Plant peroxisomes are recognized organelles that – with their capacity to generate greater amounts of H2O2 than other subcellular compartments – have a remarkable oxidative metabolism. However, over the last 15 years, new information has shown that plant peroxisomes contain other important molecules and enzymes, including nitric oxide (NO), peroxynitrite, a NADPH-recycling system, Ca2+ and lipid-derived signals, such as jasmonic acid (JA) and nitro-fatty acid (NO2-FA). This highlights the potential for complex interactions within the peroxisomal nitro-oxidative metabolism, which also affects the status of the cell and consequently its physiological processes. In this review, we provide an update on the peroxisomal interactions between all these molecules. Particular emphasis will be placed on the generation of the free-radical NO, which requires the presence of Ca2+, calmodulin and NADPH redox power. Peroxisomes possess several NADPH regeneration mechanisms, such as those mediated by glucose-6-phosphate dehydrogenase (G6PDH) and 6-phosphogluconate dehydrogenase (6PGDH) proteins, which are involved in the oxidative phase of the pentose phosphate pathway, as well as that mediated by NADP-isocitrate dehydrogenase (ICDH). The generated NADPH is also an essential cofactor across other peroxisomal pathways, including the antioxidant ascorbate–glutathione cycle and unsaturated fatty acid β-oxidation, the latter being a source of powerful signaling molecules such as JA and NO2-FA.

Peroxisomes were first described at the cellular level in animal and plant organs in the 1950s and 1960s (Rhodin, 1954; Frederick et al., 1968), when their biochemical composition also began to be analyzed. They have been shown to have a basic oxidative metabolism, characterized by high hydrogen peroxide (H2O2) content and the presence of catalase (De Duve et al., 1960; Tolbert et al., 1968). This enzyme, which is exclusively found in peroxisomes, is considered to be one of the main antioxidant enzymes and part of the first line of defense against oxidative stress. Subsequent advances in the study of these organelles have revealed that they are extraordinary subcellular compartments characterized by great endogenous metabolic adaptability and highly active interactions with other subcellular compartments (van der Zand et al., 2006; Mast et al., 2015; Wanders et al., 2016). This adaptability is even more remarkable for plant peroxisomes, because their enzymatic composition reflects the functionality of the respective tissue and can change depending on the cell type in each plant organ (root, cotyledon or leaf), the stage of development or external conditions (Goto-Yamada et al., 2015; Reumann and Bartel, 2016). An excellent example of plant peroxisome adaptability and flexibility is the dynamic and reversible interactions between leaf peroxisomes and chloroplasts, which varies depending on illumination conditions. Here, the external appearance of peroxisomes changes from a spherical (dark conditions) to an elongated elliptical (light conditions) shape in order to increase the contact surface with chloroplast, thus facilitating effective metabolite interchange (Oikawa et al., 2015; Corpas, 2015).  The use of new biochemical, proteomic and cellular approaches has increased our knowledge of the components and functions of plant peroxisomes (Palma et al., 2009; Hu et al., 2012; Kataya et al., 2015; Reumann and Bartel, 2016; Cui et al., 2016; Corpas et al., 2017). Furthermore, H2O2 and nitric oxide (NO) are not only found in chloroplasts and mitochondria, but are also present in peroxisomes. This raises the question as to whether they function as signals between the different cellular compartments. This review provides an update on the metabolic interconnections between some of the newly identified components in peroxisomes, including NO, Ca2+, NADPH and nitro fatty acid (NO2-FA). Thus, the involvement of enzymes, such as NADP-dehydrogenases (NADP-DHs), NAD kinases (NADKs) and Nudix hydrolase 19 (NUDX19, also known as NUDT19), contributes to the regulation of peroxisomal redox status mediating a change in the NADPH:NADP and GSH:GSSG ratio. These molecules are also connected to the enzymatic generation of NO, which involves NADPH, calmodulin (CaM) and Ca2+. In addition, the presence of lipid-derived signals, such as jasmonic acid (JA) and NO2-FA, opens up new avenues of research into the importance of peroxisomes and their interaction with other cellular compartments.

NO generation in peroxisomes

NO is a versatile molecule involved in numerous physiological processes and in the mechanism of defense against both biotic and abiotic stresses (Domingos et al., 2015; Simontacchi et al., 2015). Given the prominence of peroxisomal oxidative metabolism over the last 15 years, research has especially focused on the potential presence of NO and related molecules called reactive nitrogen species (RNS) in peroxisomes. We have analyzed the generation of this molecule and its modulation of proteins whose function in peroxisomes is modified upon post-translational modification, such as nitration and S-nitrosation (Corpas and Barroso, 2014a; Chaki et al., 2015; Begara-Morales et al., 2016; Corpas et al., 2017), as well as its potential role as a signaling molecule promoting intracellular communication. NO has been detected in peroxisomes in different plant species using techniques such as spin trapping electron paramagnetic resonance (EPR) spectroscopy and fluorescence probes (Corpas et al., 2004, 2009). In addition, there is biochemical evidence to show that peroxisomal NO is generated by an L-arginine-dependent NO synthase (NOS)-like activity in plant cells (Barroso et al., 1999; Corpas and Barroso, 2016). Interestingly, NOS proteins have also been detected in animal peroxisomes (Stolz et al., 2002), suggesting that this enzyme is a constitutive component of peroxisomes, although the identity of the plant NOS protein is still unknown. However, it is known that the NOS protein contains a type 2 peroxisomal-targeting signal (PTS2), which is imported into the peroxisomal matrix; this domain is present in enzymes responsible for NO generation in plant and animal systems (Loughran et al., 2013; Corpas and Barroso, 2014b). Plant peroxisomal NOS-like activity has been found to require additional molecules – especially NADPH, Ca2+ and CaM, similar to what is found during the regulation of NOS activity in animal cells (Barroso et al., 1999; Corpas et al., 2004; Alderton et al., 2001). Peroxins (PEXs), which are peroxisomal biogenesis factors located in the cytosol and the peroxisomal membrane itself, are involved in matrix protein transport (Cross et al., 2016). In addition, the import of peroxisomal NOS proteins requires PEX12, as well as both of the PTS2 receptors PEX5 and PEX7. Through the use of Ca2+ channel blockers and CaM antagonists, we have also shown that this import is strictly dependent on both Ca2+ and CaM (Corpas and Barroso, 2014b). Fig. 1 shows a working model of the known elements involved in the peroxisomal protein import mechanism responsible for NO generation.

Fig. 1.

Schematic model of known elements involved in the import mechanism of NOS which generates NO in the peroxisome. The NOS protein contains a PTS2 domain. Import of the peroxisomal NOS protein requires Ca2+, CaM and the involvement of several peroxins (PEXs), including PEX12, which is located in the membrane, and both the PTS receptors (PEX5 and PEX7), located in the cytosol. Following the import of the NOS protein into the matrix, PTS2 is cleaved by a glyoxysomal-processing protease (GPP). Ca2+ needed for NOS activity could move through a porin ion channel located in the peroxisomal membrane. NO can mediate post-translational modifications (nitration and S-nitrosation) and functions as a signaling molecule outside the peroxisomes. Dashed lines with arrow indicate the signaling processes. This figure has been modified Corpas and Barroso (2014b) with permission from Wiley © 2015 Federation of European Biochemical Societies.

Fig. 1.

Schematic model of known elements involved in the import mechanism of NOS which generates NO in the peroxisome. The NOS protein contains a PTS2 domain. Import of the peroxisomal NOS protein requires Ca2+, CaM and the involvement of several peroxins (PEXs), including PEX12, which is located in the membrane, and both the PTS receptors (PEX5 and PEX7), located in the cytosol. Following the import of the NOS protein into the matrix, PTS2 is cleaved by a glyoxysomal-processing protease (GPP). Ca2+ needed for NOS activity could move through a porin ion channel located in the peroxisomal membrane. NO can mediate post-translational modifications (nitration and S-nitrosation) and functions as a signaling molecule outside the peroxisomes. Dashed lines with arrow indicate the signaling processes. This figure has been modified Corpas and Barroso (2014b) with permission from Wiley © 2015 Federation of European Biochemical Societies.

Ca2+ and CaM function in plant peroxisomes

It is well established that Ca2+ and CaM are key regulatory elements in many cellular processes in animals and plants (Hepler, 2005; Clapham, 2007; Perochon et al., 2011). The presence of these factors in plant peroxisomes, as well as their potential metabolic functions, has recently begun to be studied. Ca2+ can move through membranes using voltage-dependent anion channels, among others (Gincel et al., 2001; Jammes et al., 2011). These channels are a class of porin ion channel, and their presence has been described in peroxisomal membranes (Fig. 1) (Reumann et al., 1997, 1998; Corpas et al., 2000). However, the presence of Ca2+ and its effect on catalase have only recently been assessed in plant peroxisomes (Costa et al., 2010; Yang and Poovaiah, 2012). The presence of Ca2+ also closely correlates with that of Ca2+-dependent protein kinase (CDPK) CPK1 in Arabidopsis peroxisomes. CPK1 appears to be not only involved in lipid metabolism and oxidative stress (Dammann et al., 2003), but also in mediating pathogen resistance (Coca and San Segundo, 2010). Similarly, CDPK2, which is involved in pollen tube growth, has been found in Petunia inflata peroxisomes (Yoon et al., 2006; Guo et al., 2013), which is indicative of a direct involvement of Ca2+ and peroxisomal NO in this process (Prado et al., 2004, 2008).

Additionally, Arabidopsis peroxisomes have been reported to contain the calmodulin-like (CML) protein CML3 (Chigri et al., 2012). CML3 can modulate Ca2+ signals, as it contains amino acid sequences commonly found in canonical CaM-binding motifs (Perochon et al., 2011). It mediates the dimerization of the Arabidopsis thaliana peroxisomal ATP-independent serine endopeptidase (DEG15), which belongs to the family of degradation of periplasmic protein (DEG) proteases; these are also orthologs of glyoxysomal-processing protease (GPP) in the watermelon (Helm et al., 2007). In its monomeric form, GPP exerts general protease activity; in its dimeric form, it is capable of specifically cleaving PTS2 domains (Fig. 1) (Schuhmann et al., 2008; Dolze et al., 2013). The PTS2 domain is present in plant peroxisomal enzymes, such as malate dehydrogenase, citrate synthase, acyl-CoA oxidase and thiolase; after cleavage by this protease, the proteins become functional (Mullen, 2002; Johnson and Olsen, 2003; Pracharoenwattana et al., 2007; Meng et al., 2014).

Recently, we reported that different CaM antagonists cause a reduction in peroxisomal Ca2+ levels, which adversely affects peroxisomal protein import (Corpas and Barroso, 2017b). Several peroxisomal enzymes, including catalase, hydroxypyruvate reductase (HPR) and glycolate oxidase (GOX), involved in photorespiration, were also found to be downregulated upon such treatment (Corpas and Barroso, 2017b). GOX is an important source of H2O2, suggesting that there is a close correlation between Ca2+ levels and H2O2. Therefore, it is plausible to assume that peroxisomal Ca2+ is crucial for normal function of several organelles. Table 1 summarizes the effects of Ca2+ on the peroxisomal protein function.

Table 1.

Plant peroxisomal proteins that modulate and are affected by Ca2+ content

Plant peroxisomal proteins that modulate and are affected by Ca2+ content
Plant peroxisomal proteins that modulate and are affected by Ca2+ content

Thus, the newly identified peroxisomal molecules NO, Ca2+ and CaM show metabolic interactions, which evokes new questions regarding the physiological role of peroxisomes in both optimal and adverse environmental conditions.

The reducing power of the NADPH cofactor is well known to be essential for normal cell growth and proliferation. As mentioned previously, NADPH is also necessary as an electron donor for the generation of NO in peroxisomes. In plants, the main source of NADPH in photosynthetic cells is the ferrodoxin-NADP reductase (FNR) enzyme. However, in non-photosynthetic cells, such as roots, or in photosynthetic cells during the dark phase, a family of NADP-DHs regenerates NADPH. These NADP-DHs include two enzymes of the oxidative step of the pentose phosphate pathway, the glucose-6-phosphate dehydrogenase (G6PDH) and 6-phosphogluconate dehydrogenase (6PGDH) proteins, as well as the NADP-malic enzyme (NADP-ME), also known as NADP-malate dehydrogenase, and NADP-isocitrate dehydrogenase (NADP-ICDH) (Corpas and Barroso, 2014c). All these NADP-DHs have different isozymes that are present in the respective subcellular compartments, including the cytosol, the chloroplast, mitochondria and peroxisomes.

NADP-DHs in peroxisomes

A set of NADPH-recycling enzymes have been detected in plant peroxisomes, including the above mentioned NADP-ICDH (Corpas et al., 1999; Reumann et al., 2007; Leterrier et al., 2016), as well as G6PDH and 6PGDH proteins (Corpas et al., 1998; Fernández-Fernández and Corpas, 2016; Hölscher et al., 2016) (see also Table 2). These NADP-DHs operate in coordination with certain import mechanisms in order to supply the substrate corresponding to each enzyme (Fig. 2). With regard to isocitrate, which is necessary for NADP-ICDH activity, the most plausible import mechanism is through the isocitrate–2-oxoglutarate shuttle (Reumann, 2000; Rottensteiner and Theodoulou, 2006; Linka and Theodoulou, 2013). However, the import system for glucose 6-phosphate (G6P), which is used by the G6PDH enzyme, has – to our knowledge – not been described in the peroxisomal membrane. One suspects the presence of a G6P translocator, similar to what is known to occur in plastids (Kunz et al., 2010). Nevertheless, an alternative import route for G6P could be through a porin in the peroxisomal membrane (Fig. 2) (Reumann et al., 1998).

Table 2.

Plant peroxisomal enzymes that consume or regenerate NADPH

Plant peroxisomal enzymes that consume or regenerate NADPH
Plant peroxisomal enzymes that consume or regenerate NADPH
Fig. 2.

Schematic representation of peroxisomal NADPH recycling pathways. NADPH is regenerated by several NADP-DHs, including G6PDH, 6PGDH and NADP-ICDH. The substrates for these dehydrogenases may be imported through different mechanisms: through a porin for G6P or through an isocitrate–2-oxoglutarate (2-OG) shuttle in the case of isocitrate. Both are located in the peroxisomal membrane. NADPH is used in several processes, including β-oxidation, NO generation by an L-arginine-dependent NOS activity and conversion of GSSG into GSH by GR activity. NADPH could be hydrolyzed to reduced nicotinamide mononucleotide (NMNH) plus 2′,5′-ADP by NUDX19 activity. The NADP+ pool is provided by NAD+ phosphorylation through NADK activity. NAD+ may be imported into the peroxisomal matrix by a NAD+ carrier (PXN) present in the membrane. PG, 6-phosphogluconolactone; 6PG, 6-phosphogluconate; Rib 5P, ribulose 5-phosphate.

Fig. 2.

Schematic representation of peroxisomal NADPH recycling pathways. NADPH is regenerated by several NADP-DHs, including G6PDH, 6PGDH and NADP-ICDH. The substrates for these dehydrogenases may be imported through different mechanisms: through a porin for G6P or through an isocitrate–2-oxoglutarate (2-OG) shuttle in the case of isocitrate. Both are located in the peroxisomal membrane. NADPH is used in several processes, including β-oxidation, NO generation by an L-arginine-dependent NOS activity and conversion of GSSG into GSH by GR activity. NADPH could be hydrolyzed to reduced nicotinamide mononucleotide (NMNH) plus 2′,5′-ADP by NUDX19 activity. The NADP+ pool is provided by NAD+ phosphorylation through NADK activity. NAD+ may be imported into the peroxisomal matrix by a NAD+ carrier (PXN) present in the membrane. PG, 6-phosphogluconolactone; 6PG, 6-phosphogluconate; Rib 5P, ribulose 5-phosphate.

Another way to maintain NADPH regeneration is to provide an appropriate NADP+ pool, which can be affected through the NAD+ metabolism. Recently, the membrane-located peroxisomal NAD+ carrier (PXN), which enables NAD+ to be imported in exchange for adenosine monophosphate (AMP), has been described in Arabidopsis (Bernhardt et al., 2012; van Roermund et al., 2016). In addition, NAD kinase (NADK) activity is known to catalyze the direct phosphorylation of NAD+ to NADP+ (Pollak et al., 2007) (Fig. 2). Of the three NADKs identified in Arabidopsis (Turner et al., 2005), NADK3 has been reported to be localized in the peroxisomal matrix (Waller et al., 2010), where it contributes to the generation of the peroxisomal NADP+ pool (Chai et al., 2006; Li et al., 2014) and maintains peroxisomal NADP-DH activity (Fig. 2).

In addition to the capacity to regenerate NADPH, the peroxisomal NADP-DHs are involved in other potential biological processes, such as leaf senescence, stomata movements and pollen-tube–ovule interactions. For example, leaf senescence is a highly regulated process characterized by antioxidative capacity loss, in which peroxisomes are involved (Pastori and del Río, 1997; Jiménez et al., 1998; Zimmermann et al., 2006). With regard to peroxisomal NADP-ICDH activity, a comparative analysis of its kinetic parameters in young and senescent leaves of pea plants has revealed that the NADP-ICDH affinity (Km) for their isocitrate substrate decreased drastically during leaf senescence. This leads to over a 20-fold increase in isocitrate affinity, while the maximum rate (Vmax) remains unchanged (Corpas et al., 1999). There are two possible explanations for this finding: first, that this allows the peroxisomal intracellular pool of isocitrate to outcompete the glyoxylate cycle enzyme isocitrate lyase, which appears in peroxisomes during leaf senescence (Pastori and del Río, 1997); or second, that the increase provides a larger supply of NADPH to eliminate excess H2O2 produced during senescence, when catalase activity decreases sharply (Corpas et al., 1999; Zimmermann et al., 2006). Stomatal movement is regulated by a complex network of signals, including Ca2+, NO, H2O2 and phytohormones (abscisic acid, JA and salicylic acid) (Murata et al., 2015). Recently Arabidopsis null mutants for the peroxisomal NADP-ICDH were shown to have stomatal movement malfunction (Leterrier et al., 2016). In this regard, it is worth pointing out that senescence and stomatal movement are also regulated by NO levels in peroxisomes (Corpas et al., 2004; Leterrier et al., 2016). Another example is peroxisomal 6PGDH in Arabidopsis, whose activity is necessary during growth and developmental processes, such as male gametophyte growth and pollen-tube–ovule interaction (Hölscher et al., 2016). Interestingly, Ca2+ and peroxisomal NO are also directly involved in the latter processes (Prado et al., 2004, 2008).

Function of peroxisomal NADPH

Apart from the involvement of NADPH in generating NO, as mentioned above, this reducing agent is also required for several other peroxisomal enzymatic activities that are involved in various pathways (see Table 2). For example, 2,4-dienoyl-CoA reductase (DECR) activity catalyzes the conversion of 2-trans,4-trans-dienoyl-CoA+NADPH into 3-trans-enoyl-CoA+NADP+ (Behrends et al., 1988; Goepfert and Poirier, 2007). This peroxisomal DECR is part of the alternative β-oxidation pathway present in plant peroxisomes, which is necessary to degrade unsaturated fatty enoyl-CoA esters with double bonds in even- and odd-numbered carbons (Fig. 3) (Rottensteiner and Theodoulou, 2006). Furthermore, 12-oxophytodienoate reductase (OPR) enzymes catalyze the conversion of 12-oxo-phytodienoic acid (OPDA)+NADPH into 3-oxo-2[2′(Z)-pentenyl]-cyclopentane-1-octanoate (OPC-8:0) and NADP+. In Arabidopsis, OPR3 is located in peroxisomes and is involved in JA biosynthesis, which has a role in signaling during development and in the response to adverse stresses, such as mechanical wounding and pathogens (Schaller et al., 2000; Strassner et al., 2002; Baker et al., 2006; Tani et al., 2008; Gfeller et al., 2010, see also below).

Fig. 3.

Model of the metabolic integration of NADPH, NO and Ca2+ in plant peroxisomes. The NADPH pool is necessary for different processes, including L-arginine-dependent NOS activity, GSH regeneration through GR activity (part of the ascorbate–glutathione cycle) and β-oxidation. NOS activity requires the presence of Ca2+, CaM and NADPH to generate NO, which can react with reduced glutathione (GSH) to form GSNO, or mediate the formation of NO2-FAs, such as NO2-Ln. Alternatively, the electrophilic property of NO2-Ln may lead to nitroalkylation of proteins containing nucleophilic residues (Cys, His or Lys). This could contribute to the signaling actions, involving NO2-FAs, and the observed antioxidant properties of these molecules. NO can also function as a signal molecule outside the peroxisomes or positively modulate the expression of OPR3, which mediates JA biosynthesis. NADPH is further needed for DECR activity, which is involved in unsaturated fatty acid β-oxidation. Potential excess NADPH could be hydrolyzed to reduced nicotinamide NMNH plus 2′,5′-ADP by NUDX19 activity. By contrast, the NADP+ pool can be maintained via NAD+ phosphorylation through NADK activity. NAD+ may be imported into the peroxisome by a NAD+ carrier (PXN) present in the membrane. On the other hand, Ca2+ could move through a porin ion channel located in the peroxisomal membrane. Peroxisomes contain several NADPH-recycling enzymes, including G6PDH, 6PGDH and NADP-ICDH. OPDA, 12-oxo-phytodienoic acid. Dashed lines with arrow indicate the signaling processes.

Fig. 3.

Model of the metabolic integration of NADPH, NO and Ca2+ in plant peroxisomes. The NADPH pool is necessary for different processes, including L-arginine-dependent NOS activity, GSH regeneration through GR activity (part of the ascorbate–glutathione cycle) and β-oxidation. NOS activity requires the presence of Ca2+, CaM and NADPH to generate NO, which can react with reduced glutathione (GSH) to form GSNO, or mediate the formation of NO2-FAs, such as NO2-Ln. Alternatively, the electrophilic property of NO2-Ln may lead to nitroalkylation of proteins containing nucleophilic residues (Cys, His or Lys). This could contribute to the signaling actions, involving NO2-FAs, and the observed antioxidant properties of these molecules. NO can also function as a signal molecule outside the peroxisomes or positively modulate the expression of OPR3, which mediates JA biosynthesis. NADPH is further needed for DECR activity, which is involved in unsaturated fatty acid β-oxidation. Potential excess NADPH could be hydrolyzed to reduced nicotinamide NMNH plus 2′,5′-ADP by NUDX19 activity. By contrast, the NADP+ pool can be maintained via NAD+ phosphorylation through NADK activity. NAD+ may be imported into the peroxisome by a NAD+ carrier (PXN) present in the membrane. On the other hand, Ca2+ could move through a porin ion channel located in the peroxisomal membrane. Peroxisomes contain several NADPH-recycling enzymes, including G6PDH, 6PGDH and NADP-ICDH. OPDA, 12-oxo-phytodienoic acid. Dashed lines with arrow indicate the signaling processes.

Another peroxisomal enzyme that requires NADPH is glutathione reductase (GR; At3g24170). This enzyme is part of the ascorbate–glutathione cycle, which facilitates the regeneration of the antioxidant glutathione (GSH) in the following reaction: GSSG+NADPH→GSH+NADP+. GR, whose activity has been observed to increase during natural leaf senescence and in response to abiotic stresses, has been detected in different plant species (Jiménez et al., 1998; Romero-Puertas et al., 2006; Kataya and Reumann, 2010). Similarly, its product, peroxisomal GSH, is increased under stress conditions and can reach a concentration of ∼5 mM in Arabidopsis leaf peroxisomes (Koffler et al., 2013; Zechmann, 2014). GSH can also react with NO to form S-nitrosoglutathione (GSNO), which may function as an intracellular NO reservoir; it mediates post-translational modifications, such as S-nitrosation. This affects the function of the protein, as previously described for catalase, ascorbate peroxidase and monodehydroascorbate peroxidase (Begara-Morales et al., 2016). This highlights connections between the peroxisomal metabolism of NADPH, antioxidant molecules and some the NO-related RNS molecules (Barroso et al., 2013; Corpas et al., 2017).

The last group of enzymes involved in the use of NADPH are pyrophosphatases that contain a highly conserved amino acid sequence or Nudix box, Gx5Ex7REUxEExGU, in which U is a hydrophobic amino acid such as Val, Ile or Leu (Ogawa et al., 2008). NUDX19 hydrolyzes NADPH to reduced nicotinamide mononucleotide (NMNH) plus 2′,5′-ADP. Amino acid sequence analysis suggests that Arabidopsis NUDX19 is located in both peroxisomes and chloroplasts (Ogawa et al., 2008; Lingner et al., 2011). Specifically, it has a C-terminal PTS1 tripeptide (SSL), which is regarded as a weak, but effective, type 1 PTS (PTS1). Indeed, the peroxisomal presence of Arabidopsis NUDX19 has been experimentally confirmed in Arabidopsis (Lingner et al., 2011). Owing to its NADPH pyrophosphohydrolase activity, NUDX19 can modulate the cellular levels and redox states of this pyridine nucleotide, which could be involved in NADP-DH regulation and hormonal signaling. Recent studies demonstrate that the loss of NUDX19 function affects NADPH content and mediates the increased resistance to arsenic and photooxidative stresses in Arabidopsis (Corpas et al., 2016; Maruta et al., 2016). The presence of NUDX19 in peroxisomes and chloroplasts points to metabolic and physical interactions during photorespiration between these organelles (Oikawa et al., 2015). Interestingly, in the Arabidopsis NUDX19 knockout mutant, a significant increase in all NADP-dehydrogenase activities (NADP-ICDH, G6PDH and 6PGDH) is triggered in roots and leaves, especially under induced arsenic stress (Corpas et al., 2016). This highlights the importance of NUDX19 in NADPH homeostasis in plants.

The cofactor NADPH could therefore be a necessary link in the redox metabolism of peroxisomes and must be regenerated to function normally; NADPH also represents a connection between several pathways, such as NO generation, the antioxidant system and β-oxidation.

Plant peroxisomes house all the components necessary for the β-oxidation of fatty acids (FAs); this oxidation is essential for the mobilization of the lipid reservoir required for seedling development (Graham, 2008). Additionally, peroxisomes are also important for the metabolism of several lipid-derived signals that have a crucial physiological role in the entire plant. The import of FAs into peroxisomes is mediated by ATP-binding cassette (ABC) transporter 1 [ABCC1, also known as COMATOSE (CTS)], which is involved in the import of other molecules such as 12-oxo-phytodienoic acid (OPDA), a precursor of jasmonic acid (JA), as well as 4-dichlorophenoxybutyric acid and indole butyric acid (IBA); the latter of which are precursors of the phytohormone auxin, which is involved in plant growth and development (Baker et al., 2015).

As mentioned above, JA is a linolenic acid-derived cyclopentanone mainly involved in mediating wounding responses (Theodoulou et al., 2005; Delker et al., 2007; Gfeller et al., 2010; León, 2013). JA and its derivatives also modulate a wide range of activities, such as plant development, as well as abiotic (including salinity, drought and UV radiation) and biotic stress tolerance; these activities also provide a defense against pathogen infection and insect attacks (Wasternack and Song, 2017). In this context, the interplay between NO, phytohormones (indoleacetic acid, abscisic acid and ethylene) and other plant growth regulators (polyamines) has been well established (Freschi, 2013; Asgher et al., 2017). This interplay could be of physiological significance in the peroxisomal metabolism, during which these molecules are generated. Indeed, in Arabidopsis, there is evidence that NO increases JA production by increasing the expression of the OPR3 gene in peroxisomes (Fig. 3) (Huang et al., 2004; Mur et al., 2013; Hussain et al., 2016). A similar effect of NO on JA has been observed in tomato roots in response to nematode infection (Zhou et al., 2015). OPR3 activity also affects the metabolism of other molecules, including those of glucosinolates and tryptophan (Cao et al., 2016). Tryptophan is a precursor of serotonin and indoleacetic acid (IAA), which are involved in plant growth and stress responses (Zolman et al., 2008; Schlicht et al., 2013; Spiess and Zolman, 2013). Therefore, apart from the important role played by peroxisomal β-oxidation in lipid mobilization during germination, there are other derived lipids, such as JA, whose synthesis specifically requires NADPH for OPR3 activity. However, expression of the OPR3 gene is also stimulated by NO, which establishes another connection point between NADPH, JA and NO in peroxisomes (Fig. 3).

NO2-FAs are derived from NO and produced by the addition of a nitro group (–NO2) to unsaturated fatty acids. These molecules were first shown to have important physiological functions in animals, such as cardiac ischemic injury protection, regulation of inflammatory processes and reduction of atherosclerosis (Freeman et al., 2008; Trostchansky et al., 2013). In higher plants, the presence of NO2-FAs was hypothesized a few years ago (Sánchez-Calvo et al., 2013). Recently, we have reported on the endogenous presence and function of NO2-FAs in Arabidopsis, with the most prominent NO2-FA being nitro-linolenic acid (NO2-Ln) (Mata-Pérez et al., 2016a). The highest NO2-Ln content, which diminished during plant development, was found in Arabidopsis seeds. NO2-Ln activates a molecular chaperone network in response to diverse abiotic stresses and has been shown to have a powerful signaling function. NO2-Ln has the capacity to either increase or decrease the gene expression of hundreds of genes with a remarkable induction capacity, and ∼25% of the total upregulated genes encode heat-shock proteins (Mata-Pérez et al., 2016a). Recently, mass spectrometry analysis of the peroxisomal fraction isolated from the leaves of pea plants has revealed that peroxisomes also contain NO2-Ln (Mata-Perez et al., 2017). This demonstrates that there is a highly active NO metabolism in plant peroxisomes, as these molecules may function as a NO reservoir with the capacity to release NO (Mata-Pérez et al., 2016b). Alternatively, the electrophilic property of NO2-Ln could allow it to contribute to nitroalkylation of a protein that contains nucleophilic residues (Cys, His or Lys), which could contribute to the signaling action of NO2-FAs. Taken together, the physiological function of these molecules in plant peroxisomes raises issues that need to be addressed and could provide new evidence to show that these organelles have a potential signaling function mediated by GSNO and NO2-FAs.

The metabolism of plant peroxisomes is characterized by adaptability and flexibility, and depends on the organ, stage of development and response to adverse environmental conditions. Consequently, peroxisomes should be regarded as nodes at the crossroads of different metabolic pathways. Examples of these pathways are photorespiration, lipid mobilization and the nitro-oxidative metabolism, which endow cells with signaling molecules, including H2O2, NO, GSNO, NO2-FA and JA. Fig. 3 summarizes a model for the interplay between the various molecules described in this review of plant peroxisomes. Their NADPH content is affected by NADP-DHs, NADKs and NUDX19, which help to regulate the peroxisomal redox status. Moreover, NADPH, synthesis of NO, ascorbate–glutathion cycle and β-oxidation are important for several essential metabolic pathways; NADPH also mediates the generation of lipid-derived signals (JA and NO2-FA).

However, several points regarding these new peroxisomal molecules need to be addressed in future research: the mechanisms that control their transport through the peroxisomal membrane of Ca2+ and the substrates specific to NADP-DHs; the identification of the protein responsible for NO synthesis; the peroxisomal proteins susceptible to post-translational modifications mediated by NO (nitration, S-nitrosation or nitroalkylation); the specific function of peroxisomal NO2-FAs under physiological and stress conditions; and the cellular function of peroxisomal molecules as compared to molecules from other subcellular compartments. The answers to all these points could provide a better understanding of the cellular role of peroxisomes as a source of intra- and inter-cellular signaling molecules (H2O2, NO, NO2-FA and JA). Such an understanding may also be useful for researchers working on peroxisomes in other organisms, including yeast and mammals, in order to provide a more complete picture of the signals involved in biochemical communication between organelles.

We would also like to thank Michael O'Shea for proofreading the text.

Funding

Research in our laboratories is supported by an European Regional Development Fund (ERDF) grant co-financed by the Ministerio de Economía y Competitividad (projects AGL2015-65104-P and BIO2015-66390-P) and the Junta de Andalucía (groups BIO192 and BIO286) in Spain.

Alderton
,
W. K.
,
Cooper
,
C. E.
and
Knowles
,
R. G.
(
2001
).
Nitric oxide synthases: structure, function and inhibition
.
Biochem. J.
357
,
593
-
615
.
Asgher
,
M.
,
Per
,
T. S.
,
Masood
,
A.
,
Fatma
,
M.
,
Freschi
,
L.
,
Corpas
,
F. J.
and
Khan
,
N. A.
(
2017
).
Nitric oxide signaling and its crosstalk with other plant growth regulators in plant responses to abiotic stress
.
Environ. Sci. Pollut. Res. Int.
24
,
2273
-
2285
.
Baker
,
A.
,
Graham
,
I. A.
,
Holdsworth
,
M.
,
Smith
,
S. M.
and
Theodoulou
,
F. L.
(
2006
).
Chewing the fat: beta-oxidation in signalling and development
.
Trends Plant Sci.
11
,
124
-
132
.
Baker
,
A.
,
Carrier
,
D. J.
,
Schaedler
,
T.
,
Waterham
,
H. R.
,
van Roermund
,
C. W.
and
Theodoulou
,
F. L.
(
2015
).
Peroxisomal ABC transporters: functions and mechanism
.
Biochem. Soc. Trans.
43
,
959
-
965
.
Barroso
,
J. B.
,
Corpas
,
F. J.
,
Carreras
,
A.
,
Sandalio
,
L. M.
,
Valderrama
,
R.
,
Palma
,
J. M.
,
Lupiáñez
,
J. A.
and
del Río
,
L. A.
(
1999
).
Localization of nitric-oxide synthase in plant peroxisomes
.
J. Biol. Chem.
274
,
36729
-
36733
.
Barroso
,
J. B.
,
Valderrama
,
R.
and
Corpas
,
F. J.
(
2013
).
Immunolocalization of S-nitrosoglutathione, S-nitrosoglutathione reductase and tyrosine nitration in pea leaf organelles
.
Acta Physiol. Plant
35
,
2635
-
2640
.
Begara-Morales
,
J. C.
,
Sánchez-Calvo
,
B.
,
Chaki
,
M.
,
Valderrama
,
R.
,
Mata-Pérez
,
C.
,
Padilla
,
M. N.
,
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2016
).
Antioxidant systems are regulated by nitric oxide-mediated post-translational modifications (NO-PTMs)
.
Front. Plant Sci.
7
,
152
.
Behrends
,
W.
,
Thieringer
,
R.
,
Engeland
,
K.
,
Kunau
,
W.-H.
and
Kindl
,
H.
(
1988
).
The glyoxysomal beta-oxidation system in cucumber seedlings: identification of enzymes required for the degradation of unsaturated fatty acids
.
Arch. Biochem. Biophys.
263
,
170
-
177
.
Bernhardt
,
K.
,
Wilkinson
,
S.
,
Weber
,
A. P. M.
and
Linka
,
N.
(
2012
).
A peroxisomal carrier delivers NAD+ and contributes to optimal fatty acid degradation during storage oil mobilization
.
Plant J.
69
,
1
-
13
.
Cao
,
J.
,
Li
,
M.
,
Chen
,
J.
,
Liu
,
P.
and
Li
,
Z.
(
2016
).
Effects of MeJA on Arabidopsis metabolome under endogenous JA deficiency
.
Sci. Rep.
6
,
37674
.
Chai
,
M.-F.
,
Wei
,
P.-C.
,
Chen
,
Q.-J.
,
An
,
R.
,
Chen
,
J.
,
Yang
,
S.
and
Wang
,
X. C.
(
2006
).
NADK3, a novel cytoplasmic source of NADPH, is required under conditions of oxidative stress and modulates abscisic acid responses in Arabidopsis
.
Plant J.
47
,
665
-
674
.
Chaki
,
M.
,
Álvarez de Morales
,
P.
,
Ruiz
,
C.
,
Begara-Morales
,
J. C.
,
Barroso
,
J. B.
,
Corpas
,
F. J.
and
Palma
,
J. M.
(
2015
).
Ripening of pepper (Capsicum annuum) fruit is characterized by an enhancement of protein tyrosine nitration
.
Ann. Bot.
116
,
637
-
647
.
Chigri
,
F.
,
Flosdorff
,
S.
,
Pilz
,
S.
,
Kölle
,
E.
,
Dolze
,
E.
,
Gietl
,
C.
and
Vothknecht
,
U. C.
(
2012
).
The Arabidopsis calmodulin-like proteins AtCML30 and AtCML3 are targeted to mitochondria and peroxisomes, respectively
.
Plant Mol. Biol.
78
,
211
-
222
.
Clapham
,
D. E.
(
2007
).
Calcium signaling
.
Cell
131
,
1047
-
1058
.
Coca
,
M.
and
San Segundo
,
B.
(
2010
).
AtCPK1 calcium-dependent protein kinase mediates pathogen resistance in Arabidopsis
.
Plant J.
63
,
526
-
540
.
Corpas
,
F. J.
(
2015
).
Peroxisomes: dynamic shape-shifters
.
Nat. Plants
1
,
15039
.
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2014a
).
Peroxynitrite (ONOO-) is endogenously produced in arabidopsis peroxisomes and is overproduced under cadmium stress
.
Ann. Bot.
113
,
87
-
96
.
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2014b
).
Peroxisomal plant nitric oxide synthase (NOS) protein is imported by peroxisomal targeting signal type 2 (PTS2) in a process that depends on the cytosolic receptor PEX7 and calmodulin
.
FEBS Lett.
588
,
2049
-
2054
.
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2014c
).
NADPH-generating dehydrogenases: their role in the mechanism of protection against nitro-oxidative stress induced by adverse environmental conditions
.
Front. Environ. Sci.
2
,
55
.
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2016
).
Nitric oxide synthase-like activity in higher plants
.
Nitric Oxide
.
2016
,
S1089-8603(16)30215-4
.
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2017a
).
Lead-induced stress, which triggers the production of nitric oxide (NO) and superoxide anion (O2·-) in Arabidopsis peroxisomes, affects catalase activity
.
Nitric Oxide Biol. Chem.
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2017b
).
Calmodulin (CaM) antagonist affects peroxisomal functionality by disrupting both peroxisomal Ca2+ and protein import
.
J. Cell Sci.
Corpas
,
F. J.
,
Barroso
,
J. B.
,
Sandalio
,
L. M.
,
Distefano
,
S.
,
Palma
,
J. M.
,
Lupiáñez
,
J. A.
and
del Río
,
L. A.
(
1998
).
A dehydrogenase-mediated recycling system of NADPH in plant peroxisomes
.
Biochem. J.
330
,
777
-
784
.
Corpas
,
F. J.
,
Barroso
,
J. B.
,
Carreras
,
A.
,
Quirós
,
M.
,
León
,
A. M.
,
Romero-Puertas
,
M. C.
,
Esteban
,
F. J.
,
Valderrama
,
R.
,
Palma
,
J. M.
,
Sandalio
,
L. M.
,
Gómez
,
M.
and
del Río
,
L. A.
(
2004
).
Cellular and subcellular localization of endogenous nitric oxide in young and senescent pea plants
.
Plant Physiol.
136
,
2722
-
2733
.
Corpas
,
F. J.
,
Hayashi
,
M.
,
Mano
,
S.
,
Nishimura
,
M.
and
Barroso
,
J. B.
(
2009
).
Peroxisomes are required for in vivo nitric oxide accumulation in the cytosol following salinity stress of Arabidopsis plants
.
Plant Physiol.
151
,
2083
-
2094
.
Corpas
,
F. J.
,
Barroso
,
J. B.
,
Sandalio
,
L. M.
,
Palma
,
J. M.
,
Lupiáñez
,
J. A.
and
del Río
,
L. A.
(
1999
).
Peroxisomal NADP-dependent isocitrate dehydrogenase. characterization and activity regulation during natural senescence
.
Plant Physiol.
121
,
921
-
928
.
Corpas
,
F. J.
,
Sandalio
,
L. M.
,
Brown
,
M. J.
,
del Río
,
L. A.
and
Trelease
,
R. N.
(
2000
).
Identification of porin-like polypeptide(s) in the boundary membrane of oilseed glyoxysomes
.
Plant Cell Physiol.
41
,
1218
-
1228
.
Corpas
,
F. J.
,
Aguayo-Trinidad
,
S.
,
Ogawa
,
T.
,
Yoshimura
,
K.
and
Shigeoka
,
S.
(
2016
).
Activation of NADPH-recycling systems in leaves and roots of Arabidopsis thaliana under arsenic-induced stress conditions is accelerated by knock-out of Nudix hydrolase 19 (AtNUDX19) gene
.
J. Plant Physiol.
192
,
81
-
89
.
Corpas
,
F. J.
,
Barroso
,
J. B.
,
Palma
,
J. M.
and
Rodriguez-Ruiz
,
M.
(
2017
).
Plant Peroxisomes: a nitro-oxidative cocktail
.
Redox Biol.
11
,
535
-
542
.
Costa
,
A.
,
Drago
,
I.
,
Behera
,
S.
,
Zottini
,
M.
,
Pizzo
,
P.
,
Schroeder
,
J. I.
,
Pozzan
,
T.
and
Lo Schiavo
,
F.
(
2010
).
H2O2 in plant peroxisomes: an in vivo analysis uncovers a Ca(2+)-dependent scavenging system
.
Plant J.
62
,
760
-
772
.
Cross
,
L. L.
,
Ebeed
,
H. T.
and
Baker
,
A.
(
2016
).
Peroxisome biogenesis, protein targeting mechanisms and PEX gene functions in plants
.
Biochim. Biophys. Acta
1863
,
850
-
862
.
Cui
,
S.
,
Hayashi
,
Y.
,
Otomo
,
M.
,
Mano
,
S.
,
Oikawa
,
K.
,
Hayashi
,
M.
and
Nishimura
,
M.
(
2016
).
Sucrose production mediated by lipid metabolism suppresses the physical interaction of peroxisomes and oil bodies during germination of Arabidopsis thaliana
.
J. Biol. Chem.
291
,
19734
-
19745
.
Dammann
,
C.
,
Ichida
,
A.
,
Hong
,
B.
,
Romanowsky
,
S. M.
,
Hrabak
,
E. M.
,
Harmon
,
A. C.
,
Pickard
,
B. G.
and
Harper
,
J. F.
(
2003
).
Subcellular targeting of nine calcium-dependent protein kinase isoforms from Arabidopsis
.
Plant Physiol.
132
,
1840
-
1848
.
De Duve
,
C.
,
Beaufay
,
H.
,
Jacques
,
P.
,
Rahman-Li
,
Y.
,
Sellinger
,
O. Z.
,
Wattiaux
,
R.
and
De Coninck
,
S.
(
1960
).
Intracellular localization of catalase and of some oxidases in rat liver
.
Biochim. Biophys. Acta
40
,
186
-
187
.
Delker
,
C.
,
Zolman
,
B. K.
,
Miersch
,
O.
and
Wasternack
,
C.
(
2007
).
Jasmonate biosynthesis in Arabidopsis thaliana requires peroxisomal beta-oxidation enzymes-additional proof by properties of pex6 and aim1
.
Phytochemistry
68
,
1642
-
1650
.
Dolze
,
E.
,
Chigri
,
F.
,
Höwing
,
T.
,
Hierl
,
G.
,
Isono
,
E.
,
Vothknecht
,
U. C.
and
Gietl
,
C.
(
2013
).
Calmodulin-like protein AtCML3 mediates dimerization of peroxisomal processing protease AtDEG15 and contributes to normal peroxisome metabolism
.
Plant Mol. Biol.
83
,
607
-
624
.
Domingos
,
P.
,
Prado
,
A. M.
,
Wong
,
A.
,
Gehring
,
C.
and
Feijo
,
J. A.
(
2015
).
Nitric oxide: a multitasked signaling gas in plants
.
Mol Plant.
8
,
506
-
520
.
Fernández-Fernández
,
Á. D.
and
Corpas
,
F. J.
(
2016
).
In silico analysis of Arabidopsis thaliana peroxisomal 6-phosphogluconate dehydrogenase
.
Scientifica (Cairo)
2016
,
3482760
.
Frederick
,
S. E.
,
Newcomb
,
E. H.
,
Vigil
,
E. L.
and
Wergin
,
W. P.
(
1968
).
Fine-structural characterization of plant microbodies
.
Planta
81
,
229
-
252
.
Freeman
,
B. A.
,
Baker
,
P. R. S.
,
Schopfer
,
F. J.
,
Woodcock
,
S. R.
,
Napolitano
,
A.
and
d'Ischia
,
M.
(
2008
).
Nitro-fatty acid formation and signaling
.
J. Biol. Chem.
283
,
15515
-
15519
.
Freschi
,
L.
(
2013
).
Nitric oxide and phytohormone interactions: current status and perspectives
.
Front. Plant Sci.
4
,
398
.
Gfeller
,
A.
,
Dubugnon
,
L.
,
Liechti
,
R.
and
Farmer
,
E. E.
(
2010
).
Jasmonate biochemical pathway
.
Sci. Signal.
3
,
cm3
.
Gincel
,
D.
,
Zaid
,
H.
and
Shoshan-Barmatz
,
V.
(
2001
).
Calcium binding and translocation by the voltage-dependent anion channel: a possible regulatory mechanism in mitochondrial function
.
Biochem. J.
358
,
147
-
155
.
Goepfert
,
S.
and
Poirier
,
Y.
(
2007
).
β-oxidation in fatty acid degradation and beyond
.
Curr. Opin. Plant Biol.
10
,
245
-
251
.
Goto-Yamada
,
S.
,
Mano
,
S.
,
Yamada
,
K.
,
Oikawa
,
K.
,
Hosokawa
,
Y.
,
Hara-Nishimura
,
I.
and
Nishimura
,
M.
(
2015
).
Dynamics of the light-dependent transition of plant peroxisomes
.
Plant Cell Physiol.
56
,
1264
-
1271
.
Graham
,
I. A.
(
2008
).
Seed storage oil mobilization
.
Annu. Rev. Plant Biol.
59
,
115
-
142
.
Guo
,
F.
,
Yoon
,
G. M.
and
McCubbin
,
A. G.
(
2013
).
PiSCP1 and PiCDPK2 localize to peroxisomes and are involved in pollen tube growth in Petunia Inflata
.
Plants (Basel)
2
,
72
-
86
.
Helm
,
M.
,
Lück,
,
C.
,
Prestele
,
J.
,
Hierl
,
G.
,
Huesgen
,
P. F.
,
Frühlich
,
T
,
Arnold
,
G. J.
,
Adamska
,
I.
,
Görg
,
A.
,
Lottspeich
,
F.
and
Gietl
,
C.
(
2007
).
Dual specificities of the glyoxysomal/peroxisomal processing protease Deg15 in higher plants
.
Proc. Natl. Acad. Sci. U S A.
104
,
11501
-
11506
.
Hepler
,
P. K.
(
2005
).
Calcium: a central regulator of plant growth and development
.
Plant Cell
17
,
2142
-
2155
.
Hölscher
,
C.
,
Lutterbey
,
M. C.
,
Lansing
,
H.
,
Meyer
,
T.
,
Fischer
,
K.
and
von Schaewen
,
A.
(
2016
).
Defects in peroxisomal 6-phosphogluconate dehydrogenase isoform PGD2 prevent gametophytic interaction in Arabidopsis thaliana
.
Plant Physiol.
171
,
192
-
205
.
Hu
,
J.
,
Baker
,
A.
,
Bartel
,
B.
,
Linka
,
N.
,
Mullen
,
R. T.
,
Reumann
,
S.
and
Zolman
,
B. K.
(
2012
).
Plant peroxisomes: biogenesis and function
.
Plant Cell.
24
,
2279
-
2303
.
Huang
,
X.
,
Stettmaier
,
K.
,
Michel
,
C.
,
Hutzler
,
P.
,
Mueller
,
M. J.
and
Durner
,
J.
(
2004
).
Nitric oxide is induced by wounding and influences jasmonic acid signaling in Arabidopsis thaliana
.
Planta
218
,
938
-
946
.
Hussain
,
A.
,
Mun
,
B. G.
,
Imran
,
Q. M.
,
Lee
,
S. U.
,
Adamu
,
T. A.
,
Shahid
,
M.
,
Kim
,
K. M.
and
Yun
,
B. W.
(
2016
).
Nitric oxide mediated transcriptome profiling reveals activation of multiple regulatory pathways in Arabidopsis thaliana
.
Front. Plant Sci.
7
,
975
.
Jammes
,
F.
,
Hu
,
H.-C.
,
Villiers
,
F.
,
Bouten
,
R.
and
Kwak
,
J. M.
(
2011
).
Calcium-permeable channels in plant cells
.
FEBS J.
278
,
4262
-
4276
.
Jiménez
,
A.
,
Hernández
,
J. A.
,
Pastori
,
G.
,
del Rio
,
L. A.
and
Sevilla
,
F.
(
1998
).
Role of the ascorbate-glutathione cycle of mitochondria and peroxisomes in the senescence of pea leaves
.
Plant Physiol.
118
,
1327
-
1335
.
Johnson
,
T. L.
and
Olsen
,
L. J.
(
2003
).
Import of the peroxisomal targeting signal type 2 protein 3-ketoacyl-coenzyme a thiolase into glyoxysomes
.
Plant Physiol.
133
,
1991
-
1999
.
Kunz
,
H. H.
,
Häusler
,
R. E.
,
Fettke
,
J.
,
Herbst
,
K.
,
Niewiadomski
,
P.
,
Gierth
,
M.
,
Bell
,
K.
,
Steup
,
M.
,
Flügge
,
U. I.
and
Schneider
,
A.
(
2010
).
The role of plastidial glucose-6-phosphate/phosphate translocators in vegetative tissues of Arabidopsis thaliana mutants impaired in starch biosynthesis
.
Plant Biol. (Stuttg.)
12
,
115
-
128
.
Kataya
,
A. R. A.
and
Reumann
,
S.
(
2010
).
Arabidopsis glutathione reductase 1 is dually targeted to peroxisomes and the cytosol
.
Plant Signal Behav.
5
,
171
-
175
.
Kataya
,
A. R. A.
,
Schei
,
E.
and
Lillo
,
C.
(
2015
).
MAP kinase phosphatase 1 harbors a novel PTS1 and is targeted to peroxisomes following stress treatments
.
J. Plant Physiol.
179
,
12
-
20
.
Koffler
,
B. E.
,
Bloem
,
E.
,
Zellnig
,
G.
and
Zechmann
,
B.
(
2013
).
High resolution imaging of subcellular glutathione concentrations by quantitative immunoelectron microscopy in different leaf areas of Arabidopsis
.
Micron
45
,
119
-
128
.
León
,
J.
(
2013
).
Role of plant peroxisomes in the production of jasmonic acid-based signals
.
Subcell. Biochem.
69
,
299
-
313
.
Leterrier
,
M.
,
Barroso
,
J. B.
,
Valderrama
,
R.
,
Begara-Morales
,
J. C.
,
Sánchez-Calvo
,
B.
,
Chaki
,
M.
,
Luque
,
F.
,
Viñegla
,
B.
,
Palma
,
J. M.
and
Corpas
,
F. J.
(
2016
).
Peroxisomal NADP-isocitrate dehydrogenase is required for Arabidopsis stomatal movement
.
Protoplasma
253
,
403
-
415
.
Li
,
W. Y.
,
Wang
,
X.
,
Li
,
R.
,
Li
,
W.-Q.
and
Chen
,
K. M.
(
2014
).
Genome-wide analysis of the NADK gene family in plants
.
PLoS ONE
9
,
e101051
.
Lingner
,
T.
,
Kataya
,
A. R.
,
Antonicelli
,
G. E.
,
Benichou
,
A.
,
Nilssen
,
K.
,
Chen
,
X. Y.
,
Siemsen
,
T.
,
Morgenstern
,
B.
,
Meinicke
,
P.
and
Reumann
,
S.
(
2011
).
Identification of novel plant peroxisomal targeting signals by a combination of machine learning methods and in vivo subcellular targeting analyses
.
Plant Cell
23
,
1556
-
1572
.
Linka
,
N.
and
Theodoulou
,
F. L.
(
2013
).
Metabolite transporters of the plant peroxisomal membrane: known and unknown
.
Subcell. Biochem.
69
,
169
-
194
.
Loughran
,
P. A.
,
Stolz
,
D. B.
,
Barrick
,
S. R.
,
Wheeler
,
D. S.
,
Friedman
,
P. A.
,
Rachubinski
,
R. A.
,
Watkins
,
S. C.
and
Billiar
,
T. R.
(
2013
).
PEX7 and EBP50 target iNOS to the peroxisome in hepatocytes
.
Nitric Oxide
31
,
9
-
19
.
Maruta
,
T.
,
Ogawa
,
T.
,
Tsujimura
,
M.
,
Ikemoto
,
K.
,
Yoshida
,
T.
,
Takahashi
,
H.
,
Yoshimura
,
K.
and
Shigeoka
,
S.
(
2016
).
Loss-of-function of an Arabidopsis NADPH pyrophosphohydrolase, AtNUDX19, impacts on the pyridine nucleotides status and confers photooxidative stress tolerance
.
Sci. Rep.
6
,
37432
.
Mast
,
F. D.
,
Rachubinski
,
R. A.
and
Aitchison
,
J. D.
(
2015
).
Signaling dynamics and peroxisomes
.
Curr. Opin. Cell Biol.
35
,
131
-
136
.
Mata-Pérez
,
C.
,
Sánchez-Calvo
,
B.
,
Padilla
,
M. N.
,
Begara-Morales
,
J. C.
,
Luque
,
F.
,
Melguizo
,
M.
,
Jiménez-Ruiz
,
J.
,
Fierro-Risco
,
J.
,
Peñas-Sanjuán
,
A.
,
Valderrama
,
R.
, et al. 
(
2016a
).
Nitro-fatty acids in plant signaling: nitro-linolenic acid induces the molecular chaperone network in Arabidopsis
.
Plant Physiol.
170
,
686
-
701
.
Mata-Pérez
,
C.
,
Sánchez-Calvo
,
B.
,
Begara-Morales
,
J. C.
,
Carreras
,
A.
,
Padilla
,
M. N.
,
Melguizo
,
M.
,
Valderrama
,
R.
,
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2016b
).
Nitro-linolenic acid is a nitric oxide donor
.
Nitric Oxide
57
,
57
-
63
.
Mata-Pérez
,
C.
,
Sánchez-Calvo
,
B.
,
Padilla
,
M. N.
,
Begara-Morales
,
J. C.
,
Valderrama
,
R.
,
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2017
).
Nitro-fatty acids in plant signaling: new key mediators of nitric oxide metabolism
.
Redox Biol.
11
,
554
-
561
.
Meng
,
W.
,
Hsiao
,
A. S.
,
Gao
,
C.
,
Jiang
,
L.
and
Chye
,
M. L.
(
2014
).
Subcellular localization of rice acyl-CoA-binding proteins (ACBPs) indicates that OsACBP6::GFP is targeted to the peroxisomes
.
New Phytol.
203
,
469
-
482
.
Mullen
,
R. T.
(
2002
).
Targeting and import of matrix proteins into peroxisomes
. In
Plant Peroxisomes, Biochemistry, Cell Biology and Biotechnological Applications
(ed.
A.
Baker
and
I. A.
Graham
), pp.
339
-
383
.
Springer
,
ISBN 978-90-481-6007-5
.
Mur
,
L. A.
,
Prats
,
E.
,
Pierre
,
S.
,
Hall
,
M. A.
and
Hebelstrup
,
K. H.
(
2013
).
Integrating nitric oxide into salicylic acid and jasmonic acid/ethylene plant defense pathways
.
Front. Plant Sci.
4
,
215
.
Murata
,
Y.
,
Mori
,
I. C.
and
Munemasa
,
S.
(
2015
).
Diverse stomatal signaling and the signal integration mechanism
.
Annu. Rev. Plant Biol.
66
,
369
-
392
.
Ogawa
,
T.
,
Yoshimura
,
K.
,
Miyake
,
H.
,
Ishikawa
,
K.
,
Daisuke
,
I.
,
Tanabe
,
N.
and
Shigeoka
,
S.
(
2008
).
Molecular characterization of organelle-type Nudix Hydrolases in Arabidopsis
.
Plant Physiol.
148
,
1412
-
1414
.
Oikawa
,
K.
,
Matsunaga
,
S.
,
Mano
,
S.
,
Kondo
,
M.
,
Yamada
,
K.
,
Hayashi
,
M.
,
Kagawa
,
T.
,
Kadota
,
A.
,
Sakamoto
,
W.
,
Higashi
,
S.
, et al. 
(
2015
).
Physical interaction between peroxisomes and chloroplasts elucidated by in situ laser analysis
.
Nat. Plants.
1
,
15035
.
Palma
,
J. M.
,
Corpas
,
F. J.
and
del Río
,
L. A.
(
2009
).
Proteome of plant peroxisomes: new perspectives on the role of these organelles in cell biology
.
Proteomics
9
,
2301
-
2312
.
Pastori
,
G. M.
and
del Rio
,
L. A.
(
1997
).
Natural senescence of pea leaves: an activated oxygen-mediated function for peroxisomes
.
Plant Physiol.
113
,
411
-
418
.
Perochon
,
A.
,
Aldon
,
D.
,
Galaud
,
J.-P.
and
Ranty
,
B.
(
2011
).
Calmodulin and calmodulin-like proteins in plant calcium signaling
.
Biochimie
93
,
2048
-
2053
.
Pollak
,
N.
,
Dölle
,
C.
and
Ziegler
,
M.
(
2007
).
The power to reduce: pyridine nucleotides-small molecules with a multitude of functions
.
Biochem. J.
402
,
205
-
218
.
Pracharoenwattana
,
I.
,
Cornah
,
J. E.
and
Smith
,
S. M.
(
2007
).
Arabidopsis peroxisomal malate dehydrogenase functions in beta-oxidation but not in the glyoxylate cycle
.
Plant J.
50
,
381
-
390
.
Prado
,
A. M.
,
Porterfield
,
D. M.
and
Feijó
,
J. A.
(
2004
).
Nitric oxide is involved in growth regulation and re-orientation of pollen tubes
.
Development
131
,
2707
-
2714
.
Prado
,
A. M.
,
Colaço
,
R.
,
Moreno
,
N.
,
Silva
,
A. C.
and
Feijó
,
J. A.
(
2008
).
Targeting of pollen tubes to ovules is dependent on nitric oxide (NO) signaling
.
Mol. Plant.
1
,
703
-
714
.
Reumann
,
S.
(
2000
).
The structural properties of plant peroxisomes and their metabolic significance
.
Biol. Chem.
381
,
639
-
648
.
Reumann
,
S.
and
Bartel
,
B.
(
2016
).
Plant peroxisomes: recent discoveries in functional complexity, organelle homeostasis, and morphological dynamics
.
Curr. Opin. Plant Biol.
34
,
17
-
26
.
Reumann
,
S.
,
Bettermann
,
M.
,
Benz
,
R.
and
Heldt
,
H. W.
(
1997
).
Evidence for the presence of a porin in the membrane of glyoxysomes of castor bean
.
Plant Physiol.
115
,
891
-
899
.
Reumann
,
S.
,
Maier
,
E.
,
Heldt
,
H. W.
and
Benz
,
R.
(
1998
).
Permeability properties of the porin of spinach leaf peroxisomes
.
Eur. J. Biochem.
251
,
359
-
366
.
Reumann
,
S.
,
Babujee
,
L.
,
Ma
,
C. L.
,
Wienkoop
,
S.
,
Siemsen
,
T.
,
Antonicelli
,
G. E.
,
Rasche
,
N.
,
Luder
,
F.
,
Weckwerth
,
W.
and
Jahn
,
O.
(
2007
).
Proteome analysis of Arabidopsis leaf peroxisomes reveals novel targeting peptides, metabolic pathways, and defense mechanisms
.
Plant Cell
19
,
3170
-
3193
.
Rhodin
,
J.
(
1954
).
Correlation of ultrastructural organization and function in normal and experimentally changed proximal convoluted tubule cells of the mouse kidney
.
Doctoral Thesis
,
Karolinska Institu
,
Aktiebolaget Godvil
,
Stochholm
.
Romero-Puertas
,
M. C.
,
Corpas
,
F. J.
,
Sandalio
,
L. M.
,
Leterrier
,
M.
,
Rodríguez-Serrano
,
M.
,
del Río
,
L. A.
and
Palma
,
J. M.
(
2006
).
Glutathione reductase from pea leaves: response to abiotic stress and characterization of the peroxisomal isozyme
.
New Phytol.
170
,
43
-
52
.
Rottensteiner
,
H.
and
Theodoulou
,
F. L.
(
2006
).
The ins and outs of peroxisomes: co-ordination of membrane transport and peroxisomal metabolism
.
Biochim. Biophys. Acta
1763
,
1527
-
1540
.
Sánchez-Calvo
,
B.
,
Barroso
,
J. B.
and
Corpas
,
F. J.
(
2013
).
Hypothesis: Nitro-fatty acids play a role in plant metabolism
.
Plant Sci.
199-200
,
1
-
6
.
Schaller
,
F.
,
Biesgen
,
C.
,
Müssig
,
C.
,
Altmann
,
T.
and
Weiler
,
E. W.
(
2000
).
12-Oxophytodienoate reductase 3 (OPR3) is the isoenzyme involved in jasmonate biosynthesis
.
Planta
210
,
979
-
984
.
Schlicht
,
M.
,
Ludwig-Müller
,
J.
,
Burbach
,
C.
,
Volkmann
,
D.
and
Baluska
,
F.
(
2013
).
Indole-3-butyric acid induces lateral root formation via peroxisome-derived indole-3-acetic acid and nitric oxide
.
New Phytol.
200
,
473
-
482
.
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
.
Simontacchi
,
M.
,
Galatro
,
A.
,
Ramos-Artuso
,
F.
and
Santa-María
,
G. E.
(
2015
).
Plant survival in a changing environment: the role of nitric oxide in plant responses to abiotic stress
.
Front. Plant Sci.
6
,
977
.
Spiess
,
G. M.
and
Zolman
,
B. K.
(
2013
).
Peroxisomes as a source of auxin signaling molecules
.
Subcell. Biochem.
69
,
257
-
281
.
Stolz
,
D. B.,
,
Zamora
,
R.
,
Vodovotz
,
Y.
,
Loughran
,
P. A.
,
Billiar
,
T. R.
,
Kim
,
Y. M.
,
Simmons
,
R. L.
and
Watkins
,
S. C.
(
2002
).
Peroxisomal localization of inducible nitric oxide synthase in hepatocytes
.
Hepatology
36
,
81
-
93
.
Strassner
,
J.
,
Schaller
,
F.
,
Frick
,
U. B.
,
Howe
,
G. A.
,
Weiler
,
E. W.
,
Amrhein
,
N.
,
Macheroux
,
P.
and
Schaller
,
A.
(
2002
).
Characterization and cDNA-microarray expression analysis of 12-oxophytodienoate reductases reveals differential roles for octadecanoid biosynthesis in the local versus the systemic wound response
.
Plant J.
32
,
585
-
601
.
Tani
,
T.
,
Sobajima
,
H.
,
Okada
,
K.
,
Chujo
,
T.
,
Arimura
,
S.
,
Tsutsumi
,
N.
,
Nishimura
,
M.
,
Seto
,
H.
,
Nojiri
,
H.
and
Yamane
,
H.
(
2008
).
Identification of the OsOPR7 gene encoding 12-oxophytodienoate reductase involved in the biosynthesis of jasmonic acid in rice
.
Planta
227
,
517
-
526
.
Theodoulou
,
F. L.
,
Job
,
K.
,
Slocombe
,
S. P.
,
Footitt
,
S.
,
Holdsworth
,
M.
,
Baker
,
A.
,
Larson
,
T. R.
and
Graham
,
I. A.
(
2005
).
Jasmonoic acid levels are reduced in COMATOSE ATP-binding cassette transporter mutants: implications for transport of jasmonate precursors into peroxisomes
.
Plant Physiol.
137
,
835
-
840
.
Tolbert
,
N. E.
,
Oeser
,
A.
,
Kisaki
,
T.
,
Hageman
,
R. H.
and
Yamazaki
,
R. K.
(
1968
).
Peroxisomes from spinach leaves containing enzymes related to glycolate metabolism
.
J. Biol. Chem.
243
,
5179
-
5184
.
Trostchansky
,
A.
,
Bonilla
,
L.
,
González-Perilli
,
L.
and
Rubbo
,
H.
(
2013
).
Nitro-fatty acids: formation, redox signaling, and therapeutic potential
.
Antioxid Redox Signal.
19
,
1257
-
1265
.
Turner
,
W. L.
,
Waller
,
J. C.
and
Snedden
,
W. A.
(
2005
).
Identification, molecular cloning and functional characterization of a novel NADH kinase from Arabidopsis thaliana (thale cress)
.
Biochem. J.
385
,
217
-
223
.
van der Zand
,
A.
,
Braakman
,
I.
,
Geuze
,
H. J.
and
Tabak
,
H. F.
(
2006
).
The return of the peroxisome
.
J. Cell Sci.
119
,
989
-
994
.
van Roermund
,
C. W. T.
,
Schroers
,
M. G.
,
Wiese
,
J.
,
Facchinelli
,
F.
,
Kurz
,
S.
,
Wilkinson
,
S.
,
Charton
,
L.
,
Wanders
,
R. J. A.
,
Waterham
,
H. R.
,
Weber
,
A. P. M.
, et al. 
(
2016
).
The peroxisomal NAD carrier from Arabidopsis imports NAD in exchange with AMP
.
Plant Physiol.
171
,
2127
-
2139
.
Waller
,
J. C.
,
Dhanoa
,
P. K.
,
Schumann
,
U.
,
Mullen
,
R. T.
and
Snedden
,
W. A.
(
2010
).
Subcellular and tissue localization of NAD kinases from Arabidopsis: compartmentalization of de novo NADP biosynthesis
.
Planta
231
,
305
-
317
.
Wanders
,
R. J.
,
Waterham
,
H. R.
and
Ferdinandusse
,
S.
(
2016
).
Metabolic interplay between peroxisomes and other subcellular organelles including mitochondria and the endoplasmic reticulum
.
Front. Cell Dev. Biol.
3
,
83
.
Wasternack
,
C.
and
Song
,
S.
(
2017
).
Jasmonates: biosynthesis, metabolism, and signaling by proteins activating and repressing transciption
.
J. Exp. Bot.
68
,
1303
-
1321
.
Yang
,
T.
and
Poovaiah
,
B. W.
(
2002
).
Hydrogen peroxide homeostasis: activation of plant catalase by calcium/calmodulin
.
Proc. Natl. Acad. Sci. USA
99
,
4097
-
4102
.
Yoon
,
G. M.
,
Dowd
,
P. E.
,
Gilroy
,
S.
and
McCubbin
,
A. G.
(
2006
).
Calcium-dependent protein kinase isoforms in Petunia have distinct functions in pollen tube growth, including regulating polarity
.
Plant Cell.
18
,
867
-
878
.
Yoshimura
,
K.
and
Shigeoka
,
S.
(
2015
).
Versatile physiological functions of the Nudix hydrolase family in Arabidopsis
.
Biosci. Biotechnol. Biochem.
79
,
354
-
366
.
Zechmann
,
B.
(
2014
).
Compartment-specific importance of glutathione during abiotic and biotic stress
.
Front. Plant Sci.
5
,
566
.
Zhou
,
J.
,
Jia
,
F.
,
Shao
,
S.
,
Zhang
,
H.
,
Li
,
G.
,
Xia
,
X.
,
Zhou
,
Y.
,
Yu
,
J.
and
Shi
,
K.
(
2015
).
Involvement of nitric oxide in the jasmonate-dependent basal defense against root-knot nematode in tomato plants
.
Front. Plant Sci
.
6
,
193
.
Zimmermann
,
P.
,
Heinlein
,
C.
,
Orendi
,
G.
and
Zentgraf
,
U.
(
2006
).
Senescence-specific regulation of catalases in Arabidopsis thaliana (L.) Heynh
.
Plant Cell Environ.
29
,
1049
-
1060
.
Zolman
,
B. K.
,
Martinez
,
N.
,
Millius
,
A.
,
Adham
,
A. R.
and
Bartel
,
B.
(
2008
).
Identification and characterization of Arabidopsis indole-3-butyric acid response mutants defective in novel peroxisomal enzymes
.
Genetics
180
,
237
-
251
.

Competing interests

The authors declare no competing or financial interests.