactions depicted have been shown as necessary or sufficient to induce pexophagy in plants. (B) Following recruitment of the isolation membrane, cargo is fully engulfed to form an autophagosome. Contents are delivered through fusion of the outer autophagosomal membrane with the vacuolar membrane. In the vacuole, resident proteases and lipases degrade the peroxisomal proteins and lipids to allow nutrient recycling.

Figure 3
Dismantling the peroxisomal proteome

(A) Peroxisomes are targeted for pexophagy by recruitment of autophagy machinery, including a double-bilayer isolation membrane marked by ATG8 (violet circles). Recruitment could occur through direct interaction of peroxisomal proteins with ATG8 or via an intermediary receptor (purple ovals) that recognizes peroxisomal proteins or their ubiquitinated (Ub) derivates. Illustrated targeting mechanisms are based on receptor-mediated partners observed in non-plant systems or demonstrated ATG8 interactions in plants; none of the interactions depicted have been shown as necessary or sufficient to induce pexophagy in plants. (B) Following recruitment of the isolation membrane, cargo is fully engulfed to form an autophagosome. Contents are delivered through fusion of the outer autophagosomal membrane with the vacuolar membrane. In the vacuole, resident proteases and lipases degrade the peroxisomal proteins and lipids to allow nutrient recycling.

Close modal

Pexophagy occurs even under optimal growth conditions, as indicated by elevated peroxisome abundance and protein levels in atg mutant seedlings [100,101]. Interestingly, proteomic analysis of Arabidopsis atg mutants reveals that peroxisomal proteins are more stabilized than proteins from other organelles under normal growth conditions [102], perhaps reflecting the damaging oxidative environment of the peroxisome compared to other organelles. Indeed, inactive catalase is found in aggregated peroxisomes in Arabidopsis atg seedlings, signifying that these peroxisomes are oxidized [100]. Moreover, stomatal opening defects are associated with decreased catalase activity and increased ROS in guard cells of atg2 mutants [103]. Additionally, maize atg mutants under stress accumulate peroxisomal β-oxidation enzymes, but do not process additional lipids, suggesting that the accumulating peroxisomes may be inactive [36].

In addition to quality control, protein turnover via pexophagy supports developmental transitions in peroxisomal reactions. For example, young pumpkin seedlings are populated by numerous β-oxidative peroxisomes processing stored fats, but with the onset of photosynthesis, leaf tissue gains peroxisomes that support photorespiration [104]. Both thiolase, a β-oxidation enzyme, and MLS, a glyoxylate cycle enzyme, are stabilized in Arabidopsis atg mutant seedlings [88], suggesting that pexophagy plays a role in this transition.

The LON2 peroxisome-localized protease plays a regulatory role in pexophagy [88]. Intriguingly, excessive pexophagy underlies the physiological and molecular defects in lon2 null mutants. Arabidopsis lon2 suppressors include dozens of mutations spanning six ATG genes, which completely restore peroxisomal function and morphology [88,105]. A transgene expressing a protease-disabled LON2 suppresses the excessive pexophagy in lon2, while a version disabling the chaperone domain does not [90]. Thus, it appears that the chaperone function of LON2, rather than the peptidase function, restricts pexophagy in wild-type plants. The specific LON2 clients that promote pexophagy when LON2 is disabled remain to be identified.

Other proteins have been implicated in plant pexophagy. The actin-related proteins 2/3 (ARP2/3) complex, which nucleates actin polymerization, appears to localize to subperoxisomal surface puncta that occassionally co-localize with ATG8 in Arabidopsis [106]. Peroxisomes are metabolically functional, more abundant, and enlarged in ARP2/3 complex mutants [106], hinting at a pexophagy-related role for the complex, and by extension, actin.

Beyond LON2 and ARP2/3, our knowledge of proteins specifically involved in plant pexophagy is sparse; receptors linking peroxisomes to the autophagy machinery remain elusive. Although several pexophagy receptors are identified in yeast [107,108], these proteins lack apparent homologs in plants. In the case of the mammalian pexophagy receptor NBR1 [109], the plant homolog serves as an autophagy receptor for polyubiquitinated aggregates [110] and exocyst-positive organelles [111]. However, a role for NBR1 in mediating the excessive pexophagy observed in lon2 mutants has been ruled out [105].

Like the pexophagy receptors, the peroxisomal proteins recognized by these receptors remain to be identified in plants. Several peroxins in the membrane are implicated in pexophagy in other systems, including PEX3 [108,112] and PEX14 [113] (Figure 3A). Moreover, the RING peroxins are positioned to ubiquitinate PMPs, which can serve as an autophagy trigger. In mammalian cells, starvation increases PEX2 expression, which promotes pexophagy via ubiquitination of peroxisomal proteins [114]. PEX5 is a pexophagy receptor ligand in ubiquitin-mediated mammalian pexophagy models (reviewed in [115,116]) (Figure 3A). Moreover, perturbing the AAA ATPase retrotranslocation complex (PEX1, PEX6, and PEX26) increases pexophagy in human cells, and this increase can be eliminated by reducing PEX5 levels [117,118], suggesting that the AAA ATPase prevents pexophagy by preventing ubiquitinated PEX5 accumulation in the membrane.

Peroxin roles in regulating pexophagy might be conserved in plants, though direct evidence remains elusive. The peroxisome-defective phenotypes of Arabidopsis pex1 and pex6 mutants are partially rescued by preventing autophagy [21,119]. This rescue suggests that, as in mammals [117,118], impeding the plant AAA ATPase results in autophagic targeting of peroxisomes that are at least partially functional.

One approach to discover proteins mediating pexophagy is to identify peroxisomal proteins that interact with autophagy proteins. ATG8 is a lipidated ubiquitin-like protein that marks the isolation membrane that envelops autophagy cargo to form autophagosomes [120,121]. Selective-autophagy receptors often facilitate cargo engulfment by binding ATG8 [122] via either an ATG8-interacting motif (AIM) or a ubiquitin-interacting motif (UIM)-like sequence [123]. These motifs bind at an LDS (LIR/AIM docking site) or UDS (UIM-like docking site) on ATG8 [123]. Bioinformatic analysis reveals nine Arabidopsis peroxisomal proteins with predicted AIMs, including PEX6 and PEX10, which also interact with ATG8f in biomolecular fluorescent complementation assays [124]. Moreover, PEX10 interacts with the ATG8e LDS in yeast two-hybrid and dot blot binding assays [123]. Whether PEX10 association with autophagy machinery is necessary or sufficient for pexophagy in vivo remains unexplored.

While precise mechanisms directing pexophagy remain elusive, identifying conditions in which pexophagy occurs can provide new avenues to study the process. Peroxisomal oxidation impacts pexophagy in multiple systems [100,125,126]. In human cells, a ROS-activated kinase promotes ROS-dependent PEX5 ubiquitination, permitting recognition by an autophagy receptor [125]. The identification of specific plant proteins mediating pexophagy will assist in determining whether the general oxidative state of the peroxisome, the state and configuration of a specific protein, or both, triggers pexophagy in plants. Other stress conditions might provide avenues to study peroxisomal proteome and population changes. For example, an ephemeral peroxisome population is induced by concerted proliferation and pexophagy during cadmium stress [127,128]. Additionally, salt induces peroxisome proliferation [129,130] and autophagy [131], but it remains unknown if selective pexophagy occurs during salt stress. It will be interesting to monitor the impact on the peroxisomal proteome and plant health as new conditions and mutants that impact the peroxisome population are identified.

Plant peroxisomal proteostasis requires coordination of processes from transcription to whole-organelle degradation, yet many nuclear factors that regulate these processes and protein mediators remain unidentified. Deployment of transcriptomics and proteomics using various mutants with peroxisome and autophagy dysfunction may help elucidate the interconnected processes impacting peroxisomal proteostasis. In addition, comparing data from different time points and developmental stages will illuminate the interplay of transcriptional regulation, protein function, inter-organelle communication, and responses. Such high-content datasets will provide a foundation for informative modeling studies and may also help identify substrates of the various peroxisomal proteases (Table 1), including pexophagy regulators.

The recent demonstration of peroxisome-associated translation in yeast [54,55] highlights the gaps in our knowledge of peroxisomal protein targeting and delivery in plants. Though de novo formation from the ER is assumed, early steps in plant peroxisome biogenesis remains largely unexplored. The role of ER trafficking in providing PMPs to budding pre-peroxisomes and pre-existing peroxisome has not been determined in plants. The gatekeeping mechanisms for membrane proteins at ER exit sites and the identity of proteins involved in addition to early-acting peroxins remains largely unknown for any organism. Interestingly, the roles for biogenesis peroxins PEX19 and PEX3 have been expanded in mammals to include ER-targeting of lipid droplet proteins, hinting at connected biogenesis [132]. Should these organelles share exit sites, findings on lipid droplet formation could provide new targets for peroxisome biogenesis and ER trafficking analysis.

The recent description of intralumenal vesicles in peroxisomes [5] prompts many questions about peroxisomal protein targeting, including possible use of these vesicles to compartmentalize PMPs, lumenal proteins, and peroxisomal metabolism. Future experiments may reveal whether PEX5, PEX11, and PEX14 occupy sub-peroxisomal domains in plants as observed in human cells [133], whether additional peroxins sort differentially, and whether this sub-organellar localization or ARP2/3 sub-domains on the peroxisomal surface [106] involves intralumenal vesicles [5].

Future experiments will undoubtedly leverage advances in CRISPR systems for targeted gene disruption and modification in plants [134]. This approach will allow studies of proteins that have previously been limited by the lack of viable mutants or the presence of duplicated genes. Moreover, CRISPR experiments allow targeting of specific functional domains. For instance, disrupting predicted AIM sequences could identify plant peroxins necessary for pexophagy. CRISPR studies will also facilitate expansion of plant peroxisome studies beyond Arabidopsis, where much of our current knowledge is based, to include other plants, including crop species.

The fluidity and adaptability of peroxisome function across development, tissues, and environments is a key attribute of the organelle. The observed versatility inspires and necessitates further exploration of how peroxisomes regulate their membrane and lumenal protein populations. Illuminating peroxisomal proteostasis remains a dynamic frontier in plant biology.

  • Peroxisomes alter their proteome in profound ways to support plant vitality based on tissue, condition, and developmental stage.

  • Recent discoveries suggest that peroxisomes share more attributes with other organelles than previously appreciated, such as protein sorting, membrane complexity, and localized translation.

  • Despite important contributions to our understanding of peroxisomal proteostasis, high-throughput analyses are a largely untapped reservoir for future studies elucidating additional facets of peroxisomal proteostasis.

The authors declare that there are no competing interests associated with the manuscript.

Peroxisome research in the authors’ lab is supported by the National Institutes of Health [grant number R35GM130338 (to B.B.)]; the National Science Foundation [grant number 1907069 (to D.M.)]; and the Welch Foundation [grant number C-1309 (to B.B.)].

D.M., K.S., and B.B. wrote and edited the article.

We thank Gabrielle Buck, Makaela Jackson, Isabella Kreko, Roxanna Llinas, Dereth Phillips, Ana Swearingen, Melissa Traver, Andrew Woodward, and Zachary Wright for critical comments on the manuscript.

AAA

ATPase associated with various cellular activities

ABI5

ABA (abscisic acid) insensitive 5

AIM

ATG8-interacting motif

ATG

autophagy-related

CAT

catalase

CML

calmodulin-like

CRISPR

clustered regularly interspaced short palindromic repeats

ER

endoplasmic reticulum

GBF1

G-BOX binding factor 1

HSP

heat shock protein

HYH

HY5-homolog

ICL

isocitrate lyase

ILV

intralumenal vesicle

LDS

LIR/AIM-docking site

LIR

LC3-interacting region (LC3 is a mammalian ATG8 ortholog)

MAPK

mitogen-activated protein kinase

MLS

malate synthase

PEX

peroxin

PHY

phytochrome

PMP

peroxisomal membrane protein

PTS

peroxisome-targeting signal

RING

really interesting new gene

ROS

reactive oxygen species

UDS

UIM-like docking site

UIM

ubiquitin-interacting motif

1.
Hoepfner
D.
,
Schildknegt
D.
,
Braakman
I.
,
Philippsen
P.
and
Tabak
H.F.
(
2005
)
Contribution of the endoplasmic reticulum to peroxisome formation
.
Cell
122
,
85
95
[PubMed]
2.
Lee
M.S.
,
Mullen
R.T.
and
Trelease
R.N.
(
1997
)
Oilseed isocitrate lyases lacking their essential type 1 peroxisomal targeting signal are piggybacked to glyoxysomes
.
Plant Cell
9
,
185
197
[PubMed]
3.
McNew
J.A.
and
Goodman
J.M.
(
1994
)
An oligomeric protein is imported into peroxisomes in vivo
.
J. Cell Biol.
127
,
1245
1257
[PubMed]
4.
Schrader
M.
,
Bonekamp
N.A.
and
Islinger
M.
(
2012
)
Fission and proliferation of peroxisomes
.
Metab. Funct. Biog. Peroxisomes Health Dis.
1822
,
1343
1357
5.
Wright
Z.J.
and
Bartel
B.
(
2020
)
Peroxisomes form intralumenal vesicles with roles in fatty acid catabolism and protein compartmentalization in Arabidopsis
.
Nat. Commun.
11
,
6221
[PubMed]
6.
Kunau
W.-H.
,
Bühne
S.
,
de la Garza
M.
,
Kionka
C.
,
Mateblowski
M.
,
Schultz-Borchard
U.
et al.
(
1988
)
Comparative enzymology of β-oxidation
.
Biochem. Soc. Trans.
16
,
418
420
[PubMed]
7.
Huang
H.
,
Liu
B.
,
Liu
L.
and
Song
S.
(
2017
)
Jasmonate action in plant growth and development
.
J. Exp. Bot.
68
,
1349
1359
[PubMed]
8.
Pan
R.
,
Liu
J.
and
Hu
J.
(
2019
)
Peroxisomes in plant reproduction and seed-related development
.
J. Integr. Plant Biol.
61
,
784
802
[PubMed]
9.
Strader
L.C.
,
Culler
A.H.
,
Cohen
J.D.
and
Bartel
B.
(
2010
)
Conversion of endogenous indole-3-butyric acid to indole-3-acetic acid drives cell expansion in Arabidopsis seedlings
.
Plant Physiol.
153
,
1577
1586
[PubMed]
10.
De Rybel
B.
,
Audenaert
D.
,
Xuan
W.
et al.
(
2012
)
A role for the root cap in root branching revealed by the non-auxin probe naxillin
.
Nat. Chem. Biol.
8
,
798
805
[PubMed]
11.
Theodoulou
F.L.
and
Eastmond
P.J.
(
2012
)
Seed storage oil catabolism: a story of give and take
.
Curr. Opin. Plant Biol.
15
,
322
328
[PubMed]
12.
Kao
Y.-T.
,
Gonzalez
K.L.
and
Bartel
B.
(
2018
)
Peroxisome function, biogenesis, and dynamics in plants
.
Plant Physiol.
176
,
162
177
[PubMed]
13.
Nishimura
M.
,
Takeuchi
Y.
,
De Bellis
L.
and
Hara-Nishimura
I.
(
1993
)
Leaf peroxisomes are directly transformed to glyoxysomes during senescence of pumpkin cotyledons
.
Protoplasma
175
,
131
137
14.
Zentgraf
U.
,
Andrade-Galan
A.G.
and
Bieker
S.
(
2022
)
Specificity of H2O2 signaling in leaf senescence: is the ratio of H2O2 contents in different cellular compartments sensed in Arabidopsis plants?
Cell. Mol. Biol. Lett.
27
,
4
[PubMed]
15.
Boisson-Dernier
A.
,
Frietsch
S.
,
Kim
T.-H.
,
Dizon
M.B.
and
Schroeder
J.I.
(
2008
)
The peroxin loss-of-function mutation abstinence by mutual consent disrupts male-female gametophyte recognition
.
Curr. Biol.
18
,
63
68
[PubMed]
16.
Fan
J.
,
Quan
S.
,
Orth
T.
,
Awai
C.
,
Chory
J.
and
Hu
J.
(
2005
)
The Arabidopsis PEX12 gene is required for peroxisome biogenesis and is essential for development
.
Plant Physiol.
139
,
231
239
[PubMed]
17.
Goto
S.
,
Mano
S.
,
Nakamori
C.
and
Nishimura
M.
(
2011
)
Arabidopsis ABERRANT PEROXISOME MORPHOLOGY9 is a peroxin that recruits the PEX1-PEX6 complex to peroxisomes
.
Plant Cell
23
,
1573
1587
[PubMed]
18.
Hu
J.
,
Aguirre
M.
,
Peto
C.
,
Alonso
J.
,
Ecker
J.
and
Chory
J.
(
2002
)
A role for peroxisomes in photomorphogenesis and development of Arabidopsis
.
Science
19.
Lin
Y.
,
Sun
L.
,
Nguyen
L.V.
,
Rachubinski
R.A.
and
Goodman
H.M.
(
1999
)
The Pex16p homolog SSE1 and storage organelle formation in Arabidopsis seeds
.
Science
284
,
328
330
[PubMed]
20.
McDonnell
M.M.
,
Burkhart
S.E.
,
Stoddard
J.M.
,
Wright
Z.J.
,
Strader
L.C.
and
Bartel
B.
(
2016
)
The early-acting peroxin PEX19 is redundantly encoded, farnesylated, and essential for viability in Arabidopsis thaliana
.
PLoS ONE
11
,
e0148335
[PubMed]
21.
Rinaldi
M.A.
,
Fleming
W.A.
,
Gonzalez
K.L.
,
Park
J.
,
Ventura
M.J.
,
Patel
A.B.
et al.
(
2017
)
The PEX1 ATPase stabilizes PEX6 and plays essential roles in peroxisome biology
.
Plant Physiol.
174
,
2231
2247
[PubMed]
22.
Sparkes
I.A.
,
Brandizzi
F.
,
Slocombe
S.P.
,
El-Shami
M.
,
Hawes
C.
and
Baker
A.
(
2003
)
An Arabidopsis pex10 null mutant is embryo lethal, implicating peroxisomes in an essential role during plant embryogenesis
.
Plant Physiol.
133
,
1809
1819
[PubMed]
23.
Kao
Y.-T.
,
Fleming
W.A.
,
Ventura
M.J.
and
Bartel
B.
(
2016
)
Genetic interactions between PEROXIN12 and other peroxisome-associated ubiquitination components
.
Plant Physiol.
172
,
1643
1656
[PubMed]
24.
Khan
B.R.
and
Zolman
B.K.
(
2010
)
pex5 mutants that differentially disrupt PTS1 and PTS2 peroxisomal matrix protein import in Arabidopsis
.
Plant Physiol.
154
,
1602
1615
[PubMed]
25.
Woodward
A.W.
,
Fleming
W.A.
,
Burkhart
S.E.
,
Ratzel
S.E.
,
Bjornson
M.
and
Bartel
B.
(
2014
)
A viable Arabidopsis pex13 missense allele confers severe peroxisomal defects and decreases PEX5 association with peroxisomes
.
Plant Mol. Biol.
86
,
201
214
[PubMed]
26.
Zolman
B.K.
and
Bartel
B.
(
2004
)
An Arabidopsis indole-3-butyric acid-response mutant defective in PEROXIN6, an apparent ATPase implicated in peroxisomal function
.
Proc. Natl. Acad. Sci.
101
,
1786
1791
27.
Gonzalez
K.L.
,
Fleming
W.A.
,
Kao
Y.-T.
,
Wright
Z.J.
,
Venkova
S.V.
,
Ventura
M.J.
et al.
(
2017
)
Disparate peroxisome-related defects in Arabidopsis pex6 and pex26 mutants link peroxisomal retrotranslocation and oil body utilization
.
Plant J.
92
,
110
128
[PubMed]
28.
Mano
S.
,
Nakamori
C.
,
Nito
K.
,
Kondo
M.
and
Nishimura
M.
(
2006
)
The Arabidopsis pex12 and pex13 mutants are defective in both PTS1- and PTS2-dependent protein transport to peroxisomes
.
Plant J.
47
,
604
618
[PubMed]
29.
Bi
C.
,
Ma
Y.
,
Wu
Z.
,
Yu
Y.-T.
,
Liang
S.
,
Lu
K.
et al.
(
2017
)
Arabidopsis ABI5 plays a role in regulating ROS homeostasis by activating CATALASE 1 transcription in seed germination
.
Plant Mol. Biol.
94
,
197
213
[PubMed]
30.
Breeze
E.
,
Harrison
E.
,
McHattie
S.
et al.
(
2011
)
High-resolution temporal profiling of transcripts during Arabidopsis leaf senescence reveals a distinct chronology of processes and regulation
.
Plant Cell
23
,
873
894
[PubMed]
31.
Desai
M.
and
Hu
J.
(
2008
)
Light induces peroxisome proliferation in Arabidopsis seedlings through the photoreceptor Phytochrome A, the transcription factor HY5 homolog, and the peroxisomal protein PEROXIN11b
.
Plant Physiol.
146
,
1117
1127
[PubMed]
32.
Dinneny José
R.
,
Long Terri
A.
,
Wang Jean
Y.
,
Jung Jee
W.
,
Daniel
Mace
,
Solomon
Pointer
et al.
(
2008
)
Cell identity mediates the response of Arabidopsis roots to abiotic stress
.
Science
320
,
942
945
[PubMed]
33.
Li
F.
,
Chung
T.
,
Pennington
J.G.
,
Federico
M.L.
,
Kaeppler
H.F.
,
Kaeppler
S.M.
et al.
(
2015
)
Autophagic recycling plays a central role in maize nitrogen remobilization
.
Plant Cell
27
,
1389
1408
[PubMed]
34.
Long
T.A.
,
Tsukagoshi
H.
,
Busch
W.
,
Lahner
B.
,
Salt
D.E.
and
Benfey
P.N.
(
2010
)
The bHLH transcription factor POPEYE regulates response to iron deficiency in Arabidopsis roots
.
Plant Cell
22
,
2219
2236
[PubMed]
35.
Lopez-Huertas
E.
,
Charlton
W.L.
,
Johnson
B.
,
Graham
I.A.
and
Baker
A.
(
2000
)
Stress induces peroxisome biogenesis genes
.
EMBO J.
19
,
6770
6777
[PubMed]
36.
McLoughlin
F.
,
Augustine
R.C.
,
Marshall
R.S.
,
Li
F.
,
Kirkpatrick
L.D.
,
Otegui
M.S.
et al.
(
2018
)
Maize multi-omics reveal roles for autophagic recycling in proteome remodeling and lipid turnover
.
Nat. Plants
4
,
1056
1070
[PubMed]
37.
McLoughlin
F.
,
Marshall
R.S.
,
Ding
X.
,
Chatt
E.C.
,
Kirkpatrick
L.D.
,
Augustine
R.C.
et al.
(
2020
)
Autophagy plays prominent roles in amino acid, nucleotide, and carbohydrate metabolism during fixed-carbon starvation in maize
.
Plant Cell
32
,
2699
2724
[PubMed]
38.
Queval
G.
,
Issakidis-Bourguet
E.
,
Hoeberichts
F.A.
,
Vandorpe
M.
,
Gakière
B.
,
Vanacker
H.
et al.
(
2007
)
Conditional oxidative stress responses in the Arabidopsis photorespiratory mutant cat2 demonstrate that redox state is a key modulator of daylength-dependent gene expression, and define photoperiod as a crucial factor in the regulation of H2O2-induced cell death
.
Plant J.
52
,
640
657
[PubMed]
39.
Rodríguez-Serrano
M.
,
Romero-Puertas
M.C.
,
Sparkes
I.
,
Hawes
C.
,
del Río
L.A.
and
Sandalio
L.M.
(
2009
)
Peroxisome dynamics in Arabidopsis plants under oxidative stress induced by cadmium
.
Free Radic. Biol. Med.
47
,
1632
1639
[PubMed]
40.
Rodríguez-Serrano
M.
,
Romero-Puertas
M.C.
,
Sanz-Fernández
M.
,
Hu
J.
and
Sandalio
L.M.
(
2016
)
Peroxisomes extend peroxules in a fast response to stress via a reactive oxygen species-mediated induction of the peroxin PEX11a
.
Plant Physiol.
171
,
1665
1674
[PubMed]
41.
Sandalio
L.M.
,
Peláez-Vico
M.A.
,
Molina-Moya
E.
and
Romero-Puertas
M.C.
(
2021
)
Peroxisomes as redox-signaling nodes in intracellular communication and stress responses
.
Plant Physiol.
186
,
22
35
[PubMed]
42.
Su
T.
,
Wang
P.
,
Li
H.
et al.
(
2018
)
The Arabidopsis catalase triple mutant reveals important roles of catalases and peroxisome-derived signaling in plant development
.
J. Integr. Plant Biol.
60
,
591
607
[PubMed]
43.
Ebeed
H.T.
,
Stevenson
S.R.
,
Cuming
A.C.
and
Baker
A.
(
2018
)
Conserved and differential transcriptional responses of peroxisome associated pathways to drought, dehydration and ABA
.
J. Exp. Bot.
69
,
4971
4985
[PubMed]
44.
Giri
M.K.
,
Singh
N.
,
Banday
Z.Z.
,
Singh
V.
,
Ram
H.
,
Singh
D.
et al.
(
2017
)
GBF1 differentially regulates CAT2 and PAD4 transcription to promote pathogen defense in Arabidopsis thaliana
.
Plant J.
91
,
802
815
[PubMed]
45.
Desai
M.
,
Pan
R.
and
Hu
J.
(
2017
)
Arabidopsis forkhead-associated domain protein 3 negatively regulates peroxisome division
.
J. Integr. Plant Biol.
59
,
454
458
[PubMed]
46.
Rackles
E.
,
Witting
M.
,
Forné
I.
et al.
(
2021
)
Reduced peroxisomal import triggers peroxisomal retrograde signaling
.
Cell Rep
34
,
108653
[PubMed]
47.
Fang
Y.
,
Morrell
J.C.
,
Jones
J.M.
and
Gould
S.J.
(
2004
)
PEX3 functions as a PEX19 docking factor in the import of class I peroxisomal membrane proteins
.
J. Cell Biol.
164
,
863
875
[PubMed]
48.
Chen
Y.
,
Pieuchot
L.
,
Loh
R.A.
,
Yang
J.
,
Kari
T.M.A.
,
Wong
J.Y.
et al.
(
2014
)
Hydrophobic handoff for direct delivery of peroxisome tail-anchored proteins
.
Nat Commun
5
,
5790
[PubMed]
49.
Hua
R.
,
Gidda
S.K.
,
Aranovich
A.
,
Mullen
R.T.
and
Kim
P.K.
(
2015
)
Multiple domains in PEX16 mediate its trafficking and recruitment of peroxisomal proteins to the ER
.
Traffic
16
,
832
852
[PubMed]
50.
Burkhart
S.E.
,
Llinas
R.J.
and
Bartel
B.
(
2019
)
PEX16 contributions to peroxisome import and metabolism revealed by viable Arabidopsis pex16 mutants
.
J. Integr. Plant Biol.
61
,
853
870
[PubMed]
51.
Nito
K.
,
Kamigaki
A.
,
Kondo
M.
,
Hayashi
M.
and
Nishimura
M.
(
2007
)
Functional classification of Arabidopsis peroxisome biogenesis factors proposed from analyses of knockdown mutants
.
Plant Cell Physiol.
48
,
763
774
[PubMed]
52.
Hunt
J.E.
and
Trelease
R.N.
(
2004
)
Sorting pathway and molecular targeting signals for the Arabidopsis peroxin 3
.
Biochem. Biophys. Res. Commun.
314
,
586
596
[PubMed]
53.
Tian
L.
,
Chou
H.-L.
,
Fukuda
M.
,
Kumamaru
T.
and
Okita
T.W.
(
2020
)
mRNA localization in plant cells
.
Plant Physiol.
182
,
97
109
[PubMed]
54.
Zipor
G.
,
Haim-Vilmovsky
L.
,
Gelin-Licht
R.
,
Gadir
N.
,
Brocard
C.
and
Gerst
J.E.
(
2009
)
Localization of mRNAs coding for peroxisomal proteins in the yeast, Saccharomyces cerevisiae
.
Proc. Natl. Acad. Sci.
106
,
19848
55.
Dahan
N.
,
Bykov Yury
S.
,
Boydston Elizabeth
A.
et al.
(
2022
)
Peroxisome function relies on organelle-associated mRNA translation
.
Sci. Adv.
8
,
eabk2141
[PubMed]
56.
Flynn
C.R.
,
Mullen
R.T.
and
Trelease
R.N.
(
1998
)
Mutational analyses of a type 2 peroxisomal targeting signal that is capable of directing oligomeric protein import into tobacco BY-2 glyoxysomes
.
Plant J.
16
,
709
720
[PubMed]
57.
Reumann
S.
,
Ma
C.
,
Lemke
S.
and
Babujee
L.
(
2004
)
AraPerox. A database of putative Arabidopsis proteins from plant peroxisomes
.
Plant Physiol.
136
,
2587
2608
[PubMed]
58.
Reumann
S.
and
Chowdhary
G.
(
2018
)
Prediction of peroxisomal matrix proteins in plants
. In
Proteomics of Peroxisomes: Identifying Novel Functions and Regulatory Networks
(
del Río
L.A.
and
Schrader
M.
, eds),
125
138
,
Springer
,
Singapore
59.
Pan
R.
and
Hu
J.
(
2018
)
Proteome of plant peroxisomes
. In
Proteomics of Peroxisomes: Identifying Novel Functions and Regulatory Networks
(
del Río
L.A.
and
Schrader
M.
, eds), pp.
3
45
,
Springer
,
Singapore
60.
Ast
J.
,
Stiebler
A.
,
Freitag
J.
and
Bölker
M.
(
2013
)
Dual targeting of peroxisomal proteins
.
Front. Physiol.
4
297
[PubMed]
61.
Al-Hajaya
Y.
,
Karpinska
B.
,
Foyer
C.H.
and
Baker
A.
(
2022
)
Nuclear and peroxisomal targeting of catalase
.
Plant Cell Environ.
45
,
1096
1108
[PubMed]
62.
Vierstra
R.D.
(
2009
)
The ubiquitin-26S proteasome system at the nexus of plant biology
.
Nat. Rev. Mol. Cell Biol.
10
,
385
397
[PubMed]
63.
Traver
M.S.
,
Bradford
S.E.
,
Olmos
J.L.
,
Wright
Z.J.
,
Miller
M.D.
,
Xu
W.
et al.
(
2022
)
The structure of the Arabidopsis PEX4-PEX22 peroxin complex—insights into ubiquitination at the peroxisomal membrane
.
Front. Cell Dev. Biol.
10
,
838923
[PubMed]
64.
Zolman
B.K.
,
Monroe-Augustus
M.
,
Silva
I.D.
and
Bartel
B.
(
2005
)
Identification and functional characterization of Arabidopsis PEROXIN4 and the interacting protein PEROXIN22
.
Plant Cell
17
,
3422
3435
[PubMed]
65.
Kao
Y.-T.
and
Bartel
B.
(
2015
)
Elevated growth temperature decreases levels of the PEX5 peroxisome-targeting signal receptor and ameliorates defects of Arabidopsis mutants with an impaired PEX4 ubiquitin-conjugating enzyme
.
BMC Plant Biol.
15
,
224
224
[PubMed]
66.
Kaur
N.
,
Zhao
Q.
,
Xie
Q.
and
Hu
J.
(
2013
)
Arabidopsis RING peroxins are E3 ubiquitin ligases that interact with two homologous ubiquitin receptor proteins
.
J. Integr. Plant Biol.
55
,
108
120
[PubMed]
67.
Platta
H.W.
,
El Magraoui
F.
,
Bäumer
B.E.
,
Schlee
D.
,
Girzalsky
W.
and
Erdmann
R.
(
2009
)
Pex2 and Pex12 function as protein-ubiquitin ligases in peroxisomal protein import
.
Mol. Cell. Biol.
29
,
5505
5516
[PubMed]
68.
Ling
Q.
,
Huang
W.
,
Baldwin
A.
and
Jarvis
P.
(
2012
)
Chloroplast biogenesis is regulated by direct action of the ubiquitin-proteasome system
.
Science
338
,
655
659
[PubMed]
69.
Pan
R.
,
Satkovich
J.
and
Hu
J.
(
2016
)
E3 ubiquitin ligase SP1 regulates peroxisome biogenesis in Arabidopsis
.
Proc. Natl. Acad. Sci.
113
,
E7307
E7316
70.
Ling
Q.
,
Li
N.
and
Jarvis
P.
(
2017
)
Chloroplast ubiquitin E3 ligase SP1: Does it really function in peroxisomes?
Plant Physiol.
175
,
586
588
[PubMed]
71.
Lingard
M.J.
,
Monroe-Augustus
M.
and
Bartel
B.
(
2009
)
Peroxisome-associated matrix protein degradation in Arabidopsis
.
Proc. Natl. Acad. Sci.
106
,
4561
72.
Burkhart
S.E.
,
Lingard
M.J.
and
Bartel
B.
(
2013
)
Genetic dissection of peroxisome-associated matrix protein degradation in Arabidopsis thaliana
.
Genetics
193
,
125
141
[PubMed]
73.
Burkhart
S.E.
,
Kao
Y.-T.
and
Bartel
B.
(
2014
)
Peroxisomal ubiquitin-protein ligases Peroxin2 and Peroxin10 have distinct but synergistic roles in matrix protein import and Peroxin5 retrotranslocation in Arabidopsis
.
Plant Physiol.
166
,
1329
1344
[PubMed]
74.
Esser
C.
,
Alberti
S.
and
Höhfeld
J.
(
2004
)
Cooperation of molecular chaperones with the ubiquitin/proteasome system
.
Ubiquitin-Proteasome Syst.
1695
,
171
188
75.
Fernández-Fernández
M.R.
,
Gragera
M.
,
Ochoa-Ibarrola
L.
,
Quintana-Gallardo
L.
and
Valpuesta
J.M.
(
2017
)
Hsp70 - a master regulator in protein degradation
.
FEBS Lett.
591
,
2648
2660
[PubMed]
76.
Wimmer
B.
,
Lottspeich
F.
,
van der Klei
I.
,
Veenhuis
M.
and
Gietl
C.
(
1997
)
The glyoxysomal and plastid molecular chaperones (70-kDa heat shock protein) of watermelon cotyledons are encoded by a single gene
.
Proc. Natl. Acad. Sci.
94
,
13624
77.
Diefenbach
J.
and
Kindl
H.
(
2000
)
The membrane-bound DnaJ protein located at the cytosolic site of glyoxysomes specifically binds the cytosolic isoform 1 of Hsp70 but not other Hsp70 species
.
Eur. J. Biochem.
267
,
[PubMed]
78.
Arai
Y.
,
Hayashi
M.
and
Nishimura
M.
(
2008
)
Proteomic analysis of highly purified peroxisomes from etiolated soybean cotyledons
.
Plant Cell Physiol.
49
,
526
539
[PubMed]
79.
Li
G.
,
Li
J.
,
Hao
R.
and
Guo
Y.
(
2017
)
Activation of catalase activity by a peroxisome-localized small heat shock protein Hsp17.6CII
.
J. Genet. Genomics
44
,
395
404
[PubMed]
80.
Ma
C.
,
Haslbeck
M.
,
Babujee
L.
,
Jahn
O.
and
Reumann
S.
(
2006
)
Identification and characterization of a stress-inducible and a constitutive small heat-shock protein targeted to the matrix of plant peroxisomes
.
Plant Physiol.
141
,
47
60
[PubMed]
81.
Quan
S.
,
Yang
P.
,
Cassin-Ross
G.
,
Kaur
N.
,
Switzenberg
R.
,
Aung
K.
et al.
(
2013
)
Proteome analysis of peroxisomes from etiolated Arabidopsis seedlings identifies a peroxisomal protease involved in β-oxidation and development
.
Plant Physiol.
163
,
1518
1538
[PubMed]
82.
Ostersetzer
O.
,
Kato
Y.
,
Adam
Z.
and
Sakamoto
W.
(
2007
)
Multiple intracellular locations of Lon protease in Arabidopsis: Evidence for the localization of AtLon4 to chloroplasts
.
Plant Cell Physiol.
48
,
881
885
[PubMed]
83.
Reumann
S.
,
Quan
S.
,
Aung
K.
et al.
(
2009
)
In-depth proteome analysis of Arabidopsis leaf peroxisomes combined with in vivo subcellular targeting verification indicates novel metabolic and regulatory functions of peroxisomes
.
Plant Physiol.
150
,
125
143
[PubMed]
84.
Tsilibaris
V.
,
Maenhaut-Michel
G.
and
Van Melderen
L.
(
2006
)
Biological roles of the Lon ATP-dependent protease
.
Res. Microbiol.
157
,
701
713
[PubMed]
85.
Wohlever
M.L.
,
Baker
T.A.
and
Sauer
R.T.
(
2014
)
Roles of the N domain of the AAA+ Lon protease in substrate recognition, allosteric regulation and chaperone activity
.
Mol. Microbiol.
91
,
66
78
[PubMed]
86.
Zhang
F.
,
Hu
M.
,
Tian
G.
,
Zhang
P.
,
Finley
D.
,
Jeffrey
P.D.
et al.
(
2009
)
Structural insights into the regulatory particle of the proteasome from Methanocaldococcus jannaschii
.
Mol. Cell
34
,
473
484
[PubMed]
87.
Lee
I.
and
Suzuki
C.K.
(
2008
)
Functional mechanics of the ATP-dependent Lon protease- lessons from endogenous protein and synthetic peptide substrates
.
Biochim. Biophys. Acta
1784
,
727
735
[PubMed]
88.
Farmer
L.M.
,
Rinaldi
M.A.
,
Young
P.G.
,
Danan
C.H.
,
Burkhart
S.E.
and
Bartel
B.
(
2013
)
Disrupting autophagy restores peroxisome function to an Arabidopsis lon2 mutant and reveals a role for the LON2 protease in peroxisomal matrix protein degradation
.
Plant Cell
25
,
4085
4100
[PubMed]
89.
Lingard
M.J.
and
Bartel
B.
(
2009
)
Arabidopsis LON2 Is necessary for peroxisomal function and sustained matrix protein import
.
Plant Physiol.
151
,
1354
1365
[PubMed]
90.
Goto-Yamada
S.
,
Mano
S.
,
Nakamori
C.
,
Kondo
M.
,
Yamawaki
R.
,
Kato
A.
et al.
(
2014
)
Chaperone and protease functions of LON Protease 2 modulate the peroxisomal transition and degradation with autophagy
.
Plant Cell Physiol.
55
,
482
496
[PubMed]
91.
Schuhmann
H.
,
Huesgen
P.F.
,
Gietl
C.
and
Adamska
I.
(
2008
)
The DEG15 serine protease cleaves peroxisomal targeting signal 2-containing proteins in Arabidopsis
.
Plant Physiol.
148
,
1847
1856
[PubMed]
92.
Helm
M.
,
Lück
C.
,
Prestele
J.
et al.
(
2007
)
Dual specificities of the glyoxysomal/peroxisomal processing protease Deg15 in higher plants
.
Proc. Natl. Acad. Sci.
104
,
11501
93.
Dolze
E.
,
Chigri
F.
,
Höwing
T.
,
Hierl
G.
,
Isono
E.
,
Vothknecht
U.C.
et al.
(
2013
)
Calmodulin-like protein AtCML3 mediates dimerization of peroxisomal processing protease AtDEG15 and contributes to normal peroxisome metabolism
.
Plant Mol. Biol.
83
,
607
624
[PubMed]
94.
Clausen
T.
,
Southan
C.
and
Ehrmann
M.
(
2002
)
The HtrA family of proteases: Implications for protein composition and cell fate
.
Mol. Cell
10
,
443
455
[PubMed]
95.
Krojer
T.
,
Sawa
J.
,
Schäfer
E.
,
Saibil
H.R.
,
Ehrmann
M.
and
Clausen
T.
(
2008
)
Structural basis for the regulated protease and chaperone function of DegP
.
Nature
453
,
885
890
[PubMed]
96.
Schuhmann
H.
and
Adamska
I.
(
2012
)
Deg proteases and their role in protein quality control and processing in different subcellular compartments of the plant cell
.
Physiol. Plant.
145
,
224
234
[PubMed]
97.
Young
P.G.
and
Bartel
B.
(
2016
)
Pexophagy and peroxisomal protein turnover in plants
.
Biochim. Biophys. Acta
1863
,
999
1005
98.
Bassham
D.C.
(
2007
)
Plant autophagy—more than a starvation response
.
Curr. Opin. Plant Biol.
10
,
587
593
[PubMed]
99.
Hanaoka
H.
,
Noda
T.
,
Shirano
Y.
,
Kato
T.
,
Hayashi
H.
,
Shibata
D.
et al.
(
2002
)
Leaf senescence and starvation-induced chlorosis are accelerated by the disruption of an Arabidopsis autophagy gene
.
Plant Physiol.
129
,
1181
1193
[PubMed]
100.
Shibata
M.
,
Oikawa
K.
,
Yoshimoto
K.
,
Kondo
M.
,
Mano
S.
,
Yamada
K.
et al.
(
2013
)
Highly oxidized peroxisomes are selectively degraded via autophagy in Arabidopsis
.
Plant Cell
25
,
4967
4983
[PubMed]
101.
Kim
J.
,
Lee
H.
,
Lee
H.N.
,
Kim
S.-H.
,
Shin
K.D.
and
Chung
T.
(
2013
)
Autophagy-related proteins are required for degradation of peroxisomes in Arabidopsis hypocotyls during seedling growth
.
Plant Cell
25
,
4956
4966
[PubMed]
102.
Havé
M.
,
Luo
J.
,
Tellier
F.
,
Balliau
T.
,
Cueff
G.
,
Chardon
F.
et al.
(
2019
)
Proteomic and lipidomic analyses of the Arabidopsis atg5 autophagy mutant reveal major changes in endoplasmic reticulum and peroxisome metabolisms and in lipid composition
.
New Phytol.
223
,
1461
1477
[PubMed]
103.
Yamauchi
S.
,
Mano
S.
,
Oikawa
K.
,
Hikino
K.
,
Teshima
K.M.
,
Kimori
Y.
et al.
(
2019
)
Autophagy controls reactive oxygen species homeostasis in guard cells that is essential for stomatal opening
.
Proc. Natl. Acad. Sci.
116
,
19187
19192
104.
Nishimura
M.
,
Yamaguchi
J.
,
Mori
H.
,
Akazawa
T.
and
Yokota
S.
(
1986
)
Immunocytochemical analysis shows that glyoxysomes are directly transformed to leaf peroxisomes during greening of pumpkin cotyledons
.
Plant Physiol.
81
,
313
316
[PubMed]
105.
Young
P.G.
,
Passalacqua
M.J.
,
Chappell
K.
,
Llinas
R.J.
and
Bartel
B.
(
2019
)
A facile forward-genetic screen for Arabidopsis autophagy mutants reveals twenty-one loss-of-function mutations disrupting six ATG genes
.
Autophagy
15
,
941
959
[PubMed]
106.
Martinek
J.
,
Cifrová
P.
,
Vosolsobě
S.
,
Krtková
J.
,
Sikorová
L.
,
Malínská
K.
et al.
,
bioRxiv
, (
2022
)
ARP2/3 complex associates with peroxisomes to participate in pexophagy in plants
107.
Farré
J.-C.
,
Manjithaya
R.
,
Mathewson
R.D.
and
Subramani
S.
(
2008
)
PpAtg30 tags peroxisomes for turnover by selective autophagy
.
Dev. Cell
14
,
365
376
[PubMed]
108.
Motley
A.M.
,
Nuttall
J.M.
and
Hettema
E.H.
(
2012
)
Pex3-anchored Atg36 tags peroxisomes for degradation in Saccharomyces cerevisiae
.
EMBO J.
31
,
2852
2868
[PubMed]
109.
Deosaran
E.
,
Larsen
K.B.
,
Hua
R.
et al.
(
2013
)
NBR1 acts as an autophagy receptor for peroxisomes
.
J. Cell Sci.
126
,
939
952
[PubMed]
110.
Jung
H.
,
Lee
H.N.
,
Marshall
R.S.
,
Lomax
A.W.
,
Yoon
M.J.
,
Kim
J.
et al.
(
2020
)
Arabidopsis cargo receptor NBR1 mediates selective autophagy of defective proteins
.
J. Exp. Bot.
71
,
73
89
[PubMed]
111.
Ji
C.
,
Zhou
J.
,
Guo
R.
,
Lin
Y.
,
Kung
C.-H.
,
Hu
S.
et al.
(
2020
)
AtNBR1 is a selective autophagic receptor for AtExo70E2 in Arabidopsis
.
Plant Physiol.
184
,
777
791
[PubMed]
112.
Yamashita
S.
,
Abe
K.
,
Tatemichi
Y.
and
Fujiki
Y.
(
2014
)
The membrane peroxin PEX3 induces peroxisome-ubiquitination-linked pexophagy
.
Autophagy
10
,
1549
1564
[PubMed]
113.
Jiang
L.
,
Hara‐Kuge
S.
,
Yamashita
S.
and
Fujiki
Y.
(
2015
)
Peroxin Pex14p is the key component for coordinated autophagic degradation of mammalian peroxisomes by direct binding to LC3-II
.
Genes Cells
20
,
36
49
[PubMed]
114.
Sargent
G.
,
van Zutphen
T.
,
Shatseva
T.
,
Zhang
L.
,
Giovanni
V.D.
,
Bandsma
R.
et al.
(
2016
)
PEX2 is the E3 ubiquitin ligase required for pexophagy during starvation
.
J. Cell Biol.
214
,
677
690
[PubMed]
115.
Eberhart
T.
and
Kovacs
W.J.
(
2018
)
Pexophagy in yeast and mammals: an update on mysteries
.
Histochem. Cell Biol.
150
,
473
488
[PubMed]
116.
Li
J.
and
Wang
W.
(
2021
)
Mechanisms and functions of pexophagy in mammalian cells
.
Cells
10
,
1094
[PubMed]
117.
Law
K.B.
,
Bronte-Tinkew
D.
,
Di Pietro
E.
,
Snowden
A.
,
Jones
R.O.
,
Moser
A.
et al.
(
2017
)
The peroxisomal AAA ATPase complex prevents pexophagy and development of peroxisome biogenesis disorders
.
Autophagy
13
,
868
884
[PubMed]
118.
Park
N.Y.
,
Jo
D.S.
,
Park
S.J.
et al.
(
2021
)
Depletion of HNRNPA1 induces peroxisomal autophagy by regulating PEX1 expression
.
Biochem. Biophys. Res. Commun.
545
,
69
74
[PubMed]
119.
Gonzalez
K.L.
,
Ratzel
S.E.
,
Burks
K.H.
,
Danan
C.H.
,
Wages
J.M.
,
Zolman
B.K.
et al.
(
2018
)
A pex1 missense mutation improves peroxisome function in a subset of Arabidopsis pex6 mutants without restoring PEX5 recycling
.
Proc. Natl. Acad. Sci.
115
E3163
E3172
120.
Ohsumi
Y.
(
2001
)
Molecular dissection of autophagy: Two ubiquitin-like systems
.
Nat. Rev. Mol. Cell Biol.
2
,
211
216
[PubMed]
121.
Yang
Y.
,
Xiang
Y.
and
Niu
Y.
(
2021
)
An overview of the molecular mechanisms and functions of autophagic pathways in plants
.
Plant Signal Behav.
16
,
1977527
[PubMed]
122.
Marshall
R.S.
and
Vierstra
R.D.
(
2018
)
Autophagy: the master of bulk and selective recycling
.
Annu. Rev. Plant Biol.
69
,
173
208
[PubMed]
123.
Marshall
R.S.
,
Hua
Z.
,
Mali
S.
,
McLoughlin
F.
and
Vierstra
R.D.
(
2019
)
ATG8-binding UIM proteins define a new class of autophagy adaptors and receptors
.
Cell
177
,
766
781
,
e24
[PubMed]
124.
Xie
Q.
,
Tzfadia
O.
,
Levy
M.
,
Weithorn
E.
,
Peled-Zehavi
H.
,
Parys
T.V.
et al.
(
2016
)
hfAIM: A reliable bioinformatics approach for in silico genome-wide identification of autophagy-associated Atg8-interacting motifs in various organisms
.
Autophagy
12
,
876
887
[PubMed]
125.
Zhang
J.
,
Tripathi
D.N.
,
Jing
J.
et al.
(
2015
)
ATM functions at the peroxisome to induce pexophagy in response to ROS
.
Nat. Cell Biol.
17
,
1259
1269
[PubMed]
126.
Lee
J.N.
,
Dutta
R.K.
,
Maharjan
Y.
,
Liu
Z.
,
Lim
J.-Y.
,
Kim
S.-J.
et al.
(
2018
)
Catalase inhibition induces pexophagy through ROS accumulation
.
Biochem. Biophys. Res. Commun.
501
,
696
702
[PubMed]
127.
Calero-Muñoz
N.
,
Exposito-Rodriguez
M.
,
Collado-Arenal
A.M.
et al.
(
2019
)
Cadmium induces reactive oxygen species-dependent pexophagy in Arabidopsis leaves
.
Plant Cell Environ.
42
,
2696
2714
[PubMed]
128.
Terrón-Camero
L.C.
,
Rodríguez-Serrano
M.
,
Sandalio
L.M.
and
Romero-Puertas
M.C.
(
2020
)
Nitric oxide is essential for cadmium-induced peroxule formation and peroxisome proliferation
.
Plant Cell Environ.
43
,
2492
2507
[PubMed]
129.
Mitsuya
S.
,
El-Shami
M.
,
Sparkes
I.A.
,
Charlton
W.L.
,
Lousa
C.D.M.
,
Johnson
B.
et al.
(
2010
)
Salt stress causes peroxisome proliferation, but inducing peroxisome proliferation does not improve NaCl tolerance in Arabidopsis thaliana
.
PLoS ONE
5
,
e9408
e9408
[PubMed]
130.
Frick
E.M.
and
Strader
L.C.
(
2018
)
Kinase MPK17 and the Peroxisome Division Factor PMD1 influence salt-induced peroxisome proliferation
.
Plant Physiol.
176
,
340
351
[PubMed]
131.
Liu
Y.
,
Xiong
Y.
and
Bassham
D.C.
(
2009
)
Autophagy is required for tolerance of drought and salt stress in plants
.
Autophagy
5
,
954
963
[PubMed]
132.
Schrul
B.
and
Kopito
R.R.
(
2016
)
Peroxin-dependent targeting of a lipid-droplet-destined membrane protein to ER subdomains
.
Nat. Cell Biol.
18
,
740
751
[PubMed]
133.
Galiani
S.
,
Waithe
D.
,
Reglinski
K.
,
Cruz-Zaragoza
L.D.
,
Garcia
E.
,
Clausen
M.P.
et al.
(
2016
)
Super-resolution microscopy reveals compartmentalization of peroxisomal membrane proteins
.
J. Biol. Chem.
291
,
16948
16962
[PubMed]
134.
Grützner
R.
,
Martin
P.
,
Horn
C.
,
Mortensen
S.
,
Cram
E.J.
,
Lee-Parsons
C.W.T.
et al.
(
2021
)
High-efficiency genome editing in plants mediated by a Cas9 gene containing multiple introns
.
Plant Commun.
2
,
100135
[PubMed]
135.
Preisig-Müller
R.
,
Muster
G.
and
Kindl
H.
(
1994
)
Heat shock enhances the amount of prenylated Dnaj protein at membranes of glyoxysomes
.
Eur. J. Biochem.
219
,
57
63
[PubMed]

Author notes

*

These authors contributed equally to this work.

This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY-NC-ND).