uva-dare (digital academic repository) novel roles …...2.3 gene expression and subcellular...

137
UvA-DARE is a service provided by the library of the University of Amsterdam (http://dare.uva.nl) UvA-DARE (Digital Academic Repository) Novel roles for phospholipase C in plant stress signalling and development Zhang, Q. Link to publication Creative Commons License (see https://creativecommons.org/use-remix/cc-licenses): Other Citation for published version (APA): Zhang, Q. (2017). Novel roles for phospholipase C in plant stress signalling and development. General rights It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open content license (like Creative Commons). Disclaimer/Complaints regulations If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You will be contacted as soon as possible. Download date: 19 Sep 2020

Upload: others

Post on 22-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

UvA-DARE is a service provided by the library of the University of Amsterdam (http://dare.uva.nl)

UvA-DARE (Digital Academic Repository)

Novel roles for phospholipase C in plant stress signalling and development

Zhang, Q.

Link to publication

Creative Commons License (see https://creativecommons.org/use-remix/cc-licenses):Other

Citation for published version (APA):Zhang, Q. (2017). Novel roles for phospholipase C in plant stress signalling and development.

General rightsIt is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s),other than for strictly personal, individual use, unless the work is under an open content license (like Creative Commons).

Disclaimer/Complaints regulationsIf you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, statingyour reasons. In case of a legitimate complaint, the Library will make the material inaccessible and/or remove it from the website. Please Askthe Library: https://uba.uva.nl/en/contact, or a letter to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam,The Netherlands. You will be contacted as soon as possible.

Download date: 19 Sep 2020

Page 2: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Novel roles for phospholipase Cin plant stress signalling and development

Qianqian Zhang

No

vel roles fo

r ph

osp

ho

lipase C

in p

lant stress sig

nallin

g an

d d

evelop

men

t Qian

qian

Zh

ang

20

17

Page 3: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Novel roles for phospholipase C in plant stress signalling and development

ACADEMISCH PROEFSCHRIFT

ter verkrijging van de graad van doctor

aan de Universiteit van Amsterdam

op gezag van de Rector Magnificus

prof. dr. ir. K.I.J. Maex

ten overstaan van een door het College voor Promoties ingestelde commissie,

in het openbaar te verdedigen in de Agnietenkapel

op donderdag 1 juni 2017, te 10:00 uur

door

Qianqian Zhang

geboren te Hebei, China

Page 4: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Promotiecommissie:

Promotor: Prof. dr. M.A. Haring, Universiteit van Amsterdam

Copromotor: Dr. T. Munnik, Universiteit van Amsterdam

Overige leden: Prof. dr. C.S. Testerink, Universiteit van Amsterdam

Prof. dr. B.J.C. Cornelissen, Universiteit van Amsterdam

Prof. dr. H.J. Bouwmeester, Universiteit van Amsterdam

Prof. dr. I. Heilmann, Martin Luther University, Halle-Wittenberg, Germany

Dr. L. Bentsink, Wageningen UR

Faculteit der Natuurwetenschappen, Wiskunde en Informatica

The research described in this thesis was carried out in the Plant Physiology lab of the

Swammerdam Institute for Life Sciences of the Faculty of Sciences, University of Amsterdam.

This work was funded in part by a scholarship from the China Scholarship Council to

Qiainqian Zhang.

Page 5: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Table of Content

Chapter 1 5

General introduction

Chapter 2 23

A role for Arabidopsis phospholipase C3 (PLC3) in seed germination, lateral root formation and stomatal closure

Chapter 3 61

Functional characterization of PLC5 in Arabidopsis thaliana - knock-down affects lateral root initiation while overexpression stunts root hair growth and enhances drought tolerance

Chapter 4 91

Role for Arabidopsis PLC7 in stomatal closure, mucilage adherence, and leaf serration

Chapter 5 119

General discussion

Summary 129

Samenvatting 131

Acknowledgements 133

List of publications 135

Page 6: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses
Page 7: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

5

Chapter 1 General Introduction Qianqian Zhang,1,2 Michel A. Haring1 & Teun Munnik1,2

1Section Plant Physiology, 2Section Plant Cell Biology, Swammerdam Institute for Life Sciences, University of Amsterdam, Science Park 904, Amsterdam, 1098 XH, The Netherlands

Page 8: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

6

Contents

1. PLC signaling in plants

2. PLC enzymes

2.1 PLC family

2.2 Enzymatic activity

2.3 Gene expression and subcellular localization

3. PLC in plant stress response and development

3.1 PLC in plant stress responses

3.1.1 Osmotic-,heat- and cold stress

3.1.2 Abscisic acid (ABA)

3.1.3 Biotic stress

3.2 PLC in plant development

4. Scope of this thesis: dissection of the role of Arabidopsis PLC3, PLC5 and PLC7 in plant

signaling

4.1 Characteristics of the Arabidopsis PLC3, PLC5 and PLC7 genes

4.2 Techniques to study PLC signaling in plant

5. References

Page 9: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

7

1. PLC signaling in plants

In nature, plants live in an open environment and face various challenges, including heat- and cold

stress, drought, salinity, wounding by herbivores, or infection by pathogens. Since plants cannot run

away, they must quickly recognize and respond to outside signals and adjust themselves to the

challenging surroundings. This process, called 'signal transduction', starts at the cellular level. Signal

perception is through proteins or other molecules located in the plasma membrane, which is a selective

barrier between the interior cell and the external environment (Fig. 1a). The plasma membrane is

comprised of a lipid bilayer with associated integral and peripheral membrane proteins. The lipid part

consists of phospholipids (~40-60%), sphingolipids (~10-20%) and sterols (~40-50%) (Furt et al.,

2011). The bilayered structure results from the biophysical properties of the phospholipids, which are

composed of a hydrophilic, phosphate-containing head group and, in general, two hydrophobic

glycerol-fatty acyl tails (Fig. 1b). The fatty acids tend to face each other inwards while the hydrophilic

head groups interact with the aqueous surroundings on the outside (Fig. 1a). The most abundant

phospholipids in the plasma membrane are phosphatidylethanolamine (PE) and phosphatidylcholine

(PC), which together account for ~70-80%, followed by phosphatidylglycerol (PG) and

phosphatidylinositol (PI) (~5-10% each). Together, they are called structural phospholipids, because

they make-up the mass- and structure of membranes, which is not only the plasma membrane, but

includes all intracellular organelles, i.e endoplasmic reticulum (ER), Golgi, vacuole, mitochondria,

peroxisomes and plastids. In each membrane, phospholipids play a key role in the dynamics and

maintenance of the fluid, bilayer structure, with the desaturation (double bonds) and saturation (no

double bonds) of the fatty acids determining a great deal of the fluidity of membranes (Van Hooren and

Munnik, 2017).

In addition to the structural phospholipoids, there are also some minor phospholipids, which are

present in relatively low quantities and have signaling rather that structural role. These include,

phosphatidic acid (PA), phosphatidylinositol 4-phosphate (PIP) and phosphatidylinositol 4,5-

bisphosphate (PIP2), which play important roles in vesicular trafficking (endo- and exocytosis) and

signal transduction. Although the study of phospholipid signaling in plants is relatively young, their

importance in plant stress and development is emerging and holds great potential for future research

(Munnik & Nielsen, 2011; Testerink & Munnik, 2011; Gillaspy, 2013; Gujas & Rodriguez-Villalon,

2016; Heilmann, 2016).

Phospholipid signaling systems would not function without various phospholipases, i.e.:

phospholipase A1 (PLA1), PLA2, phospholipase C (PLC) and phospholipase D (PLD), which attack

different positions of a phospholipid (Fig. 1b; Hong et al., 2016). In this thesis, research is focused on

PLC that is involved in signaling. It specifically hydrolyzes phosphoinositides and is referred to PI-PLC.

Plants also contain non-specific PLCs, called NPCs, which act on structural lipids like PC and PE, and

mainly function in membrane dynamics.

Page 10: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

8

Figure 1. Schematic representation of the plasma membrane consisting of phospholipids and embedded proteins. (a) The plasma membrane consists of two phospholipid layers with their hydrophobic part facing each other and the hydrophilic heads facing outward. (b) Phospholipid has a three-carbon glycerol backbone containing two fatty acid chains and a phosphate-containing head group. The phospholipid X group can vary, as listed in the blue box. The cleavage sites of phospholipases are indicated by arrows.

Page 11: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

9

PLC cleaves PIP2 into inositol 1,4,5- triphosphate (IP3) and diacylglycerol (DAG), which, in

mammalian systems, are crucial second messengers, regulating various cellular processes. IP3 is water

soluble and diffuses into cytosol to bind the IP3 receptor, which is a ligand-gated Ca2+ channel at the ER

that releases Ca2+ into the cytosol, leading to a transient increase in cytosolic Ca2+ that triggers a

cascade of downstream changes. The lipid moiety of PIP2, i.e. DAG, remains in the plasma membrane

where it recruits and activates members of the protein kinase C (PKC) family that phosphorylate

various downstream protein targets, altering numerous biochemical pathways. Alternatively, DAG can

activate so-called TRP (Transient receptor potential) channels, which triggers Ca2+ influx across the

plasma membrane.

Figure 2. Comparison between mammalian and plant PLC signaling pathways. In mammalian cell, PLC is regulated by a G protein after receiving the signal. PLC cleaves PIP2 into IP3 and DAG. IP3 releases Ca2+ via Ca2+ channel and DAG activates PKC, leading to activation of downstream signaling cascades. In higher plants, G protein, targets for IP3 and DAG are missing, indicated by the red crosses. Instead, plant seems to use IP6 and PA as second messengers. The picture is modified from Munnik et al. (2009)

While the above is typical for the mammalian PLC signaling pathway, a number of

differences have been found for the plant PLC system in terms of second messengers, regulation and

targets (Fig. 2). Perhaps the most striking difference is that plants are lacking homologs of the IP3

receptor, PKC or TRP-channel (Munnik, 2014). Earlier, PLC and IP3 had been coupled to Ca2+

signaling in ABA-induced stomatal closure (Blatt et al., 1990; Gilroy et al., 1990; Allen and Sanders,

1994; Lee et al., 1996; Staxen et al., 1999), but later this was shown to result from its phosphorylation

into IP6, with the latter being responsible for the release of the intracellular Ca2+ (Lemtiri-Chlieh et al.,

2000, 2003; Munnik and Vermeer, 2010). To produce IP6 from IP3, two inositolpolyphosphate kinases,

IPK2 and IPK1, are required (Munnik and Vermeer, 2010; Gillaspy, 2013). Apart from releasing Ca2+,

IP6 has also been implicated in auxin signaling where it binds the auxin receptor, TIR1 (Tan et al.,

2007), mRNA export (Lee et al., 2015), phosphate homeostasis (Kuo et al., 2014; Puga et al., 2014;

Wild et al., 2016) and disease resistance (Murphy et al., 2008). Meanwhile, inositolpolyphosphates

Page 12: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

10

(IPPs) other than IP6 have been implicated in signaling. For example, IP4 has been shown to regulate a

chloride channel (Zonia et al., 2002), while IP5 was discovered in the crystal structure of COI1, which

is the receptor for jasmonate signaling, (Sheard et al., 2010). Recently, another IPK, i.e. VIH2, has been

identified, which is responsible for the production of pyrophosphorylated IPPs, i.e. IP7 and IP8. The

latter are emerging as important signaling molecules in plants, as well as in animals and fungi (York,

2006; Michell, 2008; Burton et al., 2009; Shears, 2009; Desai et al., 2014; Laha et al., 2015, 2016;

Thota and Bhandari, 2015). Importantly, cellular roles for IP6 should not be confused with its role in

plants as phosphate-storage molecule in seeds (phytate), like Ca2+ signaling should not be confused with

its abundant presence in bones.

Likewise, not DAG but its phosphorylated product, phosphatidic acid (PA) is emerging as the

plant lipid second messenger (Munnik, 2001, 2014; Testerink & Munnik, 2005; Wang et al., 2006;

Arisz et al., 2009; Pokotylo et al., 2014). The latter requires the help of DAG kinase (DGK). The

metabolism of PA is complex though, as it is also an intermediate in glycerolipid biosynthesis and is

directly produced via PLD hydrolysis of PC, PE and PG (Arisz et al., 2009; Arisz and Munnik, 2013).

2. PLC enzymes

2.1 PLC family

Eukaryotic cells contain multiple PLC isoforms and they fall into six subfamilies (β, γ, δ, ε, η and ζ)

based on sequence similarity and domain structure (Fig. 3; Munnik and Testerink, 2009).

Figure 3. Domain structure and organization of PLC isozymes. In eukaryotes there are six PLC isoforms (β, γ, δ, ε, η and ζ). Plant PLCs resemble PLCζ, which is the simplest PLC that has a minimal core structure: EF- hand, X-Y-domain and C2 domain. All other PLC isoforms contain a Pleckstrin homology (PH) domain and some of them have additional domains, such as a PDZ-binding motif in PLC β and η; a Ras GEF guanine nucleotide-exchange factor domain and an RA domain in PLC ε; an SH2/SH3 domain in PLC γ.

Mammalian cells contain all six subfamilies, resulting in 13 PLC isoforms. Most subfamilies share the

conserved structure of an N-terminal pleckstrin homology (PH) domain, Ca2+ binding EF-hands,

catalytic X-and Y domains, and a lipid binding-C2 domain in the C-terminal, except for PLCζ, which

lacks the PH domain that is known to bind phosphoinositides and certain proteins, which differs per PH

domain (Rebecchi and Pentyala, 2000; Bunney and Katan, 2011). The PLC family has also been studied

in various plant species, including Arabidopsis (Hunt et al., 2004; Tasma et al., 2008; Xia et al., 2017),

Page 13: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

11

tomato (Vossen et al., 2010), rice (Singh et al., 2013; Li et al., 2017), soybean (Wang et al., 2015) and

maize (Wang et al., 2008). The plant PLC protein structure is similar to the mammalian PLCδ- and

PLCζ subfamilies, lacking a PH domain and one of the conserved EF-hand lobes (Munnik, 2014;

Pokotylo et al., 2014; Hong et al., 2016) and containing a plant-specific N-terminal region. Arabidopsis

contains 9 PLC genes, which can be divided into 4 different subfamilies, based on phylogenetic

analysis of the encoded proteins (Fig. 4; Hunt et al., 2004). Among them, PLC8 and PLC9 are the most

divergent as they contain mutations in the Y domain that would render them inactive as enzymes

(Mueller-Roeber and Pical, 2002; Hunt et al., 2004; Tasma et al., 2008). Nonetheless, PLC9 has been

implicated in heat-stress tolerance (Zheng et al., 2012).

Figure 4. Phylogenetic tree of the Arabidopsis PLC protein family together with mammalian δ and ζ isozymes. Bootstrap values for nodes that had > 50% support in a bootstrap analysis of 1000 replicates are shown. Data show is from Hunt. et al. (2004). 2.2 Enzymatic activity

PLC catalyzes the hydrolysis of phosphoinositides by attacking the phosphodiester bond at the glycerol

side to generate a phosphorylated-inositol headgroup and DAG. In vitro, the enzyme activity requires

Ca2+ and the substrate preference depends on the Ca2+ concentration. For example, at low, physiological

Ca2+ concentrations (100 nM-10 µΜ), PLC hydrolyses PI4P and PI(4,5)P2 equally well. At higher, non-

physiological Ca2+ concentrations (mM level), PLC also hydrolyses PI. However, PLCs cannot

hydrolyze phosphoinositides that are phosphorylated at the D3-position of the inositol ring (i.e. PI3P,

PI(3,4)P2, and PI(3,5)P2) nor other phospholipids (Munnik et al., 1998). Based on crude protein extracts

from different species and tissues, both membrane-associated and soluble PLCs have been described

(Munnik et al., 1998). Membrane-associated PLCs preferred PI4P and PI(4,5)P2 as substrates requiring

Page 14: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

12

low µM Ca2+, whereas soluble PLCs preferred PI at higher (mM) Ca2+. The pH optimum for both PLC

forms was between 6 and 7 (Munnik et al., 1998).

Mammalian PLCs are known to mainly use PI(4,5)P2 as in vivo substrate, which is facilitated

by the fact that mammalian cells have relatively high concentrations of PI(4,5)P2 in their plasma

membranes, and because the PLCs are equipped with PIP2-binding domains, which is either the PH

domain in PLCβ, -γ, -δ, -ε, -η or the XY linker in PLCζ (Swann and Lai, 2016). For plants, the in vivo

substrate remains unclear. The PI(4,5)P2 levels in plants are 30-100 times lower and the PLCs lack a PH

domain or the XY-linker sequence from PLCζ. Since PI4P can be hydrolyzed equally well in vitro and

plants contain relatively high amounts of it in the plasma membrane in vivo (Munnik et al., 1998;

Vermeer et al., 2009), the latter could be the common substrate for plant PLCs.

2.3 Gene expression and (sub)cellular localization

PLC genes from different plant species (Arabidopsis, rice, maize, tomato, potato, pea, mung been, etc.)

are expressed in various tissues (leave, stem, flower, root) during development, and levels are

influenced by biotic- and abiotic stresses (Hirayama et al., 1995; Hunt et al., 2004; Kim et al., 2004;

Lin et al., 2004; Das et al., 2005; Vergnolle et al., 2005; Liu, Liu, et al., 2006; Tasma et al., 2008; Song

et al., 2008; Sui et al., 2008; Vossen et al., 2010; Zheng et al., 2012; Singh et al., 2013; Wang et al.,

2015; Li et al., 2017). Several putative regulatory elements have been identified in the promoter regions

of the Arabidopsis PLCs, confirming their potential involvement in developmental and stress responses

(Tasma et al., 2008; Hsieh et al., 2013; Singh et al., 2013).

Based on subcellular-fractionation and activity assays, but also by imaging of fluorescent-

protein fusions, PLCs have been localized to the plasma membrane in various species, including

Arabidopsis PLC2 (Otterhag et al., 2001) and PLC9 (Zheng et al., 2012), Petunia PLC1 (Dowd et al.,

2006), tobacco PLC3 (Helling et al., 2006) and mung bean PLC3 (Kim et al., 2004). PLCs have also

been found in the cytoplasm (Shi et al., 1995; Rupwate and Rajasekharan, 2012; Singh et al., 2013), but

it remains unknown by which genes these are encoded.

Page 15: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

13

3. PLC in plant stress and development

Although PLC signaling in plants is still obscure, various roles in plant development and stress have

been proposed (Munnik, 2014; Pokotylo et al., 2014).

3.1 PLC in plant-stress responses

Over the years, several papers on the involvement of PLC in stress signaling have appeared, bringing

new insights of its role in plants (Munnik and Vermeer, 2010; Munnik, 2014; Pokotylo et al., 2014).

3.1.1 Osmotic-,heat- and cold stress

Upon abiotic stress, such as salinity, drought, heat and cold, several PLC genes are induced and

increased PLC activity has been reported in different plant systems (Hirayama et al., 1995; Ruelland et

al., 2002; Hunt et al., 2004; Das et al., 2005; Vergnolle et al., 2005; Zhai et al., 2005; Skinner et al.,

2005; Liu, Huang, et al., 2006; Sui et al., 2008; Tasma et al., 2008; Munnik and Vermeer, 2010; Wang

et al., 2015; Li et al., 2017). Accumulation of IP3 and increases of cytosolic Ca2+ have been correlated

and claimed to reflect PLC’s activity (DeWald et al., 2001; Takahashi et al., 2001; Ruelland et al., 2002;

Im et al., 2007; König et al., 2007; Zheng et al., 2012; Gao et al., 2014; Li et al., 2017). However, these

IP3 measurements were mainly based on so-called 3H-IP3-displacement assays, which in plants are quite

inaccurate because they contain very low IP3 levels, while IPPs that can affect the displacement assay,

are very abundant (see Munnik and Vermeer, 2010). Hence, whether "measured IP3" originates from

PLC hydrolysis or results from changes in other IPPs is still debatable. Yet increases of IP3 upon heat or

salt stress are typically accompanied by increases of PIP2 (Pical et al., 1999; DeWald et al., 2001; Zonia

and Munnik, 2004; Liu, et al., 2006; van Leeuwen et al., 2007; Darwish et al., 2009; Mishkind et al.,

2009; Horvath et al., 2012; Simon et al., 2014). The latter is likely due to the activation of PIPK rather

than inhibition of PLC or PIP2 phosphatases (Mishkind et al., 2009; Zarza, 2017). Theses PIP2

responses are relatively slow, however, accumulating after 15-30 min while Ca2+ responses already

occur within minutes, so these do not correlate at all. A decrease of PIP has also been reported in some

cases (Cho et al., 1993; Pical et al., 1999; DeWald et al., 2001; Zonia and Munnik, 2004; van Leeuwen

et al., 2007; Darwish et al., 2009; Mishkind et al., 2009; Vermeer et al., 2009; Horvath et al., 2012;

Munnik, 2014; Zarza et al., unpulished). Whether this reflects activation of PIPK or hydrolysis by PLC

remains unknown too. During salt stress, PIP2 has also been implicated in the formation of clathrin-

coated vesicles (CCV) (König et al., 2008). Whether PIP2 is used for IP3 or functions as a signaling

molecule itself remains to be addressed.

3.1.2 Abscisic acid (ABA)

PLC signaling in response to ABA has been studied for many years, especially in relation to the

induction of stomatal closure. Initially, most studies tried to link IP3/Ca2+-signaling with this event

(Gilroy et al., 1990; Blatt et al., 1990; Lee et al., 1996; Staxen et al., 1999; Blatt, 2000; Schroeder et al.,

Page 16: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

14

2001). However, the finding that (i) microinjected IP3 was converted into IP6 within seconds, (ii) that

IP6 was 10-100 times more efficient than IP3 in releasing intracellular Ca2+ , and (iii) that ABA triggered

IP6 formation in vivo within minutes of ABA treatment (Flores and Smart, 2000; Lemtiri-Chlieh et al.,

2000, 2003), raised new ideas for PLC signaling in response to ABA (Zonia and Munnik, 2006;

Munnik, 2014) However, IP6- activated Ca2+ channel (an "IP6 receptor") has still not been identified.

Apart from guard cells, ABA-induced IP3 responses have been reported for Arabidopsis seedlings

(Burnette et al., 2003; Perera et al., 2006). However, these studies only focused on IP3 and no other

IPPs were measured.

Whether PIP2 levels change upon ABA treatment remains unclear (Munnik and Vermeer,

2010). Nevertheless, PIP2 has been shown to inhibit anion channels (Lee et al., 2007) and K+ efflux

channels (Ma et al., 2009), which would facilitate stomatal opening. The importance of PLC in ABA

signaling is illustrated by several reports. For example, silencing of a PLC in tobacco made the plant

less sensitive to ABA-induced stomatal closure, leading to a wilting phenotype (Hunt et al., 2003).

Similarly, silencing of Arabidopsis PLC1 was shown to be important for secondary ABA responses

(Sanchez and Chua, 2001). Several PLC genes are also induced by ABA treatment (Hirayama et al.,

1995; Lin et al., 2004; Tasma et al., 2008; Pokotylo et al., 2014).

3.1.3 Biotic stress

A role for PLC in plant defense has also been emerging (Laxalt and Munnik, 2002; Vossen et al., 2010;

Canonne et al., 2011; Abd-El-Haliem et al., 2012, 2016, Gonorazky et al., 2014, 2016; D’Ambrosio et

al., 2017). Based on changes in PPI-labeling profiles, PLC signaling is thought to be induced by various

pathogen-associated molecular patterns (PAMPs), also called elicitors. These include xylanase, flagellin

and chitin (Luit et al., 2000; Den Hartog et al., 2003; Yamaguchi et al., 2005). It is also triggered by

many other chemical and biological inducers, such as benzothiadiazole (BTH), salicylic acid (SA),

jasmonic acid (JA) (Song and Goodman, 2002) and avirulence (Avr) proteins (De Jong et al., 2004;

Andersson et al., 2006). Rapid PA production through PLC/DGK activity has been demonstrated to

occur after pathogen attack (Luit et al., 2000; Laxalt and Munnik, 2002; Hartog et al., 2003; De Jong et

al., 2004; Canonne et al., 2011) and some of the responses have been shown involve nitric oxide (NO)

signaling (Laxalt et al., 2007; Lanteri et al., 2011; Raho et al., 2011). Interestingly, PA accumulation is

accompanied by reactive oxygen species (ROS) production which are playing an important role in

defense response (Yamaguchi et al., 2003; De Jong et al., 2004). In addition, it has been suggested that

PA might activate plant defenses via mitogen-activated protein kinase (MAPK) signaling cascades

(Laxalt and Munnik, 2002; Canonne et al., 2011). Through gene silencing, several tomato PLCs have

been shown to be involved in disease resistance (Vossen et al., 2010; Gonorazky et al., 2014, 2016;

Abd-El-Haliem et al., 2016; D’Ambrosio et al., 2017).

3.2 PLC in plant development

Page 17: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

15

In addition to stress, PLC is also involved in various aspects of plant growth- and development.

Arabidopsis PLC2 has been found to regulate reproductive development(Li et al., 2015; Di Fino et al.,

2017) and a similar function was observed for PLC1 in Torenia fournieri (Song et al., 2008). Petunia

PLC1 and tobacco PLC3 regulate tip growth in pollen tubes (Dowd et al., 2006; Helling et al., 2006).

During pollen tube elongation, PLC was found at the flanks of tip, whereas PIP2 mainly accumulated in

the apex (Ischebeck et al., 2010; Grierson et al., 2014). It is well known that PIP2 is essential for polar

tip growth, and this distribution of PLC would keep the PIP2 gradient directed to the tip, which is

important for polarized growth (Ischebeck et al., 2010; Heilmann and Ischebeck, 2016). In

Physcomitrella patens, loss of PLC1 resulted in insensitivity to cytokinin and exhibited a paler

phenotype, caused by reduced chlorophyll. The response to gravitropism was also reduced in mutant

filaments (Repp et al., 2004). IP3 has been suggested to play an important role in gravitropism (Perera

et al., 2001). However, this was based on the same inaccurate (section 3.1.1) IP3 kit and before the

realization that plants lack IP3 receptor. Auxin plays an important role in gravitropism too and as such

IP3/TIR1 could play a potential role here as well. In Brassica napus, overexpression of PLC2 enhanced

photosynthesis, changed hormone distributions, caused an earlier shift to reproductive phase and

decreased maturation time (Georges et al., 2009).

4 Scope of this thesis: dissection of the role of Arabidopsis PLC3, PLC5 and PLC7 in plant

signaling

To be able to study the role of PLC signaling in plants more precisely, we needed additional molecular

tools. While the model system Arabidopsis contains 9 PLCs (section 2.1), very little was known about

their individual function/contribution in plant stress and development. Hence, we used a reversed-

genetics approach, using Arabidopsis T-DNA insertion lines from different collections. As such, we

found that plc3 mutants exhibited a lateral-root phenotype (Chapter 2). Since the phenotype was quite

mild and PLC3-promotor GUS lines revealed specific expression in the vascular tissue, with a

particular segmentation pattern that indicated a role for this PLC in lateral root formation, we decided to

search for redundant PLCs in this process. As such we found PLC5 (Chapter 3) and PLC7 (Chapter 4),

with each revealing novel features.

4.1 Characteristics of the Arabidopsis PLC3, PLC5 and PLC7 genes

PLC3 (At4g38530), PLC5 (At5g58690) and PLC7 (At3g55940) belong to different subfamilies (Fig. 4).

Figure 5 shows the multiple sequence alignment of the amino acid sequences, with the four conserved

domains underlined in colour. No obvious differences between the PLCs were noted except for a stretch

of negatively-charged aspartic acids (Asp, D) in the linker-region between domain X and -Y of PLC7.

Predicted proteins were for ~45-53% identical (Fig. 5b). Expression analyses based on Genevestigator

data indicate that all 3 PLCs are expressed throughout development, with PLC7 being relatively lowest

but quite abundant in siliques, i.e. during seed development (Fig.6).

Page 18: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

16

Figure 5. Sequence alignment of Arabidopsis PLC3, PLC5 and PLC7. (a) Multiple sequence alignment of the amino acid sequences. Amino acids are color coded, with Red indicating hydrophobic; Green, polar; Blue, negatively charged; Pink, positively charged amino acids. The four conserved domains are underlined with colors too. Blue, EF-hand; Red, X domain; Pink, Y domain; Green, C2 domain. '*' represents identical residues in all sequences; ':' means conserved substitutions between similar residues have occurred; '.' indicates the semi-conserved substitutions between similar residues. Multiple sequence alignment by MUSCLE (3.8) (http://www.ebi.ac.uk/Tools/msa/muscle/). (b) Amino acid identity of PLC3, PLC5 and PLC7.

(b)

Page 19: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

17

Figure 6. Gene expression of PLC3, PLC5 and PLC7 based on Genevestigator. (a) Ten developmental stages from data selection AT_AFF_ATH1-2 showing 3 measures or 3 PLC genes. (b) Similar to (a) but showing the relative expression levels of each gene per developmental stage. 4.2 Techniques to study PLC signaling in plant

To study the role of PLC in plant cell signaling, biochemical and genetic approaches have been used in

this study. Here, these two techniques will be briefly described.

PLC activity can be determined in vivo by the measuring the turnover of its substrate, ie. PIP

and PIP2, and the appearance of products, i.e. IPPs and DAG. The latter, can be witnessed from its

conversion into PA via DGK. Since the concentration of the PPIs is very low, a radioactive 32Pi-

(32PO43+-) labeling technique is required to visualize the PPI and PA (DAG is no labeled). When taken

up by cells, the 32P first labels the ATP-pool and is then quickly incorporated into those phospholipids

that are produced via lipid kinases using this ATP, such as PI- and PIP kinase. The 32P-label is also

incorporated into structural phospholipids but this pathway is much slower (Arisz and Munnik, 2013;

Arisz et al., 2013). In fact, relatively short- and long-labeling times, can be used to distinguish between

PA formation from either PLC/DGK (short labeling, requiring DAG kinase activity and ATP) or de

novo- and PLD pathways (require long labeling) (Arisz and Munnik, 2013; Arisz et al., 2013).

Phospholipids can be separated by thin layer chromatography (TLC) and phosphoinositides can be

Page 20: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

18

further analyzed by High Performance Liquid Chromatography (HPLC) but then the fatty acids are

removed first, and the resulting glycerolphosphoinosides separated by a strong anion-exchanger

(Munnik, 2013; Munnik and Zarza, 2013). For IPP measurements, 3H-Myo-Inositol labeling is used in

combination with HPLC anion-exchange analyses (Lemtiri-Chlieh et al., 2000; Stevenson-Paulik et al.,

2005; Perera et al., 2008; Laha et al. 2015). These studies showed that IP1, IP2 and IP6 are the most

predominant IPPs, whereas IP3 levels are extremely low.

PLC function can also be studied by genetically modulating its gene expression, preferably in

the popular model plant system Arabidopsis thaliana (thale cress). With a fully mapped genome

sequence and short life cycle, many genetic tools have been developed, including Agrobacterium

tumefaciens-mediated gene transformation (Krysan et al., 1999), in which T-DNA insertion knock-out

(KO), knock down (KD) and overexpression (OE) mutants, as well as GUS/GFP reporter lines can be

obtained.

References Abd-El-Haliem A, Meijer H, Tameling W, Vossen J, and Joosten M (2012) Defense activation triggers differential expression

of phospholipase-C (PLC) genes and elevated temperature induces phosphatidic acid (PA) accumulation in tomato. Plant Signal Behav 7:1073–1078.

Abd-El-Haliem AM, Vossen JH, van Zeijl A, Dezhsetan S, Testerink C, Seidl MF, Beck M, Strutt J, Robatzek S, and Joosten MHAJ (2016) Biochemical characterization of the tomato phosphatidylinositol-specific phospholipase C (PI-PLC) family and its role in plant immunity. Biochim Biophys Acta - Mol Cell Biol Lipids 1861:1365–1378.

Allen GJ, and Sanders D (1994) Osmotic stress enhances the competence of Beta vulgaris vacuoles to respond to inositol 1,4,5-trisphosphate. Plant J 6:687–695.

Andersson MX, Kourtchenko O, Dangl JL, Mackey D, and Ellerström M (2006) Phospholipase-dependent signalling during the AvrRpm1- and AvrRpt2-induced disease resistance responses in Arabidopsis thaliana. Plant J 47:947–959.

Arisz SA, and Munnik T (2013) Distinguishing phosphatidic acid pools from De Novo synthesis, PLD, and DGK, in Plant Lipid Signaling Protocols (Munnik T, and Heilmann I eds) pp 55–62.

Arisz SA, Testerink C, and Munnik T (2009) Plant PA signaling via diacylglycerol kinase. Biochim Biophys Acta 1791:869–875.

Arisz SA, van Wijk R, Roels W, Zhu JK, Haring MA, and Munnik T (2013) Rapid phosphatidic acid accumulation in response to low temperature stress in Arabidopsis is generated through diacylglycerol kinase. Front Plant Sci 4:1–15.

Assmann SM, and Jegla T (2016) Guard cell sensory systems: recent insights on stomatal responses to light, abscisic acid, and CO2. Curr Opin Plant Biol 33:157–167.

Blatt MR (2000) Cellular signaling and volume control in stomatal movements in plants. Annu Rev Cell Dev Biol 16:221–241. Blatt MR, Thiel G, and Trentham DR (1990) Reversible inactivation of K+ channels of Vicia stomatal guard cells following

the photolysis of caged inositol 1,4,5-trisphosphate. Nature 346:766–769. Bunney TD, and Katan M (2011) PLC regulation: Emerging pictures for molecular mechanisms. Trends Biochem Sci 36:88–

96. Burnette RN, Gunesekera BM, and Gillaspy GE (2003) An Arabidopsis inositol 5-phosphatase gain-of-function alters abscisic

acid signaling. Plant Physiol 132:1011–1019. Burton A, Hu X, and Saiardi A (2009) Are inositol pyrophosphates signalling molecules? J Cell Physiol 220:8–15. Canonne J, Froidure-Nicolas S, and Rivas S (2011) Phospholipases in action during plant defense signaling. Plant Signal

Behav 6:13–18. D’Ambrosio JM, Couto D, Fabro G, Scuffi1 D, Lamattina L, Munnik T, Álvarez ME, Andersson MX, Zipfel C, and Laxalt

AM (2017) Phosphoinositide-specific phospholipase C2 (PLC2)associates with RBOHD and modulates ROS production and MAMP-triggered immunity in Arabidopsis. Plant J submitted.

Darwish E, Testerink C, Khalil M, El-Shihy O, and Munnik T (2009) Phospholipid signaling responses in salt-stressed rice leaves. Plant Cell Physiol 50:986–997.

Das S, Hussain A, Bock C, Keller WA, and Georges F (2005) Cloning of Brassica napus phospholipase C2 (BnPLC2), phosphatidylinositol 3-kinase (BnVPS34) and phosphatidylinositol synthase1 (BnPtdIns S1) - Comparative analysis of the effect of abiotic stresses on the expression of phosphatidylinositol signal transdu. Planta 220:777–784.

De Jong CF, Laxalt AM, Bargmann BOR, De Wit PJGM, Joosten MHAJ, and Munnik T (2004) Phosphatidic acid accumulation is an early response in the Cf-4/Avr4 interaction. Plant J 39:1–12.

Desai M, Rangarajan P, Donahue JL, Williams SP, Land ES, Mandal MK, Phillippy BQ, Perera IY, Raboy V, and Gillaspy GE (2014) Two inositol hexakisphosphate kinases drive inositol pyrophosphate synthesis in plants. Plant J 80:642–653.

DeWald DB, Torabinejad J, Jones C, Shope JC, Cangelosi R, Thompson JE, Prestwich GD, and Hama H (2001) Rapid

Page 21: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

19

accumulation of phosphatidylinositol 4,5-bisphosphate and inositol 1,4,5-trisphosphate correlates with calcium mobilization in salt-stressed arabidopsis. Plant Physiol 126:759–769.

Di Fino L, D’Ambrosio J, Tejos R, van Wijk R, Lamattina T, Munnik T, Pagnussat G, and Laxalt AM (2017) Arabidopsis phosphatidylinositol-phospholipase C2 (PLC2) is involved in female gametogenesis and embryo development. Planta 2:1–12.

Dowd PE, Coursol S, Skirpan AL, Kao T, and Gilroy S (2006) Petunia phospholipase C1 is involved in pollen tube growth. Plant Cell 18:1438–1453.

Flores S, and Smart CC (2000) Abscisic acid-induced changes in inositol metabolism in Spirodela polyrrhiza. Planta 211:823–832.

Furt F, Simon-Plas F, and Mongrand S (2011) Lipids of the Plant Plasma Membrane, in The Plant Plasma Membrane (Murphy AS et al. eds) pp 3–30, Springer Berlin Heidelberg, Berlin, Heidelberg.

Gao K, Liu YL, Li B, Zhou RG, Sun DY, and Zheng SZ (2014) Arabidopsis thaliana phosphoinositide-specific phospholipase C isoform 3 (AtPLC3) and AtPLC9 have an additive effect on thermotolerance. Plant Cell Physiol 55:1873–1883.

Georges F, Das S, Ray H, Bock C, Nokhrina K, Kolla VA, and Keller W (2009) 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 32:1664–1681.

Gillaspy GE (2013) The role of phosphoinositides and inositol phosphates in plant cell signaling, in Lipid-mediated protein signaling (Capelluto DGS ed) pp 141–157.

Gilroy S, Read ND, and Trewavas a J (1990) Elevation of cytoplasmic calcium by caged calcium or caged inositol triphosphate initiates stomatal closure. Nature 346:769–771.

Gonorazky G, Guzzo MC, and Laxalt AM (2016) Silencing of the tomato phosphatidylinositol-phospholipase C2 (SlPLC2) reduces plant susceptibility to Botrytis cinerea. Mol Plant Pathol 2:1–10.

Gonorazky G, Ramirez L, Abd-El-Haliem A, Vossen JH, Lamattina L, ten Have A, Joosten MHAJ, and Laxalt AM (2014) The tomato phosphatidylinositol-phospholipase C2 (SlPLC2) is required for defense gene induction by the fungal elicitor xylanase. J Plant Physiol 171:959–965.

Grierson C, Nielsen E, Ketelaar T, and Schiefelbein J (2014) Root Hairs. Arab B 12:1–25. Hartog M Den, Verhoef N, and Munnik T (2003) Nod factor and elicitors activate different phospholipid signaling pathways in

suspension-cultured Alfalfa cells. Plant Physiol 132:311–317. Heilmann I (2016) Phosphoinositide signaling in plant development. Development 143:2044–2055. Heilmann I, and Ischebeck T (2016) Male functions and malfunctions: the impact of phosphoinositides on pollen development

and pollen tube growth. Plant Reprod 29:3–20, Springer Berlin Heidelberg. Helling D, Possart A, Cottier S, Klahre U, and Kost B (2006) Pollen tube tip growth depends on plasma membrane

polarization mediated by tobacco PLC3 activity and endocytic membrane recycling. Plant Cell 18:3519–3534. Hirayama T, Ohtot C, Mizoguchi T, and Shinozaki K (1995) A gene encoding a phosphatidylinositol-specific phospholipase C

is induced by dehydration and salt stress in Arabidopsis thaliana. Proc Natl Acad Sci USA 92:3903–3907. Hong Y, Zhao J, Guo L, Kim SC, Deng X, Wang G, Zhang G, Li M, and Wang X (2016) Plant phospholipases D and C and

their diverse functions in stress responses. Prog Lipid Res 62:55–74. Horvath I, Glatz A, Nakamoto H, Mishkind ML, Munnik T, Saidi Y, Goloubinoff P, Harwood JL, and Vigh L (2012) Heat

shock response in photosynthetic organisms: Membrane and lipid connections. Prog Lipid Res 51:208–220. Hou Q, Ufer G, and Bartels D (2016) Lipid signalling in plant responses to abiotic stress. Plant, Cell Environ 39:1029–1048. Hsieh EJ, Cheng MC, and Lin TP (2013) Functional characterization of an abiotic stress-inducible transcription factor

AtERF53 in Arabidopsis thaliana. Plant Mol Biol 82:223–237. Hunt L, Mills LN, Pical C, Leckie CP, Aitken FL, Kopka J, Mueller-Roeber B, McAinsh MR, Hetherington AM, and Gray JE

(2003) Phospholipase C is required for the control of stomatal aperture by ABA. Plant J 34:47–55. Hunt L, Otterhag L, Lee JC, Lasheen T, Hunt J, Seki M, Shinozaki K, Sommarin M, Gilmour DJ, Pical C, and Gray JE (2004)

Gene-specific expression and calcium activation of Arabidopsis thaliana phospholipase C isoforms. New Phytol 162:643–654.

Im YJ, Perera IY, Brglez I, Davis AJ, Stevenson-Paulik J, Phillippy BQ, Johannes E, Allen NS, and Boss WF (2007) Increasing plasma membrane phosphatidylinositol(4,5)bisphosphate biosynthesis increases phosphoinositide metabolism in Nicotiana tabacum. Plant Cell 19:1603–1616.

Ischebeck T, Seiler S, and Heilmann I (2010) At the poles across kingdoms: Phosphoinositides and polar tip growth. Protoplasma 240:13–31.

Kim YJ, Kim JE, Lee JH, Lee MH, Jung HW, Bahk YY, Hwang BK, Hwang I, and Kim WT (2004) The Vr-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.). FEBS Lett 556:127–136.

König S, Ischebeck T, Lerche J, Stenzel I, and Heilmann I (2008) Salt-stress-induced association of phosphatidylinositol 4,5-bisphosphate with clathrin-coated vesicles in plants. Biochem J 415:387–399.

König S, Mosblech A, and Heilmann I (2007) Stress-inducible and constitutive phosphoinositide pools have distinctive fatty acid patterns in Arabidopsis thaliana. FASEB J 21:1958–1967.

Krysan PJ, Young JC, and Sussman MR (1999) T-DNA as an insertional mutagen in Arabidopsis. Plant Cell 11:2283–2290. Kuo HF, Chang TY, Chiang SF, Wang WD, Charng YY, and Chiou TJ (2014) Arabidopsis inositol pentakisphosphate 2-

kinase, AtIPK1, is required for growth and modulates phosphate homeostasis at the transcriptional level. Plant J 80:503–515.

Laha D, Johnen P, Azevedo C, Dynowski M, Wei M, Capolicchio S, Mao H, Iven T, Steenbergen M, Freyer M, Gaugler P, de Campos MKF, Zheng N, Feussner I, Jessen HJ, Van Wees SCM, Saiardi A, and Schaaf G (2015) VIH2 regulates the synthesis of inositol pyrophosphate InsP8 and jasmonate-dependent defenses in Arabidopsis. Plant Cell 27:1082–1097.

Page 22: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

20

Laha D, Parvin N, Dynowski M, Johnen P, Mao H, Bitters ST, Zheng N, and Schaaf G (2016) Inositol polyphosphate binding specificity of the jasmonate receptor complex. Plant Physiol 171:2364–2370.

Lanteri ML, Lamattina L, and Laxalt AM (2011) Mechanisms of xylanase-induced nitric oxide and phosphatidic acid production in tomato cells. Planta 234:845–855.

Laxalt AM, and Munnik T (2002) Phospholipid signalling in plant defence. Curr Opin Plant Biol 5:332–338. Laxalt AM, Raho N, Ten Have A, and Lamattina L (2007) Nitric oxide is critical for inducing phosphatidic acid accumulation

in xylanase-elicited tomato cells. J Biol Chem 282:21160–21168. Lee HS, Lee DH, Cho HK, Kim SH, Auh JH, and Pai HS (2015) 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 27:417–431. Lee Y, Choi YB, Suh S, Lee J, Assmann SM, Joe CO, Kelleher JF, and Crain RC (1996) Abscisic acid-induced

phosphoinositide turnover in guard cell protoplasts of Vicia faba. Plant Physiol 110:987–996. Lee Y, Kim YW, Jeon BW, Park KY, Suh SJ, Seo J, Kwak JM, Martinoia E, Hwang I, and Lee Y (2007) Phosphatidylinositol

4,5-bisphosphate is important for stomatal opening. Plant J 52:803–816. Lemtiri-Chlieh F, MacRobbie EAC, and Brearley CA (2000) Inositol hexakisphosphate is a physiological signal regulating the

K+-inward rectifying conductance in guard cells. Proc Natl Acad Sci U S A 97:8687–8692. Lemtiri-Chlieh F, MacRobbie EAC, Webb AAR, Manison NF, Brownlee C, Skepper JN, Chen J, Prestwich GD, and Brearley

CA (2003) Inositol hexakisphosphate mobilizes an endomembrane store of calcium in guard cells. Proc Natl Acad Sci U S A 100:10091–10095.

Li L, He Y, Wang Y, Zhao S, Chen X, Ye T, Wu Y, and Wu Y (2015) Arabidopsis PLC2 is involved in auxin-modulated reproductive development. Plant J 84:504–515.

Li L, Wang F, Yan P, Jing W, Zhang C, Kudla J, and Zhang W (2017) A phosphoinositide-specific phospholipase C pathway elicits stress-induced Ca 2+ signals and confers salt tolerance to rice. New Phytol, doi: 10.1111/nph.14426.

Lin WH, Ye R, Ma H, Xu ZH, and Xue HW (2004) DNA chip-based expression profile analysis indicates involvement of the phosphatidylinositol signaling pathway in multiple plant responses to hormone and abiotic treatments. Cell Res 14:34–45.

Liu HT, Gao F, Cui SJ, Han JL, Sun DY, and Zhou RG (2006) Primary evidence for involvement of IP3 in heat-shock signal transduction in Arabidopsis. Cell Res 16:394–400.

Liu HT, Huang WD, Pan QH, Weng FH, Zhan JC, Liu Y, Wan SB, and Liu YY (2006) Contributions of PIP2-specific-phospholipase C and free salicylic acid to heat acclimation-induced thermotolerance in pea leaves. J Plant Physiol 163:405–416.

Liu HT, Liu YY, Pan QH, Yang HR, Zhan JC, and Huang WD (2006) Novel interrelationship between salicylic acid, abscisic acid, and PIP 2-specific phospholipase C in heat acclimation-induced thermotolerance in pea leaves. J Exp Bot 57:3337–3347.

Luit AH Van Der, Piatti T, Doorn A Van, Musgrave A, Felix G, Boller T, and Munnik T (2000) Elicitation of Suspension-Cultured Tomato Cells Triggers the Formation of Phosphatidic Acid and Diacylglycerol Pyrophosphate 1. Plant Physiol 123:1507–1515.

Ma Y, Szostkiewicz I, Korte A, Moes D, Yang Y, Christmann A, and Grill E (2009) Regulators of PP2C phosphatase activity function as abscisic acid sensors. Science 324:1064–1069.

Mendrinna A, and Persson S (2015) Root hair growth: it’s a one way street. F1000Prime Rep 7:23–27. Michell RH (2008) Inositol derivatives: evolution and functions. Nat Rev 9:151–161. Mishkind M, Vermeer JEM, Darwish E, and Munnik T (2009) Heat stress activates phospholipase D and triggers PIP

accumulation at the plasma membrane and nucleus. Plant J 60:10–21. Mueller-Roeber B, and Pical C (2002) Inositol phospholipid metabolism in Arabidopsis. Characterized and putative isoforms

of inositol phospholipid kinase and phosphoinositide-specific phospholipase C. Plant Physiol 130:22–46. Munemasa S, Hauser F, Park J, Waadt R, Brandt B, and Schroeder JI (2015) Mechanisms of abscisic acid-mediated control of

stomatal aperture. Curr Opin Plant Biol 28:154–162. Munnik T (2013) Analysis of D3-,4-,5-Phosphorylated Phosphoinositides Using HPLC, in Plant Lipid Signaling Protocols

(Munnik T, and Heilmann I eds) pp 17–24. Munnik T (2001) Phosphatidic acid: an emerging plant lipid second messenger. Trends Plant Sci 6:227–233. Munnik T (2014) PI-PLC: Phosphoinositide-phospholipase C in plant signalling, in Phospholipases in Plant Signaling (Wang

X ed) pp 27–54. Munnik T, Irvine RF, and Musgrave A (1998) Phospholipid signalling in plants. Biochim Biophys Acta - Lipids Lipid Metab

1389:222–272. Munnik T, and Nielsen E (2011) Green light for polyphosphoinositide signals in plants. Curr Opin Plant Biol 14:489–497. Munnik T, and Testerink C (2009) Plant phospholipid signaling: “in a nutshell”. J Lipid Res 50:S260–S265. Munnik T, and Vermeer JEM (2010) Osmotic stress-induced phosphoinositide and inositol phosphate signalling in plants.

Plant Cell Environ 33:655–669. Munnik T, and Zarza X (2013) Analyzing Plant Signaling Phospholipids Through 32Pi-Labeling and TLC, in Plant Lipid

Signaling Protocols (Munnik T, and Heilmann I eds) pp 3–15. Murphy AM, Otto B, Brearley CA, Carr JP, and Hanke DE (2008) A role for inositol hexakisphosphate in the maintenance of

basal resistance to plant pathogens. Plant J 56:638–652. Otterhag L, Sommarin M, and Pical C (2001) N-terminal EF-hand-like domain is required for phosphoinositide- specific

phospholipase C activity in Arabidopsis thaliana. FEBS Lett 497:165–170. Perera IK, Hilmann I, Chang SC, Boss WF, and Kaufman PB (2001) A role for inositol 1,4,5-trisphosphate in gravitropic

signaling and the retention of cold-perceived gravistimulation of oat shoot pulvini. Plant Physiol 125:1499–1507. Perera IY, Hung C-Y, Moore CD, Stevenson-Paulik J, and Boss WF (2008) Transgenic Arabidopsis plants expressing the type

1 inositol 5-phosphatase exhibit increased drought tolerance and altered abscisic acid signaling. Plant Cell 20:2876–

Page 23: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generalintroduction

21

2893. Perera IY, Hung C, Brady S, Muday GK, and Boss WF (2006) A Universal Role for Inositol 1 , 4 , 5-Trisphosphate-Mediated

Signaling in Plant Gravitropism. Plant Physiol 140:746–760. Pical C, Westergren T, Dove SK, Larsson C, and Sommarin M (1999) Salinity and hyperosmotic stress induce rapid increases

in phosphatidylinositol 4,5-bisphosphate, diacylglycerol pyrophosphate, and phosphatidylcholine in Arabidopsis thaliana cells. J Biol Chem 274:38232–38240.

Pokotylo I, Kolesnikov Y, Kravets V, Zachowski A, and Ruelland E (2014) Plant phosphoinositide-dependent phospholipases C: Variations around a canonical theme. Biochimie 96:144–157.

Puga MI, Mateos I, Charukesi R, Wang Z, Franco-Zorrilla JM, de Lorenzo L, Irigoyen ML, Masiero S, Bustos R, Rodríguez J, Leyva A, Rubio V, Sommer H, and Paz-Ares J (2014) SPX1 is a phosphate-dependent inhibitor of Phosphate Starvation Response 1 in Arabidopsis. Proc Natl Acad Sci U S A 111:14947–14952.

Raho N, Ramirez L, Lanteri ML, Gonorazky G, Lamattina L, ten Have A, and Laxalt AM (2011) Phosphatidic acid production in chitosan-elicited tomato cells, via both phospholipase D and phospholipase C/diacylglycerol kinase, requires nitric oxide. J Plant Physiol 168:534–539.

Rebecchi MJ, and Pentyala SN (2000) Structure , Function , and Control of Phosphoinositide-Specific Phospholipase C. 80:1291–1335.

Repp A, Mikami K, Mittmann F, and Hartmann E (2004) Phosphoinositide-specific phospholipase C is involved in cytokinin and gravity responses in the moss Physcomitrella patens. Plant J 40:250–259.

Ruelland E, Cantrel C, Gawer M, Kader JC, and Zachowski A (2002) Activation of phospholipases C and D is an early response to a cold exposure in Arabidopsis suspension cells. Plant Physiol 130:999–1007.

Rupwate SD, and Rajasekharan R (2012) C2 domain is responsible for targeting rice phosphoinositide specific phospholipase C. Plant Mol Biol 78:247–258.

Sanchez JP, and Chua NH (2001) Arabidopsis PLC1 is required for secondary responses to abscisic acid signals. Plant Cell 13:1143–1154.

Schroeder JI, Allen GJ, Hugouvieux V, Kwak JM, and Waner D (2001) Guard cell signal transduction. Annu Rev Plant Physiol plant mol biol 52:627–658.

Sheard LB, Tan X, Mao H, Withers J, Ben-Nissan G, Hinds TR, Kobayashi Y, Hsu F-F, Sharon M, Browse J, He SY, Rizo J, Howe GA, and Zheng N (2010) Jasmonate perception by inositol-phosphate-potentiated COI1-JAZ co-receptor. Nature 468:400–405, Nature Publishing Group, a division of Macmillan Publishers Limited. All Rights Reserved.

Shears SB (2009) Diphosphoinositol polyphosphates: metabolic messengers? Mol Pharmacol 76:236–252. Shi J, Gonzales RA, and Bhattacharyya MK (1995) Characterization of a plasma membrane-associated phosphoinositide-

specific phospholipase C from soybean. Simon MLA, Platre MP, Assil S, Van Wijk R, Chen WY, Chory J, Dreux M, Munnik T, and Jaillais Y (2014) A multi-

colour/multi-affinity marker set to visualize phosphoinositide dynamics in Arabidopsis. Plant J 77:322–337. Singh A, Kanwar P, Pandey A, Tyagi AK, Sopory SK, Kapoor S, and Pandey GK (2013) Comprehensive Genomic Analysis

and Expression Profiling of Phospholipase C Gene Family during Abiotic Stresses and Development in Rice. PLoS One 8:1–14.

Skinner DZ, Bellinger BS, Halls S, Baek K-H, Garland-Campbell K, and Siems WF (2005) Phospholipid Acyl Chain and Phospholipase Dynamics during Cold Acclimation of Winter Wheat. Crop Sci 45:1858–1867.

Song F, and Goodman RM (2002) 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 61:31–40.

Song M, Liu S, Zhou Z, and Han Y (2008) TfPLC1, a gene encoding phosphoinositide-specific phospholipase C, is predominantly expressed in reproductive organs in Torenia fournieri. Sex Plant Reprod 21:259–267.

Staxen I, Pical C, Montgomery LT, Gray JE, Hetherington a M, and McAinsh MR (1999) Abscisic acid induces oscillations in guard-cell cytosolic free calcium that involve phosphoinositide-specific phospholipase C. Proc Natl Acad Sci U S A 96:1779–1784.

Stevenson-Paulik J, Bastidas RJ, Chiou S-T, Frye R a, and York JD (2005) Generation of phytate-free seeds in Arabidopsis through disruption of inositol polyphosphate kinases. Proc Natl Acad Sci U S A 102:12612–12617.

Sui Z, Niu L, Yue G, Yang A, and Zhang J (2008) Cloning and expression analysis of some genes involved in the phosphoinositide and phospholipid signaling pathways from maize (Zea mays L.). Gene 426:47–56.

Swann K, and Lai FA (2016) The sperm phospholipase C-ζ and Ca2+ signalling at fertilization in mammals. Biochem Soc Trans 44:267–72.

Takahashi S, Katagiri T, Hirayama T, Yamaguchi-Shinozaki K, and Shinozaki K (2001) Hyperosmotic stress induces a rapid and transient increase in inositol 1,4,5-trisphosphate independent of abscisic acid in Arabidopsis cell culture. Plant Cell Physiol 42:214–222.

Tan X, Calderon-Villalobos LI, Sharon M, Zheng C, Robinson C V, Estelle M, and Zheng N (2007) Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature 446:640–645.

Tasma IM, Brendel V, Whitham S., and Bhattacharyya MK (2008) Expression and evolution of the phosphoinositide-specific phospholipase C gene family in Arabidopsis thaliana. Plant Physiol Biochem 46:627–637.

Testerink C, and Munnik T (2011) Molecular, cellular, and physiological responses to phosphatidic acid formation in plants. J Exp Bot 62:2349–2361.

Testerink C, and Munnik T (2005) Phosphatidic acid: a multifunctional stress signaling lipid in plants. Trends Plant Sci 10:368–375.

Thota SG, and Bhandari R (2015) The emerging roles of inositol pyrophosphates in eukaryotic cell physiology. J Biosci 40:593–605.

Van Hooren M, and Munnik T (2017) The plant plasma membrane. eLS Press. van Leeuwen W, Vermeer JEM, Gadella TWJ, and Munnik T (2007) Visualization of phosphatidylinositol 4,5-bisphosphate in

Page 24: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter1

22

the plasma membrane of suspension-cultured tobacco BY-2 cells and whole Arabidopsis seedlings. Plant J 52:1014–1026.

Vergnolle C, Vaultier MN, Taconnat L, Renou JP, Kader JC, Zachowski A, and Ruelland E (2005) The cold-induced early activation of phospholipase C and D pathways determines the response of two distinct clusters of genes in Arabidopsis cell suspensions. Plant Physiol 139:1217–1233.

Vermeer JEM, Thole JM, Goedhart J, Nielsen E, Munnik T, and Gadella TWJ (2009) Imaging phosphatidylinositol 4-phosphate dynamics in living plant cells. Plant J 57:356–372.

Vossen JH, Abd-El-Haliem A, Fradin EF, Van Den Berg GCM, Ekengren SK, Meijer HJG, Seifi A, Bai Y, Ten Have A, Munnik T, Thomma BPHJ, and Joosten MHAJ (2010) 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 62:224–239.

Wang CR, Yang AF, Yue GD, Gao Q, Yin HY, and Zhang JR (2008) Enhanced expression of phospholipase C 1 (ZmPLC1) improves drought tolerance in transgenic maize. Planta 227:1127–1140.

Wang F, Deng Y, Zhou Y, Dong J, Chen H, Dong Y, Wang N, Li X, and Li H (2015) Genome-wide analysis and expression profiling of the phospholipase c gene family in soybean (glycine max). PLoS One 10.

Wang X, Devaiah SP, Zhang W, and Welti R (2006) Signaling functions of phosphatidic acid. Prog Lipid Res 45:250–278. Wild R, Gerasimaite R, Jung J-Y, Truffault V, Pavlovic I, Schmidt A, Saiardi A, Jessen HJ, Poirier Y, Hothorn M, and Mayer

A (2016) Control of eukaryotic phosphate homeostasis by inositol polyphosphate sensor domains. Science (80- ) 352:986 LP-990.

Yamaguchi T, Minami E, and Shibuya N (2003) Activation of phospholipases by N-acetylchitooligosaccharide elicitor in suspension-cultured rice cells mediates reactive oxygen generation. Physiol Plant 118:361–370.

Yamaguchi T, Minami E, Ueki J, and Shibuya N (2005) Elicitor-induced activation of phospholipases plays an important role for the induction of defense responses in suspension-cultured rice cells. Plant Cell Physiol 46:579–587.

York JD (2006) Regulation of nuclear processes by inositol polyphosphates. Biochim Biophys Acta - Mol Cell Biol Lipids 1761:552–559.

Zarza X (2017) Polyamine metabolism and activation of lipid signaling pathways in Arabidopsis thaliana. Zhai S, Sui Z, Yang A, and Zhang J (2005) Characterization of a novel phosphoinositide-specific phospholipase C from Zea

mays and its expression in Escherichia coli. Biotechnol Lett 27:799–804. Zheng S, Liu Y, Li B, Shang Z, Zhou R, and Sun D (2012) Phosphoinositide-specific phospholipase C9 is involved in the

thermotolerance of Arabidopsis. Plant J 69:689–700. Zonia L, Cordeiro S, Tupy J, and Feijo JA (2002) 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 14:2233–2249. Zonia L, and Munnik T (2006) Cracking the Green Paradigm: Functional Coding of Phosphoinositide Signals in Plant Stress

Responses, in Biology of Inositols and Phosphoinositides: Subcellular Biochemistry (Majumder AL, and Biswas BB eds) pp 207–237.

Zonia L, and Munnik T (2004) Osmotically induced cell swelling versus cell shrinking elicits specific changes in phospholipid signals in tobacco pollen tubes. Plant Physiol 134:813–823.

Page 25: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

23

Chapter 2 A role for Arabidopsis phospholipase C3 (PLC3) in seed germination, lateral root formation and stomatal closure Qianqian Zhang,1,2 Ringo van Wijk,1,2 Muhammad Shahbaz,1# Wendy Roels,1 Bas van Schooten,1 Joop E.M. Vermeer,1,3 Aisha Guardia, 4 Denise Scuffi, 4 Carlos García-Mata,4 Debabrata Laha,5 Phoebe Williams,6 Leo A.J. Willems,7 Wilco Ligterink,7 Susanne Hoffmann-Benning,8 Glenda Gillaspy,6 Gabriel Schaaf,5 Michel A. Haring,1 Ana M. Laxalt,4 & Teun Munnik1,2

1Section Plant Physiology or 2Section Plant Cell Biology, Swammerdam Institute for Life

Sciences, University of Amsterdam, Science Park 904, Amsterdam, 1098 XH, The Netherlands, 3Dept. of Plant and Microbial Biology, University of Zürich, Zürich, Switzerland,

4Instituto de

Investigaciones Biológicas (IIB-CONICET-UNMdP), Universidad Nacional de Mar del Plata, Mar del Plata, Argentina,

5Center for Plant Molecular Biology, University of Tübingen,

Tübingen, Germany, 6Dept. of Biochemistry, Virginia Polytechnic Institute and State

University, Blacksburg, VA, USA, 7

Lab. of Plant Physiology, Wageningen University, Droevendaalsesteeg 1, 6708 PB, Wageningen, The Netherlands, and

8Dept. of Biochemistry &

Molecular Biology, Michigan State University, East Lansing, MI, USA, #Current address: Dept. of Botany, University of Agriculture, Faisalabad, Pakistan.

Page 26: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

24

ABSTRACT

Phospholipase C (PLC) is best known for its role in generating second messengers by hydrolysis

of phosphatidylinositol 4,5-bisphosphate (PIP2) in mammalian cells. In plants however, PLC’s

role is less clear as plants lack the prime targets for both inositol 1,4,5-trisphosphate (i.e. a ligand-

gated Ca2+ channel) and diacylglycerol (i.e. protein kinase C and TRP-type ion channels). The

genome of Arabidopsis thaliana encodes for 9 PLC genes. Here, we analyzed the role of PLC3.

Promoter-GUS analyses revealed that PLC3 is specifically expressed in the vascular tissue (most

likely phloem) of roots, leaves and flowers, but also in guard cells and at the base of trichomes.

Knock-out mutants of PLC3 were found to be affected in seed germination, root development and

stomatal closure. Using in vivo 32Pi-lipid labeling analyses, we found that ABA stimulated the

formation of PIP2 in wild type germinating seeds, seedlings and guard cell-enriched leaf peels, but

not in plc3 mutants. The latter displayed decreased sensitivity to ABA during seed-germination

inhibition and ABA induced-stomatal closure. Overexpression of PLC3 enhanced drought

tolerance and decreased stomatal aperture. Together, our results uncovered novel roles for PLC3

in ABA signaling and plant development.

Key words: PLC; seed germination; lateral root formation; stomatal closure; drought tolerance.

Page 27: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

25

INTRODUCTION

Phospholipase C (PLC) is well known for its role in phospholipid signaling in animals. In this classical

paradigm, extracellular receptor occupation leads to the activation of intracellular PLC, which

hydrolyzes the minor lipid phosphatidylinositol 4,5-bisphosphate (PIP2) to produce two second

messengers, inositol 1,4,5- trisphosphate (IP3) and diacylglycerol (DAG). IP3 diffuses into the cytosol

where it triggers the release of Ca2+ from an intracellular store via an IP3 receptor that is a ligand-gated

Ca2+ channel, whereas DAG remains in the plasma membrane where it recruits and activates members

of the protein kinase C (PKC) family or stimulates TRP- (transient receptor potential-) ion channels.

Subsequent changes in Ca2+ and phosphorylation status affect multiple protein targets and hence,

downstream cellular processes (Irvine, 2006; Michell, 2008; Balla, 2013).

Less is clear about the PLC-signaling paradigm in plants (Munnik, 2014). Most importantly, all

higher plant genomes sequenced so far, lack homologs of an IP3 receptor, PKC or TRP channel, which

are supposed to be the primary targets of this signaling system (Wheeler and Brownlee, 2008; Munnik

and Testerink, 2009; Munnik and Vermeer, 2010; Munnik and Nielsen, 2011; Munnik, 2014; Heilmann,

2016). Initially, microinjected IP3 had been shown to release Ca2+ from an intracellular store (Gilroy et

al., 1990; Blatt et al., 1990; Allen and Sanders, 1994) indicating that plants cells exhibited a genuine IP3

receptor (Hunt and Gray, 2001) but Brearley's lab later provided evidence that this IP3 is

phosphorylated into IP6 within seconds, and that the latter compound is likely to be responsible for the

store-operated Ca2+ release (Lemtiri-Chlieh et al., 2000, 2003; Munnik and Vermeer, 2010). Similarly,

not DAG but its phosphorylated product, phosphatidic acid (PA) is emerging as the plant lipid-second

messenger (Munnik, 2001; Testerink and Munnik, 2005; Arisz et al., 2009; Pokotylo et al., 2014;

Munnik, 2014; Vermeer et al., 2017).

Evidence that PLC is important for plants has come from various studies. Silencing of PLC1 in

Arabidopsis and tobacco has indicated a role for a PLC in ABA signaling and stomatal closure

(Sanchez and Chua, 2001a; Hunt et al., 2003). ABA also induces the expression of some PLC genes

(Hirayama et al., 1995; Lin et al., 2004; Tasma et al., 2008; Pokotylo et al., 2014). A link between

ABA and polyphosphoinositide (PPI) turnover has been reported, but the data is quite controversial

(Munnik and Vermeer, 2010). Nonetheless, ABA has been shown to trigger IP6 responses within

minutes in potato guard cell protoplasts and duck weed turions (Flores and Smart, 2000; Lemtiri-Chlieh

et al., 2000, 2003), and to elevate intracellular Ca2+ levels in a variety of plant (Lee et al., 1996; Staxen

et al., 1999; Blatt, 2000; Schroeder et al., 2001; Munemasa et al., 2015; Assmann and Jegla, 2016)

Whether the formation of IP6 is PLC-dependent and related to ABA-mediated signaling is still unknown.

Apart from ABA, PLC signaling has been linked to several other abiotic stresses, including salt,

drought (mimicked by sorbitol, mannitol or PEG) and heat stress. Interestingly, these stresses are also

known to trigger an increase in the level of PIP2 (Pical et al., 1999; DeWald et al., 2001; Zonia and

Munnik, 2004; van Leeuwen et al., 2007; Darwish et al., 2009; Mishkind et al., 2009; Horvath et al.,

Page 28: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

26

2012). In some cases, IP3 responses were reported, but none addressed IP6 levels or other

inositolpolyphosphates (IPPs) that are emerging as signalling molecules, i.e. IP5, IP7 and IP8 (Takahashi

et al., 2001; Huang, et al., 2006; Liu, et al., 2006; Zheng et al., 2012; Gilaspy 2013; Laha et al., 2015,

2016). Decreases of PIP have also been reported (Cho et al., 1993; Pical et al., 1999; DeWald et al.,

2001; Vermeer et al., 2009; Zarza et al., unpublished) and theoretically, PLC could use PIP as a

substrate as well. In vitro, both PPIs are hydrolyzed equally well and in vivo, there is enough PIP in the

plasma membrane of plants, where PIP2 is typically missing, and in general only present at very low

concentrations (20-100x less than mammalian cells; Munnik et al., 1994; 1998a,b Munnik, 2014; Zarza

et al., unpublished). Interestingly, overexpression of PLC in maize, canola and tobacco has been shown

to increase the plant's tolerance to salinity-, drought- and/or cold stress (Wang et al., 2008; Georges et

al., 2009; Tripathy et al., 2011; Nokhrina et al., 2014), although it is not yet clear how the plant

achieves this (Das et al., 2005; Georges et al., 2009).

PLC has also been implicated in plant-microbe interactions (Laxalt and Munnik, 2002) both

symbiotic and pathogenic (Luit et al., 2000; Hartog et al., 2003; De Jong et al., 2004; Gonorazky et al.,

2014, 2016). Some of the pathogenic responses have been shown comprise nitric oxide (NO) signalling

(Lanteri et al., 2011; Raho et al., 2011). Recently, Vossen et al. (2010) presented the first genetic

evidence for PLC's contribution in tomato disease resistance (Vossen et al., 2010).

Apart from stress, PLC signalling has also been connected to various growth- and

developmental responses. For example, Arabidopsis PLC2 and Torenia fournieri PLC1 are involved in

auxin modulated-reproductive development (Song et al., 2008; Li et al., 2015; Di Fino et al., 2017),

while petunia PLC1 and tobacco PLC3 regulate tip growth in pollen tubes (Dowd et al., 2006; Helling

et al., 2006). In Physcomitrella patens, PLC1 has been shown to play a role in the cytokinin- and

gravity response (Repp et al., 2004).

The Arabidopsis genome encodes 9 PLC genes (Mueller-roeber and Pical, 2002). So far, no

developmental disorders other than the reproduction mentioned above have been reported for

Arabidopsis KO mutants, presumably due to genetic redundancy. Here, we show that PLC3 plays

various, yet subtle roles in plant development and ABA signaling, and that overexpression increases

drought tolerance.

Page 29: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

27

RESULTS

Loss of PLC3 affects root development

To investigate PLC3 function, we isolated two homozygous T-DNA insertion mutants of PLC3, plc3-2

(SALK_037453) and plc3-3 (SALK_054406) with T-DNA inserts in exon 3 and intron 3 located in the

X-domain, respectively (Fig. 1a). Reduction of PLC3 expression was verified by both reverse

transcription (RT)-PCR (Fig. 1b) and Q-PCR (Fig. 1c).

Growing seedlings on ½MS plates, a subtle difference in root system architecture between wild

type and plc3 mutants was found. Both PLC3 deficient lines exhibited slightly shorter primary roots

(~5%) and developed less (~15%) lateral roots than wild type. The mutants also showed low lateral root

density (~10%) (Figs. 1d, 1e). Figure 1. Effect of PLC3 knockout on seedling root development (a) Representation of PLC3 gene and T-DNA insertion positions of plc3-2 and plc3-3. Filled boxed and lines represent exons and introns, respectively. Open, grey boxes and triangle represent untranslated region, X- and Y- domains and T-DNA insertions, respectively. (b) Confirmation of reduction of PLC3 expression in plc3 lines by cDNA amplification. PLC3-specific primers were used to detect PLC3 mRNA by RT-PCR. TUBULINα4 (TUB) was used as loading control. (c) PLC3 expression level in wild-type, plc3-2 and plc3-3 lines measured by Q-PCR. Relative expression is based on comparison to expression of the SAND gene. Values are means ± SD (n = 3) for one representative experiment. (d) Seedling morphology of wild-type and plc3 mutants. Seeds were germinated on ½ MS with 0.5% sucrose for 4 days, then transferred to ½MS plates without sucrose. Photographs were taken 12 days after germination (DAG). (e) Primary root (PR) length, lateral root (LR) number and lateral root (LR) density at 12 DAG. Values are means ± SE of three independent experiments (n>20). Asterisk (*) indicate significance at P<0.05 compared to wt based on Student’s t test.

Page 30: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

28

Expression of PLC3 during plant development

Changes in transcript level of PLC3 in various organs and upon induction by hormones or abiotic stress

have been reported using quantitative RT-PCR (Hunt et al., 2004; Tasma et al., 2008). To investigate

this further, we generated β-glucuronidase- (GUS-) and YFP- reporter lines, driven by a 2.4 kb PLC3

promoter fragment (PLC3pro:GUS-YFP). As shown in Figure 2, GUS activity was mainly found in the

vasculature, throughout all stages of development, including seedling, cotyledons, leaves, hypocotyl,

flower, incl. stamen and style, and during seed development. Interestingly, the base of the trichomes

revealed GUS expression (Fig. 2j-k), which again appears to be linked to the vascular system (Fig. 2k).

The expression in the main root was not homogenous. At the distal side of the root maturation zone, the

GUS expression tended to be 'segmented', while in the apical maturation zone it was continuous, and

expression stopped near the transition zone (Fig. 2c-i). Interestingly, lateral roots always emerged from

a segment, but not every segment led to a lateral root (Fig. 2c, d). To search for a potential correlation,

we analysed GUS expression in seedlings grown on a plate that was positioned in a 45º angle, which

forces lateral roots to emerge at the curved sites. Under these conditions, less segments were found but

all lateral roots did emerge from a segment (Supplemental Fig. S1). A similar segmented pattern was

observed in tertiary root formation (Fig. 2e). Together these results confirm that PLC3 is expressed

throughout the plant, but the expression is mainly restricted to the vasculature (Hunt et al., 2004; Tasma

et al., 2008).

To obtain more detailed information about the PLC3 expression within the vasculature, optical

cross- and longitudinal sections were made by confocal microscopy (Supplemental Fig. S2). From this

data, YFP expression appeared to be localized to the phloem and this correlated with data from the eFP

browser, where PLC3 seems to be predominantly expressed in the companion cells

(http://www.bar.utoronto.ca/efp/cgi-bin/efpWeb.cgi).

Page 31: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

29

Figure 2. PLC3pro:GUS-YFP expression analyses in Arabidopis seedlings and mature tissues. GUS activity was present in the vasculature of 2-d old- (a) and 10-d old seedlings including, shoot and root (b-i). GUS staining was also observed in vascular tissue of mature plants (j), trichome base (indicated by arrows) (j, k), hydathodes (j), silique (l), developing seed chalaza (m) and different parts of the flower (n), including style, filament, receptacle and pedicel (indicated by arrows).

Page 32: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

30

Analysis of PPI-, PA- and IP6 levels in Arabidopsis seedlings

To determine whether loss-of-PLC3 caused changes in the level of PLC substrates (i.e. PIP and PIP2) or

products (i.e IP6 or other IPPs), various isotope labelling studies were performed. Since PLC-generated

DAG can be rapidly converted into PA(Munnik et al., 1998; Ruelland et al., 2002; Arisz et al., 2009,

2013), we also measured PA levels. Lipids were analyzed by labelling five-day-old seedlings O/N with 32Pi. As shown in Figures 3a and 3b, wt and plc3 seedlings were found to contain similar amounts of

PIP2, PIP and PA. For the IPP analyses, 3H-Inositol labelling of seedlings and HPLC analyses were

performed, but also here, no significant changes in the level of IP6 or lower IPPs were found. On closer

inspection of the extremely low levels of the pyrophosphate-IPPs (PP-IPPs), i.e IP7 and IP8, we did

observe some differences, however. Both plc3 mutants were found to contain ~30% less IP8 than wt

(Fig. 3d). While the latter analyses were performed on seedlings of 18-days old (11d + 7d of labelling),

we also tested younger seedlings with shorter labelling times (4d old + 4d labelling). Again no

differences in IPPs were found but in this case, plc3 mutants were found to contain ~30-40% less IP7

(see Supplemental Fig. S4).

Figure 3. PPI-, PA- and IPP levels in wild type- and plc3-mutant seedlings. For lipid analyses, five-days old seedlings were labelled with 32PO4

3- overnight and the next day their lipids were extracted and separated by TLC. (a) Autoradiograph of a typical experiment. Each lane represents the extract of 3 seedlings. (b) Quantification of 32P-labeled PIP2-, PIP- and PA levels in wild-type and plc3 mutants. Values are calculated as the percentage of total 32P-labeled phospholipids and represented as means ± SD (n=3). This experiment was repeated twice with similar results (c) Inositol polyphosphates (IPP) levels in wild type and plc3-mutant seedlings. (d) IP8 in wild type and plc3-mutant seedlings. Eleven-day old seedlings were labelled with 3H-myo-inositol for 7 days, after which the IPPs were extracted and resolved by HPLC-SAX chromatography. Fractions were collected every minute, and the radioactivity was determined by liquid scintillation counting. The quantities are expressed as percentage of total. Data shown are the means ± SD (n=10) of one representative experiment. Similar results were obtained in an independent experiment.

Page 33: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

31

Loss of PLC3 affects seed germination

We noticed another subtle phenotype: plc3 mutants always germinated slower on agar plates. Normally,

we imbibed seeds on ½MS plates in the dark at 4 °C for 48 h, after which they are transferred to the

light. After 24 h in the light, plc3-2 and plc3-3 mutants had germinated 54% and 60% less than wt,

respectively, and after 28 h this was 17% and 34% (Fig. 4a). These results also primed us to check the

pPLC3::GUS expression during seed germination. As shown in Fig. 4b, GUS activity was found during

testa rupture and radical emergence in the embryo cotyledons and shoot apical meristem, confirming a

role for PLC3 in seed germination.

To investigate whether this was due to changes in sugar quantity or composition, soluble

carbohydrates in plc3 mutants and wt seeds were measured. As shown in Fig. 4C, small changes in

trehalose and stachyose were observed, but these differences were not found to be significant (P<0.05).

Data from three different seed batches was analysed.

Figure 4. PLC3 is expressed in germinating seeds and plc3 mutants exhibit a delayed germination rate and soluble carbohydrates content in seeds of wild-type and plc3 mutants (a) Seed-germination rate was determined by radical emergence and scored in wild-type and plc3 mutants. Seeds were stratified on ½MS with 0.5% sucrose plates at 4°C for 2 days and allowed to germinate at 22°C. Data shown are the means ± SD for one representative experiment (n=55 seeds for each genotype). Asterisks (*) mark that plc3 values are significantly different from wild-type based on Student’s t-test (P< 0.05). (b) GUS activity was determined in embryo cotyledons during seed germination from testa rupture until radical emergence (20-28hrs after transfer from 4°C to 22°C). These experiment were repeated twice with similar outcome. (c) Soluble carbohydrates were extracted from dry seeds and analyzed by Dionex HPLC. Sugar quantities were corrected by means of an internal standard and transformed to µg of sugar per mg of dry material. Values are the means of triplicates ± SE of three independent seed batches.

Page 34: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

32

Decreased sensitivity to ABA in plc3 mutants

Since germination could involve ABA sensitivity, this was analysed in more detail. Wild type and plc3

mutants were germinated on ½MS plates containing different concentrations of ABA (0, 1 and 2 µM).

In the absence of ABA, plc3 mutants germinated slower than wt as described above (Fig. 4a). However,

in the presence of ABA, plc3 mutants were found to germinate faster than wt (Fig. 5a). For example,

after 40 h at 1 µM ABA, 12.5% of plc3-2 and 10.5% of plc3-3 seeds had germinated whereas only 2.5%

of the wt seeds. In time, these differences remained or even increased (Fig. 5a, left panel). At higher

ABA concentrations (2 µM), seed germination was inhibited more strongly, but again, both plc3

mutants had higher germination rates than wt (Fig. 5a, right panel). These results indicated that plc3

mutants are less sensitive to ABA during germination.

Since ABA is also involved in guard-cell closure (Munemasa et al., 2015), and since antisense-

PLC expression in Nicotiana tabacum had been shown to reduce the stomatal-closure response upon

ABA treatment (Hunt et al., 2003), we decided to investigate this further for PLC3. As shown in Fig. 5b,

GUS activity of the pPLC3::GUS line indicated that PLC3 was indeed active in guard cells. To

investigate its putative involvement in ABA-induced stomatal closure, epidermal peels of wt and plc3

mutants were treated with different concentrations of ABA (i.e. 0, 0.1, 1 and 10 µM). In the absence of

ABA, no significant differences in the stomatal aperture between wild type and plc3 mutants were

found (Fig. 5c). However, with increasing concentrations of ABA, both plc3-2 and plc3-3 showed

reduced closure responses compared to wt. In summary, these results indicate that loss-of-PLC3 leads

to decreased sensitivity to ABA, in both germinating seeds and guard cells. Figure 5. Decreased ABA sensitivity of plc3 mutants in seed germination and stomatal movement. (a) Seeds germination rate of wild-type and plc3 mutants in the presence of 1 or 2 µM ABA. Seeds were germinated on ½MS with 0.5% sucrose plates with different concentration of ABA at 22 °C after 2 days of stratification at 4 °C. Germination is defined by radical emergence and was scored at the indicated times. Data shown are the means ± SD from one representative experiment of at least 3 experiments (n=55 seeds for each genotype). Asterisks (*) mark that plc3 value are significantly different from wild-type based on Student’s t-test (P< 0.05). This experiment was repeated 3 times with similar results. (b) PLC3pro:GUS-YFP expression in guard cells, using epidermal leaf peels of 3 weeks-old Arabidopsis plants . (c) Effect of ABA on stomatal aperture in wild-type and plc3-2 (left) or plc3-3 (right). Epidermal strips were incubated in opening buffer

with light for 3 h until stomata were fully open. Strips were then treated with different concentration of ABA for 90 mins, after which stomata were digitized and the aperture width measured. Data was analyzed by 2-way ANOVA. Statistically significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method). Values are means ± SE of at least three independent experiments (n ≥ 100).

Page 35: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

33

ABA triggers PIP2 formation in germinating seeds, seedlings and guard cells

Since loss-of-PLC3 increased the ABA insensitivity during seed germination and stomatal closure, we

decided to analyse the phospholipid levels in more detail in these tissues and to analyse the effect of

ABA. First the effect in germinating seeds was analysed (Fig. 6). In the absence of ABA, no major

differences in the structural phospholipids (not shown) or PIP and PA (Fig. 6a, b) were found between

wild type and plc3 mutants. However, PIP2 levels in plc3 mutants were significantly higher than in wild

type seeds (i.e. 23% and 22% for plc3-2 and plc3-3, respectively). Upon ABA treatment, no major

changes in PIP or PA were found (Fig. 6a, b). However, while a significant increase in PIP2 (27%) was

found in wild-type seedlings upon ABA treatment, both plc3 mutants lacked this response (Fig. 6b). Figure 6. PPI- and PA levels in germinating seeds and effect of ABA in wild type and plc3 mutants.

Seeds of wild-type and plc3 mutants were pre-germinated on ½ MS with 0.5% sucrose plates until testa ruptured, then labelled with 32PO4

3- for 24 h, after which they were treated for 2 h with buffer ± 100 µM ABA. (a) Autoradiograph of a typical experiment is shown, each lane representing the extract of ± 200 seeds. (b) Quantification of the 32P-levels of PIP2, PIP and PA. Three independent experiments were performed; data shown are means ± SD (n=3) from one representative experiment. Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method).

Next, we analysed the response in guard cells (Fig. 7). For this we used epidermal-leaf peels

that are enriched in guard cells (Munnik & Laxalt, 2013) from wt and mutant leaves, and labelled these

with 32Pi for 3 hrs. Longer labelling times were found to have a negative effect on the viability of the

guard cells (not shown). Similar to what we found for germinating seeds, plc3 mutants contained

slightly higher PIP2 levels (Fig. 7), while PIP and PA and the major structural phospholipids levels

remained unchanged (Fig. 7; not shown). With ABA (15 min treatment), again a significant increase of

PIP2 was observed for wild type, but not in the plc3 mutants (Fig. 7). A small decrease in PIP and

increase in PA was found, but these changes were not statistically significant (P<0.05). No changes in

the structural phospholipids were found. We also tested the effect of ABA after 2, 5, 30 and 60 min, but

no clear differences were found before or after treatment in both wild type and plc3 mutants.

Similarly, the effect of ABA on seedlings was analysed. Time-course analyses in wt seedlings

revealed an increase in PIP2, which was found to be significant after ~30-60 min of treatment

(Supplemental Fig. S6). PIP, PA and structural-phospholipids levels remained the same during that

period (Supplemental Fig. S6 b-d). Testing plc3 mutants after 1 h of ABA treatment showed in this case

Page 36: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

34

an accumulation of PIP2, which was not significantly different from wild type (Supplemental Fig. S6f).

PIP, PA and structural-phospholipid levels remained similar. Figure 7. PPI and PA levels in leaf peels and effect of ABA in wild-type and plc3

mutants. Three-week-old rosette leaf peels from wild-type and plc3 mutants were 32Pi-labeled for 3h and then treated in buffer ± 100 µM ABA for 15 min. Lipids were then extracted and separated by TLC. Radioactivity levels in PIP2, PIP and PA were determined as percentage of total phospholipids. Three independent experiments were performed. Data shown are the means ± SD (n=3) from one representative experiment. Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method).

Over expression of PLC3 enhances drought tolerance

Overexpression of PLC has been shown to promote drought tolerance in maize, canola and tobacco

(Wang et al., 2008; Georges et al., 2009; Tripathy et al., 2011). It is unknown whether specific PLCs

were chosen for this or whether any PLC can achieve this. For Arabidopsis, this is unknown either.

Hence, we generated transgenic plants overexpressing PLC3 under the control of the UBQ10 promoter.

Transgenic plants were selected from T0 to T3 and independent homozygous lines were obtained.

Two homozygous lines, PLC3-OE9 and PLC3-OE16, were selected for further studies,

overexpressing PLC3 48-fold and 20-fold, respectively (Fig. 8a). No obvious phenotypes were

observed comparing wild type and the PLC-OE lines on either agar plates or soil. In soil, four weeks

old plants from PLC3-OE lines were found to be more drought tolerant than wt (Fig. 8b) and to show

significantly higher survival rates (Fig. 8b, c).

During drought stress, the shoot FW of wild type decreased by ~21%, which was less in the

PLC3-OE9 and PLC3-OE16 lines (17% and 12%, respectively)(Fig. 8d). Their DW, however, was

higher with or without drought (Fig. 8e). PLC3-overexpression lines also lost less water when water

loss of detached 4-week-old rosettes were compared (Fig. 8f).

ABA synthesis is stimulated by dehydration stress and known to induce stomatal closure to

reduce water loss (Sean et al., 2010). In the absence of ABA, stomatal aperture of PLC3-OEs was

found to be strongly reduced by ~30%. Upon ABA treatment (0.1 µM), stomata closed rapidly for all

Page 37: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

35

genotypes, but the aperture of the PLC3-OEs was still significantly smaller than wild type. Above 1 µM,

this difference was not observed (Fig. 8g).

Figure 8. Overexpression of PLC3 enhances drought tolerance. (a) PLC3 expression levels in wild-type and two homozygous PLC3 overexpression lines, PLC3-OE9 and PLC3-OE16 as measured by Q-PCR and based on the expression of the SAND reference gene. Values are means ± SD (n = 3) for one experiment. (b) Phenotype of wild type- and PLC3-OE plants. Four-week old soil-grown plants were exposed to drought stress by water withholding for 2 weeks. (c) Survival rates were determined by counting the visible, green plants after re-watering. (d, e) Fresh- and dry weights were determined from shoots under control and drought (1 week water withholding) conditions. (f) Water loss of detached rosette. Water loss was measured at indicated time points and expressed as a percentage of the initial fresh weight. Values are means ± SD for one representative experiment (n=36). (g) ABA-induced stomatal closure in wild-type, PLC3 OE9 (left), PLC3 OE16 (right) plants. Epidermal peels from 3-weeks old plants incubated in opening buffer with light for 3 h until stomata were fully open. Peels were then treated with different concentration of ABA for 90 mins, after which stomata were digitized and the aperture width measured. Values are means ± SE of at least three independents (n >100). Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method).

Page 38: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

36

Higher accumulation of PIP2 in PLC3 overexpressing plants under osmotic stress

To determine whether overexpression of PLC3 caused any changes in PPI- and/or PA levels, 32P-labelling experiments were performed on seedlings, and the effect of 600 mM sorbitol was tested to mimic water stress. As shown in Fig. 9, no major differences between wild type and PLC3-OE lines were found under control conditions. However, upon sorbitol treatment, a much stronger PIP2 response was observed in the OE lines. In wt, PIP2 levels increased by about 300%, while in the OE lines a ~600% increase was witnessed. The PA response appeared slightly higher (200% vs 300%) but this was not statistically significant. The osmotic stress-induced decrease in PIP was similar to wild type (Fig. 9b). These results suggest that PLC3-OE lines are capable of enhancing the PIP2 response under osmotic stress.

Figure 9. Osmotic stress triggers PIP2 and PA responses in wild type- and PLC3-OE seedlings. Six-day-old seedlings were 32P-labeled for 3h and then treated with buffer ± 600 mM sorbitol for another 30 min. Extracted lipids were analyzed by TLC and quantified through phosphoimaging. (a) Typical TLC profile with each lane representing the extract of 3 seedlings. (b) 32P-levels in PIP2, PIP and PA of wild-type and PLC3-OE lines #9 and #16 with anbd without sorbitol. Three independent experiments were performed. Data shown are the means ± SE (n=3) from three independent experiments. Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05).

Page 39: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

37

DISCUSSION

In this paper, new roles for PLC in plant stress and development are described. Using loss-of-function

mutants in Arabidopsis, we found that AtPLC3 is involved in seed germination, root development,

stomatal movement and ABA signaling, whereas overexpression of PLC3 enhanced the plant's

tolerance to drought stress. While these findings underline the importance of PLC signaling in plant

stress and development, we still know very little of how this is achieved at the molecular level.

Theoretically, there are several possibilities. First of all, PLC can produce DAG and IP2 or IP3,

(depending on whether PIP or PIP2 is used as a PLC substrate), and while plants lack the classical

targets of the mammalian paradigm (i.e. IP3 receptor, PKC), it is likely that their phosphorylated

products, i.e. PA and higher IPPs (incl. PP-IPPs) fulfill this second-messenger role in plants. Various

biological processes have already been linked to these molecules, and several protein targets involved

in signal transduction and metabolism have been identified too (see below). In guard cells, IP6 has been

show to release Ca2+ (Lemtiri-Chlieh et al., 2000, 2003), so the PLC system in plants could potentially

do that.

In non-stressed cells, it is more likely that PLC will hydrolyse PIP than PIP2. The concentration

of the latter in plasma membranes is extremely low in plants (30-100 fold lower than mammalians;

Munnik et al., 1994; Meijer and Munnik, 2003), while PI4P concentrations appear comparable to those

found in mammalian cells (Munnik, 2014). Moreover, in order to make IP6 out of IP2 or IP3 involves the

same two inositolpolyphosphate kinases (IPKs). IPK2 is an inositol multiphosphate kinase that can

phosphorylate the 3-, 5-, and 6- position of the inositol ring to produce IP5. IPK1 specifically

phosphorylates IP5 at the 2-position to produce IP6. VIH2 is a recently discovered IPK that is

responsible for the production of the pyrophosphorylated IPPs, i.e. IP8. Like in animal- and yeast cells,

these compounds are emerging as important signaling molecules in plants (York, 2006; Michell, 2008;

Burton et al., 2009; Shears, 2009; Desai et al., 2014; Laha et al., 2015).

Another function of PLC could be to attenuate PIP2 signalling. While the concentration of this

lipid is extremely low under control conditions, PIP2 is readily produced in response to certain

hormones or stress signals, where it is suggested to fulfill a second messenger itself, regulating various

aspects of plant growth, development, and stress signaling (Gillaspy, 2013; Rodriguez-Villalon et al.,

2015; Heilmann, 2016; Zarza et al., unpublished). Potential targets include proteins involved in ion

transport (e.g. K+ channels), membrane trafficking (endo/exocytosis, e.g. clathrin and Exo70) and

cytoskeletal organization (e.g. small G-protein, Rop) (Gillaspy, 2013; Munnik, 2014; Heilmann, 2016).

In vitro, plant PLCs hydrolyze PI4P and PI(4,5)P2 equally well (Munnik, 2014), and since PI4P is

emerging as a lipid second messenger too (Vermeer et al., 2009; Munnik and Nielsen, 2011; Heilmann,

2016), under certain conditions and in particular cells, PLC could also function as attenuator of PI4P

signalling. As far as we know, PLCs are unable to use D3-phosphorylated inositol lipids as a substrate

[i.e. PI3P and PI(3,5)P2] or PI5P (Munnik, 2014). Whether the newly linked-PLC3 functions observed

Page 40: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

38

here, reflect PLC's role as second messenger producer or -attenuator (or both), needs to be established.

Below, a broader perspective of our results is given and some potential molecular mechanisms

discussed.

Role for PLC3 in seed germination

Promoter-GUS expression in germinating seeds (Fig. 4b), together with the delayed germination

phenotype of both plc3 mutants (Fig. 4a), indicates a role for PLC3 in seed germination. Since ABA is

known to play an important role in this (Nambara et al., 2010; Nakashima & Yamaguchi-Shinozaki,

2013), we investigated whether the delayed germination of the plc3 seeds was caused by

hypersensitivity to ABA. Surprisingly, plc3 mutants were found to be less sensitive to ABA (Fig. 5a).

Such results are in agreement with Sanchez & Chua (2001), who found that the ABA sensitivity of seed

germination and downstream-gene expression was lost when PLC1 was silenced in Arabidopsis. Guard

cells of plc3 mutants were also found to be less sensitive to ABA, which could point to a more general

role for PLC3 in ABA signaling (see below; Fig. 5). At least, the above results indicate that the basal,

delayed germination rate in plc3 mutants is unlikely to be caused by ABA hypersensitivity.

Gibberellin (GA) is another important hormone involved in seed germination (Yamaguchi and

Kamiya, 2001). In contrast to ABA, GA promotes seed germination, and there is data to suggest that

this could involve PLC signaling too, i.e. induced PLC expression, changes in PPIs, and increased IP3

levels (Murthy et al., 1989; Chen et al., 1997; Kashem et al., 2000; Villasuso et al., 2003; Fleet et al.,

2009; Luo et al., 2012). We tested whether plc3 mutants were affected in GA responsiveness by

comparing their germination to wt, with and without 1 mM GA (Supplemental Fig. S7). Although the

initial germination rate of plc3-2 and plc3-3 mutants with GA was still slower than wild type

(Supplemental Fig. S7a and b), after 24h, the fold-increase in plc3 mutants was around 20% more than

in wild type. This was found in three independent experiments, despite the fact that the difference was

not statistically significant (Supplemental Fig. S7c). These results point to a possible role for PLC3 in

GA signaling, even though the hypersensitivity of the plc3 mutants to GA does not explain their slower

germination phenotype. One hypothesis could be that PLC has a positive effect on GA levels and

therefor lower GA levels in the mutant, which could delay germination. By adding external GA, some

of this inhibition might then be released, which would become visible as hypersensitivity.

Results from our phospholipid measurements revealed that germinating plc3 seeds contained

significantly higher levels of PIP2 (Fig. 6), which would be consistent with a loss of PLC3 that would

normally hydrolyze this lipid to produce IP3. Unfortunately, the latter is very difficult to measure

because seeds contain tiny amounts of IP3 and huge amounts of IP6, and are also extremely difficult to

label with 3H-inositol (Stevenson-Paulik et al., 2005). Seeds hardly take-up this label, and this is

probably also the reason why young seedlings require relatively long labelling times (i.e. 4-11 days vs

hrs with 32Pi; see Methods). Seeds typically store high amounts of IP6 during their development, where

it is used as supply of phosphate (e.g. for DNA, ATP, membranes and sugars) and inositol (IPPs, PPIs,

Page 41: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

39

precursor of cell wall sugars) (Munnik and Nielsen, 2011; Valluru and Van den Ende, 2011) when the

seed germinates and the embryo develops into a seedling while growing in the dark. This so-called

'storage' IP6 is easily confused with 'signalling' IP6 (Munnik and Vermeer, 2010), but has totally

different functions and is probably even differentially localized within cells or tissues. It is difficult, if

not impossible, to distinguish between these two IP6 sources at the moment (Munnik and Vermeer,

2010; Gillaspy, 2011; Munnik and Nielsen, 2011). During seed germination, IP6 is rapidly broken down

to IP3 (Luo et al., 2012) and this could be an alternative explanation for what was assumed to be PLC-

generated IP3 (Murthy et al., 1989; Chen et al., 1997; Kashem et al., 2000; Villasuso et al., 2003; Fleet

et al., 2009; Luo et al., 2012).

Another set of molecules related to inositol metabolism are Raffinose Family Oligosaccharides

(RFOs), which serve as desiccation protectant in seeds, as transport sugar in the phloem and as storage

sugar in various tissues (Sengupta et al., 2015). In Arabidopsis seeds, RFOs are required for the rapid

germination in the dark (Gangl and Tenhaken, 2016). RFOs are sucrose derivatives to which a

galactosyl unit is attached via galactinol (Gol). The latter is produced via UDP-galactose and myo-

inositol by the enzyme, galactinol synthase (GolS). To make RFOs, free myo-inositol is required and

this is predominantly formed through cyclization of glycolytic glucose 6-phosphate (G6P) into inositol-

3-phosphate (Ins3P) by myo-inositol-3-phosphate synthase (MIPS) and subsequent dephosphorylation

by inositol mono- phosphatase (InsPase). Theoretically, however, inositol could also be generated via

dephosphorylation of PLC-generated IPPs (Munnik and Vermeer, 2010). We analysed the soluble

carbohydrate composition in seeds and in the phloem sap but found no significant differences between

wt and plc3 mutants. Of course, changes could be very local so it is possible these differences remain

unobserved.

During seed development, PLC3 was expressed at the chalaza, the non-micropylar end of the

seed, likely the chalaza endosperm and/or seed coat (Fig. 2m). Nutrients from the mother plant are

transported via the vascular tissue through the chalaza into the nucellus. The vascular and chalaza

expression of PLC3 might be necessary for nutrient transportation. Alternatively, PLC might be

involved in the production of IP6 for the storage of essential minerals. Developing seeds store these

minerals in three locations, i.e. in the protein storage vacuoles of the embryo, and transiently in the

endoplasmic reticulum (ER) and vacuolar compartments of the chalaza endosperm. X-ray analysis and

enzyme treatments have suggested that these minerals are stored as IP6-salts with distinct cation (Mg,

Mn, Zn, K, and Ca) composition per compartment (Otegui et al., 2002). As such, loss of PLC3 may

affect embryo development, germination and even plant development.

Role for PLC3 in lateral root formation and auxin signaling

Loss-of-function PLC3 mutants displayed shorter primary roots and fewer lateral roots (Fig. 1). The

latter was due to less initiation sites, not development (data not shown). Promoter-GUS analyses

indicated a very typical, segmented, PLC3-expression pattern at the lateral root-emerging site, whereby

Page 42: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

40

lateral roots always emerged from a segment, but not every segment resulted in a lateral root. Normally,

we grow our plates vertically in an angle of 70°. By tilting the agar plate more horizontally (45°), roots

start to wiggle more and tend to grow a lateral root at every bend. Using the latter setup for GUS

analyses, revealed that lateral roots only emerged from these segments, but that the number of segments

was drastically reduced, which was almost 1:1 with the lateral roots whereas with the 70° setup,

typically two or three segments were found near the lateral root. These results may indicate that PLC3

expression is required just before the lateral root is initiated, and that the primary root at the 70° setup is

less determined as to where and when it will produce the lateral root compared to the 45° setup where

this decision is forced at the bending sites (Ditengou et al., 2008). That the phenotype is quite mild may

indicate that redundant PLCs are involved. Using the eFP browser data, we found that in addition to

PLC3, expression of PLC2, PLC5 and PLC7 is also present in the phloem and/or companion cells.

Interestingly, the initiation of tertiary roots revealed a very similar GUS-expression pattern,

showing segments in the lateral roots from which tertiary roots emerged. Root growth and -branching

are main events of root development. Root growth requires cell proliferation and division in the

meristematic zone, and cell expansion in the elongation zone. Lateral root formation involves three

major steps, which are initiation, primordial organogenesis, and emergence (Benková and Bielach,

2010). Auxin has been shown to be required in both primary root growth and lateral root formation

(Péret, De Rybel, et al., 2009; Péret, Larrieu, et al., 2009; Benková and Bielach, 2010). The signaling

pathway of auxin perception is well characterized. Auxin promotes the degradation of the

transcriptional repressor Aux/IAA, resulting in massive auxin responsive-gene expression. The auxin

receptor, TIR1 is a F-box protein and complex with SCF (ubiquitin protein ligase), which promotes

ubiquitin-dependent proteolysis of Aux/IAAs (Kepinski and Leyser, 2005). Interestingly, IP6 has

recently been found in the crystal structure of TIR1 where it is thought to be required for auxin binding

and TIR1 function (Tan et al., 2007). Where the IP6 is coming from is unknown, but potentially this

could be formed though PLC3-generated IP2- or IP3 formation at the above mentioned 'segments' and

subsequent phosphorylation into IP6. Less PLC-generated IP6 in plc3 mutants would then lead to less

auxin responsiveness during root development (Fig.10A). Redundant PLCs are likely to take over most

of PLC3's function(s) though.

In contrast to the germinating seeds, no differences in PPI- or PA levels were found in 32Pi-

prelabeled seedlings (Fig. 3b). However, since PLC3 is expressed in a limited number of cells

(especially in the phloem companion cells), analyses of whole seedlings might dilute any difference.

We also did not find differences in IP6 either. Tiny differences in the pyro-IPP levels were found,

however, with lower levels of IP7 and IP8 in the plc3 mutants depending on their age (Fig. 3c;

supplemental Fig. 4). Both IP7 and IP8 are implicated as novel signalling molecules (Laha et al., 2015,

2016) for which there is already lots of evidence in yeast and animals (York, 2006; Michell, 2008;

Burton et al., 2009; Shears, 2009). That PLC3 could be involved in generating such signaling molecules

is exciting, but requires further analysis. Similarly, on the role of IP6. In guard cells, IP6 may be

Page 43: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

41

responsible for the release of intracellular Ca2+ and since the latter is also important for auxin signaling

(Zhang et al., 2011), this line of research is worth perusing too. The main bottleneck, still, after the first

discoveries over 25 years ago (Blatt et al., 1990; Gilroy et al., 1990; Allen and Sanders, 1994), is the

identification of a genuine IP6- (or other IPP-) gated channel (Lemtiri-Chlieh et al., 2000, 2003).

As discussed above, PLC3 could also be involved in inositol-based RFO metabolism. Since RFOs

are important for carbohydrate transport- and storage, potentially they could be involved in loading

sucrose to sink organs, e.g in lateral root (Van den Ende, 2013; Sengupta et al., 2015; Gangl and

Tenhaken, 2016). Analysing the sugar composition of the phloem sap revealed increased amounts of

sucrose in the plc3 mutants and slightly decreased levels of myo-inositol levels, although the latter

differences were not significant (Supplemental Fig. 3b). If sucrose is not properly transported to, or into,

the lateral root via a PLC depended-RFO pathway, then sucrose levels could indeed be higher in the

phloem sap and theoretically could affect root growth and lateral root formation.

Role for PLC3 in stomatal closure and ABA signaling

PLC has been linked to ABA signaling in several reports (Hirayama et al., 1995; Sanchez and Chua,

2001; Hunt et al., 2003; Sui et al., 2008). In Arabidopsis, a number of PLC genes are induced upon

ABA treatment ( Hunt et al., 2004; Lin et al., 2004; Tasma et al., 2008). We tested our plc3 mutants for

their response to exogenous ABA with respect to inhibition of seed germination and ABA-mediated

stomatal closure along with wild type. Results showed that down-regulation of PLC3 decreased the

ABA sensitivity for both responses (Fig. 5). Similar results have been found for germinating seeds of

PLC1-silenced Arabidopsis plants (Sanchez and Chua, 2001), and in guard cells of PLC-silenced

tobacco plants(Hunt et al., 2003; Mills et al., 2004).

We also tested the effect of ABA on the turnover of phospholipids in germinating seeds and

guard cell-enriched leaf peels and found an increase of PIP2 in both tissues after ABA stimulation,

which was strongly reduced or even lost in the plc3 mutants (Figs. 6 and 7). We speculate that PLC3 is

activated by ABA, thereby increasing the hydrolysis of PIP2 and the subsequent replenishment of the

pool by PIPK. Increased turnover of PIP2 is ideally reflected by this type of 32P-labeling experiment

(Munnik et al., 1994; Munnik and Zarza, 2013).

In Figure 10B, a model is presented of how PLC3 and PIP2 could be involved in regulating

stomatal movement. The latter is controlled by changes in turgor of the surrounding guard cells. During

stomatal opening and -closing, ion channels and cytosolic Ca2+ oscillations play key roles in this process,

and these transporters need to be tightly regulated. Over the years, many genes and proteins have been

implicated (Ward et al., 2009; Roelfsema et al., 2012; Munemasa et al., 2015; Assmann and Jegla,

2016). Here, we would like to draw the attention of how PPIs and IPPs could regulate stomatal

movement. During light induced-stomatal opening, the H+-ATPase pump is activated, which causes

hyperpolarization of the plasma membrane and the opening of the voltage-gated K+‐influx channel,

KAT1. The subsequent influx of K+ lowers the water potential and drives the net influx of water into

Page 44: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

42

the guard cell (Dietrich et al., 2001; Schroeder et al., 2001; Ward et al., 2009; Roelfsema et al., 2012).

Meanwhile, ABA-INSENSITIVE 1 (ABI1), a type 2C protein phosphatase (PP2C), inhibits SNF1-

Related kinase (SnRK2, i.e. OST1) activity, which in its active form activates the slow anion channel 1

(SLAC1). PIP5K4 is essential for stomatal opening (Lee et al., 2007). This lipid kinase generates PIP2,

which has been shown to inhibit SLAC1 (Lee et al., 2007) and the K+-efflux channel (Ma et al., 2009)

co-facilitating the low water potential, the subsequent influx of water, and the opening of stomata.

Upon ABA, the PYR/PYL-receptor dimer dissociates and forms PYR‐ or PYL-ABA

complexes(Ma et al., 2009; Park et al., 2014) that bind PP2C (Melcher et al., 2009; Miyazono et al.,

2009; Nishimura, 2009; Santiago et al., 2009; Yin et al., 2009), which can then no longer inhibit the

protein kinase activity of SnRK2/OST1 (Hirayama and Umezawa, 2010). As a consequence, OST1 can

now auto-phosphorylate itself (Soon et al., 2012) and activate SLAC1 (Kulik et al., 2011), which

results in a decrease of intracellular Cl-. Activated PLC, hydrolyses PIP2, thereby releasing the

inhibition of SLAC1 and the K+-efflux channel, but also generates increased amounts of IP3 and IP6

through IPK1 and IPK2. The IP6 can release Ca2+ from internal stores (Lemtiri-Chlieh et al., 2000, 2003;

Munnik, 2014), which inhibits the K+‐influx channel (Lemtiri-Chlieh et al., 2000) and co-activates

SLAC1 (Siegel et al., 2009). Together these activities cause the net efflux of K+ and Cl-, which

decreases the water potential and causes water to leave the guard cells and stomata to close. PLC3

seems to be one of the PLC genes involved in this process. Although plc3 mutants still respond to ABA

by closing their stomata, the response is significantly reduced. How the PLCs involved are activated

still remains elusive. Ca2+ is a potential factor since it stimulates PLC activity in vitro (Munnik et al.,

1998), but this would first require an influx of Ca2+ into cytosol via another pathway. As such, PLC and

PIP2 would act as facilitators in these biophysical events. The redundancy of 9 PLCs and 11 PIPKs may

prevent more clear phenotypes.

Page 45: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

43

A

B

Figure 10. Models for the role of PLC in (A) lateral root formation, and (B) regulating stomatal aperture. Abbreviations: KAT1: voltage-gated K+influx channel; ABI1: ABA-INSENSITIVE 1; PP2C: type 2C protein phosphatase; SnRK2: SNF1-Related kinase; SLAC1: slow anion channel 1

Dark/Drought/ABA Light

Page 46: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

44

Overexpression of PLC3 enhances drought tolerance

Plants cope with drought stress via many different strategies (Zhu, 2002, 2016; Osakabe et al., 2013;

Mickelbart et al., 2015) and recently various lipid signalling pathways have been reported to be

involved in (Munnik and Meijer, 2001; Zhu, 2002; Meijer and Munnik, 2003; Munnik and Vermeer,

2010; Hou et al., 2016). Moreover, overexpression of a PLC in maize, tobacco and canola have been

shown to improve drought tolerance (Wang et al., 2008; Georges et al., 2009; Tripathy et al., 2011). To

investigate whether overexpression of Arabidopsis PLC3 could advance drought tolerance,

homozygous pUBQ10::PLC3 overexpression lines were generated. Under control conditions, these

plants appeared similar to wild type, but upon drought stress the PLC3-OE lines clearly performed

better (Fig. 8b, 8c). They lose less water than wild type, which is likely due to increased number of

closed stomata, as there was no difference in stomata number. The molecular mechanism behind this

may well reflect what we discussed above and what is summarized in Figure 10.

In an attempt to mimic this in seedlings by using sorbitol, we found that PIP2 and PA

accumulated dramatically upon this water stress, and that the accumulation was much stronger in the

PLC3-OE lines (Fig. 9). This may again reflect the increased turnover of PIP2 and phosphorylation of

DAG, which is readily picked-up by these 32P-labelling experiments. We also measured IPP levels with 3H-inositol labeling, but found no differences there between wt and PLC3-OE lines.

Besides guard-cell regulation, PA and PIP2 may accumulate in various other cells and tissues,

since the UBQ10 promoter is constitutively expressed, which is totally different from the endogenous

PLC3 expression in the vasculature. Both lipids have been implicated as second messengers, playing

roles in reorganization of the cytoskeleton, endo- and exocytosis, vesicular trafficking and ion channel

regulation (Stevenson et al., 2000; Martin, 2001; van Leeuwen et al., 2007; Heilmann, 2016), which are

all important cellular events. Therefore, PIP2 and PA are very likely to play an important role in the

plant's response to control water stress. Further unraveling of the molecular mechanisms involved here

requires identification and characterization of some of the main targets of these lipid second messengers,

but also for the IPPs, with IP6 and the PP-IPPs in particular. How PLC is activated remains also an

important issue to address.

Apart from osmotic stress, heat stress also triggers a PIP2 and PA responses (Mishkind et al.,

2009; Horvath et al., 2012). Recently AtPLC3 and AtPLC9 were claimed to be involved in heat stress.

Their T-DNA insertion lines lacked the IP3 response and exhibited decreased thermotolerance while

overexpression lines showed more heat resistance (Zheng et al., 2012; Gao et al., 2014). Problem here

is AtPLC9 is predicted to be “non-active” due to the lack of conserved amino acids in the X-Y domain

that are required for the catalytic activity (Hunt et al., 2004). IP3 was measured with the commercial

displacement assay, this may reflect changes in the flux of other IPPs (Munnik, 2014). However, if the

inactive AtPLC9 could bind PIP2, then its competition with active PLCs might regulate PIP2’s function

as a second messenger.

Page 47: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

45

For many years, PLC/IP3/Ca2+ pathway has been claimed to involved in gravitropism (Perera et

al., 1999, 2006; Stevenson et al., 2000; Boss et al., 2010). We tested our plc3 mutants response to

gravitropism by changing the root growth direction (rotate the plate by 90°; Supplemental Fig. S8 a,b).

The roots of both wild type and plc3 mutants bended around 90°due to gravitropism and no obvious

difference in bending degree between them (Supplemental Fig. S8c). The reason could be PLCs

redundancy. However, most evidence for PLC/IP3/Ca2+ enrollment in gravitropism is based on IP3

measurements using the commercial IP3- displacement kit, which might reflect the flux of other IPPs as

well (Munnik, 2014). This is equally interesting and deserves further investigation.

MATERIALS AND METHODS

Plant material

Arabidopsis thaliana (Columbia-0) T-DNA insertion mutants plc3-2 (SALK_037453) and plc3-3

(SALK_054406) were obtained from SALK collection (signal.salk.edu). Homozygous plants were

identified by PCR in F2 generation by using gene-specific primers in combination with left border

primer LBa (Supplemental Table1).

RNA extraction and RT-PCR

The expression levels of plc3 mutants were confirmed by RT-PCR. Total RNA was extracted with

Trizol reagent (Invitrogen, Carlsbad, CA) as described previously (Pieterse, 1998). RNA (5 µg) was

converted to cDNA using oligo-dT18 primers, dNTPs and SuperScript III Reverse Transcriptase

(Invitrogen) according to the manufacturer’s instructions. PLC3 and TUBLINα4 were PCR amplified

for 40 and 30 cycles respectively with gene specific primers (Supplemental Table1).

Cloning and plant transformation

To generate the PLC3pro:GUSYFP fusion, a 2437 bp PLC3 promoter region was amplified from

genomic DNA using PLC3promHindIIIfw 5’- CCCAAGCTTCAAGTCGCCGAACGAGACATC-3’

and PLC3promNheIrev 5’ -CTGCTCTTCTTCTTCTTACTTGTTAG-3’ and cloned in HindIII/XbaI

digested pJV-GUSYFP. The PLC3pro:GUSYF cassette was transferred to pGreen0179 using NotI.

MultiSite Gateway Three-Fragment Vector Construction Kit (www.lifetechnologies.com) was used to

generate UBQ10pro:PLC3, PLC3 cDNA was cloned into pGreen0125 expression vector. The procedure

followed MultiSite Gateway Three-Fragment Vector Construction Kit user guide

(https://tools.thermofisher.com/content/sfs/manuals/multisite_gateway_man.pdf). Constructs were

transferred into Agrobacterium tumefaciens strain GV3101, which was used to transform wild type

plant by floral dip (Clough and Bent, 1998). Homozygous lines were selected in T3 generation and used

for further experiments.

Page 48: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

46

Real-time quantitative RT-PCR

The primer pairs used for conformation of PLC3 (At4g38530) expression level were: 5’-

TCCAGATTTCTTCGTCAAGATTGGA-3’ (forward) and 5’-

TATAGGAAACCACTGATCGACAGC-3’ (reverse). 1µg total RNA from 10-day-old seedlings was

used for cDNA synthesis as described before. Q-PCR was performed with ABI 7500 Real-Time PCR

System (Applied Biosystem). The relative gene expression was determined by comparative threshold

cycle value. Transcript levels were normalized by the level of SAND (At2g28390; forward primer: 5’-

AAC TCT ATG CAG CAT TTG ATC CAC T-3’, reverse primer: 5’-TGA TTG CAT ATC TTT ATC GCC

ATC-3’) (Han et al., 2013). Three biological replicates and two technical replicates were used for the

values of means and standard deviations.

Histochemical analyses for GUS activity

GUS staining was performed according to the method described by Jefferson et al. (1987) with minor

modifications. Transgenic plants carrying PLC3pro:GUSYFP were grown for indicated times and

specific tissues were taken and incubated in X-Gluc reaction solution containing 1mg/ml 5-bromo-4-

chloro-3 indolyl-β-D-glucuronic acid (X-gluc), 50 mM phosphate buffer pH 7.0 and 0.1% TX-100. The

materials were incubated overnight at 37°C. The next day, the solution was replaced by 70% ethanol to

destain the tissue. Plant tissues were viewed under a stereo microscope (Leica MZFLIII) and

photographed (ThorLabs CCD camera).

Confocal microscopy

Arabidopsis PLC3pro:GUSYFP seedlings were grown 5 days and then transferred to object slides

containing a fixed cover slide, separated by a spacer of approximately 0.32 mm. This allows seedlings

to grow in liquid medium (½ MS and 1% sucrose, pH 5.8) for 1-2 days and could directly be used for

microscopy. Microscopy was performed using a Zeiss LSM 510 CLSM (confocal laser scanning

microscope) (Carl-Zeiss GMBH, Jena, Germany), implemented on an inverted microscope (Axiovert

100, Carl-Zeiss GMBH, Jena, Germany). For imaging YFP, we used confocal configurations as

described before (Vermeer et al., 2006).

Seed germination

Mature seeds were harvested and stored at room temperature. Seeds were surface sterilized in a

desiccator by using 20 ml thin bleach and 1ml 37% HCl for 3 hours and then were sown on square petri

dish containing 30 ml medium consisting of ½ Murashi-Skoog (½ MS), 0.5% sucrose, pH 5.8, and 1.2 %

daishin agar with or without ABA or GA at indicated concentrations under 4 °C in dark for two nights

and transferred to long day condition (22 °C, 16 h of light and 8h of dark). Germination was scored as

radical emergence at indicated time points by using a binocular microscope (Leica MZFLIII).

Page 49: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

47

Root growth

Seeds were sterilized and stratified as described above. Plates were transfer to long day condition and

placed vertically under an angle of 70°. Four-day-old seedlings with comparable size were transfer to ½

MS ager plate. The plates were scanned 12 days after germination (Epson Perfection V700 scanner).

Primary root length and lateral root number from each genotype were quantified by ImageJ analysis

software (National Institute of Health (NIH)). Lateral root density was expressed as the lateral number

per primary length (LR number/ PR length). For gravitropic responses, seedlings were grown on ½MS

plates with 0.5% sucrose for 4 days. Plates were then rotated by 90°and scanned 2 days later. Bending

was expressed as curvature angle, which was quantified by imageJ analysis software (NIH) (Perera et

al., 2006).

Stomatal aperture

The stomatal aperture measurement was performed according to Distéfano et al. (2012) with minor

changes. The stomatal aperture treatments were performed on epidermal strips excised from the abaxial

side of fully expanded Arabidopsis leaves. Epidermal peels from leaves of 3-week-old plants grown at

22°C under 16 h of light and 8h of dark were stripped and immediately floated in opening buffer (5 mM

MES-KOH, pH 6.1, and 50 mM KCl) for 3 h. The strips were subsequently maintained in the same

opening buffer and exposed to different ABA concentration. After 90 min, stomata were digitized using

a Nikon DS-Fi 1 camera coupled to a Nikon Eclipse Ti microscope. The stomatal aperture width was

measured using ImageJ software (NIH).

32Pi-phospholipid labelling, extraction and analysis

Different types of tissues were labelled. For germinating seeds: Seeds were sterilized and stratified on

½ MS (pH 5.8) as described and germinated under long day condition for around 20h when testa

ruptured. Germinating seeds were then transferred to 200 µl buffer (2.5 mM MES, pH 5.8, 1 mM KCl)

containing 5-10 µCi 32PO43− (32Pi) (carrier free; Perklin-Elmer) in 2 ml Eppendorf microcentrifuge tube

for 24 h. Samples were then treated with 200 µl buffer with or without ABA for the times and

concentrations indicated.

Epidermal leaf peels: Leaves of 3-week-old plants grown at 22°C under 16 h of light and 8 h of

dark were stripped and immediately floated on 100 µl opening buffer (10 mM MES, pH 6.1 and 50 mM

KCl) containing 32Pi (5-10 µCi) in a 48-wells cell culture plate (Greiner bio-one) for 3 h. Samples were

treated with 400 µl buffer (10 mM MES-KOH, pH 6.1, 2.5 µM CaCl2) with or without ABA for the

times and concentrations indicated.

Seedlings: Five-day-old seedlings were transferred to 200 µl labeling buffer (2.5 mM MES-

KOH, pH 5.8, 1 mM KCl) containing 32Pi (5-10 µCi) in 2 ml Eppendorf tubes and labeled overnight

Page 50: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

48

(~16 h) or 3h. Samples were treated the next day by adding 200 µl labeling buffer with or without ABA

or Sorbitol for times and concentrations indicated.

All treatments were stopped by adding perchloric acid at a final concentration of 5% (v/v) for

5-10 min, after which the material was transferred to 400 µl of CHCl3/MeOH/HCl [50:100:1 v/v)] to

extract the lipids. After 15 min, 400 µl of CHCl3 was added followed by 200 µl of 0.9 % (w/v) NaCl to

separate the extract into two phases. Lipid fractions were washed and concentrated as described earlier

(Munnik & Zarza, 2013). Lipids were separated by thin-layer chromatography (TLC) using an alkaline

solvent system, containing: chloroform/methanol/28% ammonia/water [90:70:4:16 (v/v)] (Munnik et al.,

1994). Radioactive phospholipids were visualized by autoradiography and quantified by

phosphoimaging (Molecular Dynamics, Sunnyvale, CA, USA). Individual phospholipid levels are

expressed as the percentage of the total 32P-lipid fraction.

Inositol phosphates labeling, extraction and HPLC Analyses

For the measurement of inositol polyphosphates (IPPs), two different procedures were followed. The

first is based on the method described in Laha et al. (2015) with minor modifications. Seedlings were

grown under short day (22 °C, 12 h of light and 12h of dark) and sterile conditions in plant media (½

MS, 2% sucrose, pH 5.7, 0.6% phytagel) for 11 days and then 10 seedlings were transferred to 2ml

liquid medium (¼ MS, pH 5.7, 0.3% phytagel) containing 3H-myo-inositol (80 µCi, Biotrend, ART-

0261-5, Cologne, Germany) for 7 days. Seedlings were washed two times with water before harvesting

and then snap-freezed into liquid N2. IPPs were extracted (Azevedo and Saiardi, 2006) and resolved by

strong anion exchange chromatography HPLC (using the partisphere SAX 4.6 x 125mm column;

Whatman) at a flow rate of 0.5 mL/ min, using a shallow gradient formed by buffer A (1 mM EDTA)

and buffer B (1 mM EDTA and 1.3 M Ammonium Phosphate, pH 3.8 with H3PO4). Fractions were

collected every minute and radioactivity quantified by liquid scintillation counting. The results are

expressed as percentage of total. The latter was determined by counting all fractions from 13 min to the

end of the run.

Alternatively, IPP were determined as described by Desai et al., (2014) with some

modifications. Seedlings were grown in ½ MS with 0.8 % agar under long day condition (100 µE light

with a 16 h day and 8 h night cycle) for 4 days. Fifteen seedlings were incubated with 50 µl medium (1x

MS, 1% sucrose, pH 5.7) and 100 µl of aqueous myo [2-3H(N)]-inositol (100 µCi, American

Radiolabeled Chemicals Cat. #ART 0116A, specific activity 20 Ci/mmol) was added to each tube. The

tubes were incubated with supplemental light for 4 days. IPPs were extracted as Azevedo and Saiardi

(2006) described, by vortexing the tissue with glass beads in extraction buffer (25 mM EDTA, 10

mg/ml IP6 and 1M HClO4). Samples were then neutralized to ~pH 6 to 8 with 250 mM EDTA, 1M

K2CO3. Samples were dried to a volume of 70 µl and separated using a binary HPLC pump (Beckman

Coulter) equipped with a Partisphere-SAX (4.6 x 125 mm) column, which was connected to a guard

cartridge. The elution gradient was set up as described by Azevedo and Saiardi (2006) using the same

Page 51: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

49

buffers as above at a flow rate of 1ml/min. An on-line IN/US radiation detector was used to generate

chromatograms. Four ml of Ultima-Flo AP scintillation cocktail (Perkin Elmer, Waltham, MA, USA)

was added to each 1 ml eluted fraction post-detector to quantify the radioactivity of the eluted fractions

using the 3H window of a Beckman Coulter LS6500 Scintillation Counter. Scintillation counts were

graphed using MicroSoft Excel. The 3H-myo-Ins cpm incorporated into total IPPs was calculated by

taking the sum of cpm of all fractions and subtracting the peak of free 3H-Ins cpm. The amount of each

IPP was calculated as follows: [(Σ cpms in peak) / (total IPP)]*100].

Thesis for printing.docxDrought tolerance assays

Determination of survival rates, fresh weight (FW) and dry weight (DW) under water deficit condition

and water loss were performed as described previously (Hua et al., 2012; Osakabe et al., 2013) with

some changes. Seeds were stratified under 4°C, dark for 2 nights and sown on soil pot (4.5 cm x 4.5 cm

x 7.5 cm) directly. Nine plants were grown in each pot with certain amount of soil (80 g) under short

day condition (22 °C with 12 h light/12 h dark) for 4 weeks and then subjected to dehydration by

withholding them for water for 2 weeks, while control plants were normally watered. And then plants

were photographed. The plants were re-watered for another week and photographed. The surviving

green plants were counted and survival rate was determined by the percentage of green plants compared

to total plants. Each experiment used 36 plants per genotype and experiments were repeated at least 3

times.

To determine the FW and DW under dehydration stress, plants were grown under short day conditions

as described above for 4 weeks and experienced 1 week dehydration by water withholding, while

control plants were normally watered. Rosettes FWs were scored immediately after detachment. After

complete drying, dry DWs were also determined. Eigtheen plants from each genotype were used for

measurement and experiments were repeated for 3 times.

To assay the water-loss, rosettes from 4-week-old plants were detached and FW determined

every one hour by weighing. Water content was calculated as a percentage from the initial FW. Twenty

plants were used for each experiment and each experiment was repeated at least 3 times.

Soluble carbohydrates measurement in seeds

Soluble carbohydrates were determined as described by (Ribeiro et al., 2014) with minor modifications.

Three milligrams of dry seeds were transferred to a 2 mL Eppendorf tube and homogenized in 1 mL of

methanol (80% v/v) with the addition of 40 µg of melezitose as internal standard. Samples were

incubated in a water bath for 15 minutes at 76°C and dried by vacuum centrifugation. Then, 500 µL of

milliQ water was added, thoroughly vortexed and centrifuged for 5 min at 17.000 g in an Eppendorf

centrifuge. The supernatant was analyzed with a Dionex HPLC system (ICS 5000 + DC) using a

CarboPac PA1, 4 x 250-mm column (Dionex) preceded by a guard column (CarboPac PA1, 4 × 50 mm).

Mono-, di-, and trisaccharides were separated by elution in an increasing concentration of NaOH (20-

350 mM) with a flow rate of 1 mL per minute. Peaks were identified by coelution of standards.

Page 52: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

50

Phloem sap soluble carbohydrates measurement

Phloem exudates were extracted and analyzed as described earlier (Guelette et al., 2012; Tetyuk et al,

2013; modified from Roessner et al., 2000). The hydrophilic fraction (sugars, amino acids, small

molecules) was dried and derivatized using methoxyamine in pyridine and BSTFA (N,O-bis-trimethyl-

silyltrifluoroacetamide) with 1% trimethylchlorosilane. The metabolites were subsequently separated

using an Agilent 5890N GC system, coupled to a 5973 inert MSD using a DB-5MS column (J&W

Scientific, 30m x 0.32mm ID x 0.25 µm film; temperature program: 70-5, 5/min-320-5). Identification

of compounds occurred using the NIST library (NIST Standard Reference Database;

https://www.nist.gov/srd/nist-standard-reference-database-1a-v14) in combination with co-elution of

standards. Peak area determination was performed using QuanLynx, a quantification software within

the MassLynx software (Waters). The experiment was repeated five times.

ACKNOWLEDGEMENTS

We thank Max van Hooren for critically reading this manuscript. This work was financially supported

by a grant from the China Scholarship Council (CSC File No. 201206300058 to Q.Z.) and by the

Netherlands Organization for Scientific Research (NWO/ALW 813.06.003, 864.05.001, and 867.15.020

to T.M.).

REFERENCES

Allen GJ, and Sanders D (1994) Osmotic stress enhances the competence of Beta vulgaris vacuoles to respond to inositol

1,4,5-trisphosphate. Plant J 6:687–695. Arisz SA, Testerink C, and Munnik T (2009) Plant PA signaling via diacylglycerol kinase. Biochim Biophys Acta 1791:869–

875. Arisz SA, van Wijk R, Roels W, Zhu JK, Haring MA, and Munnik T (2013) Rapid phosphatidic acid accumulation in response

to low temperature stress in Arabidopsis is generated through diacylglycerol kinase. Front Plant Sci 4:1–15. Assmann SM, and Jegla T (2016) Guard cell sensory systems: recent insights on stomatal responses to light, abscisic acid, and

CO2. Curr Opin Plant Biol 33:157–167. Balla T (2013) Phosphoinositides: tiny lipids with giant impact on cell regulation. Physiol Rev 93:1019–1137. Benková E, and Bielach A (2010) Lateral root organogenesis - from cell to organ. Curr Opin Plant Biol 13:677–683. Blatt MR (2000) Cellular signaling and volume control in stomatal movements in plants. Annu Rev Cell Dev Biol 16:221–241. Blatt MR, Thiel G, and Trentham DR (1990) Reversible inactivation of K+ channels of Vicia stomatal guard cells following

the photolysis of caged inositol 1,4,5-trisphosphate. Nature 346:766–769. Boss WF, Sederoff HW, Im YJ, Moran N, Grunden AM, and Perera IY (2010) Basal Signaling Regulates Plant Growth and

Development. Plant Physiol 154:439–443. Burton A, Hu X, and Saiardi A (2009) Are inositol pyrophosphates signalling molecules? J Cell Physiol 220:8–15. Chen X, Chang M, Wang B, and Wu B (1997) Cloning of a Ca2+-ATPase gene and the role of cytosolic Ca2+ in the

gibberellin-dependent signaling pathway in aleurone cells. Plant J 11:363–371. Cho MH, Shears SB, and Boss WF (1993) Changes in phosphatidylinositol metabolism in response to hyperosmotic stress in

Daucus carota L. cells grown in suspension culture. Plant Physiol 103:637–647. Clough SJ, and Bent AF (1998) Floral dip: A simplified method for Agrobacterium-mediated transformation of Arabidopsis

thaliana. Plant J 16:735–743. Darwish E, Testerink C, Khalil M, El-Shihy O, and Munnik T (2009) Phospholipid signaling responses in salt-stressed rice

leaves. Plant Cell Physiol 50:986–997. Das S, Hussain A, Bock C, Keller WA, and Georges F (2005) Cloning of Brassica napus phospholipase C2 (BnPLC2),

phosphatidylinositol 3-kinase (BnVPS34) and phosphatidylinositol synthase1 (BnPtdIns S1) - Comparative analysis of the effect of abiotic stresses on the expression of phosphatidylinositol signal transdu. Planta 220:777–784.

De Jong CF, Laxalt AM, Bargmann BOR, De Wit PJGM, Joosten MHAJ, and Munnik T (2004) Phosphatidic acid accumulation is an early response in the Cf-4/Avr4 interaction. Plant J 39:1–12.

Page 53: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

51

Desai M, Rangarajan P, Donahue JL, Williams SP, Land ES, Mandal MK, Phillippy BQ, Perera IY, Raboy V, and Gillaspy GE (2014) Two inositol hexakisphosphate kinases drive inositol pyrophosphate synthesis in plants. Plant J 80:642–653.

DeWald DB, Torabinejad J, Jones C a, Shope JC, Cangelosi a R, Thompson JE, Prestwich GD, and Hama H (2001) Rapid accumulation of phosphatidylinositol 4,5-bisphosphate and inositol 1,4,5-trisphosphate correlates with calcium mobilization in salt-stressed arabidopsis. Plant Physiol 126:759–769.

Di Fino L, D’Ambrosio J, Tejos R, van Wijk R, Lamattina T, Munnik T, Pagnussat G, and Laxalt AM (2017) Arabidopsis phosphatidylinositol-phospholipase C2 (PLC2) is involved in female gametogenesis and embryo development. Planta 2:1–12.

Dietrich P, Sanders D, and Hedrich R (2001) The role of ion channels in light-dependent stomatal opening. J Exp Bot 52:1959–1967.

Distéfano AM, Scuffi D, García-Mata C, Lamattina L, and Laxalt AM (2012) Phospholipase D?? is involved in nitric oxide-induced stomatal closure. Planta 236:1899–1907.

Ditengou FA, Teale WD, Kochersperger P, Flittner KA, Kneuper I, van der Graaff E, Nziengui H, Pinosa F, Li X, Nitschke R, Laux T, and Palme K (2008) Mechanical induction of lateral root initiation in Arabidopsis thaliana. Proc Natl Acad Sci U S A 105:18818–18823.

Dowd PE, Coursol S, Skirpan AL, Kao T, and Gilroy S (2006) Petunia phospholipase C1 is involved in pollen tube growth. Plant Cell 18:1438–1453.

Fleet CM, Ercetin ME, and Gillaspy GE (2009) Inositol phosphate signaling and gibberellic acid. Plant Signal Behav 4:73–74. Flores S, and Smart CC (2000) Abscisic acid-induced changes in inositol metabolism in Spirodela polyrrhiza. Planta 211:823–

832. Gangl R, and Tenhaken R (2016) Raffinose family oligosaccharides act as galactose stores in seeds and are required for rapid

germination of Arabidopsis in the Dark. Front Plant Sci 7:1–15. Gao K, Liu YL, Li B, Zhou RG, Sun DY, and Zheng SZ (2014) Arabidopsis thaliana phosphoinositide-specific phospholipase

C isoform 3 (AtPLC3) and AtPLC9 have an additive effect on thermotolerance. Plant Cell Physiol 55:1873–1883. Georges F, Das S, Ray H, Bock C, Nokhrina K, Kolla VA, and Keller W (2009) 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 32:1664–1681.

Gillaspy GE (2011) The cellular language of myo-inositol signaling. New Phytol 192:823–839. Gillaspy GE (2013) The role of phosphoinositides and inositol phosphates in plant cell signaling, in Lipid-mediated protein

signaling (Capelluto DGS ed) pp 141–157. Gilroy S, Read ND, and Trewavas a J (1990) Elevation of cytoplasmic calcium by caged calcium or caged inositol

triphosphate initiates stomatal closure. Nature 346:769–771. Gonorazky G, Guzzo MC, and Laxalt AM (2016) Silencing of the tomato phosphatidylinositol-phospholipase C2 (SlPLC2)

reduces plant susceptibility to Botrytis cinerea. Mol Plant Pathol 2:1–10. Gonorazky G, Ramirez L, Abd-El-Haliem A, Vossen JH, Lamattina L, ten Have A, Joosten MHAJ, and Laxalt AM (2014) The

tomato phosphatidylinositol-phospholipase C2 (SlPLC2) is required for defense gene induction by the fungal elicitor xylanase. J Plant Physiol 171:959–965.

Han B, Yang Z, Samma MK, Wang R, and Shen W (2013) Systematic validation of candidate reference genes for qRT-PCR normalization under iron deficiency in Arabidopsis. BioMetals 26:403–413.

Hartog M Den, Verhoef N, and Munnik T (2003) Nod factor and elicitors activate different phospholipid signaling pathways in suspension-cultured Alfalfa cells. Plant Physiol 132:311–317.

Heilmann I (2016) Phosphoinositide signaling in plant development. Development 143:2044–2055. Helling D, Possart A, Cottier S, Klahre U, and Kost B (2006) Pollen tube tip growth depends on plasma membrane

polarization mediated by tobacco PLC3 activity and endocytic membrane recycling. Plant Cell 18:3519–3534. Hirayama T, Ohtot C, Mizoguchi T, and Shinozaki K (1995) A gene encoding a phosphatidylinositol-specific phospholipase C

is induced by dehydration and salt stress in Arabidopsis thaliana. Proc Natl Acad Sci USA 92:3903–3907. Hirayama T, and Umezawa T (2010) The PP2C–SnRK2 complex. Plant Signal Behav 5:160–163. Horvath I, Glatz A, Nakamoto H, Mishkind ML, Munnik T, Saidi Y, Goloubinoff P, Harwood JL, and Vigh L (2012) Heat

shock response in photosynthetic organisms: Membrane and lipid connections. Prog Lipid Res 51:208–220. Hou Q, Ufer G, and Bartels D (2016) Lipid signalling in plant responses to abiotic stress. Plant, Cell Environ 39:1029–1048. Hua D, Wang C, He J, Liao H, Duan Y, Zhu Z, Guo Y, Chen Z, and Gong Z (2012) A plasma membrane receptor kinase,

GHR1, mediates abscisic acid- and hydrogen peroxide-regulated stomatal movement in Arabidopsis. Plant Cell 24:2546–61.

Hunt L, and Gray JE (2001) ABA signalling: A messenger’s FIERY fate. Curr Biol 11:968–970. Hunt L, Mills LN, Pical C, Leckie CP, Aitken FL, Kopka J, Mueller-Roeber B, McAinsh MR, Hetherington AM, and Gray JE

(2003) Phospholipase C is required for the control of stomatal aperture by ABA. Plant J 34:47–55. Hunt L, Otterhag L, Lee JC, Lasheen T, Hunt J, Seki M, Shinozaki K, Sommarin M, Gilmour DJ, Pical C, and Gray JE (2004)

Gene-specific expression and calcium activation of Arabidopsis thaliana phospholipase C isoforms. New Phytol 162:643–654.

Irvine RF (2006) Nuclear inositide signalling-expansion, structures and clarification. Biochim Biophys Acta 1761:505–508. Jefferson RA, Kavanagh TA, and Bevan MW (1987) GUS fusions: beta-glucuronidase as a sensitive and versatile gene fusion

marker in higher plants. EMBO J 6:3901–3907. Kashem M, Itoh K, Iwabuchi S, Hori H, and Mitsui T (2000) Possible involvement of phosphoinositide-Ca2+ signaling in the

regulation of alpha-amylase expression and germination of rice seed (Oryza sativa L.). Plant cell Physiol 41:399–407. Kepinski S, and Leyser O (2005) The Arabidopsis F-box protein TIR1 is an auxin receptor. Nature 435:446–451. Kulik A, Wawer I, Krzywińska E, Bucholc M, and Dobrowolska G (2011) SnRK2 protein kinases—key regulators of plant

Page 54: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

52

response to abiotic stresses. Omi A J Integr Biol 15:859–872. Laha D, Johnen P, Azevedo C, Dynowski M, Weiß M, Capolicchio S, Mao H, Iven T, Steenbergen M, Freyer M, Gaugler P,

de Campos MKF, Zheng N, Feussner I, Jessen HJ, Van Wees SCM, Saiardi A, and Schaaf G (2015) VIH2 regulates the synthesis of inositol pyrophosphate InsP8 and jasmonate-dependent defenses in Arabidopsis. Plant Cell 27:1082–1097.

Laha D, Parvin N, Dynowski M, Johnen P, Mao H, Bitters ST, Zheng N, and Schaaf G (2016) Inositol polyphosphate binding specificity of the jasmonate receptor complex. Plant Physiol 171:2364–2370.

Lanteri ML, Lamattina L, and Laxalt AM (2011) Mechanisms of xylanase-induced nitric oxide and phosphatidic acid production in tomato cells. Planta 234:845–855.

Laxalt AM, and Munnik T (2002) Phospholipid signalling in plant defence. Curr Opin Plant Biol 5:332–338. Lee Y, Choi YB, Suh S, Lee J, Assmann SM, Joe CO, Kelleher JF, and Crain RC (1996) Abscisic acid-induced

phosphoinositide turnover in guard cell protoplasts of Vicia faba. Plant Physiol 110:987–996. Lee Y, Kim YW, Jeon BW, Park KY, Suh SJ, Seo J, Kwak JM, Martinoia E, Hwang I, and Lee Y (2007) Phosphatidylinositol

4,5-bisphosphate is important for stomatal opening. Plant J 52:803–816. Lemtiri-Chlieh F, MacRobbie EAC, and Brearley CA (2000) Inositol hexakisphosphate is a physiological signal regulating the

K+-inward rectifying conductance in guard cells. Proc Natl Acad Sci U S A 97:8687–8692. Lemtiri-Chlieh F, MacRobbie EAC, Webb AAR, Manison NF, Brownlee C, Skepper JN, Chen J, Prestwich GD, and Brearley

CA (2003) Inositol hexakisphosphate mobilizes an endomembrane store of calcium in guard cells. Proc Natl Acad Sci U S A 100:10091–10095.

Li L, He Y, Wang Y, Zhao S, Chen X, Ye T, Wu Y, and Wu Y (2015) Arabidopsis PLC2 is involved in auxin-modulated reproductive development. Plant J 84:504–515.

Lin WH, Ye R, Ma H, Xu ZH, and Xue HW (2004) DNA chip-based expression profile analysis indicates involvement of the phosphatidylinositol signaling pathway in multiple plant responses to hormone and abiotic treatments. Cell Res 14:34–45.

Liu HT, Huang WD, Pan QH, Weng FH, Zhan JC, Liu Y, Wan SB, and Liu YY (2006) Contributions of PIP2-specific-phospholipase C and free salicylic acid to heat acclimation-induced thermotolerance in pea leaves. J Plant Physiol 163:405–416.

Liu HT, Liu YY, Pan QH, Yang HR, Zhan JC, and Huang WD (2006) Novel interrelationship between salicylic acid, abscisic acid, and PIP 2-specific phospholipase C in heat acclimation-induced thermotolerance in pea leaves. J Exp Bot 57:3337–3347.

Luit AH Van Der, Piatti T, Doorn A Van, Musgrave A, Felix G, Boller T, and Munnik T (2000) Elicitation of Suspension-Cultured Tomato Cells Triggers the Formation of Phosphatidic Acid and Diacylglycerol Pyrophosphate 1. Plant Physiol 123:1507–1515.

Luo Y, Xie W, and Luo F (2012) Effect of several germination treatments on phosphatases activities and degradation of phytate in Faba Bean (Vicia faba L.) and Azuki Bean (Vigna angularis L.). J Food Sci 77:1023–1029.

Ma Y, Szostkiewicz I, Korte A, Moes D, Yang Y, Christmann A, and Grill E (2009) Regulators of PP2C phosphatase activity function as abscisic acid sensors. Science 324:1064–1069.

Martin TF. (2001) PI(4,5)P2 regulation of surface membrane traffic. Curr Opin Cell Biol 13:493–499. Meijer HJG, and Munnik T (2003) Phospholipid-based signaling in plants. Annu Rev Plant Biol 54:265–306. Melcher K, Ng L-M, Zhou XE, Soon F-F, Xu Y, Suino-Powell KM, Park S-Y, Weiner JJ, Fujii H, Chinnusamy V, Kovach A,

Li J, Wang Y, Li J, Peterson FC, Jensen DR, Yong E-L, Volkman BF, Cutler SR, Zhu J-K, and Xu HE (2009) A gate-latch-lock mechanism for hormone signalling by abscisic acid receptors. Nature 462:602–608, Nature Publishing Group.

Michell RH (2008) Inositol derivatives: evolution and functions. Nat Rev 9:151–161. Mickelbart M V., Hasegawa PM, and Bailey-Serres J (2015) Genetic mechanisms of abiotic stress tolerance that translate to

crop yield stability. Nat Rev Genet 16:237–251. Mills LN, Hunt L, Leckie CP, Aitken FL, Wentworth M, McAinsh MR, Gray JE, and Hetherington AM (2004) The effects of

manipulating phospholipase C on guard cell ABA-signalling. J Exp Bot 55:199–204. Mishkind M, Vermeer JEM, Darwish E, and Munnik T (2009) Heat stress activates phospholipase D and triggers PIP

accumulation at the plasma membrane and nucleus. Plant J 60:10–21. Miyazono K-I, Miyakawa T, Sawano Y, Kubota K, Kang H-J, Asano A, Miyauchi Y, Takahashi M, Zhi Y, Fujita Y, Yoshida

T, Kodaira K-S, Yamaguchi-Shinozaki K, and Tanokura M (2009) Structural basis of abscisic acid signalling. Nature 462:609–614, Nature Publishing Group.

Mueller-roeber B, and Pical C (2002) Inositol phospholipid metabolism in Arabidopsis . characterized and putative isoforms of inositol phospholipid kinase and phosphoinositide-specific phospholipase C. Plant Physiol 130:22–46.

Munemasa S, Hauser F, Park J, Waadt R, Brandt B, and Schroeder JI (2015) Mechanisms of abscisic acid-mediated control of stomatal aperture. Curr Opin Plant Biol 28:154–162.

Munnik T (2001) Phosphatidic acid: an emerging plant lipid second messenger. Trends Plant Sci 6:227–233. Munnik T (2014) PI-PLC: Phosphoinositide-phospholipase C in plant signalling, in Phospholipases in Plant Signaling (Wang

X ed) pp 27–54. Munnik T, and Laxalt A. (2013) Measuring PLD Activity In Vivo, in Plant Lipid Signaling Protocols (Munnik T, and

Heilmann I eds) pp 219–231. Munnik T, and Meijer HJ. (2001) Osmotic stress activates distinct lipid and MAPK signalling pathways in plants. FEBS Lett

498:172–178. Munnik T, Musgrave A, and de Vrije T (1994) Rapid turnover of polyphosphoinositides in carnation flower petals. Planta

193:89–98. Munnik T, and Nielsen E (2011) Green light for polyphosphoinositide signals in plants. Curr Opin Plant Biol 14:489–497. Munnik T, and Testerink C (2009) Plant phospholipid signaling: “in a nutshell”. J Lipid Res 50:S260–S265.

Page 55: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

53

Munnik T, van Himbergen JAJ, ter Riet B, Braun F-J, Irvine RF, van den Ende H, and Musgrave A (1998) Detailed analysis of the turnover of polyphosphoinositides and phosphatidic acid upon activation of phospholipases C and D in Chlamydomonas cells treated with non-permeabilizing concentrations of mastoparan. Planta 207:133–145.

Munnik T, and Vermeer JEM (2010) Osmotic stress-induced phosphoinositide and inositol phosphate signalling in plants. Plant Cell Environ 33:655–669.

Munnik T, and Zarza X (2013) Analyzing Plant Signaling Phospholipids Through 32Pi-Labeling and TLC, in Plant Lipid Signaling Protocols (Munnik T, and Heilmann I eds) pp 3–15.

Murthy PP, Renders JM, and Keranen LM (1989) Phosphoinositides in barley aleurone layers and gibberellic Acid-induced changes in metabolism. Plant Physiol 91:1266–1269.

Nakashima K, and Yamaguchi-Shinozaki K (2013) ABA signaling in stress-response and seed development. Plant Cell Rep 32:959–970.

Nambara E, Okamoto M, Tatematsu K, Yano R, Seo M, and Kamiya Y (2010) Abscisic acid and the control of seed dormancy and germination. Seed Sci Res 20:55–67.

Nishimura N (2009) Structural mechanism of abscisic acid bind and signaling by dimeric PYR1. Science (80- ) 326:1373–1380.

Nokhrina K, Ray H, Bock C, and Georges F (2014) Metabolomic shifts in Brassica napus lines with enhanced BnPLC2 expression impact their response to low temperature stress and plant pathogens. GM Crops Food 5:120–131.

Osakabe Y, Arinaga N, Umezawa T, Katsura S, Nagamachi K, Tanaka H, Ohiraki H, Yamada K, Seo S-U, Abo M, Yoshimura E, Shinozaki K, and Yamaguchi-Shinozaki K (2013) Osmotic stress responses and plant growth controlled by potassium transporters in Arabidopsis. Plant Cell 25:609–624.

Otegui MS, Capp R, and Staehelin LA (2002) Developing seeds of Arabidopsis store different minerals in two types of vacuoles and in the endoplasmic reticulum. Plant Cell 14:1311–1327.

Park S, Fung P, Nishimura N, Jensen DR, Fujii H, Zhao Y, Lumba S, Santiago J, Rodrigues A, Chow TF, Alfred SE, Bonetta D, Finkelstein R, Provart NJ, Desveaux D, Rodriguez PL, McCourt P, Zhu J-K, Schroeder JI, Volkmann BF, and Cutler SR (2014) Abscisic acid inhibits type 2C protein phosphatases via the PYR/PYL family of START proteins. Science (80- ) 324:1068–1069.

Perera IY, Heilmann I, and Boss WF (1999) Transient and sustained increases in inositol 1,4,5-trisphosphate precede the differential growth response in gravistimulated maize pulvini. Proc Natl Acad Sci USA 96:5838–5843.

Perera IY, Hung C, Brady S, Muday GK, and Boss WF (2006) A Universal Role for Inositol 1, 4, 5-Trisphosphate-Mediated Signaling in Plant Gravitropism. Plant Physiol 140:746–760.

Péret B, De Rybel B, Casimiro I, Benková E, Swarup R, Laplaze L, Beeckman T, and Bennett MJ (2009) Arabidopsis lateral root development: an emerging story. Trends Plant Sci 14:399–408.

Péret B, Larrieu A, and Bennett MJ (2009) Lateral root emergence: A difficult birth. J Exp Bot 60:3637–3643. Pical C, Westergren T, Dove SK, Larsson C, and Sommarin M (1999) Salinity and hyperosmotic stress induce rapid increases

in phosphatidylinositol 4,5-bisphosphate, diacylglycerol pyrophosphate, and phosphatidylcholine in Arabidopsis thaliana cells. J Biol Chem 274:38232–38240.

Pieterse CMJ (1998) A novel signaling pathway controlling induced systemic resistance in Arabidopsis. Plant Cell 10:1571–1580.

Pokotylo I, Kolesnikov Y, Kravets V, Zachowski A, and Ruelland E (2014) Plant phosphoinositide-dependent phospholipases C: Variations around a canonical theme. Biochimie 96:144–157.

Raho N, Ramirez L, Lanteri ML, Gonorazky G, Lamattina L, ten Have A, and Laxalt AM (2011) Phosphatidic acid production in chitosan-elicited tomato cells, via both phospholipase D and phospholipase C/diacylglycerol kinase, requires nitric oxide. J Plant Physiol 168:534–539.

Repp A, Mikami K, Mittmann F, and Hartmann E (2004) Phosphoinositide-specific phospholipase C is involved in cytokinin and gravity responses in the moss Physcomitrella patens. Plant J 40:250–259.

Ribeiro PR, Fernandez LG, de Castro RD, Ligterink W, and Hilhorst HW (2014) Physiological and biochemical responses of Ricinus communis seedlings to different temperatures: a metabolomics approach. BMC Plant Biol 14:1–14.

Rodriguez-Villalon A, Gujas B, van Wijk R, Munnik T, and Hardtke CS (2015) Primary root protophloem differentiation requires balanced phosphatidylinositol-4,5-biphosphate levels and systemically affects root branching. Development 142:1437–1446.

Roelfsema MRG, Hedrich R, and Geiger D (2012) Anion channels: master switches of stress responses. Trends Plant Sci 17:221–9, Elsevier Ltd.

Ruelland E, Cantrel C, Gawer M, Kader JC, and Zachowski A (2002) Activation of phospholipases C and D is an early response to a cold exposure in Arabidopsis suspension cells. Plant Physiol 130:999–1007.

Sanchez JP, and Chua NH (2001) Arabidopsis PLC1 is required for secondary responses to abscisic acid signals. Plant Cell 13:1143–1154.

Santiago J, Rodrigues A, Saez A, Rubio S, Antoni R, Dupeux F, Park SY, Márquez JA, Cutler SR, and Rodriguez PL (2009) Modulation of drought resistance by the abscisic acid receptor PYL5 through inhibition of clade A PP2Cs. Plant J 60:575–588.

Schroeder JI, Allen GJ, Hugouvieux V, Kwak JM, and Waner D (2001) Guard cell signal transduction. Annu Rev Plant Physiol plant mol biol 52:627–658.

Sean R, Pedro L, Ruth R, and Suzanne R (2010) Abscisic acid:emergence of a core signaling network. Annu Rev Plant Biol 61:651–679.

Sengupta S, Mukherjee S, Basak P, and Majumder AL (2015) significance of galactinol and raffinose family oligosaccharide synthesis in plants. Front Plant Sci 6:1–11.

Shears SB (2009) Diphosphoinositol polyphosphates: metabolic messengers? Mol Pharmacol 76:236–252. Siegel RS, Xue S, Murata Y, Yang Y, Nishimura N, Wang A, and Schroeder JI (2009) Calcium elevation-dependent and

Page 56: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

54

attenuated resting calcium-dependent abscisic acid induction of stomatal closure and abscisic acid-induced enhancement of calcium sensitivities of S-type anion and inward-rectifying K channels in Arabidopsis guard cells. Plant J 59:207–220.

Song M, Liu S, Zhou Z, and Han Y (2008) TfPLC1, a gene encoding phosphoinositide-specific phospholipase C, is predominantly expressed in reproductive organs in Torenia fournieri. Sex Plant Reprod 21:259–267.

Soon F-F, Ng L-M, Zhou XE, West GM, Kovach A., Tan MHE, Suino-Powell KM, He Y, Xu Y, Chalmers MJ, Brunzelle JS, Zhang H, Yang H, Jiang H, Li J, Yong E-L, Cutler S, Zhu J-K, Griffin PR, Melcher K, and Xu HE (2012) Molecular Mimicry Regulates ABA Signaling by SnRK2 Kinases and PP2C Phosphatases. Science (80- ) 335:85–88.

Staxen I, Pical C, Montgomery LT, Gray JE, Hetherington M, and McAinsh MR (1999) Abscisic acid induces oscillations in guard-cell cytosolic free calcium that involve phosphoinositide-specific phospholipase C. Proc Natl Acad Sci U S A 96:1779–1784.

Stevenson-Paulik J, Bastidas RJ, Chiou S-T, Frye R a, and York JD (2005) Generation of phytate-free seeds in Arabidopsis through disruption of inositol polyphosphate kinases. Proc Natl Acad Sci U S A 102:12612–12617.

Stevenson JM, Perera IY, Heilmann I, Persson S, and Boss WF (2000) Inositol signaling and plant growth. Trends Plant Sci 5:252–258.

Sui Z, Niu L, Yue G, Yang A, and Zhang J (2008) Cloning and expression analysis of some genes involved in the phosphoinositide and phospholipid signaling pathways from maize (Zea mays L.). Gene 426:47–56.

Takahashi S, Katagiri T, Hirayama T, Yamaguchi-Shinozaki K, and Shinozaki K (2001) Hyperosmotic stress induces a rapid and transient increase in inositol 1,4,5-trisphosphate independent of abscisic acid in Arabidopsis cell culture. Plant Cell Physiol 42:214–222.

Tan X, Calderon-Villalobos LI a, Sharon M, Zheng C, Robinson C V, Estelle M, and Zheng N (2007) Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature 446:640–645.

Tasma IM, Brendel V, Whitham S a., and Bhattacharyya MK (2008) Expression and evolution of the phosphoinositide-specific phospholipase C gene family in Arabidopsis thaliana. Plant Physiol Biochem 46:627–637.

Testerink C, and Munnik T (2005) Phosphatidic acid: a multifunctional stress signaling lipid in plants. Trends Plant Sci 10:368–375.

Tripathy MK, Tyagi W, Goswami M, Kaul T, Singla-Pareek SL, Deswal R, Reddy MK, and Sopory SK (2011) Characterization and Functional Validation of Tobacco PLC Delta for Abiotic Stress Tolerance. Plant Mol Biol Report 30:488–497.

Valluru R, and Van den Ende W (2011) Myo-inositol and beyond - emerging networks under stress. Plant Sci 181:387–400, Elsevier Ireland Ltd.

Van den Ende W (2013) Multifunctional fructans and raffinose family oligosaccharides. Front Plant Sci 4:1–11. van Leeuwen W, Vermeer JEM, Gadella TWJ, and Munnik T (2007) Visualization of phosphatidylinositol 4,5-bisphosphate in

the plasma membrane of suspension-cultured tobacco BY-2 cells and whole Arabidopsis seedlings. Plant J 52:1014–1026.

Vermeer JEM, Thole JM, Goedhart J, Nielsen E, Munnik T, and Gadella TWJ (2009) Imaging phosphatidylinositol 4-phosphate dynamics in living plant cells. Plant J 57:356–372.

Vermeer JEM, van Leeuwen W, Tobeña-Santamaria R, Laxalt AM, Jones DR, Divecha N, Gadella TWJ, and Munnik T (2006) Visualization of PtdIns3P dynamics in living plant cells. Plant J 47:687–700.

Villasuso AL, Molas ML, Racagni G, Abdala G, and Machado-Domenech E (2003) Gibberellin signal in barley aleurone: Early activation of PLC by G protein mediates amylase secretion. Plant Growth Regul 41:197–205.

Vossen JH, Abd-El-Haliem A, Fradin EF, Van Den Berg GCM, Ekengren SK, Meijer HJG, Seifi A, Bai Y, Ten Have A, Munnik T, Thomma BPHJ, and Joosten MHAJ (2010) 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 62:224–239.

Wang CR, Yang AF, Yue GD, Gao Q, Yin HY, and Zhang JR (2008) Enhanced expression of phospholipase C 1 (ZmPLC1) improves drought tolerance in transgenic maize. Planta 227:1127–1140.

Ward JM, Mäser P, and Schroeder JI (2009) Plant ion channels: gene families, physiology, and functional genomics analyses. Annu Rev Physiol 71:59–82.

Wheeler GL, and Brownlee C (2008) Ca2+ signalling in plants and green algae - changing channels. Trends Plant Sci 13:506–514.

Yin P, Fan H, Hao Q, Yuan X, Wu D, Pang Y, Yan C, Li W, Wang J, and Yan N (2009) Structural insights into the mechanism of abscisic acid signaling by PYL proteins. Nat Struct Mol Biol 16:1230–1236.

York JD (2006) Regulation of nuclear processes by inositol polyphosphates. Biochim Biophys Acta - Mol Cell Biol Lipids 1761:552–559.

Zarza X, Atanasov KE, Marco F, Arbona V, Carrasco P, Kopka J, Fotopoulos V, Munnik T, Gomez-Cadenas A, Tiburcio AF, and Alczar R (2016) Polyamine oxidase 5 loss-of-function mutations in Arabidopsis thaliana trigger metabolic and transcriptional reprogramming and promote salt stress tolerance. Plant, Cell Environ 1–16.

Zhang J, Vanneste S, Brewer PB, Michniewicz M, Grones P, Kleine-Vehn J, Löfke C, Teichmann T, Bielach A, Cannoot B, Hoyerová K, Chen X, Xue HW, Benková E, Zažímalová E, and Friml J (2011) Inositol trisphosphate-induced Ca2+ signaling modulates auxin transport and pin polarity. Dev Cell 20:855–866.

Zheng S, Liu Y, Li B, Shang Z, Zhou R, and Sun D (2012) Phosphoinositide-specific phospholipase C9 is involved in the thermotolerance of Arabidopsis. Plant J 69:689–700.

Zhu J (2016) Abiotic Stress Signaling and Responses in Plants. Cell 167:313–324. Zhu J (2002) Salt and Drought stress signal transduction in plants. Annu Rev Plant Biol 53:247–273. Zonia L, and Munnik T (2004) Osmotically induced cell swelling versus cell shrinking elicits specific changes in phospholipid

signals in tobacco pollen tubes. Plant Physiol 134:813–823.

Page 57: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

55

Page 58: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

56

SUPPLEMENTAL DATA Supplemental Table S1. Primers for the identification of PLC3 T-DNA insertion mutants and for PLC3 RT-PCR.

Primers Sequence PLC3_Forward TGCTGAAGTTCGTCATGGCAG PLC3_Reverse GTCCACCCAACATGAGGATCG LBa* TGGTTCACGTAGTGGGCCATCG TUBLINα4_Forward CCAGCCACCAACAGTTGTTC TUBLINα4_Reverse CACAAGACGAGATTATAGAGA

• LBa and PLC3_Forward combination is for T-DNA insertion identification

Supplemental Figure S1. PLC3pro:GUS-YFP expression in seedling grown at 45° angle. To find a stronger correlation between lateral root formation and the segmented PLC3-GUS expression, seedlings were grown on ½MS agar plates at an angle of 45° for 10 days to force later root formation at the curved sites of the primary root. (a) Cartoon of the setup. (b) Cartoon of the curvy seedlings generated and the lateral root formation at the curved sites (blue circles). (c) Histological GUS analysis of PLC3pro:GUS-YFP seedlings grown for 10-d at a 45° angle. Using this setup, less segments without lateral root were found (red circles). These results also show that GUS activity is not homogenously expressed throughout the root vasculature, going from segmented (top), to complete GUS possitive (middle) to no GUS activity (root tip, transition zone).

Page 59: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

57

Supplemental Figure S2. Confocal analysis of PLC3pro:GUS-YFP expression. Confocal image of longitudinal section (a) and cross section (b) of 5-d old seedlings. (c) eFP browser database of PLC3 expression in old and young root tissues.

Supplemental Figure S3. Soluble carbohydrates content in the phloem sap of wild-type and plc3 mutants. Phloem was isolated from 6-week-old Arabidopsis plants and their carbohydrates analyzed and quantified by GC-MS . Values are the means of triplicates ± SD from 3 independent experiments.

Supplemental Figure S4. Inositolpolyphosphate levels in wild-type and plc3 mutants. (a) Inositolpolyphosphate levels in wild-type and plc3 mutants. (b) IP7 in wild type and plc3-mutant seedlings. Four-days old seedlings were labelled with [2-3H(N)]-inositol for 4 days after which IPPs were extracted and resolved by HPLC-SAX analysis. Fractions were collected each minute and analyzed by liquid scintillation counting. The 3H-myo-Ins cpm incorporated into all inositolphosphates (total IPs) was calculated by taking the sum of cpm of all fractions and subtracting the 3H-Ins peak cpm. The amount of each IP was calculated as follows: [(Σ cpms in peak) / (total InsPs)]*100]. Data shown are means ± SE (n=10) from three independent experiments.

Page 60: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

58

Supplemental Figure S5. Effect of ABA on root development in wild-type and plc3 mutants. (a) Seedling morphology of wild-type and plc3 under normal and ABA conditions. Seeds were germinated on ½MS with 0.5% sucrose for 4 days and then transferred to ½ MS with and without ABA (10 µM). Photographs were taken 12 d after germination. (b) Relative primary root- (PR-) and lateral root (LR) growth were calculated as a percentage of the length under control condition. Three independent experiments were performed. Data shown are the means ± SD (n>10) for one representative experiment.

Page 61: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC3inArabidopsis

59

Supplemental Figure S6. Effect of ABA on phospholipid-signaling responses in Arabidopsis wt- and plc3 seedlings. (a-d) Time-course of ABA response in wt seedlings. Six-day-old seedlings were 32Pi-labeled for 3h and then treated with buffer with or without 100 µM ABA for different periods of time (0, 2, 4, 8, 16, 32 and 64 min). Lipids were extracted and separate by TLC. Radioactivity was visualized by autoradiography (a) and quantified by phosphoimaging (b-d). Lipids are expressed as fold-increase with respect to control. Values are the means of triplicates ± SD for one representative experiment. Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method). (e-h) ABA response in plc3 mutants. Five-day-old wt- and plc3 seedlings were 32Pi-labeled for 3h and then treated with buffer or 100 µM ABA for 1 h. (e) Autoradiograph of TLC. (f-h) Quantification of PIP2, PIP and PA. Data shown are the means ± SE of three independent experiments. Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method).

Page 62: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter2

60

Supplemental Figure S7. Effect of GA on seed germination of wild-type and plc3 mutants Seeds germination rate of wild-type and plc3 mutants in the absence (a) or presence of 1 µM GA (b). Seeds were germinated on ½MS with 0.5% sucrose plates with or without GA at 22 °C after 2 days of stratification at 4 °C. Germination is defined by radical emergence and was scored at the indicated times. In (c), the relative effect of GA on seed germination is calculated. Data shown are the means ± SE of 3 independent experiments (n=55 seeds for each genotype). Asterisks (*) mark that plc3 value are significantly different from wild-type based on Student’s t-test (P< 0.05).

Supplemental Figure S8. Gravitropic response of roots from wild-type and plc3 mutant seedlings (a,b) Seedling morphology of wild-type and plc3 mutants under gravitropic stimulation. Seedlings were grown on ½ MS with 0.5% sucrose plate for 4 days, plates were then rotated by 90°. Photographs were taken 2 days after plate turning (6-d old seedling). (c) Bending was expressed as curvature angle. Values are means ± SD for one representative experiment (n>20). Three independent experiments were performed.

Page 63: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

61

Chapter 3

Functional characterization of PLC5 in Arabidopsis thaliana - knock-down affects lateral root initiation while overexpression stunts root hair growth and enhances drought tolerance

Qianqian Zhang,1,2 Ringo van Wijk,1,2 Muhammad Shahbaz,1# Joop E.M. Vermeer1,3, Aisha Guardia,4 Denise Scuffi, 4 Carlos García-Mata,4 Wim Van den Ende,5 Susanne Hoffmann-Benning,6 Michel A. Haring,1 Ana M. Laxalt,4 & Teun Munnik1,2

1Section Plant Physiology or 2Section Plant Cell Biology, Swammerdam Institute for Life Sciences, University of Amsterdam, Science Park 904, Amsterdam, 1098 XH, The Netherlands, 3Dept. of Plant and Microbial Biology, University of Zürich, Zürich, Switzerland, 4Instituto de Investigaciones Biológicas (IIB-CONICET-UNMdP), Universidad Nacional de Mar del Plata, Mar del Plata, Argentina, 5Laboratory of Molecular Plant Biology, University of Leuven, Leuven, Belgium, 6Dept. of Biochemistry & Molecular Biology, Michigan State University, East Lansing, MI, USA. #Current address: Dept. of Botany, University of Agriculture, Faisalabad, Pakistan

Page 64: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

62

ABSTRACT

In animal cells, phospholipase C (PLC) is known to hydrolyse phosphatidylinositol 4,5-

bisphosphate (PIP2) to generate the second messengers, inositol 1,4,5-trisphosphate (IP3) and

diacylglycerol (DAG), which release intracellular Ca2+ and activate protein kinase C (PKC),

respectively, orchestrating a wide range of cellular- and physiological processes. Plants contain

PLCs too but lack IP3 receptors and PKC, and it still remains largely unclear what its

physiological role is in plants and what molecular targets it has. The genome of Arabidopsis

thaliana encodes 9 PLC genes. Earlier work on PLC2, PLC3 and PLC9 revealed roles for PLC in

gametogenesis, ABA signalling, lateral root formation and heat stress tolerance. Here, we

functionally characterised the role of PLC5. Promoter-GUS analyses revealed that this gene is

predominantly expressed in vascular tissue, most likely the phloem, including roots, leaves and

flowers, but expression was also detected in the root-apical meristem, in guard cells and in

trichomes. We only managed to find one homozygous T-DNA insertion line, plc5-1, which turned-

out to be a knock-down mutant, suggesting that a KO mutant is probably lethal. Growth of plc5-1

plants on agar plates consistently exhibited a ~20% reduction in their lateral root formation. The

latter was caused by a decrease in initiation rather than emergence of the lateral roots. PLC3 was

found to be required for lateral root formation earlier (Chapter 2), but a double plc3plc5 mutant

did not intensify the phenotype, indicating the involvement of possible additional redundant

PLCs. Complementation of plc5-1 with the PLC5-wt gene, expressed behind its own promoter,

restored growth and rescued the lateral root phenotype. UBQ10-Overexpression of PLC5 did not

affect lateral root development, but was found to stunt root hair growth, to decrease the stomatal

aperture and to increase their tolerance to drought stress. In vivo 32Pi-labeling analyses of PLC's

substrate/product lipids revealed no differences in plc5-1 seedlings, however, PLC5-

overexpression lines clearly exhibited reduced levels of PIP- and PIP2 and increased levels of

phosphatidic acid (PA), the latter likely produced through phosphorylation of PLC-generated

DAG. Inducible overexpression of PIP5K3 in PLC5 overexpressor lines recovered the stunted

root-hair growth and restored PIP2 level. These results provide independent evidence for PIP2's

role in polar tip growth of root hairs, confirm PLC's involvement in lateral root formation, and

that overexpression of PLC seems to increase drought tolerance in general.

Key words: PLC5; lateral root formation; root hair; drought tolerance.

Page 65: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

63

INTROUDCTION

Phospholipase C (PLC) signaling is implicated in various cellular events and plays crucial roles in all

eukaryotic cells. The pathway is best known from animal systems where it generates the second

messengers, inositol 1,4,5- trisphosphate (IP3) and diacylglycerol (DAG), which are formed though

PLC-catalyzed hydrolysis of the minor phospholipid, phosphatidylinositol 4,5-bisphosphate (PIP2). The

water-soluble IP3 diffuses into the cytosol where it triggers the release of Ca2+ from the ER via a ligand

gated-Ca2+ channel, while the lipid DAG, remains in the plasma membrane where it recruits and

activates members of the protein kinase C (PKC) family or stimulates TRP (transient receptor potential)

type- ion channels. Stimulation of the PLC pathway activates multiple downstream signaling cascades,

regulating multiple cellular processes (Irvine, 2006; Michell, 2008; Balla, 2013).

Plants also contain PLCs, but the signaling pathway is likely different from animal systems, as

plants lacks the primary targets for both IP3 and DAG, i.e., the IP3 receptor, PKC and TRP channels

(Zonia & Munnik, 2006; Wheeler and Brownlee, 2008; Munnik, 2014). Also different, is that its

potential substrate, PIP2 is hardly present in plant plasma membranes, while its precursor,

phosphatidylinositol 4-monophosphate (PIP) is relative abundant (Munnik and Testerink, 2009; Munnik

and Vermeer, 2010; Munnik and Nielsen, 2011; Munnik, 2014; Heilmann, 2016). In vitro, plant PLCs

do hydrolyze both PIP and PIP2 but in vivo it is still unknown what the normal substrate is (Munnik,

2014). Biochemical subfractionation analyses showed that most (>90%) PLC activity is present in the

plasma membrane fraction (Munnik et al., 1998a).

IP3 was initially linked to the release of intracellular Ca2+ (Gilroy et al., 1990; Blatt et al., 1990;

Allen and Sanders, 1994; Hunt and Gray, 2001), but later this was shown to be caused by IP6 that was

produced by phosphorylating IP3 (Lemtiri-Chlieh et al., 2000, 2003; Munnik and Vermeer, 2010).

Similarly, not DAG, but its phosphorylated product, phosphatidic acid (PA) has been emerging as the

plant lipid-second messenger (Munnik, 2001; Testerink and Munnik, 2005; Arisz et al., 2009; Pokotylo

et al., 2014; Munnik, 2014; Heilmann 2016). In plants, PA can be further phosphorylated into

diacylglycerolpyrophosphate (DGPP) by PA kinase, an enzyme that is lacking from animals but is

present in fungi, oomycetes and trypanosomes (van Schooten et al., 2006a). Whether the formation of

DGPP reflects an attenuation of the PA signal or the formation of a new signal (this lipid is normally

not there either), remains unknown. Meanwhile, various other inositolpolyphosphates (IPPs) than IP6

are emerging as signaling molecules, also in fungi and animals, where they have been implicated in ion

channel binding, phosphate sensing, transcription and embryonic development etc. In plants, IP4 has

been proposed to regulate a chloride channel (Zonia et al., 2002), while IP5 and IP6 were discovered in

the crystal structure of TIR1 and COI1, which are receptors for auxin- and jasmonate signaling,

respectively(Tan et al., 2007; Sheard et al., 2010). Gle1, an mRNA export factor, is an IP6-binding

protein that has recently been identified as a key activator of the ATPase/RNA helicase, LOS4 (low

expression of osmotically responsive genes 4), similar to the Gle1-IP6-Dbp5 (a LOS4 homolog)

Page 66: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

64

paradigm in yeast (Lee et al., 2015). Recently, SPX-domain containing proteins were also identified as

interactors with IP6 and many of these proteins are involved in phosphate homeostasis(Kuo et al., 2014;

Puga et al., 2014; Wild et al., 2016). Besides Ca2+ release in guard cells upon ABA stimulation

(Lemtiri-Chlieh et al., 2003), IP6 production has also been linked to disease resistance (Murphy et al.,

2008). Recently, it has been discovered that IP6 can also be pyrophosphorylated to IP7 and IP8, which

have also been implicated in signaling, including JA and plant defense (Laha et al., 2015, 2016;

Williams et al., 2015). Besides the lipid-generated pathway via PLC, IPPs can also be synthesized by

the conversion of glucose 6-phosphate into inositol 3-phosphate, which is then stepwise phosphorylated

by various IPP kinases (Munnik and Vermeer, 2010).

Plant PLCs have been implicated in various abiotic stress responses. PLC gene expression is

typically induced by various stresses, including salt, drought, heat, and cold stress (Hirayama et al.,

1995; Hunt et al., 2004; Lin et al., 2004; Das et al., 2005; Vergnolle et al., 2005; Zhai et al., 2005;

Skinner et al., 2005; Liu et al., 2006a; Tasma et al., 2008; Sui et al., 2008). Some of these have been

correlated with changes in IP3 (DeWald et al., 2001; Takahashi et al., 2001; Ruelland et al., 2002;

Zheng et al., 2012; Gao et al., 2014), but we currently know that these measurements are strongly

subject to artifacts and to changes in the flux of IP6, the most abundant IPP (for discussion, see Munnik

and Vermeer, 2010; Munnik 2014). Increases in PA have also been reported and some of these were

indeed generated (or at least partly) by DAG kinase (DGK) rather than phospholipase D (PLD), which

is another important PA generator (Arisz et al., 2009, 2013; Arisz and Munnik, 2013).

Increases in PIP2 upon salt- or heat stress have been reported (DeWald et al., 2001; van

Leeuwen et al., 2007; Mishkind et al., 2009; Simon et al., 2014), and these are mainly due to activation

of a PI4P 5-kinase (PIPK) rather than inhibition of a PLC (Mishkind et al., 2009; Zarza, 2017; Munnik

Lab, unpublished). Decreased PIP levels have been reported in response to osmotic- and temperature

stresses (Cho et al., 1993; Pical et al., 1999; DeWald et al., 2001; Zonia and Munnik, 2004; van

Leeuwen et al., 2007; Darwish et al., 2009; Mishkind et al., 2009; Vermeer et al., 2009; Horvath et al.,

2012; Munnik, 2014; Zarza et al., 2016). While many of these responses have been correlated to PLC

signaling, none of them have ever been functionally linked.

Overexpression of Arabidopsis PLC3 increased the plant’s tolerance to drought stress (Chapter

2; Zhang et al., 2017), which was consistent with earlier studies on tomato, canola and maize (Wang et

al., 2008; Georges et al., 2009; Tripathy et al., 2011).

PLC has also been associated to biotic-stress responses (Luit et al., 2000; Hartog et al., 2003;

De Jong et al., 2004; Vossen et al., 2010; Gonorazky et al., 2014, 2016), some in relation to nitric oxide

(NO) signaling (Laxalt et al., 2007; Lanteri et al., 2011; Raho et al., 2011). Genetic evidence for PLC’s

involvement in disease resistance has been obtained for tomato (Vossen et al., 2010; Gonorazky et al.,

2014, 2016; Abd-El-Haliem et al., 2016) and recently for Arabidopsis PLC2 (D’Ambrosio et al., 2017).

In both cases, PLC seems to be involved in the production of reactive oxygen species (ROS); for

AtPLC2 this involved RBOHD (D’Ambrosio et al., 2017) .

Page 67: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

65

While plant-PLC signaling is often linked to stress, there is also evidence for its involvement in

growth- and developmental responses. For example, PLC is important for pollen tube growth in petunia

and tobacco (Dowd et al., 2006; Helling et al., 2006) and affects male- and female gametophyte

development in Arabidopsis and Torenia fournieri (Song et al., 2008; Li et al., 2015; Di Fino et al.,

2017). In Physcomitrella, PLC is involved in cytokinin- and gravity responses (Repp et al., 2004). In

higher plants, IP3 changes have been correlated to gravitropism and Ca2+ signaling too (Perera et al.,

1999, 2006; Stevenson et al., 2000; Boss et al., 2010) but it is unclear whether this involves a PLC.

Loss-of-PLC3 in Arabidopsis did not alter gravitropism response (chapter2; Zhang et al., 2017). The

latter did, however, reveal reduced lateral root formation, reduced seed germination, and reduced

sensitivity to ABA with respect to stomatal aperture and the inhibition of seed germination (chapter2;

Zhang et al., 2017). Arabidopsis plc3- and plc9 mutants also exhibit reduced thermotolerance responses

(Zheng et al., 2012; Gao et al., 2014).

The Arabidopsis genome encodes 9 PLC genes (Hunt et al., 2004; Tasma et al., 2008; Munnik,

2014; Pokotylo et al., 2014). While AtPLC3 is mainly expressed in the vasculature, most likely in the

phloem and companion cells (Chapter 2; Zhang et al., 2017), we searched for other PLC genes

specifically expressed in the phloem, and this resulted in the identification of AtPLC5, which belongs to

a different subfamily than AtPLC3 (Hunt et al., 2004; Tasma et al. 2008). Here, we provide evidence

that AtPLC5 plays a role in lateral root development and that its overexpression increases the plant’s

tolerance to drought. Interestingly, this overexpression led to a stunted root-hair phenotype, which is

likely caused by the increased hydrolysis of PIP2 at the tip of the root hair, required for its growth.

Page 68: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

66

RESULTS

Knockdown of PLC5 affects root development

Previously, we found that PLC3 knock-down mutants were affected in their primary- and lateral root

development, and that this was linked to specific PLC3 expression in the phloem/companion cells

within the root (Chapter 2; Zhang et al., 2017). In the eFP browser, we found that Arabidopsis PLC5 is

also predicted to be expressed in the phloem/companion cells (eFPbrowser At5g58690) To functionally

characterize PLC5, we tried to obtain homozygous T-DNA insertion mutants, but uncovered only one,

i.e. plc5-1 (SALK_144469) (Fig. 1a). Q-PCR analysis revealed that it was a knock-down and not a

knock-out mutant (Fig. 1b). As no other T-DNA insertion mutants were found, it may be that PLC5-KO

mutants are actually lethal.

Figure 1. Phenotypic analysis of plc5-1 seedlings and complementation by wild-type PLC5 gene. (a) Representation of the PLC5 gene and T-DNA insertion position of plc5-1. Filled boxes and lines represent exons and introns, respectively. Open boxes and triangle represent untranslated regions and T-DNA insertion, respectively. (b) Q-PCR analysis of PLC5 expression level in wild type, plc5-1 and two complementation lines, PLC5#2 and #4 (in plc5-1 background using SAND as a reference gene. Values are means ± SD (n=3). (c) Seedling morphology of wild type, plc5-1 and complementation lines. Seeds were germinated on ½MS with 0.5% sucrose for 4 days, and then transferred to ½MS plates without sucrose. Photographs were taken 12 days after germination (DAG). (d) Primary root (PR) length and lateral root (LR) number at 12 DAG. Values are means ± SE of three independent experiments (n>20). Asterisk (*) indicate significance at P<0.05 compared to wild type, based on Student’s t test.

Page 69: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

67

Analyzing plc5-1 mutant seedlings revealed shorter primary roots (10 %) and contained less

(19 %) lateral roots than wild type (Fig. 1c-e), a phenotype that was also found for plc3 mutants

(Chapter 2; Zhang et al., 2017). Expression and phenotype could, however, be complemented by

expressing the PLC5 gene driven by its own promoter in the plc5-1 mutant background in independent

T3 lines (Fig. 1c-e), confirming PLC5's role in root development.

Detailed analyses of the different lateral root stages indicated that the plc5-1 phenotype

concerned lateral root initiation rather than their emergence (Supplemental Fig. S1). To analyze

plc3plc5-double mutants, plc3-2 (Chapter 2; Zhang et al., 2017) was crossed with plc5-1, and

homozygous T3 lines generated, but as shown in Supplemental Figure S2, the lateral root phenotype

was only marginally enhanced (22 % fewer lateral roots), indicating other bottle necks, or that

additional PLCs are involved. In the eFP browser, we additionally found PLC2 and PLC7 also to be

expressed in the phloem/companion cells, however, plc3plc5plc7-triple mutants as well as plc2-single

mutants were found to be homozygous lethal (Munnik lab, unpublished; Di Fino et al., 2017).

Expression of PLC5 during plant development

Earlier, analyses of PLC5 expression by Q-PCR revealed some variation in the different organs and

upon hormone- or stress treatments (Tasma et al., 2008). To investigate this locally in more detail, a

PLC5-promoter β-glucuronidase- (GUS-) reporter line, kindly provided by Dr. Julie Gray (Hunt et al.,

2004) was analysed. As shown in Figure 2a, the pPLC5-GUS expression was already apparent during

germination (28h after transfer to 22ºC) in the cotyledon, hypocotyl and root of the embryo (Fig. 2a).

During further development, GUS activity was mainly found in the vasculature throughout all stages,

i.e. root, cotyledons, leaves, hypocotyl, flower, incl. stamen, style, receptacle and pedicel (Fig. 2b-l).

Interestingly, GUS activity was also visualized in the whole trichome (Fig. 2i), which is different from

PLC3 where it is only expressed at the trichome base. Expression in the root was not homogenous. Like

PLC3 (Chapter 2, Zhang et al., 2017), it tended to be 'segmented' at the distal side of the root maturation

zone, while the expression was continuous in the apical maturation zone, and stopped near the transition

zone, but appeared again in the root tip (Fig. 2d-g). The latter is again different from PLC3, which was

never found to be expressed at the root tip. Strikingly, lateral roots were always found to emerge from a

colored segment, but not every segment led to a lateral root (Fig. 2c; Chapter 2, Zhang et al., 2017).

Moreover, both segmented- and root tip expressions were also observed during the formation of tertiary

roots (Fig. 2e). Lastly, we also found GUS activity in guard cells (Fig. 2m).

Together, these results confirm that PLC5 is expressed throughout the plant (Hunt et al., 2004;

Tasma et al., 2008), but that expression is clearly restricted to the vasculature, trichomes and guard

cells.

Page 70: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

68

Figure 2. pPLC5::GUS expression analyses in Arabidopsis seedlings and mature tissues. (a) GUS activity was present in embryo cotyledons and roots during seed germination, from testa rupture until radical emergence (28 hrs after transfer from 4°C to 22°C). GUS staining was observed in the vasculature of 2-d old- (b) and 10-d old seedlings, including leaf (c) and roots (d-g). GUS activity was also found in vascular tissue of mature 3 weeks old plants (h), trichomes (i), hydathodes (indicated by arrows) (j), guard cells (m), siliques (l) and in different parts of the flower (k), including style, filament, receptacle and pedicel.

Analysis of PPI- and PA levels in plc5-1 mutant Arabidopsis seedlings

To determine whether the knockdown-of-PLC5 caused any changes in the levels of PLC's substrates

(i.e. PIP and PIP2) or -product (i.e. the conversion of DAG into PA) (Munnik et al., 1998b; Ruelland et

al., 2002; Arisz et al., 2009, 2013), seedlings were 32Pi- labeled O/N and their lipids extracted and

analyzed. As shown in Figures 3a and 3b, no significant differences in PIP2, PIP and PA levels were

found between wt and plc5-1 seedlings. Figure 3. PPI- and PA levels in wild type- and plc5-1 seedlings. Five-day old seedlings were labelled with 32PO4

3- overnight and the next day their lipids extracted and separated by TLC. (a) Autoradiograph of a typical experiment, each lane representing 1/5th of the

extract of three seedlings. (b) Quantification of 32P-labelled PIP2-, PIP- and PA levels in wild type and plc5-1. Values are calculated as the percentage of total 32P-labelled phospholipids and are represented as means ± SD (n=3). The experiment was repeated twice with similar results.

Page 71: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

69

Overexpression of PLC5 increases tolerance to drought

Previous results revealed that overexpression of PLC3 enhanced drought tolerance (Chapter 2; Zhang et

al., 2017). Similar phenotypes were obtained earlier, when PLC was overexpressed in maize, canola or

tobacco (Wang et al., 2008; Georges et al., 2009; Tripathy et al., 2011). It is unknown whether specific

PLCs are required for this, or whether overexpression of any PLC could achieve this. Hence, we

generated transgenic plants overexpressing PLC5 under the control of the ubiquitin (UBI10) promoter.

Transgenic plants were selected from T0 to T3, and independent homozygous lines were obtained.

Two homozygous lines, PLC5-OE2 and PLC5-OE3, were selected for further studies,

overexpressing PLC5 for 300- and 100-fold, respectively (Fig. 4a). In general, overexpression of PLC5

caused a slight inhibition of their growth. Soil grown PLC-OE plants exhibited a 13.9% -17.2% shorter

inflorescence lengths and smaller rosette size (31.4%). Accordingly, the fresh weight of rosettes was

less too (~46%)(see Supplemental Fig. S3). Nonetheless, PLC5-OE lines were consistently (at least

three independent experiments) found to be more drought tolerant than wild type (Fig. 4b) and when the

water loss of detached rosettes of 4-week-old plants was analysed, it was clear that the PLC5-OE lines

lost less water that wt (Fig. 9c).

ABA plays a key role in the plant’s response to dehydration, including the induction of stomatal

closure to reduce the water loss through evaporation (Sean et al., 2010). Previously, we found that the

stomatal aperture of PLC3-OE lines was strongly reduced compared to wt in the absence of ABA, but

responded similarly to increasing ABA concentrations (Chapter 2; Zhang et al., 2017).

To check the stomatal response in PLC5-OE lines, leaf peels were isolated as before and the

ABA sensitivity analyzed. As shown in Figure 4d, stomatal opening in both PLC5-OE lines was

significantly reduced compared to wt under control conditions, like PLC3-OE lines (Chapter 2, Fig 5).

However, upon ABA treatment (0.1 µM), the stomates closed rapidly for all genotypes, but the aperture

of the two PLC5-OE lines was significantly smaller than in wild type. Above 1 µM, this difference was

maintained in PLC5-OE3, but was lost in PLC5-OE2 (Fig. 4d).

Page 72: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

70

Figure 4. Overexpression of PLC5 increases tolerance to drought. Overexpression lines were generated and the expression level of PLC5 determined by Q-PCR, relative to the expression of SAND. Values are means ± SD (n = 3) for one representative experiment. At least three experiments have been repeated with similar result. (b) Phenotype of 4-weeks old wild type- and PLC5 OE plants, grown on soil and exposed to drought by withholding water for 2 weeks. (c) Water loss of detached rosette. Water loss was measured at indicated time points and expressed as a percentage of the initial fresh weight. Values are means ± SD for one representative experiment (n=36). At least three experiments has been repeated with similar result. (d) ABA-induced stomatal closure in wild type, PLC5 OE2 (left), PLC5 OE3 (right) plants. Leaves from 3-weeks old plants were striped and incubated in opening buffer with light for 3 h until stomata were fully open. Peels were then treated with different concentration of ABA for 90 mins, after which stomata were digitized and the aperture width measured. Values are means ± SE of at least three independent experiments (n > 100). Asterisk (*) marks that PLC5-OE levels are significantly different from wild-type, based on Student’s t-test (P< 0.05).

PLC5-overexpressing plants exhibit increased PPI responses

To determine whether overexpression of PLC5 caused any changes in PPI- and/or PA levels, 32P-

labeling experiments were performed on seedlings (3h labeling) and the effect of sorbitol tested to

mimic water stress. Interestingly, under control conditions, PLC5-OE lines exhibited a clear reduction

in PIP2- and PIP levels, by about 80- and 20% respectively, and an increase in PA by 30% (Fig. 5),

indicating a constitutively higher PLC activity in vivo. O/N 32P-labeling conditions gave similar results

(Supplemental Fig. S4). Upon sorbitol treatment, however, a much stronger relative increase in PIP2

was observed in the OE lines. While PIP2 levels increased about 4-fold in wt, in the OE lines a massive,

12-fold increase was witnessed, although the absolute levels of both PIP2 and PIP remained below that

of wild type. The relative increase in PA and PIP was similar, however, for both wt and OE lines (Fig.

5b).

Page 73: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

71

Figure 5. PLC5 OE lines have altered PIP-, PIP2- and PA responses, both in control conditions and upon osmotic stress. Six-day-old seedlings were 32Pi-labeled for 3h and then treated with buffer ± 600 mM sorbitol for 30 min. Lipids were extracted, analyzed by TLC and quantified by phosphoimaging. (a) Typical TLC profile with each lane representing 1/5th of the extract of 3 seedlings. (b) 32P-levels of PIP2, PIP and PA of wild-type and PLC5 OE lines #2 and #3 under control conditions and with sorbitol. Data shown are the means ± SE (n=3) from three independent experiments. Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05).

Overexpression of PLC5 stunts root hair growth by hydrolyzing essential PIP2 at the tip Overexpression of PLC5 resulted in slightly smaller shoots in terms of inflorescence- and rosette sizes.

Also the root system was smaller, i.e. a shorter primary root length, fewer lateral roots, and shorter

lateral roots (see Supplemental Figs. S3 e-h). Analyzing the root architecture in more detail, led to

another interesting discovery, namely that root hairs were found to be stunted in the OE lines (Figs. 6a,

b). Measuring individual root hairs revealed a ~90 to ~80% reduction in root hair length for PLC5-OE2

and PLC5-OE3, respectively (Fig. 6c). Figure 6. Overexpression of PLC5 affects the growth of root hairs. (a, b) Root hair phenotypes of wild type and PLC5-OE lines #2 and #4 of 6 days old seedlings (c) Root hair-length measurements of wild-type and PLC5 OE lines. Values are means ± SE of three independent experiments (n>200). Asterisk (*) indicates significance at P<0.05 compared to wild type, based on Student’s t test. Bar = 0.5 mm.

Earlier, PIP2 was found to promote root hair elongation at the tip of growing root hairs, and that

PIP5K3 was involved in its generation(van Leeuwen et al., 2007; Kusano et al., 2008; Stenzel et al.,

2008; Grierson et al., 2014). Since the PLC5-OE lines had reduced PIP2 levels, i.e 83% in PLC5 OE2

and 77% in PLC5 OE3, which corresponded to the severity of the root hair phenotype (Fig. 6a) and

phenocopied the reduced root hair length of four independent T-DNA insertion pip5k3 mutants, we

Page 74: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

72

hypothesized that the root hair phenotype in PLC5-OE lines was caused by the continuous removal of

PIP2 from the tip that is essential for its growth. To confirm that PIP2 is a key regulator for root hair

elongation, we crossed PLC5-OE2 with an estradiol-inducible over-expressor of PIP5K3 (ER8-PIP5K3)

that is known to increase PIP2 and to induce massive root hair formation (Kusano et al., 2008). T3

transgenic of PLC5-OE2 x ER8-PIP5K3 were selected and grown together with wild type, and the

individual mutant lines, PLC5-OE2 and ER8-PIP5K3, for four days on ½MS plates and then transferred

to ½MS plates with and without 10 µM estradiol for another three days. Without estradiol, the ER8-

PIP5K3 lines showed similar root hair growth as wild type, while, the crossed line, PLC5-OE2 x ER8-

PIP5K3 clearly exhibited the reduced root hair phenotype described above (Figs. 7a and 7b). After

estradiol induction, the root hair length significantly increased in both ER8-PIP5K3 and PLC5-OE2 x

ER8-PIP5K3 lines, but did not change in wild type or PLC5-OE2 (Figs. 7a and 7b). Determining the

PIP2 levels in all above lines, revealed that without estradiol PLC5-OE2 and PLC5-OE2 ER8-PIP5K3

lines showed PIP2 levels that were significantly lower than wild type and ER8-PIP5K3 line. However,

upon induction by estradiol, PIP2 levels in both ER8-PIP5K3 and PLC5-OE2 ER8-PIP5K3 went up

sharply, while remained the same for wild type and PLC5-OE2 (Fig. 7c and 7d). Increasing PIP2 levels

by estradiol-induced overexpression of PIPK3 clearly recovered root hair growth in the PLC5-OE2

background.

Figure 7. Root hair phenotype in PLC5-OE lines is rescued by inducible overexpression of PIP5K3. (a) Root hair phenotypes of wild type, PLC5-OE2, ER8-PIP5K3 and PLC5-OE2 x ER8-PIP5K3 after estradiol induction. ER8-PIP5K3 is an estradiol-inducible overexpression line (ref). Seeds were first germinated on ½MS plates supplemented with 0.5% sucrose for 4 days, and then transferred to plates containing ± 10 µM estradiol. Seedlings were scanned three days after transferring. Bar = 0.5

mm. (b) Quantification of root hair length after estradiol induction. (c) PIP2 levels after estradiol induction. For the latter, seedlings were grown on ½MS plates with 0.5% sucrose for 4 days and then transferred to the plates containing ± 10 µM estradiol for three days, after which they were labelled overnight with 32Pi to measure the changes in PIP2 (d). Values are calculated as the percentage of total 32P-labeled phospholipids and represented as means ± SD (n=3). The experiment was repeated twice with similar results.

Page 75: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

73

DISCUSSION

Earlier, we found that loss-of-function mutants of the phloem/companion cell-specific Arabidopsis

PLC3 were affected in seed germination, root development, and ABA sensitivity, and that ectopic

overexpression resulted in plants with increased drought tolerance (Chapter 2, Zhang., et al, 2017). In

this study, PLC5 that is also expressed in phloem/companion cells, but belongs to a different subclade

of the Arabidopsis PLC family was analyzed. Knock-down of PLC5 affected primary- and lateral root

development and overexpression again enhanced the plant's tolerance to drought. In this case, PLC5

overexpression led to an additional phenotype, i.e. a strong inhibition of root hair growth. Together,

these findings underline the fact that PLCs play subtle roles in plant development and stress signaling.

How PLC achieves this, is still largely unknown, but there are several possibilities. Plants lack

the primary targets for IP3 and DAG, so these molecules are unlikely to fulfill a second messenger role,

even though we cannot completely rule this out, as plants may have evolved distinct targets than

animals. Nonetheless, there is accumulating evidence that the phosphorylated products of IP3 and DAG,

which includes various IPPs and PP-IPPs and the lipids, PA and DGPP, are acting act the plant

signaling molecules (van Schooten et al., 2006b; Arisz et al., 2009; Munnik and Vermeer, 2010;

Testerink and Munnik, 2011; Gillaspy, 2013; Hou et al., 2016). The reason for this remains unknown,

but it is striking since seedlings are quite active in PA signaling (Wang et al., 2006; Testerink and

Munnik, 2011; Hou et al., 2016).

Overexpression of PLC5 resulted in decreased PIP- and PIP2 levels and increased levels of PA,

however not of DGPP. While DGPP responses are relatively abundant in algae and cell suspensions

(Wissing et al., 1994, 1995, 1992, 1993; Wissing and Behrbohm, 1993; Munnik et al., 1996, 2000;

Pical et al., 1999; van der Luit et al., 2000; Munnik and Meijer, 2001; Meijer and Munnik, 2003; Meijer

et al., 2017), in Arabidopsis seedlings this lipid is hardly detectable, even under stress (Arisz et al.,

2013).

While PIP2 is the authentic PLC substrate in animal systems, in plants, PIP2 concentrations are

extremely low and hardly detectable in plasma membranes where most of the PLC activity is believed

to reside(Munnik et al., 1998a, 1998b; Meijer and Munnik, 2003; van Leeuwen et al., 2007; Munnik,

2014; Simon et al., 2014; Tejos et al., 2014). In contrast, PI4P is 30-100 times more abundant, highly

enriched in plasma membranes (Munnik et al., 1994; Vermeer et al., 2009; Vermeer and Munnik, 2013;

Simon et al., 2014, 2016), and is hydrolyzed equally in vitro (Munnik et al., 1998a; Munnik, 2014).

Hence, in vivo, PI4P may actually be the common PLC substrate for plants. Obviously, this may be

very different in stressed cells, where PIP2 levels do go up, e.g. in response to ABA, salt stress or heat

(Takahashi et al., 2001; DeWald et al., 2001; van Leeuwen et al., 2007; Mishkind et al., 2009; Darwish

et al., 2009; Zhang et al., 2017) or during events where local PIP2 turnover is high, but levels too low to

be visualized and measured. Indeed, PI4P levels have been reported to drop in response to stimuli, e.g.

upon salt and cold stress ( Cho et al., 1993; Pical et al., 1999; DeWald et al., 2001; Ruelland et al.,

Page 76: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

74

2002;Vermeer et al., 2009; Zarza et al., 2016), even though it still remains to be shown whether this

reflects hydrolysis by PLC or PI4P phosphatase or an activation of PIPK, for example. Nonetheless,

PLC hydrolysis of PI4P would still generate DAG and IP2 that can be converted into PA and IPPs

through DGK, IPK2 and IPK1, respectively (Munnik and Vermeer, 2010), and into PP-IPPs via VIP2

(Laha et al., 2015, 2016), to activate downstream-signaling responses.

While basal levels of PIP2 in plant cells are extremely low, this lipid is clearly emerging as a

second messenger itself, involving various stress and developmental responses, including tip growth in

pollen and root hairs, vascular differentiation, salt- and heat stress, ion-channel regulation, cytoskeletal

organization, and membrane trafficking, including endo- and exocytosis (Ischebeck et al., 2010;

Gillaspy, 2013; Rodriguez-Villalon et al., 2015; Heilmann, 2016) . Potential targets include small G-

proteins like, K+ channels, clathrin-adaptor proteins, and Exo70 (Ischebeck et al., 2010; Munnik and

Nielsen, 2011; Gillaspy, 2013; Munnik, 2014; Heilmann, 2016). Similarly, PI4P has been emerging as a

lipid second messenger too under certain conditions, and in particular cells (Stevenson et al., 2000;

Meijer and Munnik, 2003; Vermeer et al., 2009; Munnik and Nielsen, 2011; Heilmann, 2016; Simon et

al., 2016) . In such cases, PLC could also function as an attenuator of signaling. However, whether PLC

attenuates second messengers or produces them (or likely both!), clearly requires more research and

will be difficult to address anyway because of the huge redundancy of both PPI/PA pathways (e.g.

Arabidopsis has 9 PLC, 11 PIPK, 12 PIK, 7 DGK and 12 PLD genes), but also due to the highly

localized events, such as within the vascular system (phloem, companion cells) in this case.

Role for PLC5 in root development and auxin signaling?

Knock-down of PLC5 led to shorter primary roots and fewer lateral roots, which was functionally

complemented by expressing PLC5 behind its own promotor. Promotor-GUS analyses indicated that

PLC5 is predominantly expressed in the root vasculature, which was not homogenous but revealed

some form of segmentation. Moreover, lateral roots always emerged from a segment, even though not

every segment gave a lateral root, which was similar to what we found for PLC3 (Chapter 2; Zhang et

al., 2017). Interestingly, the emergence of tertiary roots revealed the same segmentation pattern,

indicating that tertiary roots initiate from PLC segments in lateral roots (Figs. 2d,e). That the lateral root

phenotype is quite mild in either plc3 (knockout) or plc5 (knockdown) mutants pointed to gene

redundancy. We therefore created plc3plc5-double mutants, but they displayed a similar reduction in

primary-root length and lateral-root number as the single mutants (Supplemental Fig. S2). These results

indicate that more PLCs are involved in this phenotype. Re-analyzing the eFP browser further, we

found two other PLCs expressed in the phloem vasculature, i.e., PLC2 and PLC7 (Chapter 4; Di Fino et

al., 2017; Van Wijk, Laxalt and Munnik, unpublished). Unfortunately, plc3plc5plc7-triple mutants or

plc2-single mutants were found to be homozygous lethal (Di Fino et al., 2017; Chapter 4; Munnik lab,

unpublished). Inducible-silencing should offer new perspectives here.

Page 77: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

75

Root growth and -branching are important events in root development and the phytohormone,

auxin plays a key role in this (Péret, De Rybel, et al., 2009; Péret, Larrieu, et al., 2009; Benková and

Bielach, 2010). Auxin regulates massive changes in gene expression by promoting the degradation of

the transcriptional repressor, Aux/IAA through its receptor, TIR1. The latter is an F-box protein that

forms an SCF complex that functions as a multi-protein E3 ubiquitin-ligase complex, which catalyzes

the ubiquitination of Aux/IAAs destined for proteasomal degradation (Kepinski and Leyser, 2005).

Interestingly, the crystal structure of TIR1 was found to contain IP6, which is anticipated to regulate

auxin binding and TIR1 activity (Tan et al., 2007; Chapter 3; Zhang et al., 2017). So far, it has

remained unknown where this IP6 would be coming from, but earlier we proposed that it could be

formed though a PLC3-generated IP2/IP3 response, with subsequent phosphorylation into IP6 at the PLC

segments from which the lateral roots emerge (Chapter 3; Zhang et al., 2017). Our data here, indicates a

similar role for PLC5. Double-plc3plc5 mutants were expected to produce less PLC-generated IP6 and

hence, exhibit reduced auxin responsiveness and root development, but likely redundant PLCs, such as

PLC2 and PLC7, take over, even though it should be noted that plc5-1 is a knock-down mutant.

Alternatively, chemical redundancy could play a role, e.g. by making IP6 via de novo synthesis (Munnik

and Vermeer, 2010; Gillaspy, 2011; 2013). To functionally couple IP6 with TIR1, it will be important

to determine the amount of IP6 in TIR1 in both wt and plc-mutant backgrounds. For the latter, we may

require induced PLC-silencing lines in combination with KD- and KO mutants.

An alternative explanation for the root phenotype in plc5 could be related to the metabolism of

inositol-based Raffinose Family Oligosaccharides (RFOs). RFOs are derived from sucrose to which a

galactosyl unit is added via galactinol (Gol). Gol is synthesised from myo-inositol and UDP-galactose,

catalyzed by galactinol synthase (GolS). Since RFOs are important for carbohydrate transport- and

storage, they could be involved in loading sugars to sink organs such as lateral roots (Van den Ende,

2013; Sengupta et al., 2015). The free myo-inositol that RFOs synthesis requires is normally generated

via glucose 6-phosphate (G6P), which forms inositol-3-phosphate (Ins3P) by myo-inositol-3-phosphate

synthase (MIPS), which is subsequently dephosphorylated by inositol mono-phosphatase (InsPase).

Theoretically, inositol could also be generated via dephosphorylation of PLC-generated IPPs (Munnik

and Vermeer, 2010), especially since PLC3 and PLC5 are both expressed at phloem companion cells.

However, like the phloem-sap analyses of plc3 mutants (Chapter 2, Chang et a., 2017), we could not

detect significant differences in plc5-1, even though myo-inositol levels tended to be a bit lower

(Supplemental Fig. S5). Nonetheless, with all this PLC redundancy and the fact that our plc5-1 mutant

is not a full KO, we cannot completely rule out this pathway yet, even though biochemically, in terms

of amounts, it is less likely. CRISPR-Cas9 generated full KO lines of PLC3, 5 and 7 could help solving

this puzzle.

Page 78: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

76

Overexpression of PLC5 enhances drought tolerance

Plants have developed many strategies to cope with drought stress (Zhu, 2002, 2016; Osakabe et al.,

2013; Mickelbart et al., 2015) and lipid signaling has been implicated in its responses ( Munnik and

Meijer, 2001; Zhu, 2002; Meijer and Munnik, 2003; Wang et al., 2007;Munnik and Vermeer, 2010;

Hou et al., 2016). Earlier, overexpression of PLC has been shown to improve drought tolerance in

maize, tobacco and canola (Wang et al., 2008; Georges et al., 2009; Tripathy et al., 2011). We

confirmed this for Arabidopsis by overexpressing AtPLC3 (Chapter 2, Zhang et al., 2017) and found it

again here for PLC5. The latter are both phloem-expressed PLCs, but we don't know whether this is a

requirement or whether the any PLC could achieve this. Overexpression of a non-phloem specific

AtPLC will shed light on this.

Under control conditions, PLC5-OE plants grew slightly slower, and eventually developed

smaller in size and produced less biomass (Supplemental Fig. S4). However, upon drought stress, they

performed much better (Figs. 4b,c), also better than PLC3-OE lines (not shown). PLC5-OE plants loose

significantly less water than wild type, which is likely related to the increased number of closed stomata

that we found under normal growth conditions (Fig. 4d), as there was no difference in the number of

stomata (not shown). The reduction in stomatal aperture could also be the reason why PLC5-OE plants

tend to be smaller, as CO2 uptake and sugar production could be affected (Cornic and Massacci, 1996).

Similarly, PLC5-OE lines led to a stunted root-hair growth (Fig. 6) (see discussion below), which could

be another explanation for their smaller size, as nutrient uptake could be affected (Leitner et al., 2009).

To validate this, we measured glucose-, fructose- and sucrose levels in shoots and roots of 10-day old

seedlings but found that PLC5-OE lines even contained slightly higher carbohydrate concentrations

(Supplemental Fig. S6). Whether this is related to their drought tolerance or a consequence of their

lower water content remains unclear.

PLC5-OE lines displayed reduced PIP- and PIP2 levels and increased PA levels, likely

reflecting increased PLC activity. Earlier, PIP2 has been proposed to be important for stomatal opening

by lowering the water potential through inhibition of SLAC1 (Lee et al., 2007) and K+-efflux channels

(Ma et al., 2009), which would favor the influx of water and open the stomata. Hence, the closed-

stomata phenotype of PLC5-OE lines could be due to the strongly reduced PIP2 levels.

In an attempt to mimic drought stress in seedlings and to analyze the effect on the PA- and PPI

levels, we used sorbitol. Interestingly, PIP2 levels dramatically increased upon this osmotic stress, and

the accumulation was much stronger in the PLC5-OE lines (Fig. 5). The latter may reflect the increased

turnover of PIP2 due to enhanced PLC5 hydrolysis that would readily be picked-up by these 32P-

labeling experiments. Since UBQ10 drives PLC5 expression in all cells, which is totally different from

the endogenous, limited PLC5 expression in the vasculature, enhanced PLC-signal formation (PA and

IPPs) or PIP2 attenuation may affect many cells, tissues and processes relevant to control responses to

osmotic stress (Munnik and Vermeer, 2010).

Page 79: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

77

Similarly, PLC5-OE lines led to a stunted root-hair growth (Fig. 6) (see discussion below), and

this could be another explanation for their smaller size, as nutrient uptake could be affected (Leitner et

al., 2009). Strikingly, these lipid changes were not observed in the PLC3-OE lines, which did not reveal

a root-hair phenotype either. It is not clear from the amino-acid sequence why PLC5 would be more

active than PLC3. All AtPLCs contain the same domain structure, i.e. two EF-hands, a catalytic and

CalB/C2 domain. So there must be subtle changes in the enzymatic properties, caused by Ca2+

sensitivity, Km or Vmax values, for example, or by interacting other proteins.

PLC-dependent PIP2 generation is essential for root hair tip growth

Overexpression of PLC5 also resulted in a very-short-root hair phenotype (Fig. 6), which is very likely

due to the low amounts of basal PIP2 required for tip growth (Fig. 5b; supplemental Fig. 6).

Interestingly, these lipid changes and root-hair phenotype were not observed in the PLC3-OE lines,

indicating differences in PLC3/5 activity or regulation. Evidence that this phenotype was coupled to

PIP2 came from the induced-overexpression of PIP5K3 in PLC5-OE2 line, which restored PIP2 levels

and rescued the root hair phenotype (Fig. 7). T-DNA insertion mutants of PIP5K3 exhibit a similar

short-root hair phenotype as PLC5-OE lines, and this lipid kinase has been shown to be responsible for

generating the PIP2 at the plasma membrane of the growing tip (van Leeuwen et al., 2007; Kusano et al.,

2008; Stenzel et al., 2008; Ischebeck et al., 2008).

Similar results have been found in tobacco pollen tubes (Ischebeck et al., 2008), whose tip

growth resembles that of root hairs( Ovečka et al., 2005; Zonia and Munnik, 2008;Ischebeck et al.,

2010; Grierson et al., 2014). Reduced PLC1 activity in petunia pollen tubes led to arrested- and

depolarized pollen tubes, which was accompanied by a disorganization of the actin cytoskeleton (Dowd

et al., 2006). On the other hand, overexpression of NtPLC3 was shown to reduce pollen tube length

(Helling et al., 2006). FP-tagged NtPLC3 was localized at the flanks of the growing tip, while the PIP2,

visualized by a biosensor, accumulated at the apex (Helling et al., 2006), similar to what was found in

growing root hairs (van Leeuwen et al., 2007; Kusano et al., 2008). We found no aberrant root hair

morphology in the plc5-1 mutant, which could be due to the fact that this is not a KO mutant and/or due

to PLC redundancy.

How PIP2 is exactly involved in this polar tip growth remains unknown, but is likely to involve

a complex signaling network between membrane trafficking and cytoskeletal dynamics (Ovečka et al.,

2005; Ischebeck et al., 2010; Munnik and Nielsen, 2011; Grierson et al., 2014). Tip growth is sustained

by exocytosis of vesicles containing growth materials, like polysaccharides and proteins, for growing

cell wall and membrane (Grierson et al., 2014). In animal cells, PIP2 is involved in priming exocytosis

and vesicle fusion by binding EXO70, a subunit form the exocyst complex (Aikawa and Martin, 2003;

Munson and Novick, 2006; Liu et al., 2007). Arabidopsis contains 23 EXO70s, and the exocyst

complex is essential for pollen tube germination and growth (Synek et al., 2006; Hála et al.,

2008)(Synek et al., 2006;Hála et al., 2008). The local accumulation of PIP2 has also been shown to

Page 80: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

78

correlate with vesical secretion (Ischebeck et al., 2008). Apart from exocytosis, endocytosis is also

important for cell growth and regulation and PIP2 has been implicated in clathrin-mediated endocytosis

(CME) as clathrin-adaptor proteins bind and are recruited by PIP2 (Zhao et al., 2010; Baisa et al., 2013;

Ischebeck et al., 2013). To deliver and return vesicles to and from the cell expanding areas, the actin

cytoskeleton also plays crucial roles and PIP2 has been suggested to participate too (Gungabissoon et al.,

1998; Wasteneys and Yang, 2004; Logan and Mandato, 2006). An important PIP2 target in regulating

this is Rho GTPases (Ras-related monomeric small G proteins), which in plant are called Rop (Rho of

plants). They are crucial regulators of tip growth (Kost, 2008) proposed to control the actin

cytoskeleton and membrane trafficking (Lee et al., 2008) and there is evidence that PIP2 regulates Rop

in organizing this (Kost et al., 1999; Klahre et al., 2006; Kost, 2008; Ischebeck et al., 2011). In

summary, it is clear that PIP2 can regulate root hair growth via various cellular processes and it is not

unlikely that PLC(s), plays a role it its attenuation and/or additional second messenger formation.

Clearly, further studies are required to decipher the molecular downstream components of PPIs, IPPs

and PA, and what the role of PLC is herein. Characterization of KO, KD and OE mutants provides a

start at least.

MATERIALS AND METHODS Plant material

Arabidopsis thaliana (Columbia-0) T-DNA insertion mutant plc5-1 (SALK_144469), was obtained

from SALK collection (signal.salk.edu). Homozygous plants were identified by PCR in F2- and F3

generations using gene-specific primers (forward primer 5’-TGGAAACTCGCAGGATATGTC-3’;

reverse primer 5’-TTGCGTCTTTGATATTCAGGG-3’) and by the combination between reverse

primer and left border primer LBb1.3 (5’-ATTTTGCCGATTTCGGAAC-3’). A double mutant,

plc3plc5 (plc3/5) was created by crossing plc3-2 (SALK_037453) and plc5-1 (SALK_144469) and

selecting homozygous lines in F2- and F3 generations. The pPLC5::GUS line was kindly provided by

Dr. Julie E. Gray (Hunt et al., 2004).

Root growth

Seeds were surface sterilized in a desiccator by using 20 ml thin bleach and 1 ml 37% HCl for 3 hours,

and then sowed on square petri dishes containing 30 ml ½-strength Murashi-Skoog (½MS) medium (pH

5.8), supplemented with 0.5 % sucrose, and 1.2 % Daishin agar (Duchefa Biochemie). Plates were

stratified at 4 °C in the dark for two nights and then transferred to long-day conditions (22 °C, 16 h of

light, 8h of dark), placed vertically under an angle of 70°. Four-day-old seedlings with comparable size

were then transferred to ½MS-ager plates without sucrose and scanned 10 - 12 days after germination

(Epson Perfection V700 scanner). Primary root length, lateral root number, and average lateral root

length was quantified for each genotype through imageJ-analysis software (National Institutes of

Page 81: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

79

Health). For root hair visualization, seedlings were grown on ½MS with 0.5% sucrose for seven days

and viewed under a stereo microscope (Leica MZFLIII) and photographed (ThorLabs CCD camera).

More than 200 root hairs per seedling were quantified and ~10 seedlings for each genotype were used

for the measurement. For inducible overexpression, seedlings were grown on ½MS with 0.5% sucrose

plate for four days and then transferred to agar medium supplemented with 10 µM β-estradiol for

another three days.

Cloning and plant transformation

MultiSite Gateway Three-Fragment Vector Construction Kit (www.lifetechnologies.com) was used to

generate pPLC5::PLC5. Oligonucleotide primers (5’-

GGGGACAACTTTGTATAGAAAAGTTGCTTTATAATAGATTAAGAAGCTTCATATC-3’ and 5’-

GGGGACTGCTTTTTTGTACAAACTTGCTCTTCAAAAAGTTCCTGCAATTTAG-3’), including

attB4 and attB1r sites, were used to PCR amplify a region of approximately 770-bp upstream of the

predicted PLC5 ATG start codon. The amplified fragment was cloned into the donor vector by using

BP Clonase II enzyme mix to create entry clone BOX1. Oligonucleotide primers (5’-

GGGGACAAGTTTGTACAAAAAAGCAGGCTATGAAGAGAGATATGGGGAGTTAC -3’ and 5’-

GGGGACCACTTTGTACAAGAAAGCTGGGTTAAGAAAGTGAAACCGCATGAGAA -3’),

including attB1 and attB2 sites, were used to PCR amplify PLC5 CDS. The amplified fragment was

cloned into the donor vector by using BP Clonase II enzyme mix to create entry clone BOX2. BOX3

was pGEM-TNOS entry clone containing attR2 and attL2 sites, which was obtained from university of

Amsterdam Plant Physiology published construct resources. The three entry clones and a destination

vector (pGreen0125) were used in MultiSite Gateway LR recombination reaction to create expression

clone (Multi gateway protocol).

To generate PLC5 overexpression line, pUBQ10::PLC5 was constructed. The PLC5 CDS was

amplified from cDNA using the following primers: AtPLC5-BsrGI-fw (5’-

GAGCTGTACAATGAAGAGAGATATGGGG-3’) and AtPLC5-T-BamHI(5’-

CGGGATCCTTAAAGAAAGTGAAACCGCATGAG-3’). The PCR product was transformed into

pJET1.2, sequenced and digested with BsrGI and BamHI. After gel extraction the BsrGI-AtPLC5-

BamHI fragment was cloned into the BsrGI/BamHI digested pGreenII0029JV-pUBQ10 mcs vector.

All constructs were transformed into Agrobacterium tumefaciens strain GV3101, which was

used to transform either Arabidopsis (Col-0) wild type plants or the plc5-1 mutant background by floral

dip (Clough and Bent, 1998). Homozygous lines were selected in T3 and used for further experiments.

RNA extraction and Q-PCR

The primer pairs to measure the PLC5 (At5g58690) expression level were: 5’-

CTTTCAACATGCAGGGCTATGGAAG-3’ and 5’- GAGATTATTGTTCATCATAAAGTCCGG-3’.

Total RNA was extracted with Trizol reagent (Invitrogen, Carlsbad, CA) as described (Pieterse, 1998).

Page 82: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

80

One- and a half µg of total RNA from 10-day-old seedlings was converted to cDNA using oilgo-dT18

primers, dNTPs, and SuperScript III Reverse Transcriptase (Invitrogen), according to the

manufacturer’s instructions. Q-PCR was performed with ABI 7500 Real-Time PCR System (Applied

Biosystem). The relative gene expression was determined by comparative threshold cycle value.

Transcript levels were normalized by the level of SAND (At2g28390; forward primer: 5’-AAC TCT

ATG CAG CAT TTG ATC CAC T-3’, reverse primer: 5’-TGA TTG CAT ATC TTT ATC GCC ATC-

3’). Three biological replicates and two technical replicates were used for the values of means and

standard deviations (Han et al., 2013).

Histochemical analyses for GUS activity

Transgenic plants carrying pPLC5::GUS were grown for the times indicated in the legend. Specific

tissues were taken and incubated in X-Gluc reaction solution containing 1mg/ml 5-bromo-4-chloro-3

indolyl-β-D-glucuronic acid (X-gluc), 50 mM phosphate buffer pH 7.0 and 0.1% TX-100 and incubated

overnight at 37°C as described before (Chapter 2; Zhang et al., 2017). The next day, solutions were

replaced by 70% ethanol to de-stain the tissue after which the material was analysed under a stereo

microscope (Leica MZFLIII) and photographed (ThorLabs CCD camera).

Stomatal aperture

Stomatal-aperture measurements were performed according to Distéfano et al., (2012) with minor

changes. Epidermal strips were excised from the abaxial side of fully expanded Arabidopsis leaves of 3-

week-old plants grown at 22°C under 16 h of light and 8h of dark, and immediately floated in opening

buffer (5 mM MES-KOH, pH 6.1, and 50 mM KCl) for 3 h. Strips were then treated with (0-10 µM)

Stomata were digitized using a Nikon DS-Fi 1 camera, coupled to a Nikon Eclipse Ti microscope. The

stomatal-aperture width was subsequently measured using ImageJ software (National Institute of

Health).

32Pi-phospholipid labelling, extraction and analysis

Different types of tissues were used, whole seedlings and epidermal leaf peels. For Seedlings: Five-day-

old seedlings were transferred to 200 µl labeling buffer (2.5 mM MES-KOH, pH 5.8, 1 mM KCl)

containing 32Pi (5-10 µCi) in 2 ml Eppendorf tubes and labeled overnight (~16 h). Samples were treated

the next day by adding 200 µl labeling buffer with or without sorbitol for the times and concentrations

indicated. For epidermal leaf peels: Leaves of 3-week-old plants grown at 22°C under 16 h of light and

8 h of dark were stripped and immediately floated on 100 µl opening buffer (10 mM MES, pH 6.1 and

50 mM KCl) containing 32Pi (5-10 µCi) in a 48-wells cell culture plate (Greiner bio-one) for 3 h. All

treatments were stopped by adding perchloric acid at a final concentration of 5% (v/v) for 5-10 min,

after which the material was transferred to 400 µl of CHCl3/MeOH/HCl [50:100:1 v/v)] to extract the

Page 83: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

81

lipids. After 15 min, 400 µl of CHCl3 was added, followed by 200 µl of 0.9 % (w/v) NaCl to induce two

phases. The organic lipid fractions were washed and concentrated as described earlier (Munnik & Zarza,

2013). Lipids were separated by thin-layer chromatography (TLC) using an alkaline solvent system,

containing: chloroform/methanol/28% ammonia/water [90:70:4:16 (v/v)] (Munnik et al., 1994).

Radioactive phospholipids were visualized by autoradiography and quantified by phosphoimaging

(Molecular Dynamics, Sunnyvale, CA, USA). Individual phospholipid levels are expressed as the

percentage of the total 32P-lipid fraction.

Drought tolerance

Drought assays were performed as described earlier (Hua et al., 2012; Osakabe et al., 2013) with some

modifications. Seeds were stratified at 4°C in the dark for 2 nights and sown in soil. Each pot (4.5 x 4.5

x 7.5 cm) contained 80 g of soil and nine plants, which were grown under short day conditions at 22 °C

with 12 h light/12 h dark for 4 weeks, and then subjected to dehydration by withholding them for water

for 2 weeks, while control plants were normally watered. Each experiment used 36 plants per genotype

and experiments were repeated at least 3 times.

To assay the water-loss, rosettes from 4-week-old plants were detached and the fresh weight

(FW) determined by weighing them every hour. Water content was calculated as a percentage from the

initial FW. Twenty plants were used for each experiment and each experiment was repeated at least 3

times.

Soluble carbohydrates measurement in seedlings

Soluble carbohydrates were determined as described before (Vergauwen et al., 2000) with

some minor modifications. Ten-day old Arabidopsis seedlings, grown on ½MS plates without

sucrose, were separated in shoot- and root parts and immediately frozen in liquid nitrogen.

Samples were then grinded and their fresh weights recorded for further analysis. Soluble sugars

were extracted in water and immediately boiled in a water bath. Carbohydrates were separated

by anion-exchange chromatography and quantified by pulsed-amperometric detec- tion

(Dionex, Sunnyvale Ca, USA). Mannitol was used as an internal standard. Factors for Glc, Fru,

Suc were obtained by injecting pure compounds. A CarbopacTM PA-100 guard and CarbopacTM

PA-100 (4×250) in series were equilibrated with 90 mM NaOH for 24 min. Regeneration was

5 min with 500 mM Na-acetate and 10 min with 500 mM NaOH. Values shown are the means

± SD (n=3) for one representative experiment that was repeated twice.

Page 84: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

82

Phloem sap soluble carbohydrates measurement Phloem exudates were extracted and analyzed as described earlier (Roessner et al., 2000; Greco et al.,

2012; Tetyuk et al., 2013). The hydrophilic fraction (sugars, amino acids, small molecules) was dried

and derivatized using methoxyamine in pyridine and BSTFA (N,O-bis-trimethyl-silyltrifluoroacetamide)

with 1% trimethylchlorosilane. The metabolites were subsequently separated using an Agilent 5890N

GC system, coupled to a 5973 inert MSD using a DB-5MS column (J&W Scientific, 30m x 0.32mm ID

x 0.25 µm film; temperature program: 70-5, 5/min-320-5). Identification of compounds occurred using

the NIST library (NIST Standard Reference Database; https://www.nist.gov/srd/nist-standard-reference-

database-1a-v14) in combination with co-elution of standards. Peak area determination was performed

using QuanLynx, a quantification software within the MassLynx software (Waters). The experiment

was repeated five times.

ACKNOWLEDGEMENTS

This work was financially supported by the China Scholarship Council (CSC File No. 201206300058 to

Q.Z) and by the Netherlands Organization for Scientific Research (NWO; 867.15.020 to T.M.).

REFERENCES Abd-El-Haliem AM, Vossen JH, van Zeijl A, Dezhsetan S, Testerink C, Seidl MF, Beck M, Strutt J, Robatzek S, and Joosten

MHAJ (2016) Biochemical characterization of the tomato phosphatidylinositol-specific phospholipase C (PI-PLC) family and its role in plant immunity. Biochim Biophys Acta - Mol Cell Biol Lipids 1861:1365–1378.

Aikawa Y, and Martin TFJ (2003) ARF6 regulates a plasma membrane pool of phosphatidylinositol(4,5)bisphosphate required for regulated exocytosis. J Cell Biol 162:647–659.

Allen GJ, and Sanders D (1994) Osmotic stress enhances the competence of Beta vulgaris vacuoles to respond to inositol 1,4,5-trisphosphate. Plant J 6:687–695.

Arisz SA, and Munnik T (2013) Distinguishing phosphatidic acid pools from De Novo synthesis, PLD, and DGK, in Plant Lipid Signaling Protocols (Munnik T, and Heilmann I eds) pp 55–62.

Arisz SA, Testerink C, and Munnik T (2009) Plant PA signaling via diacylglycerol kinase. Biochim Biophys Acta 1791:869–875.

Arisz SA, van Wijk R, Roels W, Zhu JK, Haring MA, and Munnik T (2013) Rapid phosphatidic acid accumulation in response to low temperature stress in Arabidopsis is generated through diacylglycerol kinase. Front Plant Sci 4:1–15.

Baisa GA, Mayers JR, and Bednarek SY (2013) Budding and braking news about clathrin-mediated endocytosis. Curr Opin Plant Biol 16:718–725.

Balla T (2013) Phosphoinositides: tiny lipids with giant impact on cell regulation. Physiol Rev 93:1019–1137. Benková E, and Bielach A (2010) Lateral root organogenesis - from cell to organ. Curr Opin Plant Biol 13:677–683. Blatt MR, Thiel G, and Trentham DR (1990) Reversible inactivation of K+ channels of Vicia stomatal guard cells following

the photolysis of caged inositol 1,4,5-trisphosphate. Nature 346:766–769. Boss WF, Sederoff HW, Im YJ, Moran N, Grunden AM, and Perera IY (2010) Basal Signaling Regulates Plant Growth and

Development. Plant Physiol 154:439–443. Cho MH, Shears SB, and Boss WF (1993) Changes in phosphatidylinositol metabolism in response to hyperosmotic stress in

Daucus carota L. cells grown in suspension culture. Plant Physiol 103:637–647. Clough SJ, and Bent AF (1998) Floral dip: A simplified method for Agrobacterium-mediated transformation of Arabidopsis

thaliana. Plant J 16:735–743. Cornic G, and Massacci A (1996) Leaf Photosynthesis Under Drought Stress, in Photosynthesis and the Environment (Baker

NR ed) pp 347–366. D’Ambrosio JM, Couto D, Fabro G, Scuffi1 D, Lamattina L, Munnik T, Álvarez ME, Andersson MX, Zipfel C, and Laxalt

AM (2017) Phosphoinositide-specific phospholipase C2 (PLC2)associates with RBOHD and modulates ROS production and MAMP-triggered immunity in Arabidopsis. Plant J submitted.

Darwish E, Testerink C, Khalil M, El-Shihy O, and Munnik T (2009) Phospholipid signaling responses in salt-stressed rice leaves. Plant Cell Physiol 50:986–997.

Das S, Hussain A, Bock C, Keller WA, and Georges F (2005) Cloning of Brassica napus phospholipase C2 (BnPLC2), phosphatidylinositol 3-kinase (BnVPS34) and phosphatidylinositol synthase1 (BnPtdIns S1) - Comparative analysis of

Page 85: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

83

the effect of abiotic stresses on the expression of phosphatidylinositol signal transdu. Planta 220:777–784. De Jong CF, Laxalt AM, Bargmann BOR, De Wit PJGM, Joosten MHAJ, and Munnik T (2004) Phosphatidic acid

accumulation is an early response in the Cf-4/Avr4 interaction. Plant J 39:1–12. DeWald DB, Torabinejad J, Jones C a, Shope JC, Cangelosi a R, Thompson JE, Prestwich GD, and Hama H (2001) Rapid

accumulation of phosphatidylinositol 4,5-bisphosphate and inositol 1,4,5-trisphosphate correlates with calcium mobilization in salt-stressed arabidopsis. Plant Physiol 126:759–769.

Distéfano AM, Scuffi D, García-Mata C, Lamattina L, and Laxalt AM (2012) Phospholipase D?? is involved in nitric oxide-induced stomatal closure. Planta 236:1899–1907.

Dowd PE, Coursol S, Skirpan AL, Kao T, and Gilroy S (2006) Petunia phospholipase C1 is involved in pollen tube growth. Plant Cell 18:1438–1453.

Gao K, Liu YL, Li B, Zhou RG, Sun DY, and Zheng SZ (2014) Arabidopsis thaliana phosphoinositide-specific phospholipase C isoform 3 (AtPLC3) and AtPLC9 have an additive effect on thermotolerance. Plant Cell Physiol 55:1873–1883.

Georges F, Das S, Ray H, Bock C, Nokhrina K, Kolla VA, and Keller W (2009) 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 32:1664–1681.

Gillaspy GE (2013) The role of phosphoinositides and inositol phosphates in plant cell signaling, in Lipid-mediated protein signaling (Capelluto DGS ed) pp 141–157.

Gilroy S, Read ND, and Trewavas a J (1990) Elevation of cytoplasmic calcium by caged calcium or caged inositol triphosphate initiates stomatal closure. Nature 346:769–771.

Gonorazky G, Guzzo MC, and Laxalt AM (2016) Silencing of the tomato phosphatidylinositol-phospholipase C2 (SlPLC2) reduces plant susceptibility to Botrytis cinerea. Mol Plant Pathol 2:1–10.

Gonorazky G, Ramirez L, Abd-El-Haliem A, Vossen JH, Lamattina L, ten Have A, Joosten MHAJ, and Laxalt AM (2014) The tomato phosphatidylinositol-phospholipase C2 (SlPLC2) is required for defense gene induction by the fungal elicitor xylanase. J Plant Physiol 171:959–965.

Greco M, Chiappetta A, Bruno L, and Bitonti MB (2012) In Posidonia oceanica cadmium induces changes in DNA methylation and chromatin patterning. J Exp Bot 63:695–709.

Grierson C, Nielsen E, Ketelaar T, and Schiefelbein J (2014) Root Hairs. Arab B 12:1–25. Gungabissoon RA, Jiang CJ, Drebak BK, Maciver SK, and Hussey PJ (1998) Interaction of maize actin-depolymerising factor

with actin and phosphoinositides and its inhibition of plant phospholipase C. Plant J 16:689–696. Hála M, Cole R, Synek L, Drdová E, Pečenková T, Nordheim A, Lamkemeyer T, Madlung J, Hochholdinger F, Fowler JE,

and Žárský V (2008) An Exocyst Complex Functions in Plant Cell Growth in Arabidopsis and Tobacco. Plant Cell 20:1330–1345.

Han B, Yang Z, Samma MK, Wang R, and Shen W (2013) Systematic validation of candidate reference genes for qRT-PCR normalization under iron deficiency in Arabidopsis. BioMetals 26:403–413.

Hartog M Den, Verhoef N, and Munnik T (2003) Nod factor and elicitors activate different phospholipid signaling pathways in suspension-cultured Alfalfa cells. Plant Physiol 132:311–317.

Heilmann I (2016) Phosphoinositide signaling in plant development. Development 143:2044–2055. Helling D, Possart A, Cottier S, Klahre U, and Kost B (2006) Pollen tube tip growth depends on plasma membrane

polarization mediated by tobacco PLC3 activity and endocytic membrane recycling. Plant Cell 18:3519–3534. Hirayama T, Ohtot C, Mizoguchi T, and Shinozaki K (1995) A gene encoding a phosphatidylinositol-specific phospholipase C

is induced by dehydration and salt stress in Arabidopsis thaliana. Proc Natl Acad Sci USA 92:3903–3907. Horvath I, Glatz A, Nakamoto H, Mishkind ML, Munnik T, Saidi Y, Goloubinoff P, Harwood JL, and Vigh L (2012) Heat

shock response in photosynthetic organisms: Membrane and lipid connections. Prog Lipid Res 51:208–220. Hou Q, Ufer G, and Bartels D (2016) Lipid signalling in plant responses to abiotic stress. Plant, Cell Environ 39:1029–1048. Hua D, Wang C, He J, Liao H, Duan Y, Zhu Z, Guo Y, Chen Z, and Gong Z (2012) A plasma membrane receptor kinase,

GHR1, mediates abscisic acid- and hydrogen peroxide-regulated stomatal movement in Arabidopsis. Plant Cell 24:2546–2561.

Hunt L, and Gray JE (2001) ABA signalling: A messenger’s FIERY fate. Curr Biol 11:968–970. Hunt L, Otterhag L, Lee JC, Lasheen T, Hunt J, Seki M, Shinozaki K, Sommarin M, Gilmour DJ, Pical C, and Gray JE (2004)

Gene-specific expression and calcium activation of Arabidopsis thaliana phospholipase C isoforms. New Phytol 162:643–654.

Irvine RF (2006) Nuclear inositide signalling-expansion, structures and clarification. Biochim Biophys Acta 1761:505–508. Ischebeck T, Seiler S, and Heilmann I (2010) At the poles across kingdoms: Phosphoinositides and polar tip growth.

Protoplasma 240:13–31. Ischebeck T, Stenzel I, and Heilmann I (2008) Type B phosphatidylinositol-4-phosphate 5-kinases mediate Arabidopsis and

Nicotiana tabacum pollen tube growth by regulating apical pectin secretion. Plant Cell 20:3312–3330. Ischebeck T, Stenzel I, Hempel F, Jin X, Mosblech A, and Heilmann I (2011) Phosphatidylinositol-4,5-bisphosphate

influences Nt-Rac5-mediated cell expansion in pollen tubes of Nicotiana tabacum. Plant J 65:453–468. Ischebeck T, Werner S, Krishnamoorthy P, Lerche J, Meijón M, Stenzel I, Löfke C, Wiessner T, Im YJ, Perera IY, Iven T,

Feussner I, Busch W, Boss WF, Teichmann T, Hause B, Persson S, and Heilmann I (2013) Phosphatidylinositol 4,5-bisphosphate influences PIN polarization by controlling clathrin-mediated membrane trafficking in Arabidopsis. Plant Cell 25:4894–4911.

Kepinski S, and Leyser O (2005) The Arabidopsis F-box protein TIR1 is an auxin receptor. Nature 435:446–451. Klahre U, Becker C, Schmitt AC, and Kost B (2006) Nt-RhoGDI2 regulates Rac/Rop signaling and polar cell growth in

tobacco pollen tubes. Plant J 46:1018–1031. Kost B (2008) Spatial control of Rho (Rac-Rop) signaling in tip-growing plant cells. Trends Cell Biol 18:119–127.

Page 86: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

84

Kost B, Lemichez E, Spielhofer P, Hong Y, Tolias K, Carpenter C, and Chua N (1999) Tube Growth. 145:317–330. Kuo HF, Chang TY, Chiang SF, Wang W Di, Charng YY, and Chiou TJ (2014) Arabidopsis inositol pentakisphosphate 2-

kinase, AtIPK1, is required for growth and modulates phosphate homeostasis at the transcriptional level. Plant J 80:503–515.

Kusano H, Testerink C, Vermeer JEM, Tsuge T, Shimada H, Oka A, Munnik T, and Aoyama T (2008) The Arabidopsis phosphatidylinositol phosphate 5-kinase PIP5K3 is a key regulator of root hair tip growth. Plant Cell 20:367–380.

Laha D, Johnen P, Azevedo C, Dynowski M, Wei M, Capolicchio S, Mao H, Iven T, Steenbergen M, Freyer M, Gaugler P, de Campos MKF, Zheng N, Feussner I, Jessen HJ, Van Wees SCM, Saiardi A, and Schaaf G (2015) VIH2 regulates the synthesis of inositol pyrophosphate InsP8 and jasmonate-dependent defenses in Arabidopsis. Plant Cell 27:1082–1097.

Laha D, Parvin N, Dynowski M, Johnen P, Mao H, Bitters ST, Zheng N, and Schaaf G (2016) Inositol polyphosphate binding specificity of the jasmonate receptor complex. Plant Physiol 171:2364–2370.

Lanteri ML, Lamattina L, and Laxalt AM (2011) Mechanisms of xylanase-induced nitric oxide and phosphatidic acid production in tomato cells. Planta 234:845–855.

Laxalt AM, Raho N, Ten Have A, and Lamattina L (2007) Nitric oxide is critical for inducing phosphatidic acid accumulation in xylanase-elicited tomato cells. J Biol Chem 282:21160–21168.

Lee HS, Lee DH, Cho HK, Kim SH, Auh JH, and Pai HS (2015) 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 27:417–431.

Lee JY, Szumlanski A, Nielsen E, and Yang Z (2008) Rho-GTPase-dependent filamentous actin dynamics coordinate vesicle targeting and exocytosis during tip growth. J Cell Biol 181:1155–1168.

Lee Y, Kim YW, Jeon BW, Park KY, Suh SJ, Seo J, Kwak JM, Martinoia E, Hwang I, and Lee Y (2007) Phosphatidylinositol 4,5-bisphosphate is important for stomatal opening. Plant J 52:803–816.

Leitner D, Klepsch S, Ptashnyk M, Marchant A, Kirk GJD, Schnepf A, and Roose T (2009) A dynamic model of nutrient uptake by root hairs. New Phytol 9999:792–802.

Lemtiri-Chlieh F, MacRobbie EAC, and Brearley CA (2000) Inositol hexakisphosphate is a physiological signal regulating the K+-inward rectifying conductance in guard cells. Proc Natl Acad Sci U S A 97:8687–8692.

Lemtiri-Chlieh F, MacRobbie EAC, Webb AAR, Manison NF, Brownlee C, Skepper JN, Chen J, Prestwich GD, and Brearley CA (2003) Inositol hexakisphosphate mobilizes an endomembrane store of calcium in guard cells. Proc Natl Acad Sci U S A 100:10091–10095.

Li L, He Y, Wang Y, Zhao S, Chen X, Ye T, Wu Y, and Wu Y (2015) Arabidopsis PLC2 is involved in auxin-modulated reproductive development. Plant J 84:504–515.

Liu HT, Huang WD, Pan QH, Weng FH, Zhan JC, Liu Y, Wan SB, and Liu YY (2006) Contributions of PIP2-specific-phospholipase C and free salicylic acid to heat acclimation-induced thermotolerance in pea leaves. J Plant Physiol 163:405–416.

Liu J, Zuo X, Yue P, and Guo W (2007) Phosphatidylinositol 4,5-bisphosphate mediates the targeting of the exocyst to the plasma membrane for exocytosis in mammalian cells. Mol Biol Cell 18:4483–4492.

Logan MR, and Mandato CA (2006) Regulation of the actin cytoskeleton by PIP2 in cytokinesis. Biol Cell 98:377–388. Luit AH Van Der, Piatti T, Doorn A Van, Musgrave A, Felix G, Boller T, and Munnik T (2000) Elicitation of Suspension-

Cultured Tomato Cells Triggers the Formation of Phosphatidic Acid and Diacylglycerol Pyrophosphate 1. Plant Physiol 123:1507–1515.

Ma Y, Szostkiewicz I, Korte A, Moes D, Yang Y, Christmann A, and Grill E (2009) Regulators of PP2C phosphatase activity function as abscisic acid sensors. Science 324:1064–1069.

Meijer HJG, and Munnik T (2003) Phospholipid-based signaling in plants. Annu Rev Plant Biol 54:265–306. Meijer HJG, van Himbergen JAJ, Musgrave A, and Munnik T (2017) Acclimation to salt modifies the activation of several

osmotic stress-activated lipid signalling pathways in Chlamydomonas. Phytochemistry 135:64–72. Michell RH (2008) Inositol derivatives: evolution and functions. Nat Rev 9:151–161. Mickelbart M V., Hasegawa PM, and Bailey-Serres J (2015) Genetic mechanisms of abiotic stress tolerance that translate to

crop yield stability. Nat Rev Genet 16:237–251. Mishkind M, Vermeer JEM, Darwish E, and Munnik T (2009) Heat stress activates phospholipase D and triggers PIP

accumulation at the plasma membrane and nucleus. Plant J 60:10–21. Mueller-roeber B, and Pical C (2002) Inositol phospholipid metabolism in Arabidopsis . characterized and putative isoforms of

inositol phospholipid kinase and phosphoinositide-specific phospholipase C. Plant Physiol 130:22–46. Munnik T (2001) Phosphatidic acid: an emerging plant lipid second messenger. Trends Plant Sci 6:227–233. Munnik T (2014) PI-PLC: Phosphoinositide-phospholipase C in plant signalling, in Phospholipases in Plant Signaling (Wang

X ed) pp 27–54. Munnik T, De Vrije T, Irvine RF, and Musgrave A (1996) Identification of diacylglycerol pyrophosphate as a novel metabolic

product of phosphatidic acid during G-protein activation in plants. J Biol Chem 271:15708–15715. Munnik T, Irvine RF, and Musgrave A (1998) Phospholipid signalling in plants. Biochim Biophys Acta - Lipids Lipid Metab

1389:222–272. Munnik T, and Meijer HJ. (2001) Osmotic stress activates distinct lipid and MAPK signalling pathways in plants. FEBS Lett

498:172–178. Munnik T, Meijer HJG, Riet B Ter, Hirt H, Frank W, Bartels D, and Musgrave A (2000) Hyperosmotic stress stimulates

phospholipase D activity and elevates the levels of phosphatidic acid and diacylglycerol pyrophosphate. Plant J 22:147–154.

Munnik T, Musgrave A, and de Vrije T (1994) Rapid turnover of polyphosphoinositides in carnation flower petals. Planta 193:89–98.

Munnik T, and Nielsen E (2011) Green light for polyphosphoinositide signals in plants. Curr Opin Plant Biol 14:489–497. Munnik T, and Testerink C (2009) Plant phospholipid signaling: “in a nutshell”. J Lipid Res 50:S260–S265.

Page 87: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

85

Munnik T, van Himbergen JAJ, ter Riet B, Braun F-J, Irvine RF, van den Ende H, and Musgrave A (1998) Detailed analysis of the turnover of polyphosphoinositides and phosphatidic acid upon activation of phospholipases C and D in Chlamydomonas cells treated with non-permeabilizing concentrations of mastoparan. Planta 207:133–145.

Munnik T, and Vermeer JEM (2010) Osmotic stress-induced phosphoinositide and inositol phosphate signalling in plants. Plant Cell Environ 33:655–669.

Munson M, and Novick P (2006) The exocyst defrocked, a framework of rods revealed. Nat Struct Mol Biol 13:577–581. Murphy AM, Otto B, Brearley CA, Carr JP, and Hanke DE (2008) A role for inositol hexakisphosphate in the maintenance of

basal resistance to plant pathogens. Plant J 56:638–652. Osakabe Y, Arinaga N, Umezawa T, Katsura S, Nagamachi K, Tanaka H, Ohiraki H, Yamada K, Seo S-U, Abo M, Yoshimura

E, Shinozaki K, and Yamaguchi-Shinozaki K (2013) Osmotic stress responses and plant growth controlled by potassium transporters in Arabidopsis. Plant Cell 25:609–624.

Ovečka M, Lang I, Baluška F, Ismail A, Illeš P, and Lichtscheidl IK (2005) Endocytosis and vesicle trafficking during tip growth of root hairs. Protoplasma 226:39–54.

Perera IY, Heilmann I, and Boss WF (1999) Transient and sustained increases in inositol 1,4,5-trisphosphate precede the differential growth response in gravistimulated maize pulvini. Proc Natl Acad Sci USA 96:5838–5843.

Perera IY, Hung C, Brady S, Muday GK, and Boss WF (2006) A Universal Role for Inositol 1, 4, 5-Trisphosphate-Mediated Signaling in Plant Gravitropism. Plant Physiol 140:746–760.

Péret B, De Rybel B, Casimiro I, Benková E, Swarup R, Laplaze L, Beeckman T, and Bennett MJ (2009) Arabidopsis lateral root development: an emerging story. Trends Plant Sci 14:399–408.

Péret B, Larrieu A, and Bennett MJ (2009) Lateral root emergence: A difficult birth. J Exp Bot 60:3637–3643. Pical C, Westergren T, Dove SK, Larsson C, and Sommarin M (1999) Salinity and hyperosmotic stress induce rapid increases

in phosphatidylinositol 4,5-bisphosphate, diacylglycerol pyrophosphate, and phosphatidylcholine in Arabidopsis thaliana cells. J Biol Chem 274:38232–38240.

Pieterse CMJ (1998) A novel signaling pathway controlling induced systemic resistance in Arabidopsis. Plant Cell 10:1571–1580.

Pokotylo I, Kolesnikov Y, Kravets V, Zachowski A, and Ruelland E (2014) Plant phosphoinositide-dependent phospholipases C: Variations around a canonical theme. Biochimie 96:144–157.

Puga MI, Mateos I, Charukesi R, Wang Z, Franco-Zorrilla JM, de Lorenzo L, Irigoyen ML, Masiero S, Bustos R, Rodríguez J, Leyva A, Rubio V, Sommer H, and Paz-Ares J (2014) SPX1 is a phosphate-dependent inhibitor of Phosphate Starvation Response 1 in Arabidopsis. Proc Natl Acad Sci U S A 111:14947–14952.

Raho N, Ramirez L, Lanteri ML, Gonorazky G, Lamattina L, ten Have A, and Laxalt AM (2011) Phosphatidic acid production in chitosan-elicited tomato cells, via both phospholipase D and phospholipase C/diacylglycerol kinase, requires nitric oxide. J Plant Physiol 168:534–539.

Repp A, Mikami K, Mittmann F, and Hartmann E (2004) Phosphoinositide-specific phospholipase C is involved in cytokinin and gravity responses in the moss Physcomitrella patens. Plant J 40:250–259.

Rodriguez-Villalon A, Gujas B, van Wijk R, Munnik T, and Hardtke CS (2015) Primary root protophloem differentiation requires balanced phosphatidylinositol-4,5-biphosphate levels and systemically affects root branching. Development 142:1437–1446.

Roessner U, Wagner C, Kopka J, Trethewey RN, and Willmitzer L (2000) Simultaneous analysis of metabolites in potato tuber by gas chromatography-mass spectrometry. Plant J 23:131–142.

Ruelland E, Cantrel C, Gawer M, Kader JC, and Zachowski A (2002) Activation of phospholipases C and D is an early response to a cold exposure in Arabidopsis suspension cells. Plant Physiol 130:999–1007.

Sean R, Pedro L, Ruth R, and Suzanne R (2010) Abscisic acid:emergence of a core signaling network. Annu Rev Plant Biol 61:651–679.

Sengupta S, Mukherjee S, Basak P, and Majumder AL (2015) significance of galactinol and raffinose family oligosaccharide synthesis in plants. Front Plant Sci 6:1–11.

Sheard LB, Tan X, Mao H, Withers J, Ben-Nissan G, Hinds TR, Kobayashi Y, Hsu F-F, Sharon M, Browse J, He SY, Rizo J, Howe GA, and Zheng N (2010) Jasmonate perception by inositol-phosphate-potentiated COI1-JAZ co-receptor. Nature 468:400–405.

Simon MLA, Platre MP, Assil S, Van Wijk R, Chen WY, Chory J, Dreux M, Munnik T, and Jaillais Y (2014) A multi-colour/multi-affinity marker set to visualize phosphoinositide dynamics in Arabidopsis. Plant J 77:322–337.

Simon ML, MP P, MM M-B, L A, T S, V B, MC C, and Jaillais Y (2016) A PtdIns(4)P-driven electrostatic field controls cell membrane identity and signalling in plants. Nat plants 2:1–10.

Skinner DZ, Bellinger BS, Halls S, Baek K-H, Garland-Campbell K, and Siems WF (2005) Phospholipid Acyl Chain and Phospholipase Dynamics during Cold Acclimation of Winter Wheat. Crop Sci 45:1858–1867.

Song M, Liu S, Zhou Z, and Han Y (2008) TfPLC1, a gene encoding phosphoinositide-specific phospholipase C, is predominantly expressed in reproductive organs in Torenia fournieri. Sex Plant Reprod 21:259–267.

Stenzel I, Ischebeck T, König S, Hołubowska A, Sporysz M, Hause B, and Heilmann I (2008) The type B phosphatidylinositol-4-phosphate 5-kinase 3 is essential for root hair formation in Arabidopsis thaliana. Plant Cell 20:124–141.

Stevenson JM, Perera IY, Heilmann I, Persson S, and Boss WF (2000) Inositol signaling and plant growth. Trends Plant Sci 5:252–258.

Sui Z, Niu L, Yue G, Yang A, and Zhang J (2008) Cloning and expression analysis of some genes involved in the phosphoinositide and phospholipid signaling pathways from maize (Zea mays L.). Gene 426:47–56.

Synek L, Schlager N, Eliáš M, Quentin M, Hauser MT, and Žárský V (2006) AtEXO70A1, a member of a family of putative exocyst subunits specifically expanded in land plants, is important for polar growth and plant development. Plant J 48:54–72.

Page 88: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

86

Takahashi S, Katagiri T, Hirayama T, Yamaguchi-Shinozaki K, and Shinozaki K (2001) Hyperosmotic stress induces a rapid and transient increase in inositol 1,4,5-trisphosphate independent of abscisic acid in Arabidopsis cell culture. Plant Cell Physiol 42:214–222.

Tan X, Calderon-Villalobos LI a, Sharon M, Zheng C, Robinson C V, Estelle M, and Zheng N (2007) Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature 446:640–645.

Tasma IM, Brendel V, Whitham S a., and Bhattacharyya MK (2008) Expression and evolution of the phosphoinositide-specific phospholipase C gene family in Arabidopsis thaliana. Plant Physiol Biochem 46:627–637.

Tejos R, Sauer M, Vanneste S, Palacios-Gomez M, Li H, Heilmann M, van Wijk R, Vermeer JEM, Heilmann I, Munnik T, and Friml J (2014) Bipolar Plasma Membrane Distribution of Phosphoinositides and Their Requirement for Auxin-Mediated Cell Polarity and Patterning in Arabidopsis. Plant Cell 26:2114–2128.

Testerink C, and Munnik T (2011) Molecular, cellular, and physiological responses to phosphatidic acid formation in plants. J Exp Bot 62:2349–2361.

Testerink C, and Munnik T (2005) Phosphatidic acid: a multifunctional stress signaling lipid in plants. Trends Plant Sci 10:368–375.

Tetyuk O, Benning UF, and Hoffmann-Benning S (2013) Collection and Analysis of Arabidopsis Phloem Exudates Using the EDTA-facilitated Method. J Vis Exp 1–11.

Tripathy MK, Tyagi W, Goswami M, Kaul T, Singla-Pareek SL, Deswal R, Reddy MK, and Sopory SK (2011) Characterization and Functional Validation of Tobacco PLC Delta for Abiotic Stress Tolerance. Plant Mol Biol Report 30:488–497.

Van den Ende W (2013) Multifunctional fructans and raffinose family oligosaccharides. Front Plant Sci 4:1–11. van der Luit a H, Piatti T, van Doorn a, Musgrave a, Felix G, Boller T, and Munnik T (2000) Elicitation of suspension-

cultured tomato cells triggers the formation of phosphatidic acid and diacylglycerol pyrophosphate. Plant Physiol 123:1507–1516.

van Leeuwen W, Vermeer JEM, Gadella TWJ, and Munnik T (2007) Visualization of phosphatidylinositol 4,5-bisphosphate in the plasma membrane of suspension-cultured tobacco BY-2 cells and whole Arabidopsis seedlings. Plant J 52:1014–1026.

van Schooten B, Testerink C, and Munnik T (2006a) Signalling diacylglycerol pyrophosphate, a new phosphatidic acid metabolite. Biochim Biophys Acta - Mol Cell Biol Lipids 1761:151–159.

van Schooten B, Testerink C, and Munnik T (2006b) Signalling diacylglycerol pyrophosphate, a new phosphatidic acid metabolite. Biochim Biophys Acta 1761:151–159.

Vergauwen R, Van den Ende W, and Van Laere a (2000) The role of fructan in flowering of Campanula rapunculoides. J Exp Bot 51:1261–1266.

Vergnolle C, Vaultier MN, Taconnat L, Renou JP, Kader JC, Zachowski A, and Ruelland E (2005) The cold-induced early activation of phospholipase C and D pathways determines the response of two distinct clusters of genes in Arabidopsis cell suspensions. Plant Physiol 139:1217–1233.

Vermeer JEM, and Munnik T (2013) Using Genetically Encoded Fluorescent Reporters to Image Lipid Signalling in Living Plants, in Plant Lipid Signaling Protocols (Munnik T, and Heilmann I eds) pp 283–289.

Vermeer JEM, Thole JM, Goedhart J, Nielsen E, Munnik T, and Gadella TWJ (2009) Imaging phosphatidylinositol 4-phosphate dynamics in living plant cells. Plant J 57:356–372.

Vossen JH, Abd-El-Haliem A, Fradin EF, Van Den Berg GCM, Ekengren SK, Meijer HJG, Seifi A, Bai Y, Ten Have A, Munnik T, Thomma BPHJ, and Joosten MHAJ (2010) 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 62:224–239.

Wang CR, Yang AF, Yue GD, Gao Q, Yin HY, and Zhang JR (2008) Enhanced expression of phospholipase C1 (ZmPLC1) improves drought tolerance in transgenic maize. Planta 227:1127–1140.

Wang X, Devaiah SP, Zhang W, and Welti R (2006) Signaling functions of phosphatidic acid. Prog Lipid Res 45:250–278. Wang X, Zhang W, Li W, and Mishra G (2007) Phospholipid Signaling In Plant Response To Drought And Salt Stress, in

Advances in Molecular Breeding Toward Drought and Salt Tolerant Crops (Jenks MA et al. eds) pp 183–192, Springer Netherlands, Dordrecht.

Wasteneys GO, and Yang Z (2004) New views on the plant cytoskeleton. Plant Physiol 136:3884–3891. Wheeler GL, and Brownlee C (2008) Ca2+ signalling in plants and green algae - changing channels. Trends Plant Sci 13:506–

514. Wild R, Gerasimaite R, Jung J-Y, Truffault V, Pavlovic I, Schmidt A, Saiardi A, Jessen HJ, Poirier Y, Hothorn M, and Mayer

A (2016) Control of eukaryotic phosphate homeostasis by inositol polyphosphate sensor domains. Science 352:986 LP-990.

Williams SP, Gillaspy GE, and Perera IY (2015) Biosynthesis and possible functions of inositol pyrophosphates in plants. Front Plant Sci 6:1–12.

Wissing J, Heim S, and Wagner KG (1992) Diacylglycerol kinase from suspension cultured plant cells�: purification and properties. Plant Physiol 98:1148–1153.

Wissing JB, and Behrbohm H (1993) Diacylglycerol pyrophosphate, a novel phospholipid compound. FEBS Lett 315:95–99. Wissing JB, Kornak B, Funke A, Riedel B, Forschung B, and Braunschweig D (1994) Phosphatidate Kinase , A Nove1

Enzyme in Phospholipid Metabolism. Plant Physiol 105:903–909. Wissing JB, Kornak B, Funke A, Riedel B, Forschung B, and Braunschweig D- (1993) Phosphatidate Kinase , A Nove1

Enzyme in Phospholipid Metabolism. Plant Physiol 102:1243–1249. Wissing JB, Riedel.B, and Behrhohm H (1995) Diacylgtycerol- and phosphatidic acid-kinase studies in plant cell suspension

cultures. Biochem Soc Trans 32:867–871. Zarza X, Atanasov KE, Marco F, Arbona V, Carrasco P, Kopka J, Fotopoulos V, Munnik T, Gomez-Cadenas A, Tiburcio AF,

Page 89: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

87

and Alczar R (2016) Polyamine oxidase 5 loss-of-function mutations in Arabidopsis thaliana trigger metabolic and transcriptional reprogramming and promote salt stress tolerance. Plant, Cell Environ 1–16.

Zhai S, Sui Z, Yang A, and Zhang J (2005) Characterization of a novel phosphoinositide-specific phospholipase C from Zea mays and its expression in Escherichia coli. Biotechnol Lett 27:799–804.

Zhao Y, Yan A, Feijó J a, Furutani M, Takenawa T, Hwang I, Fu Y, and Yang Z (2010) Phosphoinositides regulate clathrin-dependent endocytosis at the tip of pollen tubes in Arabidopsis and tobacco. Plant Cell 22:4031–4044.

Zheng S, Liu Y, Li B, Shang Z, Zhou R, and Sun D (2012) Phosphoinositide-specific phospholipase C9 is involved in the thermotolerance of Arabidopsis. Plant J 69:689–700.

Zhu J (2016) Abiotic Stress Signaling and Responses in Plants. Cell 167:313–324. Zhu J (2002) Salt and Drought stress signal transduction in plants. Annu Rev Plant Biol 53:247–273. Zonia L, Cordeiro S, Tupy J, and Feijo JA (2002) 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 14:2233–2249. Zonia L, and Munnik T (2004) Osmotically induced cell swelling versus cell shrinking elicits specific changes in phospholipid

signals in tobacco pollen tubes. Plant Physiol 134:813–823. Zonia L, and Munnik T (2008) Vesicle trafficking dynamics and visualization of zones of exocytosis and endocytosis in

tobacco pollen tubes. J Exp Bot 59:861–873.

Page 90: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

88

SUPPLEMENTAL DATA

Supplemental Figure S1. Developmental stages of lateral root formation in plc5-1 and wild-type seedlings. Developmental stages of lateral root primordia from wild type and plc5-1 seedlings. Values are means ± SE of three independent experiments (n>20)

Supplemental Figure S2. Root development in seedlings of plc3plc5-double mutant. (a, b) Q-PCR analysis of PLC3- and PLC5-expression levels in wild type and plc3plc5. Relative expression is based on the expression of SAND. Values are means ± SD (n = 3) for one representative experiment. (c) Seedling morphology of wild-type and plc3plc5. Seeds were germinated on ½MS medium supplemented with 0.5% sucrose for 4 days, then transferred to ½MS plates without sucrose. Photographs were taken 12 days after germination (DAG). (d, e) Primary root (PR) length and lateral root (LR) number at 12 DAG. Values are means ± SE of three independent experiments (n>20). Asterisk (*) indicates significance at P<0.05 compared to wild type, based on Student’s t test.

Page 91: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC5inArabidopsis

89

Supplemental Figure S3. Phenotypic analysis of wild type- and PLC5 OE lines grown on soil or agar plates. Eight-weeks old wild type- and PLC5 OE plants were grown on soil under long day condition and the whole plants (a) and rosettes without inflorescences (b) were photographed. The inflorescence length was measured (c), and the fresh weight of rosette was determined (d). (e) Seedling morphology of wild type and PLC5-OE lines grown on agar plates. Seeds were germinated on ½ MS with 0.5% sucrose for 4 days, then transferred to ½ MS plates without sucrose. Photographs were taken 10 days after germination (DAG). (f) Primary root (PR) length, lateral root (LR) number and average lateral root (ALR) length at 10 DAG. All experiments were repeated at least three times. Values are means ± SD for one representative experiment (n=36). Asterisk (*) marks that PLC5 OEs value is significantly different from wild-type based on Student’s t-test (P< 0.05)

Page 92: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter3

90

Supplemental Figure S4. PPI- and PA levels in wild type and PLC5-OE lines. Five-days old seedlings were 32Pi-labelled overnight, and the next day their lipids extracted, separated by TLC and quantified by phosphoimaging. (a) Autoradiograph of a typical TLC, each lane representing 1/5th of the extract of 3 seedlings. (b) Quantification of 32P-labeled PIP2-, PIP- and PA levels in wild type or PLC5 OE lines. Values are calculated as the percentage of total 32P-labeled phospholipids, and are represented as means ± SD (n=3). This experiment was repeated twice with similar results.

Supplemental Figure S5. Soluble carbohydrates content in phloem sap of wild type and plc5-1. Phloem sap was isolated from 6 weeks-old Arabidopsis plants and their carbohydrates analyzed and quantified by GC-MS. Values are the means of triplicates ± SD from 3 independent experiments.

Supplemental Figure S6. Sugar content in seedlings of wild-type and PLC5 OE lines. Soluble carbohydrates were extracted from 10-day old Arabidopsis seedlings, including shoot (a) and root (b), analyzed by anion-exchange chromatography, and quantified by pulsed-amperometric detection. Mannitol was used as an internal standard. Values are the means ± SD (n=3) for one representative experiment. The experiment was repeated twice with similar outcome

Page 93: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

91

Chapter 4 Role for Arabidopsis PLC7 in stomatal closure, mucilage adherence, and leaf serration

Qianqian Zhang,1,2 Ringo van Wijk,1,2 Mart Lamers,1,* Francisca Reyes Maarquez,3 Aisha Guardia, 4 Denise Scuffi, 4 Carlos García-Mata,4 Wilco Ligterink,3 Michel A. Haring,1 Ana M. Laxalt,4 & Teun Munnik1,2

1Section Plant Physiology or 2Plant Cell Biology, Swammerdam Institute for Life Sciences, University of Amsterdam, Science Park 904, Amsterdam, 1098 XH, The Netherlands, 3Lab. of Plant Physiology, Wageningen University, Droevendaalsesteeg 1, 6708 PB, Wageningen, The Netherlands, 4Instituto de Investigaciones Biológicas (IIB-CONICET-UNMdP), Universidad Nacional de Mar del Plata, Mar del Plata, Argentina.

Page 94: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

92

ABSTRACT

Phospholipase C (PLC) signaling in plants is likely different from the well-established animal

paradigm as the prime targets for IP3 and DAG are absent, and plants contains very small

amounts of the putative PLC substrate, phosphatidylinositol 4,5-bisphosphate (PIP2). Nonetheless,

the genome of Arabidopsis thaliana encodes 9 PLC genes. To increase our understanding of PLC

signaling in plants, we have started to analyze various knock-out (KO), knock-down (KD) and

overexpression lines of Arabidopsis. Here, we functionally characterized Arabidopsis PLC7.

Promoter-GUS analyses revealed that PLC7 is specifically expressed in the phloem of roots, leaves

and flowers and also in trichomes and hydathodes. Two T-DNA insertion mutants were obtained,

with plc7-3 being a KO- and plc7-4 a KD line. In contrast to earlier-characterized phloem-

expressed PLC mutants, plc3 and plc5, plc7 mutants revealed no defects in primary- or lateral

root development. Nonetheless, like plc3 mutants, plc7 mutants were found to exhibit a reduced

sensitivity to ABA during stomatal closure. Double-knockout plc3 plc7 lines were found to be

lethal, whereas plc5 plc7 (plc5/7) mutants were viable, and revealed several new phenotypes, not

observed earlier in the single mutants. These include a defect in seed mucilage, enhanced leaf

serration, and an increased tolerance to drought. PLC7 overexpression lines showed an enhanced

drought tolerance, similar to what was found for PLC3 and PLC5 overexpression lines. In vivo 32Pi-labeling of seedlings treated with sorbitol, to mimic drought stress, revealed increased PIP2

responses in both drought tolerant plc5/7 and PLC7-OE mutants. Together, these results reveal

several novel functions for PLCs in plant stress and development. Potential mechanisms for this

are discussed.

Key words: PLC; seed mucilage; leaf serration; ABA sensitivity; drought tolerance.

Page 95: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

93

INTRODUCTION

Phospholipase C (PLC) plays a key role in mammalian signal transduction. Its intracellular activation

through transmembrane receptor occupation leads to the hydrolysis of phosphatidylinositol 4,5-

bisphosphate (PIP2) to produce two second messengers, inositol 1,4,5- trisphosphate (IP3) and

diacylglycerol (DAG). IP3 triggers the release of Ca2+ from an intracellular store via ligand-gated Ca2+

channel, while DAG recruits and activates protein kinase C (PKC) and stimulates TRP- (transient

receptor potential-) channels. The subsequent increase in Ca2+ and change in phosphorylation status of

numerous targets regulates multiple cellular processes that orchestrate the cell's response to the initial

stimulus (Irvine, 2006; Michell, 2008; Balla, 2013).

The PLC signaling pathway in plants is still enigmatic (Munnik, 2014). While higher plant

genomes encode all genes required for the metabolism of this pathway, they all lack the prime targets

for IP3 and DAG, i.e. the IP3 receptor, PKC and TRP channels (Wheeler and Brownlee, 2008; Munnik

and Testerink, 2009; Munnik and Vermeer, 2010; Munnik and Nielsen, 2011; Munnik, 2014; Heilmann,

2016). Instead, plants seem to phosphorylate IP3 and DAG into inositolpolyphosphates (IPPs) and

phosphatidic acid (PA), respectively, which are believed to function as second messengers (Gillaspy,

2013; Munnik, 2014; Heilmann, 2016). IP3 was initially thought to release Ca2+ from intracellular stores

when microinjected (Gilroy et al., 1990; Blatt et al., 1990; Allen and Sanders, 1994; Hunt and Gray,

2001), but later it was shown that this IP3 is phosphorylated to IP6 within seconds, and that the latter

compound was 10-100 fold more effective in releasing Ca2+ (Lemtiri-Chlieh et al., 2000, 2003).

Meanwhile, various other signaling functions for IP6 and other IPPs have been emerging, including the

pyro-phosphorylated IP7 and IP8. In yeast and mammalian cells, these IPP molecules play important

roles in various nuclear processes, including gene transcription, chromatin remodeling, mRNA export

and DNA repair, and are involved in a wide range of cellular processes, including osmoregulation,

phosphate homeostasis, vesicular trafficking, apoptosis, insulin- and immune signaling, cell cycle

regulation, and ribosome synthesis(Monserrate and York, 2010; Thota and Bhandari, 2015; Williams et

al., 2015). In plants, besides releasing Ca2+ in guard cells, IP6 has also been shown to bind the auxin

receptor, TIR1 (Tan et al., 2007), which is proposed to functionally regulate the SCFTIR1 ubiquitin-

ligase complex to control downstream auxin mediated-gene expression (Chapter 2; Chapter 3; Zhang et

al., 2017). Similarly, COI1, the receptor for jasmonate signaling was found to bind IP5 (Sheard et al.,

2010), with functional significance for plant immunity(Mosblech et al., 2008, 2011; Murphy et al.,

2008), even though it could be that the pyrophosphorylated form of IP5, i.e. PP-IP5 (= IP7), is

biologically responsible for this (Laha et al., 2015, 2016). GLE1, an mRNA export factor, was recently

identified as an important IP6 target in Arabidopsis involved in Pi homeostasis (Lee et al., 2015). SPX

domain-containing proteins have recently been identified to bind IPPs, including IP6 and many of these

proteins are involved in Pi signaling (Kuo et al., 2014; Puga et al., 2014; Wild et al., 2016). For PA,

several plant targets have been identified over the years, including protein kinases, proteins

Page 96: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

94

phosphatases, small G-proteins, RBOH (NADPH oxidase), GAPDH, ion channels and actin-binding

proteins (Wang et al., 2006; Mosblech et al., 2011), and PA has been implicated to regulate many

cellular processes, including vesicular trafficking, cytoskeleton dynamics, and ion-channel regulation

(Li et al., 2009, 2012, Pleskot et al., 2010, 2017; Testerink and Munnik, 2011; Thomas and Staiger,

2014). Note that PA is not only generated via PLC and DAG kinase (DGK), it can also be indirectly

formed via other DAG-generating enzymes, like non-specific PLCs (NPC; Munnik, 2014), or directly,

through phospholipase D (PLD) hydrolysis of structural phospholipids (Wang et al., 2006; Li et al.,

2009; Liu et al., 2013).

How, when, and whether PLC signaling is involved in generating PA and IPPs in the above

events is still largely unknown and mostly based on correlations and indirect evidence. Hence, tools to

genetically manipulate PLC levels will be very helpful to functional characterize its role(s). In this way,

silencing of PLC revealed its importance in plant defense in tomato and Arabidopsis (Vossen et al.,

2010; D’Ambrosio et al., 2017), in cytokines- and gravity signaling in Physcomitrella (Repp et al.,

2004), in ABA signaling and stomatal control in tobacco and Arabidopsis (Sanchez and Chua, 2001;

Hunt et al., 2003) but also in development, where Arabidopsis PLC2 plays a key role in gametogenesis,

reproduction and root development (Li et al., 2015; Laxalt et al., 2016; R. van Wijk and T. Munnik, In

prep.). In petunia and tobacco, PLC regulates pollen tube-tip growth (Dowd et al., 2006; Helling et al.,

2006). Alternatively, several Arabidopsis T-DNA insertion mutants have meanwhile revealed several

roles for PLC3, PLC5 and PLC9 in seed germination, (lateral) root development, ABA signaling and

heat stress tolerance ( Chapter 2 and 3; Zheng et al., 2012; Gao et al., 2014). On the other hand,

overexpression of PLC has been shown to increase drought tolerance in maize, canola, tobacco and

Arabidopsis (Wang et al., 2008; Georges et al., 2009; Tripathy et al., 2011; Chapter 2 and 3). Even

though it is still unknown how PLC exactly achieves all the above, it is very important that these

molecular tools become available for physiological and biochemical studies.

The Arabidopsis genome encodes 9 PLC genes, which are subdivided into four clades (Hunt et

al. 2004; Tasma et al., 2008; Pokotylo et al., 2014). Earlier, we found that knock-out (KO) mutants of

PLC3 or a knock-down (KD) mutant of PLC5, exhibited small defects in root development, i.e. 5-10%

reduced growth of the primary root and a 10-20% reduction in the number of lateral roots. Interestingly,

even though these PLCs come from different subfamilies, plc3 plc5-double mutants did not intensify

the phenotype, indicating other PLCs could be involved. Since PLC3 and PLC5 were specifically

expressed in phloem-companion cells and revealed a segmented expression pattern in the root form

which lateral roots emerged, we search for other Arabidopsis PLCs specially expressed in the phloem,

and found PLC7. This gene belongs to a different subfamily than PLC3 and PLC5. Promotor-GUS

analyses confirmed its expression in all vascular tissues, but also in trichomes and hydatodes, however

plc7-3 KO- or plc7-4 KD mutants revealed no defects in root or shoot development. Creating double

mutants revealed that the combination plc3/7 was lethal and that plc5/7 did not exhibit defects in root

development either. Nonetheless, several new phenotypes appeared for the latter mutant, including non-

Page 97: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

95

adherent mucilage, enhanced leaf serration and increased drought resistance. In addition, like PLC3 and

PLC5, overexpression of PLC7 also leads to an increase in drought tolerance.

RESULTS

Expression of PLC7 during plant development

Histochemical assays on pPLC7-GUS-SYFP reporter lines indicated that PLC7 was mainly expressed in

the vasculature throughout all developmental stages, including root, cotyledons, leaves, hypocotyl,

flower (stamen, style, petal, sepal, receptacle and pedicel) and silique septum (Fig. 1a-j, n and o), which

is similar to the expression pattern of PLC3 (Zhang et al, 2017; Chapter 2) and PLC5 (Chapter 3).

However, there are also some differences. During seed germination PLC7 expression was mainly

observed in the hypocotyl (28h after transfer to 22ºC, Fig. 1a), which remained like this in 2-day old

seedlings (Fig. 1b) and then spread to the vasculature throughout the plant upon further development

(Fig. 1c-j). Interestingly, GUS expression was quite abundant in hydathodes, in young seedling, but also

mature plants (Fig. 1b-e, l, arrows). Unlike PLC3 and PLC5, PLC7 did not display the characteristic,

segmented expression in the root vasculature. Instead, expression was homogenous in both main root-

and lateral root vasculature (Fig. 1f, g). The expression stopped near the transition zone of the root (Fig.

1h). Strong GUS staining was visible in trichomes (Fig. 1j, k), which is similar to that of PLC5, but

stronger and different from PLC3, that is only expressed in trichome basal cells. No GUS activity was

detected in guard cells, as was found for pPLC3::GUS-YFP and pPLC5::GUS. These results confirm

that PLC7 is expressed throughout the plant (Tasma et al., 2008), but the expression is mostly restricted

to the vasculature.

Page 98: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

96

Figure 1. pPLC7::GUS-SYFP expression in seedlings and mature tissues of Arabidopsis. GUS staining was found in: Embryo, 28 hrs after stratification (a), vascular tissue of 2-d old seedlings (b) and in 10-d old seedlings, including, cotelydons and roots (c-h), mature plant (i), trichomes (i), hydathodes (indicated by arrows) (c-e, i), silique (o) and flower (n), including style, filament, receptacle and pedicel. No staining was detected in guard cells (m)

No change in root system architecture in plc7 and plc5/7 mutants Knock-out/down mutants of PLC3 or PLC5 were affected in their primary- and lateral root growth

formation. However, a double plc3/5 mutant did not reveal a stronger effect, indicating that there might

be a third PLC involved. (Zhang et al., 2017; Chapters 2, 3). To investigate the role of PLC7 we

identified two homozygous T-DNA insertion lines, plc7-3 (SALK_030333) and plc7-4 (SALK_148821)

(Fig. 2a). Reduction of PLC7 expression in plc7-3 and plc7-4 mutants was confirmed by Q-PCR (Fig.

2b) and revealed that the former is a KO- and the latter a KD mutant. However, the root architecture of

12 days old seedlings of these plc7 mutants did not differ from the wild-type (Fig. 2c, d).

Page 99: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

97

Figure 2. In contrast to plc3- and plc5 mutants, KO- or KD mutants of PLC7 are not affected in root development. (a) T-DNA insertion positions (triangles) in the PLC7 gene of plc7-3 and plc7-4 lines. Filled boxes and lines represent exons and introns, respectively, while open boxes represent untranslated regions. (b) PLC7 expression levels in wild-type, plc7-3 and plc7-4 measured by Q-PCR using SAND as a reference gene. Values are the means ± SD (n = 3) of a representative experiment that was independently repeated at least three times. (c) Seedling morphology of wild-type, plc7 mutants. Seeds were first germinated on ½MS with 0.5% sucrose for 4 days and then transferred to ½MS plates without sucrose. Photographs were taken 12 days after germination (DAG). (d) Primary root (PR) length and (e) lateral root (LR) number at 12 DAG. Values are means ± SE of three independent experiments (n>20). Asterisk (*) indicate significance at P<0.05 compared to wild-type, based on Student’s t test.

To analyze gene redundancy, we tried to generate plc3/5/7-triple mutants by crossing plc7-3

with the plc3/5-double mutant. However, the plc3/7 combination turned out to be lethal and only

homozygous plc5/7-double mutants could be obtained (Fig. 3a). As shown in Figures 3b-d, no

significant difference was found between the root system of plc5/7- and wild-type seedlings.

Figure 3. A plc5/plc7 double mutant is viable and is not affected in root architecture. (a) PLC5 (left) and PLC7 (right) expression in wild-type and plc5/7-double mutant. Relative expression is based on the expression of SAND measured by Q-PCR. Values are means ± SD (n = 3) of a representative experiment that was repeated at least three times. (b) Seedling morphology of wild-type and plc5/7. Seeds were germinated on ½MS with 0.5% sucrose for 4 days and then transferred to ½MS plates without sucrose. Photographs were taken 12 days after germination (DAG). (c) Primary root (PR) length and (d) lateral root (LR) number at 12 DAG. Values are means ± SE of three independent experiments (n>20). Asterisk (*) indicate significance at P<0.05 compared to wild-type, based on Student’s t test.

plc5/7 mutant displays loose seed coat mucilage and reduced cellulose rays

While imbibing seeds for stratification, we noticed that the volume of the seed pellet of plc5/7 seeds

was always smaller than wild type after overnight incubation (Fig. 4a). Upon imbibition, the seed coat-

epidermal cells normally extrude a mucilage that forms two transparent layers (adherent and non-

Page 100: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

98

adherent layers) around the seed. To examine whether the smaller volume of the plc5/7 mutant was

caused by a mucilage defect, we stained imbibed seeds with Ruthenium red (Fig. 4b), which stains

pectins, the main component of mucilage (Western et al., 2000; Macquet et al., 2007). Compared to

wild-type, the adherent and non-adherent layers were more expanded in the plc5/7 mutant (Fig. 4b, top

panel) and when seeds were mildly shaken to remove the non-adherent layer, or treated with EDTA,

plc5/7 seeds lost the adherent layer completely(Fig. 4b, middle and lower panel, respectively).

Increased solubility of the pectins has been linked to perturbation of cellulose deposition

(Harpaz-Saad et al., 2011; Mendu et al., 2011; Sullivan et al., 2011; Yu et al., 2014; Ben-Tov et al.,

2015; Basu et al., 2016; Hu, Li, Wang, et al., 2016; Hu, Li, Yang, et al., 2016). To test this, wild-type

and plc5/7 seeds were stained with Calcofluor White (CFW, for cellulose and other β-glucans staining;

(Fig. 4c, left panel) and Pontamine S4B (cellulose-specific dye; (Fig. 4c, right panel) (Willats et al.,

2001; Anderson et al., 2010; Wallace and Anderson, 2012). In wild-type seeds, the primary cell wall

remnants, and rays extending from the columella were stained by both dyes (Fig. 4c). The plc5/7 seeds

showed a similar staining pattern, but the CFW intensity was lower and the rays were clearly reduced

compared to wild-type. These results point to a role for PLC5 and PLC7 in cellulose ray formation,

which is a completely novel function for PLC.

Figure 4. Seeds of plc5/7 double mutant exhibit mucilage phenotype. (a) Seeds of plc5/7 swell less during imbibition. Equal amounts of dry seeds of wild type and mutant were immersed in water overnight and photographed the next day. (b) Ruthenium red staining of wild type- and plc5/7 seeds, without shaking (top), with shaking (middle) and EDTA treatment. With shaking shows adherent mucilage layer; without shaking displays both adherent and non-adherent mucilage layers. (c) Cellulose staining by Calcofluor white (left panel) or Pontamine B (right panel) in wild type- and plc5/7 seeds. Confocal images of whole seeds, cross section, and close-up views (top, middle and bottom, respectively) are shown. Bars represent 2 mm (a), 0.1mm (b), 0.1mm (top and middle rows) or 0.025 mm (bottom row) (c).

PLC5 and PLC7 expression in developing seed

To further investigate the expression pattern of PLC5 and PLC7 during seed development,

pPLC5::GUS and pPLC7::GUS-SYFP lines were used for additional histochemical analyzes (Fig. 5).

At 4 day after pollination (DAP), some GUS activity was found in pPLC5::GUS (Fig. 5a, left panel)

Page 101: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

99

while a strong staining in the seed coat and chalazal area of pPLC7::GUS-SYFP was observed (Fig. 5b).

Later in development (at 8 and 10 DAP), GUS activity became stronger, with PLC5 expression

appearing in the seed coat and funiculus (Fig. 5a) and PLC7 staining became stronger in seed coat and

chalazal (Fig. 5b).

Figure 5. PLC5 and PLC7 expression during seed development. (a) GUS activity analysis in pPLC5::GUS developing seeds. Expression was found in the funiculus at 8 days after pollination (DAP) and seed coat at 10 DAP. (b) GUS activity analysis in pPLC7::GUS developing seeds. Staining was found in the chalazal and in the seed coat. Bar = 0.1 mm

Analysis of PPI- and PA levels in developing and germinating seeds

To analyze the levels of the potential PLC substrates (i.e. PIP and PIP2) and product (PA conversion of

PLC-generated DAG), 32Pi-labelling (24h) of wild-type- and plc5/7 mutant seeds at 10 DAP were

compared with germinating, mature seeds. As shown in Figure 6, wild-type and plc5/7 seeds were

found to contain similar amounts of PIP2, PIP and PA in both stages. Interestingly, PIP- and PA levels

were much higher in developing seeds while PIP2 levels were similar in both stages.

Figure 6. PPI- and PA levels in developing- and germinating

(mature) seeds of wild type and plc5/7. (a) Developing seeds, ~200 seeds at 10 DAP of wild type and plc5/7 were labelled with 32PO4

3- for 24 h and their lipids extracted, separated by TLC and quantified by Phosphoimaging. (b) Mature seeds (~200) of wild type and plc5/7 were pre-germinated on ½MS with 0.5% sucrose plates until testa ruptured, then labelled with 32PO4

3- for 24 h. Lipids were then extracted, separated and quantified as above. 32P-levels of PIP2, PIP and PA are expressed as percentage of total 32P-phospholipids. Three independent experiments were performed; data shown are means ± SD (n=3) from one representative experiment.

Page 102: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

100

Leaves of plc5/7 plants display enhanced level of leaf serration

Growing plc5/7 mutants on soil for longer periods of time revealed a novel phenotype in leaf-edge

patterning (serration). Overall, the level of serration in successive rosette leaves was significantly

increased in plc5/7 mutant, which appeared to be stronger in the proximal part of the blade than in the

distal part (Fig. 7a, b). To quantify this in more detail, we measured various parameters of the 8th leaf

(Fig. 7c-e) of 4 -weeks old rosettes of both genotypes. No changes in blade length were observed

between wild-type and plc5/7, but the blade width, -perimeter and -area were slightly, although not

significantly, bigger in plc5/7 (Fig 7d). The serration number was not changed in the plc5/7 mutant, but

the serration level (indicated by the ratio between the distance from midvein to tip and the distance

from midvein to sinus, see Fig. 7c) was significantly higher in three successive teeth (Fig. 7f).

Recent studies have indicated that Arabidopsis leaf-margin development is controlled by a

balance between microRNA164A (MIR164A) and CUP-SHAPED COTYLEDON2 (CUC2) (Nikovics et

al., 2006). Hence, we compared the expression of MIR164A and CUC2 in wild-type and plc5/7 leaves.

As shown in Figure 7g, plc5/7 leaves were consistently (3 independent experiments) found to contain

higher levels of CUC2 and lower levels of MIR164A, resulting in a significant increase in the

CUC2/MIR164A ratio.

Page 103: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

101

Figure 7. Leaves of plc5/7 plants display enhanced leaf serration (a) Phenotype of wild type- and plc5/7 rosette. Rosettes of 4-week old plants grown in short day conditions, were cut and photographed immediately. Bar = 1 cm. (b) Leaf series of 4-week old wild type and plc5/7 plant. (c) Cartoon to demonstrate the leaf parameters measured of the 8th leaf. (d) Quantification of blade size including, length, width, perimeter and area. (e,f) Quantification of leaf serration number (e) and level (f) in wild type and plc5/7 mutant. (g) Expression of CUC2 and MIR164A and their ratio in wild type and plc5/7 mutant. Expression is relative to the expression of OTC. Three independent experiments were performed. Data in (g) represents the means ± SD (n=3) from one representative experiment that was repeated twice with similar results. Asterisk (*) indicate significance at P<0.05 compared to wt, based on Student’s t test.

The plc5/7 mutant displays enhanced tolerance to drought

When plants were left in the greenhouse without watering, we noticed that plc5/7 mutants appeared to

be more drought tolerant while the single mutants behave like wt (data not shown). Performing multiple

drought tolerance assays indeed confirmed this (Fig. 8a), and detached rosettes of 4-week-old plc5/7

plants were found to loose less water than wild-type (Fig. 8b).

ABA plays a key role during the response to dehydration stress and is known to induce stomatal

closure to reduce the loss of water (Sean et al., 2010). Hence, we checked the stomatal-closure of plc5/7

and wild-type in response to ABA. As shown in Figure 8c, plc5/7 has less-opened stomata compared to

Page 104: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

102

wild-type without ABA, but when ABA was applied, stomata closed in genotypes, although the plc5/7

mutant appeared to be less responsive.

Previously, we found that overexpression of PLC3 or PLC5 enhanced drought tolerance in

Arabidopsis (Zhang et al., 2017; Chapters 2, 3), which was earlier also revealed for maize, canola and

tobacco (Wang et al., 2008; Georges et al., 2009; Tripathy et al., 2011). We therefore wondered

whether the increase in drought tolerance in plc5/7 was a result of the overexpression of any of the

other (redundant) 7 PLCs. However, no strong overexpression of any PLC was found in plc5/7 (Fig. 8d).

In fact, PLC1, PLC2 and PLC4 were found to be slightly down-regulated (Fig. 8d).

Figure 8. Soil-grown plc5/7 mutant are more tolerant to drought stress. (a) Six-weeks old plants of wild type- or plc5/7 mutant plants, grown on soil and exposed to drought by withholding water for the last two weeks and then were photographed. (b) Water loss of detached rosette of normally watered, 4-weeks old plants. Water loss was measured at indicated time points and expressed as a percentage of the initial fresh weight. Values are means ± SD for one representative experiment (n=36). (c) Effect of ABA on the stomatal aperture in leaf strips of wild type and plc5/7 plants. Leaves from 3-weeks old plants were stripped and the peels were incubated in opening buffer with light for 3 h until stomata were fully open. Peels were then treated with different concentrations of ABA for 90 mins, after which stomata were digitized and aperture widths measured. Values are means ± SE of at least three independent experiments (n >100). Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method). (d) PLC1- PLC9 expression level in wild type and plc5/7 mutant plants as measured by QPCR, relative to the amount of SAND. Values are means ± SD (n = 3) for one representative experiment that was repeated two times with similar results.

Page 105: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

103

Overexpression of PLC7 increases drought tolerance

As mentioned above, overexpression of Arabidopsis PLC3 and PLC5, which are from different

subfamilies than PLC7, resulted in enhanced drought tolerance (Zhang et al., 2017; Chapters 2 and 3).

To check the effect of PLC7 overexpression, homozygous T3 plants were generated. Two lines, PLC7-

OE9 and PLC7-OE12, which overexpressed PLC7 around 80- to 100-fold, respectively, were selected

for further studies, (Fig. 9a). Both lines were indeed found to be more drought tolerant than wild-type

(Fig. 9b), and lost slightly less water when rosettes of 4-week-old plants were detached (Fig. 9c). The

stomatal aperture and their response to ABA was found to be similar to wild-type (Fig. 9d), which is

different to PLC3- and PLC5-OE plants that had stomata that opened less wide than the wt.

Figure 9. Overexpression of PLC7 enhances drought tolerance. (a) PLC7 expression levels in wild type and two homozygous PLC7 overexpression lines, PLC7-OE9 and PLC7-OE12 as measured by Q-PCR and based on the expression of the SAND reference gene. Values are means ± SD (n = 3) for one representative experiment. (b) Phenotype of six week old plants from wild type and PLC7-overexpresion lines, #OE9 and #OE12, after two weeks of drought stress. (c) Water loss of detached rosette of four weeks old plants. Water loss was measured at indicated time points and expressed as a percentage of the initial fresh weight. Values are means ± SD for one representative experiment (n=36). (d) Stomatal aperture in leaf peels of wild type, PLC7-OE9 (left), PLC7-OE12 (right) and the effect of ABA. Leaves from 3-weeks old plants were stripped and peels were incubated in opening buffer with light for 3 h until stomata were fully open. Peels were then treated with different concentrations of ABA for 90 mins, after which stomata were digitized and the aperture width measured. Values are means ± SE of at least three independents (n > 100). Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method).

Page 106: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

104

Phospholipid responses upon osmotic stress

To analyze the phospholipid responses in plc5/7 and the PLC-OE lines, 32P-labeling experiments (3 hrs

labeling) were performed on seedlings, and the effect of sorbitol treatment was tested to mimic osmotic

stress. Both plc5/7 and wild-type showed similar levels of PPI- and PA in absence of sorbitol (Figs. 10a,

b). Upon sorbitol treatment, a stronger PIP2 response was typically observed in the plc5/7 lines though

numbers appear not to be statistically different. While PIP2 levels increased by ~4 times in wild-type,

plc5/7 seedlings typically exhibited a 6-times increase. No difference in PA- or PIP response was

observed (Fig. 10b).

PLC7-OE lines revealed no difference in PPI or PA levels compared to wild-type in control

conditions (Fig. 10c, d). However, with sorbitol, again a stronger PIP2 response was observed in PLC7-

OE lines, ~6-times vs ~4-times for wild-type, while PA- and PIP responses were similar (Fig. 10c, d).

These results suggest that both plc5/7 and PLC7-OE plants boost more PIP2 in response to osmotic

stress than wild-type, similar to what we found earlier for PLC3- and PLC5-OE lines (Chapters 2, 3).

Figure 10. Both plc5/7 and PLC7 over-expressers displayed stronger PIP2 response upon osmotic stress. Six-day-old seedlings were 32P-labeled for 3h and then treated with buffer ± 600 mM sorbitol for another 30 min. Extracted lipids were analyzed by TLC and quantified through phosphoimaging. (a and c). Typical TLC profile with each lane representing the extract of 3 seedlings. (b, d) 32P-levels of PIP2, PIP and PA of wild-type and plc5/7 (b) or PLC7-OEs (line #9 and #16) (d) with and without sorbitol. Three independent experiments were performed. Data shown are the means ± SE (n=3) from three independent experiments. Data was analyzed by 2-way ANOVA. Statistical significant differences between normal and sorbitol conditions in wild type and plc5/7 or PLC7-OEs are indicated by letters (P<0.05).

Page 107: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

105

DISCUSSION

Knockout of PLC7 does not affect root architecture

Previously, we demonstrated that Arabidopsis PLC3 and PLC5 were both involved in lateral root

formation and that a plc3/5-double mutant did not exhibit a more severe phenotype than single mutants

(Chapters 2, 3). We therefore speculated that another PLC must be involved. Since promotor-GUS

analyses revealed specific expression in phloem companion cells and a typical segmented pattern in the

vascular of the primary root from which lateral roots emerged, we specifically searched for other

phloem-specific PLCs. Using the eFP browser, we found two potential candidates, PLC2 and PLC7.

Since T-DNA insertion mutants of PLC2 were lethal (D’Ambrosio et al., 2017; Di Fino et al., 2017; R.

van Wijk and Munnik, unpublished), we focused on PLC7. Two independent homozygous T-DNA

insertion mutants were obtained, with plc7-3 being a KO- and plc7-4 a KD line. Both mutants exhibited

normal root architecture. In an attempt to create double- and triple mutants of the redundant PLCs, we

crossed the plc3-2/plc5-1 (plc3/5; Chapter 3) with plc7-3, but this only resulted in viable plc5/7 double

mutants, as the combination plc3/plc7 turned out to be lethal (not shown). The plc5/7 double mutant,

however, did not reveal any significant changes in root morphology (Figs. 2, 3). While pPLC7-GUS

analyses confirmed the vascular expression of PLC7, the typical segmented pattern found for PLC3 and

PLC5 was not found. These results may indicate that PLC2 is the redundant PLC in root development,

and preliminary results from this laboratory, using induced silencing (R. van Wijk and T. Munnik,

Unpublished), confirm this. Interestingly, PLC7 was also expressed in hydathodes and in seeds, which

correlates well with the two new phenotypes that were found for plc5/7-double mutants, which have

never been observed before. These include a mucilage phenotype in seeds and a serration phenotype in

leaves. The latter may correlate with PLC7's expression at hydathodes. PLC7 was also strongly

expressed in trichomes. PLC5 is also expressed there, although less, and PLC3 is typically expressed

only in the trichome basal cells of developing leaves (Chapters 2, 3). We checked for trichome

phenotypes in individual plc3, plc5 or plc7 and plc3/5- and plc5/7- mutants, but found no obvious effect

(number, shape) was visible. Again, this could be due to redundancy, or PLC is involved in content of,

or transport to the trichomes, which would not lead to an obvious phenotype.

Strikingly, and new to PLC loss-of-function lines, is that plc5/7 mutants were more drought

tolerant, a phenotype that is typically found when PLC is overexpressed (Chapter 2 and 3; Wang et al.,

2008; Georges et al., 2009; Tripathy et al., 2011). However, we found no upregulation of redundant

PLCs in the plc5/7-mutant background. If anything, some down regulation of PLC1, PLC2 and PLC4.

RNASeq analyses of the plc mutant and OE lines may shed light on potential pathways that are up- or

down-regulated to explain various phenotypes. Most importantly, altering expression of PLC genes

results in clear phenotypes and this will further help elucidating the role of PLC in plant signaling and

development.

Page 108: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

106

While several new pieces of the PLC-signaling puzzle (new phenotypes) in plants have been

found, it remains unclear how this is achieved at the cellular and molecular level. It is clear that plants

lack the prime targets for IP3 and DAG, and there are several indications that responses are coupled via

inositolpolyphosphates (IPPs) and/or PA (Wang et al., 2006; Arisz et al., 2009, 2013; Munnik and

Vermeer, 2010; Munnik and Nielsen, 2011; Testerink and Munnik, 2011; Gillaspy, 2013; Munnik, 2014;

Laha et al., 2015; Heilmann, 2016; Hou et al., 2016). Alternatively, since PIP or PIP2 are emerging as

second messengers themselves, PLC could function as an attenuator of signaling too. In that respect, it

is interesting to notice that the increased drought-tolerant phenotype in plc5/7 and the PLC3, -5, and -7

overexpression lines correlates with a stronger PIP2 response in seedlings osmotically stressed with

sorbitol. How this would work remains unclear, but these lines could somehow be primed for an

enhanced PIP2 turnover.

Role for PLC5 and PLC7 in seed mucilage

Mucilage is a complex of polysaccharides, which is from seed coat epidermal cells when seeds are

exposed to an aqueous environment(Western et al., 2000). The major component of mucilage is pectin,

of which polygalacturonic acid (PGA) and rhamnogalacturonan I (RGI) are the most common

compounds found in cell walls and mucilage (Carpita and Gibeaut, 1993; Cosgrove, 1997). In addition

to pectin, mucilage also contains some minor monosaccharides, i.e. arabinose, galactose, glucose,

xylose and mannose, which are equally important in controlling mucilage properties(Voiniciuc, et al.,

2015). The extruded mucilage contains two layers, a non-adherent outer layer and an adherent inner

layer. The former is easily removed from the seed (Harpaz-Saad et al., 2011; Mendu et al., 2011;

Sullivan et al., 2011). The latter is hard to detach, even chemically (Macquet et al., 2007). Our result

showed that inner mucilage of plc5/7 seed coats was easily lost (Fig. 4b). Many factors are involved in

maintaining the adherence of the inner mucilage, with cellulose being the most important (Harpaz-Saad

et al., 2011; Sullivan et al., 2011; Yu et al., 2014; Ben-Tov et al., 2015; Voiniciuc, Schmidt, et al.,

2015; Hu, Li, Wang, et al., 2016; Hu, Li, Yang, et al., 2016). Staining with Calcofluor white and

Pontamin S4B indicated that certain cellulose rays were missing from plc5/7 seed-coat mucilage (Fig.

4c). We checked for sugar-composition aberrations in whole seeds, but found no changes compared to

wild-type (Supplemental Fig. S1). Possibly, the distribution between non-adherent and adherent

mucilage layers might be different, which needs to be further determined. The adherence of mucilage

by cellulose is mainly caused by its crystallization by self-association via hydrogen bonding (Sullivan et

al., 2011), and several genes have been recently reported to regulate cellulose crystallization, i.e.

COBRA-LIKE2, IRX14 and IRX7 etc (Ben-Tov et al., 2015; Hu, Li, Wang, et al., 2016; Hu, Li, Yang, et

al., 2016).

Histochemical analysis of the GUS reporter lines indicated that both PLC5 and PLC7 are

expressed during seed development (Fig. 5). Until now, no mucilage deficiency has been linked to PLC,

and we only observed the mucilage phenotype in the plc5/7-double mutant of all our PLC knockout

Page 109: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

107

mutants. The defect in both PLC5 and PLC7 probably breaks the balance for mucilage maintenance by

altering cellulose deposition and/or crystallization in the inner mucilage. How the enzyme PLC could

be involved in all this will require additional research. Potentially, PLC could be required for mucilage

secretion or for the localization or activity of cellulose synthases, for example. PA, PIP and PIP2 have

been implicated to play essential roles in vesicular trafficking, -fusion and -fission. Even though no

difference was found in PPI- or PA levels in either developing or mature seeds (Fig. 6), reduced

amounts of PLC5 and lack of PLC7 might cause crucial local changes in lipid or IPP molecules.

Role for PLC in leaf serration

Leaf shape is defined by the pattern and degree of indentation at the margin area, thus distinguishing

plant species (Tsukaya, 2006). The patterning involves a complex crosstalk between hormone signaling

and genetic regulators (Byrne, 2005; Fleming, 2005). Leaf serration is an important patterning event

and its development involves auxin maxima at the protrusion of each serrated section (Hay and Tsiantis,

2006). Based on recent genetic studies, the auxin efflux carrier PIN-FORMED 1 (PIN1) and CUP-

SHAPED COTYLEDON 2 (CUC2) have been identified as two key factors required for the formation

of leaf serration (Hay et al., 2006; Nikovics et al., 2006). PIN1 asymmetrically localizes on plasma

membranes and directionally transports auxin, creating auxin maxima that direct the outgrowth of the

serrations (Hay et al., 2006; Scarpella et al., 2006). CUC2 is a transcription factor that is post-

transcriptionally downregulated in leaves by MIR164A (Nikovics et al., 2006). CUC2 expression is

limited to the sinus where the serration starts, and the promotion of serration outgrowth is through cell

division, not by suppression of sinus growth (Kawamura et al., 2010). CUC2 has also been indicated to

regulate the polarized localization of PIN1 in convergence points at the leaf margin, playing a role in

establishing, maintaining and/or enhancing auxin maxima that result in leaf serration (Kawamura et al.,

2010; Bilsborough et al., 2011). In a feedback loop, auxin downregulates CUC2, both transcriptionally

and post-transcriptionally through activation of MIR164A (Bilsborough et al., 2011).

The plc5/7-double mutant revealed a mildly-enhanced leaf-serration phenotype (Fig. 7) Subsequent

measurement of CUC2 and MIR164A expression revealed a an up-regulation of CUC2 and down-

regulation MIR164A, which is consistent with the enhanced serration (Kawamura et al., 2010;

Bilsborough et al., 2011). Promotor-GUS analyses showed that both PLC5 and PLC7 were expressed at

leaf hydathodes, a secretory tissue that secretes water through the leaf margin that is also associated

with leaf serration (Tsukaya and Uchimiya, 1997) and auxin response maxima (Scarpella et al., 2006).

Hence, it is possible that PLC5 and PLC7 together contribute to the regulation of leaf serration, also

because there was no such phenotype found in single mutants. We also measured PPI-and PA levels in

plc5/7 rosette leaves but, like seeds and seedlings, found no changes (data not shown; Figs. 6 and 10). It

is possible that differences are limited to particular cells or tissues and that most of the cells have

normal levels/responses. How PLC5 and PLC7 are involved in leaf serration requires further

investigation. IP6 could play a positive role in auxin signaling through activation of TIR1 (Chapter 2).

Page 110: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

108

However, increased leaf serration would then be expected to result from increased IP6 formation, which

not what we would directly expect from a plc5/7 mutant. On the other hand, PIP2 has been suggested to

play a role in the polarized localization of PIN1(Ischebeck et al., 2013; Tejos et al., 2014), in which

PLC could also play a role. As mentioned earlier, down-regulation of these two PLCs (plc5 KO is likely

lethal; Chapter 3) could result in a local upregulation of another PLC, even though overall PLC

expression levels (Fig. 8d) do not confirm such a model.

Both PLC7 overexpression and plc5/7 down-regulation enhance drought tolerance

Overexpression of PLC in maize, tobacco and canola improved drought tolerance (Wang et al., 2008;

Georges et al., 2009; Tripathy et al., 2011). Similarly, we found that overexpression of PLC3 or PLC5

increased the drought tolerance of Arabidopsis. To investigate the effect of PLC7 overexpression,

homozygous pUBQ10::PLC7 lines were generated. Meanwhile, we took plc5-1, plc7-3 and plc5/7-

double mutants along to see how they would behave under drought stress. While plants appeared

normal under control conditions, except for plc5/7 showing more serrated leaves, upon drought stress

both plc5/7- and PLC7-OE lines performed better (Figs. 8a, 9b), and lost less water than wild type (Figs.

8b, 9c), even though the latter was not statistically significant for PLC7-OE lines (Fig. 9c).

ABA synthesis is stimulated by dehydration stress and known to induce stomatal closure to

reduce the water loss (Sean et al., 2010). Previous result indicated that PLC3- and PLC5-OE lines

showed a 'less-open stomata' phenotype in the absence of ABA, and that they exhibited a reduced

closing response to ABA compared to wild-type (Chapters 2 and 3). Overexpression of PLC7 did not

show 'less-open stomata' under control conditions, and their response to ABA was like wild type.

Stomata of plc5/7 plants were less open but their response to ABA appeared normal, maybe slightly

decreased, which is in contrast to plc5- and plc7-single mutants, which have a normal opening at

control conditions, and plc7 is less sensitive to ABA (Supplemental Fig. S2; Chapter 3). Interestingly,

PLC7 itself was not expressed in guard cells (Fig. 1), but the gene is rapidly upregulated in guard cells

upon ABA treatment (Bauer et al., 2013). Hence, it could be that the drought tolerant phenotype in the

plc5/7 is a consequence of a local upregulation (for instance in the guard cells) of one or more

redundant PLCs, which may remain undetectable when whole seedling levels are measured (Fig. 9).

Osmotic stress-induced phospholipid signaling responses have been reported in many studies

(Munnik and Vermeer, 2010; Meijer et al., 2017). We measured PPI- and PA levels in response to

sorbitol to mimic drought stress in plc5/7, PLC7-OE and wild-type seedlings using 32P-labeling. Under

control conditions, no difference in PPI- and PA levels were found among the genotypes (Fig. 10).

However, in response to osmotic stress, the PIP2 response in both plc5/7 and PLC7-OE lines were

stronger than wild-type (Fig. 10b and 10d). Apart from second messenger precursor, PIP2 could also

play a role as signaling molecule itself, for example in the reorganization of the cytoskeleton, in endo-

and exocytosis, or ion channel regulation (Stevenson et al., 2000; Martin, 2001; van Leeuwen et al.,

2007; Heilmann, 2016). Identification and characterization of some of PIP2's main targets will be

Page 111: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

109

essential to start unraveling the molecular mechanisms involved. Whether the above response in plc5/7

is a consequence of enhanced, local expression of another PLC, or a consequence of differential gene

expression and hence drought tolerant response, requires further studies.

MATERIALS AND METHODS

Plant material

Arabidopsis thaliana (Col-0) was used throughout. T-DNA insertion mutants plc7-3 (SALK_030333)

and plc7-4 (SALK_148821) were obtained from the SALK collection (signal.salk.edu). Homozygous

plants were identified by PCR in F2 generation using gene-specific primers. For the identification of

plc7-3, we used forward primer 5’- GATTTGGGTGATAAAGAAGTTTGG -3’; reverse primer 5’-

CTCCACACAATCTCAGCATTAC-3’) and left border primer LBb1.3 (5’-

ATTTTGCCGATTTCGGAAC-3’, in combination with the forward primer). For plc7-4 identification,

forward primer 5’- TCCTTCCTGTTATCCATGACG -3’; reverse primer 5’-

TTGAAGAAAGCATCAAGGTGG -3’) and left border primer LBb1.3 (5’-

ATTTTGCCGATTTCGGAAC-3’, in combination with the reverse primer) were used. To generate

double mutant plc5/7, plc5-1 and plc7-3 were used for crossing.

Root growth

Seeds were surface sterilized in a desiccator using 20 ml thin bleach and 1ml 37% HCl for 3 hours, and

then sowed on square petri dishes containing 30 ml of ½ strength of Murashige and Skoog (½MS)

medium (pH 5.8), 0.5% sucrose, and 1.2 % Daishin agar (Duchefa Biochemie). Plates were stratified at

4 °C in the dark for two days, and then transferred to long day conditions (22 °C, 16 h of light and 8h of

dark) in a vertical position, under an angle of 70°. Four-day-old seedlings of comparable size were then

transferred to ½MS ager plates without sucrose and allowed to grow further for another 6-8 days. Plates

were then scanned with an Epson Perfection V700 scanner and primary root length, lateral root number

and average lateral root length from each genotype quantified through ImageJ software (National

Institutes of Health).

Page 112: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

110

Cloning and plant transformation

To generate pPLC7::GUS-SYFP reporter line, the PLC7 promoter was amplified from genomic DNA

using the following primers: PLC7proH3fw (5’-

CCCAAGCTTGATCCTATCAATATTCCTAATTCAGC-3’) and PLC7proNheIrev (5’-

CTAGCTAGCTTGAACAATTCCTCAAGTG-3’). The PCR product was cloned into pGEM-T easy

and sequenced. A HindIII-pPLC7-NheI fragment was then ligated into pJV-GUS-SYFP, cut with

HindIII and NheI. A pPLC7::GUS-SYFP1 fragment, cut with NotI and transferred to pGreenII-0229.

The MultiSite Gateway Three-Fragment Vector Construction Kit (www.lifetechnologies.com)

was used to generate PLC7 overexpression lines driven by the UBQ10 promotor (pUBQ10::PLC7).

Oligonucleotide primers (5’-

GGGGACAACGTTTGTACAAAAAAGCAGGCTATGTCGAAGCAAACATACAAAGT-3’ and 5’-

GGGGACCACTTTGTACAAGAAAGCTGGGTCACAAACTCCAACCGCACAAGAA-3’) including

attB1 and attB2 sites, were used to PCR from PLC7 cDNA and was cloned into the donor vector

(pDONR207) by using BP Clonase II enzyme mix to create entry clone BOX2. BOX1 was pGEM-

pUBQ10 entry clone containing attL4 and attR1 sites. BOX3 was pGEM-TNOS entry clone containing

attR2 and attL2 sites. Both entry clone were obtained from university of Amsterdam plant physiology

publish construct source. The three entry clones (BOX1, BOX2 and BOX3) and a destination vector

(pGreen0125) were used in MultiSite Gateway LR recombination reaction to create the expression

clone. For details, check: (Multi gateway protocol).

All constructs were transformed into Agrobacterium tumefaciens, strain GV3101, which was

subsequently used to transform Arabidopsis plants by floral dip (Clough and Bent, 1998). Homozygous

lines were selected in T3 generation and used for future experiments.

RNA extraction and Q-PCR

The primer pairs to check for PLC1- to PLC9-expression levels were obtained from Tasma et al. (2008).

Similarly, identification of CUC2- and MIR164A-expression levels were performed as described by

Bilsborough et al. (2011). Total RNA was extracted with Trizol reagent (Invitrogen, Carlsbad, CA) as

described by Pieterse (1998). Total RNA (1.5 µg) from 10-day-old seedlings, or 4-week old rosette

leaves, were converted to cDNA using oilgo-dT18 primers, dNTPs and SuperSccript III Reverse

Transcriptase (Invitrogen), according to the manufacturer’s instructions. Q-PCR was performed with

ABI 7500 Real-Time PCR System (Applied Biosystem). The relative gene expression was determined

by comparative threshold cycle value. Transcript levels were normalized by the level of SAND

(At2g28390; forward primer: 5’-AACTCTATGCAGCATTTGATCCACT-3’, reverse primer: 5’-

TGAAGGGACAAAGGTTGTGTATGTT-3’), or OTC (At1g75330; forward primer: 5’-

TGAAGGGACAAAGGTTGTGTATGTT-3’, reverse primer: 5’-CGCAGACAAAGTGGAATGGA-

3’) (Bilsborough et al., 2011; Han et al., 2013). Three biological- and two technical replicates were

performed to obtain the values for means and standard deviations (Han et al., 2013).

Page 113: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

111

Histochemical analyses for GUS activity

GUS staining was performed as described previously (Chapter 2; Zhang et al., 2017). Briefly,

transgenic plants carrying pPLC7::GUS-SYFP were grown for indicated times, after which specific

tissues were taken and incubated in an X-Gluc reaction solution containing 1mg/ml 5-bromo-4-chloro-3

indolyl-β-D-glucuronic acid (X-gluc), 50 mM phosphate buffer (pH 7.0), and 0.1% TX-100. Material

was incubated overnight at 37°C and the next day, cleared by 70% ethanol, and kept in that solution.

GUS staining was visualized under a stereo microscope (Leica MZFLIII) and digitalized with a

ThorLabs, CCD camera.

Seed staining and microscopy

To visualize seed coat mucilage, mature dry seeds were stained as described in Macquet et al. (2007).

Seeds were directly incubated in 0.03% (w/v) Ruthenium red, or after imbibition in 0.5 M EDTA, pH

8.0, for 90 min. For the latter, seeds were washed with water to remove the EDTA and then stained for

20 min with Ruthenium red. Stained seeds were routinely observed with a light microscope (Aristoplan;

Leitz). To visualize the seed surface and the adherent mucilage (AM) layer by confocal microscopy,

Calcofluor white (0.01%) and Pontamine S4B were used as staining solutions (Western et al., 2000;

Saez-Aguayo et al., 2014). Optical sections were obtained with an Olympus LX81 spectral confocal

laser-scanning microscope. A 405 nm diode laser line was used to excite Calcofluor white and the

fluorescence emission was detected between 412 and 490 nm. For Pontamine S4B a 561 nm diode laser

line was used and the detection was done between 570 and 650 nm. For comparisons of the signal

intensity within one experiment, the laser gain values were fixed. Three different batches of seeds were

analyzed and all of them showed the same phenotype. LUT green fire blue filter and LUT fire filter

(Image J) were applied to the Calcofluor white and Pontamine S4B images, respectively.

Determination of sugar composition by HPAEC-PAD

Sugar composition of wild-type and mutant seeds were measured as described previously

(Saez-Aguayo et al., 2017). 20 mg of seeds were ground in liquid nitrogen and extracted twice

in 80% ethanol with agitation for 1 h at room temperature followed by removal of lipids by

washing twice with methanol:chloroform (1:1) and twice with acetone. The final AIR was

dried overnight at RT and two mg was hydrolyzed for 1 h with 450 µL 2 M trifluoroacetic acid

(TFA) at 121 °C, using inositol as an internal control. TFA was evaporated at 60 °C with

nitrogen, samples washed two times with 250 µL of 100% isopropanol, and dried in a speed-

vac. Samples were dissolved in 200 µL Ultrapure water, clarified with a syringe filter (0.45

µm) and transferred to a new tube for analysis by High Performance Anion Exchange

Chromatography with Pulsed Amperometric Detection (HPAEC-PAD).

Page 114: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

112

The sugar quantification was done as described by Sáez-Aguayo et al. (2017). A

Dionex ICS3000 ion-chromatography system, equipped with a PAD detector, a CarboPac PA1

(4mm x 250mm) analytical column, and a CarboPac PA1 (4mm x 50mm) guard column were

used. The separation of neutral sugars was performed at 40ºC with a flow rate of 1 mL/min

using an isocratic gradient of 20 mM NaOH for 20 min followed by a wash with 200 mM

NaOH for 10 min. After every run, the column was re-equilibrated in 20 mM NaOH for 10

min. Separation of acidic sugars was performed by using 150 mM NaOAc and 100 mM NaOH

for 10 min at a flow rate of 1 mL/min at 40°C. Standard curves of neutral sugars (D-Fuc, L-

Rha, L-Ara, D-Gal, D-Glc, D-Xyl, and D-Man) or acidic sugars (D-GalA and D-GlcA) were

used for quantitation.

Leaf-shape analysis

Rosettes from 4-week old plants grown under long day condition were detached and photographed

immediately. Leaves were removed from the rosette and adhered to white paper and then scanned

(Epson Perfection V700 scanner). The 8th leaf was used for calculation. Blade length, width, perimeter,

area, serration number and serration levels were calculated from silhouettes using ImageJ software.

Leaf-serration levels are expressed as the distance from tip-to-midvein divided by the distance from

sinus-to-midvein, for indicated tooth (2nd-4th) (Kawamura et al., 2010).

Stomatal aperture

The stomatal aperture measurement was performed according to Distéfano et al., (2012) with minor

modifications. The stomatal aperture treatments were performed on epidermal strips excised from the

abaxial side of fully expanded Arabidopsis leaves. Epidermal peels from leaves of 3-week-old plants,

grown at 22°C under 16 h of light and 8h of dark, were stripped and immediately floated in opening

buffer (5 mM MES-KOH (pH 6.1), 50 mM KCl) for 3 h. The strips were subsequently maintained in

the same opening buffer and exposed to different ABA concentration. After 90 min, stomata were

digitized using a Nikon DS-Fi 1 camera, coupled to a Nikon Eclipse Ti microscope. The stomatal-

aperture width was measured using ImageJ software (National Institute of Health). 32Pi-phospholipid labelling, extraction and analysis

Developing seeds at 10 DAP were carefully removed from silique. Mature seeds were sterilized and

stratified on ½MS (pH 5.8) as described and germinated under long day conditions for around 20 h

when testa ruptured. Both developing and germinating seeds were then transferred to 200 µl labeling

buffer (2.5 mM MES, pH 5.8, 1 mM KCl) containing 5-10 µCi 32PO43- (32Pi) (carrier free; Perklin-Elmer)

in a 2 ml Eppendorf, Safelock microcentrifuge tube for 24 h.

Five-day-old seedlings were labeled similarly, except for 3 h labeling time. Samples were

treated by adding 200 µl labeling buffer with or without 600 mM sorbitol for 30 mins.

Page 115: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

113

Labeling an subsequent treatments were stopped by adding perchloric acid (final concentration,

5% by vol.) for 5-10 min, after which lipids were extracted and the phosphoinositides and PA separated

from the rest of the phospholipids by thin-layer chromatography (TLC) as described in detail by

Munnik & Zarza (2013). Radioactive phospholipids were visualized by autoradiography and quantified

by phosphoimaging (Molecular Dynamics, Sunnyvale, CA, USA). Individual phospholipid levels are

expressed as the percentage of the total 32P-lipid fraction.

Drought tolerance assays

Drought assay were performed as described previously (Hua et al., 2012; Osakabe et al., 2013) with

some changes. Seeds were stratified at 4°C in the dark and sowed in soil pots (4.5 cm x 4.5 cm x 7.5 cm)

containing equal amounts (80 g) of soil. Nine plants per pot were grown under short-day conditions (22

°C with 12 h light/12 h dark) for 4 weeks, and then subjected to dehydration by withholding them for

water for 2 weeks, while control plants were normally watered. At this point, plants were photographed.

Each experiment used 36 plants per genotype and experiments were repeated at least 3 times.

To assay water-loss, rosettes from 4-week-old plants were detached and FW determined every

hour by weighing. Water content was calculated as the percentage from the initial FW. Twenty plants

were used for each experiment, which was repeated at least 3 times.

ACKNOWLEDGEMENTS

This work was financially supported by the China Scholarship Council (CSC File No. 201206300058 to

Q.Z) and by the Netherlands Organization for Scientific Research (NWO; 867.15.020 to T.M.).

REFERENCES

Allen GJ, and Sanders D (1994) Osmotic stress enhances the competence of Beta vulgaris vacuoles to respond to inositol

1,4,5-trisphosphate. Plant J 6:687–695. Anderson CT, Carroll A, Akhmetova L, and Somerville C (2010) Real-Time Imaging of Cellulose Reorientation during Cell

Wall Expansion in Arabidopsis Roots. Plant Physiol 152:787–796. Arisz SA, Testerink C, and Munnik T (2009) Plant PA signaling via diacylglycerol kinase. Biochim Biophys Acta 1791:869–

875. Arisz SA, van Wijk R, Roels W, Zhu JK, Haring MA, and Munnik T (2013) Rapid phosphatidic acid accumulation in response

to low temperature stress in Arabidopsis is generated through diacylglycerol kinase. Front Plant Sci 4:1–15. Balla T (2013) Phosphoinositides: tiny lipids with giant impact on cell regulation. Physiol Rev 93:1019–1137. Basu D, Tian L, DeBrosse T, Poirier E, Emch K, Herock H, Travers A, and Showalter AM (2016) Glycosylation of a

Fasciclin-Like Arabinogalactan-Protein (SOS5) mediates root growth and seed mucilage adherence via a cell wall receptor-like kinase (FEI1/ FEI2) Pathway in Arabidopsis. PLoS One 11:1–27.

Bauer H, Ache P, Lautner S, Fromm J, Hartung W, Al-Rasheid KAS, Sonnewald S, Sonnewald U, Kneitz S, Lachmann N, Mendel RR, Bittner F, Hetherington AM, and Hedrich R (2013) The stomatal response to reduced relative humidity requires guard cell-autonomous ABA synthesis. Curr Biol 23:53–57.

Ben-Tov D, Abraham Y, Stav S, Thompson K, Loraine A, Elbaum R, de Souza A, Pauly M, Kieber JJ, and Harpaz-Saad S (2015) COBRA-LIKE2, a member of the glycosylphosphatidylinositol-anchored COBRA-LIKE family, plays a role in cellulose deposition in arabidopsis seed coat mucilage secretory cells. Plant Physiol 167:711–24.

Bilsborough GD, Runions A, Barkoulas M, Jenkins HW, Hasson A, and Galinha C (2011) Model for the regulation of Arabidopsis thaliana leaf margin development. PNAS 108:3424–3429.

Blatt MR, Thiel G, and Trentham DR (1990) Reversible inactivation of K+ channels of Vicia stomatal guard cells following the photolysis of caged inositol 1,4,5-trisphosphate. Nature 346:766–769.

Byrne ME (2005) Networks in leaf development. Curr Opin Plant Biol 8:59–66.

Page 116: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

114

Carpita NC, and Gibeaut DM (1993) Structural models of primary cell walls in flowering plants: consistency of molecular structure with the physical properties of the walls during growth. Plant J 3:1–30.

Clough SJ, and Bent AF (1998) Floral dip: A simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J 16:735–743.

Cosgrove DJ (1997) Assembly and enlargement of the primary cell wall in plants. Annu Rev Cell Dev Biol 13:171–201. D’Ambrosio JM, Couto D, Fabro G, Scuffi1 D, Lamattina L, Munnik T, Álvarez ME, Andersson MX, Zipfel C, and Laxalt

AM (2017) Phosphoinositide-specific phospholipase C2 (PLC2)associates with RBOHD and modulates ROS production and MAMP-triggered immunity in Arabidopsis. Plant J submitted.

Di Fino L, D’Ambrosio J, Tejos R, van Wijk R, Lamattina T, Munnik T, Pagnussat G, and Laxalt AM (2017) Arabidopsis phosphatidylinositol-phospholipase C2 (PLC2) is involved in female gametogenesis and embryo development. Planta 2:1–12.

Distéfano AM, Scuffi D, García-Mata C, Lamattina L, and Laxalt AM (2012) Phospholipase D?? is involved in nitric oxide-induced stomatal closure. Planta 236:1899–1907.

Dowd PE, Coursol S, Skirpan AL, Kao T, and Gilroy S (2006) Petunia phospholipase C1 is involved in pollen tube growth. Plant Cell 18:1438–1453.

Fleming AJ (2005) The control of leaf development. New Phytol 166:9–20. Gao K, Liu YL, Li B, Zhou RG, Sun DY, and Zheng SZ (2014) Arabidopsis thaliana phosphoinositide-specific phospholipase

C isoform 3 (AtPLC3) and AtPLC9 have an additive effect on thermotolerance. Plant Cell Physiol 55:1873–1883. Georges F, Das S, Ray H, Bock C, Nokhrina K, Kolla VA, and Keller W (2009) 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 32:1664–1681.

Gillaspy GE (2013) The role of phosphoinositides and inositol phosphates in plant cell signaling, in Lipid-mediated protein signaling (Capelluto DGS ed) pp 141–157.

Gilroy S, Read ND, and Trewavas a J (1990) Elevation of cytoplasmic calcium by caged calcium or caged inositol triphosphate initiates stomatal closure. Nature 346:769–771.

Han B, Yang Z, Samma MK, Wang R, and Shen W (2013) Systematic validation of candidate reference genes for qRT-PCR normalization under iron deficiency in Arabidopsis. BioMetals 26:403–413.

Harpaz-Saad S, McFarlane HE, Xu S, Divi UK, Forward B, Western TL, and Kieber JJ (2011) Cellulose synthesis via the FEI2 RLK/SOS5 pathway and CELLULOSE SYNTHASE 5 is required for the structure of seed coat mucilage in Arabidopsis. Plant J 68:941–953.

Hay A, Barkoulas M, and Tsiantis M (2006) ASYMMETRIC LEAVES1 and auxin activities converge to repress BREVIPEDICELLUS expression and promote leaf development in Arabidopsis. Development 133:3955–3961.

Hay A, and Tsiantis M (2006) The genetic basis for differences in leaf form between Arabidopsis thaliana and its wild relative Cardamine hirsuta. Nat Genet 38:942–947.

Heilmann I (2016) Phosphoinositide signaling in plant development. Development 143:2044–2055. Helling D, Possart A, Cottier S, Klahre U, and Kost B (2006) Pollen tube tip growth depends on plasma membrane

polarization mediated by tobacco PLC3 activity and endocytic membrane recycling. Plant Cell 18:3519–3534. Hou Q, Ufer G, and Bartels D (2016) Lipid signalling in plant responses to abiotic stress. Plant, Cell Environ 39:1029–1048. Hu R, Li J, Wang X, Zhao X, Yang X, Tang Q, He G, Zhou G, and Kong Y (2016) Xylan synthesized by Irregular Xylem 14

(IRX14) maintains the structure of seed coat mucilage in Arabidopsis. J Exp Bot 67:1243–1257. Hu R, Li J, Yang X, Zhao X, Wang X, Tang Q, He G, Zhou G, and Kong Y (2016) Irregular xylem 7 (IRX7) is required for

anchoring seed coat mucilage in Arabidopsis. Plant Mol Biol 92:25–38. Hua D, Wang C, He J, Liao H, Duan Y, Zhu Z, Guo Y, Chen Z, and Gong Z (2012) A plasma membrane receptor kinase,

GHR1, mediates abscisic acid- and hydrogen peroxide-regulated stomatal movement in Arabidopsis. Plant Cell 24:2546–61.

Hunt L, and Gray JE (2001) ABA signalling: A messenger’s FIERY fate. Curr Biol 11:968–970. Hunt L, Mills LN, Pical C, Leckie CP, Aitken FL, Kopka J, Mueller-Roeber B, McAinsh MR, Hetherington AM, and Gray JE

(2003) Phospholipase C is required for the control of stomatal aperture by ABA. Plant J 34:47–55. Irvine RF (2006) Nuclear inositide signalling-expansion, structures and clarification. Biochim Biophys Acta 1761:505–508. Ischebeck T, Werner S, Krishnamoorthy P, Lerche J, Meijón M, Stenzel I, Löfke C, Wiessner T, Im YJ, Perera IY, Iven T,

Feussner I, Busch W, Boss WF, Teichmann T, Hause B, Persson S, and Heilmann I (2013) Phosphatidylinositol 4,5-bisphosphate influences PIN polarization by controlling clathrin-mediated membrane trafficking in Arabidopsis. Plant Cell 25:4894–4911.

Kawamura E, Horiguchi G, and Tsukaya H (2010) Mechanisms of leaf tooth formation in Arabidopsis. Plant J 62:429–441. Kuo HF, Chang TY, Chiang SF, Wang W Di, Charng YY, and Chiou TJ (2014) Arabidopsis inositol pentakisphosphate 2-

kinase, AtIPK1, is required for growth and modulates phosphate homeostasis at the transcriptional level. Plant J 80:503–515.

Laha D, Johnen P, Azevedo C, Dynowski M, Weiß M, Capolicchio S, Mao H, Iven T, Steenbergen M, Freyer M, Gaugler P, de Campos MKF, Zheng N, Feussner I, Jessen HJ, Van Wees SCM, Saiardi A, and Schaaf G (2015) VIH2 regulates the synthesis of inositol pyrophosphate InsP8 and jasmonate-dependent defenses in Arabidopsis. Plant Cell 27:1082–1097.

Laha D, Parvin N, Dynowski M, Johnen P, Mao H, Bitters ST, Zheng N, and Schaaf G (2016) Inositol polyphosphate binding specificity of the jasmonate receptor complex. Plant Physiol 171:2364–2370.

Lee HS, Lee DH, Cho HK, Kim SH, Auh JH, and Pai HS (2015) 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 27:417–431.

Lemtiri-Chlieh F, MacRobbie EAC, and Brearley CA (2000) Inositol hexakisphosphate is a physiological signal regulating the K+-inward rectifying conductance in guard cells. Proc Natl Acad Sci U S A 97:8687–8692.

Page 117: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

115

Lemtiri-Chlieh F, MacRobbie EAC, Webb AAR, Manison NF, Brownlee C, Skepper JN, Chen J, Prestwich GD, and Brearley CA (2003) Inositol hexakisphosphate mobilizes an endomembrane store of calcium in guard cells. Proc Natl Acad Sci U S A 100:10091–10095.

Li J, Pleskot R, Henty-Ridilla JL, Blanchoin L, Potocký M, and Staiger CJ (2012) Arabidopsis capping protein senses cellular phosphatidic acid levels and transduces these into changes in actin cytoskeleton dynamics. Plant Signal Behav 7:1727–1730.

Li L, He Y, Wang Y, Zhao S, Chen X, Ye T, Wu Y, and Wu Y (2015) Arabidopsis PLC2 is involved in auxin-modulated reproductive development. Plant J 84:504–515.

Li M, Hong Y, and Wang X (2009) Phospholipase D- and phosphatidic acid-mediated signaling in plants. Biochim Biophys Acta - Mol Cell Biol Lipids 1791:927–935.

Liu Y, Su Y, and Wang X (2013) Phosphatidic Acid-Mediated Signaling, in Lipid-mediated Protein Signaling (Capelluto DGS ed) pp 159–176.

Macquet A, Ralet MC, Kronenberger J, Marion-Poll A, and North HM (2007) In situ, chemical and macromolecular study of the composition of Arabidopsis thaliana seed coat mucilage. Plant Cell Physiol 48:984–999.

Martin TF. (2001) PI(4,5)P2 regulation of surface membrane traffic. Curr Opin Cell Biol 13:493–499. Meijer HJG, van Himbergen JAJ, Musgrave A, and Munnik T (2017) Acclimation to salt modifies the activation of several

osmotic stress-activated lipid signalling pathways in Chlamydomonas. Phytochemistry 135:64–72. Mendu V, Griffiths JS, Persson S, Stork J, Downie AB, Voiniciuc C, Haughn GW, and DeBolt S (2011) Subfunctionalization

of Cellulose Synthases in Seed Coat Epidermal Cells Mediates Secondary Radial Wall Synthesis and Mucilage Attachment. Plant Physiol 157:441–453.

Michell RH (2008) Inositol derivatives: evolution and functions. Nat Rev 9:151–161. Monserrate JP, and York JD (2010) Inositol phosphate synthesis and the nuclear processes they affect. Curr Opin Cell Biol

22:365–373. Mosblech A, König S, Stenzel I, Grzeganek P, Feussner I, and Heilmann I (2008) Phosphoinositide and inositolpolyphosphate

signalling in defense responses of Arabidopsis thaliana challenged by mechanical wounding. Mol Plant 1:249–261, The Authors 2007. All rights reserved.

Mosblech A, Thurow C, Gatz C, Feussner I, and Heilmann I (2011) Jasmonic acid perception by COI1 involves inositol polyphosphates in Arabidopsis thaliana. Plant J 65:949–957.

Mueller-roeber B, and Pical C (2002) Inositol phospholipid metabolism in Arabidopsis . characterized and putative isoforms of inositol phospholipid kinase and phosphoinositide-specific phospholipase C. Plant Physiol 130:22–46.

Munnik T (2014) PI-PLC: Phosphoinositide-phospholipase C in plant signalling, in Phospholipases in Plant Signaling (Wang X ed) pp 27–54.

Munnik T, and Nielsen E (2011) Green light for polyphosphoinositide signals in plants. Curr Opin Plant Biol 14:489–497. Munnik T, and Testerink C (2009) Plant phospholipid signaling: “in a nutshell”. J Lipid Res 50:S260–S265. Munnik T, and Vermeer JEM (2010) Osmotic stress-induced phosphoinositide and inositol phosphate signalling in plants.

Plant Cell Environ 33:655–669. Murphy AM, Otto B, Brearley CA, Carr JP, and Hanke DE (2008) A role for inositol hexakisphosphate in the maintenance of

basal resistance to plant pathogens. Plant J 56:638–652. Nikovics K, Blein T, Peaucelle A, Ishida T, Morin H, Aida M, and Laufs P (2006) The balance between the MIR164A and

CUC2 genes controls leaf margin serration in Arabidopsis. Plant Cell 18:2929–2945. Osakabe Y, Arinaga N, Umezawa T, Katsura S, Nagamachi K, Tanaka H, Ohiraki H, Yamada K, Seo S-U, Abo M, Yoshimura

E, Shinozaki K, and Yamaguchi-Shinozaki K (2013) Osmotic stress responses and plant growth controlled by potassium transporters in Arabidopsis. Plant Cell 25:609–624.

Pieterse CMJ (1998) A novel signaling pathway controlling induced systemic resistance in Arabidopsis. Plant Cell 10:1571–1580.

Pleskot R, Li J, Žárský V, Potocký M, and Staiger CJ (2017) Regulation of cytoskeletal dynamics by phospholipase D and phosphatidic acid. Trends Plant Sci 18:496–504.

Pleskot R, Potocký M, Pejchar Př, Linek J, Bezvoda R, Martinec J, Valentová O, Novotná Z, and Žárský V (2010) Mutual regulation of plant phospholipase D and the actin cytoskeleton. Plant J 62:494–507.

Pokotylo I, Kolesnikov Y, Kravets V, Zachowski A, and Ruelland E (2014) Plant phosphoinositide-dependent phospholipases C: Variations around a canonical theme. Biochimie 96:144–157.

Puga MI, Mateos I, Charukesi R, Wang Z, Franco-Zorrilla JM, de Lorenzo L, Irigoyen ML, Masiero S, Bustos R, Rodríguez J, Leyva A, Rubio V, Sommer H, and Paz-Ares J (2014) SPX1 is a phosphate-dependent inhibitor of Phosphate Starvation Response 1 in Arabidopsis. Proc Natl Acad Sci U S A 111:14947–14952.

Repp A, Mikami K, Mittmann F, and Hartmann E (2004) Phosphoinositide-specific phospholipase C is involved in cytokinin and gravity responses in the moss Physcomitrella patens. Plant J 40:250–259.

Saez-Aguayo S, Rautengarten C, Temple H, Sanhueza D, Ejsmentewicz T, Sandoval-Ibañez O, Doñas-Cofré DA, Parra-Rojas JP, Ebert B, Lehner A, Mollet J-C, Dupree P, Scheller H V., Heazlewood JL, Reyes FC, and Orellana A (2017) UUAT1 Is a Golgi-Localized UDP-Uronic Acid Transporter that Modulates the Polysaccharide Composition of Arabidopsis Seed Mucilage. Plant Cell doi:10.1105.

Saez-Aguayo S, Rondeau-Mouro C, Macquet A, Kronholm I, Ralet MC, Berger A, Sallé C, Poulain D, Granier F, Botran L, Loudet O, de Meaux J, Marion-Poll A, and North HM (2014) Local Evolution of Seed Flotation in Arabidopsis. PLoS Genet 10:13–15.

Sanchez JP, and Chua NH (2001) Arabidopsis PLC1 is required for secondary responses to abscisic acid signals. Plant Cell 13:1143–1154.

Scarpella E, Marcos D, Friml J, Berleth T, Scarpella E, Marcos D, and Berleth T (2006) Control of leaf vascular patterning by polar auxin transport Control of leaf vascular patterning by polar auxin transport. 1015–1027.

Page 118: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

116

Sean R, Pedro L, Ruth R, and Suzanne R (2010) Abscisic acid:emergence of a core signaling network. Annu Rev Plant Biol 61:651–679.

Sheard LB, Tan X, Mao H, Withers J, Ben-Nissan G, Hinds TR, Kobayashi Y, Hsu F-F, Sharon M, Browse J, He SY, Rizo J, Howe GA, and Zheng N (2010) Jasmonate perception by inositol-phosphate-potentiated COI1-JAZ co-receptor. Nature 468:400–405, Nature Publishing Group, a division of Macmillan Publishers Limited. All Rights Reserved.

Song M, Liu S, Zhou Z, and Han Y (2008) TfPLC1, a gene encoding phosphoinositide-specific phospholipase C, is predominantly expressed in reproductive organs in Torenia fournieri. Sex Plant Reprod 21:259–267.

Stevenson JM, Perera IY, Heilmann I, Persson S, and Boss WF (2000) Inositol signaling and plant growth. Trends Plant Sci 5:252–258.

Sullivan S, Ralet MM-C, Berger A, Diatloff E, Bischoff V, Gonneau M, Marion-Poll A, and North HM (2011) CESA5 is required for the synthesis of cellulose with a role in structuring the adherent mucilage of Arabidopsis seeds. Plant Physiol 156:1725–39.

Tan X, Calderon-Villalobos LI a, Sharon M, Zheng C, Robinson C V, Estelle M, and Zheng N (2007) Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature 446:640–645.

Tasma IM, Brendel V, Whitham S a., and Bhattacharyya MK (2008) Expression and evolution of the phosphoinositide-specific phospholipase C gene family in Arabidopsis thaliana. Plant Physiol Biochem 46:627–637.

Tejos R, Sauer M, Vanneste S, Palacios-Gomez M, Li H, Heilmann M, van Wijk R, Vermeer JEM, Heilmann I, Munnik T, and Friml J (2014) Bipolar Plasma Membrane Distribution of Phosphoinositides and Their Requirement for Auxin-Mediated Cell Polarity and Patterning in Arabidopsis. Plant Cell 26:2114–2128.

Testerink C, and Munnik T (2011) Molecular, cellular, and physiological responses to phosphatidic acid formation in plants. J Exp Bot 62:2349–2361.

Thomas C, and Staiger CJ (2014) A dynamic interplay between membranes and the cytoskeleton critical for cell development and signaling. Front Plant Sci 5:335.

Thota SG, and Bhandari R (2015) The emerging roles of inositol pyrophosphates in eukaryotic cell physiology. J Biosci 40:593–605.

Tripathy MK, Tyagi W, Goswami M, Kaul T, Singla-Pareek SL, Deswal R, Reddy MK, and Sopory SK (2011) Characterization and Functional Validation of Tobacco PLC Delta for Abiotic Stress Tolerance. Plant Mol Biol Report 30:488–497.

Tsukaya H (2006) Mechanism of leaf-shape determination. Annu Rev Plant Biol 57:477–496. Tsukaya H, and Uchimiya H (1997) Genetic analyses of the formation of the serrated margin of leaf blades in Arabidopsis:

Combination of a mutational analysis of leaf morphogenesis with the characterization of a specific marker gene expressed in hydathodes and stipules. Mol Gen Genet 256:231–238.

van Leeuwen W, Vermeer JEM, Gadella TWJ, and Munnik T (2007) Visualization of phosphatidylinositol 4,5-bisphosphate in the plasma membrane of suspension-cultured tobacco BY-2 cells and whole Arabidopsis seedlings. Plant J 52:1014–1026.

Voiniciuc C, Schmidt MH-W, Berger A, Yang B, Ebert B, Scheller H V., North HM, Usadel B, and Günl M (2015) MUCILAGE-RELATED10 Produces Galactoglucomannan That Maintains Pectin and Cellulose Architecture in Arabidopsis Seed Mucilage. Plant Physiol 169:403–420.

Voiniciuc C, Yang B, Schmidt MH-W, Günl M, and Usadel B (2015) Starting to gel: how Arabidopsis seed coat epidermal cells produce specialized secondary cell walls. Int J Mol Sci 16:3452–3473.

Vossen JH, Abd-El-Haliem A, Fradin EF, Van Den Berg GCM, Ekengren SK, Meijer HJG, Seifi A, Bai Y, Ten Have A, Munnik T, Thomma BPHJ, and Joosten MHAJ (2010) 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 62:224–239.

Wallace IS, and Anderson CT (2012) Small Molecule Probes for Plant Cell Wall Polysaccharide Imaging. Front Plant Sci 3:1–8.

Wang CR, Yang AF, Yue GD, Gao Q, Yin HY, and Zhang JR (2008) Enhanced expression of phospholipase C 1 (ZmPLC1) improves drought tolerance in transgenic maize. Planta 227:1127–1140.

Wang X, Devaiah SP, Zhang W, and Welti R (2006) Signaling functions of phosphatidic acid. Prog Lipid Res 45:250–278. Western TL, Skinner DJ, and Haughn GW (2000) Differentiation of mucilage secretory cells of the Arabidopsis seed coat.

Plant Physiol 122:345–356. Wheeler GL, and Brownlee C (2008) Ca2+ signalling in plants and green algae - changing channels. Trends Plant Sci 13:506–

514. Wild R, Gerasimaite R, Jung J-Y, Truffault V, Pavlovic I, Schmidt A, Saiardi A, Jessen HJ, Poirier Y, Hothorn M, and Mayer

A (2016) Control of eukaryotic phosphate homeostasis by inositol polyphosphate sensor domains. Science (80- ) 352:986 LP-990.

Willats WGT, Mccartney L, and Knox JP (2001) Pectinas Polissacarideos. Williams SP, Gillaspy GE, and Perera IY (2015) Biosynthesis and possible functions of inositol pyrophosphates in plants.

Front Plant Sci 6:1–12. Yu L, Shi D, Li J, Kong Y, Yu Y, Chai G, Hu R, Wang J, Hahn MG, and Zhou G (2014) CELLULOSE SYNTHASE-LIKE

A2, a glucomannan synthase, is involved in maintaining adherent mucilage structure in Arabidopsis seed. Plant Physiol 164:1842–1856.

Zheng S, Liu Y, Li B, Shang Z, Zhou R, and Sun D (2012) Phosphoinositide-specific phospholipase C9 is involved in the thermotolerance of Arabidopsis. Plant J 69:689–700.

Page 119: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

RoleforPLC7inArabidopsis

117

SUPPLEMENTAL DATA

Supplemental Figure S1. Sugar composition of wild type- and plc5/7-mutant seeds.Sugars were extracted from dry seeds and quantified by HPAEC-PAD. Quantities were corrected through internal standards, and transformed into mg of sugar per gram of dry material. Values represent the means of triplicates ± SE of three independent experiments.

Supplemental Figure S2. Decreased ABA sensitivity of plc7 mutants in stomatal closure. Leaves from 3-weeks old plants were stripped and peels were incubated in opening buffer with light for 3 h until stomata were fully open. Peels were then treated with different concentrations of ABA for 90 mins, after which stomata were digitized and the aperture width measured for wild type and plc7-3 (left) or plc7-4 (right). Values are means ± SE of at least three independents (n >100). Data was analyzed by 2-way ANOVA. Statistical significant differences between genotypes are indicated by letters (P<0.05, Dunn’s method).

Page 120: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter4

118

Page 121: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generaldiscussion

119

Chapter 5 General discussion

Qianqian Zhang,1,2 Michel A. Haring,1 & Teun Munnik1,2

1Section Plant Physiology or

2Section Plant Cell Biology, Swammerdam Institute for Life

Sciences, University of Amsterdam, Science Park 904, Amsterdam, 1098 XH, The Netherlands

Page 122: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter5

120

Comparison of Arabidopsis PLC3, PLC5 and PLC7 -- similarities and differences

In this thesis, we explored the roles of three PLC genes - PLC3, PLC5 and PLC7 - in Arabidopsis

development and stress responses at the molecular, physiological and biochemical level. Arabidopsis

contains 9 PLCs, which share several conserved protein domains, including an EF-hand domain, the

catalytic X- and Y-domains, and a C2 domain (Tasma et al., 2008; Munnik, 2014). In addition, the

genes for PLC3, PLC5 and PLC7 localize to chromosome 3, 4 and 5 respectively and may have evolved

differently (Hunt et al., 2004; Tasma et al., 2008). The difference in protein structure and evolutionary

history is likely to result into functional differences between the nine PLC homologs of Arabidopsis

The discussion below will focus on the similarities and differences between PLC3, PLC5 and

PLC7 in terms of their function in plant development and stress. This will help to put the results

presented in thesis, in a broader perspective.

Expression

The PLC-expression analyses in this thesis were based on promoter-GUS analysis. The expression

patterns of PLC3, PLC5 and PLC7 have similarities, but also some differences (see Table 1). The most

striking communality in expression is that they are all vascular-expressed PLCs in both vegetative

(shoot and root) and reproductive organs (flower). Cell-specific gene profiling analyses in roots

revealed that these PLCs are expressed in phloem-related cells, e.g. companion cells

(http://bar.utoronto.ca/efp/cgi-bin/efpWeb.cgi). Phloem is responsible for transferring soluble organic

compounds, especially sucrose made by photosynthesis in the leaves, to tissues where it is required,

such as roots (Turgeon and Wolf, 2009; De Schepper et al., 2013). PLC expression detected in phloem

suggests a potential role in plant development. A closer look at the expression in root vasculature

revealed a typical "segmented-expression" pattern for PLC3 and PLC5 in roots, but not for PLC7. This

could to be linked to the lateral root formation phenotype in plants that have lost the function of PLC3

or PLC5, but not PLC7 (see discussion below).

Table 1. Promoter-GUS analysis of PLC expression.

expression vascular tissue

segment expression

in root

developing seed

chalazal

developing seed coat

mature seed

embryo

trichome guard cell

PLC3 yes yes yes no yes yes (base only) yes

PLC5 yes yes yes yes yes yes yes

PLC7 yes no yes yes yes yes no

Page 123: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generaldiscussion

121

PLC is expressed though out the whole life cycle, based on current knowledge (Hunt et al.,

2004; Tasma et al., 2008; Munnik, 2014; http://bar.utoronto.ca/efp/cgi-bin/efpWeb.cgi). We found all

three PLCs were expressed in the chalazal during seed development. This is the tissue that serves as a

channel for passing nutrients from the mother plant to the embryo (Xu et al., 2016). Therefore, PLCs

might exert the same function here as they do in the vasculature. PLC5 and PLC7 were expressed in the

developing seed coat but PLC3 was not expressed in this tissue. The plc5plc7 double mutant exhibited a

non-adherent seed coat mucilage phenotype, which might have to do with the expression of both PLC5

and PLC7 in the seed coat. The expression of all three genes was also detected in germinating seeds.

pPLC3::GUS-YFP embryo was intensively stained by X-Gluc (Chapter 2, Figure 4b), this was much

less obvious for PLC5::GUS (Chapter 3, Figure 2a) and PLC7::GUS-SYPF (Chapter 4, Figure 1a),

which could be a link to the slower germination rate of the plc3 mutants only.

Interestingly, all three PLCs were expressed in trichomes. The expression of PLC5 and PLC7

was detected in the whole trichome body, but PLC3 only in the basal cells of the trichome, which

connect to the leaf-vascular tissue and seems to be part of the vasculature. In Arabidopsis, trichomes are

non-secreting epidermal cells, which protect plants from adverse conditions and also serve as a source

of phytochemicals (Schwab et al., 2000; Pattanaik et al., 2014). A numbers of genes have been

identified to be involved in trichome development. The consistent expression of PLC5 and PLC7 in

trichomes hints at a possible role in morphogenesis and/or function, even though we did not observe

any phenotype in the single- and double plc mutants that were viable; obviously the PLCs may be

redundant. An inducible-amiPLC3 line in a plc5/7 background may reveal more on this.

In guard cells, both PLC3 and PLC5 were expressed, which confirms data provided by the

Arabidopsis eFP browser (http://bar.utoronto.ca/efp/cgi-bin/efpWeb.cgi). However, PLC7::GUS-SYFP

staining was not detected in guard cells. According to online database from eFP browser, PLC7 could

be induced in the guard cells after exogenous ABA application. We performed GUS staining on the leaf

peel from pPLC7::GUS-SYFP plants after ABA treatment, but found no GUS signal (data not shown).

The online data is based on microarray technique and guard cell protoplasts were used as plant material,

which is different with our GUS staining analysis using seedlings, or freshly peeled leaf epidermis.

The promoter-GUS analysis and light microscopy studies cannot provide information about the

PLC localization at the cellular level. The precise location in the cell will help to understand how plant

PLCs work in various biological events. Thus, making transgenic plants that express PLC-GFP fusions

driven by their own promoters is essential and will be our next step. We attempted by complementing

the plc5-1 mutant with an N-terminal YFP-PLC5, driven by the PLC5 promoter, but no fluorescence

could be observed anymore in T2- and T3 generations while antibiotic resistance, genotyping and

phenotyping (root architecture) suggested that the wild-type gene was there. We suspect that the N-

terminal FP-fusion is lethal/interferes too much and that we select for plants that have been able to

loose the FP, as similar results were obtained for PLC2 (R. van Wijk and T. Munnik, unpublished).

Page 124: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter5

122

Currently, the lab is testing this for new N- and C-terminal FP-PLC fusions (M. van Hooren, T.

Munnik).

Role for PLCs in seed germination

We discovered a number of phenotypes in plc3, plc5 and plc7 single mutants, but also double plc3/5

and plc5/7 mutants (table 2). Among the mutants studied, only plc3 exhibited a delayed seed

germination phenotype (Chapter 2, Figure 4a). Seed germination is a very complex process, involving

many factors, such as light, hormones and sugars etc. (Bentsink and Koornneef, 2008). Among these,

abscisic acid (ABA) and gibberellin (GA) are best known for their crucial roles. ABA inhibits seed

germination (Nambara et al., 2010; Nakashima & Yamaguchi-Shinozaki, 2013) while, GA acts

stimulates germination (Yamaguchi and Kamiya, 2001). We compared the responses of plc3 mutants to

both ABA (Chapter 2; Fig.5) and GA (Chapter 2; Supplemental Fig. S7) with that of wild-type seeds.

plc3 mutant seeds germinated faster than wild-type when treated with exogenous ABA, indicating an

increased ABA insensitivity. Therefore, the slower germination of plc3 compared to wild type under

normal condition cannot be caused by hypersensitivity to ABA. In response to GA, plc3 mutants

seemed to be hypersensitive, which again does not explain their slower germination phenotype. One

possibility is that the endogenous GA level in plc3 mutants is lower. Similarly, plc3 mutants’

insensitivity to ABA might correspond to a higher level of ABA in seeds. Hence, checking hormone

level in plc3 mutant seeds will be essential. Apart from hormones, sugar is another important factor

during seed germination. The link between sugar metabolism and PLC signaling could be the Raffinose

Family Oligosaccharides (RFOs), whose synthesis might require PLC-dependent inositol (Chapter 2

discussion). However, the analysis of soluble carbohydrate composition in seeds showed no significant

differences between wild type and plc3 mutants. Of course, changes could be very local, so it is also

possible that these differences remain unobserved (Chapter 2; Fig. 4c).

A role for PLC in ABA signaling

Both plc mutants and PLC-overexpression lines of PLC3, 5 and 7 appear to have changes in ABA

related physiological responses. In addition, PLC3 and PLC5 are strongly expressed in guard cells.

Both plc3 (Chapter 2; Fig. 5c) and plc7 (Chapter 4; Supplemental Fig. 3) KO mutants did not close their

stomata as much as wild type did during ABA treatment of epidermal leaf strips, which indicates an

insensitivity for ABA. Furthermore, in both PLC3 and PLC5 overexpressor lines stomata were more

closed in control conditions and while the guard cells in epidermal peels appeared to be less responsive

to ABA. PLC7-OE lines did not show these changes. Together, these results hint to an involvement of

PLC in ABA-induced stomatal closure. The T-DNA insertion plc mutants provide a tool to investigate

the link between PLC and IPPs. According to our observation, using 3H-Inositol pre-labeled seedlings

and HPLC analyses, plc3 mutants did not any show change in IP3 or IP6 levels, but a small decrease in

either IP7 or IP8, was found depending on the seedling age. Plants contain plenty of IP6, which is

predominantly formed from MIPS-generated Ins3P (Munnik and Vermeer, 2010). The possible changes

Page 125: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generaldiscussion

123

in PLC-derived IP6 in guard cells will be difficult to observe in a huge background other cells.

Nevertheless, our results suggest that higher IPP and PP-IPP levels could be involved in PLC signaling.

The measurement of IPPs and PP-IPPs will be continued in other plc mutants with ABA and other

stresses application, in this way we will get closer to the real PLC signaling pathway in plant!

Alternatively, fluorescent biosensors expressed in guard cells could help studying changes in PIP2 or

PIP levels during ABA induced stomatal closure.

Table 2. Phenotypes that are exhibited in plc knock-out or knock-down mutants

plc mutant phenotype

delayed seed germination

insensitive to ABA

shorter primary

root

fewer lateral roots

loose seed coat mucilage

stronger leaf

serration seed germination

stomatal closure

plc3 yes yes yes yes yes no no

plc5 no no no yes yes no no

plc7 no no yes no no no no

plc3/5 no no no yes yes no no

plc5/7 no no no no no yes yes

The role for PLC in lateral root formation

PLC3::GUS-YFP and PLC5::GUS displayed a segmented-expression pattern in roots, which could be

associated with the decrease in lateral root formation in plc3- and plc5 mutants. PLC7::GUS-SYFP did

not have this expression pattern, and plc7 mutants did not display the lateral root phenotype either. The

double mutant plc3/5 had a similar root phenotype as the single mutants, with only a minor increase in

severity (Chapter 3; Fig. S2), whereas no changes were found in the root architecture of the plc5/7

mutant (Chapter 4; Fig. 3), suggesting redundancy of other PLC family members. The combination of

plc3- and plc7-KO mutations was not viable, as we could not obtain homozygous lines from this cross.

Generation of induced silencing lines for PLC3 and PLC7 could help to overcome this problem.

The possible role of PLC in lateral root formation has been discussed in Chapter 2 and 3. Two

options were considered: PLC might affect the root architecture through the generation of IP6 to

influence auxin-TIR1 signaling, or via the metabolism, transport or storage of raffinose

oligosaccharides (RFOs). One function for RFOs is loading sucrose to lateral roots, which could either

facilitate their development or functions to store sugars. In plc3 mutants, phloem sap sucrose had

increased and inositol was slightly reduced in plc3 mutants compared to wild type. However, plc5 only

displayed a slightly lower inositol level (had no change in phloem sucrose). Further investigation of the

phloem sap sugar composition in plc3/5 and multiple plc mutants is required to confirm the

PLC/inositol/RFOs/LR hypothesis. Alternatively, PLC might regulate lateral root formation by PLC-

dependent IP6, which could affect TIR1 activity, resulting in less efficient auxin perception and fewer

lateral root initiation. To confirm this hypothesis, an FP-tagged version of TIR1 in a tir1 mutant

Page 126: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter5

124

background (pTIR1::gTIR-mVENUS) will be crossed with plc3- and plc5 mutants. TIR-mVENUS will

then be immunoprecipitated and IP6 presence will be validated by its conversion into 32P-IP7 by

incubating shortly-boiled fractions with an IP6 kinase (yeast KCS1, fused to GST and expressed in

E.coli) and 32P-γATP. Radioactive labeling is very sensitive, and this might allow us to detect the

decreased IP6 levels in the plc mutant backgrounds that may not be detected by other methods. In this

way, we can link PLC activity directly with IP6 production and auxin signaling through TIR1.

Overexpression of PLC enhances drought tolerance

Drought stress is a major threat that limits plant productivity in agriculture. It does not only cause

hyperosmotic stress, but also oxidative stress on cells (Zhu, 2016). Plant responses to drought stress are

complex, including the sensing and transduction of the stress signal, but also the adaptation to the

adverse environment, which involves numerous physiological-, structural-, morphological-, and

biochemical changes. PLC signaling is activated under drought stress (Munnik and Meijer, 2001;

Munnik and Vermeer, 2010; Hou et al., 2016). Both PLC’s substrates (PIP, PIP2) and products (IPP, PA

converted by DAG) have been reported to be implicated in important cellular events, such as the

reorganization of the cytoskeleton, endo- and exocytosis, vesicular trafficking, and ion-channel

regulation, including intracellular Ca2+ release (Stevenson et al., 2000; Martin, 2001; van Leeuwen et al.,

2007; Heilmann, 2016). Recently, overexpression of a PLC in maize, tobacco and canola have been

shown to improve drought tolerance (Wang et al., 2008; Georges et al., 2009; Tripathy et al., 2011).

Overexpression of PLC3, PLC5 or PLC7 were all found to enhance the drought tolerance of

Arabidopsis, with PLC5-OE showing the strongest phenotype (Table 3). Similarly, overexpression of

PLC2 and PLC4 showed an increased drought-tolerant phenotype (Munnik lab, unpublished),

indicating that overexpression of any PLC might improve the plant's tolerance to drought. When plants

experience drought, several physiological changes happen: e.g stomata close, root systems become

denser and grow deeper, compatible solutes accumulate, photosynthesis decreases, etc. (Comas et al.,

2013; Singh et al., 2015; Basu et al., 2016; Joshi et al., 2016). PLC3-OE and PLC5-OE exhibited a

smaller stomatal aperture than wild type, which could be a reason for their enhanced drought tolerance.

However, two PLC7-OE lines #9 and #12 showed no difference in stomatal aperture, but they did

exhibit the drought tolerance phenotype. It is possible though, that the difference in stomatal opening in

PLC7-OE lines #9 and #12 is too small to be observed. Therefore, more independent PLC7-OE lines

should be used for further confirmation, and preferably, we would like to measure stomatal conductance

in vivo, on a growing plant that experiences drought stress(Andrés et al., 2014). The closure of stomata

is induced by the phytohormone, ABA and this signaling pathway is central to the drought-stress

response. How PLC could be involved in ABA signaling has been well summarized in Chapter 2

(Figure 10).

Page 127: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generaldiscussion

125

Table 3. Phenotypes that are exhibited in PLC-OE lines.

PLC-OE

phenotype

drought

tolerance

stunted root

hairs

smaller stomatal

aperture

ABA sensitivity

stomata

PLC3 yes no yes less

PLC5 yes yes yes less

PLC7 yes no no no change

The root system of PLC-OE lines appears to be neither denser nor larger that would account for

absorbing more water, and thus more drought tolerant. PLC5-OE lines even grow a smaller root system

with very short root hairs, which was not found in other PLC-OE lines. Thus, the increased tolerance of

the PLC-OE lines to drought is probably not through alteration of their root systems. In PLC5-OE lines,

an increase in the accumulation of sugars was detected, which could be another possibility for drought

resistance. Sugar content of the phloem should be carried out with other PLC-OE lines. The drought

response in plants involves many genes with potential cross-talk (Basu et al., 2016; Zhu, 2016) and

PLC could be one of them. Knowing the transcripts of drought-responsive genes in PLC-OE plant will

help to discover the link between PLC and other genes in plant's response to drought stress.

An overview of PLC3, PLC5 and PLC7 phenotypes

All phenotypes found in manipulated PLC lines (KO mutants and OE lines) are summarized in Figure1.

We discovered several novel roles for PLC in plant stress and development, which provides a step

forward in understanding how PLC functions in plants. Here, we emphasize some new scenarios for

plant PLC signaling.

Firstly, apart from second messengers (IPPs and DAG/PA) producer, PLC could also function

as PIP2 and PIP signaling attenuator (Munnik and Nielsen, 2011). PIP2 and PIP are emerging as second

messenger themselves, involved in various stress- and developmental responses (Stevenson et al., 2000;

Meijer and Munnik, 2003; Vermeer et al., 2009; Ischebeck et al., 2010; Munnik and Nielsen, 2011;

Gillaspy, 2013; Rodriguez-Villalon et al., 2015; Heilmann, 2016; Simon et al., 2016).

Secondly, PLC follows the traditional signaling pathway to generate second messengers,

except that IP3 is either phosphorylated to IP6 to release Ca2+, or dephosphorylated to inositol, which is

the precursor for RFO sugar metabolism. The DAG that originates from PLC activity is converted into

PA to activate downstream targets (Munnik and Nielsen, 2011; Testerink and Munnik, 2011; Munnik,

2014).

Thirdly, higher IPPs and PP-IPPs (IP4, IP5, IP6, IP7 and IP8) function as signaling molecules

independently of Ca2+ signaling, such as ion-channel regulation (Zonia et al., 2002), hormone

Page 128: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter5

126

perception (Tan et al., 2007; Sheard et al., 2010; Laha et al., 2015, 2016), mRNA transport (Lee et al.,

2015) and phosphate homeostasis (Kuo et al., 2014; Puga et al., 2014; Wild et al., 2016).

Last but not the least, PI4P exerts great potential to be PLC’s substrate in non-stressed cells.

The authentic PIP2 is missing from most plant plasma membranes, whereas PI4P is much more

abundant (30-100 times), and most (>90%) of the PLC activity is associated with the plasma membrane

fraction. In vitro, PI4P is hydrolyzed equally well as PIP2. The formed IP2 can still be further

phosphorylated to IP6 by the same two IPKs, and the DAG be converted into PA (Munnik et al., 1998a,

1998b; Meijer and Munnik, 2003; van Leeuwen et al., 2007; Vermeer et al., 2009; Vermeer and

Munnik, 2013; Munnik, 2014; Simon et al., 2014; Tejos et al., 2014; Simon et al., 2016). Hence, we

should keep our eyes open for everything!

Figure 1. Summary of phenotypes in plc knockout and PLC overexpression lines. Left: plc knockout mutants exhibit delayed germination; fewer lateral roots; insensitivity to ABA during seed germination and ABA-induced stomatal closure; non-adherent seed coat mucilage and enhanced leaf serration phenotypes. Middle: proposed PLC signaling in plant development and stress responses Right: PLC-OE display increased drought tolerance; less opened stomata and stunted roothair growth phenotypes.

Page 129: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Generaldiscussion

127

REFERENCES

Andrés Z, Pérez-Hormaeche J, Leidi EO, Schlücking K, Steinhorst L, McLachlan DH, Schumacher K, Hetherington AM,

Kudla J, Cubero B, and Pardo JM (2014) Control of vacuolar dynamics and regulation of stomatal aperture by tonoplast potassium uptake. Proc Natl Acad Sci U S A 111:1806–1814.

Basu S, Ramegowda V, Kumar A, and Pereira A (2016) Plant adaptation to drought stress. F1000Research 5:1554. Bentsink L, and Koornneef M (2008) Seed Dormancy and Germination. Arab B 70:1–18. Comas LH, Becker SR, Cruz VM V, Byrne PF, and Dierig D a (2013) Root traits contributing to plant productivity under

drought. Front Plant Sci 4:1–16. De Schepper V, De Swaef T, Bauweraerts I, and Steppe K (2013) Phloem transport: A review of mechanisms and controls. J

Exp Bot 64:4839–4850. Georges F, Das S, Ray H, Bock C, Nokhrina K, Kolla VA, and Keller W (2009) 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 32:1664–1681.

Gillaspy GE (2013) The role of phosphoinositides and inositol phosphates in plant cell signaling, in Lipid-mediated protein signaling (Capelluto DGS ed) pp 141–157.

Heilmann I (2016) Phosphoinositide signaling in plant development. Development 143:2044–2055. Hou Q, Ufer G, and Bartels D (2016) Lipid signalling in plant responses to abiotic stress. Plant, Cell Environ 39:1029–1048. Hunt L, Otterhag L, Lee JC, Lasheen T, Hunt J, Seki M, Shinozaki K, Sommarin M, Gilmour DJ, Pical C, and Gray JE (2004)

Gene-specific expression and calcium activation of Arabidopsis thaliana phospholipase C isoforms. New Phytol 162:643–654.

Ischebeck T, Seiler S, and Heilmann I (2010) At the poles across kingdoms: Phosphoinositides and polar tip growth. Protoplasma 240:13–31.

Joshi R, Wani SH, Singh B, Bohra A, Dar ZA, Lone AA, Pareek A, and Singla-Pareek SL (2016) Transcription Factors and Plants Response to Drought Stress: Current Understanding and Future Directions. Front Plant Sci 7:1–15.

Kuo HF, Chang TY, Chiang SF, Wang W Di, Charng YY, and Chiou TJ (2014) Arabidopsis inositol pentakisphosphate 2-kinase, AtIPK1, is required for growth and modulates phosphate homeostasis at the transcriptional level. Plant J 80:503–515.

Laha D, Johnen P, Azevedo C, Dynowski M, Weiß M, Capolicchio S, Mao H, Iven T, Steenbergen M, Freyer M, Gaugler P, de Campos MKF, Zheng N, Feussner I, Jessen HJ, Van Wees SCM, Saiardi A, and Schaaf G (2015) VIH2 regulates the synthesis of inositol pyrophosphate InsP8 and jasmonate-dependent defenses in Arabidopsis. Plant Cell 27:1082–1097.

Laha D, Parvin N, Dynowski M, Johnen P, Mao H, Bitters ST, Zheng N, and Schaaf G (2016) Inositol polyphosphate binding specificity of the jasmonate receptor complex. Plant Physiol 171:2364–2370.

Lee HS, Lee DH, Cho HK, Kim SH, Auh JH, and Pai HS (2015) 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 27:417–431.

Martin TF. (2001) PI(4,5)P2 regulation of surface membrane traffic. Curr Opin Cell Biol 13:493–499. Meijer HJG, and Munnik T (2003) Phospholipid-based signaling in plants. Annu Rev Plant Biol 54:265–306. Mueller-Roeber B, and Pical C (2002) Inositol phospholipid metabolism in Arabidopsis. Characterized and putative isoforms

of inositol phospholipid kinase and phosphoinositide-specific phospholipase C. Plant Physiol 130:22–46. Munnik T (2014) PI-PLC: Phosphoinositide-phospholipase C in plant signalling, in Phospholipases in Plant Signaling (Wang

X ed) pp 27–54. Munnik T, Irvine RF, and Musgrave A (1998) Phospholipid signalling in plants. Biochim Biophys Acta - Lipids Lipid Metab

1389:222–272. Munnik T, and Meijer HJ. (2001) Osmotic stress activates distinct lipid and MAPK signalling pathways in plants. FEBS Lett

498:172–178. Munnik T, and Nielsen E (2011) Green light for polyphosphoinositide signals in plants. Curr Opin Plant Biol 14:489–497. Munnik T, van Himbergen JAJ, ter Riet B, Braun F-J, Irvine RF, van den Ende H, and Musgrave A (1998) Detailed analysis of

the turnover of polyphosphoinositides and phosphatidic acid upon activation of phospholipases C and D in Chlamydomonas cells treated with non-permeabilizing concentrations of mastoparan. Planta 207:133–145.

Munnik T, and Vermeer JEM (2010) Osmotic stress-induced phosphoinositide and inositol phosphate signalling in plants. Plant Cell Environ 33:655–669.

Nakashima K, and Yamaguchi-Shinozaki K (2013) ABA signaling in stress-response and seed development. Plant Cell Rep 32:959–970.

Nambara E, Okamoto M, Tatematsu K, Yano R, Seo M, and Kamiya Y (2010) Abscisic acid and the control of seed dormancy and germination. Seed Sci Res 20:55–67.

Pattanaik S, Patra B, Singh SK, and Yuan L (2014) An overview of the gene regulatory network controlling trichome development in the model plant, Arabidopsis. Front Plant Sci 5:259.

Puga MI, Mateos I, Charukesi R, Wang Z, Franco-Zorrilla JM, de Lorenzo L, Irigoyen ML, Masiero S, Bustos R, Rodríguez J, Leyva A, Rubio V, Sommer H, and Paz-Ares J (2014) SPX1 is a phosphate-dependent inhibitor of Phosphate Starvation Response 1 in Arabidopsis. Proc Natl Acad Sci U S A 111:14947–14952.

Rodriguez-Villalon A, Gujas B, van Wijk R, Munnik T, and Hardtke CS (2015) Primary root protophloem differentiation requires balanced phosphatidylinositol-4,5-biphosphate levels and systemically affects root branching. Development 142:1437–1446.

Page 130: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Chapter5

128

Schwab B, Folkers U, Ilgenfritz H, and Hülskamp M (2000) Trichome morphogenesis in Arabidopsis. Philos Trans R Soc Lond B Biol Sci 355:879–883.

Sheard LB, Tan X, Mao H, Withers J, Ben-Nissan G, Hinds TR, Kobayashi Y, Hsu F-F, Sharon M, Browse J, He SY, Rizo J, Howe GA, and Zheng N (2010) Jasmonate perception by inositol-phosphate-potentiated COI1-JAZ co-receptor. Nature 468:400–405.

Simon MLA, Platre MP, Assil S, Van Wijk R, Chen WY, Chory J, Dreux M, Munnik T, and Jaillais Y (2014) A multi-colour/multi-affinity marker set to visualize phosphoinositide dynamics in Arabidopsis. Plant J 77:322–337.

Simon ML, MP P, MM M-B, L A, T S, V B, MC C, and Jaillais Y (2016) A PtdIns(4)P-driven electrostatic field controls cell membrane identity and signalling in plants. Nat plants 2:1–10.

Singh M, Kumar J, Singh S, Singh VP, and Prasad SM (2015) Roles of osmoprotectants in improving salinity and drought tolerance in plants: a review. Rev Environ Sci Biotechnol 14:407–426.

Stevenson JM, Perera IY, Heilmann I, Persson S, and Boss WF (2000) Inositol signaling and plant growth. Trends Plant Sci 5:252–258.

Tan X, Calderon-Villalobos LI a, Sharon M, Zheng C, Robinson C V, Estelle M, and Zheng N (2007) Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature 446:640–645.

Tasma IM, Brendel V, Whitham S a., and Bhattacharyya MK (2008) Expression and evolution of the phosphoinositide-specific phospholipase C gene family in Arabidopsis thaliana. Plant Physiol Biochem 46:627–637.

Tejos R, Sauer M, Vanneste S, Palacios-Gomez M, Li H, Heilmann M, van Wijk R, Vermeer JEM, Heilmann I, Munnik T, and Friml J (2014) Bipolar Plasma Membrane Distribution of Phosphoinositides and Their Requirement for Auxin-Mediated Cell Polarity and Patterning in Arabidopsis. Plant Cell 26:2114–2128.

Testerink C, and Munnik T (2011) Molecular, cellular, and physiological responses to phosphatidic acid formation in plants. J Exp Bot 62:2349–2361.

Tripathy MK, Tyagi W, Goswami M, Kaul T, Singla-Pareek SL, Deswal R, Reddy MK, and Sopory SK (2011) Characterization and Functional Validation of Tobacco PLC Delta for Abiotic Stress Tolerance. Plant Mol Biol Report 30:488–497.

Turgeon R, and Wolf S (2009) Phloem transport: cellular pathways and molecular trafficking. Annu Rev Plant Biol 60:207–221.

van Leeuwen W, Vermeer JEM, Gadella TWJ, and Munnik T (2007) Visualization of phosphatidylinositol 4,5-bisphosphate in the plasma membrane of suspension-cultured tobacco BY-2 cells and whole Arabidopsis seedlings. Plant J 52:1014–1026.

Vermeer JEM, Thole JM, Goedhart J, Nielsen E, Munnik T, and Gadella TWJ (2009) Imaging phosphatidylinositol 4-phosphate dynamics in living plant cells. Plant J 57:356–372.

Wang CR, Yang AF, Yue GD, Gao Q, Yin HY, and Zhang JR (2008) Enhanced expression of phospholipase C 1 (ZmPLC1) improves drought tolerance in transgenic maize. Planta 227:1127–1140.

Wild R, Gerasimaite R, Jung J-Y, Truffault V, Pavlovic I, Schmidt A, Saiardi A, Jessen HJ, Poirier Y, Hothorn M, and Mayer A (2016) Control of eukaryotic phosphate homeostasis by inositol polyphosphate sensor domains. Science 352:986–990.

Xu W, Fiume E, Coen O, Pechoux C, Lepiniec L, and Magnani E (2016) Endosperm and Nucellus Develop Antagonistically in Arabidopsis Seeds. Plant Cell 28:doi:10.1105.

Zhu J (2016) Abiotic Stress Signaling and Responses in Plants. Cell 167:313–324. Zonia L, Cordeiro S, Tupy J, and Feijo JA (2002) 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 14:2233–2249.

Page 131: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Summary

129

Summary

For many years, efforts have been made to explore PLC signaling in plants. Compared to the classical

eukaryotic, mammalian PLC signaling pathway, a different picture is emerging for plants. Several roles

for PLC in plant development and stress responses have been claimed but genetic evidence for this is

mostly missing.

In chapter 2, we functionally characterized the role of PLC3 and found it mainly expressed in

vascular tissue, but was also present in guard cells and trichomes. Loss-of-function mutants showed

different phenotypes, including delayed seed germination, slightly less-elongated primary roots (~5%),

fewer lateral roots (15%), and decreased sensitivity to ABA in terms seed germination inhibiting and

stomatal-closure induction. ABA triggered an increase in PIP2 in various Arabidopsis tissues, however,

no difference in PLC substrate- or product levels in two independent plc3 KD mutants were found.

Overexpression of PLC3 increased the drought tolerance of plants and decreased their stomatal aperture.

In chapter 3, similar analyses were processed for PLC5, which was also predominantly

expressed in the vasculature, but also in the root apical meristem, guard cells and trichomes. Knock-

down mutant, plc5-1 displayed shorter primary roots (~10%) and exhibited fewer lateral roots (~20%),

which could be restored by the expression of the endogenous PLC5-wt gene, driven by its own

promoter. Double-plc3plc5 mutant did not intensify the root phenotype, indicating the involvement of

(a)other PLC(s). Overexpression of PLC5 enhanced drought tolerance and reduced stomatal aperture,

like PLC3, but in this case also led to a new phenotype, i.e. a stunted root-hair growth. 32P-

phospholipid labeling analyses revealed no differences between wt and plc5-1 seedlings, however,

decreased levels of PIP and PIP were found with increased levels of PA in PLC5 over-expressor lines.

Inducible overexpression of PIP5K3 in PLC5-OE line rescued the root hair phenotype and restored the

level of PIP2, providing independent evidence for PIP2's crucial role in tip growth.

In chapter4, we characterized the role of PLC7, another phloem-expressed PLC. Expression

was throughout the plant vasculature, including roots, leaves and flowers, and also appeared in

trichomes and hydatodes. We obtained a plc7-3 KO- and plc7-4 KD mutant but found no affects in root

development, which is different from the plc3- and plc5 mutants. However, like plc3 mutants, plc7

mutants exhibited a reduced sensitivity to ABA-dependent stomatal closure. Double-knockout mutants

of plc3plc7 were lethal, whereas plc5plc7 (plc5/7) mutants were viable, and revelead several new

phenotypes not observed earlier. These include a defect in seed coat mucilage, enhanced leaf serration,

and an increased tolerance to drought. The latter phenotype was previously found when PLCs were

overexpressed. Overexpression of PLC7 also led to an enhancd drought tolerant phenotype. In vivo 32Pi-labeling of seedlings treated with sorbitol to mimick drought stress revealed increased PIP2

responses in both drought tolerant plc5/7 and PLC7-OE mutants. Together, these results reveal several

novel functions for PLC in plant stress and development.

Page 132: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

130

Page 133: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Samenvatting

131

Samenvatting

De rol van PLC in signaaltransductieprocessen in planten is nog grotendeels onbekend. Het is in ieder

geval duidelijk dat dit anders moet zijn dan de klassieke, dierlijke, PLC-signaleringsroute, voornamelijk

omdat planten de targets missen voor zowel IP3 als DAG. Dit roept veel vragen op zoals, dienen er

soms andere moleculen als second messengers, en hoe werkt dat dan? Is PIP2 het enige substraat voor

PLC of wordt standaard het meer abundante PIP gebruikt? En is de functie van PLC om PPI signalering

te dempen of om second messengers te produceren? Meer bewijs is verzameld voor rollen van PLC in

de ontwikkeling van de plant en stressreacties, langzaam waardoor meer en meer stukken van deze

puzzel op tafel.

In hoofdstuk 2 wordt de rol van PLC3 functioneel gekarakteriseerd. Het gen komt vooral tot

expressie in vaatweefsel, maar is ook aanwezig in sluitcellen en trichomen. Mutanten met een defect

PLC3 gen hebben een vertraagde zaadkieming, minder lange hoofdwortel (~ 5%), minder zijwortels

(15%), en een verminderde gevoeligheid voor ABA tijdens zaadkieming en bij het sluiten van

huidmondjes. ABA stimuleert de toename in PIP2 hoeveelheden in verschillende weefsels, maar er

wordt geen verschil gevonden in PLC's substraat- of productniveaus van plc3 mutanten. Overexpressie

van PLC3 in Arabidopisis planten blijkt de droogtetolerantie te verhogen en de opening van

huidmondjes te verlagen.

In hoofdstuk 3 werden soortgelijke analyses uitgevoerd voor PLC5, welke ook overwegend

in het vaatstelsel tot expressie komt, maar ook werd gevonden in het meristeem van de worteltop, in

sluitcellen en in trichomen. Een knock-down, plc5-1 mutant heeft kortere hoofdwortels (~ 10%) en

minder zijwortels (~ 20%), welke weer kan worden hersteld door expressie van het wild-type PLC5 gen,

aangedreven door de eigen promotor. Dubbel plc3plc5-mutanten gaven geen sterker fenotype,

mogelijke doordat andere redundante PLCs de rol overnemen. Overexpressie van PLC5 verhoogd de

droogtetolerantie en verminderd de opening van huidmondjes. Daarnaast werd nog een nieuw fenotype

gevonden, namelijk een remming van de wortelhaargroei. 32Pi-fosfolipidenlabelinganalyses lieten geen

vershil zien tussen wt and plc5-1 zaailingen, echter een aanzienlijk verminderde PIP- en PIP2 niveau en

verhoogde PA niveau werden in PLC5 over-expressie mutanten gevonden. Induceerbare overexpressie

van PIP5K3 in deze PLC5-overexpressielijn, kon de wortelhaargroei en de PIP2 niveaus herstellen,

welke onafhankelijk bewijs levert voor de cruciale rol die PIP2 speelt in tip groei.

In Hoofdstuk 4 werd de rol van PLC7, weer een andere floëem-specifieke PLC gen

gekaraketriseerd. Het gen komt in het gehele vaatstelsel tot expressie, inclusief wortels, bladeren, en

bloemen, maar ook in trichomen en in de waterporiën van de bladeren (hydatoden). Mutanten, i.e.

plc7-3 KO en plc7-4 KD, bleken geen defect in de wortelontwikkeling te hebben, anders dan in de

plc3- en plc5 mutanten. Echter, net als plc3 mutanten, bleken plc7 mutanten ook een verlaagde

Page 134: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Samenvatting

132

gevoeligheid voor ABA te hebben voor het sluiten van huidmondjes. Dubbel-KO mutanten van

plc3plc7 bleken lethaal te zijn, maar plc5plc7 (plc5/7) mutanten waren wel levensvatbaar. Deze laatste

lieten echter verschillende nieuwe fenotypes zien die nog nooit niet eerder waren waargenomen, te

weten: een defect in het slijm van de zaadhuid, een toename in de gezaagdheid van de bladeren, en een

verhoogde tolerantie voor droogtestress. Dit laatste fenotype werd eerder gevonden waneer juist PLCs

to overexpressie werden gebracht. En ook overexpressie van PLC7 laat wederom zien de

droogtetolerantie te verbeteren. In vivo 32Pi-labeling van zaailingen die behandeld werden met sorbitol

om droogtestress na te bootsen, blijken sterkere PIP2 responsen te vertonen in zowel de droogte

tolerante plc5/7- als de PLC7-OE mutanten. Samen onthullen deze resultaten een aantal nieuwe functies

voor planten PLCs in stressreacties en ontwikkeling.

Page 135: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Acknowledgments

133

Acknowledgments

The journey of my PhD is coming to the end. Looking back, there have been excitement but also

frustration. Treasures, in the end. There are people who play important role along this road and I want

to say: “thank you!”

Teun and Michel, as my daily supervisor and promoter, you are definitely big roles to make

my PhD dream come true. Teun, I always appreciate the opportunity you offered to me for joining your

lab and your supportive attitude for scientific challenges I encountered. The way you see the beauty of

science inspires me. The way you train people makes me grow from a naive student to a young scientist.

You devoted yourself to thesis writing correction and made it very readable, which deserves a nobel

literature prize. Michel, your constructive ideas are always valuable and your guide through my PhD

path directly accelerates the whole process. Your encouragement makes me feel confident. I appreciate

these all. Also Christa, Rob and Petra, thanks for all the constructive suggestions and you together

create a wonderful scientific environment here.

Ringo, most of my lab techniques are learned from you. And you always try the best to save me

from lab disaster. Well, not only me, you help everyone, whenever we have a guest, its your show time.

As to our never-stop coming guests, I did share special memories with them, Wojciech, Kelly, Necla

and of course Michael -- our VIP, thank you very much for providing me information of labs doing

applied research. Our lab is expending with the joining of Max, a brilliant boy who continues PLC

work and I believe you will make the project shining. And Pamela, a very very very nice girl, it was

fun to have you around in my last bit of PhD. I’ve been sitting in the office for more than four years,

seeing people left and come. All office mates, thanks for the interesting office chatting and sharing - -

jokes, food... It's that you make this office vividly! All lab mates from UVA plants groups, thanks for

your encouragement words when I showed my experiment-failed face and your generous attitude when

I messed the lab. I find half of the greenhouse staffs (Ludek, Harold, Thijs and Fernando) are artists!

Would it be my destination? Seriously, thanks for taking care of my plants. Lunchtime is one of the

most beautiful moments during a working day. Fariza, thanks for your company for these years,

although you left me, the precious memory will be forever and hope you get your PhD sooner. The

lunch group was then rebuilt by a bunch of Chinese ladies, Lingxue, Pulv, Xiaotang, Yanxia, Yanting

and Yutao, thanks you for sharing food and gossips.

Page 136: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Acknowledgments

134

I won’t survive in the lab without you Xavi, my Catalan brother with wild beard but gentle eyes

who is suspected by dutch police as drug dealer or terrorist on street sometimes. No matter what do you

look like, your care for people, your romantic attitude to life make you a charming man. No wonder,

Carlos is so fond of you and this Chilean moustache dude, as smart as fox, funny as Charlie Chaplin,

brought lots of joy to my PhD life. And Marc, a gentleman-looking French dude, thoughtful, French

humour always, sitting on my left hand, typing with music, having visitors all the time, because you

know statistic analysis. Having you guys around is my huge luck. Your friendship does value a lot for

me and I definitely will miss the old crazy time!

As a Chinese, traveling far away to Netherlands, homesick is non-avoiding. Here, special

thanks are given to two JJ ladies. Jie, shy but funny, has endless interesting stories to tell when you’re

in good mood. TV show is not amusing anymore when you are around. Jiesen, playing lots of roles,

lunch mate, office mate, and lab mate but most important a friend, is a very rare species who never

gossips as a girl. Thanks for both of your company when I felt alone, thanks for your listening when I

complain, thanks for your help when I was in trouble. Thanks for your true friendship. I would also like

to thank another two wonderful ladies Yang and Xue. Although we met occasionally, I really enjoyed

the time we had. Your words always inspire me and open my mind. I hope our friendships will extend

out of Netherlands.

在此,我将最真挚的感谢献于辛勤养育我的父亲和母亲。感谢你们给予我无私的爱以及

无微不至的关怀。感谢你们一直以来的激励与支持。父母之恩,水不可溺,火不可灭。

我的挚爱彦邦,遇君,幸也。于我,你既是伴侣,亦是良师益友。生活上,你呵护备至;

工作上,你排忧解难。感谢, 一路上有你。

Page 137: UvA-DARE (Digital Academic Repository) Novel roles …...2.3 Gene expression and subcellular localization 3. PLC in plant stress response and development 3.1 PLC in plant stress responses

Listofpublications

135

List of publications

Zhang Q, van Wijk R, Shahbaz M, Roels W, van Schooten B, E.M. Vermeer J, Guardia A, Scuffi D,

García-Mata C, Laha D, Williams P, A.J. Willems L, Ligterink W, Hoffmann-Benning S, Gillaspy G,

Schaaf G, Haring M.A, Laxalt A.M., & Munnik T

A role for Arabidopsis phospholipase C3 (PLC3) in seed germination, lateral root formation and

stomatal closure. (In preparation)

Zhang Q, van Wijk R, Shahbaz M, E.M. Vermeer J, Guardia A, Scuffi D, García-Mata C, Van den

Ende W, Hoffmann-Benning S, Haring M.A, Laxalt A.M., & Munnik T

Functional characterization of PLC5 in Arabidopsis thaliana - knock-down affects lateral root initiation

while overexpression stunts root hair growth and enhances drought tolerance. (In preparation)

Zhang Q, van Wijk R, Lamers M, Maarquez F.R, Guardia A, Scuffi D, García-Mata C, Ligterink W,

Haring M.A, Laxalt A.M., & Munnik T

Role for Arabidopsis PLC7 in stomatal closure, mucilage adherence, and leaf serration. (In preparation)