Ca2+ is a second messenger in many physiological and phytopathological processes. Peroxisomes are subcellular compartments with an active oxidative and nitrosative metabolism. Previous studies have demonstrated that peroxisomal nitric oxide (NO) generation is dependent on Ca2+ and calmodulin (CaM). Here, we used Arabidopsis thaliana transgenic seedlings expressing cyan fluorescent protein (CFP) containing a type 1 peroxisomal-targeting signal motif (PTS1; CFP–PTS1), which enables peroxisomes to be visualized in vivo, and also used a cell-permeable fluorescent probe for Ca2+. Analysis by confocal laser-scanning microscopy (CLSM) enabled us to visualize the presence of endogenous Ca2+ in the peroxisomes of both roots and guard cells. The presence of Ca2+ in peroxisomes and the import of CFP–PTS1 are drastically disrupted by both CaM antagonist and glutathione (GSH). Furthermore, the activity of three peroxisomal enzymes (catalase, glycolate oxidase and hydroxypyruvate reductase) containing PTS1 was clearly affected in these conditions, with a decrease of between 41 and 51%. In summary, data show that Ca2+ and CaM are strictly necessary for protein import and normal functionality of peroxisomal enzymes, including antioxidant and photorespiratory enzymes, as well as for NO production.

Ca2+ is a cellular second messenger whose homeostasis is highly regulated in cells through different components including Ca2+ pumps, channels, cation exchangers and Ca2+-binding proteins, since this cation regulates many cellular activities (Choi et al., 2012, 2014; Charpentier and Oldroyd, 2013; Cheval et al., 2013; Steinhorst and Kudla, 2013; Simeunovic et al., 2016). Moreover, crosstalk among Ca2+ and other signaling molecules, including H2O2 and nitric oxide (NO), which affects gene expression and Ca2+-dependent protein kinase, has also been described (Rentel and Knight, 2004; González et al., 2012; Han et al., 2014; Niu and Liao, 2016). Ca2+ signals can remain local or be propagated throughout an entire cell within milliseconds to seconds (Monshausen, 2012). Ca2+ has been detected in almost all cellular compartments including the cytosol, chloroplasts, mitochondria and nucleus. But, there can be a 10,000-fold difference between cytoplasmic and non-cytoplasmic Ca2+ concentrations, which allows the formation of Ca2+ signals by rapid alterations of cytoplasmic Ca2+ levels via membrane-localized Ca2+-permeable channels (Stael et al., 2012). However, information regarding Ca2+ in peroxisomes is scarce; to our knowledge, only one study has claimed that it is present in plants (Costa et al., 2010). Nevertheless, its potential function in peroxisomes has been analysed, as Ca2+ is necessary for key antioxidant peroxisomal enzymes, such as catalase (Yang and Poovaiah, 2002; Schmidt et al., 2006; Costa et al., 2013), and for peroxisomal NO generation (Corpas et al., 2009; Corpas and Barroso, 2014b).

Peroxisomes are single-membrane-bound subcellular compartments present in almost all types of eukaryotic cells. They contain catalase and H2O2-producing flavin oxidases as indispensable enzymatic elements (Reumann et al., 2007; Palma et al., 2009; Corpas, 2015). These organelles are characterized by metabolic flexibility, as their enzymatic content can vary according to the organism (i.e. animal, plant or yeast), cell or tissue type (i.e. liver, root or leaf), development stage and external environmental conditions (Mullen et al., 2001; del Río et al., 2002; Hayashi and Nishimura, 2006; Pracharoenwattana and Smith, 2008; Hu et al., 2012).

In plant cells, peroxisomes house a large number of antioxidative enzymes, such as catalase, superoxide dismutase, components of the ascorbate–glutathione cycle and several NADP dehydrogenases, which are involved in different functions (del Río et al., 2002; Fernández-Fernández and Corpas, 2016; Hölscher et al., 2016; Leterrier et al., 2016; Corpas et al., 2017). On the other hand, accumulating data have shown that plant peroxisomes have the capacity to generate NO through a L-arginine-dependent nitric oxide synthase (NOS) activity, which strictly depends on NADPH and requires calmodulin (CaM) and Ca2+ (Barroso et al., 1999; Corpas et al., 2004). Furthermore, peroxisomal NO, together with other related reactive nitrogen species (RNS) such as peroxynitrite, has been shown to participate in the response to abiotic stresses such as salinity (Corpas et al., 2009) and heavy metals such as cadmium or lead (Corpas and Barroso, 2014a, 2016). In addition, it has been demonstrated that the protein responsible for generating NO appears to be imported by a type 2 peroxisomal-targeting signal (PTS2) in a process that depends on the cytosolic receptor PEX7, and CaM and Ca2+ (Corpas and Barroso, 2014b).

With the aim of gaining a deeper understanding of cross-talk between Ca2+ and the peroxisomal NO metabolism, a pharmacological study was carried out using different NO donors, glutathione (GSH) and a CaM antagonist in Arabidopsisthaliana seedlings. The data show that Ca2+ is present in the peroxisomes of Arabidopsis stomata and root cells, and its presence is drastically disrupted by a CaM antagonist, which also blocks protein import into peroxisomes. Furthermore, the activity of typical peroxisomal enzymes that containing a type 1 PTS motif (PTS1), such as catalase, glycolate oxidase and hydroxypyruvate reductase, was negatively affected by the CaM antagonist, suggesting that CaM must also be necessary for plant peroxisome functionality.

Location of Ca2+ in the peroxisomes of roots and guard cells as analysed by CLSM

Using Arabidopsis seedlings expressing CFP containing a PTS1 motif (CFP–PTS1), the potential endogenous presence of Ca2+ in peroxisomes was studied by using the fluorescence probe Fluo-3 AM. Although, theoretically, there is no overlap between the excitation and emission wavelengths of CFP with the specific fluorescent probe used to detect Ca2+ (Fluo-3 AM), potential overlap was evaluated at the experimental level. Fig. S1A,D shows in vivo confocal laser-scanning microscopy (CLSM) visualization of peroxisomes in the guard cells and root tips of transgenic Arabidopsis seedlings expressing CFP–PTS1 (excitation maximum at 458 nm and emission maximum at 475 nm). The peroxisomes appeared in the form of spherical spots in the cells. Fig. S1B,E, shows the same field observed by CLSM using an excitation wavelength of 514 nm and an emission wavelength of 610 nm but in the absence of the fluorescent probe (Fluo-3 AM) used to detect Ca2+ in which it is not possible to detect any fluorescence signal.

Fig. 1 shows in vivo CLSM visualization of peroxisomes, Ca2+ and chloroplasts in the guard and root cells of transgenic Arabidopsis seedlings expressing CFP–PTS1. Peroxisomes appeared in the form of spherical spots of a green color in guard and root cells (Fig. 1A,F, respectively). On the other hand, Fig. 1B,G show the same fields analyzed using Fluo-3 AM as fluorescence probe, which also enabled Ca2+ (red color) to be detected. The red fluorescence appeared in both types of spherical spots, with a pattern similar to that of CFP–PTS1, and also throughout the cells although with less intensity. Fig. 1C shows chlorophyll autofluorescence (purple color), enabling us to detect the localization of chloroplasts as spherical spots in guard cells; however, these were larger and in a different position in the cells from the spots corresponding to peroxisomes (Fig. 1A). Fig. 1D,H contains merged images of the overlap of the corresponding panels, showing a complete overlap of the punctate patterns of peroxisomes with the Ca2+ signals; in addition, the presence of the red color was localized throughout the cytosol, as well as being observed in other parts both inside and between the cells, for example, in the cell wall of roots cells (Fig. 1G) and colocalized with chloroplasts (see Fig. 2B,C). Fig. 1E,I shows the bright field of the guard and root cells, respectively. Apparently, peroxisomes have a high Ca2+ content in relation to rest of the cells, and this seems to be logical considering that the concentration of cytosolic Ca2+ usually is lower (around 100 nM) in comparison with the non-cytoplamisc Ca2+, which has a millimolar concentration (Monshausen et al., 2008; Stael et al., 2012). However, the Ca2+ signal has a wide distribution, such that it could be observed in images taken at a lower magnification. For example, Fig. S2 shows root cells taken at a lower magnification, illustrating the general distribution of Ca2+ signal using Fluo-3 AM as fluorescence probe.

Fig. 1.

In vivo detection of Ca2+, peroxisomes and chloroplasts in Arabidopsis guard and root cells. Representative CLSM images of (A–E) guard cells and (F–I) root cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1. Ca2+ is shown in red, peroxisomes in green and chloroplasts in purple. (A,F) Punctuate structures (green) visualized through CFP–PTS1 fluorescence, indicating the localization of peroxisomes. (B,G) Fluorescence (red) attributable to the detection of Ca2+ in the same cell area. (C) Chlorophyll autofluorescence (purple) demonstrating location of chloroplasts. (D,H) Merged images of A–C and F–H, respectively. (E,I) Bright-field image of guard and root cells. Ca2+ (red color) was detected by using 5 µM Fluo-3 AM. Arrows indicate representative punctuate spots corresponding to the colocalization of Ca2+ with peroxisomes. Asterisks indicate localization of Ca2+ in the cytosol or cell wall.

Fig. 1.

In vivo detection of Ca2+, peroxisomes and chloroplasts in Arabidopsis guard and root cells. Representative CLSM images of (A–E) guard cells and (F–I) root cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1. Ca2+ is shown in red, peroxisomes in green and chloroplasts in purple. (A,F) Punctuate structures (green) visualized through CFP–PTS1 fluorescence, indicating the localization of peroxisomes. (B,G) Fluorescence (red) attributable to the detection of Ca2+ in the same cell area. (C) Chlorophyll autofluorescence (purple) demonstrating location of chloroplasts. (D,H) Merged images of A–C and F–H, respectively. (E,I) Bright-field image of guard and root cells. Ca2+ (red color) was detected by using 5 µM Fluo-3 AM. Arrows indicate representative punctuate spots corresponding to the colocalization of Ca2+ with peroxisomes. Asterisks indicate localization of Ca2+ in the cytosol or cell wall.

Fig. 2.

In vivo detection of Ca2+, peroxisomes and chloroplasts in Arabidopsis guard cells upon pre-incubation with different chemicals. Representative CLSM images of guard cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1 pre-incubated with different chemicals visualized as in Fig. 1. Ca2+ is shown in red, peroxisomes in green and chloroplasts in purple. Control plants (A–E) and plants treated with (F–J) 2 mM S-nitrosoglutathione (GSNO) as an NO donor, (K–O) 2 mM DEA NoNoate as an NO donor, (P–T) 2 mM reduced glutathione (GSH) or (U–Y) 300 µM of the CaM antagonist TFP, are shown. Arrows indicate punctuate spots corresponding to the colocalization of Ca2+ with peroxisomes and angles (^) indicate colocalization of Ca2+ with chloroplasts.

Fig. 2.

In vivo detection of Ca2+, peroxisomes and chloroplasts in Arabidopsis guard cells upon pre-incubation with different chemicals. Representative CLSM images of guard cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1 pre-incubated with different chemicals visualized as in Fig. 1. Ca2+ is shown in red, peroxisomes in green and chloroplasts in purple. Control plants (A–E) and plants treated with (F–J) 2 mM S-nitrosoglutathione (GSNO) as an NO donor, (K–O) 2 mM DEA NoNoate as an NO donor, (P–T) 2 mM reduced glutathione (GSH) or (U–Y) 300 µM of the CaM antagonist TFP, are shown. Arrows indicate punctuate spots corresponding to the colocalization of Ca2+ with peroxisomes and angles (^) indicate colocalization of Ca2+ with chloroplasts.

Exogenous application of NO, GSH and CaM antagonist – effect on peroxisomal Ca2+ and import of CFP–PTS1 into peroxisomes

Previous data have demonstrated that peroxisomal NO generation depends on Ca2+ (Corpas et al., 2009) and that the import of the peroxisomal NOS-like protein responsible for this NO generation also depends on Ca2+ and CaM (Corpas and Barroso, 2014b). With the aim of gaining a greater insight into the relationship between peroxisomal NO and Ca2+ metabolism, a pharmacological study was carried out using different NO donors (GSNO and DEA NONOate) and the CaM antagonist trifluoperazine (TFP). These analyses were performed in both the guard and root cells of transgenic Arabidopsis seedlings expressing CFP–PTS1 (Figs 2 and 3, respectively).

Fig. 3.

In vivo detection of Ca2+ and peroxisomes in Arabidopsis root tip cells upon pre-incubation with different chemicals. Representative CLSM images of root tip cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1 pre-incubated with different chemicals, visualized as in Fig. 1. Ca2+ is shown in red and peroxisomes in green. Control plants (A–D) and plants treated with (E–H) 2 mM S-nitrosoglutathione (GSNO) as an NO donor, (I–L) 2 mM DEA NoNoate as an NO donor, (M–P) 2 mM reduced glutathione (GSH) or (Q–T) 300 µM of the calmodulin antagonist TFP, are shown. Arrows indicate representative punctuate spots corresponding to the colocalization of Ca2+ with peroxisomes.

Fig. 3.

In vivo detection of Ca2+ and peroxisomes in Arabidopsis root tip cells upon pre-incubation with different chemicals. Representative CLSM images of root tip cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1 pre-incubated with different chemicals, visualized as in Fig. 1. Ca2+ is shown in red and peroxisomes in green. Control plants (A–D) and plants treated with (E–H) 2 mM S-nitrosoglutathione (GSNO) as an NO donor, (I–L) 2 mM DEA NoNoate as an NO donor, (M–P) 2 mM reduced glutathione (GSH) or (Q–T) 300 µM of the calmodulin antagonist TFP, are shown. Arrows indicate representative punctuate spots corresponding to the colocalization of Ca2+ with peroxisomes.

Fig. 2A–E corresponds to the untreated Arabidopsis seedlings in which it can be observed that Ca2+ clearly colocalized with peroxisomes and chloroplasts (Fig. 2D). Fig. 2F–J shows the effect of using S-nitrosoglutathione (GSNO) as an NO donor. It can be seen that this molecule significantly affects the import of CFP–PTS1, since the green signal appears to diffuse to all the guard cells, with only very few green punctuate spots being observed (Fig. 2F), and also leads to a sharp reduction in Ca2+ content (Fig. 2G). However, the chloroplasts do not appear to be affected (Fig. 2H). Figs 2I,J shows the merged image of the corresponding panels and the bright field, respectively. To corroborate this potential effect, another NO donor, DEA NONOate, was used. In this case, a smaller number of green spots corresponding to CFP–PTS1 content was observed (Fig. 2K) in comparison to the untreated sample (Fig. 2A); in addition, the red signal corresponding to the Ca2+ was diffused to the guard cells but showed an accumulation in the inner walls of the guard cells (Fig. 2L). In this regard, it must be mentioned that NO can have a wide spectrum of actions because it can directly or indirectly modulate post-translational modifications affecting the function of the target molecules (Frungillo et al., 2014; Yun et al., 2016; Begara-Morales et al., 2014, 2015, 2016; Mata-Pérez et al., 2016).

Taking in consideration that GSNO releases both NO and glutathione (GSH), the putative effect of GSH was also assessed as an internal control. We found that the pre-treatment of seedlings with 2 mM GSH prevents the import of CFP–PTS1 since the green fluorescence signal appears to diffuse through the guard cells (Fig. 2P); the red color corresponding to the Ca2+ signal was also diffused to the guard cells but with a strong accumulation in the inner walls of the guard cells (Fig. 2Q). On the other hand, Ca2+ in the chloroplasts seems to be less affected (Fig. 2R,S). Fig. 2T shows the corresponding bright field image. CaM is a Ca2+-binding protein involved in the signaling process. The chemical TFP, a CaM antagonist (Vandonselaar et al., 1994; Sengupta et al., 2007), was therefore used to evaluate its possible effect on both peroxisomal protein import and Ca2+. Fig. 2U shows that pretreatment with TFP significantly affects the import of CFP–PTS1 since the green fluorescence signal appears to diffuse to all the guard cells. On the other hand, Fig. 2V shows that it also lowers the Ca2+ signal (red color), with very few red punctuate signals being observed, which do not seem to correspond to peroxisomes. Ca2+ signal was observed in the chloroplasts (Fig. 2W). Fig. 2X,Y show the merged image of the corresponding panels and the bright field, respectively. Thus, these results collectively suggest that NO also affects both peroxisomal protein import and endogenous Ca2+, although the effect is less strong than that seen with GSH and TFP.

Fig. 3 illustrates an analysis of root cell similar to that carried out on guard cells. Fig. 3A,B shows peroxisomes and Ca2+ detected in untreated Arabidopsis roots, in which it can be observed that Ca2+ clearly colocalizes with peroxisomes (Fig. 3C,D). Fig. 3E–H shows the impact of GSNO. Thus, it can be observed that this molecule did not affect the import of CFP–PTS1, as peroxisomes appear in the form of green spherical spots (Fig. 3E); however, the red signal (corresponding to Ca2+) appears to diffuse through the cells and also appears in a smaller number of red spherical spots (Fig. 3F). Fig. 3G,H shows the merged image of the corresponding panels and the bright field, respectively. To corroborate this potential effect, another NO donor, DEA NONOate, was used. In this case, the green color was also observed to appear in green spherical spots and also with some diffuse green signals (Fig. 3I). However, the red fluorescence signal corresponding to the Ca2+ signal appears to be only very weakly present in spherical spots (Fig. 3J). Fig. 3K,L shows the merged image of the corresponding panels and the bright field, respectively. These effects of NO donors (GSNO and DEA NONOate) in roots are quite different to that observed in guard cells (Fig. 2F,K). This could be due to the fact that guard cells are specialized cells in the epidermis of leaves involved in stomata movement, and its regulation there is dependent on a complex interplay between NO, Ca2+, H2O2, abscisic acid (ABA), salicylic acid and Ca2+-dependent protein kinases (Desikan et al., 2004; Li et al., 2009; Zhang et al., 2009; Hao et al., 2010; Wang et al., 2014; Murata et al., 2015), where peroxisomes seem also to be involved (Leterrier et al., 2016). Therefore, we hypothesize that the application of exogenous NO could alter this equilibrium, with the guard cells more affected that the root cells.

As performed above, GSH was used as an internal control for GSNO. We found that the pre-treatment of seedlings with 2 mM GSH avoids the import of CFP–PTS1 since the green signal appears to diffuse through the cells, especially in the cells of the root cap (Fig. 3M); the red color corresponding to the Ca2+ signal was also diffused through the cells in the elongation area of the primary root (Fig. 3N). Fig. 3O,P shows the merged image of the corresponding panels and the bright field, respectively. Fig. 3Q shows that pretreatment with TFP (CaM antagonist) drastically affects the import of CFP–PTS1, with the green color appearing to diffuse in the cells of the root cap. On the other hand, Fig. 3R shows a very weak red signal, indicating that the CaM antagonist very significantly affects Ca2+ content in root cells. Fig. 3S,T shows the merged image of the corresponding panels and the bright field, respectively. Fig. 4 shows the change in the fluorescence intensity of Ca2+ (red color) in guard and root cells of Arabidopsis seedlings pretreated with GSNO, DEA NONOate, GSH and TFP which corresponds to results shown in Figs 2 and 3. In this sense, it can be seen that GSH is apparently the chemical that had a stronger effect in Ca2+ content and distribution in guard and root cells.

Fig. 4.

Quantification of the relative amount of Ca2+ in guard and root cells of upon treatment with different chemicals. The fluorescence intensity of the red Fluo-3 AM signal (detecting Ca2+) in guard and root cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1 pre-incubated with different chemicals [2 mM GSNO or 2 mM DEA NoNoate (NO donors); 2 mM reduced glutathione (GSH); 300 µM TFP] as shown in the images in Figs 2 and 3. The fluorescence is expressed as mean±s.e.m. [arbitrary units (A.U.)], as determined using Image J software. *P<0.05 compared with the corresponding control for guard or root cells.

Fig. 4.

Quantification of the relative amount of Ca2+ in guard and root cells of upon treatment with different chemicals. The fluorescence intensity of the red Fluo-3 AM signal (detecting Ca2+) in guard and root cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1 pre-incubated with different chemicals [2 mM GSNO or 2 mM DEA NoNoate (NO donors); 2 mM reduced glutathione (GSH); 300 µM TFP] as shown in the images in Figs 2 and 3. The fluorescence is expressed as mean±s.e.m. [arbitrary units (A.U.)], as determined using Image J software. *P<0.05 compared with the corresponding control for guard or root cells.

Although the chemical TFP acts as a CaM antagonist (Vandonselaar et al., 1994; Sengupta et al., 2007), it might also induce non-specific effects that could not be directly related to Ca2+/CaM metabolism, such as inhibition of electron transport and phosphorylation in plant mitochondria (Dunn et al., 1984), inhibition of Ca2+ ATPase, proton excretion or membrane properties (Lichtman et al., 1982; Barr et al., 1990; Wilson, 1994; Sengupta et al., 2007). Consequently, to affirm the effect of TFP as a CaM antagonist in our experimental Arabidopsis model, another CaM antagonist was used, in particular N-(6-aminohexyl)-5-chloro-1-naphthalenesulfonamide (W7), which has been previously described to have a similar effect to TFP in animal and plant cells (Sengupta et al., 2007; Ma et al., 2008). Fig. S3 illustrates the CLSM in vivo detection of Ca2+ (red signal) and peroxisomes (green signal) in root tip cells of 5-day-old transgenic Arabidopsis seedlings expressing CFP–PTS1 pre-incubated with either of the CaM antagonists TFP (Fig. S3D–F) or W7 (Fig. S3G–I). In both cases and, in comparison with the untreated seedlings (Fig. S3A,B), the green signal appears to be diffuse in the cells of the root cap with a very weak red signal, indicating that both CaM antagonists caused identical effects. On the other hand, pre-incubation with the chemical W5 (Fig. S3J– L), which is an inactive form of W7, shows that Ca2+ clearly colocalizes with peroxisomes, similar to the localization observed in untreated seedlings (Fig. S3A–C).

Additionally, to evaluate whether the CaM antagonist TFP could affect the cell viability, Arabidopsis seedlings were stained with propidium iodide (PI). This chemical is a cationic dye that does not readily cross intact membranes but that can penetrate throughout the meristem and bind to cell walls, thereby showing an outline of living cells; PI binds to nuclear structures, causing bright punctate fluorescence. Fig. S4 shows the PI staining of root cells of 5-day-old transgenic Arabidopsis seedling expressing CFP–PTS1 that were pre-incubated with TFP. It can be observed that the CaM antagonist (Fig. S4D–F) does not affect the cell viability when compared to untreated seedlings (Fig. S4A–D).

Given that the CaM antagonist TFP is the chemical with the most drastic effect on both peroxisomal protein import and Ca2+ content, we also performed in vitro assays with this CaM antagonist to measure the activity of representative peroxisomal enzymes, including catalase, glycolate oxidase and hydroxypyruvate reductase, using 14-day-old Arabidopsis seedlings pre-incubated with 300 µM TFP for 2 h at 25°C in darkness. Fig. 5A–C shows that these three activities were considerably affected, with a reduction of 45% for catalase, 41% for glycolate oxidase and 51% for hydroxypyruvate reductase. For the purposes of characterization, we also determined whether the CaM antagonist is capable of modifying the gene expression of these peroxisomal enzymes. The Arabidopsis genome contains three catalase genes (CAT1, CAT2 and CAT3) (Du et al., 2008), whose gene expression increased ∼2.0-, 1.4- and 1.5-fold, respectively, following pretreatment of Arabidopsis seedlings with TFP for 2 h (Fig. 5D). The CaM antagonist also causes an induction of HPR1 expression of ∼2.8-fold. Finally, glycolate oxidase is encoded by five genes, three of which, GOX1, GOX2 and GOX3, were selected for analysis in this study. While GOX1 and GOX3 expression seems to be unaffected, GOX2 expression increased 1.6-fold after pretreatment with the CaM antagonist (Fig. 5E).

Fig. 5.

Activity and gene expression of peroxisomal enzymes in response to a CaM antagonist. (A–C) Peroxisomal enzymatic activity assays for (A) catalase, (B) glycolate oxidase (GOX) activity and (C) hydroxypyruvate reductase (HPR) activity. (D,E) Semi-quantitative RT-PCR analyses for (D) CAT gene expression, and (E) HPR and GOX gene expression. Semi-quantitative RT-PCR was performed on total RNA isolated using ACT2 as internal control. Representative agarose electrophoresis gels of the amplification products were visualized by GelRed™ staining under UV light. T/C indicates the relative level of the each amplification product (T) over the ACT2 internal control (C) after normalization to the control samples, and it expresses the fold change in with respect to the untreated control. In all panels, 14-day-old Arabidopsis seedlings were exposed to 300 µM TFP for 2 h at 25°C in darkness. Results are the mean±s.e.m. of three different experiments. *P<0.05 in relation to control values.

Fig. 5.

Activity and gene expression of peroxisomal enzymes in response to a CaM antagonist. (A–C) Peroxisomal enzymatic activity assays for (A) catalase, (B) glycolate oxidase (GOX) activity and (C) hydroxypyruvate reductase (HPR) activity. (D,E) Semi-quantitative RT-PCR analyses for (D) CAT gene expression, and (E) HPR and GOX gene expression. Semi-quantitative RT-PCR was performed on total RNA isolated using ACT2 as internal control. Representative agarose electrophoresis gels of the amplification products were visualized by GelRed™ staining under UV light. T/C indicates the relative level of the each amplification product (T) over the ACT2 internal control (C) after normalization to the control samples, and it expresses the fold change in with respect to the untreated control. In all panels, 14-day-old Arabidopsis seedlings were exposed to 300 µM TFP for 2 h at 25°C in darkness. Results are the mean±s.e.m. of three different experiments. *P<0.05 in relation to control values.

Ca2+ is a highly important second messenger whose presence in different plant subcellular compartments, including the cytosol, chloroplasts, mitochondria, nucleus and vacuoles, has been well studied (Raychaudhury et al., 2006; Monshausen et al., 2008; Dodd et al., 2010; Kudla et al., 2010; Behera et al., 2015; Xu et al., 2015). However, information concerning Ca2+ in peroxisomes, particularly in plants, is scarce. In a study using human HeLa cells transiently expressing a peroxisome-targeted the GFP-based Ca2+ indicator D3cpv-SKL, the presence of Ca2+ in animal peroxisomes was shown for the first time, with the peroxisomal membrane representing a significant barrier to Ca2+ diffusion into peroxisomes (Drago et al., 2008). In the same year, Lasorsa et al. (2008) performed additional experiments showing that peroxisomes contain between 20- and 50-fold more Ca2+ than the cytosol (reaching a maximum of 100 µM). To our knowledge, only one study in the literature analyses Ca2+ homeostasis in plant peroxisomes; this study used the GFP-based Cameleon Ca2+ indicator D3cpv-KVK-SKL and demonstrated that peroxisomal Ca2+ affected catalase activity (Costa et al., 2010). The present study therefore aimed to gain a deeper understanding of the significance of Ca2+ in the peroxisomal metabolism of plants.

CaM is necessary for normal peroxisomal Ca2+ homeostasis and for peroxisomal protein import

CaM is a protein of ∼17 kDa that is expressed in all eukaryotic cells and has the capacity to bind Ca2+ (Bouché et al., 2005). This protein, together with CaM-like (CML) proteins, is the primary Ca2+ sensor that controls the activity of various target proteins and consequently diverse cellular functions, including development processes and responses to adverse biotic and abiotic conditions (Yang and Poovaiah, 2003; Bouché et al., 2005; Reddy et al., 2011; Cheval et al., 2013; Zhao et al., 2015; Zeng et al., 2015). For instance, in the Arabidopsis genome, seven distinct CaM and 50 CMLs genes have been predicted (McCormack et al., 2005). Recently, Arabidopsis thaliana (At)CML3 (also known as AGD11) has been demonstrated to be targeted to peroxisomes (Chigri et al., 2012), where it appears to mediate the dimerization of peroxisomal processing protease AtDEG15 (Dolze et al., 2013). Moreover, there is another family of proteins which act as Ca2+ sensors called Ca2+-dependent protein kinases (CDPKs) (Boudsocq and Sheen, 2013) and in Arabidopsis the isoform AtCPK1 is targeted to peroxisomes (Dammann et al., 2003) and seems to mediate pathogen resistance (Coca and San Segundo, 2010).

Using an alternative fluorescent probe to detect Ca2+, the present study corroborates the presence of Ca2+ in the peroxisomes of two different tissues and cells – guard cells in green cotyledons and root cells – suggesting that Ca2+ must be necessary for the normal physiological functioning of plant peroxisomes, as has been shown for other subcellular compartments. Moreover, pharmacological techniques using different chemicals have enabled us to demonstrate that peroxisomal Ca2+ must be regulated by CaM and GSH and also, though less intensely, by NO. Thus, the CaM antagonist TFP blocks peroxisomal function due to its negative effect on peroxisomal protein import, as shown by CFP–PTS1 remaining in the cytosol. Additionally, the presence of Ca2+ in peroxisomes suggests that a CaM or CML protein is strictly necessary for the peroxisomal homeostasis of Ca2+, although, to our knowledge there is no experimental information about the potential effect of a CaM antagonist on the peroxisomal AtCML3 function. In any case, this is a reasonable assumption, since the function of CaM is to transport and regulate Ca2+ levels. Thus, these data are closely in line with previous findings that show that Ca2+ is required for peroxisomal NO production (Corpas et al., 2009) and that enzyme protein import responsible for its generation is blocked by CaM antagonist (Corpas and Barroso, 2014b).

Hitherto, there is clear evidence to support the cross-talk between NO and Ca2+/CaM in plant cells (Sang et al., 2008; Ma et al., 2012). However, in the present study, we find that NO appears to be less involved in controlling peroxisomal Ca2+ content; by contrast, peroxisomal Ca2+ is necessary and indispensable for the production of NO in peroxisomes. It has been demonstrated that the activity of peroxisomal protein L-arginine-dependent nitric oxide synthase (NOS) responsible for NO generation requires Ca2+ (Barroso et al., 1999; Corpas et al., 2004). More recently, it has been reported that the import of this protein into the peroxisome also depends on Ca2+ and CaM (Corpas et al., 2009; Corpas and Barroso, 2014b).

Reduced glutathione (GSH) constitutes an important antioxidant, and it is present in plant peroxisomes (Zechmann and Müller, 2010) where it is indispensable for the functioning of glutathione reductase (GR), a component of the ascorbate–glutathione cycle that functions in these organelles (Romero-Puertas et al., 2006; Reumann and Corpas, 2010). Under our experimental conditions, GSH was used as control during the analysis of the GSNO treatment; however, our findings were surprising as GSH, which caused a general delocalization of the Ca2+ signal and a marked intensification throughout the cells, drastically affected both peroxisomal protein import and the peroxisomal localization of Ca2+. This suggests that a change in the redox state or a possible S-glutathionylation process could be involved in the mechanism of peroxisomal protein import; the latter is regulated by a large number of cytosolic and peroxisomal proteins, known as peroxins (PEXs), which are encoded by at least 22 PEX genes in Arabidopsis (Nito et al., 2007; Brown and Baker, 2008; Hu et al., 2012).

Ca2+ deficiency negatively affects the activity of catalase and photorespiratory enzymes but increases their gene expression

As part of the potential physiological functionality of Ca2+, the CaM antagonist TFP was observed to negatively affect the activity of the antioxidant catalase enzyme, as well as the glycolate oxidase and hydroxypyruvate reductase enzymes, which are both involved in photorespiration (Rojas et al., 2012). The decrease in their activity could be due to the blocking of the peroxisomal import of the corresponding proteins containing a PTS1; this is closely in line with the results obtained, since the CaM antagonist TFP blocks the import of CFP–PTS1. However, a previous study has demonstrated that the activity of plant catalase is stimulated by both Ca2+ and CaM (Yang and Poovaiah, 2002; Afiyanti and Chen, 2014). Moreover, with the aid of a recombinant AtCat3 in a CaM-binding assay, the same authors demonstrated that CaM specifically binds to AtCat3 in the presence of Ca2+; protein sequence analysis of Cat3 specifically enabled them to identify a CaM-binding region of 37 amino acids between Gly415 and Val451 (Yang and Poovaiah, 2002). More recently, an increase in intraperoxisomal Ca2+, which occurs in response to adverse conditions accompanied by oxidative stress, has been shown to stimulate catalase activity (Costa et al., 2010). However, to our knowledge, there is no information available on any potential relationship between CaM/Ca2+ and glycolate oxidase and hydroxypyruvate reductase activity. Nevertheless, the results show that their activity was also adversely affected, with the blocking of their import into the peroxisomes being the most plausible cause. Thus, the gene expression of these enzyme groups either increased (CAT1 to CAT3, HPR1 and GOX2) or appeared to be unaffected (GOX1 and GOX3), which could be a way of compensating for the loss of these enzyme activities.

In summary, our findings indicate that the presence of Ca2+ is necessary for the physiological functioning of peroxisomes in root and guard cells, since its absence negatively affects the activity of key enzymes as well as the generation of NO, as has previously been demonstrated (Corpas et al., 2009). Moreover, we suggest that Ca2+/CaM is also necessary for the import of peroxisomal proteins containing a PTS1. However, a drastic alteration in the redox state and/or S-glutathionylation of peroxisomal protein import of peroxins could be involved. In this latter case, further analyses would be necessary to determine which peroxins are affected by this redox regulation.

Plant material and growth conditions

Arabidopsis thaliana transgenic seeds expressing cyan fluorescent protein (CFP) containing a type 1 peroxisomal-targeting signal (PTS1) (Nelson et al., 2007) were grown on Petri plates for 5 days or 14 days under a cycle of 16 h light (under a light intensity of 100 µE m−2 s−1) at 22°C and 8 h dark at 18°C as previously described (Corpas and Barroso, 2014a). For microscope analyses, 5-day-old seedlings were grown on vertically positioned plates. For enzymatic and gene expression assays, 14-day-old seedlings were grown on horizontally positioned plates.

Detection of Ca2+ in transgenic plants expressing CFP–PTS1 by CLSM

Ca2+ was detected with 5 µM Fluo-3 AM (Biotium) prepared in 10 mM Tris-HCl (pH 7.4) (Zhang et al., 1998; Kanchiswamy et al., 2010, 2014). The Arabidopsis seedlings were incubated in darkness with 5 µM Fluo-3 AM at 25°C for 1 h. Each sample was then washed twice in the same buffer for 15 min and mounted on a microscope slide for examination with a CLSM (Leica TCS SP5 II). For Fluo-3 AM, the excitation laser wavelength was 514 nm, with an emission of 610 nm and a 40-nm band pass width (590–630 nm). For CFP, the excitation laser wavelength was 458 nm, with an emission of 475 nm and a 40-nm band pass width (455–495 nm). Chlorophyll autofluorescence was studied as follows: chlorophyll a and b, excitation of 429 and 450 nm, respectively; emission of 650 and 670 nm, respectively.

Pharmacological assays of peroxisomal Ca2+

Arabidopsis seedlings (5 days old) expressing CFP–PTS1 were pre-incubated for 120 min at 25°C in darkness with different chemicals prepared in 10 mM Tris-HCl (pH 7.4) including 2 mM diethylamine NONOate (DEA NoNoate) and 2 mM S-nitrosoglutathione (GSNO), which are both NO donors (Broniowska et al., 2013; Begara-Morales et al., 2014), and 2 mM glutathione (GSH) and 300 µM trifluoperazine (TFP, a calmodulin antagonist) (Vandonselaar et al., 1994). Additionally, seedlings were also pre-incubated with 200 µM N-(6-aminohexyl)-5-chloro-1-naphthalenesulfonamide (W7), which is another CaM antagonist (Ma et al., 2008), and 200 µM N-(6-aminohexyl)-1-naphthalenesulfonamide (W5), which is the inactive W7 structural analog (Li et al., 2004). The seedlings were then incubated with 5 µM Fluo-3 AM for 1 h at 25°C in darkness to detect and visualize Ca2+ using CLSM. In all cases, the images obtained by CLSM from control and treated Arabidopsis seedlings were kept constant during the course of the experiment in order to produce comparable data. For each treatment, at least three independent experiments were made using, at least, four or five seedlings in each experiment.

Cell wall integrity

Plant cell integrity was evaluated with propidium iodide (PI) staining (Truernit and Haseloff, 2008). Arabidopsis seedlings were incubated for 5 min at room temperature in a 10 μg/ml PI solution (P4170 Sigma) in water. For PI, the excitation laser wavelength was 536 nm, with an emission of 617 nm and a 40-nm band pass width (597–637 nm).

Crude extracts of Arabidopsis seedlings

Crude extracts of 14-day-old Arabidopsis seedlings were frozen in liquid N2 and ground in a mortar with a pestle. The powder was suspended in a homogenizing medium containing 50 mM Tris-HCl (pH 7.8, ratio 1:3; w/v) containing 0.1 mM EDTA, 0.2% (v/v) Triton X-100 and 10% (v/v) glycerol. Homogenates were centrifuged at 27,000 g for 20 min, and the supernatants were used for the enzymatic assays.

Enzyme activity assay

Catalase activity (EC 1.11.1.6) was assayed by measuring the disappearance of H2O2 as described by Aebi (1984). NADH-dependent hydroxypyruvate reductase (HPR; EC 1.1.1.29) was determined according to the method used by Schwitzguébel and Siegenthaler (1984). Glycolate oxidase (GOX; EC 1.1.3.1) was assayed as previously described (Kerr and Groves, 1975), by measuring the formation of glyoxylate-phenylhydrazone.

RNA isolation and semiquantitative RT-PCR

Total RNA was extracted with Trizol according to instructions provided by Gibco BRL (Life Technologies); 5 μg of total RNA were used to produce cDNA for the reverse transcriptase (RT) reaction by adding 0.5 mM dNTPs, poly-dT23, 5x RT buffer, 40 U RNase inhibitor (Invitrogen) and 200 U Reverse Transcriptase (Thermo Fisher) in a final volume of 20 μl. The reaction was carried out at 50°C for 30 min.

Semiquantitative RT-PCR amplification of actin 2 (ACT2) cDNA from Arabidopsis was chosen as control. CAT1, CAT2, CAT3, HPR1, GOX1, GOX2 and GOX3 cDNAs were amplified by a PCR as previously described (Corpas and Barroso, 2016). PCR products were then detected after electrophoresis in 2.8% (w/v) agarose gels and by staining with GelRed™. Quantification of the bands was performed using a Gel Doc system (Bio-Rad Laboratories) coupled with a high-sensitivity charge-coupled device (CCD) camera. The T/C ratio indicates the relative level of the each amplification product (T) over the ACT2 used as internal control (C) after normalization to the control samples, and it expresses the fold change with respect to the untreated control (Marone et al., 2001).

Other assays

Protein concentration was determined with the aid of the Bio-Rad Protein Assay (Hercules, CA) using bovine serum albumin as a standard. For microscope analyses, around five to ten 5-day-old Arabidopsis seedlings were used for each experiment and treatment. For enzymatic activity and gene expression assays, ∼150 14-day-old Arabidopsis seedlings were used for each experiment. Data are means of at least three sets of independent experiments with three repetitions each. To estimate the statistical significance between means, the data was analyzed by the Student's test. Relative fluorescence was quantified by using ImageJ software.

Authors are grateful to Mr. Carmelo Ruiz-Torres for his excellent technical support. The authors thank Dr Juan D Alché (CSIC) for his advice on the analysis of images. Technical and human support provided by CICT of Universidad de Jaén (UJA, MINECO, Junta de Andalucía, FEDER) is also acknowledged. We also thank Michael O'Shea for proofreading the text.

Author contributions

Conceptualization: F.J.C; Methodology: F.J.C., J.B.B.; Formal analysis and investigation: F.J.C., J.B.B.; Writing - original draft preparation: F.J.C.; Writing - review and editing: F.J.C., J.B.B.; Funding acquisition: F.J.C., J.B.B.

Funding

Research in F.J.C. and J.B.B. laboratory is supported by a European Regional Development Fund (ERDF) grant co-financed by the Ministerio de Economía y Competitividad (projects AGL2015-65104-P and BIO2015-66390-P, respectively), and Consejería de Economía, Innovación, Ciencia y Empleo, Junta de Andalucía (groups BIO192 and BIO286) in Spain.

Aebi
,
H.
(
1984
).
Catalase in vitro
.
Methods Enzymol.
105
,
121
-
126
.
Afiyanti
,
M.
and
Chen
,
H.-J.
(
2014
).
Catalase activity is modulated by calcium and calmodulin in detached mature leaves of sweet potato
.
J. Plant Physiol.
171
,
35
-
47
.
Barr
,
R.
,
Böttger
,
M.
,
Frederick
,
L.
and
Crane
,
F. L.
(
1990
).
The effect of selected inhibitors on plasma membrane redox reactions and proton excretion by carrot cells
.
Plant Sci.
69
,
33
-
38
.
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
.
Begara-Morales
,
J. C.
,
Sánchez-Calvo
,
B.
,
Luque
,
F.
,
Leyva-Pérez
,
M. O.
,
Leterrier
,
M.
,
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2014
).
Differential transcriptomic analysis by RNA-Seq of GSNO-responsive genes between Arabidopsis roots and leaves
.
Plant Cell Physiol.
55
,
1080
-
1095
.
Begara-Morales
,
J. C.
,
Sánchez-Calvo
,
B.
,
Chaki
,
M.
,
Mata-Pérez
,
C.
,
Valderrama
,
R.
,
Padilla
,
M. N.
,
López-Jaramillo
,
J.
,
Luque
,
F.
,
Corpas
,
F. J.
and
Barroso
,
J. B.
(
2015
).
Differential molecular response of monodehydroascorbate reductase and glutathione reductase by nitration and S-nitrosylation
.
J. Exp. Bot.
66
,
5983
-
5996
.
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
.
Behera
,
S.
,
Wang
,
N.
,
Zhang
,
C.
,
Schmitz-Thom
,
I.
,
Strohkamp
,
S.
,
Schültke
,
S.
,
Hashimoto
,
K.
,
Xiong
,
L.
and
Kudla
,
J.
(
2015
).
Analyses of Ca2+ dynamics using a ubiquitin-10 promoter-driven Yellow Cameleon 3.6 indicator reveal reliable transgene expression and differences in cytoplasmic Ca2+ responses in Arabidopsis and rice (Oryza sativa) roots
.
New Phytol.
206
,
751
-
760
.
Bouché
,
N.
,
Yellin
,
A.
,
Snedden
,
W. A.
and
Fromm
,
H.
(
2005
).
Plant-specific calmodulin-binding proteins
.
Annu. Rev. Plant Biol.
56
,
435
-
466
.
Boudsocq
,
M.
and
Sheen
,
J.
(
2013
).
CDPKs in immune and stress signaling
.
Trends Plant Sci.
18
,
30
-
40
.
Broniowska
,
K. A.
,
Diers
,
A. R.
and
Hogg
,
N.
(
2013
).
S-nitrosoglutathione
.
Biochim. Biophys. Acta
1830
,
3173
-
3181
.
Brown
,
L.-A.
and
Baker
,
A.
(
2008
).
Shuttles and cycles: transport of proteins into the peroxisome matrix (review)
.
Mol. Membr. Biol.
25
,
363
-
375
.
Charpentier
,
M.
and
Oldroyd
,
G. E. D.
(
2013
).
Nuclear calcium signaling in plants
.
Plant Physiol.
163
,
496
-
503
.
Cheval
,
C.
,
Aldon
,
D.
,
Galaud
,
J.-P.
and
Ranty
,
B.
(
2013
).
Calcium/calmodulin-mediated regulation of plant immunity
.
Biochim. Biophys. Acta
1833
,
1766
-
1771
.
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
.
Choi
,
W.-G.
,
Swanson
,
S. J.
and
Gilroy
,
S.
(
2012
).
High-resolution imaging of Ca2+, redox status, ROS and pH using GFP biosensors
.
Plant J.
70
,
118
-
128
.
Choi
,
W.-G.
,
Toyota
,
M.
,
Kim
,
S.-H.
,
Hilleary
,
R.
and
Gilroy
,
S.
(
2014
).
Salt stress-induced Ca2+ waves are associated with rapid, long-distance root-to-shoot signaling in plants
.
Proc. Natl. Acad. Sci. USA
111
,
6497
-
6502
.
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
).
What is the role of hydrogen peroxide in plant peroxisomes?
Plant Biol.
17
,
1099
-
1103
.
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.
(
2016
).
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.
,
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.
, et al. 
(
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.
,
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
.
Costa
,
A.
,
Drago
,
I.
,
Zottini
,
M.
,
Pizzo
,
P.
and
Pozzan
,
T.
(
2013
).
Peroxisome Ca2+ homeostasis in animal and plant cells
.
Subcell. Biochem.
69
,
111
-
133
.
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
.
del Río
,
L. A.
,
Corpas
,
F. J.
,
Sandalio
,
L. M.
,
Palma
,
J. M.
,
Gómez
,
M.
and
Barroso
,
J. B.
(
2002
).
Reactive oxygen species, antioxidant systems and nitric oxide in peroxisomes
.
J. Exp. Bot.
53
,
1255
-
1272
.
Desikan
,
R.
,
Cheung
,
M.-K.
,
Bright
,
J.
,
Henson
,
D.
,
Hancock
,
J. T.
and
Neill
,
S. J.
(
2004
).
ABA, hydrogen peroxide and nitric oxide signalling in stomatal guard cells
.
J. Exp. Bot.
55
,
205
-
212
.
Dodd
,
A. N.
,
Kudla
,
J.
and
Sanders
,
D.
(
2010
).
The language of calcium signaling
.
Annu. Rev. Plant Biol.
61
,
593
-
620
.
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
.
Drago
,
I.
,
Giacomello
,
M.
,
Pizzo
,
P.
and
Pozzan
,
T.
(
2008
).
Calcium dynamics in the peroxisomal lumen of living cells
.
J. Biol. Chem.
283
,
14384
-
14390
.
Du
,
Y.-Y.
,
Wang
,
P.-C.
,
Chen
,
J.
and
Song
,
C.-P.
(
2008
).
Comprehensive functional analysis of the catalase gene family in Arabidopsis thaliana
.
J. Integr. Plant Biol.
50
,
1318
-
1326
.
Dunn
,
P. P. J.
,
Slabas
,
A. R.
,
Cottingham
,
I. R.
and
Moore
,
A. L.
(
1984
).
Trifluoperazine inhibition of electron transport and adenosine triphosphatase in plant mitochondria
.
Arch. Biochem. Biophys.
229
,
287
-
294
.
Fernández-Fernández
,
Á. D.
and
Corpas
,
F. J.
(
2016
).
In silico analysis of Arabidopsis thaliana peroxisomal 6-Phosphogluconate dehydrogenase
.
Scientifica
2016
,
3482760
.
Frungillo
,
L.
,
Skelly
,
M. J.
,
Loake
,
G. J.
,
Spoel
,
S. H.
and
Salgado
,
I.
(
2014
).
S-nitrosothiols regulate nitric oxide production and storage in plants through the nitrogen assimilation pathway
.
Nat. Commun.
5
,
5401
.
González
,
A.
,
Cabrera
,
M. D. L. A.
,
Henríquez
,
M. J.
,
Contreras
,
R. A.
,
Morales
,
B.
and
Moenne
,
A.
(
2012
).
Cross talk among calcium, hydrogen peroxide, and nitric oxide and activation of gene expression involving calmodulins and calcium-dependent protein kinases in Ulva compressa exposed to copper excess
.
Plant Physiol.
158
,
1451
-
1462
.
Han
,
S.
,
Wang
,
C.-W.
,
Wang
,
W.-L.
and
Jiang
,
J.
(
2014
).
Mitogen-activated protein kinase 6 controls root growth in Arabidopsis by modulating Ca2+-based Na+ flux in root cell under salt stress
.
J. Plant Physiol.
171
,
26
-
34
.
Hao
,
F.
,
Zhao
,
S.
,
Dong
,
H.
,
Zhang
,
H.
,
Sun
,
L.
and
Miao
,
C.
(
2010
).
Nia1 and Nia2 are involved in exogenous salicylic acid-induced nitric oxide generation and stomatal closure in Arabidopsis
.
J. Integr. Plant Biol.
52
,
298
-
307
.
Hayashi
,
M.
and
Nishimura
,
M
. (
2006
).
Arabidopsis thaliana: a model organism to study plant peroxisomes
.
Biochim Biophys Acta.
1763
,
1382
-
1391
.
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
.
Kanchiswamy
,
C. N.
,
Takahashi
,
H.
,
Quadro
,
S.
,
Maffei
,
M. E.
,
Bossi
,
S.
,
Bertea
,
C.
,
Zebelo
,
S. A.
,
Muroi
,
A.
,
Ishihama
,
N.
,
Yoshioka
,
H.
, et al. 
(
2010
).
Regulation of Arabidopsis defense responses against Spodoptera littoralis by CPK-mediated calcium signaling
.
BMC Plant Biol.
10
,
97
.
Kanchiswamy
,
C. N.
,
Malnoy
,
M.
,
Occhipinti
,
A.
and
Maffei
,
M. E.
(
2014
).
Calcium imaging perspectives in plants
.
Int. J. Mol. Sci.
15
,
3842
-
3859
.
Kerr
,
M. W.
and
Groves
,
D.
(
1975
).
Purification and properties of glycollate oxidase from Pisum sativum leaves
.
Phytochemistry
14
,
359
-
362
.
Kudla
,
J.
,
Batistic
,
O.
and
Hashimoto
,
K.
(
2010
).
Calcium signals: the lead currency of plant information processing
.
Plant Cell
22
,
541
-
563
.
Lasorsa
,
F. M.
,
Pinton
,
P.
,
Palmieri
,
L.
,
Scarcia
,
P.
,
Rottensteiner
,
H.
,
Rizzuto
,
R.
and
Palmieri
,
F.
(
2008
).
Peroxisomes as novel players in cell calcium homeostasis
.
J. Biol. Chem.
283
,
15300
-
15308
.
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
,
B.
,
Liu
,
H. T.
,
Sun
,
D. Y.
and
Zhou
,
R. G.
(
2004
).
Ca(2+) and calmodulin modulate DNA-binding activity of maize heat shock transcription factor in vitro
.
Plant Cell Physiol.
45
,
627
-
634
.
Li
,
J. H.
,
Liu
,
Y. Q.
,
,
P.
,
Lin
,
H. F.
,
Bai
,
Y.
,
Wang
,
X. C.
and
Chen
,
Y. L.
(
2009
).
A signaling pathway linking nitric oxide production to heterotrimeric G protein and hydrogen peroxide regulates extracellular calmodulin induction of stomatal closure in Arabidopsis
.
Plant Physiol.
150
,
114
-
124
.
Lichtman
,
A. H.
,
Segel
,
G. B.
and
Lichtman
,
M. A.
(
1982
).
Effects of trifluoperazine and mitogenic lectins on calcium ATPase activity and calcium transport by human lymphocyte plasma membrane vesicles
.
J. Cell. Physiol.
111
,
213
-
217
.
Ma
,
W.
,
Smigel
,
A.
,
Tsai
,
Y.-C.
,
Braam
,
J.
and
Berkowitz
,
G. A.
(
2008
).
Innate immunity signaling: cytosolic Ca2+ elevation is linked to downstream nitric oxide generation through the action of calmodulin or a calmodulin-like protein
.
Plant Physiol.
148
,
818
-
828
.
Ma
,
F.
,
Lu
,
R.
,
Liu
,
H.
,
Shi
,
B.
,
Zhang
,
J.
,
Tan
,
M.
,
Zhang
,
A.
and
Jiang
,
M.
(
2012
).
Nitric oxide-activated calcium/calmodulin-dependent protein kinase regulates the abscisic acid-induced antioxidant defence in maize
.
J. Exp. Bot.
63
,
4835
-
4847
.
Marone
,
M.
,
Mozzetti
,
S.
,
De Ritis
,
D.
,
Pierelli
,
L.
and
Scambia
,
G.
(
2001
).
Semiquantitative RT-PCR analysis to assess the expression levels of multiple transcripts from the same sample
.
Biol. Proc. Online
3
,
19
-
25
.
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. 
(
2016
).
Nitro-fatty acids in plant signaling: nitro-linolenic acid induces the molecular chaperone network in Arabidopsis
.
Plant Physiol.
170
,
686
-
701
.
McCormack
,
E.
,
Tsai
,
Y.-C.
and
Braam
,
J.
(
2005
).
Handling calcium signaling: arabidopsis CaMs and CMLs
.
Trends Plant Sci.
10
,
383
-
389
.
Monshausen
,
G. B.
(
2012
).
Visualizing Ca2+ signatures in plants
.
Curr. Opin. Plant Biol.
15
,
677
-
682
.
Monshausen
,
G. B.
,
Messerli
,
M. A.
and
Gilroy
,
S.
(
2008
).
Imaging of the Yellow Cameleon 3.6 indicator reveals that elevations in cytosolic Ca2+ follow oscillating increases in growth in root hairs of Arabidopsis
.
Plant Physiol.
147
,
1690
-
1698
.
Mullen
,
R. T.
,
Flynn
,
C. R.
and
Trelease
,
R. N.
(
2001
).
How are peroxisomes formed? The role of the endoplasmic reticulum and peroxins
. Trends Plant Sci.
6
,
256
-
261
.
Murata
,
Y.
,
Mori
,
I. C.
and
Munemasa
,
S.
(
2015
).
Diverse stomatal signaling and the signal integration mechanism
.
Annu. Rev. Plant Biol.
66
,
369
-
392
.
Nelson
,
B. K.
,
Cai
,
X.
and
Nebenführ
,
A.
(
2007
).
A multicolored set of in vivo organelle markers for co-localization studies in Arabidopsis and other plants
.
Plant J.
51
,
1126
-
1136
.
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
.
Niu
,
L.
and
Liao
,
W.
(
2016
).
Hydrogen peroxide signaling in plant development and abiotic responses: crosstalk with nitric oxide and calcium
.
Front. Plant Sci.
7
,
230
.
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
.
Pracharoenwattana
,
I.
and
Smith
,
S. M.
(
2008
).
When is a peroxisome not a peroxisome?
Trends Plant Sci.
13
,
522
-
525
.
Raychaudhury
,
B.
,
Gupta
,
S.
,
Banerjee
,
S.
and
Datta
,
S. C.
(
2006
).
Peroxisome is a reservoir of intracellular calcium
.
Biochim. Biophys. Acta
1760
,
989
-
992
.
Reddy
,
A. S. N.
,
Ali
,
G. S.
,
Celesnik
,
H.
and
Day
,
I. S.
(
2011
).
Coping with stresses: roles of calcium- and calcium/calmodulin-regulated gene expression
.
Plant Cell
23
,
2010
-
2032
.
Rentel
,
M. C.
and
Knight
,
M. R.
(
2004
).
Oxidative stress-induced calcium signaling in Arabidopsis
.
Plant Physiol.
135
,
1471
-
1479
.
Reumann
,
S.
,
Babujee
,
L.
,
Ma
,
C.
,
Wienkoop
,
S.
,
Siemsen
,
T.
,
Antonicelli
,
G. E.
,
Rasche
,
N.
,
Lüder
,
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
.
Reumann
,
S.
and
Corpas
,
F. J.
(
2010
).
The peroxisomal ascorbate–glutathione pathway: molecular identification and insights into its essential role under environmental stress conditions
. In
Ascorbate-Glutathione Pathway and Stress Tolerance in Plants
(ed.
Anjum
, et al. 
), pp.
387
-
404
.
Springer.
Rojas
,
C. M.
,
Senthil-Kumar
,
M.
,
Wang
,
K.
,
Ryu
,
C.-M.
,
Kaundal
,
A.
and
Mysore
,
K. S.
(
2012
).
Glycolate oxidase modulates reactive oxygen species-mediated signal transduction during nonhost resistance in Nicotiana benthamiana and Arabidopsis
.
Plant Cell
24
,
336
-
352
.
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
.
Sang
,
J.
,
Zhang
,
A.
,
Lin
,
F.
,
Tan
,
M.
and
Jiang
,
M.
(
2008
).
Cross-talk between calcium-calmodulin and nitric oxide in abscisic acid signaling in leaves of maize plants
.
Cell Res.
18
,
577
-
588
.
Schmidt
,
M.
,
Grief
,
J.
and
Feierabend
,
J.
(
2006
).
Mode of translational activation of the catalase (cat1) mRNA of rye leaves (Secale cereale L.) and its control through blue light and reactive oxygen
.
Planta
223
,
835
-
846
.
Schwitzguébel
,
J.-P.
and
Siegenthaler
,
P.-A.
(
1984
).
Purification of peroxisomes and mitochondria from spinach leaf by percoll gradient centrifugation
.
Plant Physiol.
75
,
670
-
674
.
Sengupta
,
P.
,
Ruano
,
M. J.
,
Tebar
,
F.
,
Golebiewska
,
U.
,
Zaitseva
,
I.
,
Enrich
,
C.
,
McLaughlin
,
S.
and
Villalobo
,
A.
(
2007
).
Membrane-permeable calmodulin inhibitors (e.g. W-7/W-13) bind to membranes, changing the electrostatic surface potential: dual effect of W-13 on epidermal growth factor receptor activation
.
J. Biol. Chem.
282
,
8474
-
8486
.
Simeunovic
,
A.
,
Mair
,
A.
,
Wurzinger
,
B.
and
Teige
,
M.
(
2016
).
Know where your clients are: subcellular localization and targets of calcium-dependent protein kinases
.
J. Exp. Bot.
67
,
3855
-
3872
.
Stael
,
S.
,
Wurzinger
,
B.
,
Mair
,
A.
,
Mehlmer
,
N.
,
Vothknecht
,
U. C.
and
Teige
,
M.
(
2012
).
Plant organellar calcium signalling: an emerging field
.
J. Exp. Bot.
63
,
1525
-
1542
.
Steinhorst
,
L.
and
Kudla
,
J.
(
2013
).
Calcium and reactive oxygen species rule the waves of signaling
.
Plant Physiol.
163
,
471
-
485
.
Truernit
,
E.
and
Haseloff
,
J.
(
2008
).
A simple way to identify non-viable cells within living plant tissue using confocal microscopy
.
Plant Methods
4
,
15
.
Vandonselaar
,
M.
,
Hickie
,
R. A.
,
Quail
,
J. W.
and
Delbaere
,
L. T. J.
(
1994
).
Trifluoperazine-induced conformational change in Ca(2+)-calmodulin
.
Nat. Struct. Biol.
1
,
795
-
801
.
Wang
,
P.
,
Du
,
Y.
,
Hou
,
Y.-J.
,
Zhao
,
Y.
,
Hsu
,
C.-C.
,
Yuan
,
F.
,
Zhu
,
X.
,
Tao
,
W. A.
,
Song
,
C.-P.
and
Zhu
,
J.-K.
(
2014
).
Nitric oxide negatively regulates abscisic acid signaling in guard cells by S-nitrosylation of OST1
.
Proc. Natl. Acad. Sci. USA
112
,
613
-
618
.
Wilson
,
S. B.
(
1994
).
Complex V of plant mitochondria - another target for agrochemicals?
Biochem. Soc. Trans.
22
,
73S
.
Xu
,
H.
,
Martinoia
,
E.
and
Szabo
,
I.
(
2015
).
Organellar channels and transporters
.
Cell Calcium
58
,
1
-
10
.
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
.
Yang
,
T.
and
Poovaiah
,
B. W.
(
2003
).
Calcium/calmodulin-mediated signal network in plants
.
Trends Plant Sci.
8
,
505
-
512
.
Yun
,
B.-W.
,
Skelly
,
M. J.
,
Yin
,
M.
,
Yu
,
M.
,
Mun
,
B.-G.
,
Lee
,
S.-U.
,
Hussain
,
A.
,
Spoel
,
S. H.
and
Loake
,
G. J.
(
2016
).
Nitric oxide and S-nitrosoglutathione function additively during plant immunity
.
New Phytol.
211
,
516
-
526
.
Zechmann
,
B.
and
Müller
,
M.
(
2010
).
Subcellular compartmentation of glutathione in dicotyledonous plants
.
Protoplasma
246
,
15
-
24
.
Zeng
,
H.
,
Xu
,
L.
,
Singh
,
A.
,
Wang
,
H.
,
Du
,
L.
and
Poovaiah
,
B. W.
(
2015
).
Involvement of calmodulin and calmodulin-like proteins in plant responses to abiotic stresses
.
Front. Plant Sci.
6
,
600
.
Zhang
,
W.-H.
,
Rengel
,
Z.
and
Kuo
,
J.
(
1998
).
Determination of intracellular Ca2+ in cells of intact wheat roots: loading of acetoxymethyl ester of Fluo-3 under low temperature
.
Plant J.
15
,
147
-
151
.
Zhang
,
Y.
,
Zhu.
H.
,
Zhang
,
Q.
,
Li
,
M.
,
Yan
,
M.
,
Wang
,
R.
,
Wang
,
L.
,
Welti
,
R.
,
Zhang
,
W.
and
Wang
,
X.
(
2009
).
Phospholipase dalpha1 and phosphatidic acid regulate NADPH oxidase activity and production of reactive oxygen species in ABA-mediated stomatal closure in Arabidopsis
.
Plant Cell
.
21
,
2357
-
2377
.
Zhao
,
Q.
,
Zhang
,
C.
,
Jia
,
Z.
,
Huang
,
Y.
,
Li
,
H.
and
Song
,
S.
(
2015
).
Involvement of calmodulin in regulation of primary root elongation by N-3-oxo-hexanoyl homoserine lactone in Arabidopsis thaliana
.
Front. Plant Sci.
5
,
807
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information