Download EARLY SENESCENCE 1 Encodes a SCAR

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

History of herbalism wikipedia , lookup

Plant tolerance to herbivory wikipedia , lookup

Cultivated plant taxonomy wikipedia , lookup

Arabidopsis thaliana wikipedia , lookup

Historia Plantarum (Theophrastus) wikipedia , lookup

Ornamental bulbous plant wikipedia , lookup

History of botany wikipedia , lookup

Plant use of endophytic fungi in defense wikipedia , lookup

Plant secondary metabolism wikipedia , lookup

Plant defense against herbivory wikipedia , lookup

Hydroponics wikipedia , lookup

Xylem wikipedia , lookup

Venus flytrap wikipedia , lookup

Plant stress measurement wikipedia , lookup

Leaf wikipedia , lookup

Plant physiology wikipedia , lookup

Embryophyte wikipedia , lookup

Plant morphology wikipedia , lookup

Sustainable landscaping wikipedia , lookup

Plant evolutionary developmental biology wikipedia , lookup

Glossary of plant morphology wikipedia , lookup

Transcript
Plant Physiology Preview. Published on August 4, 2015, as DOI:10.1104/pp.15.00991
1
Running Head: ES1 affects water loss
2
3
Corresponding authors:
4
Dali Zeng
5
6
State Key Laboratory of Rice Biology, China National Rice Research Institute, Hangzhou
7
310006, China
8
9
Tel: (+0571)6337 0537
10
Fax: (+0571)6337 0389
11
E-mail: [email protected]
12
13
Research Area: Genes, Development and Evolution
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Copyright 2015 by the American Society of Plant Biologists
14
15
EARLY SENESCENCE 1 Encodes a SCAR-like Protein2 that Affects Water
16
Loss in Rice
17
Yuchun Rao†, Yaolong Yang†, Jie Xu†, Xiaojing Li, Yujia Leng, Liping Dai, Lichao Huang,
18
Guosheng Shao, Deyong Ren, Jiang Hu, Longbiao Guo, Jianwei Pan, Dali Zeng*
19
20
State Key Laboratory of Rice Biology, China National Rice Research Institute, Hangzhou
21
310006, China (Y.R., Y.Y., J.X., Y.L., L.D., L.H., G.S., D.R., J.H., L.G. and D.Z.); College of
22
Chemistry and Life Sciences, Zhejiang Normal University, Jinhua 321004, China (Y.R., X.L
23
and J.P.); Key Laboratory of Crop Physiology, Ecology and Genetic Breeding, Ministry of
24
Education, Jiangxi Agricultural University, Nanchang 330045, China (Y.Y. and J.X.)
25
26
27
One sentence summary:
28
The SCAR/WAVE-like protein2 affects water loss by regulating stomatal density in rice
29
30
Total word count: 10652
31
Total characters (including space): 67804
32
2
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
33
34
35
Footnotes
36
37
1
38
(31201183, 31221004, 31171531, 31171520, and 91435105), the State Key Basic Research
39
Program (2013CBA01403), the Ministry of Agriculture of China for transgenic research
40
(No.2013ZX08009003-001), and the China Postdoctoral Science Foundation
41
(2014M561108).
This work was supported by grants from the National Natural Science Foundation of China
42
43
44
*Corresponding author e-mail: [email protected]
45
46
47
48
1
49
this article in accordance with the policy described in the Instructions for Authors
50
(www.plantphysiol.org) is: Dali Zeng ([email protected]).
51
† These authors contributed equally to this work.
The author responsible for the distribution of materials integral to the findings presented in
52
53
54
55
56
57
58
59
3
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
60
61
62
Abstract
63
The global problem of drought threatens agricultural production and constrains the
64
development of sustainable agricultural practices. In plants, excessive water loss causes
65
drought stress and induces early senescence. In this study, we isolated a rice (Oryza sativa)
66
mutant, designated as early senescence 1 (es1), which exhibits early leaf senescence. The
67
es1-1 leaves undergo water loss at the seedling stage (as reflected by whitening of the leaf
68
margin and wilting) and display early senescence at the three-leaf stage. We used map-based
69
cloning to identify ES1, which encodes a SCAR-like protein 2, a component of the
70
SCAR/WAVE complex involved in actin polymerization and function. The es1-1 mutants
71
exhibited significantly higher stomatal density. This resulted in excessive water loss and
72
accelerated water flow in es1-1, also enhancing the water absorption capacity of the roots and
73
water transport capacity of the stems, as well as promoting in vivo enrichment of metal ions
74
co-transported with water. Expression of ES1 is higher in the leaves and leaf sheaths than in
75
other tissues, consistent with its role in controlling water loss from leaves. GFP-ES1 fusion
76
proteins were ubiquitously distributed in the cytoplasm of plant cells. Collectively, our data
77
suggest that ES1 is important for regulating water loss in rice.
78
79
Key words: ES1; water loss; SCAR; stomatal density
80
81
82
INTRODUCTION
Rice is a major world-wide food crop, but it consumes more water than most crops
83
(Bruce et al., 2015), with water consumption for rice cultivation accounting for
84
approximately 65% of agricultural water usage. Rice provides a staple food for about 3
85
billion people, while using an estimated 24–30% of the world's developed freshwater
86
resources (Bouman et al., 2007). Severe water shortages restrict the expansion of rice
87
production and hinder the irrigation of existing paddy fields (Zhu et al., 2013). One effective
88
way to overcome water shortages is to reduce water loss in rice plants, thus allowing
89
cultivation of this key crop in environments with less water (Nguyen et al., 1997).
90
In plants, the stomata on the leaf surface work as the main channels for discharge of
4
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
91
water and entry of carbon dioxide, and thus strongly affect physiological processes such as
92
transpiration and photosynthesis. Previous studies showed that mutations in some genes
93
could affect stomatal density or differentiation, such as AtTMM (Yang et al., 1995),
94
AtSCRM2 (Kanaoka et al., 2008), AtSDD1(Von Groll et al., 2002), AtYODA2(Bergmann et
95
al.,2004). Stomata show a regular distribution on rice leaves (Huang et al., 2009) during plant
96
growth and development, various environmental factors affect the density and size of
97
stomata, as well as the chlorophyll contents of rice leaves. For example, rice leaves that
98
develop under water stress show substantial fewer stomata compared with leaves that develop
99
under well-watered conditions (Huang et al., 2009). Changes in stomatal density and
100
morphology affect water loss (Boonrueng et al., 2013). In rice, OsSRO1 (Similar to RCD-
101
One) enhances drought tolerance by regulating stomatal closure (You et al., 2013) while the
102
zinc finger protein DST (Drought and Salt Tolerance) functions in drought and salt tolerance
103
by adjusting stomatal aperture (Huang et al., 2009). In Arabidopsis thaliana, overexpression
104
of the Mg-chelatase H subunit in guard cells confers drought tolerance by promoting stomatal
105
closure (Tsuzuki et al., 2013). Thus, stomatal closure and low stomatal density enhance
106
drought tolerance by reducing water loss in plants.
107
The epicuticular wax layer in plants acts as the first barrier to environmental conditions
108
by reducing water loss due to transpiration and preventing plant damage due to excessively
109
strong sunlight (Riederer et al., 2001). Leaf transpiration involves both stomatal and cuticular
110
transpiration. Stomatal conductance controls stomatal transpiration, but the physicochemical
111
properties of the leaf surface mainly control cuticular transpiration. For example, the
112
composition, thickness, and microstructure of the cuticular wax affect water permeability and
113
transport (Buschhaus et al., 2012; Svenningsson, 1988; Xu et al., 1995). OsDWA1 (rice
114
Drought-induced Wax Accumulation 1), OsGL1 (GLOSSY1), and OsWR1 (rice Wax
115
synthesis Regulatory gene 1) affect drought tolerance by regulating deposition or
116
biosynthesis of cuticular wax (Islam et al., 2009; Wang et al., 2012; Zhou et al., 2013; Zhu et
117
al., 2013).
118
Besides, leaf trichomes can also affect water loss (Konrad et al., 2015), and leaf
119
trichomes closely linked with actin cytoskeleton. The involvement of the actin cytoskeleton
120
in controlling directional cell expansion in trichomes has received much attention (Zhang et
121
al., 2005). Generally, genes that affect cytoplasmic organization can be studied by screening
122
leaf trichome mutants (Qiu et al., 2002). In Arabidopsis, a reproducible morphogenetic
123
program directs the polarized development of trichome branches (Le et al., 2006; Mathur et
124
al., 1999; Syzmanski et al., 1999). Some of these genes affect the cytoskeleton and also affect
5
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
125
the morphology of normal plant cells, especially epidermal cells. For example, mutation of
126
SPIKE1 in Arabidopsis causes epidermal cells to show simple arrangement and
127
morphologies, i.e., all cells dividing along a single axis (Qiu et al., 2002).
128
To study the molecular mechanisms underlying water loss in rice, we isolated and
129
characterized the es1-1 (early senescence 1-1) rice mutant, which showed excessive water
130
loss and early senescence phenotypes. Map-based cloning data showed that ES1 encodes a
131
SCAR-like protein 2, and its Arabidopsis thaliana homolog affects the polymerization of
132
actin. The es1-1 mutants showed obvious changes in leaf trichomes, similar to Arabidopsis.
133
However, few studies have reported a connection between the actin cytoskeleton and water
134
loss in Arabidopsis. Our results demonstrated a critical role of the actin cytoskeleton in
135
regulating water loss in rice.
136
RESULTS
137
Identification and Characterization of Early Senescence Mutants
138
To study the mechanisms of senescence in rice, we screened a large pool of mutants
139
generated in the japonica rice cultivar Nipponbare (NPB) background by mutagenesis using
140
ethyl methanesulfonate (EMS). From this pool, we identified two mutants with an early
141
senescence phenotype. The two mutants showed similar phenotypes under the same growth
142
conditions and F1 hybrid individuals produced by the two parental mutants exhibited
143
phenotypes like the parental line with the weaker phenotype (subsequently named es1-1, see
144
below), indicating that the two mutants are allelic (Supplemental Table 1); therefore, we
145
termed these mutants early senescence 1 (es1-1 and es1-2). Under normal growth conditions,
146
es1-1 plants showed severely retarded development. At the seedling stage, the es1-1 mutants
147
displayed whitish and yellowish leaf tips (Figure 1A), a hallmark of early senescence (Li et
148
al., 2014); this phenotype increased with increasing leaf age, becoming more severe at the
149
heading stage (Figure 1B). Newly developed leaves showed yellowing margins that rolled
150
inwards or formed a spiral and old leaves showed spots with a rusty color and water
151
deficiency phenotypes, including wilting (Figures 1A and 1B). Histochemical analysis
152
showed that concentrations of senescence-related substances, including hydrogen peroxide
153
(H2O2), superoxide radical (O2•-), and malondialdehyde (MDA), were higher in es1-1 leaves
154
than in the wild-type leaves (Figures 1C and 1D), indicating that the mutant does show
155
senescence phenotypes (Moradi et al., 2007). Moreover, we measured the expression of two
6
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
156
senescence marker genes, Staygreen (Sgr) and Os185, in NPB and es1-1 (Supplemental
157
Figures 1F and 1G). We found that the transcript levels of these two genes were substantially
158
higher in es1-1 than in NPB, suggesting that es1-1 caused senescence (Park et al., 2007; Lee
159
et al., 2001). We also measured plant height, finding that es1-1 plants were much shorter than
160
the wild-type plants from the seedling to the mature stages (Figures 1A and 1B, Supplemental
161
Table2). At the mature stage, es1-1 mutants showed degraded or white panicles with a low
162
seed setting rate (only 2.4%; Figure 1E, Supplemental Table 2), short internodes (Figure 1F,
163
Supplemental Table2), and brown, open glumes (Figure 1G). Scanning transmission electron
164
microscopy showed that es1-1 mutants had disordered thylakoids, compared with the neat
165
and well-ordered thylakoids observed in wild-type plants; moreover, es1-1 mutants had lower
166
chlorophyll contents than wild type NPB.
167
Compared with es1-1, the es1-2 plants had more tillers (Supplemental Figures 1A and
168
1C), higher plants (Supplemental Figures 1A and 1D), and showed a stronger early
169
senescence phenotype in the leaves (Supplemental Figure 1B). In this study, we mainly
170
focused on es1-1 mutants, unless otherwise specified.
171
172
Map-Based Cloning of ES1
To clarify whether es1-1 phenotypes were caused by a single or multiple gene(s) and
173
dominant or recessive genes, we crossed the mutant plants directly and reciprocally with
174
three wild-type cultivars. All the F1 plants showed no symptoms of early senescence. In the
175
F2 segregating populations of these three crosses, normal and early senescent plants showed a
176
typical segregation ratio of 3:1 (Supplemental Table 3). This result suggested that a single
177
recessive nuclear gene controls the early senescence phenotype of es1-1.
178
For cloning ES1, we crossed the mutant es1-1 (japonica) with a typical indica cultivar
179
(NJ6). Plants with the es1-1 phenotype were identified from the F2 generation and used as the
180
initial mapping population. For primary chromosome mapping, we used 164 pairs of simple
181
sequence repeat (SSR) primers distributed on the 12 chromosomes of rice at a distance of 10–
182
30 cM from each other to detect differences between es1-1 and NJ6. A significant
183
amplification difference was detected on chromosome 1 using the SSR marker M1, showing
184
linkage with a small population in the F2. The M1 marker was further used to analyze 320
185
mutants in the F2 population, which detected 15 recombinants. Genetic linkage analysis
186
mapped the M1 marker to a distance of 2.3 cM from ES1 (Figure 2A).
7
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
187
For fine mapping of ES1, the mutation in ES1 must be flanked by two molecular
188
markers. Thus, we designed multiple pairs of primers at positions relatively far from M1
189
(Supplemental Table 4). The marker M6 also showed polymorphisms between es1-1 and NJ6.
190
Screening of the primary mapping population (320 mutants) identified 11 recombinants that
191
differed from those identified using M1 (Figure 2A). Linkage analysis mapped the M6
192
marker to a distance of 1.8 cM from ES1. Next, we used M1 and M6 to screen the entire
193
population (2850 plants), which identified 50 and 39 recombinants, respectively (Figure 2A).
194
To further narrow the ES1 region, we developed a large number of sequence-tagged site
195
(STS) and cleaved amplified polymorphic sequence (CAPS) markers between M3 and M4.
196
Primers C1–C9 showed good polymorphisms between es1-1 and TN1. Thus, we used these
197
primers to analyze the recombinants obtained by preliminary screening and to construct fine
198
genetic and physical maps of ES1 (Figure 2B; Supplemental Table 4). The number of
199
recombinants detected by C1 and C10 decreased to 28 and 35, respectively. Similarly, the
200
number of recombinants screened by C2 and C9 decreased to 10 and 12, respectively, and
201
those detected by C3 and C8 further decreased to three and four, respectively. The
202
recombinants detected by C3 and C8 corresponded to different individual plants.
203
We also designed primers for the region between C3 and C8 to further screen the
204
recombinants. Finally, these new primers detected only one recombinant and thus these
205
markers tightly co-segregate with ES1 (Figure 2B). The physical distance between the end
206
markers C5 and C6 was 13.1 kb, which included two predicted open reading frames.
207
Sequencing analysis detected a unique mutation in LOC-Os01g11040 within the 13.1 kb
208
region in es1-1 compared with the wild-type sequence. This single-base mutation replaced G
209
with A in the first nucleotide of the third intron (nucleotide 2275; Figure 2C). To test whether
210
this mutation alters the normal splicing of the third intron, we generated es1-1 and wild-type
211
genomic cDNAs using reverse transcription. Sequencing showed that the es1-1 cDNA was 62
212
bases longer than the wild-type cDNA (Figure 2D). This increase in length coincided with the
213
length of the third intron, which contains a stop codon that results in the premature
214
termination of protein translation (Figure 2E). Similarly, sequencing analysis showed that the
215
mutation in es1-2 was a single-base substitution (G to A) at nucleotide 4660. This mutation
216
changed the tryptophan (Trp) codon into a stop codon, thus resulting in the premature
217
termination of protein translation (Figures 2E).
218
219
We then used genetic complementation to verify the identity of ES1. A 10.6-kb fragment
containing ES1 (including the upstream 2-kb and downstream 1-kb sequences) was
8
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
220
introduced into es1-1 and es1-2 mutants, and 52 and 36 transgenic plants, respectively, were
221
recovered. Exogenous ES1 rescued the phenotypes of es1-1 and es1-2 (Figure 2F), and the
222
complementation reversed the mutant phenotype of es1-1. Taken together, these data
223
demonstrate that LOC-Os01g11040 (Os01g0208600) is ES1.
224
ES1 Encodes a SCAR-Like Protein of the Arp2/3-SCAR Pathway
225
We performed a BLAST search to identify ES1 homologs in more than 20 different
226
plant species. ES1 showed high amino acid sequence similarity to proteins in other species
227
such as Oryza brachyantha (85%), Brachypodium distachyon (68%), maize (60%), and
228
Setaria italica (64%) (Figure 3E). ES1 also shares high amino acid sequence similarity to
229
proteins in both monocots and dicots, suggesting that ES1 and its homologs may perform
230
similar functions. Examination of ES1 homologs in Arabidopsis showed that ES1
231
(Os01g0208600) encodes a SCAR-like protein 2; its closest homolog in Arabidopsis forms
232
part of the actin-related 2/3 (ARP2/3) complex, which functions in regulating actin filament
233
polymerization and has been examined in detail in Arabidopsis. ARP2/3 alone is inactive in
234
Arabidopsis. A domain formed by the SCAR-like family proteins, and members of other
235
protein families (e.g., SPA1 and ABI families) are responsible for binding to and activation of
236
the Arp2/3 complex (Zhang et al., 2005). In Arabidopsis, mutation of the gene encoding
237
SCAR-like protein 2 leads to changes in leaf trichomes and stomata (Basu et al., 2005). In
238
our study, we observed substantial changes in leaf trichomes of es1-1, mainly reflected by the
239
serious degradation of the trichome tip (Figures 3A and 3B). The leaves of es1-1 had smooth
240
leaf margins without a jagged pattern (Figure 3A).
241
To examine the possible role of ES1 in regulating actin filaments, we examined the
242
organization of F-actin in the roots of wild type and es1-1 mutants by Alexa Fluor 488-
243
phalloidin staining. The actin filaments were well organized in wild type (Figure 3C). The
244
actin filaments in wild-type were complete, orderly arranged, while es1-1 showed irregular
245
arrangement. Moreover, the actin filaments in es1-1 were shorter than that in wild type, as
246
indicated by arrows (Figure 3D), which indicated that ES1 likely functions in the Arp2/3-
247
SCAR pathway in rice and that a mutation in ES1 affects the actin cytoskeleton in rice cells
248
(Dyachok et al., 2011).
9
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
249
250
Expression of ES1 and Subcellular Localization of ES1
To monitor the tissue-specific expression of ES1, we generated an ES1:GUS (β-
251
glucuronidase) reporter construct containing the 2.0-kb genomic region upstream of the start
252
codon of ES1. We detected high GUS activity in leaves, leaf sheaths, and root tips, and also
253
detected lower GUS activity in stamens and vascular bundles (Figures 4A–4E). Reverse-
254
transcription quantitative polymerase chain reaction (RT-qPCR) assays showed that ES1 was
255
ubiquitously expressed in various tissues at the tillering and heading stages, including the
256
roots, culms, leaves, sheaths, and spikelets. We observed the highest levels of ES1 expression
257
in the leaf and leaf sheath (Figure 4F). Consistent with the RT-qPCR data, ES1:GUS
258
expression was higher in rice leaves and leaf sheaths than in other plant tissues.
259
To investigate the subcellular localization of ES1, we constructed a construct to express
260
eGFP-ES1 (35S::eGFP-ES1 DNA) and introduced it into onion and tobacco epidermal cells.
261
Fluorescence signals were detected in both cell types, and the signals surrounded the
262
cytoplasm, nucleus, and cell membrane (Figures 4G, 4H, 4I, and 4J). Consistent results were
263
obtained by examining protoplasts extracted from tobacco leaves after the expression of the
264
fusion protein (Figures 4K and 4L). These results indicated that ES1 was ubiquitously
265
distributed in the cytoplasm of plant cells.
266
ES1 Affects the Density of Leaf Stomata in Rice
267
Scanning electron microscopy of es1-1 leaves showed that these leaves had smoother
268
surfaces than the leaves of wild-type plants. In addition, these leaves showed substantial
269
smaller siliceous protrusions at the seedling stage and substantially more stomata per unit leaf
270
area compared with wild-type leaves. The number of semi-open stomata in es1-1 was also
271
larger than that in wild-type (Figures 5A and 5B). Moreover, the tips of trichomes and the
272
waxy layer were severely degraded on the leaf surface of es1-1 mutants (Figure 5A).
273
Statistical data showed that, regardless of the position, stomatal density was significantly
274
higher in es1-1 than that in the wild-type. While both stomatal conductance and transpiration
275
rate were also detected, they were also significantly higher in es1-1, these data indirectly
276
suggested stomatal density in es1-1 was higher (Figure 5B). The maximum and minimum
277
stomatal apertures were greater in the wild-type compared with es1-1(Figure 5C). At the
278
seedling stage, the transcriptional level of several genes which could affect the increase of
10
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
279
stomatal density were analyzed by the real-time PCR, including OsTMM, OsSCRM,
280
OsSDD1, OsYODA2 (Du et al., 2014). The expression level of OsTMM, OsSCRM and
281
OsSDD1 were significantly higher in es1-1 than that in the wild-type, and ES1 gene also
282
showed higher expression level in es1-1(Figure 5E).
283
Leaves of wild-type and es1-1 at seedling stage were sprayed with 300 umol / L ABA,
284
after 3 hours, quantitative analysis of OsTMM, OsSCRM, OsSDD1, OsYODA2 was taken. In
285
the wild-type, the expression level of these genes almost didn’t change; but in es1-1, the
286
majority of these genes had shown a significantly declined expression level, this indicated in
287
es1-1 those genes which could affect the increase of stomatal density were sensitive to ABA.
288
We also detected the rate of vitro water loss (RWL) of seedling leaves from the same position
289
of the wild-type and es1-1 before and after ABA treatment. The results showed that, no
290
matter with or without ABA treatment, the vitro RWL value was always higher in es1-1 than
291
that in the wild-type. The vitro RWL value in the wild-type was reduced after ABA
292
treatment, and the value in es1-1 was also reduced after ABA treatment, but it was not so
293
obvious compared with the wild-type (Figure 5D). SEM assay on the leaves of wild-type and
294
es1-1 with ABA treatment showed that nearly all stomata of both wild-type and es1-1 were
295
closed. Moreover, we also examined the expression level of multiple ABA response-related
296
genes (ABA signaling genes) in wild-type and es1-1 before and after ABA treatment (Shang
297
et al., 2010), and discovered that after ABA treatment, OsDREB2A gene was up-regulated,
298
while OsABI4 gene was down-regulated. These results suggest that both wild-type and es1-1
299
were sensitive to ABA (Figure 5E).
300
301
302
The es1-1 Plants Showed Higher Water Loss Rates than Wild Type
In vitro, the rate of water loss of es1-1 was higher than that of NPB at the seedling
303
stage (Figure 6A), as the detached leaves of es1-1 shrunk into a line at 30 min, but the NPB
304
leaves remained nearly normal in width (Figure 6A). To confirm this phenomenon, we also
305
measured water loss per unit of time (g cm-2.10-3) at the tillering stage in vivo and in vitro
306
(Figures 6B and 6C), which gave similar results: the water loss per unit of time in vitro
307
(detached leaves) in different periods was uniformly higher in es1-1 than in NPB (Figure 6B).
308
The water loss per unit of time in vivo was also uniformly higher in es1-1 than in NPB at 4, 8,
309
and 14 days after tillering (Figure 6C).
310
To verify excessive water loss in es1-1, we performed a guttation experiment, where we
11
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
311
looked for the exudation of drops of xylem sap on the leaf periphery in plants grown at
312
different humidities. Hydroponic seedlings were cultured in light, in an incubator at different
313
humidities, from 30% and 90%, and the es1-1 and wild-type plants were examined for
314
guttation. At the seedling stage (two-leaf), both the wild-type plants and es1-1 were capable
315
of guttation at 30% humidity. However, higher humidity was required for guttation with plant
316
growth. Until the strong seedling stage (two-tiller), guttation was observed in wild-type
317
plants but not in es1-1 (Supplemental Figure 2). This result indicated that water loss occurred
318
more rapidly in es1-1 mutants than in the wild-type plants.
319
320
Water Absorption and Transport Capacity was Increased in es1-1
321
Mutants
322
It is unknown whether excessive water loss through leaves affects water absorption in
323
the root. To address this, we measured root traits in es1-1 and wild-type plants (Figures 7A,
324
7B, 7C, and 7D). The es1-1 plants had smaller primary and total root lengths (Figures 7E and
325
7F) but higher average root diameters, total root volumes, and total numbers of fibrous roots,
326
compared with the wild-type plants (Figures 7G, 7H, 7I, and 7J). As fibrous roots are a major
327
tissue for water adsorption in rice (Smith et al., 2012), an increase in fibrous root number
328
indicated an increase in the water absorption capacity of the roots in es1-1 mutants.
329
330
Examination of cross sections of the culm and primary root from es1-1 and NPB showed
331
that the es1-1 stems had 3–4 additional vascular bundles (Figures 8A, 8B, and 8C) and the
332
es1-1 primary roots had 2 to 3 additional conduits, compared with NPB (Figures 8D, 8E, and
333
8F). Vascular bundles and conduits are necessary for water transport (Kim et al., 2014).
334
Moreover, the water potential values were all higher in es1-1 than that in wild-type in various
335
tissues, including roots, stems, leaves, and sheath, which indicated that the ability to transport
336
water was stronger in es1-1 than in WT (Figure 8G) (Silva et al., 2011). Therefore, our results
337
indicate that es1-1 mutants have an enhanced ability to transport water.
338
339
340
Excessive Water Loss Causes in Situ Enrichment of Minerals in es1-1
In es1-1 mutants, excessive water loss through the leaves enhances the water absorption
12
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
341
capacity of the roots and the water transportation capacity of the roots and culms. This in turn
342
accelerates water circulation in the plant, thus leading to predicted in situ enrichment of metal
343
ions cotransported with water (Kim et al., 2014). To test this hypothesis in this mutant, we
344
determined the mineral concentrations of different tissues of es1-1 and wild-type plants at the
345
mature stage. Consistent with the hypothesis, the concentrations of all four minerals tested
346
(Cu, Fe, Mn, and Zn) in the examined tissues were much higher in es1-1, especially in the
347
roots and new leaves, than in wild-type plants (Table 1). The manganese (Mn) concentration
348
in es1-1 leaves (719 mg kg-1) was seven times higher than that in wild-type plants (100 mg
349
kg-1). The concentrations of the four minerals were also higher in es1-1 seeds than in NPB
350
(Table 1). These results suggested that mutation in ES1 promoted the accumulation of
351
minerals in plants. The substantial accumulation of metal ions in es1-1 plants was likely due
352
to excessive water loss and higher transpiration pull in the upper part of the plant.
353
DISCUSSION
354
In this study, we characterized es1-1, an early senescence mutant that experienced
355
excessive water loss. Map-based cloning showed that ES1 encodes a SCAR-like protein 2, a
356
component of the SCAR/WAVE complex that is involved in actin polymerization. This
357
suggests that a mutation in ES1 affects polymerization of actin and causes an abnormality in
358
the cytoskeleton that promotes a higher stomatal density (Huang et al., 2000; Xiao et al.,
359
2003; Yu et al., 2001; Zhang et al., 2012), thus leading to excessive water loss.
360
Under stress conditions, plants develop a wax layer on the leaf surface or moderately roll
361
their leaves to reduce transpiration, thereby protecting themselves from excessive water loss
362
(Hu et al., 2012; Jäger et al., 2014). Although stomata account for only 1–2% of the leaf area,
363
approximately 90% of water loss occurs through stomata (Buckley et al., 2005). Under
364
drought conditions, rice cultivars that can maintain relatively high water potential and low
365
transpiration rate in their leaves have strong drought resistance (Luo et al., 2001). These
366
drought-resistant cultivars commonly retain water in the plant body by closing their stomata.
367
Small stomatal aperture and large diffusion resistance can jointly reduce water loss due to
368
transpiration, thus avoiding the adverse effects of drought. Grill and Ziegler (1998) proposed
369
that controlling the stomata is the most important way to improve the efficiency of water use
370
in plants. In the present study, scanning electron microscopy showed that the es1-1 mutants
371
had substantially more stomata than wild-type plants. The expression level of OsTMM,
372
OsSCRM and OsSDD1 were significantly higher in es1-1 than that in the wild-type (Figure
13
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
373
5E), which was closely related to stomatal density. Stomatal density was significantly higher
374
in es1-1 than that in the wild-type. While both stomatal conductance and transpiration were
375
also significantly higher in es1-1 than that in the wild-type (Figure 5B), these data suggested
376
that es1-1 definitely lose water more quickly than wild-type. Nearly all stomata of both wild-
377
type and es1-1 were closed after ABA treatment, and the vitro RWL value in the wild-type
378
was reduced after ABA treatment, the value in es1-1 was also reduced after ABA treatment,
379
but it was not so obvious compared with the wild-type (Figure 5D). Moreover, the wax layer
380
on the leaf surface of es1-1 mutants was considerably degraded. So, we thought the excessive
381
water loss in es1-1 mainly due to increased stomatal density, meanwhile the degradation of
382
waxy layer and leaves hair may aggravate the rate of water loss.
383
Excessive water loss in the leaves inevitably depletes water from the body of es1-1
384
plants. Thus, more water needs to be absorbed through the roots to meet the metabolic
385
requirements of es1-1 plants. In this study, we analyzed the roots of es1-1 and found that the
386
total volume of root and the number of fibrous roots increased compared to wild type
387
(Figures 7I and 7J), indicating that the water absorption capacity of the roots was enhanced in
388
es1-1. Cross-sections of the primary root and stems also showed that the number of vascular
389
bundles in the stems increased by four (Figure 8C) and those in the roots increased by two to
390
three in es1-1 compared to the wild-type plants (Figure 8F). Moreover, the water potential
391
values were all higher in es1-1 than in wild-type in various tissues (Figure 8G). These
392
observations confirmed that the water transport capacity of the roots was enhanced in es1-1.
393
All the above changes inevitably accelerate water circulation and cause minerals
394
commonly cotransported with water to accumulate in the various tissues of es1-1 plants.
395
Indeed, in different tissues, the concentrations of four minerals were higher in es1-1 than in
396
NPB (Table 1). For example, the Cu concentration in es1-1 (446.3 mg kg-1) was 18-times
397
higher than in wild-type plants (23.5 mg kg-1), which could cause Cu overload in es1-1
398
plants. Enrichment of low-mobility metal elements (e.g., Cu, Fe, Mn, and Zn) in various
399
tissues of es1-1 may in turn accelerate the onset of senescence. Moreover, the higher
400
concentrations of mineral nutrients in seeds from the es1-1 mutants might produce more-
401
nutritious food; however, the accumulation of toxic elements such as arsenic might produce
402
food that could adversely affect human health.
403
In our study, we found that actin filaments were shorter in es1-1, and the continuity was
404
poor compared with wild-type (Figures 3C and 3D). Actin filaments function in many
405
endomembrane processes, such as endocytosis of plasma membrane, vacuole formation,
406
plasma membrane protrusion in crawling cells, and vesicle transport from the Golgi complex
14
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
407
(Svitkina et al., 1999; Eitzen et al., 2002; Kaksonen et al., 2003; Mathur et al., 2003; Chen et
408
al., 2004). The ARP2/3 complex functions as a key regulator of actin filament nucleation in
409
plants (Zhang et al., 2008). The ARP2/3 is inactive by itself, as actin polymerization requires
410
activation by the SCAR/WAVE complex, formed by proteins of the SCAR family. The
411
SCAR/WAVE complex is highly conserved in plants. A deficiency of the SCAR/WAVE
412
complex often leads to morphological changes in leaf trichomes and epidermal cells. This
413
phenomenon has been reported in several studies in Arabidopsis (Li et al., 2003; Mathur et
414
al., 2003; Saedler et al., 2004) but not in rice. In our study, es1-1 showed evident
415
morphological changes in leaf trichomes (Figures 3A, 3B, and 5C), confirming that the
416
function of the SCAR/WAVE complex is conserved in different species. Additional
417
components of the SCAR/WAVE complex can be identified by screening for mutants that
418
affect the leaf trichomes of the es1-1 mutant.
419
Actin polymerization occurs in the cytoplasm. In the present study, the fluorescent
420
signal of eGFP-ES1 was detected in the cytoplasm and plasma membrane of tobacco and
421
onion epidermal cells, a localization consistent with the predicted function of ES1 in affecting
422
the actin cytoskeleton. After fusing the ES1 promoter with the GUS reporter gene, GUS
423
staining was observed in various tissues, including the roots, root hairs, stems, leaves, leaf
424
sheath, and vascular bundles, which are associated with water absorption, transportation, and
425
loss. The ES1 expression level was relatively high in the leaves and leaf sheaths—the two
426
major tissues involved in water loss. All these results are consistent with the proposed
427
function of ES1.
428
The increased number of stomata would increase water loss and induce drought stress
429
in es1-1 plants. Drought stress affects photosynthetic capacity, respiration, membrane
430
permeability, osmotic adjustment, enzyme activity, and hormone levels in rice leaves, thus
431
reducing leaf area, leaf area index, and dry weight. Drought stress also prevents the
432
accumulation of dry matter and accelerates senescence (Hu et al., 2010; Ding et al., 2014;
433
Sellammal et al., 2014). In the present study, es1-1 plants showed early senescence
434
characterized by yellowish leaf color and wilting of leaf margins at the seedling stage.
435
Transmission electron microscopy showed that the volume percentage of chloroplasts in the
436
es1-1 cells was lower than that in wild-type cells; however, the number of chloroplasts
437
(irregular and fine prolate) was higher in es1-1 cells than in wild-type cells (Figure 1H). This
438
might have caused the decline in chlorophyll content in es1-1, causing the leaves to turn
439
yellowish. In es1-1, the chloroplasts contained disorganized thylakoids (Figure 1H) as
15
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
440
opposed to the well-organized thylakoids in wild-type plants. The disordered thylakoids may
441
have reduced the capacity to absorb light energy, which in turn would decrease the
442
photosynthetic rate, thus causing early senescence in plant leaves.
443
Early senescence in rice (especially in functional leaves) is extremely unfavorable for
444
rice growth and production because of its serious effects on yield and quality (Zhu et al.,
445
2012). Therefore, rice breeders generally avoid lines showing early senescence. Similarly, the
446
stay-green trait in rice also does not improve yield; although the stay-green trait promotes the
447
ability of source tissues to fix carbon, it does not necessarily enhance the contents of sink
448
tissues in plants. For example, a study on a stay-green mutant of rice (sgr[t]) reported that
449
late photosynthesis and yield traits (seed-setting rate and thousand-grain weight) were not
450
enhanced in sgr[t] mutants (Cha et al., 2002). In the present study, the early senescing mutant
451
es1-1 showed other phenotypes such as decreased number of tillers and substantially reduced
452
seed-setting rate. Our data provide new insights into the molecular mechanisms underlying
453
early senescence in plants, which will help genetically improve rice yield.
454
MATERIALS AND METHODS
455
Plant Growth and Sampling
456
The es1-1 and es1-2 mutants were isolated from the japonica variety Nipponbare that
457
was subjected to EMS treatment. After multiple generations of inbreeding, the phenotypes of
458
these two mutants were stably inherited, and homozygous mutant strains were obtained. The
459
homozygous mutant es1-1 was directly and reciprocally crossed with two typical japonica
460
cultivars (Zhonghua 11 and Chunjiang 06) and one typical indica rice cultivar (Nanjing 6)
461
providing a relatively clear genetic background to construct the genetic population. The
462
obtained seeds of the F1 generation were sown and transplanted as individual plants to
463
generate the F2 generation by inbreeding. The numbers of individuals showing wild-type and
464
early senescence phenotypes were counted at the seedling stage. The segregation ratio was
465
calculated using statistical methods. Genetic analysis was performed on es1-1, and recessive
466
individuals were retrieved for preliminary mapping of the mutation. Plant materials were bred
467
in experimental field of the China National Rice Research Institute, Hangzhou, China.
16
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
468
469
Detection of Reactive Oxygen Species and Malondialdehyde
H2O2 and O2•- were detected using nitroblue tetrazolium (NBT) and diaminobenzidine
470
(DAB), respectively, according to (Li et al., 2010; Wi et al., 2010) with some modifications.
471
Leaves of 3-week-old plants that were grown in growth chambers were incubated in 0.05%
472
NBT (Duchefa) or 0.1% DAB (Sigma) at room temperature in darkness with gentle shaking
473
for 12 hrs. Chlorophyll was cleared by treating with 90% ethanol at 80°C. The concentration
474
of malondialdehyde (MDA) was measured according to the method described by Moradi and
475
Ismail (2007) with minor modifications (Moradi et al., 2007).
476
Determination of Water Potential
477
Ten seeds of NPB and es1-1 were selected, germinated for 3 days, then cultured in an
478
incubator at 28°C for 60 days. The WP4-T Dewpoint PotentiaMeter was preheated for 30
479
min until the temperature remained stable. Various tissues such as leaf, sheath, culm, and root
480
of NPB and es1-1 were harvested from the same position on each plant and cut into short
481
pieces. After making sure the sample temperature was lower than room temperature, the
482
samples were put in the measuring apparatus, which showed the water potential value of the
483
sample. Every sample measurement had six independent repeats, and the weight of the
484
samples was equal.
485
Map-Based Cloning of ES1
486
The mutation in es1-1 was mapped in the F2 population of a cross between es1-1 and the
487
indica variety NJ6. To finely map the ES1 locus, new molecular markers were developed
488
using public databases (Supplemental Table 4). Further, ES1 was sequenced using specific
489
primers (Supplemental Table 4).
490
491
492
Fluorescence Microscopy of Actin Filaments
For F-actin observation, plants were grown in a culture chamber with 16/8 hrs
493
photoperiod at 30°C. Seedling roots were used to observe actin organization. F-actin was
494
stained using the previously described method (Olyslaegers and Verbelen, 1998; Zhang et al.,
17
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
495
2011). Seven-day-old roots were fixed in PEM buffer (100 mM PIPES, 10 mM EGTA, 0.3 M
496
mannitol, 5 mM MgSO4, pH 6.9) with 3% paraformaldehyde for 30 min and washed 3 times
497
with PEM. After washing, roots were incubated in PEM buffer with 0.66 mM Alexa Fluor
498
488-phalloidin (Life Technologies). After overnight incubation at 4°C, samples were washed
499
three times with PEM buffer and then observed in 50% glycerol with a confocal microscope
500
(Olympus FLUOVIEW FV1000).
501
502
503
Phylogenetic Analyses
A BLAST search of the SCAR-like 2 amino acid sequence was performed
504
against the database (http://www.ebi.ac.uk/Tools/sss/ncbiblast/). The phylogenetic tree was
505
constructed based on the amino acid sequences. Maximum likelihood (ML) phylogenetic
506
analysis was performed by MEGA6.0 (Tamura et al., 2013). ML analysis of the SCAR-like 2
507
family was based on the amino acid sequences using the Jones-Taylor-Thornton model,
508
which was chosen after model testing. Support for the ML trees was evaluated by 1000
509
bootstrap replicates.
510
511
512
Plasmid Construction and Plant Transformation
The cDNA sequence of ES1 was amplified using the primers cES1-F and cES1-R (Table
513
S3). This cDNA was fused in frame with eGFP in pCambia1300 (http://www.cambia.org)
514
with minor modifications to generate the transient expression vector 35S::eGFP:cES1. The
515
35S::eGFP-cES1 plasmids were transferred into Agrobacterium tumefaciens strain EHA105
516
by electroporation; the cells of rice ES1, Nicotiana tabacum, and onion were transformed
517
according to a previously described method (Hiei et al., 1994; Lv, 2010).
518
For the molecular complementation assay, a 10.6-kb genomic fragment comprising the
519
ES1 promoter region, coding sequence, and 3′ untranslated region was PCR amplified from
520
the genomic DNA of wild-type plants using pES-F and pES-R primers (Table S3). The PCR
521
amplification product was cloned into pCAMBIA1300 to generate the ES1:ES1 construct.
522
Briefly, the pCambia1300::ES1 vector was first digested with the restriction enzymes PstI
523
and XbaI. Next, the fragment containing ES1 was digested with restriction enzymes Sse83871
18
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
524
and NheI. All the restriction enzymes used above are isocaudomers. The pCambia1300::ES1
525
plasmids were transformed into A. tumefaciens strain EHA105 by electroporation. The es1-1
526
and es1-2 plants were transformed according to a published method (Hiei et al., 1994).
527
To generate the ES1:GUS fusion construct, a 2.0-kb upstream genomic region of the
528
ES1 start codon was amplified with the primers ES1-gus-F and ES1-gus-R (Table S3). The
529
amplified fragment was linked with the reporter GUS gene and together inserted in the
530
pCambia1305 vector (http://www.cambia.org).
531
GUS Staining
532
Samples (proES1::GUS) were stained with a solution containing 50 mM NaPO4 buffer, 1
533
mM 5-bromo-4-chloro-3-indolyl-β-d-glucuronic acid, 0.4 mM K3Fe(CN)6 , 0.4 mM
534
K4Fe(CN)6, and 0.1% (v/v) Triton X-100. The samples were then incubated at 37°C in the
535
dark for 10 hrs. Chlorophyll was removed with 70% ethanol for observation.
536
RNA Extraction and RT-qPCR
537
Total RNA was extracted from the roots, culms, leaves, sheaths, and spikelets of wild-
538
type and es1-1 plants at the seedling, tillering and heading stages using an RNeasy Plant Mini
539
Kit (Qiagen). Reverse transcription was handled with the ReverTra Ace qPCR-RT Kit
540
(TOYOBO), the primers are described in Supplemental Table 4. RT-qPCR analysis was
541
conducted with the StepOnePlus System (ABI) using THUNDERBIRD qPCR Mix
542
(TOYOBO), reaction procedure was 95°C 30 s; 95°C 5 s, 55°C 10 s, 72°C 15s for 40 cycles.
543
OsActin was used as a control. The 2-ΔΔCt method was used for quantification (Livak et al.,
544
2001).
545
Microscopy
546
In different periods, especially at the noon when the sunlight was the strongest (and thus
547
the stomata are likely to be closed), fresh leaf specimens were collected from the same
548
position ( 1/4, 1/2 and 3/4 of leaf blade ) in the mutant es1-1 and the control Nipponbare,
549
from plants that showed consistent growth. The specimens were cut into approximately 1 × 2
19
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
550
551
mm pieces and processed as follows for microscopic examination.
Dual fixation: The specimens were fixed in 2.5% glutaraldehyde solution, vacuumed in a
552
vacuum apparatus for approximately 20–30 min (which allowed the specimens to be
553
immersed in the fixative as completely as possible), and kept at 4°C overnight. Thereafter,
554
the fixative was discarded, and the specimens were rinsed with 0.1 M phosphate-buffered
555
saline (PBS), pH 7.0. The specimens were rinsed 3–4 times (15 min each). The specimens
556
were then fixed in 1% osmium tetroxide for at least 1 h. The fixative was discarded, and the
557
specimens were rinsed with 0.1 M PBS three times (15 min each) as described above.
558
Dehydration: The specimens were first dehydrated with a gradient of ethanol (50%, 70%,
559
80%, 90%, and 95%; 15 min to 20 min at each gradient), processed in anhydrous ethanol for
560
20 min, and treated in pure acetone for 20 min.
561
After dehydration, the specimens were immersed in a mixture of ethanol and isoamyl
562
acetate esters (volume ratio 1:1) for 30 min and were then treated with pure isoamyl acetate
563
for 1-2 h. The specimens were dried, coated, and examined using a Hitachi TM-1000
564
scanning electron microscope.
565
For transmission electron microscopy of the leaves and cross-sectional examination of
566
the leaves and roots, the specimens were pretreated as described above. After dehydration,
567
the specimens were treated with a mixture of embedding agent and pure acetone at a volume
568
ratio of 1:1 for 1 h, followed by treatment with a mixture of embedding agent and pure
569
acetone at a volume ratio of 3:1 for 3 h and pure embedding agent overnight. The specimens
570
were then embedded in small boxes and heated at 70°C for approximately 9 h. The obtained
571
embedded samples were sliced into 70-90 nm sections using a Reichert ultramicrotome and
572
stained with lead citrate solution and saturated uranyl acetate solution in 50% ethanol for 15
573
min each. The sections were examined and photographed using a Hitachi H-7650 type
574
transmission electron microscope.
575
Determination of Minerals
576
Wild-type and es1-1 plants were cultured in complete hydroponic nutrient solution up
577
to the heading stage. Tissue specimens were taken from the roots, stems, old leaves, leaf
578
sheaths, and new leaves. The weight of fresh specimens was more than 10 g for the different
20
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
579
tissues. The specimens were oven dried at 80°C for 3 days to achieve a constant weight. After
580
grinding, the dry samples (0.25 g for es1-1 and NPB, respectively) were digested with a
581
mixture of concentrated nitric acid and perchloric acid (volume ratio 1:3). The volume of the
582
digestion solution was adjusted to 25 ml with single-distilled water. The treated samples were
583
then used for elemental analysis. Concentrations of elements were determined using a full-
584
spectrum direct reading inductively coupled plasma-atomic emission spectroscopy (IRIS
585
Intrepid II XSP ICP-OES PE-2100; Thermo Electron Co., Ltd, USA; Shao et al., 2008).
586
Detection of the Rate of Water Loss
587
In vitro detection: The same leaf parts were selected from NPB and es1-1 plants at the
588
seedling stage. These leaf parts were left at room temperature and observed and photographed
589
every 10 min to examine their dehydration. At the tillering stage, the same leaf parts of NPB
590
and es1-1 were quickly weighed (W1), and the total area (S) was measured. Moreover, the
591
weights of the leaf blade (W2, W3, and W4) were also measured every 1 h. The formula
592
(W1-W2) S-1 was used to calculate the rate of water loss. Accordingly, the rate of water loss
593
at different time points was obtained.
594
In vivo detection: After accelerating germination, the wild-type and es1-1 seeds were
595
cultured in fine sand and watered twice a day (in the morning and evening). At the 1–2-leaf
596
stage, rice seedlings were washed off and wrapped in a long sponge, then placed in a closed
597
bottle with clean water. At the 3-leaf stage, clean water was replaced by complete nutrient
598
solution. After the seedlings developed new roots, all the bottles were filled with a sufficient
599
volume of nutrient solution, and the level was marked for subsequent measurement. The
600
leaves were weighed on days 1 (W1), 4 (W2), 8 (W3), and 14 (W4). The corresponding leaf
601
areas were also measured (S2, S3, and S4). Water loss per unit of time in 4 days was
602
calculated as (W1 − W2) S2-1. Six plants were used for every time interval measurement (the
603
results are presented as mean values). Analogously, (W1 − W3) S3-1 was the amount of water
604
loss per unit of time in 8 days, and (W1 − W4) S4-1 was the amount of water loss per unit of
605
time in 14 days.
21
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
606
607
Measurement of Root Traits
Hydroponic experiments were conducted using the wild-type and es1-1 plants. The seeds
608
were soaked for two days, during which time the water was changed once. Next, the seeds
609
were transferred into a germination box to accelerate germination. Three days later, seeds
610
with radicals breaking through the testa were transferred into fine sand for cultivation. After
611
incubation for 7 days, the seedlings were washed carefully to avoid damage to the roots. The
612
plants were transferred into a black bucket containing tap water. After the transition, the
613
plants were supplied with complete nutrient solution. The seedlings entered the two-leaf stage
614
after three days. At the 5-leaf stage, seedlings of the mutant and wild-type plants (10 each)
615
were taken out for root measurements. The hydroponic seedlings were rinsed and placed in a
616
transparent rectangular box containing single-distilled water and loaded onto a scanner
617
(ScanWizard Pro: ScanMarker 1000XL) for transparent scanning. The obtained images were
618
loaded into Win-RHIZO. Pro. V. 2002c for data analysis, and the results were exported to
619
MS Excel 2011 (Microsoft, USA) for analysis.
620
22
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
621
622
Accession Numbers
623
Sequence data from this article are available at the NBCI website
624
(http:www.ncbi.nlm.nih.gov) with the accession numbers: LOC_Os01g11040
625
(NP_001172227.1), OB01G1689 (XP_006643892.1), Bradi2g06610(XP_003565325.1),
626
Oryza_sativa_indica (OsI_00845), Setaria italica (Si000056m.g), Hordeum vulgare var.
627
distichum (M0XEC2), Aegilops tauschii (F775_07293), Triticum aestivum (W5CZS7), Zea
628
mays (ZEAMMB73_679274), Triticum urartu (TRIUR3_31093), Musa acuminata subsp.
629
Malaccensis (M0UB73), Medicago truncatula (MTR_8g086300), Theobroma cacao
630
(TCM_029698), Phaseolus vulgaris (PHAVU_002G238500g), Jatropha curcas
631
(JCGZ_21566), Glycine max (I1K3Z5), Populus trichocarpa (POPTR_0004s05600g),
632
Prunus persica (PRUPE_ppa000443mg), Citrus sinensis (CISIN_1g001053mg), Citrus
633
clementina (CICLE_v10014081mg), Eucalyptus grandis (EUGRSUZ_D00489), Solanum
634
lycopersicum (Solyc02g076840.2), Erythranthe guttata (MIMGU_mgv1a000504mg), Vitis
635
vinifera (VIT_10s0003g01270), Ricinus communis (RCOM_0850090), Amborella trichopoda
636
(AMTR_s00021p00061660), Morus notabilis (L484_002198), Arabidopsis lyrata
637
subsp.(Lyrata ARALYDRAFT_903078).
638
639
Supplemental Data
640
Supplemental Table 1. Allelic analysis of es1-1 and es1-2.
641
Supplemental Table 2: Phenotypic data of NPB and es1-1.
642
Supplemental Table 3. Reciprocal cross experiments between es1-1 and three crop strains
643
(ZH11, CJ06, and NJ6).
644
Supplemental Table 4. Primers used for PCR in this study.
645
Supplemental Figure 1. Phenotype of allelic mutants.
646
A, wild-type; bar = 10 cm. B, allelic mutants; bar = 10 cm (left, es1-1; right, es1-2). C,
647
Leaves of allelic mutants; bar = 5 cm (left, es1-1; right, es1-2). D, Difference in tiller number
648
between es1-1 and es1-2. E, Difference of plant height between es1-1 and es1-2. F-G,
649
Expression level of SGR and esl85 between NPB and es1-1. The results are given as means of
650
three independent assays, and error bars indicate SD. *, ** significant at the level of 5% and
651
1%, respectively.
652
Supplemental Figure 2. Comparison of guttation between wild-type and es1-1 plants. NPB
653
leaf showing guttation drops at the point of emergence (adaxial side).
23
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
654
655
ACKNOWLEDGEMENTS
656
We sincerely thank Prof.Jianru Zuo (Institute of Genetics and Developmental Biology,
657
Chinese Academy of Sciences) for sharing unpublished data. This work was supported by
658
grants from the National Natural Science Foundation of China (31201183, 31221004,
659
31171531, 31171520, 91435105) and the State Key Basic Research Program
660
(2013CBA01403), the Ministry of Agriculture of China for transgenic research
661
(No.2013ZX08009003-001), China Postdoctoral Science Foundation (2014M561108).
662
663
LITERATURE CITED
664
665
Basu D, Le J, El-Essal SE, Huang S , Zhang C, Mallery EL, Koliantz G, Staiger CJ,
666
Szymanski DB (2005) DISTORTED3/SCAR2 is a putative Arabidopsis WAVE complex
667
subunit that activates the Arp2/3 complex and is required for epidermal morphogenesis.
668
Plant Cell 17: 502-524
669
Boonrueng N, Anuntalabhochai S, Jampeetong A (2013) Morphological and anatomical
670
assessment of KDML 105 (Oryza sativa L. spp. indica) and its mutants induced by low-
671
energy ion beam. Rice Sci 20(3): 213-219
672
673
674
Bouman B, Humphreys E, Tuong T, Barker R (2007) Rice and water. Adv Agron 92:187237
Bruce A, Merle M, Maria A, Rufus L, L.Lanier N, Eliete F, Chris V (2014) Reducing
675
greenhouse gas emissions, water use, and grain arsenic levels in rice systems. Global
676
change biol 21:407-417
677
Buckley TN (2005) The control of stomata by water balance. New Phytol 168:275-292
678
Buschhaus C, Jetter R (2012) Composition and physiological function of the wax layers
679
coating Arabidopsis leaves: β-Amyrin negatively affects the intracuticular water barrier.
680
Plant Physiol 160: 1120-1129
681
Cha K-W, Lee Y-J, Koh H-J, Lee B-M, Nam Y-W, Paek N-C (2002) Isolation,
682
characterization, and mapping of the stay green mutant in rice. Theor Appl Genet 104:
683
526-532
684
Chen J, Lacomis L, Erdjument-Bromage H, Tempst P, Stamnes M (2004) Cytosol-
685
derived proteins are sufficient for Arp2/3 recruitment and ARF/coatomer-dependent actin
686
polymerization on Golgi membranes. FEBS Lett 566: 281–286
24
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
687
Ding L, Li Y, Li Y, Shen Q, Guo S (2014) Effects of drought stress on photosynthesis and
688
water status of rice leaves. Chin J Rice Sci 28: 65-70 (in Chinese with English abstract)
689
Du H, ChangY, Huang F, Xiong L (2014) GID1 modulates stomatal response and
690
submergence tolerance involving abscisic acid and gibberellic acid signaling in rice. J
691
Integr Plant Biol doi: 10.1111/jipb.12313
692
Dyachok J, Zhu L, Liao F, He J, Huq E, Blancaflor E (2011) SCAR mediates light-
693
induced root elongation in Arabidopsis through photoreceptors and proteasomes. Plant
694
Cell 23: 3610-3626
695
696
697
Eitzen G, Wang L, Thorngren N, Wickner W (2002) Remodeling of organelle-bound actin
is required for yeast vacuole fusion. J Cell Biol 158: 669–679
Jäger K, Fábián A, Eitel G, Szabó L, Deák C, Barnabás B, Papp I (2014) A morpho-
698
physiological approach differentiates bread wheat cultivars of contrasting tolerance under
699
cyclic water stress. Plant Physiol 171: 1256-1266
700
Grill E, Ziegler H (1998) A plant’s dilemma. Science 282, 252 –253.
701
Hiei Y, Ohta S, Komari T, Kumashiro T (1994) Efficient transformation of rice (Oryza
702
sativa L.) mediated by Agrobacterium and sequence analysis of the boundaries of the T-
703
DNA. Plant J 6: 271-282
704
Hu L, Wang Z, Huang B (2010) Diffusion limitations and metabolic factors associated with
705
inhibition and recovery of photosynthesis from drought stress in a C3 perennial grass
706
species. Physiol Plant 139: 93-106
707
708
Huang R, Wang X, Lou C (2000) Cytoskeletal inhibitors suppress the stomatal opening of
Vicia faba L. Induced by fusicoccin and IAA. Plant Sci 156:65-71
709
Huang X, Chao D, Gao J, Zhu M, Shi M, Lin H (2009) A previously unknown zinc finger
710
protein, DST, regulates drought and salt tolerance in rice via stomatal aperture control.
711
Genes Dev 23: 1805-1817
712
Islam MA, Du H, Ning J, Ye H, Xiong L (2009) Characterization of Glossy1-homologous
713
genes in rice involved in leaf wax accumulation and drought resistance. Plant Mol Biol
714
70: 443-456
715
716
717
718
Kaksonen M, Sun Y, Drubin DG (2003) A pathway for association of receptors, adaptors,
and actin during endocytic internalization. Cell 115: 475–487
Kanaoka M, Pillitteri L, Fujii H, Yoshida Y, Bogenschutz N, Takabayashi J, Zhu J,
Torii K (2008) SCREAM/ICE1 and SCREAM2 specify three cell-state transitional steps
25
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
719
720
721
722
leading to Arabidopsis stomatal differentiation. Plant Cell 20: 1775-1785
Kim H, Park J, Hwang (2014) Investigating water transport through the xylem network in
vascular plants. J Exp Bot 65:1895-1904
Konrad W, Burkhardt J, Ebner M, Roth-Nebelsick R (2015) Leaf pubescence as a
723
possibility to increase water use efficiency by promoting condensation. Ecohydrol 8:480-
724
492
725
Le J, Mallery EL, Zhang C, Brankle S, Szymanski DB (2006) Arabidopsis
726
BRICK1/HSPC300 is an essential WAVE-Complex subunit that selectively stabilizes the
727
Arp2/3 activator SCAR2. Curr Biol 16: 895-901
728
729
730
Lee R, Wang C, Huang L, Chen S (2001) Leaf senescence in rice plants: cloning and
characterization of senescence up-regulated genes. J Exp Bot 52: 1117-1121
Li J, Pandeya D, Nath K, Zulfugarov IS, Yoo S-C, Zhang H, Yoo J-H, Cho S-H., Koh
731
H-J., Kim D-S, Seo HS, Kang B-C, Lee C-H, Paek N-C (2010) ZEBRA-NECROSIS, a
732
thylakoid-bound protein, is critical for the photoprotection of developing chloroplasts
733
during early leaf development. Plant J 62: 713-725
734
735
736
Li S, Blanchoin L, Yang Z, Lord EM (2003) The putative Arabidopsis Arp2/3 complex
controls leaf cell morphogenesis. Plant Physiol 132: 2034–2044
Li Z, Zhang Y, Liu L, Liu Q, Bi Z, Yu N, Cheng S, Cao L (2014) Fine mapping of the
737
lesion mimic and early senescence 1(lmes1) in rice (Oryza sativa). Plant Physiol Biochem
738
80: 300-307
739
740
741
742
Livak K, Schmittgen T (2001) Analysis of relative gene expression data using real-time
quantitative PCR and the 2-ΔΔCt method. Methods 25: 402-408.
Luo L, Zhang Q (2001) The status and strategy on drought resistance of rice (Oryza sativa
L.) Chin J Rice Sci 15: 209-214 (in Chinese with English abstract)
743
Lv F (2010) Optimization for Agrobacterium Tumefaciens-mediated transformation system
744
for exogenous genes of tobacco. J Anhui Agri Sci 38: 10065-10066 (in Chinese with
745
English abstract)
746
Mathur J, Spielhofer P, Kost B, Chua N-H (1999) The actin cytoskeleton is required to
747
elaborate and maintain spatial patterning during trichome cell morphogenesis in
748
Arabidopsis thaliana. Development 126: 5559–5568
749
Mathur J, Mathur N, Kernebeck B, Hülskamp M (2003) Mutations in actin-related
26
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
750
proteins 2 and 3 affect cell shape development in Arabidopsis. Plant Cell 15: 1632–1645
751
Moradi F, Iamail A.(2007)Responses of photosynthesis, chlorophyII fluorescence and
752
ROS-scavenging systems to salt stress during seedling and reproductive stages in rice.
753
Ann Bot 99:1161-1173
754
755
756
Nguyen HT, Babu RC, Blum A (1997) Breeding for drought resistance in rice: physiology
and molecular genetics considerations. Crop Sci 37: 1426-1434
Park S, Yu J, Park J, Li J, Yoo S, Lee N, Lee S, Jeong S, Seo H, Koh H, Jeon J, Park Y,
757
Paek N (2007) The senescence-induced staygreen protein regulates chlorophyII
758
degradation. Plant Cell 19: 1649-1664
759
760
761
762
763
Qiu J-L, Jilk R, Marks MD, Szymanski DB (2002) The Arabidopsis SPIKE1 gene is
required for normal cell shape control and tissue development. Plant Cell 14: 101-118
Riederer M, Schreiber (2001) Protecting against water loss: analysis of the barrier
properties of plant cuticles. J Exp Bot 52: 2023-2032
Saedler R, Mathur N, Srinivas BP, Kernebeck B, Hülskamp M, Mathur J (2004) Actin
764
control over microtubules suggested by DISTORTED2 encoding the Arabidopsis ARPC2
765
subunit homolog. Plant Cell Physiol 45: 813–822
766
Sellammal R, Robin S, Raveendran M (2014) Association and heritability studies for
767
drought resistance under varied moisture stress regimes in backcross inbred population of
768
rice. Rice Sci 21: 150-161.
769
Shang Y, Yan L, Liu Z, Cao Z, Mei C, Xin Q, Wu F, Wang X, Du S, Jiang T, Zhang X,
770
Zhao R, Sun H, Liu R, Yu Y, Zhang D (2010) The Mg-chelatase H subunit of
771
Arabidopsis antagonizes a group of WRKY transcription repressors to relieve ABA-
772
responsive genes of inhibition. Plant Cell 22: 1909-1935
773
774
Shao G, Chen M, Wang D, Xu C, Mou R, Cao Z, Zhang X (2008) Iron manure regulation
of cadmium accumulation in rice. Chin Sci Bull (C): Life Sci 38: 180-187
775
Silva D, Favreau R, Auzmendi I, Dejong T(2011) Linking water stress effects on carbon
776
partitioning by introducing a xylem circuit into L-PEACH. Ann Bot 108:1135-1145
777
Smith S, Smet I (2012) Root system architecture: insights from Arabidopsis and cereal
778
779
crops. Phil.Trans.R.Soc.B 367:1441-1452
Svenningsson M (1988) Epi- and intracuticular lipids and cuticular transpiration rates of
780
primary leaves of eight barley (Hordeum vulgare) cultivars. Plant Physiol 73: 512-517
781
Svitkina TM, Borisy GG (1999) Arp2/3 complex and actin depolymerizing factor/cofilin in
27
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
782
dendritic organization and treadmilling of actin filament array in lamellipodia. J Cell Biol
783
145: 1009–1026
784
785
786
787
Szymanski DB, Marks MD, Wick SM (1999) Organized F-Actin is essential for normal
trichome morphogenesis in Arabidopsis. Plant Cell 11: 2331–2347
Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013) MEGA6:Molecular
evolutionary genetics analysis version6.0. Mol Bio Evol 30:2725-2729
788
Tsuzuki T, Takahashi K, Tomiyama M, Inoue S, Kinoshita T (2013) Overexpression of
789
the Mg-chelatase H subunit in guard cells confers drought tolerance via promotion of
790
stomatal closure in Arabidopsis thaliana. Front Plant Sci 4: 440
791
792
793
Von G, Berger D, Altmann T (2002) The subtilisin-like serine protease SDD1 mediates
cell-to-cell signaling during Arabidopsis stomatal development. Plant Cell 14: 1527-1539
Wang Y, Wan L, Zhang L, Zhang Z, Zhang H, Quan R, Zhou S, Huang R (2012) An
794
ethylene response factor OsWR1 responsive to drought stress transcriptionally activates
795
wax synthesis related genes and increases wax production in rice. Plant Mol Biol 78: 275-
796
288
797
Wi SJ, Jang SJ, Park KY (2010) Inhibition of biphasic ethylene production enhances
798
tolerance to abiotic stress by reducing the accumulation of reactive oxygen species in
799
Nicotiana tabacum. Mol Cells 30: 37-49
800
801
Xiao Y, Chen Y, Wang X (2003) Filaments skeleton in guard cell. Chin Bull Bot 20: 489494
802
Xu H, Gauthier L, Gossenlin A (1995) Stomatal and cuticular transpiration of greenhouse
803
tomato plants in response to high solution electrical conductivity and low soil water
804
content. J Amer Soc Hort Sci 120: 417-422
805
806
Yang M, Sack F (1995) The too many mouths and four lips mutations affect stomatal
production in Arabidopsis. Plant Cell 7: 2227-2239
807
You J, Zong W, Li X, Ning J, Hu H, Li X, Xiao J, Xiong L (2013) The SNAC1-targeted
808
gene OsSRO1c modulates stomatal closure and oxidative stress tolerance by regulating
809
hydrogen peroxide in rice. J Exp Bot 64: 569-583
810
811
Yu R, Huang R, Wang X, Yuan M (2001) Microtubule dynamics are involved in stomatal
movement of Vicia faba L. Protoplasma 216:113-118
812
Zhang C, Mallery EL, Schlueter J, Huang S, Fan Y, Brankle S, Staiger CJ, Szymanski
813
DB (2008) Arabidopsis SCARs function interchangeably to meet actin-related protein 2/3
814
activation thresholds during morphogenesis. Plant Cell 20: 995-1011
28
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
815
Zhang Q, Lin F, Mao T, Nie J, Yan M, Yuan M, Zhang W (2012)
Phosphatidic Acid
816
Regulates Microtubule Organization by Interacting with MAP65-1 in Response to Salt
817
Stress in Arabidopsis. Plant Cell 24:4555-4576
818
Zhang X, Dyachok J, Krishnakumar S, Smith LG, Oppenheimer DG (2005)
819
IRREGULAR TRICHOME BRANCH1 in Arabidopsis encodes a plant homolog of the
820
actin-related protein2/3 complex activator Scar/WAVE that regulates actin and
821
microtubule organization. Plant Cell 17: 2314-2326
822
Zhou L, Ni E, Yang J, Zhou H, Liang H, Li J, Jiang D, Wang Z, Liu Z, Zhuang C
823
(2013) Rice OsGL1-6 is involved in leaf cuticular wax accumulation and drought
824
resistance. PLoS One 8: e65139 1-12
825
826
827
Zhu LF, Yu SM, Jin QY (2012) Effects of aerated irrigation on leaf senescence at late
growth stage and grain yield of rice. Rice Sci 19(1): 44-48.
Zhu X, Xiong L (2013) Putative megaenzyme DWA1 plays essential roles in drought
828
resistance by regulating stress-induced wax deposition in rice. Proc Natl Acad Sci USA
829
110: 17790-17795
830
831
832
29
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
833
834
Table 1. Comparison of mineral contents between wild-type and es1-1 plants.
Cu(mg kg-1)
Fe(mg kg-1)
Mn(mg kg-1)
23.5±3.83
3757.7±122.55
17.75±1.75
NPB
root
446.3±18.52**
18155.3±326.27**
58.97±3.88**
es1-1
8.61±1.21
35.52±3.7
32.19±1.3
NPB
culm
20.18±4.7**
46.05±3.54**
147.99±10.06**
es1-1
6.59±1.15
136.2±8.59
92.331±5.9
NPB
shealth
16.88±2.43**
240.2±14.65**
582.2±13.22**
es1-1
9.32±2.1
228.8±15.1
100.01±7.1
old
NPB
leaf
18.27±3.88**
358.3±21.94**
719.05±34.8**
es1-1
8.02±1.16
102.1±6.78
27.51±3.32
young
NPB
leaf
17.07±2.91**
223.1±11.4**
175.81±12.24**
es1-1
4.26±0.64
4.05±0.35
31.62±2.86
NPB
seed
5.91±0.72**
6.98±1.06**
50.36±4.1**
es1-1
835 The results are given as mean ± SD of three independent assays.
836
Zn(mg kg-1)
28.49±1.12
91.26±5.88**
28.75±1.81
85.83±6.35**
42.52±3.4
65.36±7.92**
29.89±2.56
64.04±7.5**
30.48±3.44
62.22±4.28**
24.23±2.6
53.13±8.12**
*, ** significant at the level of 5% and 1%, respectively
837
838
Figures and Legends
839
Table 1. Comparison of mineral contents between wild-type and es1-1 plants.
840
Figure 1. Phenotypes of wild-type and es1-1 plants.
841
A, Plants in the seedling stage. Wild-type (NPB, left) and es1-1 (right) plants; bar = 3 cm. B,
842
Plants at the heading stage. Wild-type (NPB, left) and es1-1 (right) plants; bar = 10 cm. C,
843
Nitroblue tetrazolium (NBT) staining (up: NPB, left; es1-1, right); Diaminobenzidine (DAB)
844
staining (down: NPB, left; es1-1, right). D, Malondialdehyde (MDA) contents at the seedling
845
stage. (NPB, left; es1-1, right). E, Panicle rachis (NPB, left; es1-1, right); bar = 5 cm. F, Leaf
846
at different parts (NPB, left; es1-1, right); from left to right, internode length of the first to
847
top, internode length of the second to top, internode length of third to top, and internode
848
length of fourth to top; bar = 3 cm. G, Grain size (NPB, up; es1-1, down); bar = 1 mm. H,
849
Chloroplasts, observed by transmission electron microscopy; (NPB, up; es1-1, down); bar =
850
0.5 μm.
851
852
Figure 2. Map-based cloning of ES1, and phenotype of es1-1 complementation transgenic
853
line.
30
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
854
A, Location of es1-1 on rice chromosome 1. B, Coarse linkage map of es1-1. C, es1-1 gene
855
structure. D, Reverse transcription of es1-1. E, Schematic diagram of es1-1 and mutated
856
proteins encoded by es1-1 and es1-2. F, Phenotype of complementation transgenic lines; es1-
857
1 plant (left), Wild-type (NPB, middle) and transgenic line (right), bar = 10 cm.
858
859
Figure 3. ES1 is a SCAR homolog.
860
A, Leaf margin (left, NPB; right, es1-1). B, Trichome phenotype (macro hairs; upper, NPB;
861
down, es1-1). C, D, F-actin organization was visualized with Alexa Fluor 488-phalloidin in
862
root cells of wild type and es1. Over 40 roots were used in this analysis. Bar =5 μm (C, NPB;
863
D, es1-1). Arrows indicate the continuity of F-actin. E, Phylogenetic tree of SCAR-like 2
864
protein in several species.
865
866
Figure 4. Expression of ES1 and subcellular localization of ES1.
867
A–E, GUS staining of leaf (A), leaf sheath (B), fibrous root (C), stamen (D), and culm
868
vascular bundles (E). F, Quantitative real-time PCR analysis of ES1 expression in different
869
tissues at the tillering stage (T) and heading (H) stage. The results are given as means of three
870
independent assays, and each error bar represents the percent standard deviation of the mean.
871
G–L, GFP fluorescence visualized by confocal microscopy. Transient expression of eGFP
872
(G) and eGFP–ES1 cDNA (H) in epidermal cells of onion. 35S::eGFP::ES1 cDNA and
873
35S::GFP were transiently expressed in onion epidermal cells. Transient expression of eGFP
874
in epidermal cells (I) and protoplasts (K) and expression of eGFP–ES1 in epidermal cells (J)
875
and protoplasts (L) of Nicotiana benthamiana leaves. 35S::eGFP::ES1 and 35S::GFP were
876
transiently expressed in N. benthamiana leaves.
877
878
Figure 5. Observation of stomata between NPB and es1-1.
879
A, Stomata density of the wild-type (left) and es1-1 (right) leaves. Bar = 30 µm. the red
880
triangle indicate the position of stomata. B, Statistical analysis of stomatal parameters, such
881
as stomata density, partially-open stomata, stomatal conductance, transpiration rate. C,
882
Morphological characteristics (left) and maximum opening size (right) of stomata between
883
wild-type (up) and es1-1 (down) plants. Bar = 5 µm D, Comparison of water loss rate
884
between wild-type and es1-1 plants with and without 300um ABA treatment at the 3-leaf-
885
stage. E, qRT-PCR analysis of stomata density related genes (OsTMM, OsSCRM, OsSDD1,
886
OsYODA2), ES1, and some ABA response-related genes (OsDREB1A, OsDREB2A,
887
OsMYB2,OsABI4, OsABI5, OsABF4) in WT and es1-1 leaves before and after 300um ABA
31
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
888
treatment at the 3-leaf-stage. The values were normalized to the Actin1 gene. ** indicates
889
significance (t-test) at the P<0.01. Different letters indicate significant difference at the 1%
890
level (Duncan multiple range test)
891
892
893
Figure 6. Observation of water loss rate between wild-type and es1-1 plants.
894
A, Water loss rate of detached leaves at the seedling stage (left, NPB; right, es1-1). B, Water
895
loss rate of detached leaves at the tillering stage. C, Water loss per unit of time of living
896
leaves between NPB and es1-1 plants at the tillering stage. Error bars indicate SD (n = 15).
897
Asterisks indicate the significance of differences between NPB and es1-1 plants, as
898
determined by t-test. *0.01 ≤ P < 0.05, **P < 0.01.
899
900
Figure 7. Phenotype comparison of the roots between NPB and es1-1 plants.
901
A, Root scanning of NPB plants; bar = 5 cm. B, Root scanning of es1-1 plants; bar = 5 cm.
902
C, Enlarged figure of roots in NPB plants; bar = 5 cm. D, Enlarged figure of root in es1-1
903
plants; bar = 5 cm. E–J, Statistical analysis of the main root length (E), total root length (F),
904
total area of the root (G), average diameter of the root (H), total volume of the root (I), total
905
number of fibrous roots (J) of NPB and es1-1 plants. The results are given as means of three
906
independent assays, and error bars indicate SD. Asterisks indicate the significance of
907
differences between NPB and es1-1 plants, as determined by t-test. *0.01 ≤ P < 0.05, **P <
908
0.01.
909
910
Figure 8. Number of vascular bundles and conduits in wild-type and es1-1 plants.
911
A, Number of vascular bundles in NPB plants. B, Number of vascular bundles in es1-1
912
plants. C, Comparison of the number of vascular bundles between NPB and es1-1 plants. D,
913
Number of conduits in NPB plants. E, Number of conduits in es1-1 plants. F, Comparison of
914
the number of conduits between NPB and es1-1 plants. G, Detection of the water potential
915
value in WT and es1-1. Relative water potential value in different tissues at seedling stage of
916
WT and es1-1 analyzed by WP4-T Dewpoint PotentiaMeter. The results are given as means
917
of three independent assays, and error bars indicate SD. *, ** significant at the level of 5%
918
and 1%, respectively.
919
32
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 1. Phenotypes of wild-type and es1-1 plants.
A, Plants in the seedling stage. Wild-type (NPB, left) and es1-1 (right) plants; bar = 3 cm. B,
Plants at the heading stage. Wild-type (NPB, left) and es1-1 (right) plants; bar = 10 cm. C,
Nitroblue tetrazolium (NBT) staining (up: NPB, left; es1-1, right); Diaminobenzidine (DAB)
staining (down: NPB, left; es1-1, right). D, Malondialdehyde (MDA) contents at the seedling
stage. (NPB, left; es1-1, right). E, Panicle rachis (NPB, left; es1-1, right); bar = 5 cm. F, Leaf
at different parts (NPB, left; es1-1, right); from left to right, internode length of the first to top,
internode length of the second to top, internode length of third to top, and internode length of
fourth to top; bar = 3 cm. G, Grain size (NPB, up; es1-1, down); bar = 1 mm. H, Chloroplasts,
observed by transmission electron microscopy; (NPB, up; es1-1, down); bar = 0.5 μm.
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 2. Map-based cloning of ES1, and phenotype of es1-1 complementation transgenic
line.
A, Location of es1-1 on rice chromosome 1. B, Coarse linkage map of es1-1. C, es1-1 gene
structure. D, Reverse transcription of es1-1. E, Schematic diagram of es1-1 and mutated
proteins encoded by es1-1 and es1-2. F, Phenotype of complementation transgenic lines;
es1-1 plant (left), Wild-type (NPB, middle) and transgenic line (right), bar = 10 cm.
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 3. ES1 is a SCAR homolog.
A, Leaf margin (left, NPB; right, es1-1). B, Trichome phenotype (macro hairs; upper, NPB;
down, es1-1). C, D, F-actin organization was visualized with Alexa Fluor 488-phalloidin in
root cells of wild type and es1. Over 40 roots were used in this analysis. Bar =5 μm (C, NPB;
D, es1-1). Arrows indicate the continuity of F-actin. E, Phylogenetic tree of SCAR-like 2
protein in several species.
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 4. Expression of ES1 and subcellular localization of ES1.
A–E, GUS staining of leaf (A), leaf sheath (B), fibrous root (C), stamen (D), and culm
vascular bundles (E). F, Quantitative real-time PCR analysis of ES1 expression in different
tissues at the tillering stage (T) and heading stage (H). The results are given as means of three
independent assays, and each error bar represents the percent standard deviation of the mean.
G–L, GFP fluorescence visualized by confocal microscopy. Transient expression of eGFP (G)
and eGFP–ES1 cDNA (H) in epidermal cells of onion. 35S::eGFP::ES1 cDNA and
35S::GFP were transiently expressed in onion epidermal cells. Transient expression of eGFP
in epidermal cells (I) and protoplasts (K) and expression of eGFP–ES1 in epidermal cells (J)
and protoplasts (L) of Nicotiana benthamiana leaves. 35S::eGFP::ES1 and 35S::GFP were
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
transiently expressed in N. benthamiana leaves.
2
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 5. Observation of stomata between NPB and es1-1.
A, Stomata density of the wild-type (left) and es1-1 (right) leaves. Bar = 30 µm. the red
triangle indicate the position of stomata. B, Statistical analysis of stomatal parameters, such
as stomata density, partially-open stomata, stomatal conductance, transpiration rate. C,
Morphological characteristics (left) and maximum opening size (right) of stomata between
wild-type (up) and es1-1 (down) plants. Bar = 5 µm D, Comparison of water loss rate
between wild-type and es1-1 plants with and without 300um ABA treatment at the
3-leaf-stage. E, qRT-PCR analysis of stomata density related genes (OsTMM, OsSCRM,
OsSDD1, OsYODA2) and ES1, and some ABA response-related genes (OsDREB1A,
OsDREB2A, OsMYB2,OsABI4, OsABI5, OsABF4) in WT and es1-1 leaves before and after
300um ABA treatment at the 3-leaf-stage. The values were normalized to the Actin1 gene. **
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
indicates significance (t-test) at the P<0.01. Different letters indicate significant difference at
the 1% level (Duncan multiple range test)
2
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 6. Observation of water loss rate between wild-type and es1-1 plants.
A, Water loss rate of detached leaves at the seedling stage (left, NPB; right, es1-1). B, Water
loss rate of detached leaves at the tillering stage. C, Water loss per unit of time of living
leaves between NPB and es1-1 plants at the tillering stage. Error bars indicate SD (n = 15).
Asterisks indicate the significance of differences between NPB and es1-1 plants, as
determined by t-test. *0.01 ≤ P < 0.05, **P < 0.01.
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 7. Phenotype comparison of the roots between NPB and es1-1 plants.
A, Root scanning of NPB plants; bar = 5 cm. B, Root scanning of es1-1 plants; bar = 5 cm.
C, Enlarged figure of roots in NPB plants; bar = 5 cm. D, Enlarged figure of root in es1-1
plants; bar = 5 cm. E–J, Statistical analysis of the main root length (E), total root length (F),
total area of the root (G), average diameter of the root (H), total volume of the root (I), total
number of fibrous roots (J) of NPB and es1-1 plants. The results are given as means of three
independent assays, and error bars indicate SD. Asterisks indicate the significance of
differences between NPB and es1-1 plants, as determined by t-test. *0.01 ≤ P < 0.05, **P <
0.01.
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Figure 8. Number of vascular bundles and conduits in wild-type and es1-1 plants.
A, Number of vascular bundles in NPB plants. B, Number of vascular bundles in es1-1 plants.
C, Comparison of the number of vascular bundles between NPB and es1-1 plants. D, Number
of conduits in NPB plants. E, Number of conduits in es1-1 plants. F, Comparison of the
number of conduits between NPB and es1-1 plants. G, Detection of the water potential value
in WT and es1-1. Relative water potential value in different tissues at seedling stage of WT
and es1-1 analyzed by WP4-T Dewpoint PotentiaMeter. The results are given as means of
three independent assays, and error bars indicate SD. *, ** significant at the level of 5% and
1%, respectively.
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Table 1. Comparison of mineral contents between wild-type and es1-1 plants.
root
culm
shealth
old
leaf
young
leaf
seed
NPB
es1-1
NPB
es1-1
NPB
es1-1
NPB
es1-1
NPB
es1-1
NPB
es1-1
Cu(mg kg-1)
23.5±3.83
446.3±18.52**
8.61±1.21
20.18±4.7**
6.59±1.15
16.88±2.43**
9.32±2.1
18.27±3.88**
8.02±1.16
17.07±2.91**
4.26±0.64
5.91±0.72**
Fe(mg kg-1)
3757.7±122.55
18155.3±326.27**
35.52±3.7
46.05±3.54**
136.2±8.59
240.2±14.65**
228.8±15.1
358.3±21.94**
102.1±6.78
223.1±11.4**
4.05±0.35
6.98±1.06**
Mn(mg kg-1)
17.75±1.75
58.97±3.88**
32.19±1.3
147.99±10.06**
92.331±5.9
582.2±13.22**
100.01±7.1
719.05±34.8**
27.51±3.32
175.81±12.24**
31.62±2.86
50.36±4.1**
Zn(mg kg-1)
28.49±1.12
91.26±5.88**
28.75±1.81
85.83±6.35**
42.52±3.4
65.36±7.92**
29.89±2.56
64.04±7.5**
30.48±3.44
62.22±4.28**
24.23±2.6
53.13±8.12**
The results are given as mean ± SD of three independent assays.
*, ** significant at the level of 5% and 1%, respectively
1
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Parsed Citations
Basu D, Le J, El-Essal SE, Huang S , Zhang C, Mallery EL, Koliantz G, Staiger CJ, Szymanski DB (2005) DISTORTED3/SCAR2 is a
putative Arabidopsis WAVE complex subunit that activates the Arp2/3 complex and is required for epidermal morphogenesis. Plant
Cell 17: 502-524
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Boonrueng N, Anuntalabhochai S, Jampeetong A (2013) Morphological and anatomical assessment of KDML 105 (Oryza sativa L.
spp. indica) and its mutants induced by low-energy ion beam. Rice Sci 20(3): 213-219
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Bouman B, Humphreys E, Tuong T, Barker R (2007) Rice and water. Adv Agron 92:187-237
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Bruce A, Merle M, Maria A, Rufus L, L.Lanier N, Eliete F, Chris V (2014) Reducing greenhouse gas emissions, water use, and grain
arsenic levels in rice systems. Global change biol 21:407-417
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Buckley TN (2005) The control of stomata by water balance. New Phytol 168:275-292
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Buschhaus C, Jetter R (2012) Composition and physiological function of the wax layers coating Arabidopsis leaves: ß-Amyrin
negatively affects the intracuticular water barrier. Plant Physiol 160: 1120-1129
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Cha K-W, Lee Y-J, Koh H-J, Lee B-M, Nam Y-W, Paek N-C (2002) Isolation, characterization, and mapping of the stay green mutant
in rice. Theor Appl Genet 104: 526-532
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Chen J, Lacomis L, Erdjument-Bromage H, Tempst P, Stamnes M (2004) Cytosol-derived proteins are sufficient for Arp2/3
recruitment and ARF/coatomer-dependent actin polymerization on Golgi membranes. FEBS Lett 566: 281-286
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Ding L, Li Y, Li Y, Shen Q, Guo S (2014) Effects of drought stress on photosynthesis and water status of rice leaves. Chin J Rice
Sci 28: 65-70 (in Chinese with English abstract)
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Du H, ChangY, Huang F, Xiong L (2014) GID1 modulates stomatal response and submergence tolerance involving abscisic acid
and gibberellic acid signaling in rice. J Integr Plant Biol doi: 10.1111/jipb.12313
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Dyachok J, Zhu L, Liao F, He J, Huq E, Blancaflor E (2011) SCAR mediates light-induced root elongation in Arabidopsis through
photoreceptors and proteasomes. Plant Cell 23: 3610-3626
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Eitzen G, Wang L, Thorngren N, Wickner W (2002) Remodeling of organelle-bound actin is required for yeast vacuole fusion. J Cell
Biol 158: 669-679
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Jäger K, Fábián A, Eitel G, Szabó L, Deák C, Barnabás B, Papp I (2014) A morpho-physiological approach differentiates bread
wheat cultivars of contrasting tolerance under cyclic water stress. Plant Physiol 171: 1256-1266
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Downloaded
from282,
on June
2017 - Published by www.plantphysiol.org
Grill E, Ziegler H (1998) A plant's dilemma.
Science
252 17,
-253.
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Hiei Y, Ohta S, Komari T, Kumashiro T (1994) Efficient transformation of rice (Oryza sativa L.) mediated by Agrobacterium and
sequence analysis of the boundaries of the T-DNA. Plant J 6: 271-282
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Hu L, Wang Z, Huang B (2010) Diffusion limitations and metabolic factors associated with inhibition and recovery of photosynthesis
from drought stress in a C3 perennial grass species. Physiol Plant 139: 93-106
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Huang R, Wang X, Lou C (2000) Cytoskeletal inhibitors suppress the stomatal opening of Vicia faba L. Induced by fusicoccin and
IAA. Plant Sci 156:65-71
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Huang X, Chao D, Gao J, Zhu M, Shi M, Lin H (2009) A previously unknown zinc finger protein, DST, regulates drought and salt
tolerance in rice via stomatal aperture control. Genes Dev 23: 1805-1817
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Islam MA, Du H, Ning J, Ye H, Xiong L (2009) Characterization of Glossy1-homologous genes in rice involved in leaf wax
accumulation and drought resistance. Plant Mol Biol 70: 443-456
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Kaksonen M, Sun Y, Drubin DG (2003) A pathway for association of receptors, adaptors, and actin during endocytic internalization.
Cell 115: 475-487
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Kanaoka M, Pillitteri L, Fujii H, Yoshida Y, Bogenschutz N, Takabayashi J, Zhu J, Torii K (2008) SCREAM/ICE1 and SCREAM2
specify three cell-state transitional steps leading to Arabidopsis stomatal differentiation. Plant Cell 20: 1775-1785
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Kim H, Park J, Hwang (2014) Investigating water transport through the xylem network in vascular plants. J Exp Bot 65:1895-1904
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Konrad W, Burkhardt J, Ebner M, Roth-Nebelsick R (2015) Leaf pubescence as a possibility to increase water use efficiency by
promoting condensation. Ecohydrol 8:480-492
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Le J, Mallery EL, Zhang C, Brankle S, Szymanski DB (2006) Arabidopsis BRICK1/HSPC300 is an essential WAVE-Complex subunit
that selectively stabilizes the Arp2/3 activator SCAR2. Curr Biol 16: 895-901
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Lee R, Wang C, Huang L, Chen S (2001) Leaf senescence in rice plants: cloning and characterization of senescence up-regulated
genes. J Exp Bot 52: 1117-1121
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Li J, Pandeya D, Nath K, Zulfugarov IS, Yoo S-C, Zhang H, Yoo J-H, Cho S-H., Koh H-J., Kim D-S, Seo HS, Kang B-C, Lee C-H, Paek
N-C (2010) ZEBRA-NECROSIS, a thylakoid-bound protein, is critical for the photoprotection of developing chloroplasts during
early leaf development. Plant J 62: 713-725
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Li S, Blanchoin L, Yang Z, Lord EM (2003) The putative Arabidopsis Arp2/3 complex controls leaf cell morphogenesis. Plant Physiol
132: 2034-2044
Pubmed: Author and Title
CrossRef: Author and Title
Downloaded
Google Scholar: Author Only Title Only Author
and Title from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Li Z, Zhang Y, Liu L, Liu Q, Bi Z, Yu N, Cheng S, Cao L (2014) Fine mapping of the lesion mimic and early senescence 1(lmes1) in
rice (Oryza sativa). Plant Physiol Biochem 80: 300-307
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Livak K, Schmittgen T (2001) Analysis of relative gene expression data using real-time quantitative PCR and the 2-??Ct method.
Methods 25: 402-408.
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Luo L, Zhang Q (2001) The status and strategy on drought resistance of rice (Oryza sativa L.) Chin J Rice Sci 15: 209-214 (in
Chinese with English abstract)
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Lv F (2010) Optimization for Agrobacterium Tumefaciens-mediated transformation system for exogenous genes of tobacco. J
Anhui Agri Sci 38: 10065-10066 (in Chinese with English abstract)
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Mathur J, Spielhofer P, Kost B, Chua N-H (1999) The actin cytoskeleton is required to elaborate and maintain spatial patterning
during trichome cell morphogenesis in Arabidopsis thaliana. Development 126: 5559-5568
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Mathur J, Mathur N, Kernebeck B, Hülskamp M (2003) Mutations in actin-related proteins 2 and 3 affect cell shape development in
Arabidopsis. Plant Cell 15: 1632-1645
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Moradi F, Iamail A.(2007)Responses of photosynthesis, chlorophyII fluorescence and ROS-scavenging systems to salt stress
during seedling and reproductive stages in rice. Ann Bot 99:1161-1173
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Nguyen HT, Babu RC, Blum A (1997) Breeding for drought resistance in rice: physiology and molecular genetics considerations.
Crop Sci 37: 1426-1434
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Park S, Yu J, Park J, Li J, Yoo S, Lee N, Lee S, Jeong S, Seo H, Koh H, Jeon J, Park Y, Paek N (2007) The senescence-induced
staygreen protein regulates chlorophyII degradation. Plant Cell 19: 1649-1664
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Qiu J-L, Jilk R, Marks MD, Szymanski DB (2002) The Arabidopsis SPIKE1 gene is required for normal cell shape control and tissue
development. Plant Cell 14: 101-118
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Riederer M, Schreiber (2001) Protecting against water loss: analysis of the barrier properties of plant cuticles. J Exp Bot 52: 20232032
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Saedler R, Mathur N, Srinivas BP, Kernebeck B, Hülskamp M, Mathur J (2004) Actin control over microtubules suggested by
DISTORTED2 encoding the Arabidopsis ARPC2 subunit homolog. Plant Cell Physiol 45: 813-822
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Sellammal R, Robin S, Raveendran M (2014) Association and heritability studies for drought resistance under varied moisture
stress regimes in backcross inbred population of rice. Rice Sci 21: 150-161.
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Shang Y, Yan L, Liu Z, Cao Z, Mei C, Xin Q, Wu F, Wang X, Du S, Jiang T, Zhang X, Zhao R, Sun H, Liu R, Yu Y, Zhang D (2010) The
Downloaded
from on
June 17,
- Published
by www.plantphysiol.org
Mg-chelatase H subunit of Arabidopsis
antagonizes
a group
of 2017
WRKY
transcription
repressors to relieve ABA-responsive genes
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
of inhibition. Plant Cell 22: 1909-1935
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Shao G, Chen M, Wang D, Xu C, Mou R, Cao Z, Zhang X (2008) Iron manure regulation of cadmium accumulation in rice. Chin Sci
Bull (C): Life Sci 38: 180-187
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Silva D, Favreau R, Auzmendi I, Dejong T(2011) Linking water stress effects on carbon partitioning by introducing a xylem circuit
into L-PEACH. Ann Bot 108:1135-1145
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Smith S, Smet I (2012) Root system architecture: insights from Arabidopsis and cereal crops. Phil.Trans.R.Soc.B 367:1441-1452
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Svenningsson M (1988) Epi- and intracuticular lipids and cuticular transpiration rates of primary leaves of eight barley (Hordeum
vulgare) cultivars. Plant Physiol 73: 512-517
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Svitkina TM, Borisy GG (1999) Arp2/3 complex and actin depolymerizing factor/cofilin in dendritic organization and treadmilling of
actin filament array in lamellipodia. J Cell Biol 145: 1009-1026
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Szymanski DB, Marks MD, Wick SM (1999) Organized F-Actin is essential for normal trichome morphogenesis in Arabidopsis. Plant
Cell 11: 2331-2347
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013) MEGA6:Molecular evolutionary genetics analysis version6.0. Mol Bio
Evol 30:2725-2729
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Tsuzuki T, Takahashi K, Tomiyama M, Inoue S, Kinoshita T (2013) Overexpression of the Mg-chelatase H subunit in guard cells
confers drought tolerance via promotion of stomatal closure in Arabidopsis thaliana. Front Plant Sci 4: 440
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Von G, Berger D, Altmann T (2002) The subtilisin-like serine protease SDD1 mediates cell-to-cell signaling during Arabidopsis
stomatal development. Plant Cell 14: 1527-1539
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Wang Y, Wan L, Zhang L, Zhang Z, Zhang H, Quan R, Zhou S, Huang R (2012) An ethylene response factor OsWR1 responsive to
drought stress transcriptionally activates wax synthesis related genes and increases wax production in rice. Plant Mol Biol 78:
275-288
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Wi SJ, Jang SJ, Park KY (2010) Inhibition of biphasic ethylene production enhances tolerance to abiotic stress by reducing the
accumulation of reactive oxygen species in Nicotiana tabacum. Mol Cells 30: 37-49
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Xiao Y, Chen Y, Wang X (2003) Filaments skeleton in guard cell. Chin Bull Bot 20: 489-494
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Xu H, Gauthier L, Gossenlin A (1995) Stomatal and cuticular transpiration of greenhouse tomato plants in response to high
solution electrical conductivity and low soil water content. J Amer Soc Hort Sci 120: 417-422
Pubmed: Author and Title
CrossRef: Author and Title
Downloaded
Google Scholar: Author Only Title Only Author
and Title from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.
Yang M, Sack F (1995) The too many mouths and four lips mutations affect stomatal production in Arabidopsis. Plant Cell 7: 22272239
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
You J, Zong W, Li X, Ning J, Hu H, Li X, Xiao J, Xiong L (2013) The SNAC1-targeted gene OsSRO1c modulates stomatal closure and
oxidative stress tolerance by regulating hydrogen peroxide in rice. J Exp Bot 64: 569-583
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Yu R, Huang R, Wang X, Yuan M (2001) Microtubule dynamics are involved in stomatal movement of Vicia faba L. Protoplasma
216:113-118
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Zhang C, Mallery EL, Schlueter J, Huang S, Fan Y, Brankle S, Staiger CJ, Szymanski DB (2008) Arabidopsis SCARs function
interchangeably to meet actin-related protein 2/3 activation thresholds during morphogenesis. Plant Cell 20: 995-1011
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Zhang Q, Lin F, Mao T, Nie J, Yan M, Yuan M, Zhang W (2012) Phosphatidic Acid Regulates Microtubule Organization by Interacting
with MAP65-1 in Response to Salt Stress in Arabidopsis. Plant Cell 24:4555-4576
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Zhang X, Dyachok J, Krishnakumar S, Smith LG, Oppenheimer DG (2005) IRREGULAR TRICHOME BRANCH1 in Arabidopsis
encodes a plant homolog of the actin-related protein2/3 complex activator Scar/WAVE that regulates actin and microtubule
organization. Plant Cell 17: 2314-2326
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Zhou L, Ni E, Yang J, Zhou H, Liang H, Li J, Jiang D, Wang Z, Liu Z, Zhuang C (2013) Rice OsGL1-6 is involved in leaf cuticular wax
accumulation and drought resistance. PLoS One 8: e65139 1-12
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Zhu LF, Yu SM, Jin QY (2012) Effects of aerated irrigation on leaf senescence at late growth stage and grain yield of rice. Rice Sci
19(1): 44-48.
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Zhu X, Xiong L (2013) Putative megaenzyme DWA1 plays essential roles in drought resistance by regulating stress-induced wax
deposition in rice. Proc Natl Acad Sci USA 110: 17790-17795
Pubmed: Author and Title
CrossRef: Author and Title
Google Scholar: Author Only Title Only Author and Title
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2015 American Society of Plant Biologists. All rights reserved.