中本, 雅俊 - osaka prefecture university...

84
http://repository.osakafu-u.ac.jp/dspace/ Title Studies on essential roles of C24-ethylsterol in plant development Author(s) 中本, 雅俊 Editor(s) Citation Issue Date 2016 URL http://hdl.handle.net/10466/15130 Rights

Upload: dokiet

Post on 19-May-2019

220 views

Category:

Documents


0 download

TRANSCRIPT

http://repository.osakafu-u.ac.jp/dspace/

   

Title Studies on essential roles of C24-ethylsterol in plant development

Author(s) 中本, 雅俊

Editor(s)

Citation

Issue Date 2016

URL http://hdl.handle.net/10466/15130

Rights

Studies on essential roles of C24-ethylsterol in plant development

(植物ステロールが担う未知生理機能の解明)

Masatoshi Nakamoto

2016

1

Contents

Introduction ………. 2

Chapter1 ………. 11

Establishment and characterization of an Arabidopsis C24-ethylsterol deficient

mutant line

Chapter2 ………. 43

Clarification of the disturbed polar localization of auxin efflux carrier protein

Chapter3 ………. 53

General discussion

Supplementary Information ………. 60

Materials and Methods

References ………. 72

Acknowledgements ………. 82

2

Introduction

Sterols are isoprenoid-derived lipids that have diverse and essential functions in all

eukaryotes. Sterols are integral components of the membrane lipid bilayer, where they

regulate membrane permeability and fluidity in conjunction with phospholipids. In

animals, cholesterol alone serves this structural function, whereas plant membranes

consist of several sterols. Plants and algae present distinct sterol biosynthetic features

compared to metazoan as they produce multiple pathway end-products. In addition to

relatively low amounts of cholesterol, their sterol profiles display several

24-alkyl-5-sterols as major products. This chemical diversity of the sterol-reinforcing

component of higher plant cellular membranes is considered the result of an

evolutionary process leading to optimal membrane stability in sessile organisms (Nes

and McKean, 1977; Ramgopal and Bloch, 1983; Dufourc, 2008). The question of that

chemical diversity has driven research on plant sterol biology. Recently, the power of

Arabidopsis thaliana genetics has clearly shown that breakdown of sterol homeostasis

led to severe damage throughout the plant life cycle, embryogenesis, cell elongation,

cell wall formation, vascular differentiation, and hormone signaling (Diener et al., 2000;

Carland et al., 2002; Schrick et al., 2004; Souter et al., 2004; Men et al., 2008).

Biosynthetic defects results in a lack of major membrane sterols that accompany the

accumulation of unusual sterols, a condition reminiscent of sterolosis syndromes well

described in humans (Kelley and Herman, 2001). Plant sterol profile disruption

furthermore leads to a dysfunctional brassinosteroid (BR) biosynthesis that in turn

causes dwarfisms (Schaller, 2004).

It is widely accepted that levels of 24-alkyl-5-sterol (phytosterols) must be

3

appropriately controlled for normal plant development (Schaeffer et al., 2001). This is

determined primarily by the ratio of C24-methylsterols (campesterol) to C24-ethylsterols

(-sitosterol, stigmasterol). This ratio is tuned by the sterol side chain C24-alkylation

reactions that add two exocyclic carbon atoms provided by S-adenosyl methionine to

sterol precursors and are catalyzed by two distinct sterol methyltransferases (SMTs)

(Fig. 1). The first-type SMT (SMT1) is conserved across kingdoms and is responsible for

the C24-methylation of cycloartenol (or for example, of lanosterol in fungi), and the

second-type SMT (SMT2) is specific for Charophycean green algae and plants, and is

responsible for the C28-methylation (or C241-methylation) of 24-methylene lophenol

(Fig. 2). Thus, SMT2 differentiate C24-ethylsterols from C24-methylsterols by adding

one extra methyl group to the side-chain.

Mutations in genes that act upstream of 24-methylene lophenol, the substrate of

SMT2 and SMT3 (Fig. 1), completely impair the production of major phytosterols

compatible with survival (Clouse, 2002). Such Arabidopsis mutants include the

cph/smt1/orc mutants defective in SMT1 (the first methyltransferase step producing

24-methylene-cycloartanol from cycloartenol) (Fig. 1), that exhibit incomplete cell wall

structures, abnormal embryonic development, low fertility rates, irregular vascular

system development, and aberrant membrane localization of the PIN1/PIN3 auxin efflux

transporter proteins (Diener et al., 2000; Lindsey et al., 2003; Willemsen et al., 2003;

Schrick et al., 2004). The Arabidopsis cpi1-1 mutant is defective in the cyclopropylsterol

isomerase gene, resulting in lack of major sterols and accumulation of atypical sterols.

cpi1-1 plants display disturbed root development, unsuccessful endocytosis, loss of

PIN2 polar localization and abnormal cytokinesis (Men et al., 2008). Other Arabidopsis

mutants such as fackel/hydra2 harboring mutations in the C14-reductase (Jang et al.,

4

2000; Schrick et al., 2002), and hydra1 (Souter et al., 2002), which defects in the

8,7-isomerase, exhibit defects in root development, cell wall structure, embryonic and

seedling cell patterning, and ethylene signaling (Souter et al., 2004). In general, these

sterol mutants are thought to be defective in the generation of the auxin concentration

gradients, which are primarily established through the cell-to-cell directional transport by

auxin carrier proteins (Dhonukshe et al., 2008; Petrásek and Friml, 2009; Feraru and

Friml, 2008). The sterol involvement in cell division has also been suggested from the

disrupted cytokinesis in smt1 (Schrick et al., 2002; Schrick et al., 2004) and cpi1-1 (Men

et al., 2008) mutants. The defective vascular patterning in the shoots of hydra1 and

fackel/hydra2 mutant seedlings is associated with ectopic cell divisions (Pullen et al.,

2010), whereas the cell division during early stages of lateral root initiation in the hydra1

and fackel/hydra2 mutants is not affected (Souter et al., 2004). In addition to these

mutations in the post-squalene biosynthetic segment leading to 24-methylene lophenol,

a 4-methyl sterol intermediate substrate of land plants SMTs (Fig. 1), the loss of function

of ERG28 in Arabidopsis induces the accumulation of 4-carboxy-4-methyl-24-methylene

cycloartanol (CMMC), a sterol biosynthetic intermediate yielded from

24-methylene-cycloartanol during the demethylation reaction at C4 (Fig. 1). The

accumulated 4-carboxy-4-methyl sterol derivative, otherwise undetectable in plant

tissues when ERG28 tethers the C4 demethylation enzymatic complex, interferes with

polar auxin transport via its binding to ABC transporters and this result in serious growth

disorder (Mialoundama et al., 2013). Taken together, these results indicate that

phytosterols play critical roles in a variety of molecular events involved in plant

development throughout the life cycle. However, little is known about the mechanism by

which the sterol compositions affect the polar recycling of PIN proteins, cell division, and

5

development. Most importantly, critical roles of individual sterols could not be clarified

from the characterization of a series of Arabidopsis biosynthetic mutants described

above.

Conversely, mutations of the multi-cellular plant specific sterol methyltransferase

genes (SMT2 and SMT3 in Arabidopsis, Fig. 1) affect the C24-methylsterol to

C24-ethylsterols ratio, and so modify specifically the pathway end-product profile. An

Arabidopsis SMT2 mutant (cvp1) displayed post-embryonic defects in vascular

patterning (Carland et al., 2002). In addition, SMT2 co-suppression lines harbored

higher levels of 24-methylcholesterol (campesterol and its epimer

22(23)-dihydrobrassicasterol) and lower levels of -sitosterol and exhibit reduced apical

dominance and fertility in brassinosteroid-independent manner (Schaeffer et al., 2001).

A double mutant line (cvp1 smt3) generated by crossing a cvp1-3 (a null mutation of

SMT2) mutant and a T-DNA mutant of SMT3 (SALK_085292) displayed severe growth

abnormalities affecting processes such as floral organ development and the auxin

response (Carland et al., 2010). The sterol profile of the cvp1 smt3 mutant almost

completely lacks the C24-ethylsterols instead of increased levels of C24-methylsterols,

without influencing the total sterol content and BR biosynthesis (Carland et al., 2010). In

Arabidopsis, the role of (24R)-24-methylcholesterol (campesterol) as a precursor of

brassinosteroids was demonstrated (Fujioka and Yokota, 2003). However, essential

roles of C24-ethylsterols beyond membrane reinforcement have never been clarified in

the multi-cellular plants expressing the SMT2 gene family or in other organisms that

produce C24-ethylsterols, albeit with a different 24-ethyl-orientation (Fig. 2).

While Qian et al. (2013) reported that exogenous application of stigmasterol

(C24-ethylsterol) and campesterol (C24-methylsterol) partially rescued the stomatal

6

development and patterning of the Arabidopsis fk-J3158 mutants, (fk-J3158 is a weak

allele of the sterol C-14 reductase gene FACKEL), exogenously added sterols have

been usually ineffective in the rescue of mutant phenotypes (Diener et al., 2000;

Willemsen et al., 2003; Schrick et al., 2002; Carland et al., 2002; Carland et al., 2010),

and thus it remains unknown whether C24-ethylsterols play a crucial role in plant

development. Aiming to clarify whether C24-ethylsterols may be play essential roles or

be involved in plant development, we studied the strikingly abnormal phenotype of an

Arabidopsis line (smt2/smt2-smt3/smt3) defective in both the SMT2 and SMT3 genes,

which encode the sterol methyltransferases in charge of the second methyltransfer

reaction at the C24 position of the side chain (Fig. 1). Here, we show that -sitosterol (a

C24-ethylsterol) is necessary and sufficient not only for auxin responses but also for the

cell division plane determination.

7

Figure 1. The sterol biosynthetic pathway in Arabidopsis

The biosynthetic pathway involves two methyltransfer reactions at C24 position. The first reaction is catalyzed by SMT1, leading to the production of C24-methylsterols, and the second reaction is catalyzed by SMT2 and SMT3, yielding C24-ethylsterols.

8

9

1; Maize2, 2; Maize1, 3; Volvox, 4; Trypanosoma, 5; Thalassiosira, 6; Selaginella1, 7; Selaginella2, 8; Populus2, 9; Populus1, 10; Penium2, 11; Penium1, 12; Physco2, 13; Physco1, 14;Ostreococcus1, 15; Rice1, 16; Rice2, 17; Monosiga, 18; Micromonas1, 19; Klebs2, 20; Klebs1, 21; Cotton2, 22; Cotton1, 23; ERG6, 24; Ectocarpus, 25; Dictyostelium, 26; Cyanidioschyzon, 27; Coccomyxa, 28; Chlorokybus2; Chlorokybus1, 30; Chlamydomonas, 31; Chaetospaeridium2, 32; Chaetospaeridium1, 33; Candida, 34; BotryococcusTMT1, 35; BotryococcusSMT2, 36; AtSMT3, 37;AtSMT2, 38; AtSMT1, 39; Aspergillus, 40; Acanthamoeba.

10

Figure 2. Phylogenetic tree for sterol methyltransferase proteins.

Green dots (algae) and blue dots (terrestrial plants) indicate species containing both SMT2 and SMT3 genes.

The Klebsromidium sequences were retrieved at the Klebsormidium genome project (http://www.plantmorphogenesis.bio.titech.ac.jp/~algae_genome_project/klebsormidium/ index.html). The methyltransferases in land plants are indicated with blue dots, and those in charophycean algae are shown with green dots. The second-type sterol methyltransferases (SMT2) are shown with in green dotted line. Nonredundant database using the TBLASTN search in the National Center for Biotechnology. Sterolmethyltransferase sequences were identified in the GenBank Information with the Arabidopsis SMT2 At1g20330 protein sequence as the query. Multiple alignments of sterolmethyltransferase protein sequences were aligned using ClustalW with default parameter values and analyzed using MEGA6 software version 6.0.6 (Tamura et al., 2013). The sterolmethyltransferase sequences used for the analysis were Acanthamoeba; Acanthamoeba castellanii XM_004336492.1, Aspergillus; Aspergillus nidulans XP_001211350, BotryococcusTMT1; Botryococcus braunii AEY68256.1, BotryococcusSMT1; B. braunii AEY68259.1, Candida; Candida tropicalis XP_002548099.1, Chaetospaeridium1; Chaetospaeridium globosum HO399328.1, Chaetospaeridium2; C.

globosum HO404922.1, Chlamydomonas; Chlamydomonas reinhardtii XP_001690775.1, Chlorokybus1; Chlorokybus atmophyticus HO414429, Chlorokybus2; C.atmophyticus HO408954.1, Coccomyxa; Coccomyxa subellipsoidea XP_005649905.1, Maize1; Zea mays P_002441576.1, Maize2; Zea mays NP_001131331.2, Cotton1; Gossypium hirsutum AAZ83345.1, Cotton2; Gossypium hirsutum ABY49844.1, Cyanidioschyzon; Cyanidioschyzon merolae XP_005537914.1, Dictyostelium; Dictyostelium discoideum XP_636481.1, Ectocarpus; Ectocarpus siliculosus FP261080.1, ERG6; Saccharomyces

cerevisiae erg6 NP_013706.1, Klebs2; Klebsormidium flaccidum kfl00449, Klebs1; K.

flaccidum kfl00283, Monosiga; Monosiga brevicollis XP_001748534.1, Micromonas1; Micromonas sp RCC299 XM_002499342.1, Ostreocossus1; Ostreococcus lucimarinus CCE9901 p XM_001416678.1, Penium1; Penium margaritaceum HO583831.1, Penium2; P.

margaritaceum HO608114.1, Populus1; Populus euphratica XM_011036654, Populus2; Populus euphratica XM_011027171, Physco2; Physcomitrella patens, XM_001772181.1, Physco1; P. patens XM_001785643.1, Rice1; Oryza sativa Os07g0206700, Rice2; O. sativa Os03g0136200, SMT1; Arabidopsis thaliana At5g13710, SMT2; A. thaliana At1g20330, SMT3; A. thaliana At1g76090, Selaginella1; Selaginella moellendorffii XM_002968054.1, Selaginella2; S. moellendorffii, XM_002975434.1, Thalassiosira; Thalassiosira pseudonana CCMP1335b XP_002292358.1, Trypanosoma; Trypanosoma cruzi XP_802864, Volvox; Volvox carteri XP_002948023.1.

11

Chapter1

Establishment and characterization of an Arabidopsis C24-ethylsterol

deficient mutant line

Sterol composition of smt2 smt3 plants

We crossed two T-DNA insertion lines (GABI_433_F03 for smt2 and SALK_085292

for smt3) with each other and maintained a +/smt2-smt3/smt3 line for self-pollination to

obtain homozygous smt2/smt2-smt3/smt3 (smt2 smt3, hereafter) plants, which are

self-sterile (described below). The smt2 smt3 plants lack both the SMT2 and SMT3

transcripts (Fig. 3) and exhibit a sterol profile highly distorted towards C24-methylsterol

accumulation accompanied by an almost complete loss of C24-ethylsterols (Table 1). In

smt2 and smt2 smt3, the faint C24-ethylsterol levels could be ascribed to the weak

substrate reactivity of SMT1 toward 24-methylenelophenol (Diener et al., 2000).

Drastically reduced amount of sitosterol (that accounted for 80% of the total sterol in the

wild-type and less than 2% in smt2 smt3) led also to a complete disappearance of

sitosterol glucoside (Fig. 4), a compound commonly found in the pool of sterol

conjugates in Arabidopsis (DeBolt et al., 2009). Moreover, plant sterol profiles usually

display minor amounts of 4,4-dimethyl- and 4-methyl sterol precursors in proportions

that may vary according to species (Nes and McKean, 1977; Benveniste, 2004; Schaller,

2010). We measured in the mutants considered here a relative increase in the content

of those precursors (cycloartenol, cycloeucalenol, 4-methylfecosterol, 24-methylene

lophenol, Table 1) that represented 0.3% of the total in the wild-type and 1.3% and 3%

in smt2 and smt2 smt3 mutants, respectively. We did not detect any accumulation of

12

unusual sterols reported in plants mutant affected in genes upstream of SMT2 step,

such as the cpi1-1 (Men et al., 2008). The slight increase in 4,4-methyl- and 4-methyl-

sterols prompted a search for 4-carboxy-4-methyl sterol biosynthetic intermediates

(Mialoundama et al., 2013) that might disturb polar auxin transport in smt2 smt3. Those

derivatives were not detected in either the WT or smt2 smt3 mutant seedlings (Fig. 5).

13

Figure 3. The sterolmethyltransferase mutant used in this study

(a) A schematic diagram of the T-DNA insertion events in the SMT2 gene in the smt2

line (GABI_433_F03) and in the SMT3 gene in the smt3 line (SALK_085292).

(b) SMT2 and SMT3 gene transcription levels in the smt2, smt3, and smt2 smt3 lines

(RT-PCR).

14

a smt2, GABI_443_F03

b smt3, SALK_085292

c in percent ; one representative analysis is shown per genotype

d, three,

e,f two and

g four independent measurements.

Dry materials were saponified in methanol containing 6% (w/v) KOH for 2 hours at 80°C. The hexane extract was dried out then acetylated in toluene/acetic anhydride/pyridine for 1 hour at 70°C. The dried residue was separated by TLC with one run of CH2Cl2. Steryl acetates were

scrapped off then analyzed by GC-FID and GC-MS.

Table 1. Sterol composition of rosette leaves from Arabidopsis thaliana smt mutants

15

Figure 4. Analysis of -sitosterol and sitosteryl glucoside by UPLC-MS in 5-day old

WT and smt2 smt3 seedlings Sterol detection was done by Selected Ion Recording (SIR) MS for -sitosterol (m/z = 397.71) and sitosteryl glucoside (M+H-glucose-H2O; m/z = 397.70) based on published work (Mialoudama et al., 2013). (a) Authentic sitosterol and sitosteryl glucoside were separated by UPLC (sitosterol, t=16.27 min; sitosteryl glucoside, t=12.29min). (b) Close-up on the 12 min area of the chromatograms shown in a (relative intensity to compound elution time) to demonstrate very low levels or detection of sitosteryl glucoside in smt2 smt3 mutant seedlings. Analysis was done in biological triplicate. Representative chromatograms are shown (relative intensity to compound elution time).

16

Figure 5. Analysis of 4-carboxy sterols derivatives by UPLC-MS in 5-day-old WT and smt2

smt3 seedlings

A search of the 4-carboxy sterol biosynthetic intermediates was done by Selected Ion Recording (SIR) MS for 4-carboxymethyl cycloartenol (CMC; m/z = 439.50) and 4-carboxymethyl-24-methylene cycloartanol (CMMC; m/z = 453.48) based on published work (Mialoudama et al., 2013). (a) No detection of CMC. (b) No detection of CMMC. Analysis was done in biological triplicate. Representative chromatograms are shown (relative intensity to compound elution time).

17

Shoot phenotypes of smt2 smt3 plants

The smt2 smt3 plants exhibit phenotypes similar to those reported for the cvp1-3

smt3 mutant (Carland et al., 2010), including severe growth retardation, abnormal

flowers, aberrant vascular development, and abnormal leaf development (Fig. 6a, 6b,

6c, 6d). In addition to those phenotypes, we observed that the smt2 smt3 mutant was

self-sterile (Fig. 6e), which could be ascribed to the dysfunctional development of the

female gametogenesis (Fig. 6f) and transmitting tract in pistils (Fig. 6g). The transmitting

tract is essential for the growth of pollen tubes that deliver sperms to the egg cells. The

smt2 smt3 pollens were viable and could cross-pollinate wild type (WT) flowers. A

detailed phenotypic characterization of SMT2 cosuppressed lines (smt2COSUP) has been

reported (Schaeffer et al., 2001). The developmental progression of integuments in

smt2COSUP ovules is impaired whereas the funiculus has the same pattern in both

genotypes. The formation of integuments that protect the embryo sac is most probably

affected by a dramatic reduction of SMT2 expression and subsequently -sitosterol

production (Schaeffer et al., 2001), at the early integument formation. Eventually, a few

mature seeds are formed in the smt2COSUP siliques suggesting that the embryo sac is

not affected by the suppressed SMT2 (Schaeffer et al., 2001). The overall

measurements of seed yield indicate a 15-fold reduction in smt2COSUP compared to WT.

The cellular phenotype of smt2COSUP ovules might suggest an action of SMT2 upstream

or during the initiation stage of integuments development (Gasser et al., 1998).

Embryonic phenotypes of smt2 smt3 plants

The embryogenesis in Arabidopsis, the cells of embryo acquires specific fates in

integrated manner to form plant body. The embryo undergoes characteristic shape

18

changes brought about by different orientations of cell divisions and directional cell

expansions. The smt2 smt3 mutant display abnormal embryonic phenotypes during

early stage of embryonic development (Fig. 7). During 8-cell stage and 16-cell stage,

the shrunken suspensor and abnormal cell morphologies at the base of embryo were

observed in smt2 smt3 (Fig. 7a), suggesting aberrant cell division in early embryos.

smt2 smt3 embryos showed irregular divisions in cells at the center of globular and

heart stage embryos (Fig. 7a). Moreover the DR5:GUS expression in smt2 smt3

embryos was strikingly intense not only in the columella cell but also the whole embryo

(Fig. 7b), suggesting that the directional auxin transport was defective. Mutations in

genes that act upstream of 24-methylene lophenol (smt1/cph, fk and hyd1) which lack

total sterols exhibit embryonic defects (Schrick et al., 2002; Clouse, 2002), but those

downstream do not (cvp1, dwf1, dwf5, dwf7) (Clouse, 2002). These results indicate that

C24-ethylsterols are required for embryo development.

Root phenotypes of smt2 smt3 plants

The smt2 smt3 displays extensive abnormalities in the roots (Fig. 8), whereas

SMT2 and SMT3 genes are specifically expressed in the root elongation zone (Carland

et al., 2002). A representative smt2 smt3 seedling phenotype is characterized by a

twisted short root (Fig. 8a, 8b). The root development is completely disrupted,

demonstrating abnormal tissue architecture, occasional swollen cells, and severe

growth inhibition (Fig. 8c). The abnormal tissue architecture in smt2 smt3 roots is

visualized by the staining of membranes (FM4-64; Fig. 9a), and cell walls (propidium

iodide, PI; Fig. 9b). Abnormal cell division was prominent in the smt2 smt3 root stem

cells. FM4-64 staining of smt2 smt3 revealed disruption of the formative division

19

followed by cell lineages with deformed shapes and irregular orientations as well as an

abnormally formed cell plate. In asymmetric divisions, the division planes are governed

by the polarity of the mother cell (Rasmussen et al., 2010). Thus, abnormal cytokinesis

in the smt2 smt3 root stem cell is likely responsible for the defects in subsequent cell

division that lead to the collapsed architecture of the subsequent cell lineages (Fig. 9a).

Furthermore, smt2 smt3 roots harbor a number of PI-positive cells, including root

meristem, vascular cells, and in epidermis and cortex cells (Fig. 9b). PI does not

penetrate into living cells across the plasma membrane (PM) (Curtis et al., 2007). Thus,

the PI-positive cells in smt2 smt3 roots demonstrate not only collapsed maintenance of

root meristem but also somatic cell death. In smt2 smt3 roots, the loss of gravitropism

and the disordered orientation of starch granules suggested that the auxin-related

cellular functions were impaired (Fig. 10a, 10b). DR5:GUS expression in the smt2 smt3

was strikingly intense in the root tip and throughout the root vascular system (Fig. 10c),

suggesting that the defect was in the directional auxin transport and/or due to excess

auxin accumulation. In WT plants, the PIN2 auxin efflux transporter localized correctly,

maintaining the downward and upward directions in the epidermis and the cortex cells,

respectively (Carland et al., 2010; Blilou et al., 2005). Multiple mislocalization patterns of

PIN2-GFP proteins were evident in smt2 smt3 on the apical side, basal side, and on the

disconnected cell wall (Fig. 11a, Fig. 12). PIN2-GFP protein levels were lower in smt2

smt3 than WT roots (Fig. 11b), and GFP-positive granular structures were observed in

smt2 smt3 root cells (Fig. 11c). Thus, these PIN2 mislocalization patterns were ascribed

to failure of cell division in smt2 smt3.

20

Auxin biosynthesis and auxin degradation

The DR5:GUS expression in smt2 smt3 roots was strikingly intense, suggesting

that auxin biosynthesis and/or auxin degradation may be disturbed in smt2 smt3.

However, indole-3-acetic acid (IAA) biosynthesis genes including TAA1, YUC1, YUC4,

YUC6, CYP79B2 and CYP79B3 expression levels in smt2 smt3 seedlings did not

increase as compared to those of WT (Fig. 13). And GH3 genes which are involved in

IAA degradation (GH3.2, GH3.5 and GH3.6) expression levels were also unaffected in

smt2 smt3 seedlings (Fig. 13).These results indicated that the lack of C24-ethylsterol

didn’t affect the auxin biosynthesis and the auxin degradation.

21

Figure 6. Shoot phenotypic comparison of smt2, smt3, and smt2 smt3 mutant plants

(a) 4-week-old WT, smt2, smt3, and smt2smt3 plants. Scale bar = 5 cm. (b) Cotyledon vasculatures in WT, smt2, smt3 and smt2 smt3. Scale bar = 1 mm. (c) Abnormal leaf development in smt2 smt3 plants. Curly leaves of smt3 were observed, whereas the smt2 leaves did not display a similar phenotype. Scale bar = 5 mm. (d) Abnormal flower development in smt2 smt3. Frayed petals and Curly sepals were observed on smt2 smt3

flower. Scale bar = 1 mm. (e) smt2 smt3 plants were fertile. Abnormal siliques contained aborted seeds. Scale bar = 1 mm. (f) Toluidine blue-stained ovules in siliques from WT and

co-suppressed SMT2 plants (Schaeffer et al., 2001) (smt2COSUP

) at the mature green silique stage

(Schaeffer et al., 2001). Representative images of several siliques taken from 10 to 15

plants are shown. Scale bar = 50 m. (g) Pistil cross sections. The abnormal transmitting tract development was observed in smt2 smt3 plants (red arrows). Scale bar = 200 m.

22

Figure 7. Embryonic defects in smt2 smt3

Development of WT and smt2 smt3 embryo. (a) Deformed shapes and irregular orientations of suspensor cells and hypophysis cells (red arrow) were observed in smt2 smt3 embryo at 8-cell and 16-cell stage. Abnormal cell division pattern (red arrow) were observed at the center of globular and heart stage in smt2

smt3 (b) DR5:GUS staining of WT and smt2 smt3 embryos. smt2 smt3 embryos displayed abnormally strong expression. Scale bar = 100 m

23

Figure 8. The smt2 smt3 root phenotypes

(a) Seedlings of 5-day-old wild type (WT), smt2, smt3 and smt2 smt3 plants. (b) A representative smt2 smt3 seedling phenotype: a twisted short root and wavy cotyledons. (c) WT, smt2, smt3, and smt2 smt3 plant root morphology. Disrupted smt2 smt3 root development is completely affected, demonstrating abnormal tissue architecture, occasional swollen cells, and severe growth inhibition. Scale bar = 200 m.

24

Figure 9. Aberrant cell division and cell death in smt2 smt3 roots

4-day-old seedlings of WT and smt2 smt3 (a) WT and smt2 smt3 seedlings were treated with 50 g/ml FM4-64 for membranes staining. Comparison of the tissue architectures between WT and smt2 smt3 roots. In smt2 smt3 roots, unusual cell shapes and a loss of ordered cell orientations (arrows) are prominent. The arrowheads indicate the cortex stem cells, demonstrating the incorrect direction of stem cell division in the smt2 smt3 roots. The insets present magnified root cap images. (b) WT and smt2 smt3 seedlings were treated with 10 g/ml propidium iodide (PI) for cell wall staining. In smt2 smt3, aberrant cell wall structures and PI-positive cells were observed. The PI-positive cells represent the cell death (white arrows). The root specimens were observed using a confocal microscope (LSM700, Carl Zeiss) and the ZEN 2011 software (Zeiss Efficient Navigation 2011) at 555 nm for excitation. Fluorescence emissions of FM4-64 and PI were detected at 640 to 700 nm and at 556 to 700 nm, respectively. Scale bars = 50 m.

25

Figure 10. The smt2 smt3 root phenotype indicates disrupted auxin responses

(a)plates in the dark. From each plant line, 100 seedlings were used to monitor the gravitropic bending responses. Agar plates were rotated 90° counter-clockwise, and the gravitropic responses were determined after 24 h by scoring the angles of the bent roots that differed from 0°; this orientation corresponded to a complete vertical reorientation. The bars represent the number of seedlings that displayed differential gravitropic responses. The arrow indicates 0° (complete reorientation). (b) Abnormal starch granule arrangement in the smt2 smt3 root caps. Scale bar = 50 m. Amyloplasts were visualized with Lugol’s iodine staining solution. (c) DR5:GUS

staining of WT, smt2, smt3, and smt2 smt3 plant roots. The WT staining pattern is also observed in smt2 and smt3 mutants, whereas smt2 smt3 roots display a strongly affected GUS staining pattern. Scale bar = 100 m. The specimens were visualized using the IX71 microscopic system (Olympus, Tokyo, Japan).

26

Figure 11. Abnormal cell file and loss of PIN2-GFP polar localization in smt2 smt3 roots

(a) The PIN2-GFP-expressing line was stained with 50 g/ml FM4-64. In WT plants, PIN2-GFP localizes upward in epidermal cells and downward in both cortical (c) and epidermal cells (e). In smt2 smt3 plants, abnormal PIN2-GFP localization is observed with occasional lateral side and upside-down localizations (white arrowheads). An aberrant cell file is also prominent in smt2 smt3 roots (circled by the dotted line). (b) PIN2-GFP localization in the WT and smt2 smt3 root epidermal cells. The PIN2-GFP level was significantly lower in smt2 smt3 than in the WT. (c) Small PIN2-GFP positive particles (white arrows) were observed in the smt2 smt3 root epidermal cells. Scale bar = 10 m

27

Figure 12. PIN2-GFP localization at cell plate during cytokinesis

(a) PIN2-GFP expression was visualized with FM4-64 staining. PIN2-GFP

accumulated at the phragmoplast in both WT and smt2 smt3 roots. (b) PIN2-GFP

localization at the abnormally-formed cell plates (disconnected form; unusually

anchored form) was observed in smt2 smt3. The fluorescence images were

obtained using a confocal microscope (LSM700) and the ZEN 2011 software (Carl

Zeiss). Scale bar = 10 m.

28

Figure 13. Auxin biosynthesis and degradation in smt2 smt3 seedlings

(a) IAA biosynthesis and degradation gene transcription levels in WT and smt2 smt3

seedling were determined by semi-quantitative RT-PCR. 7-day-old seedlings germinated on MS-agar plates. PCR reaction cycles are indicated to the top of the panels. (b) auxin biosynthesis pathway.

IAA: indole-3-acetic acid, IPA: indole-3-pyruvic acid, IAOx: indole-3-acetaldoxime, IAN: indole-3-acetonitrile, IAM: indole-3-acetamide, TAM: tryptamine, IAAld: indole-3-acetaldehyde

29

Cell division is disrupted in smt2 smt3 plants

Abnormal cell division is prominent in the smt2 smt3 root stem cells (Fig. 9a). The

FM4-64 staining of smt2 smt3 roots revealed the disruption of the formative division of

root stem cells followed by cell lineages with deformed shapes and irregular orientations

as well as abnormally formed cell plates (Fig. 9a, 12b). In the asymmetric divisions, the

division planes are governed by the polarity of the mother cell (Rasmussen et al., 2010).

Thus, the abnormal cytokinesis in the smt2 smt3 root stem cells may have caused the

defects in subsequent cell divisions that led to the collapsed architecture of the following

cell lineages (Fig. 9). The FM4-64 staining of the tubulin-GFP expression line revealed

both abnormal and normal cell division in smt2 smt3 roots (Fig. 14b). The time-lapse

imaging analysis revealed delayed cytokinesis, which resulted in improper division zone

formation in smt2 smt3 (Fig. 14c). The misorientation and perpendicular formation of the

preprophase band (PPB) in abnormal cell division were visualized by immunostaining

with anti-tubulin and anti-KNOLLE syntaxin antibodies (Fig. 15a). Arabidopsis KNOLLE

is a vesicle-docking syntaxin specifically expressed during mitosis and localizes to the

cell division plane (Lauber et al., 1997). Abnormal phragmoplast formation in smt2 smt3

was also indicated by the disjunctive distributions of KNOLLE and tubulin, as well as the

disorganized and fragmented spindles (Fig. 15b). Staining with DAPI

(4',6-diamidino-2-phenylindole) revealed the presence of multinucleated cells in the

smt2 smt3 roots (Fig. 15b, 16). Mislocalization of PIN2-GFP was demonstrated at the

unusually formed division zone in smt2 smt3 root cells (Fig. 12).

In smt2 smt3, tubulin-GFP signal was observed with FM4-64 fluorescence during

the phragmoplast formation stage in a similar way as visualized in WT roots (Fig. 14).

PIN2-GFP signal co-localized with FM4-64 (Fig. 12), irrespective of the normal and

abnormal cell division. Also, the immunostaining experiment demonstrated in smt2 smt3

that tubulin assembled at cell plate shown by KNOLLE localization (Fig. 15b). These

30

results indicate that the process of phragmoplast formation and the transport of newly

synthesized PIN2 protein are not defective in smt2 smt3 roots. We did not observe

KNOLLE mis-localization to lateral PM as reported with the upstream pathway mutants,

cpi1-1 and smt1orc, and roots treated with fenpropimorph (Boutté et al., 2010), a general

sterol biosynthetic inhibitor. KNOLLE localization in the cell division plane is maintained

through sterol-dependent endocytosis mediated by a clathrin- and DYNAMIN-RELATED

PROTEIN1A-dependent mechanism (Boutté et al., 2010; Lauber et al., 1997). These

observations could be due to the intrinsic differences in the sterol compositions between

the total defects in the sterol biosynthesis (cpi1-1 and smt1orc) and the specific loss of

C24-ethylsterols (smt2 smt3). In summary, the cytokinesis in smt2 smt3 plants was

delayed, as visualized by the FM4-64 staining of the tubulin-GFP expressing line (Fig.

14c). Thus, smt2 smt3 roots are defective in the determination of the cell division plane,

which is controlled by the consecutive events comprised of the positioning, anchoring,

and guidance of the division zone (Van Damme, 2009).

31

Figure 14. Root cytokinesis were visualized in the Tubulin-GFP-expressing line

stained with FM4-64

(a) Cytokinesis in the WT GFP-tubulin-expressing lines. (b) An abnormal division zone was occasionally observed with unusual phragmoplasts in smt2 smt3 (white arrows, tubulin; white arrowhead, FM4-64 staining). (c) The microtuble positions and early and late phragmoplast formation are seen during the 30 min time course in WT roots. Cytokinesis in smt2 smt3 roots is significantly delayed and associated with the formation of twisted phragmoplasts. Images were visualized using a confocal microscope (LSM700, Carl Zeiss) and the ZEN 2011 software (Zeiss Efficient Navigation 2011). Scale bar = 10 m.

32

Figure 15. PPB and phragmoplat were visualized by immunostaining with

anti-tubulin and anti-KNOLLE

(a) Abnormal preprophase band (PPB) formation in smt2 smt3 was occasionally observed with perpendicular angles (white arrows). The nuclei (blue) are stained with DAPI (4′,’6-diamidino-2-phenylindole). (b) The cell plate formation stage was analyzed by immunostaining using anti-tubulin antibodies (green), antiserum for KNOLLE (red), and DAPI (blue). Abnormal cytokinesis in smt2 smt3 is demonstrated by incomplete division, mislocalized KNOLLE (white arrow heads), and disordered tubulin accumulation (white arrow). A fragmented cell plate (KNOLLE positive structure, white arrow heads) with double-nuclei (DAPI), and disorganized cell plate direction in two adjacent cells (two-way arrows) are shown. Scale bar = 10 m

33

Figure 16. The root tissue architecture was visualized using whole mount

immunostaining with anti-tubulin antibodies and 4',6-diamidino-2-phenylindole

(DAPI)

(a) The orientation of WT root tubulin. (b) The orientation of abnormal cortical microtubules and double-nucleus cells (white arrows) in smt2 smt3 roots. (c) A magnified image of a typical root cell in smt2 smt3 roots. (d) A magnified image of a double nucleus cell in smt2 smt3 roots. The fluorescence images were obtained using a confocal microscope (LSM700, Carl Zeiss) and the ZEN 2011 software (Zeiss

Efficient Navigation 2011). 4',6-diamidino-2-phenylindole (DAPI) (10 g/ml) was used for DNA staining. Scale bar = 10 m.

34

Partial restoration of the disrupted cell division by exogenously supplied

-sitosterol

The roots were cultured in a liquid medium supplemented with synthetic auxins to

induce the differentiation of secondary roots. Under typical culture conditions, WT roots

successfully differentiated and developed secondary roots, while the smt2 smt3 roots

yielded only disorganized tissue bumps (Fig. 17a). Intriguingly, exogenous -sitosterol

administration partially rescued the failed secondary root development (Fig. 17a),

without restoring normal C24-ethylsterol levels, including steryl glycosides, acylated

steryl esters, and acylated steryl glycosides (Fig. 18, Table 2). The

-sitosterol-dependent recovery of the second root development was partly substituted

by stigmasterol (another significant C24-ethylsterol) but not by campesterol

(C24-methylsterol) (Fig. 17a, b). Secondary root development in WT plants is

associated with DR5:GUS expression that is localized to the secondary root tips, similar

to that in intact plant roots, which completely disappeared in smt2 smt3 plants (Fig. 17c).

The irregular DR5:GUS expression pattern in the smt2 smt3 roots was also rescued by

the addition of -sitosterol (Fig. 17c), suggesting that directional auxin transport and cell

polarity were amended to enable formative cell division in smt2 smt3 roots. -sitosterol

induced successful secondary root development (Fig. 17). The root culture experiment

may be a suitable system to study the requirements of sterols in plant development.

The disturbed cytokinesis in smt2 smt3 roots was further assessed at the single cell

level. Phragmoplast structures were visualized using GFP-tubulin and FM4-64 staining

in the late cytokinesis of cells from WT and smt2 smt3 cultured roots recovering from

protoplast isolation (Fig. 19). Protoplasts initiated cytokinesis at least 96 h after the

incubation, and almost 75% of WT protoplasts completed abscission at 102 h of

incubation. Conversely, most smt2 smt3 protoplasts failed to form normal

phragmoplasts during cytokinesis (Fig. 19a, b). Cytokinesis completion was scored by

35

visually discriminating between the telophase stage and the completion of cell division

(Fig. 19c). Only 10% of smt2 smt3 protoplasts completed cytokinesis within the same

time course, and further incubation did not increase the cytokinesis achievement rates

(Fig. 19d). However, when protoplasts were isolated from smt2 smt3 roots cultured with

-sitosterol, the cytokinetic abscission rate increased to 30% (Fig. 19d). The addition of

-sitosterol to the culture medium did not elicit acute effects on protoplasts in the

recovery of disrupted cytokinesis. Cellulose fiber formation by WT and smt2 smt3

protoplasts was indistinguishable under our experimental conditions (Fig. 20) The

involvement of sterols in cellulose biosynthesis (Schrick et al., 2004; Schrick et al.,

2012) was not observed in the surfaces of WT and smt2 smt3 protoplast-derived cells

(Fig. 20b). However, aberrant cell wall structures in smt2 smt3 (Fig. 9b) were consistent

with the suggestion that C24-ethylsterols are involved in cellulose biosynthesis. These

results demonstrate that the smt2 smt3 protoplasts cannot determine the cell division

plane and that the cell population capable of normal cytokinesis increased during the

secondary root development in the smt2 smt3 roots cultured with -sitosterol. The

expression of the DR5:GUS reporter suggested that the directional auxin transport in

the smt2 smt3 roots was restored upon -sitosterol supplementation (Fig. 17). The

supplemental -sitosterol may have rescued the dysfunctional root development by

restoring auxin transport and thereby promoting the auxin distribution-dependent plant

development (Dhonukshe et al., 2008).

36

Figure 17. Lateral root development in cultured roots

(a) smt2 smt3 cultured roots developed abnormal lateral roots, leading to the formation of cell clumps. The addition of exogenous -sitosterol (3.0 g/ml) restored the lateral root growth of smt2 smt3. The addition of stigmasterol was also effective to some extent, whereas campesterol was not. (b)

Exogenous -sitosterol restored lateral root growth. Roots were cultured for 7 days and supplemented with 3 g/ml -sitosterol. In each treatment, a total of 100 lateral roots and root primordia were visually identified under a microscope (SZX-10, Olympus), and the numbers of successful root growth (longer than 200 m in length) were then scored. The y-axis compares the percentage of the successful root growth among the treatments. The results are the average of three experiments (means ± SD). The asterisks indicate statistical significance at P < 0.001 (Student’s t-test). (c) WT roots exhibited a DR5:GUS expression pattern similar to that of intact plant roots, indicating the generation of an auxin gradient. The addition of -sitosterol restored an auxin gradient and the development of lateral root growth in smt2 smt3. Scale bars = 100 m.

37

Figure 18. Analysis of -sitosterol derivatives in the -sitosterol-supplemented cultured-roots

Total sterol fractions were obtained according to Brigh and Dyer (1959) and separated by TLC to obtain free sterol fraction and sterol conjugate fractions (sterol fatty acid ester, SE, sterol glucosides; SG and acylated steryl glucosides; ASG). The FS and SE fractions were saponified, and SG and ASG fractions were subjected to the hydrolysis. All samples were analyzed with a 6890 gas chromatograph (Agilent) equipped with a HP5-MS column (J & W; 30m long, 0.32 mm id., 0.25 mm film thickness) and coupled with a 5973 mass selective detector. 1; campesterol, 2; stigmasterol, 3; -sitosterol, 4; -amyrin, 5; -amyrin, 6; lupenol, UN; Unknown peak commonly detected in the analyzed samples. Peak intensities are indicated to compare the sterol levels.

38

Table 2. Sterol profiles of the cultured roots

39

Figure 19. Abnormal smt2 smt3 root protoplast cell division

Protoplasts were prepared from tubulin-GFP-expressing lines. (a) The WT and smt2 smt3 root protoplasts at the cytokinesis stage exhibited normal phragmoplast structure formation as indicated by tubulin-GFP fluorescence. Abnormal phragmoplast formation in smt2 smt3

protoplasts (arrow head). (b) FM4-64-stained protoplasts at the cytokinesis stage. The upper panel is a representative WT protoplast image, and the lower panel is an example of abnormal phragmoplast formation (arrow heads) in smt2 smt3 protoplasts. Fragmented phragmoplasts and collapsed PM structures were observed in the smt2 smt3 protoplast (arrow heads). (c) The completion of cell division was evaluated by distinguishing the telophase-stage cells from those that had finished cytokinesis. (d) Exogenously supplied -sitosterol partially restored the delayed cell division of smt2 smt3 protoplasts prepared from roots cultured in the presence or absence of 3.0 g/ml -sitosterol. After incubation for 102 h, live cells (1000 cells from each experiment) were identified using PM staining with FM4-64. The numbers of cells at the telophase-stage ( ) and completed cytokinesis ( ) were counted individually counted. Data are presented as the mean ± standard deviation (SD) of three independent experiments. The asterisk indicates significance at P

< 0.05 (Student’s t-test). Scale bars = 10 m.

40

Figure 20. Time course of cellulose accumulation (calcofulor white staining) in

protoplasts

(a) WT and smt2 smt3 protoplasts stained with calcofulor white. There was no significant difference in the time course of the cellulose accumulation between WT and smt2 smt3 protoplasts. (b) Cellulose fiber formation on the surfaces of WT and smt2 smt3 protoplasts. The protoplasts were stained with calcofulor white after a 10-h incubation. The images of the calcofulor white-stained specimens were collected using an Olumpus IX17 fluorescent microscope (Olympus) with the excitation at 330-385 nm and emission at 415-488 nm. Scale bar = 10 m.

41

Lipid order of cell plate in smt2 smt3

The root culture experiments (Fig. 17, 18 and Table 2) demonstrated that a small

amount of -sitosterol (or stigmasterol) critically promoted the recovery of abnormal

growth of lateral roots of smt2 smt3. However, we were not able to identify metabolites

in very low amounts derived from the exogenously supplied -sitosterol in the cultured

roots. The exogenously supplied -sitosterol may accumulate at specific sites, such as

membrane microdomains in plants (Malinsky et al., 2013), in a temporary and spatially

precise manner at either the tissue or single cell level. Frescatada-Rosa et al. (2014)

monitored the membrane lipid packing state by recording generalized polarization (GP)

values (Owen et al., 2012) and observed that the lipid order increased at the cell plate

during cytokinesis in a sterol- and DRP1A- (DYNAMIN-RELATED PROTEIN1A)

dependent manner. The lipid order of cell plate in roots of cpi1-1 and roots treated with

fenpropimorph remain unchanged (Frescatada-Rosa et al. 2014). Their results

prompted us to examine whether the cell plate might be the specific site for

C24-ethylsterol accumulation. If C24-ethylsterols accumulate specifically at the cell

plate and are required for correct cytokinesis, the lipid order at the cell plate in smt2

smt3 lacking C24-ethylsterols should remain unchanged. Then we observed the lipid

order of cell plate in WT and smt2 smt3 root cells. The lipid order in smt2 smt3 shifted

toward a higher order, as observed in the WT, irrespective of correct or failed cytokinesis

(Fig. 28). By contrast, the GP values recorded in smt2 smt3 roots were generally higher

than those in the WT, regardless of cytokinetic stage, than those in the WT, suggesting

intrinsically different membrane properties between the WT and smt2 smt3 (Fig. 21).

42

Figure 21. Ratiometric live imaging to monitor the membrane lipid order

(a) During cytokinesis, the membrane lipid packing state at the cell plate changed in a similar manner in both WT and smt2 smt3 roots: initially shifted toward ordered-phase (high order) at the early stage and then gradually decreased to disordered-phase (low order) at the middle and late stages. Note that the perpendicularly aligned cells in smt2 smt3 exhibited the lipid order shift in a similar manner as observed in WT. (b) Relative changes in the ratio of GP values of the cell plate (CP) and the PM were compared among the cytokinetic early, midle, and late stages were form 10 independent cells of WT and smt2

smt3, respectively. The fluorescence imaging analysis was performed according to Owen et al. (2012). Scale bar = 10 m

43

Chapter2

Clarification of the disturbed polar localization of auxin efflux protein

smt2 smt3 plants are impaired in the polar localization and endocytic recycling of

the PIN2 auxin efflux transporter protein

In WT plants, the PIN2 auxin efflux transporter localized correctly, maintaining the

downward and upward directions in the epidermis and the cortex cells, respectively

(Carland et al., 2010; Blilou et al., 2005). Multiple mislocalization patterns of PIN2-GFP

proteins were evident in smt2 smt3 on the apical side, basal side, and on the

disconnected cell wall (Fig. 11a). PIN2-GFP protein levels were lower in smt2 smt3 than

WT roots (Fig. 11b), and GFP-positive granular structures were observed in smt2 smt3

root cells (Fig. 11c). Immunohistochemical analysis revealed PIN2-GFP protein

mislocalization and the PIN2-GFP protein mislocalization was associated with abnormal

cell divisions (Fig. 22). Thus, these PIN2 mislocalization patterns were ascribed to

failure of cell division in smt2 smt3. Furthermore, the polar distribution of the PIN2

protein was not appropriately established in the epidermal cells of smt2 smt3 roots (Fig.

11). The correct PIN2 localization at the late cytokinesis stage (Fig. 12) and the cell

plate (Fig. 22) suggest that PIN2 protein trafficking operates normally in smt2 smt3.

It has been demonstrated that the intracellular asymmetric PIN2 distribution is a

controlled through a sterol-dependent clathrin-mediated endocytosis from the PM and

polar recycling to the PM as a post-cytokinetic event (Men et al., ellular 2008;

Dhonukshe et al., 2008; Blilou et al., 2005). Initial PIN2 internalization can be studied by

inhibiting endocytotic vesicular transport with brefeldin toxin A (BFA) (Xu and Scheres,

2005; Kleine-Vehn et al., 2008) in the presence of cycloheximide (CHX). The BFA

treatment induced the aggregate formation of PIN2-GFP and the endocytic marker

44

FM4-64 fluorescence within the cells (BFA compartments), indicating that the

endocytosed endomembrane aggregated where sterols co-accumulate with an

endocytic marker ARA6 (a Rab5-GTPase homolog) and PIN2 in Arabidopsis (Grebe et

al., 2003). FM4-64 pulse labeling demonstrated that the fluorescence signals of

PIN2-GFP and FM4-64 on the PM decreased with time in both the WT and smt2 smt3

root epidermal cells (Fig. 23). Compared with the WT, PIN2-GFP was internalized more

rapidly in smt2 smt3 (Fig. 23b), and prolonged residence of the BFA compartments was

apparent (Fig. 23a). By contrast, FM4-64 internalization was more rapid in the WT than

in smt2 smt3 (Fig. 23c). These observations are in clear contrast to the total endocytotic

inhibition observed in cpi1-1 plants (Men et al., 2008). The GFP-positive but

FM4-64-negative structures (Fig. 23a) may correspond to the PIN2-GFP-positive

particles present in smt2 smt3 before drug treatment (Fig. 11c, white arrows). Upon

BFA-washout, the BFA compartments in WT plants gradually disappeared (Fig. 24a),

whereas those in smt2 smt3 mutants remained within the cells, with a number of small

GFP-positive vesicles becoming apparent (Fig. 24a), and the PIN2-GFP signal on the

PM was dramatically reduced in smt2 smt3. These observations indicate that the

endocytosis and vesicular recycling are disturbed in smt2 smt3 plants, leading to the loss

of PIN2-GFP polar localization (Fig. 24). However, the disturbed endocytosis in smt2

smt3 roots was not restored by the addition of exogenous -sitosterol (Fig. 25). We

cannot rule out the possibility that exogenous -sitosterol may be unable to restore

endocytic recycling in the presence of the protein synthesis inhibitor CHX.

The unusual patterns of tubulin-GFP and ABD2-GFP (Fig. 26, 28) in smt2 smt3

roots suggested a possible involvement of C24-ethylsterols in the orientations of

cytoskeletal proteins. Early endocytic sterol transport has been demonstrated to depend

on actin (Grebe et al., 2003). The lack of efficient maintenance of PIN2 protein levels in

smt2 smt3 (Fig. 11b) may be due to disturbed endocytic recycling (Fig. 24). The

45

retromer component SNX1 (SORTING NEXIN 1) is involved in the endocytic recycling

of PIN2 to the PM (Jaillais et al. 2006; Kleine-Vehn et al. 2008). Ambrose et al. (2013)

demonstrated the involvement of microtubules (MTs) in PIN2 recycling through the

interaction of the MT-associated protein CLASP with SNX1. The aberrant localization of

GFP–CLASP as well as unusual MT orientations in smt2 smt3 (Fig. 26 and 27) suggest

that the retromer function is interrupted under C24-ethylsterol-deficient conditions,

leading to the retardation of PIN2-GFP recycling and PIN2 mislocalization.

46

Figure 22. Immunostaining of the PIN2 protein

PIN2-GFP localization was visualized by whole mount immunostaining using anti-GFP antibodies. The normal polarity of the PIN2-GFP localization (white arrows) was frequently lost in the smt2 smt3 root epidermal cells (blue arrows). The PIN2-GFP protein was detected at the division plane in the smt2 smt3 root as well as in the WT (white triangles), suggesting that the post-cytokinetic polar distribution of PIN2-GFP was disrupted in smt2 smt3. Abnormal PIN2 localization was also associated with perpendicular cell division (blue triangle). Scale bars = 10 m.

47

Figure 23. Internalization in the root epidermal cells

(a) Five-day-old seedlings were incubated for 30 min in 0.5× MS medium containing 50 M cycloheximide and stained for 10 min with 50 g/ml FM4-64 in the presence of 50 M CHX and 50 M brefeldin A (BFA). The tissue fluorescence images were recorded at indicated times (after BFA treatment). Endocytic PIN2-GFP internalization was not inhibited in the smt2 smt3 root epidermal cells, whereas PIN2-GFP signals became prominent over time in the intracellular space in smt2 smt3. GFP-positive structures without a FM4-64 signal were also observed in smt2

smt3.(b, c) The relative changes in the GFP and FM4-64 fluorescence intensities are presented in (b) and (c), respectively. Vertical axes show the relative changes in the fluorescence intensities from the PM (WT_PM, smt2 smt3_PM) and the intracellular space (WT_In, smt2 smt3_In). Fluorescence intensities were measured in 10 independent cells from 4 different seedlings, and the average values and standard errors are shown in (b) and (c) Scale bar = 10 m

48

Figure 24. Endocytic recycling in the root epidermal cells

(a) The roots were treated with 50 M BFA for 90 min to induce the formation of BFA compartments, followed by an additional treatment with 50 M CHX and 50 M BFA

for 30 min. The tissues were then washed with 0.5× MS medium containing 50 M CHX and incubated for 90 min without BFA. The incubation periods after BFA removal are indicated.

(b) The time course of the decreased intracellular (excluding the PM) signal intensities from PIN2-GFP in the WT and smt2 smt3 demonstrated the loss of PIN2-GFP from the PM during the treatment. The average values and standard errors were determined from 20 cells of four independent seedlings of the WT and smt2 smt3 at each respective time point. The average values and the standard errors are from 20 independent cells. Scale bars = 10 m.

49

Figure 25. Effects of exogenous sterols on PIN2-GFP endocytosis

Roots were treated with 50 M CHX for 10 min,which was followed by an additional treatment with 50 M CHX and 50 M BFA in 0.5× MS medium. The confocal microscopic images were collected after a 40-min treatment. -sitosterol and campesterol were added at a concentration of 3 g/ml throughout the incubation period. The PIN2-GFP signal intensities are higher in smt2

smt3 than in WT roots. GFP-positive vesicles were smaller in the smt2 smt3 than in the WT roots.

(a, b) A representative control experiment (only ethanol added). (c, d) Exogenous -sitosterol did not reverse the inhibition of PIN2-GFP endocytosis in smt2 smt3 plants. (e, f) Exogenous campesterol did not reverse the inhibition of PIN2-GFP endocytosis in smt2 smt3. (g) Percentage of intracellular fluorescence intensity (excluding PM) after the treatment of 50 M CHX and 50 M BFA for 40 min. There was no significant difference in the intracellular fluorescence intensities in either WT or smt2 smt3 in the presence or absence of -sitosterol and campesterol. The results were obtained from 25 root cells from 6 independent seedlings for each treatment. The fluorescence images were collected using a confocal microscopy

(LSM700) and the ZEN 2011 software (Carl Zeiss). Scale bar = 10 m.

50

Figure 26. Cortical microtubules in the WT and smt2 smt3 roots

4-day-old seedlings were stained using anti-tubulin antibodies. In WT roots, the

cortical microtubules are visualized on the lateral side in the cortex (c) and the outer

cell-side in epidermis cells (e). In smt2 smt3 roots, cortical microtubules are

distributed evenly in the cortex cells, and the localized microtubule orientation is

completely abolished in the epidermal cells. Fluorescence images were collected

using a confocal microscope (LSM700; Carl Zeiss) and the ZEN 2011 software

(Zeiss Efficient Navigation 2011). Scale bar = 10 m.

51

Figure 27. Aberrant localization of GFP-CLASP in smt2 smt3 roots

Staining with 50 g/ml FM4-64 for 10 min in 0.5x MS medium,

pCLASP:GFP-CLASP seedlings were observed using a confocal microscope LSM

700 (Carl Zeiss) with the excitation at 488 nm for GFP and 555 nm for FM4-64.

Fluorescence emissions of GFP and FM4-64 were detected at 490 to 555 nm and

640 to 700 nm, respectively. c; cortex cells, e; epidermal cells. Scale bar = 10 m

52

Figure 28. Root hair actin filaments

Root hairs in the WT and the smt2 smt3 plants expressing GFP-AtFim1ABD2. Images were obtained using an Olympus IX17 fluorescent microscope (Olympus). Scale bars = 10 m.

53

Chapter 3

General discussion

Sterols serve as essential membrane constituents, and therefore, there is no doubt

that the lack of major sterols or the abnormal sterol compositions that result from

genetic defects affect physicochemical properties of membranes and normal functions

of membrane proteins, thereby influencing diverse aspects of developmental processes

in eukaryotic life. Ample evidence supports the role of plant sterols in the generation of

the auxin concentration gradients achieved through the directional transport of auxin by

carrier proteins (Dhonukshe et al., 2008; Petrásek and Friml, 2009; Feraru and Friml,

2008). It has also been reported that sterols are required for correct auxin and ethylene

crosstalk to achieve vascular development (Pullen et al., 2010). Furthermore, disrupted

cell divisions in the sterol biosynthetic mutants smt1 (Schrick et al., 2004; Schrick et al.,

2002) and cpi1-1 (Men et al., 2008), or ectopic cell division leading to abnormal vascular

patterning in the hydra1 and fackel/hydra2 mutants reinforce the implication of sterols in

major cellular processes (Pullen et al., 2010). However, the possible relationships

between given plant sterol molecular species, cell division, and phytohormones have

not been clearly deduced from studies mentioned above (Fig. 1). In this report, we show

the essentiality of C24-ethylsterols using the Arabidopsis smt2 smt3 mutant that is

defective in the Charophycean green algae and land plant specific sterol

methyltransferase and that bear therefore a strongly modified distribution of

24-alkyl-5-sterols (Fig. 1). We present evidence that C24-ethylsterols are involved in

the endocytic recycling and the polarity control of the PIN2 protein as well as the

54

determination of the cell division plane to establish polarized growth.

Cell division

In smt2 smt3 roots, aberrant cell division plane formation was observed at the step

of PPB formation, and collapsed cell division was accompanied by the uncoordinated

localizations of KNOLLE and tubulin to the cell division plane. However, once PPB

formation occurred correctly, cytokinesis appeared to proceed normally. Thus,

C24-ethylsterols may be required for the determination of the cell division plane, which

is controlled by the consecutive events comprised of the positioning, anchoring, and

guidance of the division zone (Van Damme, 2009). The root culture experiments

demonstrated that a small amount of -sitosterol (or stigmasterol) critically promoted the

recovery of abnormal secondary root growth of smt2 smt3 roots. However, we were not

able to identify metabolites in very low amounts derived from the exogenously supplied

-sitosterol in the cultured roots (Fig. 18), and we cannot rule out the possibility that

-sitosterol may be further modified to yield novel bioactive sterol structures. We have

not eliminated the possibility that the exogenously supplied -sitosterol may have

accumulated at an extremely specific site, such as the membrane microdomain in plants

(Malinsky et al., 2013). Because the intracellular trafficking of the ABCB19 auxin

transporter in Arabidopsis was interrupted within 2 h of the treatment of fenpropimorph

(Yang et al., 2013), suggesting that spatiotemporal sterol synthesis, rather than total

sterol composition, may be required for endomembrane trafficking. The observation of

lipid order of cell plate demonstrated that the lipid order in smt2 smt3 shifted toward a

higher order, as observed in the WT, irrespective of correct or failed cytokinesis (Fig. 21).

By contrast, the GP values recorded in smt2 smt3 roots were generally higher than

55

those in the WT, regardless of cytokinetic stage, suggesting intrinsically different

membrane properties between the WT and smt2 smt3 (Fig. 21). The

C24-ethylsterol-deficient membrane in smt2 smt3 may represent an actual case of

membrane organization alteration by different phytosterol compositions in a model

system (Grosjean et al., 2015). Although the total disruption of sterol biosynthesis led to

the failure of normal cell plate formation via a sterol- and DRP1A-dependent mechanism

(Frescatada-Rosa et al., 2014), the vesicle transport responsible for cell plate formation

may not be reliant on C24-ethylsterols. In addition, KNOLLE maintenance in the plane

of cell division is also dependent on a sterol- and DRP1A-dependent mechanism, which

was not affected in smt2 smt3 (depletion of C24-ethylsterols) but is inhibited when sterol

biosynthesis is totally abolished (Boutté et al., 2010). These results indicate that the

totally disrupted sterol composition and the depletion of C24-ethylsterols may have

differentially influenced the processes of cell division.

PIN2 polarity and auxin transport

In plants, mutations in the sterol biosynthetic pathway (Fig. 1) generally affect polar

auxin transport. In fact, the mutants of smt1/cph/orc have reduced polar auxin transport

due to the mislocalization of PIN1 and PIN3 (Willemsen et al., 2003), and the cpi1-1

mutation results in the loss of polar localization of PIN1 and PIN2 (Men et al., 2008). The

interference of polar auxin transport in the sterol mutants are explainable by the

involvement of sterols in the cell polarity control and endosomal vesicle trafficking (Men

et al., 2008; Feraru and Friml, 2008; Grebe et al., 2003). The unaffected PIN1 and PIN2

polarity in sterol mutants such as hyd1 and hyd2/fk (Souter et al., 2002; Pullen et al.,

2010) and the unchanged polar auxin transport in cvp1/smt2 (Carland et al., 2002) may

56

implicate multiple roles of sterols in cellular events. Our results with smt2 smt3 mutant

indicate that C24-ethylsterols are involved in the control post-cytokinetic PIN2 polarity

and auxin distribution. The PIN2-GFP protein level was significantly lower in smt2 smt3

than that in the WT (Fig. 11b). The endocytosis experiments indicate that the substantial

decrease in the level of PIN2-GFP at the PM in smt2 smt3 was attributable, at least in

part to inhibited PIN2-GFP recycling (Fig. 24). Although auxin inhibits PIN endocytic

internalization to increase PIN protein levels at the PM, thus facilitating auxin efflux

(Paciorek et al., 2005), the lower PIN2 protein levels observed in smt2 smt3 are not

consistent with a potential increase in tissue auxin levels. The initial internalization of

PIN2-GFP and the endocytic tracer FM4-64 from the PM was not inhibited but rather

slightly facilitated in smt2 smt3. Interestingly, our results are not consistent with the

strong inhibition of the internalization of both FM4-64 and PIN2 in dividing and

post-cytokinetic root epidermal cells of the cpi1-1 mutant, an upstream sterol

biosynthetic mutant exhibiting a severely altered sterol composition (Men et al., 2008).

The intracellular asymmetric PIN2 distribution, which is essential for the directional

transport of auxin, is accomplished by clathrin-mediated endocytosis from the PM and

subsequent polar recycling as a post-cytokinetic event (Blilou et al., 2005; Xu and

Scheres, 2005; Kleine-Vehn and Friml, 2008; Men et al., 2008). PIN2 was localized to

the PM just after cytokinesis (Fig. 12), implying that the polar distribution of PIN2 may be

impaired in smt2 smt3. The interaction of CLASP with SNX1 stabilizes SNX1

endosomes, such as PIN2-carrying SNX1 vesicles on MTs that prevent PIN2

degradation (Ambrose et al., 2013). Thus, the aberrant localization patterns of CLASP

and MTs (Fig. 26, 27) suggest the involvement of C24-ethylsterols in PIN2 endocytic

recycling mediated by cytoskeletal proteins. Both the organization of the MTs and the

57

actin cytoskeletons were broadly affected in smt2 smt3 roots (Fig. 26, 28). Several lines

of evidence support the possible link between the smt2 smt3 phenotype and cytoskeletal

proteins.

Appropriate auxin concentration gradients are a prerequisite for normal

development. Auxin transport may be affected by a class of 4-methyl sterol derivatives

that accumulate upon disruption of the sterol-C4-demethylation enzymatic complex

(Mialoundama et al., 2013). Arabidopsis ERG28 acts as the scaffolding platform for the

functional complex of the sterol C4 demethylation enzymes (Gachotte et al., 2001;

Darnet and Rahier, 2004). Arabidopsis erg28 mutants exhibit serious growth

abnormalities associated with unusually high levels of CMMC, the substrate of the

carboxysterol-C3-dehydrogenase/C4-decarboxylase of the C4-demethylation complex

(Rahier et al., 2006). In the erg28 mutants, the accumulated CMMC is thought to inhibit

polar auxin transport without affecting the endocytosis-dependent control of the PIN1

protein localization (Mialoundama et al., 2013). The sterol substrate of SMT2, namely,

24-methylene lophenol, is also a substrate of the C4-demethylation complex (Fig 1).

Although smt2 smt3 mutant tissues contained small amounts of 24-methylene lophenol,

as well as other 4-methyl sterols (cycloartenol, cycloeucalenol), we did not detect

4-carboxy-4-methyl derivatives in smt2 smt3 seedlings (Fig. 5). Therefore, the

accumulation of the so-called cryptic sterol biosynthetic intermediates (Mialoundama et

al., 2013) appears not to be a factor in smt2 smt3 mutants. In addition, in smt2 smt3

plants, 24-methylene lophenol accumulated to a level comparable with that detected in

the smt2 mutant (Table 1), which exhibited only minor phenotypic defects of leaf

vascular development and petal morphology (Fig. 6b, 6c). Thus, the current

experimental observations in the smt2 smt3 mutant, such as the interference of the

58

endocytosis-dependent recycling of the PIN2 protein and the defective cell division, do

not conform to the suggested involvement of CMMC in the inhibition of polar auxin

transport mediated by ABCB1 and ABCB19 auxin efflux transporters (Mialoundama et

al., 2013).

SMT genes and C24-ethylsterols

We have demonstrated that -sitosterol, rather than sterol composition, plays a

critical role in plant development. -sitosterol differs from campesterol by the presence

of one additional methyl group in the side chain, which is added by the SMT2 specific to

Charophycean green algae and land plants (Neelakandan et al., 2009) (Fig. 2).

Concerning the side chain stereochemistry, land plants may produce (24R,

24S)-24-alkylsterols (McKean and Nes, 1977). The epimeric mixture of

24-methylcholesterol (24R, campesterol; 24S, dihydrobrassicasterol) was shown to be

ubiquitous whereas 24-ethylsterols are exclusively (24R)-24-ethylsterols such as

sitosterol and stigmasterol (Nes et al., 1977; Guo et al., 1995; Schaller, 2003) (Fig. 2)

with the exception of a few plant genus able to produce limited amounts of

(24S)-24-ethylsterols however restricted in the plant lifespan (Fenner and Patterson,

1992). The occurrence of C24-ethylsterols is not restricted to land plants. For example,

Chlamydomonas synthesizes and accumulates ergosterol as the major sterol together

with a C24-ethylsterol, 7-dehydroporiferasterol, as a minor component, and a single

sterol methyltransferase gene is thought to be responsible for both the C24-methylation

and subsequent C28-methylation reactions (Desmond and Gribaldo, 2009). Volvox and

Coccomyxa (a unicellular green algae) contain single SMT genes (Fig. 2); these algal

species may also contain a C24-ethylsterol as reported for Chlamydomonas.

59

Biosynthetic routes leading to these (24S)-24-alkylsterols have been reported in details

(Nes et al., 1990) as well as the identical features of yeast and algal sterol pathways

(Salimova et al., 1999). By contrast, sequence comparisons have suggested that

Charophycean green algae, which are thought to be most closely related to land plants

(Sørensen et al., 2012; Timme et al., 2012), contain two types of SMT genes

responsible for the two-step methyl transfer reactions similar to those operating in land

plants (Fig. 1 and 2). It is thus possible that -sitosterol is produced in these multicellular

Charophycean algae.

We propose that the acquisition of the two-step SMT reactions enabled the

production of -sitosterol, which was recruited to control cell division (correct formation

of PPB and phragmoplast) and the endocytotic recycling of PIN2 to generate an auxin

concentration gradient through directional auxin transport, and thus contributed to the

evolution of polarized plant growth. In addition, our results suggest different sterol

requirements among vesicle trafficking pathways, such as the secretion of newly

synthesized PIN proteins to the PM, endocytosis from the PM, and subsequent polar

endocytic recycling. Importantly, our results indicate that the presence of a very low

amount of -sitosterol, not the normal sterol composition, is critical for plant

developmental processes. Clarification of the spatiotemporal localization of -sitosterol

and its possible binding site, as well as the possible conversion of -sitosterol into an

unknown sterol-derived signaling molecule, should provide insights into the unique role

of this sterol in plants.

60

Supplementary information

Materials and methods

Plant growth conditions

All experiments were carried out using Arabidopsis thaliana, ecotype Columbia

(Col-0). Arabidopsis plants were grown in a growth chamber maintained at 23°C under

continuous light (140–160 μmol m-2s-1). Seeds were sterilized with 70% ethanol (1 min)

and 0.8% sodium hypochlorite (10 min), and then thoroughly washed with sterilized

water. Plants in soil were supplied with a culture solution composed of 1.0 mM NH4NO3,

2.5 mM KNO3, 0.5 mM KH3PO4, 0.5 mM CaCl2·2H2O, and 0.5 mM MgSO4·7H2O

supplemented with micronutirents (50 M H3BO3, 20 M MnCl2·4H2O, 20 M

CuSO4·7H2O, 2 M ZnSO4·7H2O, 0.5 M Na2MoO·4H2O, 100 M Fe-EDTA). For

experiments under sterile conditions, seeds were germinated on agar (0.8%) plates

containing a germination medium (GM) composed of 1× Murashige and Skoog (MS)

salts and 1% (w/v) sucrose (Valvekens et al., 1998).

Transgenic plants

The T-DNA insertion mutant lines of smt2 (GABI_443_F03) and smt3

(SALK_085292) were obtained from the Arabidopsis Biological Resource Center

(ABRC) (Alonso et al., 2003). Homozygotes with regard to the T-DNA insertion events

(smt2/smt2 and smt3/smt3) were identified by polymerase chain reaction (PCR) using

gene-specific primers and T-DNA-specific primers. The homozygous plants of

smt2/smt2 and smt3/smt3 were crossed with each other to establish a double mutant

line of +/smt2-smt3/smt3 for self-pollination to generate smt2/smt2-smt3/smt3 (smt2

smt3) plants. Since the smt2 smt3 plants were self-sterile, the current experiments were

61

performed by maintaining +/smt2-smt3/smt3 plants for self-pollination to obtain smt2

smt3 plants. smt2 smt3 seedlings were easily identified by their abnormal phenotype

(Fig. 8a). Homozygous plants for each T-DNA insertion event were identified by means

of PCR screening and segregation analyses. PCR was performed using primers

designed from the coding sequences for SMT2 and SMT3 (Table 3). For confirmation of

the smt2 mutation, an SMT2 gene-specific primer (SMT2F) and a reverse primer

(SMT2R) together with a primer for T-DNA sequence (GABI LB) were used. For the

smt3 mutation, an SMT3-specific reverse primer (SMT3R) and a forward primer

(SMT3F) together with a primer for T-DNA (LBa1) were used. A DR5:GUS expression

line of A. thaliana (Ulmasov et al., 1997) was crossed to the T-DNA insertion mutant

plants (smt2/smt2, smt3/smt3, and +/smt2-smt3/smt3). The Arabidopsis line

homozygous for the expression cassette of PIN2-GFP (Blilou et al., 2005) was crossed

to the mutant lines (+/smt2-smt3/smt3). The +/smt2-smt3/smt3 plants carrying either the

DR5:GUS cassette or the PIN2-GFP cassette were maintained. A tubulin-GFP

expression line and an actin-GFP line were generated by the transformation of A.

thaliana plants using the expression plasmids containing a coding sequence for either

GFP-NtTub fusion protein (Kumagai et al., 2001) or GFP-AtFim1ABD2 (Sano et al.,

2005) by the floral dip method (Clough et al., 1998). T1 seeds were germinated on GM

agar plates containing 25 g/ml kanamycin. After selecting T2 homozygous lines, T3

plants were crossed to the +/smt2-smt3/smt3 line to obtain F1 seeds, which were

allowed self-pollination to yield F2s (tubulin-GFP expression line of +/smt2-smt3/smt3

and actin-GFP expression line of +/smt2-smt3/smt3). The +/smt2-smt3/smt3 F2 plants

(expressing either tubulin-GFP or actin-GFP) were allowed self-pollination to generate

smt2/smt2-smt3/smt3 F3 plants expressing tubulin-GFP or actin-GFP. The presence of

62

the coding sequences of the GFP-fusion proteins was analyzed by PCR using a primer

set of 35sFw for the CaMV35S promoter and sGFPRv for the GFP sequence (Table 3).

RT-PCR

Reverse-transcription (RT)-PCR analyses were carried out to determine the

expression levels of SMT2 and SMT3 genes in Arabidopsis using the gene-specific

primer sets (Table 3). Arabidopsis actin gene was amplified under the same PCR

condition using a primer pair of Act2F and Act2R as the internal control (Table 3). Total

RNA was isolated using an RNeasy Plant Mini Kit (QIAGEN). Genomic DNA

contamination was eliminated using an RNase free DNase Set (QIAGEN, Tokyo,

Japan). First-strand cDNA was synthesized using a Takara RNA PCR Kit (AMV) Ver.3.0.

(Takara, Kyoto, Japan) in a 10 l reaction mixture containing 500 ng of total RNA using

an oligo(dT)16 as the reverse primer. The RT reactions were carried out at 30°C for 10

min, 50°C for 30 min, and 95°C for 5 min, and then chilled to 5°C for 5 min. The PCR

was done using 2 l of the RT products as the template in a 10 l of reaction mixture

containing 1 mM MgCl2, 0.2 mM dNTP mixture, 0.025 unit/l Takara Ex Taq HS

(Takara), and 0.2 M of primers using a Takara RNA PCR Kit (AMV) Ver.3.0 (Takara).

The PCR was programmed in 30 cycles of 94°C for 30 s, 50°C for 30 s, and 72°C for 1

min followed by an extension period of 10 min at 72°C.

GUS Staining of transgenic plants

Embryos and five-day-old seedlings were treated with 90% (v/v) ice-cold acetone

for 1 h and submerged in GUS staining buffer (Jefferson et al., 1987) containing 100

mM sodium phosphate (pH 7.0), 10 mM EDTA, 0.5 mM potassium ferrocyanide, 0.5 mM

63

potassium ferricyanide, 0.1% (w/v) Triton X-100, and 0.5 g/ml

5-bromo-4-chloro-3-indolyl--D-glucuronide (Nacalai Tesque Inc., Kyoto Japan). The

tissues were then infiltrated with staining buffer under vacuum and incubated overnight

at 37°C. After rinsing with water, tissues were cleared with chloral hydrate:distilled

water:glycerol (8:2:1, v/v), followed by washes using 90, 70, and 50% (v/v) diluted

ethanol.

Lugol staining

Starch granules in the columella root cap were visualized by the staining using 1%

Lugol’s solution (Nacalai Tesque Inc.) (Willemsen et al., 2003). Seedlings were stained

for 3 min, rinsed with water, and cleared with chloral hydrate. Microscopic images were

collected using an IX71 system (Olympus, Tokyo, Japan).

Sterol analysis

Samples (50 mg) were frozen and homogenized in liquid N2, and total sterol

fractions were extracted for 30 min at room temperature in 5 ml of chloroform/methanol

(1:2, v/v). By adding 2 ml of 1% (w/v) KCl and 1 ml of chloroform, sterols were recovered

in the organic phase. Two milliliters of methanol/H2O (10:9, v/v) were then added to the

organic phase and was evaporated to dryness under N2 stream. The saponification was

done in 2.5 ml of 1 M KOH in methanol at 90oC for 1h, and 2 ml chloroform and 2.5 ml

H2O were added to recover the organic phase. After adding 1.25 ml of 0.5 M KOH and 6

ml H2O, the organic phase was evaporated to dryness in vacuo and used for the

trimethylsilyl (TMS)-derivatization. Sterols were analyzed by gas chromatography-mass

spectrometry (GC-MS) on a Saturn 2200 GC/MS system (Varian Inc. Palo Alto, CA) with

64

a CP-SIL5 CB LOW Bleed/MS column (30 m×0.25 mm) (Varian). The port temperature

for split injections (a split ratio of 50%) was 270°C, and helium was used as the carrier

gas at a flow rate of 1.5 ml/min. The temperature program was started with a slow rise

from 230 to 285°C (2°C/min), and from 285 to 320°C (10°C/min), and finally 10 min at

320°C. Standards consisting of cholesterol, brassicasterol, campesterol, stigmasterol,

and -sitosterol were used for quantification and identification. The peak areas were

automatically calculated using a 2200 Workstation (Varian), and sterol amounts were

determined from the ratio of the peak area of each sterol to that of 5--cholestane as an

internal standard (Schaeffer et al., 2001; Dyer et al., 1995). The sterol structures were

identified by reference to relative retention time and mass spectra. The pattern of

fragment ions with the mass-to-charge (m/z) values of 484, 394, 255, and 129 were

attributed to stigmasterol, and the fragment ions (m/z = 470, 380, 365, and 129) were

used to identify brassicasterol/crinosterol. The 24-epimers (brassicasterol and

crinosterol) were not separately analyzed under our experimental conditions.

For analysis of sterol conjugates, cultured root samples (30 mg) were frozen and

homogenized in liquid N2, and sterols (including free sterol; FS, sterol fatty acid ester;

SE, sterol glucosides; SG and acylated steryl glucosides; ASG) were extracted with

dichloromethane/methanol (1:2, v/v). The FSs and sterol conjugates were separated by

TLC. The TLC plate was developed in solvent system, containing of

dichloromethane/methanol/water, 85:15:0.5 (v/v), FS fraction elutes at Rf 0.82-0.89, SE

fraction elutes at Rf 1.00, SG fraction elutes at 0.33, ASG fractions elutes at 0.61. The

plate was sprayed with 0.01% berberine in ethanol, and examined under UV light. The

separated individual bands were scraped and extracted using

dichloromethane/methanol (2:1). The saponification of SE fraction was done in 2 ml of

65

6% KOH in methanol at 70oC for 2h, and 1.5 ml hexane and 1 ml H2O were added to

recover the organic phase. Hydrolysis of SG and ASG fractions were done in 2 ml 1 %

sulfuric acid in ethanol at 70°C for 2 h and 1.5 ml hexane and 1 ml H2O were added to

recover the organic phase. Acetylation of 3’-OH-sterols in each fractions was done in 50

l toluene, 30 l acetic anhydride and 20 l pyridine at 70°C for 2 h. Sterols were

analyzed by GC-MS.

Chromatography and mass spectrometry of sterol derivatives using

UPLC-APPI+-MS

Plant lipids (including carboxysterols and glucoside conjugates) were extracted

with a mixture of dichloromethane/methanol (2:1, v/v). Samples of 3 l (of a total extract

of 200 l) were analyzed on an Acquity UPLC system (Waters Quattro Premier XE,

Waters, Mildorf, MA, USA) equipped with a diode array detector (DAD) and an

Atmospheric Pressure Photon Ionisation (APPI) source. UV spectra were recorded from

190 to 500 nm. Chromatographic separation was achieved using an Acquity UPLC BEH

C8 column (100 x 2.1 mm, 1.7µm; Waters) coupled to an Acquity UPLC BEH C8

pre-column (2.1 x 5 mm, 1.7µm; Waters). The column was operated at 46°C with a

flow-rate of 0.3 ml/min of the following mobile phases: solvent A (75% methanol in water

with 0.01% formic acid) and solvent B (99.99% isopropanol with 0.01% formic acid). The

separation (total run time of 48 min) started with 100% of solvent A hold for 2 min,

followed by 25 min of a gradient to reach 50% of solvent A and 50% of solvent B, and

then hold for 4 min. It was followed by a gradient to reach 100 % of solvent B in 8 min,

hold for 3 min. Then another 4 min gradient was applied to reach 100% of solvent A hold

for 3 min. The column was equilibrated with solvent A for 4 min before the next run. Both

of the drying gas and nebulising gas were nitrogen generated from pressurized air in an

66

N2G nitrogen generator (Mistral, Schmidlin-dbs-AG, Switzerland). The parameters for

the MS and MS/MS detection and APPI ionization were as follows. Nebulizer gas flow,

50 l/h; desolvation gas flow, 500 l/h; APPI probe temperature, 450°C; source

temperature, 120°C; capillary voltage, 1.5 kV; cone voltage after optimization for each

sterol derivative was set to 25V; positive ionization mode; low and high mass resolution,

15 for both mass analyzers; ion energies 1 and 2 were 0.2 and 1V; entrance and exit

potential, 30V; detector (multiplier) gain, 650V. The Selected Ion Recording (SIR) MS

mode was used to determine mass transitions from the parent ion [M+ H - H2O]:

4-carboxy-4-methyl-cycloartanol (CMC, m/z: 439.50) (Mialoundama et al., 2013),

4-carboxy-4-methyl-24-methylene cycloartanol (CMMC, m/z: 453.48) (Mialoundama et

al., 2013), sitosterol (m/z: 397.71), and sitosteryl glucoside [M+ H - glucose - H2O]: (m/z:

397.70). Measurements of sterol derivatives were based on specific chromatography

retention times and total parent ion counts by comparison with authentic standards.

Data acquisition and analysis were performed with the MassLynx software (version 5.1,

Windows XP professional, Pentium PC).

Protoplasts

Five-day-old seedlings germinated on GM agar plates were transferred to 0.5× MS

liquid medium and cultured for 20 days under 16 h light/8 h dark on a rotary shaker (100

rpm) at 24°C. The roots were cut to 5 to 10 mm in length and incubated for 5 days in a

secondary root-inducing medium consisting of 0.5× MS medium supplemented with 1%

(w/v) sucrose, 3.0 g/ml IBA (indolebutyric acid) 1.0 g/ml NAA (naphthaleneacetic acid)

on a rotary shaker (80 rpm). The secondary roots were induced in the presence or

absence of 3 g/ml phytosterols (-sitosterol, stigmasterol, or campesterol). Protoplasts

were prepared principally according to the methods described previously (Jaideep et al.,

1995; Yoo et al., 2007). Briefly, the induced secondary roots were incubated for 12 h in

67

0.5× MS liquid medium containing 0.4 M glucose, 1% (w/v) CELLULASE ONOZUKA

R-10 (Yakult Pharmaceutical Industry, Tokyo, Japan), 0.25% (w/v) MACEROZYME

R-10 (Yakult). Protoplasts were released from root tissues by incubating for 12 h on a

rotary shaker (80 rpm) and filtered through a 70 m mesh filter. The isolated protoplasts

were washed with 0.4 M mannitol and cultured in a glass- bottomed dish for 100 h in a

0.5× MS protoplast culture medium containing 0.4 M glucose, 2.0 g/ml 2,4-D, 0.5 g/ml

IAA and 6.25 g/ml meropenem hydrate. The protoplasts were stained with 50 g/ml

FM4-64 for the PM visualization. In order to monitor the cellulose accumulation,

protoplasts were stained with 1 g/ml calcofluor-white M2R (Sigma, Tokyo, Japan)

(Sato et al., 2001).

Endocytosis and endocytic recycling

Five-day-old seedlings were stained with 50 g/ml FM4-64 (Cosmo Bio, Tokyo,

Japan ) in 0.5× MS medium for 10 min on ice. After washing with 0.5× MS medium,

specimens were visualized using a confocal microscope (LSM 700). Five-day-old

seedlings were incubated for 30 min in 0.5× MS medium containing 50 mM

cycloheximide (CHX, Nacalai Tesque). Seedling specimens in 0.5× MS were mounted

on glass slides with coverslips and visualized using a confocal microscope (LSM700,

Carl Zeiss). The endocytic recycling inhibitor brefeldin A (BFA) induces the formation of

the BFA compartments consisting of aggregates derived from the vesicles and tubules

from the trans-Golgi network (Blilou et al., 2005; Geldner et al., 2001; Geldner et al.,

2003). After the pretreatment with 50 M CHX in 0.5× MS medium, the seedlings were

stained for 10 min with 50 g/ml FM4-64 in 0.5× MS medium containing 50 M CHX and

50 M BFA (Wako Pure Chemicals, Osaka, Japan). The seedlings were further

incubated in 0.5× MS medium containing 50 M CHX and 50 M BFA for 30 min. The

68

time course of the formation of BFA compartments was observed using a confocal

microscope LSM700 (Carl Zeiss). For the observation of endocytic recycling, the

seedlings were treated for 90 min with 50 M BFA in 0.5× MS medium, and further

incubated in 0.5× MS medium containing 50 M CHX and 50 M BFA for 30 min. The

seedlings were washed three times with 50 M CHX in 0.5× MS medium. The time

course of the PIN2-EGFP endocytic recycling was observed for 3 h using a confocal

microscope LSM 700 (Carl Zeiss) with the excitation at 488 nm for GFP and 555 nm for

FM4-64. Fluorescence emissions of GFP and FM4-64 were detected at 490 to 555 nm

and at 640 to 700 nm, respectively.

Immunolabeling

Whole-mount immunolabeling experiments principally followed the method

described previously (Men et al., 2008). Arabidopsis seedlings were fixed in 4%

paraformaldehyde for 1 h in MTSB (50 mM PIPES, 5 mM MgSO4, 5 mM EGTA, pH 7.0).

After washing 5 times for 5 min with 0.1% Triton X-100 in MTSB, seedlings were

extensively washed 5 times with sterilized H2O. For whole-mount immunolabeling, the

paraformaldehyde-fixed seedlings were treated with 2% (w/v) driselase (Sigma) for 45

min and washed with MTSB 5 times followed by 1 h incubation in MTSB containing 10%

(w/v) DMSO and 3% (w/v) NP-40. After washing with MTSB 5 times, specimens were

incubated in 2% (w/v) BSA in MTSB for 1 h and treated with primary antibodies in MTSB

containing 3% (w/v) BSA at 4°C for overnight. After washing with MTSB for 7 times,

secondary antibodies were applied at 37°C for 1.5 h and the samples were stained with

10 g/ml DAPI in MTSB for 10 min. Antibody dilutions were: rabbit anti-KNOLLE (Lauder

et al., 1997), 1:2000; mouse anti--tubulin IgG (Invitrogen), 1:500; rabbit anti-GFP

69

(Sigma), 1:1000; Alexa Fluor 488 goat anti-mouse IgG (Invitrogen), 1:500; Alexa Fluor

488 goat anti-rabbit IgG (Invitrogen), 1:500; Alexa Fluor 568 goat anti-mouse IgG

(Invitrogen), 1:500; and Alexa Fluor 568 goat anti-rabbit IgG (Invitrogen), 1:500. The

excitation wavelengths of DAPI, Alexa Fluor 488, and Alexa Fluor 568 were 405 nm,

488 nm, and 555 nm, respectively. Fluorescence detections were performed at 395 to

540 nm for DAPI, 490 to 635 nm for Alexa Fluor 488, and 569 to 700 nm for Alexa Flour

568 using a confocal microscope (LSM700) and images were incorporated using the

ZEN 2011 software (Zeiss Efficient Navigation 2011).

Toluidine blue staining.

Arabidopsis flowers and ovules in siliques were incubated in vacuo for 24 h in a

fixation medium (formalin:acetic acid:ethanol:H2O = 1:1:9:9). The specimens were

dehydrated in a graded alcohol series of n-butanol:absolute ethanol:H2O = 1:4:5, 2:5:3,

7:10:3, and 50:49:1. The samples were left for 1 h in each grade. The specimens were

then washed with n-butanol:absolute ethanol (3:1) for 1 h and washed twice with

n-buthanol and embedded in paraffin wax. For microscopic observation, the specimens

were cut into 5 to 10 mm sections using a microtome (MTH-1 NK system, Osaka,

Japan). Cut sections were placed on slides and subjected to deparaffinization with

xylene for 2× 60 min. Then the slides were transferred to a 1:1 mixture of xylene and

absolute ethanol for 15 min followed by a series of graded ethanol 95%, 85%, 70%,

50%, and 30%. The specimens were left for 5 min in each ethanol grade. After washing

with H2O, the specimens were air dried and used for the staining with 0.05% toluidine

blue O (Wako Pure Chemicals, Osaka, Japan). The stained specimens were examined

by the microscopic visualization using an IX71 system (Olympus).

70

Lipid order

The calculation of the GP images obtained from ratiometric di-4-ANEPPDHQ

fluorescence imaging was performed using ImageJ (http://imagej.nih.gov/ij/) following

the procedure described by Owen et al. (2012) and using their custom-written macro.

The ordered (490–580 nm) and disordered (615–750 nm) phase fluorescence channels

were assigned ch00 and ch01, respectively. The threshold value for the analysis was

fixed at 15, the color scale for the output GP images was set to ’16-color’, and no

immunofluorescence mask was selected. The ImageJ macro for GP analysis generates

GP images from ordered and disordered channel images based on the following

equation:

where I represents the intensity in each pixel in the image acquired in the indicated

spectral channel (numbers in nm). To compensate for differences in the efficiency of

collection in the two channels, GP values were corrected using a G factor. In order to

obtain the G factor, the same microscope set-up employed for imaging root samples

was used to image the fluorescence of a drop of undiluted di-4-ANEPPDHQ stock

solution (2 l) at three laser powers (0.3%, 0.5% and 1%). The mean pixel intensities of

the channels ch00 (ordered) and ch01 (disordered) were extracted in ImageJ, and

corresponding GPmes values were calculated (using the previous equation with G = 1).

The G factor was then calculated according to the equation:

GPref is a reference value for di-4-ANEPPDHQ in DMSO, here fixed at -0.85 (Owen et al.,

2012). In this study, the G factor was defined as G =0.68.

71

Table 3. Primer list used in this study

72

References

Alonso, J.M., Stepanova, A.N., Leisse, T.J., Kim, C.J., Chen, H., Shinn, P., Stevenson,

D.K., Zimmerman, J., Barajas, P., Cheuk, R., Gadrinab, C., Heller, C., Jeske, A.,

Koesema, E., Meyers, C.C., Parker, H., Prednis, L., Ansari, Y., Choy, N., Deen, H.,

Geralt, M., Hazari, Formatted,N., Hom, E., Karnes, M., Mulholland, C., Ndubaku, R.,

Schmidt, I., Guzman, P., Aguilar-Henonin, L., Schmid, M., Weigel, D., Carter, D.E.,

Marchand, T., Risseeuw, E., Brogden, D., Zeko, A., Crosby, W., Berry, C.C. and

Ecker, J.R. (2003) Genome-wide insertional mutagenesis of Arabidopsis thaliana.

Science, 301, 653-657.

Ambrose, C., Ruan, Y., Gardiner, J., Tamblyn, L.M., Catching, A., Kirik, V., Marc, J.,

Overall, R. and Wasteneys, G.O. (2013) CLASP interacts with sorting nexin 1 to link

microtubules and auxin transport via PIN2 recycling in Arabidopsis thaliana. Dev.

Cell, 24, 649-659.

Benveniste, P. (2004) Biosynthesis and accumulation of sterols. Annu Rev Plant Biol,

55, 429-457.

Blilou, I., Xu, J., Wildwater, M., Willemsen, V., Paponov, I., Friml, J., Heidstra, R., Aida,

M., Palme, K. and Scheres, B. (2005) The PIN auxin efflux facilitator network

controls growth and patterning in Arabidopsis roots. Nature, 433, 39-44.

Boutté, Y., Frescatada-Rosa, M., Men, S., Chow, C.M., Ebine, K., Gustavsson, A.,

Johansson, L., Ueda, T., Moore, I., Jürgens, G. and Grebe, M. (2010) Endocytosis

restricts Arabidopsis KNOLLE syntaxin to the cell division plane during late

cytokinesis. EMBO J., 29, 546-558.

Carland, F.M., Fujioka, S. and Nelson, T. (2010) The sterol methyltransferases SMT1,

SMT2, and SMT3 influence Arabidopsis development through nonbrassinosteroid

products. Plant Physiol., 153, 741-756.

73

Carland, F.M., Fujioka, S., Takatsuto, S., Yoshida, S. and Nelson, T. (2002) The

identification of CVP1 reveals a role for sterols in vascular patterning. Plant Cell, 14,

2045-2058.

Clough, S.J. and Bent, A.F. (1998) Floral dip: a simplified method for

Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J., 16,

735-743.

Clouse, S.D. (2002) Arabidopsis Mutants Reveal Multiple Roles for Sterols in Plant

Development. Plant Cell, 14, 1995-2000.

Curtis, M.J. and Hays, J.B. (2007) Tolerance of dividing cells to replication stress in

UVB-irradiated Arabidopsis roots: requirements for DNA translation polymerases eta

and zeta. DNA Repair., 6, 1341-1358.

Darnet, S. and Rahier, A. (2004) Plant sterol biosynthesis: identification of two distinct

families of sterol 4alpha-methyl oxidases. Biochem. J., 378, 889-898.

DeBolt, S., Scheible, W.R., Schrick, K., Schaller, H., Beisson, F., Bischoff, V.,

Bouvier-Nave, P., Li, Y., Nair, M., Carroll, A., Hematy, K., Li, Y., Milne, J., Nair, M.,

Schaller, H., Zemla, M. and Somerville, C. (2009) Mutations in UDP

glucose:sterol-glucosyltransferase in Arabidopsis cause transparent testa

phenotype and suberization defect in seeds. Plant Physiol., 151, 78-87.

Derksen, J., Janssen, G.J., Wolters-Arts, M., Lichtscheidl, I., Adlassnig, W., Ovecka, M.,

Doris, F. and Steer, M. (2011) Wall architecture with high porosity is established at

the tip and maintained in growing pollen tubes of Nicotiana tabacum. Plant J., 68,

495–506.

Desmond, E. and Gribaldo, S. (2009) Phylogenomics of Sterol Synthesis: Insights into

the Origin, Evolution, and Diversity of a key Eukaryotic Feature. Genome Biol. Evol.,

1, 364-381.

74

Dhonukshe, P., Tanaka, H., Goh, T., Ebine, K., Mähönen, A.P., Prasad, K., Blilou,

I., Geldner, N., Xu, J., Uemura, T., Chory, J., Ueda, T., Nakano, A., Scheres, B.

and Friml, J. (2008) Generation of cell polarity in plants links endocytosis, auxin

distribution and cell fate decisions. Nature, 456, 962-966.

Diener, A.C., Li, H., Zhou, W., Whoriskey, W.J., Nes, W.D. and Fink, G.R. (2000) Sterol

methyltransferase 1 controls the level of cholesterol in plants. Plant Cell, 12,

853-870.

Dufourc, E.J. (2008) Sterols and membrane dynamics. J. Chem. Biol., 1, 63-77.

Dyer, R.G., Hetherington, C.S., Alberti, K.G.M.M. and Laker, M.F. (1995) Simultaneous

measurement of phytosterols (campesterol and β-sitosterol) and 7-ketocholesterol in

human lipoproteins by capillary column gas chromatography. J. Chromatography B.,

663, 1-7.

Fenner, G.P. and Patterson, G.W. (1992) 24-ethyl-orientation of the 24-ethylsterols

during the life cycle of the squash. Phytochem., 31, 73-75.

Feraru, E. and Friml, J. (2008) PIN polar targeting. Plant Physiol., 147, 1553-1559.

Frescatada-Rosa, M., Stanislas, T., Backues, S.K,, Reichardt, I, Men, S., Boutté, Y.,

Jürgens, G., Moritz, T., Bednarek, S.Y., and Grebe, M. (2014) High lipid order of

Arabidopsis cell-plate membranes mediated by sterol and DYNAMIN-RELATED

PROTEIN1A function. Plant J., 80, 745-757.

Fujioka, S. and Yokota, T. (2003) Biosynthesis and metabolism of brassinosteroids.

Annu. Rev. Plant Biol., 54, 137-164.

Gachotte, D., Eckstein, J., Barbuch, R., Hughes, T., Roberts, C. and Bard, M. (2001) A

novel gene conserved from yeast to humans is involved in sterol biosynthesis. J.

Lipid Res., 42, 150-154.

75

Gasser, C.S., Broadhvest, J. and Hauser, B.A. (1998) Genetic analysis of ovule

development. Annu. Rev. Plant Physiol. Plant Mol. Biol., 49, 1-24.

Geldner, N., Anders, N., Wolters, H., Keicher, J., Kornberger, W., Muller, P., Delbarre,

A., Ueda, T., Nakano, A. and Jürgens, G. (2003) The Arabidopsis GNOM ARF-GEF

mediates endosomal recycling, auxin transport, and auxin-dependent plant growth.

Cell, 112, 219-230.

Geldner, N., Friml, J., Stierhof, Y.D., Jürgens, G. and Palme, K. (2001) Auxin transport

inhibitors block PIN1 cycling and vesicle trafficking. Nature, 413, 425-428.

Grebe, M., Xu, J., Mobius, W., Ueda, T., Nakano, A., Geuze, H.J., Rook, M.B. and

Scheres, B. (2003) Arabidopsis sterol endocytosis involves actin-mediated

trafficking via ARA6-positive early endosomes. Curr. Biol., 13, 1378-1387.

Grosjean, K., Mongrand, S., Beney, L., Simon-Plas, F., and Gerbeau-Pissot, P. (2015)

Differential effect of plant lipids on membrane organization: specificities of

phytosphingolipids and phytosterols. J. Biol. Chem. 290, 5810-5825.

Guo, D., Venkatramesh, M. and Nes, W.D. (1995) Developmental regulation of sterol

biosynthesis in Zea mays. Lipids, 30, 203-219.

Jaideep, M., Csaba, K. and László, S. (1995) A simple method for isolation, liquid

culture, transformation and regeneration of Arabidopsis thaliana protoplasts. Plant

Cell Rep., 14, 221-226.

Jaillais, Y., Fobis-Loisy, I., Miège, C., Rollin, C. and Gaude, T. (2006) AtSNX1 defines

an endosome for auxin-carrier trafficking in Arabidopsis. Nature, 443, 106-109.

Jang, J.C., Fujioka, S., Tasaka, M., Seto, H., Takatsuto, S., Ishii, A., Aida, M., Yoshida,

S. and Sheen, J. (2000) A critical role of sterols in embryonic patterning and

meristem programming revealed by the fackel mutants of Arabidopsis thaliana.

Genes Dev., 14, 1485-1497.

76

Jefferson, R.A., Kavanagh, T.A. and Bevan, M.W. (1987) GUS fusion: -glucuronidase

as a sensitive and versatile gene fusion marker in higher plants. EMBO J., 16,

3901-3907.

Kazłowska, K., Lin, H.T., Chang, S.H. and Tsai, G.J. (2013) In Vitro and In Vivo

Anticancer Effects of Sterol Fraction from Red Algae Porphyra dentata. Evid. Based

Complement. Alternat. Med., 2013, 493869.

Kelley, R.I. and Herman, G.E. (2001) Inborn errors of sterol biosynthesis. Annu Rev

Genomics Hum Genet., 2, 299-341.

Kleine-Vehn, J. and Friml, J. (2008) Polar targeting and endocytic recycling in

auxin-dependent plant development. Annu. Rev. Cell. Dev. Biol., 24, 447-473.

Kleine-Vehn, J., Leitner, J., Zwiewka, M., Sauer, M., Abas, L., Luschnig, C. and Friml, J.

(2008) Differential degradation of PIN2 auxin efflux carrier by retromer-dependent

vacuolar targeting. Proc. Natl. Acad. Sci. U S A, 105, 17812-17817.

Kumagai, F., Yoneda, A., Tomida, T., Sano, T., Nagata, T. and Hasezawa, S. (2001)

Fate of nascent microtubules organized at the M/G1 interface, as visualized by

synchronized tobacco BY-2 cells stably expressing GFP-tubulin: time-sequence

observations of the reorganization of cortical microtubules in living plant cells. Plant

Cell Physiol., 42, 723-732.

Lauber, M.H., Waizenegger, I., Steinmann, T., Schwarz, H., Mayer, U., Hwang, I.,

Lukowitz, W. and Jürgens, G. (1997) The Arabidopsis KNOLLE protein is a

cytokinesis-specific syntaxin. J. Cell Biol., 139, 1485-1493.

Lindsey, K., Pullen, M.L. and Topping, J.F. (2003) Importance of plant sterols in pattern

formation and hormone signalling. Trends Plant Sci., 8, 521-525.

Malinsky, J., Opekarová, M., Grossmann, G. and Tanner, W. (2013) Membrane

microdomains, rafts, and detergent-resistant membranes in plants and fungi. Annu.

Rev. Plant Biol., 64, 501-529.

77

McKean, M. and Nes, W.R. (1977) Evidence for separate intermediates in the

biosynthesis of - -alkylsterols in tracheophytes. Phytochem., 16, 683-686.

Men, S., Boutté, Y., Ikeda, Y., Li, X., Palme, K., Stierhof, Y.D., Hartmann, M.A., Moritz,

T. and Grebe, M. (2008) Sterol-dependent endocytosis mediates post-cytokinetic

acquisition of PIN2 auxin efflux carrier polarity. Nat. Cell Biol., 10, 237-244.

Mialoundama, A.S., Jadid, N., Brunel, J., Di Pascoli, T., Heintz, D., Erhardt, M., Mutterer,

J., Bergdoll, M., Ayoub, D., Van Dorsselaer, A., Rahier, A., Nkeng, P., Geoffroy, P.,

Miesch, M., Camara, B. and Bouvier, F. (2013) Arabidopsis ERG28 Tethers the

Sterol C4-Demethylation Complex to Prevent Accumulation of a Biosynthetic

Intermediate That Interferes with Polar Auxin Transport. Plant Cell, 25, 4879-4893.

Neelakandan, A.K., Song, Z., Wang, J., Richards, M.H., Wu, X., Valliyodan, B., Nguyen,

H.T. and Nes, W.D. (2009) Cloning, functional expression and phylogenetic analysis

of plant sterol 24C-methyltransferases involved in sitosterol biosynthesis.

Phytochemistry, 70, 1982-1998.

Nes, W.R. and McKean, M.L. (1977) Biochemistry of steroids and other isopentenoids.

Baltimore, Md: University Park Press.

Nes, W.R., Krevitz, K., Joseph, J., Nes, W.D., Harris, B., Gibbons, G.F. and Patterson,

G.W. (1977) The phylogenetic distribution of sterols in tracheophytes. Lipids, 12,

511-527.

Nes, W.D., Norton, R.A., Crumley, F.G., Madigan, S.J. and Katz, E.R. (1990) Sterol

phylogenesis and algal evolution. Proc. Natl. Acad. Sci. U S A, 87, 7565-7569.

Nguyen, H.M., Baudet, M., Cuiné, S., Adriano, J.M., Barthe, D., Billon, E., Bruley, C.,

Beisson, F., Peltier, G., Ferro, M. and Li-Beisson, Y. (2011) Proteomic profiling of oil

bodies isolated from the unicellular green microalga Chlamydomonas reinhardtii:

with focus on proteins involved in lipid metabolism. Proteomics., 11, 4266-4273.

78

Owen, D.M., Rentero, C., Magenau, A., Abu-Siniyeh, A., and Gaus, K. (2012)

Quantitative imaging of membrane lipid order in cells and organisms. Nat. Protoc. 7,

24–35.

Paciorek, T., Zazímalová, E., Ruthardt, N., Petrásek, J., Stierhof, Y.D., Kleine-Vehn, J.,

Morris, D.A., Emans, N., Jürgens, G., Geldner, N. and Friml, J. (2005) Auxin inhibits

endocytosis and promotes its own efflux from cells. Nature, 435, 1251-1256.

Petrásek, J. and Friml, J. (2009) Auxin transport routes in plant development.

Development, 136, 2675-2688.

Pullen, M., Clark, N., Zarinkamar, F., Topping, J. and Lindsey, K. (2010) Analysis of

vascular development in the hydra sterol biosynthetic mutants of Arabidopsis. PLoS

One., 5, e12227.

Qian, P., Han, B., Forestier, E., Hu, Z., Gao, N., Lu, W., Schaller, H., Li, J. and Hou, S.

(2013) Sterols are required for cell-fate commitment and maintenance of the

stomatal lineage in Arabidopsis. Plant J., 74, 1029-1044.

Rahier, A., Darnet, S., Bouvier, F., Camara, B. and Bard, M. (2006) Molecular and

enzymatic characterizations of novel bifunctional 3beta-hydroxysteroid

dehydrogenases/C-4 decarboxylases from Arabidopsis thaliana. J. Biol. Chem., 281,

27264-27277.

Ramgopal, M. and Bloch, K. (1983) Sterol synergism in yeast. Proc. Natl. Acad. Sci. U S

A, 80, 712-715.

Rasmussen, C.G., Humphries, J.A. and Smith, L.G. (2010) Determination of symmetric

and asymmetric division planes in plant cells. Annu. Rev. Plant Biol., 62, 387-409.

Saeidnia, S., Permeh, P., Gohari, A.R. and Mashinchian-Moradi, A. (2012)

Gracilariopsis persica from Persian Gulf Contains Bioactive Sterols. Iran J. Pharm.

Res., 11, 845-849.

79

Salimova, E., Boschetti, A., Eichenberger, W. and Lutova, L. (1999) Sterol mutants of

Chlamydomonas reinhardtii: Characterisation of three strains deficient in C24(28)

reductase. Plant Physiol. Biochem., 37, 241-249.

Sano, T., Higaki, T., Oda, Y., Hayashi, T. and Hasezawa, S. (2005) Appearance of actin

microfilament ‘twin peaks’ in mitosis and their function in cell plate formation, as

visualized in tobacco BY-2 cells expressing GFP-fimbrin. Plant J., 44, 595-605.

Sato, S., Kato, T., Kakegawa, K., Ishii, T., Liu, Y.G., Awano, T., Takabe, K., Nishiyama,

Y., Kuga, S., Sato, S., Nakamura, Y., Tabata, S. and Shibata, D. (2001) Role of the

putative membrane-bound endo-1,4-beta-glucanase KORRIGAN in cell elongation

and cellulose synthesis in Arabidopsis thaliana. Plant Cell Physiol., 42, 251-263.

Schaeffer, A., Bronner, R., Benveniste, P. and Schaller, H. (2001) The ratio of

campesterol to sitosterol that modulates growth in Arabidopsis is controlled by

STEROLMETHYLTRANSFERASE 2;1. Plant J., 25, 605-615.

Schaller, H. (2003) The role of plant sterols in plant growth and development. Prog.

Lipid Res., 42, 163-175.

Schaller, H. (2004) New aspects of sterol biosynthesis in growth and development of

higher plants. Plant Physiol. Biochem., 42, 465-476.

Schaller, H. (2010) Sterol and steroid biosynthesis and metabolism in plants and

microorganisms. In Comprehensive Natural Products II Chemistry and Biology,

Mander, L. Lui, H.-W. eds, Elsevier: Oxford Vol. 1, pp 755-787.

Schrick, K., Debolt, S. and Bulone, V. (2012) Deciphering the molecular functions of

sterols in cellulose biosynthesis. Front. Plant Sci., 3, 84.

Schrick, K., Fujioka, S., Takatsuto, S., Stierhof, Y.D., Stransky, H., Yoshida, S. and

Jurgens, G. (2004) A link between sterol biosynthesis, the cell wall, and cellulose in

Arabidopsis. Plant J., 38, 227-243.

80

Schrick, K., Mayer, U., Horrichs, A., Kuhnt, C., Bellini, C., Dangl, J., Schmidt, J. and

Jürgens, G. (2000) FACKEL is a sterol C-14 reductase required for organized cell

division and expansion in Arabidopsis embryogenesis. Genes Dev., 14.1471-1484.

Schrick, K., Mayer, U., Martin, G., Bellini, C., Kuhnt, C., Schmidt, J. and Jürgens, G.

(2002) Interactions between sterol biosynthesis genes in embryonic development of

Arabidopsis. Plant J., 31, 61-73.

Sørensen, I., Rose, J.K., Doyle, J.J., Domozych, D.S. and Willats, W.G. (2012) The

Charophycean green algae as model systems to study plant cell walls and other

evolutionary adaptations that gave rise to land plants. Plant Signal Behav., 7,1-3.

Souter, M., Topping, J., Pullen, M., Friml, J., Palme, K., Hackett, R., Grierson, D. and

Lindsey, K. (2002) hydra mutants of Arabidopsis are defective in sterol profiles and

auxin and ethylene signaling. Plant Cell, 14, 1017-1031.

Souter, M.A., Pullen, M.L., Topping, J.F., Zhang, X. and Lindsey, K. (2004) Rescue of

defective auxin-mediated gene expression and root meristem function by inhibition

of ethylene signalling in sterol biosynthesis mutants of Arabidopsis. Planta., 219,

773-783.

Tamura, K., Stecher, G., Peterson, D., Filipski, A. and Kumar, S. (2013) MEGA6:

Molecular Evolutionary Genetics Analysis Version 6.0. Molecular Biology and

Evolution, 30, 2725-2729.

Timme, R.E., Bachvaroff, T.R. and Delwiche, C.F. (2012) Broad phylogenomic sampling

and the sister lineage of land plants. PLoS One, 7, e29696.

Topping, J.F., May, V.J., Muskett, P.R. and Lindsey, K. (1997) Mutations in the

HYDRA1 gene of Arabidopsis perturb cell shape and disrupt embryonic and

seedling morphogenesis. Development. 124, 4415-4424.

81

Ulmasov, T., Murfett, J., Hagen, G. and Guilfoyle, T.J. (1997) Aux/IAA proteins repress

expression of reporter genes containing natural and highly active synthetic auxin

response elements. Plant Cell, 9, 1963-1971.

Valvekens, D., Montagu, M.V. and Lusebettens, M.V. (1998) Agrobacterium

tumefaciens-mediated transformation of Arabidopsis thaliana root explants by using

kanamycin selection. Proc. Natl. Acad. Sci. U S A, 85, 5536-5540.

Van Damme, D. (2009) Division plane determination during plant somatic cytokinesis.

Curr Opin Plant Biol., 12, 745-751.

Van Damme, D. and Geelen, D. (2008) Demarcation of the cortical division zone in

dividing plant cells. Cell Biol. Int., 32, 178-187.

Willemsen, V., Friml, J., Grebe, M., van den Toorn, A., Palme, K. and Scheres, B.

(2003) Cell polarity and PIN protein positioning in Arabidopsis require STEROL

METHYLTRANSFERASE1 function. Plant Cell, 15, 612-625.

Xu, J. and Scheres, B. (2005) Dissection of Arabidopsis ADP-RIBOSYLATION

FACTOR 1 function in epidermal cell polarity. Plant Cell, 17, 525-536.

Yang, H., Richter, G.L., Wang, X., Młodzińska, E., Carraro, N., Ma, G., Jenness, M.,

Chao, D.Y., Peer, W.A. and Murphy, A.S. (2013) Sterols and sphingolipids

differentially function in trafficking of the Arabidopsis ABCB19 auxin transporter.

Plant J., 74, 37-47.

Yoo, S.D., Cho, Y.H. and Sheen, J. (2007) Arabidopsis mesophyll protoplasts: a

versatile cell system for transient gene expression analysis. Nat. Protocols, 2,

1565-1572.

82

ACNOREDGEMENTS

I am deeply grateful to my supervisor Professor Daisaku Ohta, Osaka Prefecture

University, for his guidance, invaluable discussion, continuous encouragement, and

critical review in all the time of research for and writing of this thesis.

I wish to express my since gratitude to Professor Nozomu Koizumi, Osaka

Prefecture University, and Professor Koh Aoki, Osaka Prefecture University, for their

invaluable advice and suggestions.

I also wish to express my thanks to Professor Hubert Schaller and member of

CNRS, for valuable guidance and stimulating discussion throughout this work.

I also wish to express my thanks to Dr. Masahiko Hurutani, NAIST, for valuable

guidance and stimulating discussion throughout this work.

I am thankful to Dr. Atsushi Okazawa, Osaka Prefecture University, for kind his kind

suggestions and encouragement throughout the work.

I am thankful to Dr. Takumi Ogawa, Osaka Prefecture University, for kind his kind

suggestions and encouragement throughout the work.

I also thank to member of Laboratory of Metabolomics, Osaka Prefecture University,

for their technical assistance and encouragement throughout the work.

Masatoshi Nakamoto