Abstract
Phosphoinositide-specific phospholipase C (PI-PLC) signaling is involved in various plant stress and developmental responses. Though several aspects of this lipid signaling pathway are conserved within animals and plants, clear differences have also emerged. While animal PLC signaling is characterized by the hydrolysis of PIP2 and production of IP3 and DAG as second messengers to activate Ca2+ and PKC signaling, plant PI-PLCs seem to predominantly use PIP as substrate and convert IP2 and DAG into inositolpolyphosphates and phosphatidic acid (PA) as plant second messengers. Sequencing of multiple plant genomes confirmed that plant PLC signaling evolved differently from animals, lacking homologs of the IP3 gated-Ca2+ channel, PKC and TRP channels, and with PLC enzymes resembling the PLCζ subfamily, which lacks the conserved PH domain that binds PIP2. With emerging tools in plant molecular biology, data analyses, and advanced imaging, plant PLC signaling is ready to gain momentum.
Introduction
Phosphoinositide-phospholipase C (PI-PLC or PLC) (EC 3.1.4.11) signaling was first discovered in animal systems, where activation of G-protein coupled receptors or receptor tyrosine kinases was shown to trigger the hydrolysis of the minor phospholipid phosphatidylinositol 4,5-bisphosphate (PIP2) to produce 2 second messengers: inositol 1,4,5-trisphosphate (IP3) and diacylglycerol (DAG). The water-soluble IP3 triggers an increase in cytosolic Ca2+ by activating ligand-activated Ca2+ channel at the ER, while DAG remains at the plasma membrane (PM) to recruit and activate members of the protein kinase C (PKC) family. Both messengers regulate numerous molecular events and many cellular and physiological processes (Katan and Cockcroft 2020).
Initially, evidence for the existence of a PLC-signaling system in plants rapidly followed the mammalian paradigm, that is, showing the presence of PIP2 and its precursor PI4P, the enzymes that produce them (i.e. PI 4-kinase [PI4K] and PIP 5-kinase [PIP5K]), and a PLC activity hydrolyzing them in vitro (reviewed in Munnik et al. 1998a, 1998b; Drøbak et al. 1999; Stevenson et al. 2000; Meijer and Munnik 2003). Furthermore, by showing that microinjected IP3 was able to release Ca2+ from an intracellular store and to cause stomatal closure (Blatt et al. 1990; Gilroy et al. 1990), it seemed that the plant PLC story was similar to that of animals (Munnik et al. 1998a, 1998b; Ng et al. 2001; Krinke et al. 2007). However, over the years, several fundamental differences between plant and animal PLC signaling emerged, showing these pathways evolved independently and work differently (see Box 1). Most crucially, flowering plants lack homologs for PKC or ligand gated-IP3 receptors and contain very low levels of PIP2 in their PM at basal conditions (Zonia and Munnik 2006; Munnik and Testerink 2009; Munnik and Vermeer 2010; Munnik and Nielsen 2011; Munnik 2014). So why would plants evolve a PLC system based on IP3 and DAG while both effectors and substrate are missing? The answer seems to lie in the use of PI4P as plant-PLC substrate and that the hydrolytic products IP2 and DAG are instantly converted into signaling molecules, that is, various water-soluble inositolpolyphosphates (IPPs) and the lipid second messenger, phosphatidic acid (PA) (Munnik and Vermeer 2010; Testerink and Munnik 2011; Arisz et al. 2013; Vermeer and Munnik 2013; Munnik 2014). Nevertheless, under certain conditions (stress, developmental, polar growth), PIP2 synthesis can be spatiotemporally activated for its own signaling purposes, which may also end as PLC substrate with a subsequent production of IP3 (Fig. 1; Munnik and Vermeer 2010; Munnik 2014). A summary of the crucial differences between plant and animal PLC signaling is highlighted in Box 1.
Box 1. Differences between animal and plant PLC systems.
Based on amino acid sequence and domain organization, animal PLCs have been classified into 6 subfamilies. All plant PLCs belong to the PLCζ class (Box 1). PLCζ represents the most simple isoform, containing only the core domains found in every PLC: the catalytic X and Y domains, 2 Ca2+-binding EF hands to stabilize the catalytic site, and a C2 domain that allows Ca2+-dependent binding to the membrane. Interestingly, PLCζ is the only isoform that lacks a PH domain that binds PI(4,5)P2 in the plasma membrane. Other domains known to regulate animal PLCs—small or trimeric G-proteins, or receptor tyrosine kinases—are all missing in plants (Munnik 2014). How plant PLC is regulated and activated is unknown. In vitro, eukaryotic PLCs use both PI4P and PI(4,5)P2 as substrate, but in vivo, flowering plants contain hardly any PI(4,5)P2, whose levels are typically ∼30- to 100-fold lower than in animals (Munnik and Vermeer 2010). This may also explain the lack of the PIP2-binding PH domains in plant PLCs. They also lack homologs of the IP3 receptor PKC or TRP channel, the most important animal targets of canonical PLC signaling. Plants do have relatively high amounts of PI4P in their plasma membrane, so in vivo, plant PLCs are likely to use PI4P as standard substrate (Munnik 2014; De Jong and Munnik 2021). Subsequent hydrolysis generates IP2 and DAG, for which no direct targets are known in plants. Instead, evidence is emerging that both components are rapidly converted into molecules with second messenger properties, that is, various IPPs (e.g. IP6, IP8), and PA and DGPP, respectively (Munnik and Vermeer 2010; De Jong and Munnik 2021). In animals, IPPs and PA are important signaling molecules too, but plants have exploited this much further, perhaps due to the importance of Pi for plant growth and that IPPs serve also as excellent Pi-storage molecules, adding another dimension to signaling and sensing. Plants also lean much more heavily on PLD signaling, which generates PA from structural phospholipids (Yao et al. 2024). Under certain conditions, PIP2 synthesis is activated in plants, where it seems to function as lipid second messenger—that is, by polyamines; salt, heat, and osmotic stress; and during polar tip growth of root hairs and pollen tubes or expansion of the cell plate during cell division (Heilmann and Heilmann, 2024). Subsequent hydrolysis by PLC would then not only attenuate PIP2 signaling but could also initiate IPP and PA signaling pathways.
Figure 1.
PLC signaling in plants stress and development. PIP and PIP2 are synthesized by PI4-kinase (PI4K) and PIP 5-kinase (PIP5K), respectively. Both lipids have their own signaling functions and are substrate to PLC, generating DAG and IP2 or IP3, depending on the substrate. DAG is rapidly phosphorylated by DGK to generate the lipid second messenger PA, while IP2 and IP3 are phosphorylated by inositol polyphosphate (IPP) kinases IPK2 and IPK1 to form IP5 and IP6 that function as signaling molecules or can even be further phosphorylated by ITPK1, VIH1, and VIH2 to form the pyrophosphorylated species IP7 and IP8 that are involved in signaling too. Alternatively, IPPs can be broken down to inositol for subsequent production of RFOs. Targets and cellular processes are indicated. Part of PA signaling can also be generated through PLD (not indicated). The gene coding for the IP6-mediated Ca2+ release (Lemtiri-Chlieh et al. 2003) is still unknown and indicated by (?).
Plant PLC messengers
Inositolpolyphosphates (IPPs)
PLC-generated IP2 or IP3 is rapidly phosphorylated by different IPP kinases (IPKs, IMPKs) into a range of higher IPPs, of which IP6 is most famous (Flores and Smart 2000; Lemtiri-Chlieh et al. 2000; Lemtiri-Chlieh et al. 2003; Munnik 2014; Lorenzo-Orts et al. 2020; Riemer et al. 2022). IPK2 has 3-, 5- and 6-kinase activity while IPK1 phosphorylates IP5 on the final 2- position of the inositol ring (Fig. 1). Historically, IP6 (also known as phytate) is known as Pi-storage molecule in seeds (and likely other cells, tissues), but over the years, it has emerged as an important signaling molecule and signaling precursor, as IP6 can be further phosphorylated to IP7 by ITPK1, and from IP7 to IP8 by VIH1/VIH2 kinases, both of which are implicated in signaling too (Munnik 2014; Laha et al. 2015, 2016; Wild et al. 2016; Jung et al. 2018; Zhu et al. 2019; Lorenzo-Orts et al. 2020; Riemer et al. 2022; Guan et al. 2022, Guan et al. 2023).
In guard cells, PLC-produced IP6 induces the release of intracellular Ca2+ during ABA-mediated stomatal closure (Flores and Smart 2000; Lemtiri-Chlieh et al. 2000; Lemtiri-Chlieh et al. 2003; Mills et al. 2004). IPPs are also crucial in signaling other hormones, with IP6 being essential as cofactor for auxin receptor, TIR1, and IP5 and IP8 as cofactor for the jasmonate receptor, COI1 (Tan et al. 2007; Hao and Yang 2010; Sheard et al. 2010; Mosblech et al. 2011; Calderón-Villalobos et al. 2012; Munnik 2014; Laha et al. 2015, 2016). Without IPPs, these receptors, and thus the cells and tissues expressing them, would be blind for the particular hormone, allowing yet another interesting layer of regulation that has so far been ignored. IP6, IP7, and IP8 have all been emerging to play key roles in phosphate signaling (Stevenson-Paulik et al. 2005; Williams et al. 2015; Dong et al. 2019; Zhu et al. 2019; Jia et al. 2021; Ried et al. 2021; Riemer et al. 2022), with IP6 acting as cofactor for GLE1, a nuclear pore protein that transports mRNA (Lee et al. 2015), and others being sensed by SPX domains present in multiple target proteins, including nutrient transporters (Lorenzo-Orts et al. 2020). In pollen tubes, IP4 prevents cell expansion by inhibiting a chloride-efflux channel, similar as in mammalian systems (Zonia et al. 2002). Thus, various IPPs are involved in a wide variety of plant signaling events, and potentially this may be initiated by activation of a PLC (Munnik 2014; Riemer et al. 2022). Most evidence, however, is still indirect and circumstantial (e.g. Zhang et al. 2018b, 2018c; Verslues et al. 2023).
Phosphatidic acid (PA)
PLC-generated DAG is rapidly converted into PA by DGK, which is measurable within minutes if the right method is used (Arisz et al. 2009). This PA has emerged as an important second messenger in various signaling pathways, including salt, osmotic, and cold stress but also in disease resistance and symbiosis (Munnik et al. 2000; Van der Luit et al. 2000; den Hartog et al. 2001; Meijer et al. 2001a, 2001b; 2003; Laxalt and Munnik 2002; De Jong et al. 2004; Konig et al. 2007; Laxalt et al. 2007; Distéfano et al. 2008; Mishkind et al. 2009; Sueldo et al. 2010; Raho et al. 2011; Arisz et al. 2013; Meijer et al. 2017; D'Ambrosio et al. 2018; Li et al. 2019; Kong et al. 2024). The problem of measuring PA is that multiple pools exist within cells, of which most is not related to signaling but rather functions as precursor for lipid biosynthesis of bulk phospholipids and glycolipids at the luminal sites of ER and plastids, respectively (Testerink and Munnik 2011; Li-Beisson et al. 2013). Lastly, PA can also be generated through PLD hydrolysis of structural phospholipids, which may have a signaling role too or may simply reflect membrane remodeling or breakdown during phosphate shortage or wounding (Bargmann et al. 2009; Li et al. 2009; Nakamura et al. 2009; Hong et al. 2016). With so many PA pools, and the small amounts of DAG being phosphorylated from PLC hydrolysis of the low abundant PPIs, a simple PA measurement from a cell or plant tissue lipid extract by LC- or GC-MS analysis is unlikely to measure PLC/DGK responses (Arisz et al. 2000, 2003, 2009; Munnik 2014). With differential 32Pi-labeling and transphosphatidylation assays, it is possible to distinguish between them (Munnik et al. 1998a; Arisz et al. 2009). Yet another layer of complexity is that there are other enzymes that could generate DAG in response to stress, such as NPC (non-specific phospholipase C, which hydrolyzes structural phospholipids) and triacylglycerol lipase, whose substrates are present in mass amounts and of which generated DAG could be phosphorylated by DGK too. Using a DAG biosensor, a fluorescent protein fused to the C1 domain of PKC that binds DAG, revealed hardly any free DAG at the cytosolic leaflets of membranes, with no increases upon stresses known to trigger rapid PA responses, implying that the rapid conversion of PLC-generated DAG into PA by DGK likely occurs within the same complex (Vermeer et al. 2017).
Over the years, PA signaling has been associated with many processes, such as, but not limited to, endocytosis, membrane trafficking, actin cytoskeleton dynamics, osmoregulation, PIN localization, regulation of transcription factors, and ROS production (Fig. 1) (Munnik 2001; Wang et al. 2006; Li et al. 2009; Testerink and Munnik 2011; Pleskot et al. 2013; Thomas and Staiger 2014; Wu et al. 2014; Hou et al. 2016; Ufer et al. 2017; Pokotylo et al. 2018; Yao and Xue 2018; Li et al. 2019; Li and Wang 2019; Wang et al. 2019; Kolesnikov et al. 2022). Various potential PA targets have been identified (Testerink et al. 2004; McLoughlin et al. 2012, 2013; Gao et al. 2013; Wu et al. 2014; Julkowska et al. 2015; Putta et al. 2016; Pandit et al. 2018; Yao and Xue 2018; Kim et al. 2019; Kolesnikov et al. 2022; Kong et al. 2024), and an increase of PA at the PM, or anywhere else at the cytosolic leaflets of membranes, is expected to recruit such target proteins to these sites. In this way, ABA insensitive (ABI1), a protein phosphatase, is inhibited by PA, thereby preventing the dephosphorylating of target proteins in the cytoplasm, which promotes ABA signaling (Zhang et al. 2004; Mishra et al. 2006). ROS-producing NADPH oxidases RbohD and RbohF are examples of PA targets already located at the PM but whose activity is enhanced and stabilized by PA (Zhang et al. 2009; Kong et al. 2024; Qi et al. 2024). RbohD and RbohF play essential roles in stress signaling and development (Zhang et al. 2009; Kadota et al. 2015; Chapman et al. 2019). Also activated by PA are various protein kinases, such as MPK6, MKK7, and MPK9, leading to the activation of SOS1, an Na+/H+ antiporter that transports Na+ out of the cell in response to salinity stress (Yu et al. 2010; Shen et al. 2019). SnRK2.4 and SnRK2.10 are 2 other PA-binding protein kinases related to salinity (McLoughlin et al. 2012, 2013; Julkowska et al. 2015). PA has also been suggested to play a role in the nucleus, by preventing CCA1, a circadian rhythm regulator, from binding its target promoters (Kim et al. 2019). Some of these PA responses may also be related to PLD signaling (Yao et al. 2024).
With so many PA pools and responses, the big question remains of how PA would generate any specificity. A similar question holds for Ca2+ signaling that is basically activated by any stress ever tested. In this field, specificity is thought to lie in so-called Ca2+ signatures, with differences in amplitude and frequency (Edel and Kudla 2015; Gilroy et al. 2016; Yuan et al. 2017). A much easier explanation, however, is that stress and developmental signals activate more than just 1 signaling pathway and that the combination of such pathways, plus what is specifically expressed in the particular cell or tissue as targets, together determine the specificity of the outcome. A big advantage of PA (and other lipid second messengers) is that it generates a spatiotemporal membrane-delimited signal, so that the whole process, that is, target activation and/or recruitment, complex formation, and response output etc., can occur at the same site where the signal is perceived and/or receptor is activated—for example, where a cell wall locally requires a reinforcement. Water-soluble messengers, like Ca2+ and the various IPPs, have other spatio-temporal advantages, such that they can warn and prepare the whole cell.
The PA signal can be switched off by dephosphorylation through PA phosphatases and metabolism into structural phospholipids, or by phosphorylating it into diacylglycerol pyrophosphate (DGPP) through PA kinase (PAK) (Munnik et al. 1996; Munnik 2001; van Schooten et al. 2006). Both routes would attenuate PA signaling, but with DGPP, a potential new signaling molecule would be generated, though research on PAK and DGPP targets is progressing slowly (den Hartog et al. 2001; Munnik 2001; van Schooten et al. 2006; Arisz et al. 2009; Graber et al. 2022).
PIP and PIP2
PI4P and PI(4,5)P2 are not only substrates for PLC, but they are also emerging as signaling molecules themselves (Fig. 1; Noack and Jaillais 2020; De Jong and Munnik 2021; Heilmann and Heilmann 2024), and as such, PLC may control their activity besides generating IPPs and PA (Munnik 2014).
Synthesis of these PPIs starts from phosphatidylinositol (PI), which is generated by PI synthase from CDP-DAG and myo-inositol (Lofke et al. 2008; Vermeer and Munnik 2013). PI comprises about 5% to 10% of all phospholipids, of which only a small percentage (5% to 10%) is phosphorylated into PI4P by PI4K and even a smaller percentage (2% to 3% of PIP) is phosphorylated into PI(4,5)P2 by PIP5K (Munnik et al. 1994; Munnik et al. 1998a; Vermeer and Munnik 2013).
Phosphorylation can also occur at the D3 postion of the inositol ring, creating distinct PPI isomers: PI3P, PI(3,5)P2, and PI5P [the latter through dephosphorylation of PI(3,5)P2] (Munnik and Vermeer 2010). While these PPIs play important roles in signaling too (Meijer et al. 1999; Hirano et al. 2017a; Noack and Jaillais 2017; Lee et al. 2018; Chung 2019; Hirano and Sato 2019; Liu et al. 2020a; Xing et al. 2021; Lin et al. 2022), they are not PLC substrates (Munnik et al. 1998a).
PI4P and PI(4,5)P2 are implicated in various cellular processes, often involving PA too, including endocytosis and exocytosis through interaction with AP2, clathrin, EXO70, and pattelin proteins (Naramoto et al. 2009; Zhao et al. 2010; Ischebeck et al. 2013; Noack and Jaillais 2017; Wu et al. 2017; Noack and Jaillais 2020; Doumane et al. 2021; Synek et al. 2021; Xing et al. 2021; Yperman et al. 2021; Lebecq et al. 2022). Similarly, these lipids are implicated in membrane contact sites (Brault et al. 2019; Ruiz-Lopez et al. 2021; Qian et al. 2022) and are important for (1) cell division, probably by recruiting target proteins to the cell plate during its formation (van Leeuwen et al. 2007; Vermeer et al. 2009, 2017; Lin et al. 2019); (2) cell polarity, by influencing PIN localization in cells (Mei et al. 2012; Tejos et al. 2014; Rodriguez-Villalon et al. 2015; Gujas et al. 2017; Marhava et al. 2020; Wang et al. 2023; Aliaga Fandino et al. 2024); and (3) tip growth, by localizing PIP2 at the tip of root hairs and pollen tubes (Dowd et al. 2006; Helling et al. 2006; Sousa et al. 2008; Gagne and Clark 2010; Ischebeck et al. 2010; Zonia and Munnik 2011; Yoo et al. 2012; Ischebeck et al. 2013; Stanislas et al. 2015; Simon et al. 2016; Kato et al. 2019; Fratini et al. 2021; Song et al. 2021). Both PPIs (and PA) are involved in regulating the actin cytoskeleton as well (Yoo et al. 2012; Guo et al. 2020; Doumane et al. 2021; Fratini et al. 2021). Additionally, PI4P has been found to play a role in brassinosteroid signaling by interacting with BKI1 and MAKR1-4 proteins (Jaillais and Vert 2016; Simon et al. 2016) and is involved in plant immunity and cell death (Saile et al. 2021). Lastly, PIP2 is emerging to regulate K+ channels and to control stomatal opening (Lee et al. 2007; Lee and Lee 2008; Ma et al. 2009; Zarza et al. 2020).
Several lipid-binding domains that specifically bind the headgroup of PI(4,5)P2 have been identified (De Jong and Munnik 2021). Target proteins containing such domains can be recruited to sites where PIP2 is synthesized by activation of 1 of the 11 Arabidopsis PIP5Ks (De Jong and Munnik 2021; Kuroda et al. 2021; Verslues et al. 2023; Heilmann and Heilmann 2024). Additionally, PIP2 and PI4P are implicated to play roles in membrane curvature, electrostatics, and nanodomain identity, which may further explain involvement in above pathways (Fig. 1) (Heilmann 2016a, 2016b; Simon et al. 2016; Gronnier et al. 2017; Colin and Jaillais 2020; Noack and Jaillais 2020; Fratini et al. 2021; Heilmann and Heilmann 2022, 2024), which PLC would attenuate upon activation. Recent studies also imply PIP2 signaling in the nucleus (Dieck et al. 2012; Gerth et al. 2017; Heilmann and Heilmann 2024).
Difficulties of measuring PLC activity in vivo
Typical omics analyses are unable to measure PPI or IPP levels. The reason for this is that their concentrations are low and that isomers cannot be distinguished (Munnik 2014). PA is usually present in just sufficient quantities (1% to 4% of total phospholipids), but most of this represents a precursor role in glycerolipid synthesis at ER and plastids and not what is generated by DAG phosphorylation (Arisz et al. 2013).
To measure PPI and PA responses in plant tissues, in vivo metabolic labeling with radioactive tracers (32Pi, 33Pi) is commonly used. Extracted lipids can be separated by thin layer chromatography and visualized and quantified by phosphoimaging (Munnik and Zarza 2013). If necessary, PPI isomers can be separated by HPLC-SAX analysis after removal of their fatty acids (Munnik 2013). For IPPs, tissue or seedlings are metabolically prelabeled with 3H- or 14C-myo-inositol and analyzed using HPLC-SAX analysis too. However, since concentrations of many are extremely low, they remain difficult to measure (Stevenson-Paulik et al. 2005; Laha et al. 2015; Zhang et al. 2018b; Laha et al. 2022).
With the discovery of various specific lipid-binding domains (de Jong and Munnik 2021), a number of genetically encoded-lipid biosensors have been developed by fusing encoded DNA to that of fluorescent proteins and expressing them in plants. As such, biosensors have been developed for DAG, PA, PI4P, and PI(4,5)P2, but also for PI3P, PI(3,5)P2, and PS (Vermeer et al. 2006; van Leeuwen et al. 2007; Vermeer et al. 2009; Vermeer and Munnik 2013; Simon et al. 2014; Platre and Jaillais 2016; Hirano et al. 2017b; Vermeer et al. 2017; Platre et al. 2018). Such biosensors visualize the in vivo dynamics and subcellular localization of these lipids but do not measure absolute levels.
PLC isoforms: protein structure, regulation, and localization
Arabidopsis contains 9 PLC genes (Hunt et al. 2004; Tasma et al. 2008). Several plant PLC gene families have been resolved meanwhile, including monocots maize (11), wheat (11), brachypodium (4), and rice (4); and dicots potato (4), tomato (6), tobacco (6) soybean (12), chickpea (6) cotton (12), peach (5), coffea (4), various orchids (∼2–3), poplar (8), and rapeseed (27) (Das et al. 2005; Sui et al. 2008; Vossen et al. 2010; Singh et al. 2013; Zhang et al. 2014; Zhang et al. 2015; Wang et al. 2015; Zhang et al. 2018a; Gonzalez-Mendoza et al. 2020; Iqbal et al. 2020; Sagar et al. 2020; Wang et al. 2020a; Wang et al. 2020b; Zhu et al. 2020; Chen et al. 2021; Kanchan et al. 2021; Wang et al. 2021; Luo et al. 2022; Sun et al. 2022; Perk et al. 2023). The number of PLCs in rapeseed, an important seed oil crop and close family member of Arabidopsis, is quite remarkable, even taking into account that rapeseed (Brassica napus (AACC, 2n = 38) emerged from a cross between B. rapa (AA, 2n = 20) and B. oleracea (CC, 2n = 18) (Iqbal et al. 2020).
Phylogenetic analyses show that PLCs cluster into 5 clades, with 4 of them consisting of dicotyledonous species, and the monocotyledonous species grouping together into a single clade (Perk et al. 2023). Interestingly, Arabidopsis PLC8- and PLC9 appear to be outliers, as no homologues in other species are found. AtPLC8 and −9 differ from other PLCs by containing deletions in the Y-domain and 2 mutations in the catalytic center, which are associated with a strong reduction in enzymatic activity in rat PLCδ1 (Ellis et al. 1998; Tasma et al. 2008). Whether PLC8 and PLC9 exhibit in vitro enzymatic activity is unknown; however, KO and OE mutants of PLC9 do exhibit reduced and increased thermotolerance, respectively (Zheng et al. 2012; Liu et al. 2020b), and OE of PLC9 has been shown to lead to increased drought tolerance, just like OE of AtPLC2, −3, −4, −5, and −7 (Van Wijk et al. 2018; Zhang et al. 2018b, 2018c; van Hooren et al. 2023, 2024). Hence, the enzymatic status of AtPLC8 and AtPLC9 remains unclear, let alone their precise role in Arabidopsis.
The catalytic activity of PLC rests on the X and Y domains, which together form a 3-dimensional structure, like a hollow barrel (TIM-barrel), of which the inside hosts the active catalytic sites (Box 1; Essen et al. 1996; Hunt et al. 2004; Katan and Cockcroft 2020). The X-Y linker connecting the 2 domains is a region with a high degree of structural disorder that is not conserved between plant PLCs and may have a role in regulating PLC's activity (Kopka et al. 1998; Hicks et al. 2008; Robuschi et al. 2024).
How plant PLCs are activated and regulated is still a big mystery. Prediction software identified potential sites for palmitoylation, phosphorylation, sumoylation, ubiquitination, and acylation (Pokotylo et al. 2014; Kumar et al. 2022), though in vivo, only phosphorylation and acetylation has been shown (Nuhse et al. 2007; Hemsley et al. 2013). In particular, the intrinsically disordered X-Y linker in AtPLC2 is susceptible to phosphorylation upon elicitation with flg22 or oligogalacturonic acid (Nuhse et al. 2003, 2007; Kohorn et al. 2016). Regulation through monomeric and trimeric G-proteins has been suggested (Misra et al. 2007; Khalil et al. 2011; Sun et al. 2011), but genetic and structural evidence is missing for this (Munnik 2014). Despite the scarce information available on the regulatory role of the extra-catalytic domains, both C2 domain and EF-hand motif confirm a role of Ca2+ for their activity (Otterhag et al. 2001; Hunt et al. 2004; Rupwate and Rajasekharan 2012).
Various Arabidopsis PLC genes (PLC1-3, −5, −7, and −9) are prominently expressed in vascular tissue, in particular in phloem and/or companion cells (van Wijk et al. 2018; Zhang et al. 2018b, 2018c; Van Hooren et al. 2023, 2024). Specific PLC expression has been reported for various other cell types, including guard cells (PLC2, PLC5), trichomes (PLC2, −3, −5, and −7), female gametophytes and early embryos (PLC2), and in pollen (PLC4) (Hunt et al. 2004; Kanehara et al. 2015; Di Fino et al. 2017; van Wijk et al. 2018; Zhang et al. 2018b, 2018c; van Hooren et al. 2023, 2024).
Several PLC fusions with fluorescent proteins confirmed localization at the PM, including AtPLC2, AtPLC3, AtPLC5, AtPLC9, PetPLC1, NtPLC3, VrPLC3, OsPLC1, and OsPLC3 (Kim et al. 2004; Dowd et al. 2006; Helling et al. 2006; Zheng et al. 2012; Kanehara et al. 2015; Li et al. 2017; Ren et al. 2017; Yu et al. 2021; van Hooren 2023). Some PLC fusions were found in the cytoplasm, such as OsPLC1 and OsPLC4 (Singh et al. 2013), or nucleus, including AtPLC3, OsPLC1, OsPLC3, and OsPLC4 (Singh et al. 2013; Ren et al. 2017; Yu et al. 2021), though it is unclear whether these reflect breakdown products of the fusion protein. Interestingly, our laboratories never managed to create N-terminal fusions, indicating the N terminus is crucial for PLC functionality.
Since breakdown of PIP and production of PA can already be witnessed within minutes of stress application, it is likely that most of the PLC regulation occurs at the protein level (Munnik et al. 2000; Meijer et al. 2001a, 2001b; Konig et al. 2007; Mishkind et al. 2009; Arisz et al. 2013; Meijer et al. 2017; Li et al. 2019). Thus, the PLC protein should already be there, expressed at basal conditions, and catalytically inactive. Often, treatments that activate PLC signaling also enhance its gene expression (or of other PLCs), maybe as a feedback mechanism to prime the cell/tissue for further stimulation (Hirayama et al. 1995; Hirt 1999), though it may also reflect a response that is important for acclimation later on. Such a strategy has been used to identify tomato PLCs involved in plant defense against Cladosporium fulvum (Vossen et al. 2010), Botrytis cinerea (Gonorazky et al. 2016; Perk et al. 2023), Phytophthora infestans (Perk et al. unpublished data), and during perception of pathogen derived molecules (Gonorazky et al. 2014).
Transcriptomic analysis has revealed stress-responsive expression of multiple PLC genes in all plant species studied. In most cases, transcript levels increased in response to stress, though substantial variation among PLC isoforms, type of stress, and plant species is observed. A large number of stress-responsive elements have been found in PLC promotors across plant species, including, but not limited to, DREB2A-responsive ABRE element, ethylene-responsive ERE element, and the stress-responsive MYB element (Tasma et al. 2008; Hsieh et al. 2013; Singh et al. 2013; Sagar et al. 2020; Wang et al. 2020a, 2020b; Zhu et al. 2020). Only a few PLCs were downregulated: OsPLC3 CaPLC4 and ZmPLC2 upon salt stress (Sagar et al. 2020; Zhu et al. 2020; Yu et al. 2021), ZmPLC1 and ZmPLC2 upon cold (Zhu et al. 2020), GmPLC3 and GmPLC9 upon ABA treatment (Wang et al. 2015), GmPLC6 upon hyperosmotic stress treatment with PEG (Wang et al. 2015), ZmPLC3 and ZmPLC4 with heavy metal (copper) (Zhu et al. 2020), and SlPLC1 during Botrytis infection (Gonorazky et al. 2016).
PLC in plant stress and development
Development and growth
KO and KD analyses of Arabidopsis PLCs have revealed roles for PLC in ABA and auxin signaling, lateral root formation, leaf serration, seed mucilage, and gametogenesis (Table 1). Silencing of AtPLC1 decreased ABA sensitivity during seed germination and seedling growth (Sanchez and Chua 2001), while Atplc3-KO mutants revealed reduced ABA sensitivity in both seed germination and stomatal closure and displayed a reduction in lateral root number (Zhang et al. 2018b). KO of AtPLC5 turned out lethal, whereas Atplc5-KD plants grew more slowly and exhibited a reduced number of lateral roots too (Zhang et al. 2018c). Atplc7-KO plants are less sensitive to ABA during stomatal closure, and double KO mutant of plc5 plc7 displayed defects in seed mucilage and leaf serration (van Wijk et al. 2018). In contrast, double Atplc3 plc7 KOs were lethal (van Wijk et al. 2018). Single KO of AtPLC2 is also lethal due to problems in female gametophyte and embryo development (Di Fino et al. 2017), of which parts are mediated through auxin (Li et al. 2015). A low proportion (0.8%) of plc2-2 homozygous mutants escaped lethality and showed morphological defects in megagametogenesis (Di Fino et al. 2017) and auxin signaling, with severe phenotypes related to auxin and PIN distribution in root development and gravitropism (Chen et al. 2019).
Table 1.
Arabidopsis PI-PLC genes, mutants, and their phenotypes
| PI-PLC | ATG nr. | Expression | Mutant | Phenotype | References |
|---|---|---|---|---|---|
| PLC1 | At5g58670 | KD | Antisense | Reduced inhibition of ABA on germination and growth | Sanchez and Chua (2001) |
| KO | plc1-1 | Increased trichome branching | Goldberg et al. (2024)) | ||
| OE | Underdeveloped trichomes | Goldberg et al. (2024) | |||
| PLC2 | At3g08510 | KD | amiR | Reduced stomatal MAMP-triggered immunity, reduced ROS production | D'Ambrosio et al. (2017) |
| Reduced Botrytis cinerea infection and proliferation | Robuschi et al. (2024) | ||||
| KO | plc2-2 | Problems in gametophyte and embryo development | Di Fino et al. (2017) | ||
| Unpaired auxin-modulated reproductive development | Li et al. (2015) | ||||
| Unpaired auxin-modulated root development, gravitropism | Chen et al. (2019) | ||||
| CRISPR/Cas9 | Unpaired auxin-modulated reproductive development | Li et al. (2015) | |||
| plc2-1 | Unpaired endoplasmic reticulum stress responses | Kanehara et al. (2015) | |||
| Problems in female gametophyte development | Di Fino et al. (2017) | ||||
| Reduced Botrytis cinerea infection and proliferation; Shorter leaf serration | Robuschi et al. (2024) | ||||
| OE | Increased drought tolerance | van Hooren et al. (2024) | |||
| PLC3 | At4g38530 | KD | plc3-2, plc3-3 | Defects on root development, lateral root formation, seed germination | Zhang et al. (2018a) |
| Reduced sensitivity to ABA stomatal closure and germination | Zhang et al. (2018a) | ||||
| Decreased heat resistance | Gao et al. (2014) | ||||
| OE | Increased heat resistance | Gao et al. (2014) | |||
| Increased drought tolerance | Zhang et al. (2018a); van Hooren et al. (2024) | ||||
| PLC4 | At5g58700 | KO | plc4-3 | Hyposensitivity to salt; normal salt sensitivity | Xia et al. (2017); van Hooren et al. (2023) |
| OE | Increased drought tolerance | van Hooren et al. (2024) | |||
| Hypersensitivity to salt; normal salt sensitivity | Xia et al. (2017); van Hooren et al. (2023) | ||||
| PLC5 | At5g58690 | KO | Lethal | Zhang et al. (2018b) | |
| KD | plc5-1 | Defects in primary root growth and lateral root formation | Zhang et al. (2018b) | ||
| OE | Increased drought tolerance; shorter root hairs | van Hooren et al. (2024); Zhang et al. (2018b) | |||
| PLC6 | At2g40116 | No | NA | NA | Müller-Roeber and Pical (2000); Hunt et al. (2004) |
| PLC7 | At3g55940 | KO/KD | plc7-3/plc7-4 | Reduved stomatal closure by ABA; enhanced leaf serration, seed mucilage defect | van Wijk et al. (2018); |
| OE | Increased drought tolerance | van Wijk et al. (2018); van Hooren et al. (2024) | |||
| AtPLC8 | At3g47290 | NA | NA | Hunt et al. (2004); Tasma et al. (2008) | |
| AtPLC9 | At3g47220 | KO | plc9-1, 9-2 | Decreased heat tolerance | Zheng et al. (2012) |
| OE | Increased heat tolerance | Gao et al. (2014) | |||
| Increased drought tolerance | van Hooren et al. (2024) |
Abiotic stress
Atplc3- and plc9-KO mutants lost part of their heat stress tolerance, which was additive in plc3 plc9 double mutants (Zheng et al. 2012; Gao et al. 2014). Atplc5 plc7 double mutants showed decreased stomatal movement but, strangely, showed increased survival upon drought stress (van Wijk et al. 2018; Zhang et al. 2018c). For Atplc4 mutants, an increase in salt tolerance was reported (Xia et al. 2017), though this was not reproduced (Van Hooren et al. 2023). Rice PLC-KO mutants Osplc1 and Osplc4 both showed reduced salt tolerance, with Osplc4 also showing reduced drought tolerance and Osplc1 plants having reduced grain size (Li et al. 2017; Yu et al. 2022). For Osplc3, reduced damage was found when treated with H2O2 (Yu et al. 2021).
Overexpression (OE) of PLCs clearly improves plant stress tolerance. OE of AtPLC3 or AtPLC9 increased heat tolerance in Arabidopsis (Zheng et al. 2012; Gao et al. 2014) and rice (Liu et al. 2020b). Numerous PLCs have been shown to promote drought tolerance when overexpressed. These include ZmPLC1 in maize (Wang et al. 2008), BnPLC2 in canola (Georges et al. 2009), and AtPLC2, AtPLC3, AtPLC4, AtPLC5, AtPLC7, and AtPLC9 in Arabidopsis (van Wijk et al. 2018; Zhang et al. 2018b, 2018c; van Hooren et al. 2023, 2024). OE of wheat TaPLC1 in Arabidopsis or NtPLCδ1 in tobacco improved both salt and osmotic stress tolerance (Tripathy et al. 2012; Wang et al. 2020b). OE of rice OsPLC1 and OsPLC4, or soybean GmPLC7, increased the salt and drought stress tolerance too (Li et al. 2017; Deng et al. 2019; Chen et al. 2021). Two outliers of the above pattern have been reported: (1) inducible OE of AtPLC4 was found to reduce salt tolerance of Arabidopsis seedling growth; and (2) OE of OsPLC3 in rice displayed a reduction in tolerance to either salt or drought, though it must be said that the latter mRNA levels normally decrease under salt or drought stress (Xia et al. 2017; Yu et al. 2021).
Biotic stress
PLC signaling has been implicated in ROS production during PAMP (pathogen-associated molecular patterns)- and effector-triggered immunity (PTI and ETI) responses (van der Luit et al. 2000; Laxalt and Munnik 2002; den Hartog et al. 2003; de Jong et al. 2004; Andersson et al. 2006; Laxalt et al. 2007; Raho et al. 2011; Gonorazky et al. 2014; Abd-El-Haliem and Joosten 2017; D’Ambrosio et al. 2017; Fang et al. 2023), which is mainly based on instant 32P-PA responses. Genetic evidence for the involvement of PLC came from several recent studies using KO and KD plants. In Arabidopsis, AtPLC2/DGK5-derived PA regulates the NADPH oxidase D (RbohD) activity during PTI response (Box 2; D'Ambrosio et al. 2017; Kong et al. 2024). By transient gene silencing or loss-of-function mutants generated by CRISPR/Cas9 in tomato, SlPLC2 was shown to play a crucial role in the susceptibility to Botrytis cinerea (Gonorazky et al. 2016; Perk et al. 2023) and Phytophthora infestans (Perk et al. unpublished data) and, hence, in this context should be considered as a susceptibility gene (a gene utilized by the pathogen during colonization and infection). SlPLC2-KO plants exhibit normal phenotypic development and normal susceptibility to bacterial pathogen (Perk et al. 2023). SlPLC4 is specifically required for the ETI during Avr4 perception, while SlPLC6 is associated with responses to a wider range of pathogens (Vossen et al. 2010). In Nicotiana benthamiana, NbPLC2 is required for PTI triggered by Ralstonia solanacearum or Pseudomonas fluorescens (Kiba et al. 2020). In rice, OsPLC-1 expression is associated with the resistance response to the fungus, Magnaporthe grisea (Song and Goodman 2002).
Box 2. Case study - Role for Arabidopsis PLC2 and DGK5 in plant defence.
In the early 2000s, various plant systems were shown to trigger a fast (minutes) PA response upon stimulation with PAMPs (then called elicitors). Biochemically, evidence was provided that the majority of this response was generated through phosphorylation of DAG and that this occurred upstream of the ROS burst (Van der Luit et al. 2000; De jong et al. 2004; den Hartog et al. Laxalt and Munnik, 2002). Recently, genetic evidence for this pathway was obtained for PLC2 and DGK5 in Arabidopsis. Constitutively, of all 9 AtPLCs, PLC2 exhibits the highest expression levels among tissues, with predominant expression in the vasculature (roots and leaves). PLC2 is associated with the plasma membrane and rapidly phosphorylated on Ser280 upon treatment with flg22 and other elicitors (Nuhse et al. 2003; Nuhse et al. 2007; Kohorn et al. 2016). Silencing of PLC2 reduced PAMP-triggered immunity (PTI) by reducing RbohD-dependent ROS production, and consequently reduced disease resistance (D'Ambrosio et al. 2017). Recent studies identified DGK5 as key enzyme in the PA response upon flg22 treatment, with dgk5-KO mutants showing reduced ROS production and disease resistance (Kalachova et al. 2022; Kong et al. 2024; Qi et al. 2024). PA-binding of RbohD enhanced ROS production and increased protein stability, thereby augmenting the PAMP induced-ROS production (Kong et al. 2024; Qi et al. 2024). Upon flg22 recognition by FLS2, DGK5 is activated through phosphorylation of Ser506 by the pattern recognition receptor (PRR)-associated kinase, BIK1, while MPK4 phosphorylation of Thr446 deactivates DGK5, resulting in a local and transient PA response at the plasma membrane (Kong et al. 2024). These findings provide functional and genetic evidence for the role of PLC2-DGK5 action in producing ROS as a first layer of basal defense. The relevance of PLC2's phosphorylation and identity of the protein kinase are still unknown. Consequences of the subsequent increase in IP2 remains unknown too, but likely plays a role in the production of IPPs implicated in plant defense (Murphy et al. 2008; Mosblech et al. 2011; Laha et al. 2015; Riemer et al. 2022). Meanwhile, several plant-pathogen systems have been shown to use this pathway.
Concluding remarks
Much progress has been made on PLC signaling over the last decade, yet many issues still remain to be resolved. Clearly, we have to let go of the animal paradigm and accept that plants PLC signaling has its differences. We still need to identify and characterize more downstream targets of PA and IPPs and obtain more knowledge about their spatiotemporal activation and/or recruitment in response to stimuli. The various lipid biosensors have revolutionized our ideas about the location and behavior of lipids, but it would even be better if FRET-based sensors were available to monitor concentration changes. Similarly, it would be great if biosensors for the various IPP isomers would become available to monitor their in vivo dynamics and also a new MS protocol that would measure their absolute levels; the latter also holds for the PPIs. A link to Ca2+ has already been known since the 1980s, but this field, unfortunately, has not made any progress in identifying any of these ligand-gated channel(s). Even though many potential regulators of PLC have been identified, actual evidence for regulation is still very limited, while this could greatly help understanding the function of PLCs. That many PLCs are predominantly expressed in deeper lying phloem tissues will be another challenge to functionally characterize their role in signaling. Nonetheless, with a good number of tools emerging in advanced imaging, omics, genetic screens, and data analyses with AI, PLC signaling will remain an exciting topic for future basic and applied sciences.
Advances box.
Multiple plant PLC families have been sequenced; all members belong to the same PLCζ subfamily as described in animals. It is still unknown how this isoform is regulated and activated.
PLC-KO and -OE mutants have revealed specific roles in plant stress signaling and development.
Not PI(4,5)P2 but PI4P appears to be the predominant PLC substrate.
Plant PLC signaling does not work via the canonical IP3/Ca2+ and DAG/PKC routes as described in animals. Both IP3 receptor and PKC (or TRP channels, another important set of targets) are missing from flowering plant genomes. Instead, IPPs and PA are emerging as plant second messengers downstream of PLC.
The first receptor-activated PLC/DGK pathway has been unravelled, revealing a crucial role in basal disease resistance in Arabidopsis (flg22, FLS2, BIK1, PLC2, DGK5, RbohD).
Outstanding questions box.
What controls PLC's activity status and how is the enzyme regulated at the molecular level?
Is PLC in complex with DGK to boost PA levels directly?
What is the effect of PLC phosphorylation on activity, localization, complex formation?
Do changes in PLC gene expression eventually reflect changes in PLC activity?
How does IP6 activate a Ca2+ channel that releases intracellular Ca2+ upon IP6 formation in guard cells but also elsewhere?
While several PA-binding proteins have been identified, a genuine PA target “at work” remains to be shown. How does the PA-binding domain look like?
Single KO mutants of PLC have only small or no phenotypic effect? Is this due to genetic or chemical redundancy?
Which other receptors than FLS2 lead to a direct activation of a plant PLC-signaling cascade?
Why is PLC so predominantly expressed in phloem and companion cells?
What kind of protein modification regulates plant PLCs?
Is there any difference in Km and Vmax values of PLC activity between PI4P and PI(4,5)P2?
Is PLC a bona fide attenuator of PI(4,5)P2 signaling?
Are PLC-generated IPPs required for functional auxin- and JA perception via their respective receptors, and if so, what triggers it?
Similarly, are (part of) the IPPs that play a role in phosphate signaling generated through PLC?
Acknowledgments
We thank our funding sponsors and both reviewers for their valuable comments and corrections.
Contributor Information
Ana M Laxalt, Instituto de Investigaciones Biológicas, IIB-CONICET, Universidad Nacional de Mar del Plata, Argentina.
Max van Hooren, Plant Cell Biologie, Swammerdam Institute for Life Sciences, University of Amsterdam, Amsterdam, The Netherlands.
Teun Munnik, Plant Cell Biologie, Swammerdam Institute for Life Sciences, University of Amsterdam, Amsterdam, The Netherlands.
Author contributions
All 3 authors wrote the review and constructed the figures.
Funding
This work was supported by the Netherlands Organization for Scientific Research (NWO 867.15.020 to TM), the European Union (EPPN2020 grant 731013 to TM), and Agencia Nacional de Promoción Científica y Tecnológica (ANPCyT, PICT 2021 to AML).
Data availability
The data underlying this article are available in the article and in its online supplementary material.
Dive Curated Terms
The following phenotypic, genotypic, and functional terms are of significance to the work described in this paper:
References
- Abd-El-Haliem AM, Joosten MH. Plant phosphatidylinositol-specific phospholipase C at the center of plant innate immunity. J Integr Plant Biol. 2017:59(3):164–179. 10.1111/jipb.12520 [DOI] [PubMed] [Google Scholar]
- Aliaga Fandino AC, Jelínková A, Marhava P, Petrášek J, Hardtke CS. Ectopic assembly of an auxin efflux control machinery shifts developmental trajectories. Plant Cell. 2024:36(5):1791–1805. 10.1093/plcell/koae023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andersson MX, Kourtchenko O, Dangl JL, Mackey D, Ellerström M. Phospholipase-dependent signalling during the AvrRpm1- and AvrRpt2-induced disease resistance responses in Arabidopsis thaliana. Plant J. 2006:47(6):947–959. 10.1111/j.1365-313X.2006.02844.x [DOI] [PubMed] [Google Scholar]
- Arisz SA, Testerink C, Munnik T. Plant PA signaling via diacylglycerol kinase. Biochim Biophys Acta. 2009:1791(9):869–875. 10.1016/j.bbalip.2009.04.006 [DOI] [PubMed] [Google Scholar]
- Arisz SA, Valianpour F, Van Gennip AH, Munnik T. Substrate preference of stress-activated phospholipase D in Chlamydomonas and its contribution to PA formation. Plant J. 2003:34(5):595–604. 10.1046/j.1365-313X.2003.01750.x [DOI] [PubMed] [Google Scholar]
- Arisz SA, van Himbergen JA, Musgrave A, van den Ende H, Munnik T. Polar glycerolipids of Chlamydomonas moewusii. Phytochemistry. 2000:53(2):265–270. 10.1016/s0031-9422(99)00505-1 [DOI] [PubMed] [Google Scholar]
- Arisz SA, van Wijk R, Roels W, Zhu JK, Haring MA, Munnik T. Rapid phosphatidic acid accumulation in response to low temperature stress in Arabidopsis is generated through diacylglycerol kinase. Front Plant Sci. 2013:4:1. 10.3389/fpls.2013.00001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bargmann BO, Laxalt AM, ter Riet B, Testerink C, Merquiol E, Mosblech A, Leon-Reyes A, Pieterse CM, Haring MA, Heilmann I, et al. Reassessing the role of phospholipase D in the Arabidopsis wounding response. Plant Cell Environ. 2009:32(7):837–850. 10.1111/j.1365-3040.2009.01962.x [DOI] [PubMed] [Google Scholar]
- Blatt MR, Thiel G, Trentham DR. Reversible inactivation of K+ channels of Vicia stomatal guard cells following the photolysis of caged inositol 1,4,5-trisphosphate. Nature. 1990:346(6286):766–769. 10.1038/346766a0 [DOI] [PubMed] [Google Scholar]
- Brault ML, Petit JD, Immel F, Nicolas WJ, Glavier M, Brocard L, Gaston A, Fouché M, Hawkins TJ, Crowet JM, et al. Multiple C2 domains and transmembrane region proteins (MCTPs) tether membranes at plasmodesmata. EMBO Rep. 2019:20(8):e47182. 10.15252/embr.201847182 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calderón-Villalobos LI, Lee S, De Oliveira C, Ivetac A, Brandt W, Armitage L, Sheard LB, Tan X, Parry G, Mao H, et al. A combinatorial TIR1/AFB-aux/IAA co-receptor system for differential sensing of auxin. Nat Chem Biol. 2012:8(5):477–485. 10.1038/nchembio.926 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chapman JM, Muhlemann JK, Gayomba SR, Muday GK. RBOH-dependent ROS synthesis and ROS scavenging by plant specialized metabolites to modulate plant development and stress responses. Chem Res Toxicol. 2019:32(3):370–396. 10.1021/acs.chemrestox.9b00028 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen X, Li L, Xu B, Zhao S, Lu P, He Y, Ye T, Feng YQ, Wu Y. Phosphatidylinositol-specific phospholipase C2 functions in auxin-modulated root development. Plant Cell Environ. 2019:42(5):1441–1457. 10.1111/pce.13492 [DOI] [PubMed] [Google Scholar]
- Chen ZF, Ru JN, Sun GZ, Du Y, Chen J, Zhou YB, Chen M, Ma YZ, Xu ZS, Zhang XH. Genomic-wide analysis of the PLC family and detection of GmPI-PLC7 responses to drought and salt stresses in soybean. Front Plant Sci. 2021:12:631470. 10.3389/fpls.2021.631470 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chung T. How phosphoinositides shape autophagy in plant cells. Plant Sci. 2019:281:146–158. 10.1016/j.plantsci.2019.01.017 [DOI] [PubMed] [Google Scholar]
- Colin LA, Jaillais Y. Phospholipids across scales: lipid patterns and plant development. Curr Opin Plant Biol. 2020:53:1–9. 10.1016/j.pbi.2019.08.007 [DOI] [PubMed] [Google Scholar]
- D'Ambrosio JM, Couto D, Fabro G, Scuffi D, Lamattina L, Munnik T, Andersson MX, Alvarez ME, Zipfel C, Laxalt AM. Phospholipase C2 affects MAMP-triggered immunity by modulating ROS production. Plant Physiol. 2017:175(2):970–981. 10.1104/pp.17.00173 [DOI] [PMC free article] [PubMed] [Google Scholar]
- D'Ambrosio JM, Gonorazky G, Sueldo DJ, Moraga J, Di Palma AA, Lamattina L, Collado IG, Laxalt AM. The sesquiterpene botrydial from Botrytis cinerea induces phosphatidic acid production in tomato cell suspensions. Planta. 2018:247(4):1001–1009. 10.1007/s00425-018-2843-8 [DOI] [PubMed] [Google Scholar]
- Das S, Hussain A, Bock C, Keller WA, Georges F. Cloning of brassica napus phospholipase C2 (BnPLC2), phosphatidylinositol 3-kinase (BnVPS34) and phosphatidylinositol synthase1 (BnPtdInsS1): comparative analysis of the effect of abiotic stresses on the expression of phosphatidylinositol signal transduction-related genes in B. napus. Planta. 2005:220:777–784. 10.1007/s00425-004-1389-0 [DOI] [PubMed] [Google Scholar]
- de Jong C, Laxalt AM, Bargmann BOR, de Wit PJGM, Joosten MHAJ, Munnik T. Phosphatidic acid accumulation is an early response in the Cf-4/Avr4 interaction. Plant J. 2004:39(1):1–12. 10.1111/j.1365-313X.2004.02110.x [DOI] [PubMed] [Google Scholar]
- de Jong F, Munnik T. Attracted to membranes: lipid-binding domains in plants. Plant Physiol. 2021:185(3):707–723. 10.1093/plphys/kiaa100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng X, Yuan S, Cao H, Lam SM, Shui G, Hong Y, Wang X. Phosphatidylinositol-hydrolyzing phospholipase C4 modulates rice response to salt and drought. Plant Cell Environ. 2019:42(2):536–548. 10.1111/pce.13437 [DOI] [PubMed] [Google Scholar]
- den Hartog M, Musgrave A, Munnik T. Nod factor-induced phosphatidic acid and diacylglycerol pyrophosphate formation: a role for phospholipase C and D in root hair deformation. Plant J. 2001:25(1):55–65. 10.1046/j.1365-313x.2001.00931.x [DOI] [PubMed] [Google Scholar]
- den Hartog M, Verhoef N, Munnik T. Nod factor and elicitors activate different phospholipid signaling pathways in suspension-cultured alfalfa cells. Plant Physiol. 2003:132(1):311–317. 10.1104/pp.102.017954 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dieck CB, Boss WF, Perera IY. A role for phosphoinositides in regulating plant nuclear functions. Front Plant Sci. 2012:3:50. 10.3389/fpls.2012.00050 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Di Fino LM, D’Ambrosio JM, Tejos R, van Wijk R, Lamattina L, Munnik T, Pagnussat GC, Laxalt AM. Arabidopsis phosphatidylinositol-phospholipase C2 (PLC2) is required for female gametogenesis and embryo development. Planta. 2017:245(4):717–728. 10.1007/s00425-016-2634-z [DOI] [PubMed] [Google Scholar]
- Distéfano AM, García-Mata C, Lamattina L, Laxalt AM. Nitric oxide-induced phosphatidic acid accumulation: a role for phospholipases C and D in stomatal closure. Plant Cell Environ. 2008:31(2):187–194. 10.1111/j.1365-3040.2007.01756.x [DOI] [PubMed] [Google Scholar]
- Dong J, Ma G, Sui L, Wei M, Satheesh V, Zhang R, Ge S, Li J, Zhang TE, Wittwer C, et al. Inositol pyrophosphate InsP8 acts as an intracellular phosphate signal in arabidopsis. Mol Plant. 2019:12(11):1463–1473. 10.1016/j.molp.2019.08.002 [DOI] [PubMed] [Google Scholar]
- Doumane M, Lebecq A, Colin L, Fangain A, Stevens FD, Bareille J, Hamant O, Belkhadir Y, Munnik T, Jaillais Y, et al. Inducible depletion of PI(4,5)P(2) by the synthetic iDePP system in Arabidopsis. Nat Plants. 2021:7(5):587–597. 10.1038/s41477-021-00907-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dowd PE, Coursol S, Skirpan AL, Kao TH, Gilroy S. Petunia phospholipase C1 is involved in pollen tube growth. Plant Cell. 2006:18(6):1438–1453. 10.1105/tpc.106.041582 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Drøbak BK, Dewey RE, Boss WF. Phosphoinositide kinases and the synthesis of polyphosphoinositides in higher plant cells. Int Rev Cytol. 1999:189:95–130. 10.1016/S0074-7696(08)61386-8 [DOI] [PubMed] [Google Scholar]
- Edel KH, Kudla J. Increasing complexity and versatility: How the calcium signaling toolkit was shaped during plant land colonization. Cell Calcium. 2015:57(3):231–246. 10.1016/j.ceca.2014.10.013 [DOI] [PubMed] [Google Scholar]
- Ellis MV, James SR, Perisic O, Downes CP, Williams RL, Katan M. Catalytic domain of phosphoinositide-specific phospholipase C (PLC)—mutational analysis of residues within the active site and hydrophobic ridge of PLC delta 1. J Biol Chem. 1998:273(19):11650–11659. 10.1074/jbc.273.19.11650 [DOI] [PubMed] [Google Scholar]
- Essen LO, Perisic O, Cheung R, Katan M, Williams RL. Crystal structure of a mammalian phosphoinositide-specific phospholipase Cδ. Nature. 1996:380(6575):595–602. 10.1038/380595a0 [DOI] [PubMed] [Google Scholar]
- Fang Y, Jiang J, Ding H, Li X, Xie X. Phospholipase C: diverse functions in plant biotic stress resistance and fungal pathogenicity. Mol Plant Pathol. 2023:24(9):1192–1202. 10.1111/mpp.13343 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flores S, Smart CC. Abscisic acid-induced changes in inositol metabolism in Spirodela polyrrhiza. Planta. 2000:211(6):823–832. 10.1007/s004250000348 [DOI] [PubMed] [Google Scholar]
- Fratini M, Krishnamoorthy P, Stenzel I, Riechmann M, Matzner M, Bacia K, Heilmann M, Heilmann I. Plasma membrane nano-organization specifies phosphoinositide effects on Rho-GTPases and actin dynamics in tobacco pollen tubes. Plant Cell. 2021:33(3):642–670. 10.1093/plcell/koaa035 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gagne JM, Clark SE. The Arabidopsis stem cell factor POLTERGEIST is membrane localized and phospholipid stimulated. Plant Cell. 2010:22(3):729–743. 10.1105/tpc.109.068734 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao HB, Chu YJ, Xue HW. Phosphatidic acid (PA) binds PP2AA1 to regulate PP2A activity and PIN1 polar localization. Mol Plant. 2013:6(5):1692–1702. 10.1093/mp/sst076 [DOI] [PubMed] [Google Scholar]
- Gao K, Liu YL, Li B, Zhou RG, Sun DY, Zheng SZ. Arabidopsis thaliana phosphoinositide-specific phospholipase C isoform 3 (AtPLC3) and AtPLC9 have an additive effect on thermotolerance. Plant Cell Physiol. 2014:55(11):1873–1883. 10.1093/pcp/pcu116 [DOI] [PubMed] [Google Scholar]
- Georges F, Das S, Ray H, Bock C, Nokhrina K, Kolla VA, Keller W. Over-expression of Brassica napus phosphatidylinositol-phospholipase C2 in canola induces significant changes in gene expression and phytohormone distribution patterns, enhances drought tolerance and promotes early flowering and maturation. Plant Cell Environ. 2009:32(12):1664–1681. 10.1111/j.1365-3040.2009.02027.x [DOI] [PubMed] [Google Scholar]
- Gerth K, Lin F, Daamen F, Menzel W, Heinrich F, Heilmann M. Arabidopsis phosphatidylinositol 4-phosphate 5-kinase 2 contains a functional nuclear localization sequence and interacts with alpha-importins. Plant J. 2017:92(5):862–878. 10.1111/tpj.13724 [DOI] [PubMed] [Google Scholar]
- Gilroy S, Bialasek M, Suzuki N, Gorecka M, Devireddy AR, Karpinski S, Mittler R. ROS, calcium, and electric signals: key mediators of rapid systemic signaling in plants. Plant Physiol. 2016:171(3):1606–1615. 10.1104/pp.16.00434 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gilroy S, Read ND, Trewavas AJ. Elevation of cytoplasmic calcium by caged calcium or caged inositol triphosphate initiates stomatal closure. Nature. 1990:346(6286):769–771. 10.1038/346769a0 [DOI] [PubMed] [Google Scholar]
- Goldberg A, O’Connor P, Gonzalez C, Ouren M, Rivera L, Radde N, Nguyen M, Ponce-Herrera F, Lloyd A, Gonzalez A. Genetic interaction between TTG2 and AtPLC1 reveals a role for phosphoinositide signaling in a co-regulated suite of Arabidopsis epidermal pathways. Sci Rep. 2024:14(1):9752. 10.1038/s41598-024-60530-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gonorazky G, Guzzo MC, Abd-El-Haliem AM, Joosten MH, Laxalt AM. Silencing of the tomato phosphatidylinositol-phospholipase C2 (SlPLC2) reduces plant susceptibility to Botrytis cinerea. Mol Plant Pathol. 2016:17(9):1354–1363. 10.1111/mpp.12365 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gonorazky G, Ramirez L, Abd-El-Haliem A, Vossen JH, Lamattina L, ten Have A, Joosten MH, Laxalt AM. The tomato phosphatidylinositol-phospholipase C2 (SlPLC2) is required for defense gene induction by the fungal elicitor xylanase. J Plant Physiol. 2014:171(11):959–965. 10.1016/j.jplph.2014.02.008 [DOI] [PubMed] [Google Scholar]
- Gonzalez-Mendoza VM, Sanchez-Sandoval ME, Munnik T, Hernandez-Sotomayor SMT. Biochemical characterization of phospholipases C from Coffea arabica in response to aluminium stress. J Inorg Biochem. 2020:204:110951. 10.1016/j.jinorgbio.2019.110951 [DOI] [PubMed] [Google Scholar]
- Graber Z, Owusu Kwarteng D, Lange SM, Koukanas Y, Khalifa H, Mutambuze JW, Kooijman EE. The electrostatic basis of diacylglycerol pyrophosphate-protein interaction. Cells. 2022:11(2):290. 10.3390/cells11020290 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gronnier J, Crowet JM, Habenstein B, Nasir MN, Bayle V, Hosy E, Platre MP, Gouguet P, Raffaele S, Martinez D, et al. Structural basis for plant plasma membrane protein dynamics and organization into functional nanodomains. eLife. 2017:6:e26404. 10.7554/eLife.26404 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guan Z, Chen J, Liu R, Chen Y, Xing Q, Du Z, Cheng M, Hu J, Zhang W, Mei W, et al. The cytoplasmic synthesis and coupled membrane translocation of eukaryotic polyphosphate by signal-activated VTC complex. Nat Commun. 2023:14(1):718. 10.1038/s41467-023-36466-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guan Z, Zhang Q, Zhang Z, Zuo J, Chen J, Liu R, Savarin J, Broger L, Cheng P, Wang Q, et al. Mechanistic insights into the regulation of plant phosphate homeostasis by the rice SPX2—PHR2 complex. Nat Commun. 2022:13(1):1581. 10.1038/s41467-022-29275-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gujas B, Cruz TMD, Kastanaki E, Vermeer JEM, Munnik T, Rodriguez-Villalon A. Perturbing phosphoinositide homeostasis oppositely affects root vascular differentiation in Arabidopsis thaliana roots. Development. 2017:144(19):3578–3589. 10.1242/dev.155788 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo T, Chen HC, Lu ZQ, Diao M, Chen K, Dong NQ, Shan JX, Ye WW, Huang S, Lin HX. A SAC phosphoinositide phosphatase controls rice development via hydrolyzing PI4P and PI(4,5)P2. Plant Physiol. 2020:182(3):1346–1358. 10.1104/pp.19.01131 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hao GF, Yang GF. The role of Phe82 and Phe351 in auxin-induced substrate perception by TIR1 ubiquitin ligase: a novel insight from molecular dynamics simulations. PLoS One. 2010:5(5):e10742. 10.1371/journal.pone.0010742 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heilmann I. Phosphoinositide signaling in plant development. Development. 2016a:143(12):2044–2055. 10.1242/dev.136432 [DOI] [PubMed] [Google Scholar]
- Heilmann I. Plant phosphoinositide signaling—dynamics on demand. Biochim Biophys Acta. 2016b:1861(9):1345–1351. 10.1016/j.bbalip.2016.02.013 [DOI] [PubMed] [Google Scholar]
- Heilmann M, Heilmann I. Regulators regulated: different layers of control for plasma membrane phosphoinositides in plants. Curr Opin Plant Biol. 2022:67:102218. 10.1016/j.pbi.2022.102218 [DOI] [PubMed] [Google Scholar]
- Heilmann M, Heilmann I. Getting attached to membranes - How plant signalling networks employ PtdIns(4,5)P2. Plant Physiol. 2024:kiae393. 10.1093/plphys/kiae393 [DOI] [PubMed] [Google Scholar]
- Helling D, Possart A, Cottier S, Klahre U, Kost B. Pollen tube tip growth depends on plasma membrane polarization mediated by tobacco PLC3 activity and endocytic membrane recycling. Plant Cell. 2006:18(12):3519–3534. 10.1105/tpc.106.047373 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hemsley PA, Weimar T, Lilley KS, Dupree P, Grierson CS. A proteomic approach identifies many novel palmitoylated proteins in Arabidopsis. New Phytol. 2013:197(3):805–814. 10.1111/nph.12077 [DOI] [PubMed] [Google Scholar]
- Hicks SN, Jezyk MR, Gershburg S, Seifert JP, Harden TK, Sondek J. General and Versatile Autoinhibition of PLC Isozymes. Mol Cell. 2008:31(3):383–394. 10.1016/j.molcel.2008.06.018 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirano T, Munnik T, Sato MH. Inhibition of phosphatidylinositol 3,5-bisphosphate production has pleiotropic effects on various membrane trafficking routes in Arabidopsis. Plant Cell Physiol. 2017a:58(1):120–129. 10.1093/pcp/pcw164 [DOI] [PubMed] [Google Scholar]
- Hirano T, Sato MH. Diverse physiological functions of FAB1 and phosphatidylinositol 3,5-bisphosphate in plants. Front Plant Sci. 2019:10:274. 10.3389/fpls.2019.00274 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirano T, Stecker K, Munnik T, Xu H, Sato MH. Visualization of phosphatidylinositol 3,5-bisphosphate dynamics by a tandem ML1N-based fluorescent protein probe in Arabidopsis. Plant Cell Physiol. 2017b:58(7):1185–1195. 10.1093/pcp/pcx011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirayama T, Ohto C, Mizoguchi T, Shinozaki K. A gene encoding a phosphatidylinositol-specific phospholipase C is induced by dehydration and salt stress in Arabidopsis thaliana. Proc Natl Acad Sci U S A. 1995:92(9):3903–3907. 10.1073/pnas.92.9.3903 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirt H. Transcriptional upregulation of signaling pathways: more complex than anticipated? Trends Plant Sci. 1999:4(1):7–8. 10.1016/s1360-1385(98)01359-4 [DOI] [PubMed] [Google Scholar]
- Hong Y, Zhao J, Guo L, Kim S-C, Deng X, Wang G, Zhang G, Li M, Wang X. Plant phospholipases D and C and their diverse functions in stress responses. Prog Lipid Res. 2016:62:55–74. 10.1016/j.plipres.2016.01.002 [DOI] [PubMed] [Google Scholar]
- Hou Q, Ufer G, Bartels D. Lipid signaling in plant responses to abiotic stress. Plant Cell Environ. 2016:39(5):1029–1048. 10.1111/pce.12666 [DOI] [PubMed] [Google Scholar]
- Hsieh EJ, Cheng MC, Lin TP. Functional characterization of an abiotic stress-inducible transcription factor AtERF53 in Arabidopsis thaliana. Plant Mol Biol. 2013:82(3):223–237. 10.1007/s11103-013-0054-z [DOI] [PubMed] [Google Scholar]
- Hunt L, Otterhag L, Lee J, Lasheen T, Hunt J, Seki M, Shinozaki K, Sommarin M, Gilmour D, Pical C. Gene-specific expression and calcium activation of Arabidopsis thaliana phospholipase C isoforms. New Phytol. 2004:162(3):643–654. 10.1111/j.1469-8137.2004.01069.x [DOI] [PubMed] [Google Scholar]
- Iqbal S, Ali U, Fadlalla T, Li Q, Liu H, Lu S, Guo L. Genome wide characterization of phospholipase A & C families and pattern of lysolipids and diacylglycerol changes under abiotic stresses in Brassica napus L. Plant Physiol Biochem. 2020:147:101–112. 10.1016/j.plaphy.2019.12.017 [DOI] [PubMed] [Google Scholar]
- Ischebeck T, Seiler S, Heilmann I. At the Poles across kingdoms: phosphoinositides and polar tip growth. Protoplasma. 2010:240(1–4):13–31. 10.1007/s00709-009-0093-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ischebeck T, Werner S, Krishnamoorthy P, Lerche J, Meijon M, Stenzel I, Lofke C, Wiessner T, Im YJ, Perera IY, et al. Phosphatidylinositol 4,5-bisphosphate influences PIN polarization by controlling clathrin-mediated membrane trafficking in Arabidopsis. Plant Cell. 2013:25(12):4894–4911. 10.1105/tpc.113.116582 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jaillais Y, Vert G. Brassinosteroid signaling and BRI1 dynamics went underground. Curr Opin Plant Biol. 2016:33:92–100. 10.1016/j.pbi.2016.06.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jia X, Wang L, Zeng H, Yi K. Insights of intracellular/intercellular phosphate transport and signaling in unicellular green algae and multicellular land plants. New Phytol. 2021:232(4):1566–1571. 10.1111/nph.17716 [DOI] [PubMed] [Google Scholar]
- Julkowska MM, McLoughlin F, Galvan-Ampudia CS, Rankenberg JM, Kawa D, Klimecka M, Haring MA, Munnik T, Kooijman EE, Testerink C. Identification and functional characterization of the Arabidopsis Snf1-related protein kinase SnRK2.4 phosphatidic acid-binding domain. Plant Cell Environ. 2015:38(3):614–624. 10.1111/pce.12421 [DOI] [PubMed] [Google Scholar]
- Jung JY, Ried MK, Hothorn M, Poirier Y. Control of plant phosphate homeostasis by inositol pyrophosphates and the SPX domain. Curr Opin Biotechnol. 2018:49:156–162. 10.1016/j.copbio.2017.08.012 [DOI] [PubMed] [Google Scholar]
- Kadota Y, Shirasu K, Zipfel C. Regulation of the NADPH oxidase RBOHD during plant immunity. Plant Cell Physiol. 2015:56(8):1472–1480. 10.1093/pcp/pcv063 [DOI] [PubMed] [Google Scholar]
- Kalachova T, Škrabálková E, Pateyron S, Soubigou-Taconnat L, Djafi N, Collin S, Sekereš J, Burketová L, Potocký M, Pejchar P, et al. DIACYLGLYCEROL KINASE 5 participates in flagellin-induced signaling in Arabidopsis. Plant Physiol. 2022:190(3):1978–1996. 10.1093/plphys/kiac354 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kanchan M, Ramkumar TR, Himani, Sembi JK. Genome-wide characterization and expression profiling of the phospholipase C (PLC) gene family in three orchids of economic importance. J Genet Eng Biotechnol. 2021:19(1):124. 10.1186/s43141-021-00217-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kanehara K, Yu C-Y, Cho Y, Cheong W-F, Torta F, Shui G, Wenk MR, Nakamura Y. Arabidopsis AtPLC2 is a primary phosphoinositide-specific phospholipase C in phosphoinositide metabolism and the endoplasmic reticulum stress response. PLoS Genet. 2015:11(9):e1005511. 10.1371/journal.pgen.1005511 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Katan M, Cockcroft S. Phospholipase C families: common themes and versatility in physiology and pathology. Prog Lipid Res. 2020:80:101065. 10.1016/j.plipres.2020.101065 [DOI] [PubMed] [Google Scholar]
- Kato M, Tsuge T, Maeshima M, Aoyama T. Arabidopsis PCaP2 modulates the phosphatidylinositol 4,5-bisphosphate signal on the plasma membrane and attenuates root hair elongation. Plant J. 2019:99(4):610–625. 10.1111/tpj.14226 [DOI] [PubMed] [Google Scholar]
- Khalil HB, Wang Z, Wright JA, Ralevski A, Donayo AO, Gulick PJ. Heterotrimeric Gα subunit from wheat (Triticum aestivum), GA3, interacts with the calcium-binding protein, Clo3, and the phosphoinositide-specific phospholipase C, PI-PLC1. Plant Mol Biol. 2011:77(1–2):145–158. 10.1007/s11103-011-9801-1 [DOI] [PubMed] [Google Scholar]
- Kiba A, Nakano M, Hosokawa M, Galis I, Nakatani H, Shinya T, Ohnishi K, Hikichi Y, Spoel S. Phosphatidylinositol-phospholipase C2 regulates pattern-triggered immunity in Nicotiana benthamiana. J Exp Bot. 2020:71(16):5027–5038. 10.1093/jxb/eraa233 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim SC, Nusinow DA, Sorkin ML, Pruneda-Paz J, Wang X. Interaction and regulation between lipid mediator phosphatidic acid and circadian clock regulators. Plant Cell. 2019:31(2):399–416. 10.1105/tpc.18.00675 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim YJ, Kim JE, Lee J-H, Lee MH, Jung HW, Bahk YY, Hwang BK, Hwang I, Kim WT. The V-PLC3 gene encodes a putative plasma membrane-localized phosphoinositide-specific phospholipase C whose expression is induced by abiotic stress in mung bean (Vigna radiata L.) 1. FEBS Lett. 2004:556(1–3):127–136. 10.1016/S0014-5793(03)01388-7 [DOI] [PubMed] [Google Scholar]
- Kohorn BD, Hoon D, Minkoff BB, Sussman MR, Kohorn SL. Rapid oligo-galacturonide induced changes in protein phosphorylation in arabidopsis. Mol Cell Proteomics. 2016:15(4):1351–1359. 10.1074/mcp.M115.055368 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kolesnikov Y, Kretynin S, Bukhonska Y, Pokotylo I, Ruelland E, Martinec J, Kravets V. Phosphatidic acid in plant hormonal signaling: from target proteins to membrane conformations. Int J Mol Sci. 2022:23(6):3227. 10.3390/ijms23063227 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kong L, Ma X, Zhang C, Kim SI, Li B, Xie Y, Yeo IC, Thapa H, Chen S, Devarenne TP, et al. Dual phosphorylation of DGK5-mediated PA burst regulates ROS in plant immunity. Cell. 2024:187(3):609–623. 10.1016/j.cell.2023.12.030 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Konig S, Mosblech A, Heilmann I. Stress-inducible and constitutive phosphoinositide pools have distinctive fatty acid patterns in Arabidopsis thaliana. FASEB J. 2007:21(9):1958–1967. 10.1096/fj.06-7887com [DOI] [PubMed] [Google Scholar]
- Kopka J, Pical C, Gray JE, Muller-Rober B. Molecular and enzymatic characterization of three phosphoinositide-specific phospholipase C isoforms from potato. Plant Physiol. 1998:116(1):239–250. 10.1104/pp.116.1.239 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krinke O, Novotna Z, Valentova O, Martinec J. Inositol trisphosphate receptor in higher plants: is it real? J Exp Bot. 2007:58(3):361–376. 10.1093/jxb/erl220 [DOI] [PubMed] [Google Scholar]
- Kumar M, Carr P, Turner SR. An atlas of Arabidopsis protein S-acylation reveals its widespread role in plant cell organization and function. Nat Plants. 2022:8(6):670–681. 10.1038/s41477-022-01164-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuroda R, Kato M, Tsuge T, Aoyama T. Arabidopsis phosphatidylinositol 4-phosphate 5-kinase genes PIP5K7, PIP5K8, and PIP5K9 are redundantly involved in root growth adaptation to osmotic stress. Plant J. 2021:106(4):913–927. 10.1111/tpj.15207 [DOI] [PubMed] [Google Scholar]
- Laha D, Johnen P, Azevedo C, Dynowski M, Weiss M, Capolicchio S, Mao H, Iven T, Steenbergen M, Freyer M, et al. VIH2 regulates the synthesis of inositol pyrophosphate insp8 and jasmonate-dependent defenses in Arabidopsis. Plant Cell. 2015:27(4):1082–1097. 10.1105/tpc.114.135160 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laha D, Parvin N, Dynowski M, Johnen P, Mao H, Bitters ST, Zheng N, Schaaf G. Inositol polyphosphate binding specificity of the jasmonate receptor complex. Plant Physiol. 2016:171(4):2364–2370. 10.1104/pp.16.00694 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laha NP, Giehl RFH, Riemer E, Qiu D, Pullagurla NJ, Schneider R, Dhir YW, Yadav R, Mihiret YE, Gaugler P, et al. INOSITOL (1,3,4) TRIPHOSPHATE 5/6 KINASE1-dependent inositol polyphosphates regulate auxin responses in Arabidopsis. Plant Physiol. 2022:190(4):2722–2738. 10.1093/plphys/kiac425 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laxalt AM, Munnik T. Phospholipid signalling in plant defence. Curr Opin Plant Biol. 2002:5(4):332–338. 10.1016/S1369-5266(02)00268-6 [DOI] [PubMed] [Google Scholar]
- Laxalt AM, Raho N, Have AT, Lamattina L. Nitric oxide is critical for inducing phosphatidic acid accumulation in xylanase-elicited tomato cells. J Biol Chem. 2007:282(29):21160–21168. 10.1074/jbc.M701212200 [DOI] [PubMed] [Google Scholar]
- Lebecq A, Doumane M, Fangain A, Bayle V, Leong JX, Rozier F, Marques-Bueno MD, Armengot L, Boisseau R, Simon ML, et al. The Arabidopsis SAC9 enzyme is enriched in a cortical population of early endosomes and restricts PI(4,5)P2 at the plasma membrane. eLife. 2022:11:e73837. 10.7554/eLife.73837 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee HS, Lee DH, Cho HK, Kim SH, Auh JH, Pai HS. InsP6-sensitive variants of the Gle1 mRNA export factor rescue growth and fertility defects of the ipk1 low-phytic-acid mutation in Arabidopsis. Plant Cell. 2015:27(2):417–431. 10.1105/tpc.114.132134 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee NH, Zarza X, Kim JH, Yoon MJ, Kim S-H, Lee J-H, Paris N, Munnik T, Otegui MS, Chung T. Vacuolar trafficking protein VPS38 is dispensable for autophagy. Plant Physiol. 2018:176(2):1559–1572. 10.1104/pp.17.01297 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee Y, Kim YW, Jeon BW, Park KY, Suh SJ, Seo J, Kwak JM, Martinoia E, Hwang I, Lee Y. Phosphatidylinositol 4,5-bisphosphate is important for stomatal opening. Plant J. 2007:52(5):803–816. 10.1111/j.1365-313X.2007.03277.x [DOI] [PubMed] [Google Scholar]
- Lee Y, Lee Y. Roles of phosphoinositides in regulation of stomatal movements. Plant Signal Behav. 2008:3(4):211–213. 10.4161/psb.3.4.5557 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lemtiri-Chlieh F, MacRobbie EA, Brearley CA. Inositol hexakisphosphate is a physiological signal regulating the K+-inward rectifying conductance in guard cells. Proc Natl Acad Sci U S A. 2000:97(15):8687–8692. 10.1073/pnas.140217497 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lemtiri-Chlieh F, MacRobbie EA, Webb AA, Manison NF, Brownlee C, Skepper JN, Chen J, Prestwich GD, Brearley CA. Inositol hexakisphosphate mobilizes an endomembrane store of calcium in guard cells. Proc Natl Acad Sci U S A. 2003:100(17):10091–10095. 10.1073/pnas.1133289100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J, Wang X. Phospholipase D and phosphatidic acid in plant immunity. Plant Sci. 2019:279:45–50. 10.1016/j.plantsci.2018.05.021 [DOI] [PubMed] [Google Scholar]
- Li L, He Y, Wang Y, Zhao S, Chen X, Ye T, Wu Y, Wu Y. Arabidopsis PLC2 is involved in auxin-modulated reproductive development. Plant J. 2015:84(3):504–515. 10.1111/tpj.13016 [DOI] [PubMed] [Google Scholar]
- Li L, Wang F, Yan P, Jing W, Zhang C, Kudla J, Zhang W. A phosphoinositide-specific phospholipase C pathway elicits stress-induced Ca2+ signals and confers salt tolerance to rice. New Phytol. 2017:214(3):1172–1187. 10.1111/nph.14426 [DOI] [PubMed] [Google Scholar]
- Li M, Hong Y, Wang X. Phospholipase D- and phosphatidic acid-mediated signaling in plants. Biochim Biophys Acta. 2009:1791(9):927–935. 10.1016/j.bbalip.2009.02.017 [DOI] [PubMed] [Google Scholar]
- Li W, Song T, Wallrad L, Kudla J, Wang X, Zhang W. Tissue-specific accumulation of pH-sensing phosphatidic acid determines plant stress tolerance. Nat Plants. 2019:5(9):1012–1021. 10.1038/s41477-019-0497-6 [DOI] [PubMed] [Google Scholar]
- Li-Beisson Y, Shorrosh B, Beisson F, Andersson MX, Arondel V, Bates PD, Baud S, Bird D, Debono A, Durrett TP, et al. Acyl-lipid metabolism. Arabidopsis Book. 2013:11:e0161. 10.1199/tab.0161 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin F, Krishnamoorthy P, Schubert V, Hause G, Heilmann M, Heilmann I. A dual role for cell plate-associated PI4Kbeta in endocytosis and phragmoplast dynamics during plant somatic cytokinesis. EMBO J. 2019:38(4):e100303. 10.15252/embj.2018100303 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin F, Zheng J, Xie Y, Jing W, Zhang Q, Zhang W. Emerging roles of phosphoinositide-associated membrane trafficking in plant stress responses. J Genet Genomics. 2022:49(8):726–734. 10.1016/j.jgg.2022.05.003 [DOI] [PubMed] [Google Scholar]
- Liu F, Hu W, Li F, Marshall RS, Zarza X, Munnik T, Vierstra RD. AUTOPHAGY-RELATED14 and its associated phosphatidylinositol 3-kinase complex promotes autophagy in Arabidopsis. Plant Cell. 2020a:32(12):3939–3960. 10.1105/tpc.20.00285 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu YL, Liu XY, Wang X, Gao K, Qi WW, Ren HM, Hu HR, Sun DY, Bai JT, Zheng SZ. Heterologous expression of heat stress-responsive AtPLC9 confers heat tolerance in transgenic rice. BMC Plant Biol. 2020b:20(1):514. 10.1186/s12870-020-02709-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lofke C, Ischebeck T, Konig S, Freitag S, Heilmann I. Alternative metabolic fates of phosphatidylinositol produced by phosphatidylinositol synthase isoforms in Arabidopsis thaliana. Biochem J. 2008:413(1):115–124. 10.1042/BJ20071371 [DOI] [PubMed] [Google Scholar]
- Lorenzo-Orts L, Couto D, Hothorn M. Identity and functions of inorganic and inositol polyphosphates in plants. New Phytol. 2020:225(2):637–652. 10.1111/nph.16129 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luo P, Jiang A, Zhou Y, Yang M, Zhou X, Yang Y, Yu J, Tang X. Phospholipase C is a novel regulator at the early stages of microspore embryogenesis in Nicotiana tabacum. Plant Signal Behav. 2022:17(1):2094618. 10.1080/15592324.2022.2094618 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma Y, Szostkiewicz I, Korte A, Moes D, Yang Y, Christmann A, Grill E. Regulators of PP2C phosphatase activity function as abscisic acid sensors. Science. 2009:324(5930):1064–1068. 10.1126/science.1172408 [DOI] [PubMed] [Google Scholar]
- Marhava P, Aliaga Fandino AC, Koh SWH, Jelínková A, Kolb M, Janacek DP, Breda AS, Cattaneo P, Hammes UZ, Petrášek J, et al. Plasma membrane domain patterning and self-reinforcing polarity in Arabidopsis. Dev Cell. 2020:52(2):223–235. 10.1016/j.devcel.2019.11.015 [DOI] [PubMed] [Google Scholar]
- McLoughlin F, Arisz SA, Dekker HL, Kramer G, de Koster CG, Haring MA, Munnik T, Testerink C. Identification of novel candidate phosphatidic acid-binding proteins involved in the salt-stress response of Arabidopsis thaliana roots. Biochem J. 2013:450(3):573–581. 10.1042/BJ20121639 [DOI] [PubMed] [Google Scholar]
- McLoughlin F, Galvan-Ampudia CS, Julkowska MM, Caarls L, van der Does D, Laurière C, Munnik T, Haring MA, Testerink C. The Snf1-related protein kinases SnRK2.4 and SnRK2.10 are involved in maintenance of root system architecture during salt stress. Plant J. 2012:72(3):436–449. 10.1111/j.1365-313X.2012.05089.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mei Y, Jia WJ, Chu YJ, Xue HW. Arabidopsis phosphatidylinositol monophosphate 5-kinase 2 is involved in root gravitropism through regulation of polar auxin transport by affecting the cycling of PIN proteins. Cell Res. 2012:22(3):581–597. 10.1038/cr.2011.150 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meijer HJ, Arisz SA, Van Himbergen JA, Musgrave A, Munnik T. Hyperosmotic stress rapidly generates lyso-phosphatidic acid in Chlamydomonas. Plant J. 2001a:25(5):541–548. 10.1046/j.1365-313x.2001.00990.x [DOI] [PubMed] [Google Scholar]
- Meijer HJ, Berrie CP, Iurisci C, Divecha N, Musgrave A, Munnik T. Identification of a new polyphosphoinositide in plants, phosphatidylinositol 5-monophosphate (PtdIns5P), and its accumulation upon osmotic stress. Biochem J. 2001b:360(2):491–498. 10.1042/bj3600491 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meijer HJ, Munnik T. Phospholipid-based signaling in plants. Annu Rev Plant Biol. 2003:54(1):265–306. 10.1146/annurev.arplant.54.031902.134748 [DOI] [PubMed] [Google Scholar]
- Meijer HJ, van Himbergen JA, Musgrave A, Munnik T. Acclimation to salt modifies the activation of several osmotic stress-activated lipid signaling pathways in Chlamydomonas. Phytochemistry. 2017:135:64–72. 10.1016/j.phytochem.2016.12.014 [DOI] [PubMed] [Google Scholar]
- Meijer HJG, Divecha N, van den Ende H, Musgrave A, Munnik T. Hyperosmotic stress induces rapid synthesis of phosphatidyl-D-inositol 3,5-bisphosphate in plant cells. Planta. 1999:208(2):294–298. 10.1007/s004250050561 [DOI] [Google Scholar]
- Mills LN, Hunt L, Leckie CP, Aitken FL, Wentworth M, McAinsh MR, Gray JE, Hetherington AM. The effects of manipulating phospholipase C on guard cell ABA-signaling. J Exp Bot. 2004:55(395):199–204. 10.1093/jxb/erh027 [DOI] [PubMed] [Google Scholar]
- Mishkind M, Vermeer JE, Darwish E, Munnik T. Heat stress activates phospholipase D and triggers PIP accumulation at the plasma membrane and nucleus. Plant J. 2009:60(1):10–21. 10.1111/j.1365-313X.2009.03933.x [DOI] [PubMed] [Google Scholar]
- Mishra G, Zhang W, Deng F, Zhao J, Wang X. A bifurcating pathway directs abscisic acid effects on stomatal closure and opening in Arabidopsis. Science. 2006:312(5771):264–266. 10.1126/science.1123769 [DOI] [PubMed] [Google Scholar]
- Misra S, Wu Y, Venkataraman G, Sopory SK, Tuteja N. Heterotrimeric G-protein complex and G-protein-coupled receptor from a legume (Pisum sativum): role in salinity and heat stress and cross-talk with phospholipase C. Plant J. 2007:51(4):656–669. 10.1111/j.1365-313X.2007.03169.x [DOI] [PubMed] [Google Scholar]
- Mosblech A, Thurow C, Gatz C, Feussner I, Heilmann I. Jasmonic acid perception by COI1 involves inositol polyphosphates in Arabidopsis thaliana. Plant J. 2011:65(6):949–957. 10.1111/j.1365-313X.2011.04480.x [DOI] [PubMed] [Google Scholar]
- Mueller-Roeber B, Pical C. Inositol phospholipid metabolism in arabidopsis. Characterized and putative isoforms of inositol phospholipid kinase and phosphoinositide-specific phospholipase C. Plant Physiol. 2002:130(1):22–46. 10.1104/pp.004770 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Munnik T. Phosphatidic acid: an emerging plant lipid second messenger. Trends Plant Science. 2001:6(5):227–233. 10.1016/S1360-1385(01)01918-5 [DOI] [PubMed] [Google Scholar]
- Munnik T. Analysis of D3-,4-,5-phosphorylated phosphoinositides using HPLC. Methods Mol Biol. 2013:1009:17–24. 10.1007/978-1-62703-401-2_2 [DOI] [PubMed] [Google Scholar]
- Munnik T. PI-PLC: phosphoinositide-phospholipase C in plant signaling. In: Wang X, editor. Phospholipases in plant signaling. Berlin-Heilderberg: Springer; 2014. p. 27–54. [Google Scholar]
- Munnik T, de Vrije T, Irvine RF, Musgrave A. Identification of diacylglycerol pyrophosphate as a novel metabolic product of phosphatidic acid during G-protein activation in plants. J Biol Chem. 1996:271(26):15708–15715. 10.1074/jbc.271.26.15708 [DOI] [PubMed] [Google Scholar]
- Munnik T, Irvine RF, Musgrave A. Rapid turnover of phosphatidylinositol 3-phosphate in the green alga Chlamydomonas eugametos: signs of a phosphatidylinositide 3-kinase signaling pathway in lower plants? Biochem J. 1994:298(2):269–273. 10.1042/bj2980269 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Munnik T, Irvine RF, Musgrave A. Phospholipid signaling in plants. Biochim Biophys Acta. 1998a:1389(3):222–272. 10.1016/S0005-2760(97)00158-6 [DOI] [PubMed] [Google Scholar]
- Munnik T, Meijer HJ, Ter Riet B, Hirt H, Frank W, Bartels D, Musgrave A. Hyperosmotic stress stimulates phospholipase D activity and elevates the levels of phosphatidic acid and diacylglycerol pyrophosphate. Plant J. 2000:22(2):147–154. 10.1046/j.1365-313x.2000.00725.x [DOI] [PubMed] [Google Scholar]
- Munnik T, Nielsen E. Green light for polyphosphoinositide signals in plants. Curr Opin Plant Biol. 2011:14(5):489–497. 10.1016/j.pbi.2011.06.007 [DOI] [PubMed] [Google Scholar]
- Munnik T, Testerink C. Plant phospholipid signaling: “in a nutshell”. J Lipid Res. 2009:50:S260–S265. 10.1194/jlr.R800098-JLR200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Munnik T, van Himbergen JAJ, ter Riet B, Braun F-J, Irvine RF, van den Ende H, Musgrave A. Detailed analysis of the turnover of polyphosphoinositides and phosphatidic acid upon activation of phospholipase C and -D in Chlamydomonas cells treated with non-permeabilizing concentrations of mastoparan. Planta. 1998b:207(1):133–145. 10.1007/s004250050465 [DOI] [Google Scholar]
- Munnik T, Vermeer JE. Osmotic stress-induced phosphoinositide and inositol phosphate signaling in plants. Plant Cell Environ. 2010:33(4):655–669. 10.1111/j.1365-3040.2009.02097.x [DOI] [PubMed] [Google Scholar]
- Munnik T, Zarza X. Analyzing plant signaling phospholipids through 32Pi-labeling and TLC. Methods Mol Biol. 2013:1009:3–15. 10.1007/978-1-62703-401-2_1 [DOI] [PubMed] [Google Scholar]
- Murphy AM, Otto B, Brearley CA, Carr JP, Hanke DE. A role for inositol hexakisphosphate in the maintenance of basal resistance to plant pathogens. Plant J. 2008:56(4):638–652. 10.1111/j.1365-313X.2008.03629.x [DOI] [PubMed] [Google Scholar]
- Nakamura Y, Koizumi R, Shui G, Shimojima M, Wenk MR, Ito T, Ohta H. Arabidopsis lipins mediate eukaryotic pathway of lipid metabolism and cope critically with phosphate starvation. Proc Natl Acad Sci U S A. 2009:106(49):20978–20983. 10.1073/pnas.0907173106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Naramoto S, Sawa S, Koizumi K, Uemura T, Ueda T, Friml J, Nakano A, Fukuda H. Phosphoinositide-dependent regulation of VAN3 ARF-GAP localization and activity essential for vascular tissue continuity in plants. Development. 2009:136(9):1529–1538. 10.1242/dev.030098 [DOI] [PubMed] [Google Scholar]
- Ng CK, McAinsh MR, Gray JE, Hunt L, Leckie CP, Mills L, Hetherington AM. Calcium-based signaling systems in guard cells. New Phytol. 2001:151(1):109–120. 10.1046/j.1469-8137.2001.00152.x [DOI] [PubMed] [Google Scholar]
- Noack LC, Jaillais Y. Precision targeting by phosphoinositides: how PIs direct endomembrane trafficking in plants. Curr Opin Plant Biol. 2017:40:22–33. 10.1016/j.pbi.2017.06.017 [DOI] [PubMed] [Google Scholar]
- Noack LC, Jaillais Y. Functions of anionic lipids in plants. Annu Rev Plant Biol. 2020:71(1):71–102. 10.1146/annurev-arplant-081519-035910 [DOI] [PubMed] [Google Scholar]
- Nuhse TS, Boller T, Peck SC. A plasma membrane syntaxin is phosphorylated in response to the bacterial elicitor flagellin. J Biol Chem. 2003:278(46):45248–45254. 10.1074/jbc.M307443200 [DOI] [PubMed] [Google Scholar]
- Nuhse TS, Bottrill AR, Jones AM, Peck SC. Quantitative phosphoproteomic analysis of plasma membrane proteins reveals regulatory mechanisms of plant innate immune responses. Plant J. 2007:51(5):931–940. 10.1111/j.1365-313X.2007.03192.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Otterhag L, Sommarin M, Pical C. N-terminal EF-hand-like domain is required for phosphoinositide-specific phospholipase C activity in Arabidopsis thaliana. FEBS Lett. 2001:497(2–3):165–170. 10.1016/S0014-5793(01)02453-X [DOI] [PubMed] [Google Scholar]
- Pandit S, Dalal V, Mishra G. Identification of novel phosphatidic acid binding domain on sphingosine kinase 1 of Arabidopsis thaliana. Plant Physiol Biochem. 2018:128:178–184. 10.1016/j.plaphy.2018.04.039 [DOI] [PubMed] [Google Scholar]
- Perk E, Arruebarrena Di Palma A, Colman S, Mariani O, Cerrudo I, D'Ambrosio J, Robuschi L, Pombo M, Rosli H, Villareal F, et al. CRISPR/Cas9-mediated phospholipase C 2 knock-out tomato plants are more resistant to Botrytis cinerea. Planta. 2023:257(6):17. 10.1007/s00425-023-04147-7 [DOI] [PubMed] [Google Scholar]
- Platre MP, Jaillais Y. Guidelines for the use of protein domains in acidic phospholipid imaging. Methods Mol Biol. 2016:1376:175–194. 10.1007/978-1-4939-3170-5_15 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Platre MP, Noack LC, Doumane M, Bayle V, Simon MLA, Maneta-Peyret L, Fouillen L, Stanislas T, Armengot L, Pejchar P, et al. A combinatorial lipid code shapes the electrostatic landscape of plant endomembranes. Dev Cell. 2018:45(4):465–480.e411. 10.1016/j.devcel.2018.04.011 [DOI] [PubMed] [Google Scholar]
- Pleskot R, Li J, Zarsky V, Potocky M, Staiger CJ. Regulation of cytoskeletal dynamics by phospholipase D and phosphatidic acid. Trends Plant Sci. 2013:18(9):496–504. 10.1016/j.tplants.2013.04.005 [DOI] [PubMed] [Google Scholar]
- Pokotylo I, Kolesnikov Y, Kravets V, Zachowski A, Ruelland E. Plant phosphoinositide-dependent phospholipases C: variations around a canonical theme. Biochimie. 2014:96:144–157. 10.1016/j.biochi.2013.07.004 [DOI] [PubMed] [Google Scholar]
- Pokotylo I, Kravets V, Martinec J, Ruelland E. The phosphatidic acid paradox: too many actions for one molecule class? Lessons from plants. Prog Lipid Res. 2018:71:43–53. 10.1016/j.plipres.2018.05.003 [DOI] [PubMed] [Google Scholar]
- Putta P, Rankenberg J, Korver RA, van Wijk R, Munnik T, Testerink C, Kooijman EE. Phosphatidic acid binding proteins display differential binding as a function of membrane curvature stress and chemical properties. Biochim Biophys Acta. 2016:1858(11):2709–2716. 10.1016/j.bbamem.2016.07.014 [DOI] [PubMed] [Google Scholar]
- Qi F, Li J, Ai Y, Shangguan K, Li P, Lin F, Liang Y. DGK5β-derived phosphatidic acid regulates ROS production in plant immunity by stabilizing NADPH oxidase. Cell Host Microbe. 2024:32(3):425–440. 10.1016/j.chom.2024.01.011 [DOI] [PubMed] [Google Scholar]
- Qian T, Li C, Liu F, Xu K, Wan C, Liu Y, Yu H. Arabidopsis synaptotagmin 1 mediates lipid transport in a lipid composition-dependent manner. Traffic. 2022:23(6):346–356. 10.1111/tra.12844 [DOI] [PubMed] [Google Scholar]
- Raho N, Ramirez L, Gonorazky AGM, Lanteri ML, Lamattina L, ten Have A, Laxalt AM. Phosphatidic acid production in chitosan-elicited tomato cells, via both phospholipase D and phospholipase C/diacylglycerol kinase, requires nitric oxide. J Plant Physiol. 2011:168(6):534–539. 10.1016/j.jplph.2010.09.004 [DOI] [PubMed] [Google Scholar]
- Ren H, Gao K, Liu Y, Sun D, Zheng S. The role of AtPLC3 and AtPLC9 in thermotolerance in Arabidopsis. Plant Signal Behav. 2017:12(10):e1162368. 10.1080/15592324.2016.1162368 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ried MK, Wild R, Zhu J, Pipercevic J, Sturm K, Broger L, Harmel RK, Abriata LA, Hothorn LA, Fiedler D, et al. Inositol pyrophosphates promote the interaction of SPX domains with the coiled-coil motif of PHR transcription factors to regulate plant phosphate homeostasis. Nat Commun. 2021:12(1):384. 10.1038/s41467-020-20681-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Riemer E, Pullagurla NJ, Yadav R, Rana P, Jessen HJ, Kamleitner M, Schaaf G, Laha D. Regulation of plant biotic interactions and abiotic stress responses by inositol polyphosphates. Front Plant Sci. 2022:13:944515. 10.3389/fpls.2022.944515 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Robuschi L, Mariani O, Perk EA, Cerrudo I, Villarreal F, Laxalt AM. Arabidopsis thaliana phosphoinositide-specific phospholipase C 2 is required for Botrytis cinerea proliferation. Plant Sci. 2024:340:111971. 10.1016/j.plantsci.2023.111971 [DOI] [PubMed] [Google Scholar]
- Rodriguez-Villalon A, Gujas B, van Wijk R, Munnik T, Hardtke CS. Primary root protophloem differentiation requires balanced phosphatidylinositol-4,5-bisphosphate levels and systemically affects root branching. Development. 2015:142(8):1437–1446. 10.1242/dev.118364 [DOI] [PubMed] [Google Scholar]
- Ruiz-Lopez N, Perez-Sancho J, Del Valle AE, Haslam RP, Vanneste S, Catala R, Perea-Resa C, Damme DV, Garcia-Hernandez S, Albert A, et al. Synaptotagmins at the endoplasmic reticulum-plasma membrane contact sites maintain diacylglycerol homeostasis during abiotic stress. Plant Cell. 2021:33(7):2431–2453. 10.1093/plcell/koab122 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rupwate SD, Rajasekharan R. C2 domain is responsible for targeting rice phosphoinositide specific phospholipase C. Plant Mol Biol. 2012:78(3):247–258. 10.1007/s11103-011-9862-1 [DOI] [PubMed] [Google Scholar]
- Sagar S, Biswas DK, Singh A. Genomic and expression analysis indicate the involvement of phospholipase C family in abiotic stress signaling in chickpea (Cicer arietinum). Gene. 2020:753:144797. 10.1016/j.gene.2020.144797 [DOI] [PubMed] [Google Scholar]
- Saile SC, Ackermann FM, Sunil S, Keicher J, Bayless A, Bonardi V, Wan L, Doumane M, Stobbe E, Jaillais Y, et al. Arabidopsis ADR1 helper NLR immune receptors localize and function at the plasma membrane in a phospholipid dependent manner. New Phytol. 2021:232(6):2440–2456. 10.1111/nph.17788 [DOI] [PubMed] [Google Scholar]
- Sanchez JP, Chua NH. Arabidopsis PLC1 is required for secondary responses to abscisic acid signals. Plant Cell. 2001:13(5):1143–1154. 10.1105/tpc.13.5.1143 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheard LB, Tan X, Mao H, Withers J, Ben-Nissan G, Hinds TR, Kobayashi Y, Hsu FF, Sharon M, Browse J, et al. Jasmonate perception by inositol-phosphate-potentiated COI1-JAZ co-receptor. Nature. 2010:468(7322):400–405. 10.1038/nature09430 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shen L, Zhuang B, Wu Q, Zhang H, Nie J, Jing W, Yang L, Zhang W. Phosphatidic acid promotes the activation and plasma membrane localization of MKK7 and MKK9 in response to salt stress. Plant Sci. 2019:287:110190. 10.1016/j.plantsci.2019.110190 [DOI] [PubMed] [Google Scholar]
- Simon ML, Platre MP, Assil S, van Wijk R, Chen WY, Chory J, Dreux M, Munnik T, Jaillais Y. A multi-colour/multi-affinity marker set to visualize phosphoinositide dynamics in Arabidopsis. Plant J. 2014:77(2):322–337. 10.1111/tpj.12358 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simon ML, Platre MP, Marques-Bueno MM, Armengot L, Stanislas T, Bayle V, Caillaud MC, Jaillais Y. A PtdIns(4)P-driven electrostatic field controls cell membrane identity and signaling in plants. Nat Plants. 2016:2(7):16089. 10.1038/nplants.2016.89 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singh A, Kanwar P, Pandey A, Tyagi AK, Sopory SK, Kapoor S, Pandey GK. Comprehensive genomic analysis and expression profiling of phospholipase C gene family during abiotic stresses and development in rice. PLoS One. 2013:8(4):e62494. 10.1371/journal.pone.0062494 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Song F, Goodman RM. Molecular cloning and characterization of a rice phosphoinositide-specific phospholipase C gene, OsPI-PLC1, that is activated in systemic acquired resistance. Physiol Mol Plant Pathol. 2002:61(1):31–40. 10.1006/pmpp.2002.0414 [DOI] [Google Scholar]
- Song L, Wang Y, Guo Z, Lam SM, Shui G, Cheng Y. NCP2/RHD4/SAC7, SAC6 and SAC8 phosphoinositide phosphatases are required for PtdIns4P and PtdIns(4,5)P2 homeostasis and Arabidopsis development. New Phytol. 2021:231(2):713–725. 10.1111/nph.17402 [DOI] [PubMed] [Google Scholar]
- Sousa E, Kost B, Malho R. Arabidopsis phosphatidylinositol-4-monophosphate 5-kinase 4 regulates pollen tube growth and polarity by modulating membrane recycling. Plant Cell. 2008:20(11):3050–3064. 10.1105/tpc.108.058826 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stanislas T, Huser A, Barbosa IC, Kiefer CS, Brackmann K, Pietra S, Gustavsson A, Zourelidou M, Schwechheimer C, Grebe M. Arabidopsis D6PK is a lipid domain-dependent mediator of root epidermal planar polarity. Nat Plants. 2015:1(11):15162. 10.1038/nplants.2015.162 [DOI] [PubMed] [Google Scholar]
- Stevenson JM, Perera IY, Heilmann I, Persson S, Boss WF. Inositol signaling and plant growth. Trends Plant Sci. 2000:5(6):252–258. 10.1016/S1360-1385(00)01652-6 [DOI] [PubMed] [Google Scholar]
- Stevenson-Paulik J, Bastidas RJ, Chiou ST, Frye RA, York JD. Generation of phytate-free seeds in Arabidopsis through disruption of inositol polyphosphate kinases. Proc Natl Acad Sci U S A. 2005:102(35):12612–12617. 10.1073/pnas.0504172102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sueldo DJ, Foresi NP, Casalongué CA, Lamattina L, Laxalt AM. Phosphatidic acid formation is required for extracellular ATP-mediated nitric oxide production in suspension-cultured tomato cells. New Phytol. 2010:185(4):909–916. 10.1111/j.1469-8137.2009.03165.x [DOI] [PubMed] [Google Scholar]
- Sui Z, Niu L, Yue G, Yang A, Zhang J. Cloning and expression analysis of some genes involved in the phosphoinositide and phospholipid signaling pathways from maize (Zea mays L.). Gene. 2008:426(1–2):47–56. 10.1016/j.gene.2008.09.004 [DOI] [PubMed] [Google Scholar]
- Sun J, Liu X, Pan Y. The physical interaction between LdPLCs and Arabidopsis G beta in a yeast two-hybrid system. Front Agric China. 2011:5(1):64–71. 10.1007/s11703-011-1063-9 [DOI] [Google Scholar]
- Sun Y, Li Y, Sun X, Wu Q, Yang C, Wang L. Overexpression of a phosphatidylinositol-specific phospholipase C gene from populus simonii × P. nigra improves salt tolerance in transgenic tobacco. J Plant Biol. 2022;65(5):365–376. 10.1007/s12374-022-09359-0 [DOI] [Google Scholar]
- Synek L, Pleskot R, Sekeres J, Serrano N, Vukasinovic N, Ortmannova J, Klejchova M, Pejchar P, Batystova K, Gutkowska M, et al. Plasma membrane phospholipid signature recruits the plant exocyst complex via the EXO70A1 subunit. Proc Natl Acad Sci U S A. 2021:118(36):e2105287118. 10.1073/pnas.2105287118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tan X, Calderon-Villalobos LI, Sharon M, Zheng C, Robinson CV, Estelle M, Zheng N. Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature. 2007:446(7136):640–645. 10.1038/nature05731 [DOI] [PubMed] [Google Scholar]
- Tasma IM, Brendel V, Whitham SA, Bhattacharyya MK. Expression and evolution of the phosphoinositide-specific phospholipase C gene family in Arabidopsis thaliana. Plant Physiol Biochem. 2008:46(7):627–637. 10.1016/j.plaphy.2008.04.015 [DOI] [PubMed] [Google Scholar]
- Tejos R, Sauer M, Vanneste S, Palacios-Gomez M, Li H, Heilmann M, van Wijk R, Vermeer JE, Heilmann I, Munnik T, et al. Bipolar plasma membrane distribution of phosphoinositides and their requirement for auxin-mediated cell polarity and patterning in Arabidopsis. Plant Cell. 2014:26(5):2114–2128. 10.1105/tpc.114.126185 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Testerink C, Dekker HL, Lim ZY, Johns MK, Holmes AB, Koster CG, Ktistakis NT, Munnik T. Isolation and identification of phosphatidic acid targets from plants. Plant J. 2004:39(4):527–536. 10.1111/j.1365-313X.2004.02152.x [DOI] [PubMed] [Google Scholar]
- Testerink C, Munnik T. Molecular, cellular, and physiological responses to phosphatidic acid formation in plants. J Exp Bot. 2011:62(7):2349–2361. 10.1093/jxb/err079 [DOI] [PubMed] [Google Scholar]
- Thomas C, Staiger CJ. A dynamic interplay between membranes and the cytoskeleton critical for cell development and signaling. Front Plant Sci. 2014:5:335. 10.3389/fpls.2014.00335 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tripathy MK, Tyagi W, Goswami M, Kaul T, Singla-Pareek SL, Deswal R, Reddy MK, Sopory SK. Characterization and functional validation of tobacco PLC delta for abiotic stress tolerance. Plant Mol Biol Rep. 2012:30(2):488–497. 10.1007/s11105-011-0360-z [DOI] [Google Scholar]
- Ufer G, Gertzmann A, Gasulla F, Röhrig H, Bartels D. Identification and characterization of the phosphatidic acid-binding A. thaliana phosphoprotein PLD rp1 that is regulated by PLD α1 in a stress-dependent manner. Plant J. 2017:92(2):276–290. 10.1111/tpj.13651 [DOI] [PubMed] [Google Scholar]
- van der Luit AH, Piatti T, van Doorn A, Musgrave A, Felix G, Boller T, Munnik T. Elicitation of suspension-cultured tomato cells triggers the formation of phosphatidic acid and diacylglycerol pyrophosphate. Plant Physiol. 2000:123(4):1507–1516. 10.1104/pp.123.4.1507 [DOI] [PMC free article] [PubMed] [Google Scholar]
- van Hooren M. The role of phospholipase C in plant drought tolerance. PhD Thesis, University of Amsterdam; 2023. p. 1–209 [Google Scholar]
- van Hooren M, Darwish E, Munnik T. Stress- and phospholipid signalling responses in Arabidopsis PLC4-KO and -overexpression lines under salt- and osmotic stress. Phytochemistry. 2023:216:113862. 10.1016/j.phytochem.2023.113862 [DOI] [PubMed] [Google Scholar]
- van Hooren M, van Wijk R, Vaseva II, Van Der Straeten D, Haring M, Munnik T. Ectopic expression of distinct PLC genes identifies ‘compactness’ as a possible architectural shoot strategy to cope with drought stress. Plant Cell Physiol. 2024:65(6):885–903. 10.1093/pcp/pcad123 [DOI] [PMC free article] [PubMed] [Google Scholar]
- van Leeuwen W, Vermeer JE, Gadella TW, Jr., Munnik T. Visualization of phosphatidylinositol 4,5-bisphosphate in the plasma membrane of suspension-cultured tobacco BY-2 cells and whole Arabidopsis seedlings. Plant J. 2007:52(6):1014–1026. 10.1111/j.1365-313X.2007.03292.x [DOI] [PubMed] [Google Scholar]
- van Schooten B, Testerink C, Munnik T. Signaling diacylglycerol pyrophosphate, a new phosphatidic acid metabolite. Biochim Biophys Acta. 2006:1761(2):151–159. 10.1016/j.bbalip.2005.12.010 [DOI] [PubMed] [Google Scholar]
- van Wijk R, Zhang Q, Zarza X, Lamers M, Marquez FR, Guardia A, Scuffi D, Garcia-Mata C, Ligterink W, Haring MA, et al. Role for Arabidopsis PLC7 in stomatal movement, seed mucilage attachment, and leaf serration. Front Plant Sci. 2018:9:1721. 10.3389/fpls.2018.01721 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vermeer JE, Thole JM, Goedhart J, Nielsen E, Munnik T, Gadella TW, Jr. Imaging phosphatidylinositol 4-phosphate dynamics in living plant cells. Plant J. 2009:57(2):356–372. 10.1111/j.1365-313X.2008.03679.x [DOI] [PubMed] [Google Scholar]
- Vermeer JE, van Leeuwen W, Tobena-Santamaria R, Laxalt AM, Jones DR, Divecha N, Gadella TW, Jr., Munnik T. Visualization of PtdIns3P dynamics in living plant cells. Plant J. 2006:47(5):687–700. 10.1111/j.1365-313X.2006.02830.x [DOI] [PubMed] [Google Scholar]
- Vermeer JEM, Munnik T. Using genetically encoded fluorescent reporters to image lipid signaling in living plants. Methods Mol Biol. 2013:1009:283–290. 10.1007/978-1-62703-401-2_26 [DOI] [PubMed] [Google Scholar]
- Vermeer JEM, van Wijk R, Goedhart J, Geldner N, Chory J, Gadella TWJ, Jr., Munnik T. In vivo imaging of diacylglycerol at the cytoplasmic leaflet of plant membranes. Plant Cell Physiol. 2017:58(7):1196–1207. 10.1093/pcp/pcx012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Verslues PE, Bailey-Serres J, Brodersen C, Buckley TN, Conti L, Christmann A, Dinneny JR, Grill E, Hayes S, Heckman RW, et al. Burning questions for a warming and changing world: 15 unknowns in plant abiotic stress. Plant Cell. 2023:35(1):67–108. 10.1093/plcell/koac263 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vossen JH, Abd-El-Haliem A, Fradin EF, van den Berg GC, Ekengren SK, Meijer HJ, Seifi A, Bai Y, ten Have A, Munnik T, et al. Identification of tomato phosphatidylinositol-specific phospholipase-C (PI-PLC) family members and the role of PLC4 and PLC6 in HR and disease resistance. Plant J. 2010:62(2):224–239. 10.1111/j.1365-313X.2010.04136.x [DOI] [PubMed] [Google Scholar]
- Wang C-R, Yang A-F, Yue G-D, Gao Q, Yin H-Y, Zhang J-R. Enhanced expression of phospholipase C 1 (ZmPLC1) improves drought tolerance in transgenic maize. Planta. 2008:227(5):1127–1140. 10.1007/s00425-007-0686-9 [DOI] [PubMed] [Google Scholar]
- Wang F, Deng Y, Zhou Y, Dong J, Chen H, Dong Y, Wang N, Li X, Li H. Genome-wide analysis and expression profiling of the phospholipase C gene family in soybean (Glycine max). PLoS One. 2015:10(9):e0138467. 10.1371/journal.pone.0138467 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang K, Li Y-l, Chen S. Genome-wide identification of phospholipase C related to chilling injury in peach fruit. J Plant Biochem Biotechnol. 2021:30(3):452–461. 10.1007/s13562-020-00632-2 [DOI] [Google Scholar]
- Wang P, Shen L, Guo J, Jing W, Qu Y, Li W, Bi R, Xuan W, Zhang Q, Zhang W. Phosphatidic acid directly regulates PINOID-dependent phosphorylation and activation of the PIN-FORMED2 auxin efflux transporter in response to salt stress. Plant Cell. 2019:31(1):250–271. 10.1105/tpc.18.00528 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Q, Aliaga Fandino AC, Graeff M, DeFalco TA, Zipfel C, Hardtke CS. A phosphoinositide hub connects CLE peptide signaling and polar auxin efflux regulation. Nat Commun. 2023:14(1):423. 10.1038/s41467-023-36200-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X, Devaiah SP, Zhang W, Welti R. Signaling functions of phosphatidic acid. Prog Lipid Res. 2006:45(3):250–278. 10.1016/j.plipres.2006.01.005 [DOI] [PubMed] [Google Scholar]
- Wang X, Liu Y, Li Z, Gao X, Dong J, Yang M. Expression and evolution of the phospholipase C gene family in Brachypodium distachyon. Genes Genomics. 2020a:42(9):1041–1053. 10.1007/s13258-020-00973-1 [DOI] [PubMed] [Google Scholar]
- Wang X, Liu Y, Li Z, Gao X, Dong J, Zhang J, Zhang L, Thomashow LS, Weller DM, Yang M. Genome-wide identification and expression profile analysis of the phospholipase C gene family in wheat (Triticum aestivum L.). Plants. 2020b:9(7):885. 10.3390/plants9070885 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wild R, Gerasimaite R, Jung J-Y, Truffault V, Pavlovic I, Schmidt A, Saiardi A, Jessen HJ, Poirier Y, Hothorn M, et al. Control of eukaryotic phosphate homeostasis by inositol polyphosphate sensor domains. Science. 2016:352(6288):986–990. 10.1126/science.aad9858 [DOI] [PubMed] [Google Scholar]
- Williams SP, Gillaspy GE, Perera IY. Biosynthesis and possible functions of inositol pyrophosphates in plants. Front Plant Sci. 2015:6:67. 10.3389/fpls.2015.00067 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu C, Tan L, van Hooren M, Tan X, Liu F, Li Y, Zhao Y, Li B, Rui Q, Munnik T, et al. Arabidopsis EXO70A1 recruits patellin3 to the cell membrane independent of its role as an exocyst subunit. J Integr Plant Biol. 2017:59(12):851–865. 10.1111/jipb.12578 [DOI] [PubMed] [Google Scholar]
- Wu P, Gao HB, Zhang LL, Xue HW, Lin WH. Phosphatidic acid regulates BZR1 activity and brassinosteroid signal of Arabidopsis. Mol Plant. 2014:7(2):445–447. 10.1093/mp/sst138 [DOI] [PubMed] [Google Scholar]
- Xia K, Wang B, Zhang J, Li Y, Yang H, Ren D. Arabidopsis phosphoinositide-specific phospholipase C 4 negatively regulates seedling salt tolerance. Plant Cell Environ. 2017:40(8):1317–1331. 10.1111/pce.12918 [DOI] [PubMed] [Google Scholar]
- Xing J, Zhang L, Duan Z, Lin J. Coordination of phospholipid-based signaling and membrane trafficking in plant immunity. Trends Plant Sci. 2021:26(4):407–420. 10.1016/j.tplants.2020.11.010 [DOI] [PubMed] [Google Scholar]
- Yao HY, Xue HW. Phosphatidic acid plays key roles regulating plant development and stress responses. J Integr Plant Biol. 2018:60(9):851–863. 10.1111/jipb.12655 [DOI] [PubMed] [Google Scholar]
- Yao S, Kim SC, Li J, Tang S, Wang X. Phosphatidic acid signaling and function in nuclei. Prog Lipid Res. 2024:93:101267. 10.1016/j.plipres.2023.101267 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yoo CM, Quan L, Cannon AE, Wen J, Blancaflor EB. AGD1, a class 1 ARF-GAP, acts in common signaling pathways with phosphoinositide metabolism and the actin cytoskeleton in controlling Arabidopsis root hair polarity. Plant J. 2012:69(6):1064–1076. 10.1111/j.1365-313X.2011.04856.x [DOI] [PubMed] [Google Scholar]
- Yperman K, Wang J, Eeckhout D, Winkler J, Vu LD, Vandorpe M, Grones P, Mylle E, Kraus M, Merceron R, et al. Molecular architecture of the endocytic TPLATE complex. Sci Adv. 2021:7(9):eabe7999. 10.1126/sciadv.abe7999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu L, Nie J, Cao C, Jin Y, Yan M, Wang F, Liu J, Xiao Y, Liang Y, Zhang W. Phosphatidic acid mediates salt stress response by regulation of MPK6 in Arabidopsis thaliana. New Phytol. 2010:188(3):762–773. 10.1111/j.1469-8137.2010.03422.x [DOI] [PubMed] [Google Scholar]
- Yu M, Cao C, Yin X, Liu X, Yang D, Gong C, Wang H, Wu Y. The rice phosphoinositide-specific phospholipase C3 is involved in responses to osmotic stresses via modulating ROS homeostasis. Plant Sci. 2021:313:111087. 10.1016/j.plantsci.2021.111087 [DOI] [PubMed] [Google Scholar]
- Yu M, Huang D, Yin X, Liu X, Yang D, Gong C, Wang H, Wu Y. The phosphoinositide-specific phospholipase C1 modulates flowering time and grain size in rice. Planta. 2022:256(2):29. 10.1007/s00425-022-03941-z [DOI] [PubMed] [Google Scholar]
- Yuan P, Jauregui E, Du L, Tanaka K, Poovaiah BW. Calcium signatures and signaling events orchestrate plant–microbe interactions. Curr Opin Plant Biol. 2017:38:173–183. 10.1016/j.pbi.2017.06.003 [DOI] [PubMed] [Google Scholar]
- Zarza X, Van Wijk R, Shabala L, Hunkeler A, Lefebvre M, Rodriguez-Villalon A, Shabala S, Tiburcio AF, Heilmann I, Munnik T. Lipid kinases PIP5K7 and PIP5K9 are required for polyamine-triggered K+ efflux in Arabidopsis roots. Plant J. 2020:104(2):416–432. 10.1111/tpj.14932 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang B, Wang Y, Liu JY. Genome-wide identification and characterization of phospholipase C gene family in cotton (Gossypium spp.). Sci China Life Sci. 2018a:61(1):88–99. 10.1007/s11427-017-9053-y [DOI] [PubMed] [Google Scholar]
- Zhang J, Zhang Z, Zhu D, Guan Y, Shi D, Chen Y, Li R, Wang H, Wei J. Expression and initial characterization of a phosphoinositide-specific phospholipase C from Populus tomentosa. J Plant Biochem Biotechnol. 2015:24(3):338–346. 10.1007/s13562-014-0279-1 [DOI] [Google Scholar]
- Zhang K, Jin C, Wu L, Hou M, Dou S, Pan Y. Expression analysis of a stress-related phosphoinositide-specific phospholipase C gene in wheat (Triticum aestivum L.). PLoS One. 2014:9(8):e105061. 10.1371/journal.pone.0105061 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Q, van Wijk R, Shahbaz M, Roels W, Schooten BV, Vermeer JEM, Zarza X, Guardia A, Scuffi D, Garcia-Mata C, et al. Arabidopsis phospholipase C3 is involved in lateral root initiation and ABA responses in seed germination and stomatal closure. Plant Cell Physiol. 2018b:59(3):469–486. 10.1093/pcp/pcx194 [DOI] [PubMed] [Google Scholar]
- Zhang Q, van Wijk R, Zarza X, Shahbaz M, van Hooren M, Guardia A, Scuffi D, Garcia-Mata C, Van den Ende W, Hoffmann-Benning S, et al. Knock-down of Arabidopsis PLC5 reduces primary root growth and secondary root formation while overexpression improves drought tolerance and causes stunted root hair growth. Plant Cell Physiol. 2018c:59(10):2004–2019. 10.1093/pcp/pcy120 [DOI] [PubMed] [Google Scholar]
- Zhang W, Qin C, Zhao J, Wang X. Phospholipase D alpha 1-derived phosphatidic acid interacts with ABI1 phosphatase 2C and regulates abscisic acid signaling. Proc Natl Acad Sci U S A. 2004:101(25):9508–9513. 10.1073/pnas.0402112101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y, Zhu H, Zhang Q, Li M, Yan M, Wang R, Wang L, Welti R, Zhang W, Wang X. Phospholipase Dα1 and phosphatidic acid regulate NADPH oxidase activity and production of reactive oxygen species in ABA-mediated stomatal closure in Arabidopsis. Plant Cell. 2009:21(8):2357–2377. 10.1105/tpc.108.062992 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao Y, Yan A, Feijo JA, Furutani M, Takenawa T, Hwang I, Fu Y, Yang Z. Phosphoinositides regulate clathrin-dependent endocytosis at the tip of pollen tubes in Arabidopsis and tobacco. Plant Cell. 2010:22(12):4031–4044. 10.1105/tpc.110.076760 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng SZ, Liu YL, Li B, Shang ZL, Zhou RG, Sun DY. Phosphoinositide-specific phospholipase C9 is involved in the thermotolerance of Arabidopsis. Plant J. 2012:69(4):689–700. 10.1111/j.1365-313X.2011.04823.x [DOI] [PubMed] [Google Scholar]
- Zhu J, Lau K, Puschmann R, Harmel RK, Zhang Y, Pries V, Gaugler P, Broger L, Dutta AK, Jessen HJ, et al. Two bifunctional inositol pyrophosphate kinases/phosphatases control plant phosphate homeostasis. Elife. 2019:8:e43582. 10.7554/eLife.43582 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu J, Zhou Y, Li J, Li H. Genome-wide investigation of the phospholipase C gene family in Zea mays. Front Genet. 2020:11:611414. 10.3389/fgene.2020.611414 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zonia L, Cordeiro S, Tupy J, Feijo JA. Oscillatory chloride efflux at the pollen tube apex has a role in growth and cell volume regulation and is targeted by inositol 3,4,5,6-tetrakisphosphate. Plant Cell. 2002:14(9):2233–2249. 10.1105/tpc.003830 [DOI] [PubMed] [Google Scholar]
- Zonia L, Munnik T. Cracking the green paradigm: functional coding of phosphoinositide signals in plant stress responses. Subcell Biochem. 2006:39:207–237. 10.1007/0-387-27600-9_9 [DOI] [PubMed] [Google Scholar]
- Zonia L, Munnik T. Understanding pollen tube growth: the hydrodynamic model versus the cell wall model. Trends Plant Sci. 2011:16(7):347–352. 10.1016/j.tplants.2011.03.009 [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The data underlying this article are available in the article and in its online supplementary material.



