Accepted Manuscript Title: To grow or not to grow: A stressful decision for plants Author: Rudy Dolferus PII: DOI: Reference:

S0168-9452(14)00243-X http://dx.doi.org/doi:10.1016/j.plantsci.2014.10.002 PSL 9061

To appear in:

Plant Science

Received date: Revised date: Accepted date:

4-9-2014 6-10-2014 9-10-2014

Please cite this article as: R. Dolferus, To grow or not to grow: a stressful decision for plants, Plant Science (2014), http://dx.doi.org/10.1016/j.plantsci.2014.10.002 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1 1 2 3

Review:

ip t

4 5

To grow or not to grow: a stressful decision for plants

cr

6

us

7 8

an

9 10

12

M

11 Rudy Dolferus

14 CSIRO

16

Agriculture Flagship

17 18 19 20 21

Ac ce p

15

te

d

13

GPO Box 1600

Canberra ACT 2601 Australia

Tel: +61-2-62465010

E-mail: [email protected]

22

Page 1 of 65

2 22

Contents

24 25 26

Abstract ..................................................................................................3 1. Introduction .......................................................................................4

27

1.2. Plant growth and the environment ......................................................................5

28

1.3. Can we exploit plant adaptive capacity?.............................................................6

29

1.4. The virtue of model plants ..................................................................................7

30

1.5. Are domesticated plants different? .....................................................................8

31 32

2. Genetic approaches for improving abiotic stress tolerance.............9

33

2.2. Avoidance and escape reactions .......................................................................10

34

2.3. Constitutive vs. inducible stress tolerance ........................................................11

35

2.4. QTL analysis in the genomics era.....................................................................15

36

2.5. Next generation phenotyping methods .............................................................16

37 38

3. Components of abiotic stress responses ..........................................17

39

3.2. First things first: establishment of cellular protection ......................................19

40

3.3. Taking care of metabolic adjustment................................................................20

41

3.4. Do abiotic stress response pathways overlap? ..................................................22

42

3.5. Selection for tolerance to multiple abiotic stresses...........................................23

44 45 46 47 48 49 50 51 52 53 54

us

cr

1.1. Abiotic stresses: definition and impact on agriculture........................................4

M

an

2.1. Field or controlled environment phenotyping?...................................................9

te

d

3.1. The power of transcriptomics ...........................................................................17

Ac ce p

43

ip t

23

3.6. Transgenic approaches for abiotic stress tolerance...........................................25

4. Coordination of growth responses to abiotic stress........................27 4.1. Do plants have brains? ......................................................................................27 4.2. Growth inhibition responses .............................................................................28 4.3. Growth stimulation responses...........................................................................30 4.4. An old legend born again: auxins .....................................................................32 4.5. Coordination of environmental responses ........................................................33

5. Conclusions.......................................................................................35 Acknowledgements ..............................................................................37 References ............................................................................................38 Figure Legends.....................................................................................59

55

Page 2 of 65

3 55

Abstract

57

Progress in improving abiotic stress tolerance of crop plants using classic breeding

58

and selection approaches has been slow. This has generally been blamed on the lack

59

of reliable traits and phenotyping methods for stress tolerance. In crops, abiotic stress

60

tolerance is most often measured in terms of yield-capacity under adverse weather

61

conditions. “Yield” is a complex trait and is determined by growth and developmental

62

processes which are controlled by environmental signals throughout the light cycle of

63

the plant. The use of model systems has allowed us to gradually unravel how plants

64

grow and develop, but our understanding of the flexibility and opportunistic nature of

65

plant development and its capacity to adapt growth to environmental cues is still

66

evolving. There is genetic variability for the capacity to maintain yield and

67

productivity under abiotic stress conditions in crop plants such as cereals.

68

Technological progress in various domains has made it increasingly possible to mine

69

that genetic variability and develop a better understanding about the basic mechanism

74

Ac ce p

te

d

M

an

us

cr

ip t

56

75

human nutrition, the cereals.

70 71 72 73

of plant growth and abiotic stress tolerance. The aim of this paper is not to give a detailed account of all current research progress, but instead to highlight some of the current research trends that may ultimately lead to strategies for stress-proofing crop species. The focus will be on abiotic stresses that are most often associated with climate change (drought, heat and cold) and those crops that are most important for

76 77

Keywords:

78

Abiotic stress/plant development/senescence/hormone regulation/cereals/crop yield

Page 3 of 65

4 79

1. Introduction

81

1.1. Abiotic stresses: definition and impact on agriculture

82

Plants are immobile and depend on their environment for growth and development.

83

This environment is variable and challenges plants with abiotic stress situations

84

throughout their life cycle: light (quality and quantity), mineral nutrition (depletion

85

and toxicity, salinity), temperature (heat, cold) and water availability (drought,

86

flooding). Plant development is therefore flexible and adjustable to the environment.

87

During evolution, wild plant species have learned to adapt to their natural

88

environment and this has determined their geographical distribution. In agricultural

89

environments, crop productivity is usually well controlled by agronomical practices,

90

but crop losses due to extreme and unexpected weather events are unavoidable. The

91

prospect of having to meet food demands for a 34% increase of the global population

92

by 2050 is imminent [1]. Crop yields will need boosting, but the higher frequency and

93

intensity of drought, heat and cold spells will also require crops that are better able to

98

Ac ce p

te

d

M

an

us

cr

ip t

80

99

therefore be based on a thorough understanding of the complexity of plant growth and

94 95 96 97

maintain productivity under sub-optimal conditions. The concept of “tolerance” and “sensitivity” of plants to abiotic stress situations can be difficult to measure. In model plants like Arabidopsis, tolerance is often measured as “survival”. In crop species like cereals maintenance of “yield” and “productivity” is for economical reasons more important than “survival”. The criteria to evaluate stress tolerance in crop plants must

100

developmental processes that ultimately correlate with maintenance of productivity.

101

Unfortunately, this knowledge is still evolving.

102

Page 4 of 65

5 Progress in cereal yield improvements has generally been slow and is starting to reach

104

a plateau, falling short of the annual yield increases required to meet 2050 food

105

demands [2]. In rice and wheat it is estimated that annual rate of yield increase has so

106

far been primarily achieved through improved managing practices (mechanisation)

107

rather than through breeding and genetic gain [3, 4]. During the last decade significant

108

progress has been made in improving our understanding about plant physiology and

109

molecular biology and new technologies have placed us now in a better position to

110

improve the efficiency of crop breeding. Improved knowledge and advanced new

111

technologies may now provide us with an opportunity to improve the speed and

112

efficiency of breeding to boost crop yield and abiotic stress tolerance.

an

us

cr

ip t

103

M

113

1.2. Plant growth and the environment

115

Plants continuously adjust growth and development, growing prolifically when

116

conditions are optimal and slowing down, arresting and even reversing growth (e.g.

117

abscission, senescence and cell death responses) under sub-optimal conditions - even

122

Ac ce p

te

d

114

123

is therefore not surprising that crop productivity attributes (yield, quality) are strongly

124

influenced by environmental variability (gene-environment interactions, GxE).

125

Consequently, the responsiveness of crop plants to abiotic stresses is equally variable

126

and is controlled by complex gene networks with epistatic interactions [5]. In the

127

field, crop plants are continuously challenged by a combination of stresses which are

118 119 120 121

when conditions are not life-threatening. This bidirectional growth adjustment mechanism is quite remarkable and poorly understood, but it may hold the key for improving abiotic stress tolerance. Plant growth is a measure of environmental input and adaptive capacity to a particular environment; some conditions can be controlled by humans (irrigation, fertilization etc.), but others are at the mercy of the weather. It

Page 5 of 65

6 often typical for that environment. Breeding activities are usually focused on specific

129

target environments, but this approach tends to improve adaptive traits that are

130

constitutively present and are relevant for that environment only. This approach may

131

have resulted in the loss of genetic variation from current breeding stock that would

132

allow the plant to maintain productivity under unexpected and/or more extreme stress

133

conditions. To identify germplasm that is better able to maintain productivity under

134

more challenging abiotic stress conditions, it will be necessary to increase the

135

selection standards and identify germplasm that is able to perform well under stress

136

conditions.

an

us

cr

ip t

128

137

1.3. Can we exploit plant adaptive capacity?

139

Plants in general (higher and lower plants) have a staggering capacity to adapt to

140

extreme environments and they can be found in most ecosystems of the globe. Some

141

grasses and flowering plants can be found on the Antarctic Peninsula [6], resurrection

142

plants are adapted to extremely hot and dry conditions [7, 8], while seagrasses are

147

Ac ce p

te

d

M

138

148

stress signalling and metabolic and developmental adaption mechanisms [8, 10].

149

Proof-of-concept transgenic approaches can be used to evaluate some of these

150

adaptation mechanisms (e.g., cryo- and osmo-protectants) in crop species such as

151

cereals. However, this may be difficult to achieve if genes of an entire metabolic

152

pathway need to be transferred and it may also compromise important yield and

143 144 145 146

land plants that have re-adapted to life in a marine environment, surviving conditions of low light, high salt and anoxia in the sediment [9]. Adaptation of plants to extreme environments requires complex morphological, developmental and metabolic adaptations. Exploring the molecular mechanisms of drought tolerance in resurrection plants and salt tolerance of halophytes has benefited our understanding about abiotic

Page 6 of 65

7 quality traits. Important morphological and developmental components that contribute

154

to abiotic stress tolerance simply cannot be transferred to cereals. Sourcing abiotic

155

stress tolerance traits from the available genetic variability in crop species, landraces

156

and progenitor species may be a more desirable approach.

ip t

153

157

1.4. The virtue of model plants

159

A small genome size has been an important criterion for the selection of model plants.

160

The simple dicot Arabidopsis has been a workhorse for advancing our understanding

161

of various plant biological processes, including plant development and response to

162

various abiotic stresses. Comparative genomics is starting to reveal important

163

differences between different model systems, suggesting that care is needed when

164

extrapolating information from model systems to other plants. For example, some

165

genes are missing in Arabidopsis that are present in other plants [11], while other

166

genes have diverged and evolved different functions in other plants. The control of

167

flowering and flower development differs considerably between eudicots

172

Ac ce p

te

d

M

an

us

cr

158

173

has become faster and cheaper, which has made it possible to sequence larger plant

174

genomes. In addition to the rice genome, the genome sequences of four other

175

cultivated grasses (maize, sorghum, barley and wheat; www.gramene.org) and one

176

wild grass (Brachypodium; www.brachypodium.org) are now available, providing a

177

wealth of information for comparative genomics studies into the evolution of these

168 169 170 171

(Arabidopsis) and monocots (rice, Brachypodium), even though the regulatory genes (e.g. MADS-box genes) identified in Arabidopsis are also present in monocots [1214]. In addition, it has been shown plants that many of the proteins with completely unknown function (POFs; proteins with obscure features) are species-specific and have no homologues in other species [15, 16]. In recent times, sequencing technology

Page 7 of 65

8 178

genomes (synteny, gene loss/conservation, gene divergence). This unstoppable

179

progress in sequencing technologies and genomics will ultimately reduce the reliance

180

on model systems.

ip t

181

1.5. Are domesticated plants different?

183

Domestication has turned wild ancestor varieties into cultivated crops that have a

184

different architecture and look vastly different from their progenitor species (e.g., rice:

185

Fig. 1). Selection for desirable traits has affected many plant developmental

186

processes, including yield-related traits (seed size and number), seed shattering, seed

187

dormancy, photoperiod and flowering time, palatability and overall shape and body

188

architecture [17, 18]. Comparative genomics is slowly revealing the effect of

189

domestication at the DNA level [19]. Comparison of the wild rice (Oryza rufipogon)

190

and cultivated rice (O. sativa japonica) genome sequences reveals significant gene

191

loss in the cultivated species [20]. A comparison between domesticated and wild

192

tomato revealed genes that underwent positive selection and many genes that showed

197

Ac ce p

te

d

M

an

us

cr

182

198

ancestor plants and landraces are often more tolerant to abiotic stresses and this

199

genetic variability could be re-introduced in domesticated crops. This process has

200

started for cereals such as rice [23] and maize [24] using wide crosses between

201

progenitors and interbreeding relatives. In bread wheat, the reconstruction of synthetic

202

hexaploids from the respective wild ancestors aims to achieve a similar goal [25].

193 194 195 196

shifts in gene expression levels [21]. Some gene deletions, mutations and variation in gene expression may have directly or indirectly affected abiotic stress tolerance. In beans, two DREB2 loci (dehydration-responsive element binding transcription factor) shared high levels of sequence diversity in one bean locus but no variation in the other, suggesting that domestication may have affected one of these genes [22]. Wild

Page 8 of 65

9 This process can be complicated by the lack of molecular markers for precision

204

breeding and the possible introduction of undesirable traits. These problems will be

205

discussed in more detail in the following chapter. Comparative genomics could also

206

be used to compare adaptation of crop species to abiotic stresses in different

207

environments and to compare genetic variability in stress tolerance [26]. The

208

difference in growth responses to environmental conditions can also reside in more

209

subtle changes in gene functions (e.g., base pair substitutions). Identifying those

210

differences will take additional effort.

us

cr

ip t

203

an

211

2. Genetic approaches for improving abiotic stress tolerance

213

2.1. Field or controlled environment phenotyping?

214

Controlled environments allow control over occurrence and timing of a stress during

215

plant development, as well as its duration and severity. It is also possible to

216

investigate the effect of a single abiotic stress at a time. This is a significant advantage

217

over field studies, where environmental conditions are typically variable and

219 220 221

d

te

Ac ce p

218

M

212

unpredictable. However, field environments remain difficult to simulate in growth chambers, even though technology is improving [27]. Air humidity, variation of light quantity and quality throughout the day (blue and red light enrichments at sunrise and sunset, respectively) control important plant physiological processes, yet are often

222

ignored in controlled environments. Additionally, heat load caused by light bulbs can

223

sometimes cause heat stress problems [28, 29] and soil drought and frost events are

224

extremely hard to simulate in controlled environments. In the field, environmental

225

changes that cause stress in plants most often occur over several hours (in the case of

226

heat during the day or frosts overnight) or even over several days (in the case of

Page 9 of 65

10 drought). These gradual changes are difficult to replicate in controlled environments

228

and stresses are often imposed abruptly, causing a shock situation by not allowing the

229

plant to gradually adapt to the stress. Additionally, growing plants in pots that are too

230

small affects root development, which in the case of drought stress affects the severity

231

and the speed with which the stress is imposed [30, 31]. Despite these issues,

232

controlled environments are the only tool that allows the comparison between

233

different stress responses independently and when used with due care and reasonable

234

attention, they can help to analyse stress responses in terms of sensitivity of different

235

plant developmental stages and effect of treatment duration and severity. This is very

236

important for designing phenotyping methods and to make sure that lines with

237

different flowering times are stressed at the same developmental stage when

238

comparing different lines. Communication with breeders and farmers can identify

239

germplasm that performs better/worse in field stress conditions and this material can

240

then provide an excellent benchmark to establish “realistic” stress treatment

241

conditions that give identical rankings in growth chambers. It is equally important to

242

replicate controlled environment results under field conditions.

246

Ac ce p

te

d

M

an

us

cr

ip t

227

247

flowering time [32, 33]. This is particularly troublesome when screening large

248

populations that segregate for flowering time genes. Flowering time is important for

249

optimizing grain yield in wheat, as flowering too early can result in cold and frost

250

damage and late flowering can result in poor yields due to drought and heat stress [34,

251

35]. Manipulating flowering time can also have adverse effects on yield; early-

243 244 245

2.2. Avoidance and escape reactions Selecting germplasm that is tolerant to abiotic stress under field conditions is compromised by escape or avoidance responses. In wheat, the damage caused by terminal drought can be alleviated by escaping drought through alteration of

Page 10 of 65

11 maturing varieties have less chance to accumulate biomass compared to late maturing

253

varieties, which indirectly affects grain yield in wheat [36]. Plants can also avoid

254

stress damage by adapting metabolic activity and growth rate. Accelerated growth

255

requires faster metabolism and mobilization of resources, while slowing down

256

metabolism and growth saves vital resources for passive survival of abiotic stress

257

conditions. Plants can use any of these tactics for survival in a particular environment.

258

In rice, ethylene response transcription factors (ERF) play an important role in

259

flooding tolerance [37, 38]. The ERF genes SNORKEL1 and 2 are important in deep-

260

water rice varieties, where elongation growth and outgrowing rising water levels

261

(escape response) is important for longer-term survival and grain production. In

262

contrast, another ERF-family member, SUBMERGENCE-1A (SUB1A), is important

263

in rice varieties that have to survive occasional short-term submergence and flooding

264

by transiently keeping growth and metabolic activity quiescent. Analysis of the

265

molecular basis indicates that the plant hormone ethylene plays an important role in

266

regulating elongation growth under stress conditions. The example of flooding

267

tolerance in rice illustrates the importance of regulating plant growth rate and

272

Ac ce p

te

d

M

an

us

cr

ip t

252

273

2.3. Constitutive vs. inducible stress tolerance

274

Selection for yield-related traits in field environments has dominated crop breeding.

275

Traits such as growth vigour, biomass accumulation, harvest index (reproductive

276

biomass), stem carbohydrate levels, tiller number, plant height, water use and

268 269 270 271

metabolic activity under stress conditions. Avoidance and escape reactions generally provide protection against abiotic stress through adjustment of growth rate and developmental processes. Understanding the molecular basis of these processes is important for understanding abiotic stress tolerance.

Page 11 of 65

12 transpiration efficiency, carbon isotope discrimination and root depth are important in

278

cereal breeding. These traits, together with improved management practises, have

279

improved vegetative growth of cereal crops, resulting in higher yield and productivity

280

[32, 33, 39-41]. Interestingly, for Australian wheat the yield gain was found to be

281

proportionally higher in the driest years compared to the better years even though

282

those traits were not specifically targeting drought conditions [4]. This illustrates that

283

yield-based traits that boost vegetative growth, biomass accumulation and water use

284

efficiency generally benefit plant growth and resilience and contribute to higher yields

285

under abiotic stress conditions [3-5, 41, 42]. However, unexpected and more extreme

286

abiotic stress conditions still result in massive yield penalties, indicating that growth

287

vigour and biomass accumulation does not necessarily result in a better capacity to

288

maintain that yield potential when growth conditions during reproductive

289

development are not favourable [5]. It is clear that an additional tolerance mechanism

290

is needed to convert or maintain the yield potential generated during the vegetative

291

stage to successful reproductive development and grain productivity. Sensitivity of

292

crops to various abiotic stresses are usually associated with phenotypes that are

297

Ac ce p

te

d

M

an

us

cr

ip t

277

298

compromised as growth repression saves resources for later growth when conditions

299

have returned to normal [43]. Even when this occurs, yield can never be recovered

300

because it is too late in the growing season and previously established biomass and

301

productivity has been lost. In many crops, growth repression can be an exaggeration

293 294 295 296

indicative of growth arrest, generally including leaf senescence or cell death, severe tissue necrosis (e.g., frosts and salinity), stomatal closure and arrest of photosynthesis [5]. At the reproductive stage, pollen sterility and abortion of grain development are similar growth repression phenotypes that cause major yield losses in cereals. Under shorter-duration or unexpected stress periods, survival of the plant is often not

Page 12 of 65

13 or an overly sensitive response to stress conditions. There may therefore be an

303

opportunity for increasing the threshold level at which growth repression takes place

304

in response to stresses, in order to maintain growth and productivity for as long as

305

possible. Interestingly, the two flooding tolerance mechanisms described earlier for

306

rice (section 2.2: growth acceleration versus metabolic quiescence) may be more

307

widely applicable for other abiotic stresses and the molecular understanding of

308

flooding tolerance may stimulate research on other abiotic stresses. An intriguing

309

question is whether avoidance and escape strategies should also be seen as part of the

310

plants overall strategy to tolerate abiotic stresses.

an

us

cr

ip t

302

311

To identify crops that maintain growth and yield potential under adverse growth

313

conditions it is necessary to complement constitutively expressed yield traits with

314

traits that are induced and specifically expressed under stress conditions. Identifying

315

more stress-induced traits will have a positive effect for breeding stress-tolerant crops,

316

but also for improving our understanding of the underlying physiological and

317

molecular mechanism. Under stress conditions, plants require mechanisms to protect

322

Ac ce p

te

d

M

312

323

(irrigated, rain-fed, rainout shelter plots to compare water stress conditions) [44].

324

However, controlled environments offer significant advantages if precision-

325

phenotyping is required (see section 2.1). Leaf senescence is a stress-induced

326

phenotype and has received a lot of attention as a stress-induced trait. Selecting for

318 319 320 321

their cellular machinery, metabolic adaptations and their capacity to sustain growth and development (see section 3). Selection for stress-inducible traits is more difficult to achieve, considering the unpredictability of field conditions and interference of avoidance and escape reactions. Field plots can be selected to target certain abiotic stresses (drought, heat, frost) or can be artificially modified to create stress conditions

Page 13 of 65

14 delayed foliar senescence (stay-green) and maintenance of stomatal conductance,

328

transpiration and photosynthesis during stress conditions are relatively easy

329

phenotypes to score [45]. Significant improvements in drought tolerance have resulted

330

from proof-of-concept transgenic approaches manipulating cytokinin levels,

331

confirming that this trait contributes to stress tolerance [46]. Leaf rolling is another

332

distinctive and easy to score heat and drought-induced phenotype. Reduction in leaf

333

area prevents transpiration and water loss and genetic variation for leaf rolling is

334

available in wheat and rice. However, leaf rolling does not always correlate with

335

drought tolerance, suggesting that it could be an escape rather than tolerance

336

mechanism [47, 48]. Osmotic adjustment is an inducible drought adaptation

337

mechanism that maintains leaf water potential through the synthesis of osmotically

338

active substances [44, 50]. Osmotic adjustment delays leaf senescence and leaf

339

rolling, maintains stomatal conductance and turgor pressure, thereby sustaining

340

growth under drought conditions [51]. Despite its importance, osmotic adjustment has

341

so far remained a difficult trait to phenotype [52]. Many abiotic stresses cause pollen

342

sterility and loss of fertility and grain yield in cereals [53, 54]. A phenotyping method

347

Ac ce p

te

d

M

an

us

cr

ip t

327

348

building up a higher yield potential, stress-inducible traits are essential to sustain that

349

higher yield potential under adverse environmental conditions to maintain growth and

350

reproductive development.

343 344 345 346

for cold and drought-induced pollen sterility was established using controlled environments [55, 56] and the plant hormone abscisic acid (ABA) was shown to play a role in stress-induced pollen abortion [53, 57, 58]. Stress-inducible traits are less likely to have negative effects on productivity of crops under non-stress conditions. While yield traits that improve vegetative plant growth and development contribute to

351

Page 14 of 65

15

2.4. QTL analysis in the genomics era

353

Identifying genetic variation for abiotic stress tolerance in crops requires the tedious

354

and laborious process of establishing linkage maps using DNA-based molecular

355

markers: restriction fragment length polymorphism (RFLP), amplified fragment

356

length polymorphism (AFLP), random amplification of polymorphic DNA (RAPD),

357

cleaved amplified polymorphic sequences (CAPS) and simple sequence repeat (SSR,

358

microsatellite) markers. In the last decade, the shift to high-throughput technologies

359

such as Diversity Arrays (DArT; [59]) and Single Nucleotide Polymorphism markers

360

(SNP; [60]) has made the construction of high density genomic maps easier. The

361

identification of SNP markers was boosted by the availability of the genome sequence

362

for many crop species. 160,000 SNPs were identified in the non-repetitive genome

363

fraction of 20 different rice varieties [61]. The availability of annotated genome

364

sequences and accurate high density SNP maps makes it easier to identify candidate

365

genes within QTL (Quantitative Trait Loci) regions [61, 62] and the lowering in

366

sequencing costs has made it possible to carry out genotyping by sequencing (GBS),

367

which further facilitates fine-mapping QTL [63]. Genome-wide association studies

369 370 371 372

cr

us

an

M

d

te

Ac ce p

368

ip t

352

(GWAS, [64]) also benefit from high density SNP maps and can be used for mapping abiotic stress tolerance loci. In wheat, the development of multi-parent advanced inter-cross populations (MAGIC) provides a powerful tool for mapping QTL [65].

Abiotic stress tolerance is typically controlled by a large number of QTL with

373

epistatic interactions and low phenotypic contribution and heritability. A

374

comprehensive overview of QTL for various abiotic stress-related traits can be

375

accessed at the Gramene and Plant Stress websites (archive.gramene.org/qtl/;

376

www.plantstress.com/files/qtls_for_ resistance.htm). Abiotic stress QTL mapping and

Page 15 of 65

16 genomic selection (GS) has so far not led to markers for routine use in marker-

378

assisted selection (MAS) for abiotic stress tolerance [41, 66, 67]. With the bottleneck

379

of genotyping removed, mapping of abiotic stress tolerance loci will depend on the

380

availability of reliable traits for phenotyping. The case of salinity tolerance in rice is a

381

good example that QTL analysis can lead to identification of candidate genes

382

provided that reliable phenotyping methods are available [68].

cr

ip t

377

us

383

2.5. Next generation phenotyping methods

385

Considering the difficulties involved in direct selection for abiotic stress tolerance in

386

field or controlled environments, shifting from “observable” to molecular or

387

secondary traits that are highly correlated with abiotic stress tolerance, may improve

388

reliability of phenotyping procedures [69]. Our understanding about the physiological

389

and molecular basis of stress responses has improved and technological progress in

390

the last decade has provided opportunities for high-throughput phenotyping.

391

Metabolomics is a promising technology that can now be used at a scale that is

396

Ac ce p

te

d

M

an

384

397

used to quantitatively and qualitatively evaluate components of cellular protection.

398

Hormone measurements can be used as indicators of developmental responses in

399

sensitive and tolerant germplasm (senescence or growth). Proteomics can also be used

400

for phenotyping abiotic stress responses, but protein expression profiling can be

401

technically more challenging (e.g., resolution limits of two-dimensional

392 393 394 395

compatible with population screening and mQTL mapping. This technology was used to identify genes controlling several metabolites and quality-related traits [70-72], but can also be used to map mQTL for metabolite changes associated with abiotic stresses [73, 74]. Measuring diagnostic metabolites can be informative about the physiological state of plant tissues in response to drought, heat and cold and metabolomics can be

Page 16 of 65

17 electrophoresis and detection limits of mass spectrometry) and is harder to adapt for

403

high throughput screening [75]. The development in recent years of various digital

404

imaging technologies has added even more opportunities for phenotyping [67, 76].

405

Non-destructive imaging can measure canopy properties that contribute to biomass

406

accumulation, as well as stress-related traits (photosynthesis, transpiration and leaf

407

senescence). This technology can be applied for high-throughput screening under

408

field or controlled environments [77-79]. New generation phenotyping technologies

409

are powerful but require some knowledge about the molecular and physiological basis

410

of abiotic stress phenotypes and the questions to be addressed.

an

us

cr

ip t

402

411

3. Components of abiotic stress responses

413

3.1. The power of transcriptomics

414

While GxE interactions are considered problematic and something to avoid in plant

415

breeding, molecular biologists have used differential gene expression of stress-treated

416

versus unstressed plant material as a standard method to study abiotic stress

418 419 420 421

d

te

Ac ce p

417

M

412

responses. In the recent decade, large-scale transcriptome analyses using microarrays and more recently new generation sequencing technologies (RNA-seq) have proven to be a powerful tool for identifying genes and cellular processes that are affected by abiotic stresses [80]. A massive amount of transcriptome information for different stresses and plant species is currently available in public databases.

422 423

Transcriptome information is starting to reveal how plants respond to various abiotic

424

stresses, but the full potential is still unexplored [81]. Currently, about 40% of the

425

proteins encoded by a eukaryotic genome have an unknown function [82]. It is

Page 17 of 65

18 estimated that between 18 and 38% of the eukaryotic proteome consists of proteins

427

without any defined domain or motif [15]. Interestingly, when comparing the

428

Arabidopsis and rice proteins with totally unknown features (POFs) nearly half were

429

found to be species-specific and had no homolog in the other genome [16].

430

Obviously, identifying the function of these proteins will require species-specific

431

studies and this will be a major challenge. Finding the exact physiological function for

432

members of large gene families (e.g., transcription factors) can also be a complicated

433

and time-consuming process. In model plants such as Arabidopsis and rice, insertion

434

mutagenesis using T-DNA and transposons can be used to identify gene functions and

435

support the gene annotation process. In rice, about 60.49% of the nuclear genes have

436

been tagged with T-DNA or Tos17 transposon insertions [83], but the functional

437

characterization of these insertion mutants remains a major effort. In addition, the

438

function of some genes for which the insertion mutant phenotype is lethal cannot be

439

investigated. Another limitation is gene redundancy and lack of a clear phenotype for

440

some mutations. Over-expression and RNAi technology can also be used to reveal the

441

function of candidate genes in plants that can be transformed. Transcriptome analysis

446

Ac ce p

te

d

M

an

us

cr

ip t

426

447

triggering early developmental responses, are likely to be present all the time and

448

simply require activation by upstream signals (e.g., phosphorylation). Identifying

449

those genes will require more fundamental approaches, ideally using model systems

450

in the first place (e.g., using mutagenesis approaches).

442 443 444 445

needs support from other technologies to speed up the identification of unknown gene functions. A systems biology approach combining transcriptomics with proteomics and metabolomics can help this process [84, 85]. It is also important to realize that transcriptomics focuses on differentially expressed genes, while some genes that play an important role in the early stress signal perception and transduction events, or those

Page 18 of 65

19

3.2. First things first: establishment of cellular protection

452

Bacteria, yeast and animals have a general cellular stress response mechanism that

453

protects essential macromolecules (DNA, proteins and lipids) against oxidative stress

454

and removes damaged cells using a cell death response. The conservation of this

455

response in different life forms suggests that it is an ancient protection mechanism

456

against general stress situations. The minimal cellular stress response proteome

457

consists of 44 proteins with known function, including molecular chaperones (e.g.

458

heat shock proteins), various enzymes that repair DNA damage and various proteins

459

that protect against oxidative stress and reactive oxygen species (ROS), such as

460

superoxide dismutase and glutathione antioxidant defence pathway proteins [86].

an

us

cr

ip t

451

M

461

Plants also activate a cellular protection mechanism in response to various stresses.

463

Little is known about macromolecule protection in plants, but chaperone proteins (e.g.

464

heat shock proteins) are induced by all abiotic stresses and their importance is

465

illustrated by the fact that an Escherichia coli gene encoding a cold shock protein that

466

functions as RNA chaperone can significantly improve tolerance to multiple stresses

468 469 470 471

te

Ac ce p

467

d

462

(cold, drought, heat) in transgenic rice and maize [87]. The transformation of light into chemical energy during photosynthesis and the mitochondrial electron transport chain produce damaging free radicals [88]. Regulation of intracellular redox homeostasis has been shown to control important metabolic pathways such as photosynthesis [89, 90] and is also important for regulating root and leaf

472

developmental processes [91, 92]. Superoxide, hydrogen peroxide and hydroxyl

473

radical production is induced in response to abiotic and biotic stresses and results in

474

activation of genes encoding ROS-detoxifying enzymes [93, 94]. Active oxygen

475

species such as hydrogen peroxide are generally considered as local and systemic

Page 19 of 65

20 signals in response to various stress situations [95, 96]. This indicates that plants may

477

have turned this early stress defence mechanism into a systemic warning signal to

478

protect different plant parts. Some oxidative stress-related genes are expressed in cells

479

associated with the vascular bundles [97], which is compatible with a systemic

480

signalling function of ROS [98-100]. Resistance to the ROS-generating herbicide

481

paraquat in Conyza bonariensis is correlated with a highly expressed constitutive

482

ROS detoxification system and cross-tolerance to environmental oxidants [101, 102].

483

Paraquat resistance in wheat and barley has been correlated with tolerance to water

484

stress and paraquat treatment has been evaluated as a screening system for abiotic

485

stress tolerance [103, 104]. Overexpression of peroxidase, catalase, superoxide

486

reductase and superoxide dismutase in transgenic plants has resulted in improved

487

tolerance to cold, drought, salinity and heat stress [105-108], while an ascorbate

488

deficient mutant in Arabidopsis caused a stress-sensitive phenotype [109]. Cellular

489

protection in plants may also function as an intracellular and systemic signal to

490

regulate developmental processes. As a stress defence mechanism it may be essential

491

for all other aspects of the stress response to function and it could therefore act as an

493 494 495

cr

us

an

M

d

te

Ac ce p

492

ip t

476

“enabling” mechanism that needs to be activated before other aspects of stress responses (biotic and abiotic) can be established (Fig. 2).

3.3. Taking care of metabolic adjustment

496

Changes in plant growth and development under abiotic stress conditions must be

497

associated with metabolic activity to provide the energy required to establish the

498

response. Firstly, the altered cellular environment requires changes in the cellular

499

machinery to be put in place; adaptations of translation initiation and protein folding

500

are commonly observed in stress-induced transcriptomes [110, 111]. Then, specially

Page 20 of 65

21 adapted metabolic proteins are induced early in the stress response (Fig. 2). Many

502

abiotic stresses shut down photosynthesis, while photosynthates are a crucial source

503

of energy. Sugars are transported from source to sink tissues via the phloem and are

504

important signals for growth and development, as well as response to various abiotic

505

stresses. Sugar signalling and metabolism are therefore tightly linked to growth

506

responses [112]. ABA regulates stomatal conductance and photosynthetic activity,

507

causing vegetative growth retardation. This has been shown to contribute to

508

vegetative stage abiotic stress tolerance, but this growth repression also has a negative

509

effect on reproductive processes [113]. Abiotic stresses repress the sucrose cleaving

510

enzyme cell wall invertase in anthers, preventing hexose supply for pollen

511

development and causing pollen sterility. ABA accumulation was shown to directly or

512

indirectly repress cell wall invertase expression. Tolerant wheat and rice germplasm

513

displayed different anther ABA homeostasis, maintaining lower ABA levels than

514

sensitive lines in response to cold and drought stress [55-57, 113]. Sugars and ABA

515

are known to regulate ethylene and senescence responses [114]. Sugars are an

516

important growth signal in plants and are therefore tightly connected with the decision

521

Ac ce p

te

d

M

an

us

cr

ip t

501

522

osmotic stresses such as drought, cold and salt stress and are tightly integrated with

523

signalling pathways for sugars, essential nutrients (nitrogen) and various hormones

524

[118-120]. The yeast SnRK (Sucrose non-fermenting related kinase) related protein

525

kinases KIN10 and KIN11 play a central role in coordinating sugar, stress and

517 518 519 520

to grow or repress growth and respond to abiotic stress (Fig. 2). The glycolytic enzyme hexokinase (HXK) is a cellular sugar sensor that cross-talks to several phytohormones [115]. Other cellular components involved in regulation of metabolism in plants show some similarity to yeast. The Mitogen-Activated Protein Kinase (MAPK) [116, 117] and the Salt Overly Sensitive (SOS) pathways respond to

Page 21 of 65

22 developmental signals with metabolic pathways, activating gene expression via bZIP

527

transcription factors [121, 122]. The General Control Non-repressible related protein

528

kinase (GCN-2) phosphorylates translation initiation factor 2 (eIF2α) and is able to

529

sense free amino acid levels, respond to osmotic stresses and control protein synthesis

530

[123, 124]. Both SnRK1 and GCN-2 regulate nitrate reductase and nitrogen

531

metabolism [120, 121]. Further research is required to establish how this complex

532

metabolic regulation mechanism is controlled by environmental stimuli. The

533

conservation of the kinases that regulate fundamental metabolic pathways between

534

plants, yeast and animals illustrates their evolutionary importance.

an

us

cr

ip t

526

535

3.4. Do abiotic stress response pathways overlap?

537

It has been demonstrated that treatment with one abiotic stress can provide “cross-

538

tolerance” or “hardening” to other stresses, including biotic stresses [125, 126]. Pre-

539

treatment of plants with the stress hormone ABA has a similar effect [127-129]. This

540

already suggests that there must be some functional overlap in the signalling and

545

Ac ce p

te

d

M

536

546

pollen fertility, induction of osmo-protectants by drought, cold and salinity) are also

547

shared by different stresses. Communication between sink and source tissues is

548

especially important under abiotic stress conditions when growth can be limited by

549

available resources. To fully understand the impact of abiotic stresses on plant growth

541 542 543 544

response pathways of abiotic stresses. Osmotic stresses (drought, cold, salinity, heat) involve ABA and are therefore expected to share common components. Furthermore, transcriptome analyses have confirmed that early macromolecule and oxidative stress protection is recruited by most stresses and is a general stress response (Fig. 2). Some developmental and metabolic responses (e.g., growth adaptation, leaf senescence,

Page 22 of 65

23 550

it is essential to further unravel the relationships between metabolism and

551

developmental processes.

552 Due to gene redundancy, different gene copies encoding proteins with the same

554

function can be activated under different stresses, suggesting that overlap between

555

different stresses could be larger at the protein level than at the transcript level. In

556

economical terms, it seems logical that plants will share the response to the initial

557

threat and mount stress-specific responses once the general response has created the

558

environment to make this possible (Fig. 2). A shared initial stress response may not

559

require many genes and some may have so far remained undetected in differential

560

gene expression studies. Regulation at the protein level, such as targeted protein

561

degradation using the ubiquitin proteasome, is commonly used by plant hormones to

562

regulate downstream developmental signalling [130]. It is therefore possible that a

563

rather small - but critically important - part of the overlap between different abiotic

564

stress responses has so far escaped detection.

cr

us

an

M

d

te

569

Ac ce p

565

ip t

553

570

arrest is taken in this early response mechanism and it is an important factor for

571

determining productivity and abiotic stress tolerance in crops. Selection for adaptation

572

to specific environments and productivity traits has affected many developmental

573

properties and may also have modified the early response mechanism to stress (Fig.

574

2). Selection against seed dormancy may have affected homeostasis of hormones that

566 567 568

3.5. Selection for tolerance to multiple abiotic stresses Genetic diversity for abiotic stress tolerance is more likely to occur in the early response mechanism than in the stress-specific responses that depend on the initial response (Fig. 2). The decision to continue growth or induce senescence and growth

Page 23 of 65

24 play a role in stress tolerance (ABA, GA). Adaptation to seasonal conditions and

576

different environments in cereals has modified important growth processes such as

577

duration of photosynthesis, adaptation to day-length and altered rate of leaf

578

senescence, which may also have changed the interaction between plant hormones

579

(ethylene, cytokinins). The stay-green trait has therefore been thought of as a potential

580

domestication trait [45, 131]. Analysis of stay-green QTL (e.g., sorghum) could lead

581

to positional cloning and identification of new genes involved in this process [45,

582

132].

us

cr

ip t

575

an

583

In nature, frequent exposure to a combination of stresses may have selected a shared

585

genetic adaptation mechanism to those stresses. This may be the case for heat and

586

drought stress which often occur at the same time in field conditions [126, 133]. It

587

therefore makes sense to select germplasm that is tolerant to more than one abiotic

588

stress [126], especially since different abiotic stress responses may already share a

589

common initial response mechanism (Fig. 2). Germplasm that establishes the initial

590

response successfully may also be better able to establish a stress-specific response

595

Ac ce p

te

d

M

584

596

wheat lines that are tolerant to multiple abiotic stresses individually (drought, heat,

597

shading and cold). QTL mapping using this tolerant germplasm can also be used to

598

identify overlapping general stress-response QTL, as well as QTL that are specific for

599

different stresses. Obtaining high levels of abiotic stress tolerance may ultimately

591 592 593 594

(Fig. 2). However, selecting germplasm that is tolerant to more than one stress using the combination of those stresses may be difficult to achieve because of quantitative and qualitative differences in tolerance to the combination of stresses and the difficulty to choose physiologically relevant selection conditions for the combination of stresses. We are currently using a step-wise selection procedure to first identify

Page 24 of 65

25 require combining both general stress response and stress-specific QTL. However, for

601

some traits it may be difficult to obtain tolerance to a combination of stresses; some

602

traits for drought (stomata closed to prevent water loss) and heat stress (open stomata

603

to reduce leaf temperature) are mutually exclusive [126].

604

ip t

600

3.6. Transgenic approaches for abiotic stress tolerance

606

It is common practice in molecular biology to use proof-of-concept transgenic

607

approaches (over-expression, RNAi) to evaluate candidate genes for their effect on

608

abiotic stress tolerance. Many genes have indeed been shown to improve abiotic stress

609

tolerance ([134]; see Plant Stress website for a comprehensive listing:

610

www.plantstress.com/ files/abiotic-stress_gene.htm). These include transcription and

611

regulatory factors, osmo-protectants, hormone and oxidative stress-related genes,

612

molecular chaperones, transporters and various metabolic genes. Abiotic stress

613

tolerance for most of these genes was evaluated under controlled environment or

614

glasshouse conditions, using model systems such as Arabidopsis or tobacco and some

619

Ac ce p

te

d

M

an

us

cr

605

620

tolerance may also be a contributing factor; in model plants tolerance is usually

621

measured as survival under vegetative stage stresses, while in crops maintenance of

622

productivity during reproductive stage stresses is important [136, 137]. The choice of

623

promoter to drive transgene expression is also important. Strong constitutive

624

promoters lead to ectopic expression of a transgene, potentially causing adverse

615 616 617 618

were also evaluated in cereals (rice, wheat, barley, maize). Relatively few transgenic lines have so far made any impact for improving field abiotic stress tolerance [135]. Potential explanations are that the field environment is much harsher, transgenes only partially improve the abiotic stress response or they improve response to one stress and not a combination of stresses. Differences in evaluation criteria for abiotic stress

Page 25 of 65

26 secondary effects on crop productivity. CBF/DREB transcription factors can improve

626

osmotic stress tolerance but result in stunted growth when using constitutive

627

promoters [138]; plants look normal when a strong drought-inducible promoter is

628

used that expresses the transgene only when required [139]. The quantitative and

629

qualitative properties of the promoter driving a transgene may be particularly critical

630

in the case of transcription factors, hormone metabolic and signal transduction genes.

631

Using Arabidopsis as a model system it may be extremely difficult or impossible to

632

fully evaluate and predict potential adverse effects in crop species [139]. In the case

633

of multigene families it is often difficult to find out which gene to use for

634

transformation. For instance, only a few aquaporin gene family members are affected

635

by stress and lead to improvement of drought tolerance in transgenic plants [140]. For

636

large transcription factor families, trial and error approaches can identify which gene

637

has a positive effect on stress tolerance [141]. Some transgenic approaches may result

638

in morphologically different plants where stress phenotypes are simply delayed (e.g.,

639

smaller leaf area reduce transpiration under drought), giving transgenics an unfair

640

advantage [142]. Manipulation of cytokinin levels using a stress-induced promoter led

645

Ac ce p

te

d

M

an

us

cr

ip t

625

646

multiple stresses (cold, drought and heat) in field experiments [88], suggesting that

647

focusing on the manipulation of the top of the stress signalling cascade using general

648

stress responsive genes may yield positive results.

641 642 643 644

to delayed senescence and improved drought tolerance in rice under glasshouse conditions [143]. It is possible that these transgenic lines may also perform well under field drought conditions, considering the field experience with stay-green plants. Interestingly, transgenic rice and maize plants expressing an E. coli cold shock protein

that acts as RNA chaperone in cellular protection resulted in improved tolerance to

649

Page 26 of 65

27

4. Coordination of growth responses to abiotic stress

651

4.1. Do plants have brains?

652

How are the early responses to abiotic stresses orchestrated by signals from the

653

environment? Higher animals are mobile and react to stress by escaping

654

environmental challenges. The brain processes environmental signals via the central

655

nervous system and regulates this mobility and escape reaction. The flexibility of

656

plant development in response to environmental change indicates that they have an

657

efficient systemic signalling mechanism that coordinates and orchestrates the

658

response to adverse environmental conditions. The plant vascular system bears some

659

resemblance to an animal central nervous system, sparking some speculation that

660

plants have a cellular communication mechanism similar to animals [144]. Specific

661

proteins known to play a role as neurotransmitters in animals (e.g., glutamate

662

receptors, 14-3-3 proteins) are also encoded in plant genomes but they acquired

663

different functions when plants evolved into multicellular organisms. The vascular

664

system plays an important role in coordinating growth and development between the

666 667 668 669

cr

us

an

M

d

te

Ac ce p

665

ip t

650

different plant parts [145], but the signalling mechanism is vastly different to that of animals. Plants have evolved their own systemic signals to drive growth and development (Fig. 2). Being photosynthetic organisms, they use photosynthates and the capacity to produce sugars as a resource signal for growth. They also adapted reactive oxygen species as a signal for abiotic stress. Plants have evolved their own

670

hormone signals, which are totally unlike animal hormones, to signal developmental

671

and growth responses. An emerging theme in plant biology is the observation that

672

many genes involved in hormone synthesis and signalling, e.g. those involved in ABA

673

synthesis and signalling [146, 58], are expressed in vascular parenchyma cells. This

Page 27 of 65

28 allows rapid signal perception and distribution via the vascular system, similar to a

675

nervous system but at a slower pace. Environmental signals can therefore be sensed in

676

any plant part and quickly spread throughout the plant, suggesting that plants have

677

essentially obviated the need for a central nervous system. Despite its importance, the

678

signal transmitting function of the vascular system still needs to be unravelled [147].

679

Understanding of how the signals (hormones, ROS, metabolites) themselves work is

680

gradually emerging. Auxins can be transported all over the plant using directional

681

efflux carriers and long-distance transporters and tissue-specific response mechanisms

682

make it possible to mount different auxin responses in different plant parts [148].

683

These different plant parts are pre-programmed to react differently to plant hormone

684

signals, explaining how environmental signals can have different but coordinated

685

responses. Hormonal signals therefore form an important link between the

686

environment and developmental processes. Plant evolution and global diversity is

687

testimony that plants have evolved efficient systems to manage and adapt to

688

environmental challenges.

cr

us

an

M

d

te

693

Ac ce p

689

ip t

674

694

development and have been implicated in abiotic stress responses. A lot of progress

695

has been made in recent years in understanding their function. Plant hormones can be

696

growth-retarding (ABA, ethylene, jasmonic acid) or growth-promoting (auxin,

697

gibberellic acid, cytokinin). Two recently discovered plant hormones, brassinosteroids

698

[149] and strigolactones [150] act in conjunction with auxins and can be classified as

690 691 692

4.2. Growth inhibition responses The key to understanding abiotic stress tolerance resides in understanding the plant’s capacity to accelerate/maintain or repress growth. Interaction between plant hormones must play an important role in this phenomenon. Most plant hormones play a role in

Page 28 of 65

29 growth promoting hormones, while salicylic acid functions in plant defence responses

700

to pathogens [151]. The stress hormone ABA is implicated in stomatal closure and

701

regulation of plant water balance, which impairs photosynthesis and restricts growth

702

[113]. ABA levels are up-regulated in response to osmotic stresses (drought, cold,

703

salinity) and heat stress. Higher ABA levels improve stress tolerance at the vegetative

704

level, but there is a compromise at the reproductive level. QTL analysis in maize

705

indicated that lines with higher root ABA levels had lower grain yield [152] and our

706

own work demonstrated that lower ABA levels in stressed anthers was correlated with

707

better cold and drought tolerance, as well as maintenance of anther sink strength in

708

rice [58]. The ABA signalling pathway interacts with other hormones and sugar

709

signalling via the SnRK network [98, 123]. ABA’s restriction of photosynthesis and

710

photosynthate allocation to sink tissues may help in shorter periods of abiotic stress,

711

but is destructive under longer term stress conditions such as terminal drought. The

712

opposing effect of ABA on vegetative and reproductive structures indicates that it is

713

important to understand the effect of abiotic stress during plant development. The

714

success of transgenic approaches for manipulation of abiotic stress tolerance will

719

Ac ce p

te

d

M

an

us

cr

ip t

699

720

tapetum [153, 154]. External application and stress-induced accumulation of ABA

721

results in senescence, but the role of ABA in this process is still unclear. ABA may

722

interact with the oxidative stress response that protects against senescence [155], but

723

may also cause senescence via interaction with ethylene. The role of ethylene in

715 716 717 718

depend on carefully targeting the control of ABA homeostasis to particular tissues and growth stages.

Growth repression under abiotic stress conditions is associated with induction of leaf senescence (Fig. 3) or programmed cell death responses in tissues such as the anther

Page 29 of 65

30 inducing leaf senescence and inhibition of root elongation has been well investigated.

725

Antisense repression of ethylene biosynthesis inhibits senescence, and the limitation

726

of ethylene production has in many cases resulted in improved abiotic stress tolerance

727

[156]. Ethylene can induce the biosynthesis of the growth-promoting hormone auxin

728

in a tissue-specific manner [157]. The role of ethylene can therefore also be growth-

729

promoting; low light (etiolation) and shading conditions cause elongation growth in

730

shading-sensitive plants [158]. Ethylene is therefore in a unique position to control

731

plant developmental processes: it can act as an inhibitor of growth, but also as a

732

growth promoter (Fig. 4).

an

us

cr

ip t

724

733

Jasmonic acid can also induce senescence. In Arabidopsis, jasmonate-induced

735

senescence involves induction of the transcription factor WRKY57, which is

736

repressed by the growth hormone auxin [159]. In addition, jasmonate induces

737

expression of ICE (Inducer of CBF Expression), thereby promoting freezing tolerance

738

in Arabidopsis [160]. Ethylene and jasmonate can regulate each other’s homeostasis

739

via feedback regulation, producing a fine balance between growth repression

743

Ac ce p

te

d

M

734

744

Cytokinins counteract the effect of ethylene, preventing senescence and stimulating

745

sugar metabolism and sink strength. Over-expression of the cytokinin biosynthetic

746

gene isopentenyl transferase has been used to produce plants that show delayed

747

senescence (stay-green trait), increased biomass production and improved stress

748

tolerance [45, 131]. However, the stay-green trait is not always associated with

740 741 742

(jasmonic acid) and growth stimulation (ethylene).

4.3. Growth stimulation responses The growth hormone group of cytokinins plays a role in controlling cell division.

Page 30 of 65

31 749

increased yield and productivity [131], suggesting that increased cytokinin levels

750

benefit vegetative growth but not reproductive development.

751 Gibberellins (GA) play a crucial role in the promotion of plant elongation growth. In

753

the absence of GA, elongation growth is restrained by DELLA nuclear proteins. In the

754

presence of GA, DELLA proteins bind to the GA-GID1 receptor complex, targeting it

755

for degradation by the ubiquitin-26S proteasome and thereby activating GA signalling

756

and elongation growth [161]. Some DELLA mutants are unable to bind the GA-GID1

757

complex, causing it to escape proteasome degradation. This suppresses elongation

758

growth, causing a semi-dwarf phenotype. Other DELLA mutants abolish its

759

repression activity, resulting in a tall stature (slender); these mutants are also male

760

sterile, suggesting that DELLA proteins play a role in pollen development [162].

761

Mutations in GA biosynthesis genes and DELLA proteins with a semi-dwarf

762

phenotype increase yield in cereals and have formed the basis of the Green

763

Revolution [163]. However, some GA-insensitive dwarf mutants in wheat (reduced

764

height; Rht) also have reduced pollen viability, which has been associated with

769

Ac ce p

te

d

M

an

us

cr

ip t

752

770

CBF1 was shown to induce DELLA gene expression and activate GA catabolic genes,

771

causing growth repression [168, 169]. Stress-induced accumulation of ABA

772

antagonizes GA action by controlling DELLA activity [170]. In addition, DELLA

773

proteins play a role in mounting a protective response to oxidative stress [169, 171].

765 766 767 768

reduced tolerance to abiotic stresses such as heat and drought [162, 164]. Interestingly, the growth-stimulation of GA can be counteracted by environmental stresses and hormones such as ethylene and auxins, which affect the growth restraining activity of DELLA proteins [165-167]. In Arabidopsis, the CBF/DREB cold-inducible transcription factors activate cold acclimation and freezing tolerance.

Page 31 of 65

32 774

The DELLA proteins obviously form a hub of hormonal and environmental

775

interactions that determine continuation or repression of growth in function of

776

environmental cues [172].

ip t

777

4.4. An old legend born again: auxins

779

Auxins were the first plant hormone to be discovered. The growth-promoting

780

properties of auxins have gained increasing prominence in recent years because of

781

their role in regulating development and response to abiotic stress. Auxins are

782

synthesized in the shoot apical meristem and are transported to neighbouring tissues

783

and over longer distances using efflux carriers and polar transporters respectively.

784

Auxins stimulate root growth and other tissue-specific responses throughout the plant

785

such as leaf and fruit senescence [148]. In Arabidopsis this requires cross-talk with

786

jasmonic acid signalling and the transcription factor WRKY57 [161, 173, 174].

us

an

M

d

One of the oldest known effects of auxins is the control of apical dominance and

793

Ac ce p

788

te

787

cr

778

794

have a role in stamen development and are actively synthesized in the anther,

795

controlling pollen development and anther dehiscence [178]. Apical dominance may

796

play an important role in promoting reproductive development and grain yield in

797

cereals. Auxins are synthesized in the anthers towards maturity where they play a role

798

in anther senescence [178, 179]. At the start of reproductive development in cereals

789 790 791 792

shoot branching (tillering in cereals). Cutting the main stem of a plant removes the apical meristem where auxins are made, resulting in increased branching. The branching response involves interaction with strigolactone and requires adequate sugar supply to support axillary bud outgrowth [175, 176]. It has been demonstrated that auxin treatment improves fertility in heat-stressed barley plants [177]. Auxins

Page 32 of 65

33 the shoot apex where auxins are synthesized changes into a flowering meristem. It

800

then develops into a spike containing the reproductive organs. The presence of auxin

801

biosynthesis in the floral organs at this stage might signify that apical dominance is

802

controlled by the reproductive structures. This function may be essential to direct

803

resources to the reproductive structures for seed production rather than investing them

804

in further vegetative growth. Abiotic stresses in cereals cause pollen sterility in

805

sensitive lines, resulting in increased tillering after the stress period (Fig. 3). This may

806

reflect the loss in apical dominance as a result of pollen sterility. An intriguing aspect

807

of auxins is that they regulate some aspects of plant development such as lateral root

808

development synergistically with ethylene and other hormones, while for some

809

aspects both hormones act antagonistically [180]. Recent progress in understanding

810

plant hormone action is illustrating the complexity of cross-talk between different

811

plant hormones and the importance of controlling hormone homeostasis.

812

Understanding the intricacies of these interactions is important to unravel how genetic

813

variability in the network can affect how plants adapt to environmental change.

cr

us

an

M

d

te

818

Ac ce p

814

ip t

799

819

in photomorphogenesis. Phytochromes respond to darkness (etiolation) and changes

820

in the red to far-red light ratio, which warns the plant about competing vegetation

821

(shade avoidance) [158]. Phytochromes control many growth processes from seed

822

germination to reproductive development and they are well known to modulate biotic

823

and abiotic stress responses [181].The activated form of phytochrome moves to the

815 816 817

4.5. Coordination of environmental responses Progress made in unravelling how plants react to low light conditions provided a clue as to how environmental signals regulate plant growth. Phytochrome photoreceptors react to changes in the ratio between red and far-red light and play an important role

Page 33 of 65

34 nucleus and forms a complex with members of the basic helix-loop-helix (bHLH)

825

transcription factors, the Phytochrome Interacting Factors (PIF). PIFs interact with

826

DELLA proteins; in the absence of GA, DELLA proteins bind to PIFs, preventing

827

them from regulating their target genes. In the presence of GA, DELLA proteins are

828

degraded, PIFs become functional and elongation growth is activated [161, 182, 183]

829

(Fig. 4). PIF transcription factors also activate auxin biosynthesis and their own

830

expression is controlled by the circadian clock; the Arabidopsis bZIP transcription

831

factor Hy5 (Elongated Hypocotyl) promotes photomorphogenesis by antagonizing PIF

832

action and the RING-motif E3 ligase COP1 (Constitutive Morphogenic) inhibits Hy5.

833

This pathway also influences CBF function and freezing tolerance [184]. PIFs form in

834

combination with DELLA proteins a hub for the integration of signals from various

835

hormones [183, 185-187], day-length [184, 188], light quality [158, 184], as well as

836

sugars [189] (Fig. 4). Light quality is also important for the induction of CBF

837

transcription factors and activation of cold and frost tolerance [190, 191] and PIFs

838

play a role in regulating expression of DREB transcription factors that are required for

839

drought responses [183]. Low red to far-red ratios lead to increased levels of ethylene

844

Ac ce p

te

d

M

an

us

cr

ip t

824

845

converge into a single, complicated signalling hub that orchestrates plant growth

846

responses. This environmental response hub may explain why one abiotic stress can

847

improve the response to other stresses (section 3.4).

840 841 842 843

[158], which affects the stability of the DELLA proteins [165] (Fig. 4). Induction of the ERF transcription factor SUB1A under flooding conditions prevents elongation growth by increasing DELLA levels, thereby inhibiting GA-mediated elongation growth [37, 38]. The PIF transcription factors also mediate cross-talk with ROS signalling [193]. These findings demonstrate how different environmental stimuli

848

Page 34 of 65

35

5. Conclusions

850

In the last decade important progress has been made using model plants in

851

understanding how plants grow and develop and how they respond to changes in the

852

environment. This know-how is still fragmentary and needs to be extended to crop

853

plants. In important crop species such as cereals, it is important to maintain

854

productivity under abiotic stress conditions during the reproductive stage. Some crop

855

plants appear to overreact and switch to growth arrest too quickly, even when survival

856

is not immediately under threat. One way of improving stress tolerance and grain

857

productivity in cereals would be to increase the threshold level at which plants switch

858

from promotion to arrest of growth. At the vegetative stage, the stay-green trait has

859

achieved this by selecting for delayed senescence. Maybe an equivalent of the stay-

860

green trait is required to protect the reproductive stage and grain formation in cereals.

861

Seed production itself is a stress survival mechanism; seed can survive prolonged

862

stress conditions in dehydrated state and this guarantees the plant’s next generation.

863

This potential may have been lost from crop plants, but genetic diversity to

864

reintroduce this trait may still be available in breeding lines, landraces or wild

866 867 868 869

cr

us

an

M

d

te

Ac ce p

865

ip t

849

progenitor species. This material can also be used to further improve our understanding of the hormonal interactions that control growth.

A lesson could be learned from flooding tolerance research, which showed that both growth arrest and acceleration can be beneficial – depending on the circumstances.

870

This response requires ethylene and ERF transcription factors. The molecular basis of

871

how flooding tolerance interacts with the environmental response hub can serve as a

872

guideline for other abiotic stresses. Response to shading also shares some of the hub

873

components used by flooding stress. Importantly, the example of flooding stress

Page 35 of 65

36 indicates that avoidance and/or escape reactions should not necessarily be treated as

875

different or independent from true tolerance responses. The common denominator is

876

“growth regulation”. It is crucial that we learn to understand how plants regulate

877

growth in function of environmental restraints; this may lead to strategies to

878

manipulate the threshold levels to switch from growth arrest to maintenance of

879

growth. Crop plants such as cereals, combined with current technologies, can help us

880

to reach that level of understanding.

cr

ip t

874

us

881

Having a single regulatory hub to integrate all environmental responses and regulate

883

plant growth and development makes a lot of sense, but a lot of questions still need to

884

be answered. We need to get a better understanding about systemic signalling and the

885

relationship between vegetative and reproductive growth. During the stage of

886

flowering and seed production, a plant behaves quite differently from a plant during

887

vegetative growth. Even though there is a shared response system to the environment,

888

growth signals still need to be relayed to different plant parts and the effect in

889

different plant parts can be interpreted very differently. For instance, nitrogen

894

Ac ce p

te

d

M

an

882

895

environment. The use of Green Revolution genes in cereals has shown that reducing

896

stem elongation growth using semi-dwarf genes benefits grain yield, but there is a

897

trade-off in terms of abiotic stress tolerance and pollen fertility. Some semi-dwarf

898

mutations affect the function of DELLA proteins in the central hub controlling growth

890 891 892 893

application can stimulate vegetative growth and repress reproductive development. In cereals, grain yield depends on successful interaction between both vegetative and reproductive growth. While management, agronomy and breeding practices have focused a lot on the vegetative establishment phase of cereals, relatively little is known about the control of reproductive development and its interaction with the

Page 36 of 65

37 and abiotic stress responses (Fig. 4). This raises the question whether high yield and

900

high abiotic stress tolerance are compatible – or not. There is a strong need to fully

901

understand the function of the central environmental response hub (e.g., role of PIF

902

family members, identification of still unknown components), but the use of model

903

systems only may not allow us to achieve this and genetic variation in crop species

904

should be included in these studies. Analysing this genetic variation using new

905

generation genotyping and phenotyping technologies has vastly improved and

906

identification of candidate stress tolerance genes is made easier using genomics.

907

Proof-of-function transgenic approaches may also lead to identification of genes that

908

can be used for stress-proofing cereals.

M

909

an

us

cr

ip t

899

The technological revolution of the last decade has provided renewed hope for

911

improving abiotic stress tolerance in crops such as cereals, but it is clear that this

912

effort will increasingly require close interaction between plant scientists of different

913

disciplines, including bioinformaticians and engineers.

te

918

Ac ce p

914

d

910

919

of cited references in this review paper has been limited by journal policy. The author

920

apologises to those authors whose publications were not cited in this paper.

915 916 917

Acknowledgements

R.D. is supported by grants from the Grains Research and Development corporation (GRDC, grants CSP00130, CSP00143 and CSP00175). The author thanks Jane Edlington and Holly Staniford for their help in preparing the manuscript. The number

921

Page 37 of 65

38 922

References

923

1. FAO, How to feed the world in 2050, http://www.fao.org, 2009.

924

2. G. Edmeades, T. Fischer, D. Byerlee, Can we feed the world in 2050? In: ‘Food security from sustainable agriculture, Proceedings of the 15th Agronomy

926

Conference 2010, 2010, Lincoln, New Zealand, pp. 15-19.

cr

928

3. A.J. Hall, R.A. Richards, Prognosis for genetic improvement of yield potential

and water-limited yield of major grain crops, Field Crops Res. 143 (2013) 18-33.

us

927

ip t

925

4. R.A. Richards, J.R. Hunt, J.A. Kirkegaard, J.B. Passioura, Yield improvement and

930

adaptation of wheat to water-limited environments in Australia - a case study,

931

Crop Past. Sci. 65 (2014) 676-689.

935 936 937 938 939 940 941 942 943

M

6. L.A. Bravo, M. Griffith, Characterization of antifreeze activity in Antarctic plants,

d

934

adaptation in cereals, Crit. Rev. Plant Sci. 27 (2008) 377-412.

J. Exp. Bot. 56 (2005) 1189-1196.

te

933

5. J.L. Araus, G.A. Slafer, C. Royo, D. Serret, Breeding for yield potential and stress

7. D. Bartels, F. Salamini, Desiccation tolerance in the resurrection plant Craterostigma plantagineum. A contribution to the study of drought tolerance at

Ac ce p

932

an

929

the molecular level, Plant Physiol. 127 (2001) 1346-1353.

8. T.S. Gechev, C. Dinakar, M. Benina, V. Toneva, D. Bartels, Molecular mechanisms of desiccation tolerance in resurrection plants, Cell. Mol. Life Sci. 69 (2012) 3175-3186.

9. L. Wissler et al., Back to the sea twice: identifying candidate plant genes for molecular evolution to marine life, BMC Evol. Biol. 11 (2011) 8.

944

10. S. Shabala, Learning from halophytes: physiological basis and strategies to

945

improve abiotic stress tolerance in crops, Ann. Bot. 112 (2013) 1209-1221

Page 38 of 65

39

948 949 950 951 952

effector, Plant J. 76 (2013) 800-810. 12. H. Yoshida, Y. Nagato, Flower development in rice, J. Exp. Bot. 62 (2011) 47194730.

ip t

947

11. J.P.B. Lloyd, B. Davies, SMG1 is an ancient nonsense-mediated mRNA decay

13. M. Ciaffi, A.R. Paolacci, O.A. Tanzarella, E. Porceddu, Molecular aspects of flower development in grasses, Sex. Plant Reprod. 24 (2011) 247-282.

cr

946

14. K. Matsubara, K. Hori, E. Ogiso-Tanaka, M. Yano, Cloning of quantitative trait genes from rice reveals conservation and divergence of photoperiod flowering

954

pathways in Arabidopsis and rice, Front. Plant Sci. 5 (2014) 1-7.

958 959 960

an

16. M. Gollery, J. Harper, J. Cushman, T. Mittler, R. Mittler, POFs: what we don't know can hurt us, Trends Plant Sci. 12 (2007) 492-496. 17. B.L Gross, K.M. Olsen, Genetic perspectives on crop domestication, Trends Plant Sci. 15 (2010) 529-537.

18. R.S. Meyer, M.D. Purugganan, Evolution of crop species: genetics of

966

Ac ce p

961

M

957

obscure features. Genome Biol. 7 (2006) R57.

d

956

15. M. Gollery et al., What makes species unique? The contribution of proteins with

te

955

us

953

967

21. D. Koenig et al., Comparative transcriptomics reveals patterns of selection in

968

domesticated and wild tomato, Proc. Natl Acad. Sci. USA 110 (2013) E2655-

969

2662.

962 963 964 965

domestication and diversification. Nat. Rev. Genet. 14 (2013) 840-852.

19. K.M. Olsen, J.F. Wendel, A bountiful harvest: genomic insights into crop domestication phenotypes, Annu. Rev. Plant Biol. 64 (2013) 47-70.

20. H. Sakai, T. Itoh, Massive gene losses in Asian cultivated rice unveiled by comparative genome analysis, BMC Genom. 11 (2010) 121-134.

Page 39 of 65

40 22. A.J. Cortés, D. This, C. Chavarro, S. Madriñán, M.W. Blair, Nucleotide diversity

971

patterns at the drought-related DREB2 encoding genes in wild and cultivated

972

common bean (Phaseolus vulgaris L.), Theor. Appl. Genet. 125 (2012) 1069-

973

1085.

974

ip t

970

23. B.J. Atwell, H. Wang, A.P. Scafaro, Could abiotic stress tolerance in wild

relatives of rice be used to improve Oryza sativa? Plant Sci. 215-216 (2014) 48-

976

58.

981 982 983 984

us

wheat wild relatives and landraces, J. Exp. Bot. 58 (2007) 177-186. 26. .K. Mochida, K. Shinozaki K., Unlocking Triticeae genomics to sustainably feed the future, Plant Cell Physiol. 54 (2013) 1931-1950. 27. H. Poorter et al., The art of growing plants for experimental purposes: a practical guide for the plant biologist, Func. Plant Biol. 39 (2012) 821-838. 28. R.J. Downs, H. Hellmers, Environment and the experimental control of plant

990

Ac ce p

985

an

980

25. M. Reynolds, F. Dreccer, R. Trethowan, Drought-adaptive traits derived from

M

979

utilization, J. Biosci. 37 (2012) 843-855.

d

978

24. B.M. Prasanna, Diversity in global maize germplasm: characterization and

te

977

cr

975

991

30. J.B. Passioura, the perils of pot experiments, Func. Plant Biol. 33 (2006) 1075-

986 987 988 989

992

growth, in: J.F. Sutcliffe, P. Mahlburg, Experimental Botany Vol. 6, 1975, Academic Press, London-New York-San Francisco, pp. 125-140.

29. I.G. Cummings, J.B. Reid, A. Koutoulis, Red to far-red ratio correction in plant growth chambers - growth responses and influence of thermal load on garden pea, Physiol. Plant. 131 (2007) 171-179.

1079.

Page 40 of 65

41 993

31. H. Poorter, J. Climent, D. Van Dusschoten, J. Bühler, J. Postma, Pot size matters:

994

a meta-analysis on the effect of rooting volume on plant growth, Func. Plant Biol.

995

39 (2012) 839-850. 32. D. Fleury, S. Jefferies, H. Kuchel, P. Langridge, Genetic and genomic tools to

ip t

996

improve drought tolerance in wheat, J. Exp. Bot. 61 (2010) 3211-3222.

998

33. R.A. Richards et al., Breeding for improved water productivity in temperate

1000

cereals: phenotyping, quantitative trait loci, markers and the selection environment, Func. Plant Biol. 37 (2010) 85-97.

us

999

cr

997

34. A. Greenup, W.J. Peacock, E.S. Dennis, B. Trevaskis, The molecular biology of

1002

seasonal flowering-responses in Arabidopsis and the cereals, Ann. Bot. 103

1003

(2009) 1165-1172.

M

1004

an

1001

35. B. Zheng, B. Biddulph, D. Li, H. Kuchel, S. Chapman, Quantification of the effects of VRN1 and Ppd-D1 to predict spring wheat (Triticum aestivum) heading

1006

time across diverse environments, J. Exp. Bot. 64 (2013) 3747-3761. 36. P.A. Riffkin, P.M. Evans, J.F. Chin, G.A. Kearney, Early-maturing spring wheat outperforms late-maturing winter wheat in the high rainfall environment of south-

1013

Ac ce p

1008

te

1007

d

1005

1014

39. R.A. Fischer, Understanding the physiological basis of yield potential in wheat, J.

1009 1010 1011 1012

1015 1016 1017

western Victoria. Austral. J. Agric. Res. 54 (2003) 193-202.

37. J. Bailey-Serres, L.A. Voesenek, Life in the balance: a signaling network controlling survival of flooding, Curr. Opin. Plant Biol. 13 (2010) 489-494.

38. J. Bailey-Serres et al., Making sense of low oxygen sensing, Trends Plant Sci. 17 (2012) 129-138.

Agric. Sci. 145 (2007) 99-113. 40. R.A. Fischer, Wheat physiology: a review of recent developments, Crop Pasture Sci. 62 (2011) 95-114.

Page 41 of 65

42 1018

41. N.C. Collins, F. Tardieu, R. Tuberosa, Quantitative trait loci and crop

1019

performance under abiotic stress: where do we stand? Plant Physiol. 147 (2008)

1020

469-486.

1023

ip t

1022

42. M.M. Chaves, J.P. Maroco, J.S. Pereira, Understanding plant responses to drought - from genes to the whole plant, Func. Plant Biol. 30 (2003) 239-264.

43. A. Guiboileau, R. Sormani, C. Meyer, C. Masclaux-Daubresse, Senescence and

cr

1021

death of plant organs: Nutrient recycling and developmental regulation, Comptes

1025

Rendus Biologies 333 (2010) 382-391.

44. H. Sprigg, R. Belford, S. Milroy, S.J. Bennett, D. Bowran, Adaptations for

an

1026

us

1024

growing wheat in the drying climate of Western Australia, Crop Past. Sci. 65

1028

(2014) 627-644.

M

1027

45. H. Thomas, H. Ougham, The stay-green trait, J. Exp. Bot. 65 (2014) 3889-3900.

1030

46. R.M. Rivero et al., Delayed leaf senescence induces extreme drought tolerance in

1032

47. X.R.R. Sirault, A.G. Condon, G.J. Rebetzke, G.D. Farquhar, Genetic analysis of leaf rolling in wheat, in: R. Appels, R. Eastwood, E. Lagudah, P. Langridge, M.M.

1038

Ac ce p

1033

a flowering plant, Proc. Natl Acad. Sci. USA 104 (2007) 19631-19636.

te

1031

d

1029

1039

49. J.M. Morgan, Osmoregulation and water stress in higher plants, Annu. Rev. Plant

1034 1035 1036 1037

1040

Lynne (Eds.), Proceedings of the 11th International Wheat Genetic Symposium, Brisbane, Australia.

48. S. Bunnag, P. Pongthai, Selection of rice (Oryza sativa L.) cultivars tolerant to drought stress at the vegetative stage under field conditions, Am. J. Plant Sci. 4 (2013) 1701-1708.

Physiol. 35 (1984) 299-319.

Page 42 of 65

43 1041

50. J.M. Morgan, Growth and yield of wheat lines with differing osmoregulative

1042

capacity at high soil water deficit in seasons of varying evaporative demand, Field

1043

Crops Res. 40 (1995) 143-152. 51. T.C. Hsiao, J.C. O'Toole, E.B. Yambao, N.C. Turner, Influence of osmotic

ip t

1044

adjustment on leaf rolling and tissue death in rice (Oryza sativa L.), Plant Physiol.

1046

75 (1984) 338-341.

cr

1045

52. R.C. Babu, M.S. Pathan, A. Blum, H.T. Nguyen, Comparison of measurement

1048

methods of osmotic adjustment in rice cultivars, Crop Sci. 39 (1999) 150-158.

1049

53. H.S. Saini, D. Aspinall, Sterility in wheat (Triticum aestivum L.) induced by water

an

us

1047

deficit or high temperature: possible mediation by abscisic acid. Austral. J. Plant

1051

Physiol. 9 (1982) 529-537.

1053

54. N. Powell, X. Ji, R. Ravash, J. Edlington, R. Dolferus, Yield stability for cereals in a changing climate, Func. Plant Biol. 39 (2012) 539-552.

d

1052

M

1050

55. S.N. Oliver et al., Cold-induced repression of the rice anther-specific cell wall

1055

invertase gene OSINV4 is correlated with sucrose accumulation and pollen

1056

sterility, Plant Cell Envir. 28, (2005) 1534-1551.

1058 1059 1060 1061 1062 1063 1064

Ac ce p

1057

te

1054

56. X. Ji et al., Importance of pre-anthesis anther sink strength for maintenance of grain number during reproductive stage water stress in wheat, Plant Cell Envir. 33 (2010) 926-942.

57. S.N. Oliver, E.S. Dennis, R. Dolferus, ABA regulates apoplastic sugar transport and is a potential signal for cold-induced pollen sterility in rice, Plant Cell Physiol. 48 (2007) 1319-1330. 58. X. Ji et al., Control of ABA catabolism and ABA homeostasis is important for reproductive stage stress tolerance in cereals, Plant Physiol. 156 (2011) 647-662.

Page 43 of 65

44 1065

59. D. Jaccoud, K. Peng, D. Feinstein, A. Kilian, Diversity arrays: a solid state

1066

technology for sequence information independent genotyping, Nucl. Ac. Res. 29

1067

(2001) E25.

1070

ip t

1069

60. J. Mammadov, R. Aggarwal, R. Buyyarapu, S. Kumpatla, SNP markers and their impact on plant breeding, Int. J. Plant Genom. 2012 (2012) 728398.

61. K.L. McNally et al., Genomewide SNP variation reveals relationships among

cr

1068

landraces and modern varieties of rice, Proc. Natl Acad. Sci. U.S.A 106 (2009)

1072

12273-12278.

1075

an

1074

62. M.W. Ganal et al., Large SNP arrays for genotyping in crop plants, J. Biosci. 37 (2012) 821-828.

63. J. Spindel et al, Bridging the genotyping gap: using genotyping by sequencing

M

1073

us

1071

(GBS) to add high-density SNP markers and new value to traditional bi-parental

1077

mapping and breeding populations, Theor. Appl. Genet. 126 (2013) 2699-2716.

1080 1081 1082 1083 1084 1085

te

1079

64. P.K. Gupta, P.L. Kulwal, V. Jaiswal, Association mapping in crop plants: opportunities and challenges, Adv. Genet. 85 (2014) 109-147. 65. B.E. Huang et al., A multiparent advanced generation inter-cross population for

Ac ce p

1078

d

1076

genetic analysis in wheat, Plant Biotechnol. J. 10 (2012) 826-839.

66. A. Nakaya, S.N. Isobe, Will genomic selection be a practical method for plant breeding? Ann. Bot. 110 (2012) 1303-1316.

67. J.N. Cobb, G. De Clerck, A. Greenberg, R. Clark, S. McCouch, Next-generation phenotyping: requirements and strategies for enhancing our understanding of

1086

genotype-phenotype relationships and its relevance to crop improvement, Theor.

1087

Appl. Genet. 126 (2013) 867-887.

Page 44 of 65

45 1088

68. S. Negrão, B. Courtois, N. Ahmadi, I. Abreu, N. Saibo, M. Oliveira, Recent

1089

updates on salinity stress in rice: from physiological to molecular responses, Crit.

1090

Rev. Plant Sci. 30 (2011) 329-377. 69. J. Feng et al., Characterization of metabolite quantitative trait loci and metabolic

1092

networks that control glucosinolate concentration in the seeds and leaves of

1093

Brassica napus, New Phytol. 193 (2012) 96-108.

cr

1094

ip t

1091

70. N. Carreno-Quitero et al., Untargeted metabolic quantitative trait loci analyses reveal a relationship between primary metabolism and potato tuber quality, Plant

1096

Physiol. 158 (2012) 1306-1318.

an

us

1095

71. Y. Wahyuni et al., Genetic mapping of semi-polar metabolites in pepper fruits

1098

(Capsicum sp.): towards unraveling the molecular regulation of flavonoid

1099

quantitative trait loci, Mol. Breeding 33 (2014) 503-518. 72. U. Roessner et al., Metabolic profiling allows comprehensive phenotyping of

d

1100

M

1097

genetically or environmentally modified plant systems, Plant Cell 13 (2001) 11-

1102

29.

73. C. Caldana et al., High density kinetic analysis of the metabolomics and

1108

Ac ce p

1103

te

1101

1109

75. I.A. Abreu et al., Coping with abiotic stress: Proteome changes for crop

1104 1105 1106 1107

1110 1111 1112

transcriptomic response of Arabidopsis to eight environmental conditions, Plant J. 67 (2011) 869-884.

74. C.B. Hill et al., Whole-genome mapping of agronomic and metabolomics trait to identify novel quantitative trait loci in bread wheat grown in a water-limited environment, Plant Physiol. 162 (2013) 1266-1281.

improvement, J. Prot. 93 (2013) 145-168. 76. F. Fiorani, U. Schurr, Future scenarios for plant phenotyping, Annu. Rev. Plant Biol. 64 (2013) 267-291.

Page 45 of 65

46 1113

77. N. Honsdorf, T.J. March, B. Berger, M. Tester, K. Pillen, High-throughput

1114

phenotyping to detect drought tolerance QTL in wild barley introgression lines,

1115

Plos One 9 (2014) e97047.

1117

78. B. Masuka, J.L. Araus, B. Das, K. Sonder, J.E. Cairns, Phenotyping for abiotic stress tolerance in maize, J. Integr. Plant Biol. 54 (2012) 238-249.

ip t

1116

79. A. Ballvora et al., “Deep phenotyping” of early plant response to abiotic stress

1119

using non-invasive approaches in barley, in: G. Zhang et al. (Eds.), Advance in

1120

barley sciences: proceeding of 11th International Barley Genetics Symposium,

1121

Chapter 26, Zhejiang University and Springer Science, Dordrecht, pp. 317-326.

an

us

cr

1118

80. O. Morozova, M.A. Marra, Applications of next-generation sequencing

1123

technologies in functional genomics, Genomics 92 (2008) 255-264.

1124

M

1122

81. J. Kilian, F. Peschke, K.W. Berendzen, K. Harter, D. Wanke, Prerequisites, performance and profits of transcriptional profiling the abiotic stress response,

1126

Biochim. Biophys. Acta 1819 (2012) 166-175. 82. K. Horan et al., Annotating genes of known and unknown function by large-scale coexpression analysis, Plant Physiol. 147 (2008) 41-57.

1133

Ac ce p

1128

te

1127

d

1125

1134

Genet. 7 (2011) e1002020.

1129 1130 1131 1132

1135 1136

83. N. Wang, T. Long, W. Yao, L. Xiong, Q. Zhang, C. Wu, Mutant resources for the functional analysis of the rice genome, Mol. Plant 6 (2013) 596-604.

84. G.R. Cramer, K. Urano, S. Delrot, M. Pezzotti, K. Shinozaki, Effects of abiotic stress on plants: a systems biology perspective, BMC Plant Biol. 11 (2011) 163.

85. Y-S. Seo et al., Towards establishment of a rice stress response interactome, PLoS

86. D. Kültz, Molecular and evolutionary basis of the cellular stress response, Ann. Rev. Physiol. 67 (2005) 225-257.

Page 46 of 65

47 1137

87. P. Castiglioni et al., Bacterial RNA chaperones confer abiotic stress tolerance in

1138

plants and improved grain yield in maize under water-limited conditions. Plant

1139

Physiol. 147 (2008) 446-455. 88. F.J. Schmitt et al., Reactive oxygen species: re-evaluation of generation,

ip t

1140

monitoring and role in stress-signaling in phototrophic organisms, Biochim.

1142

Biophys. Acta. 1837 (2014) 835-848.

1145

Ann. Bot. 99 (2007) 3-8.

us

1144

89. S. Debolt, V. Melino, C.M. Ford, Ascorbate as a biosynthetic precursor in plants,

90. A. Nunes-Nesi, R. Sulpice, Y. Gibon, A.R. Fernie, The enigmatic contribution of

an

1143

cr

1141

mitochondrial function in photosynthesis, J. Exp. Bot. 59, 1675-1684.

1147

91. Z. Chen, D.R. Gallie, Dehydroascorbate reductase affects leaf growth,

1149

development and function, Plant Physiol. 142 (2006) 775-787. 92. E. Olmos, G. Kiddle, T.K. Pellny, S. Kumar, C.H. Foyer, Modulation of plant

d

1148

M

1146

morphology, root architecture, and cell structure by low vitamin C in Arabidopsis

1151

thaliana, J. Exp. Bot. 57 (2006) 1645-1655.

1153 1154 1155 1156 1157

93. G. Noctor, A. Mhamdi, C.H. Foyer, The roles of reactive oxygen metabolism in

Ac ce p

1152

te

1150

drought: not so cut and dried, Plant Physiol. 164 (2014) 1636-1648.

94. M. Fujita et al., Crosstalk between abiotic and biotic stress responses: a current view from the points of convergence in the stress signaling networks, Curr. Opin. Plant Biol. 9 (2006) 436-442.

95. G.M. Pastori, C.H. Foyer, Common components, networks, and pathways of

1158

cross-tolerance to stress. The central role of “redox” and abscisic acid-mediated

1159

controls, Plant Physiol. 129 (2002) 460-468.

1160 1161

96. G. Kocsy et al., Redox control of plant growth and development, Plant Sci. 211 (2013) 77-91.

Page 47 of 65

48 1162

97. D. Hérouart, M. Van Montagu, D. Inzé, Developmental and environmental

1163

regulation of the Nicotiana plumbaginifolia cytosolic Cu/Zn-superoxide dismutase

1164

promoter in transgenic tobacco, Plant Physiol. 104 (1994) 873-880.

1168 1169

ip t

1167

and auxin perspective, Plant Cell Environ. 35 (2012) 321-333.

99. A. Baxter, R. Mittler, N. Suzuki, ROS as key players in plant stress signaling, J.

cr

1166

98. V.B. Tognetti, P. Mühlenbock, F. Van Breusegem, Stress homeostasis - the redox

Exp. Bot. 65 (2014) 1229-1240. 100.

A. Krishnamurthy, B. Rathinasabapathi, Oxidative stress tolerance in plants -

us

1165

novel interplay between auxin and reactive oxygen species signaling, Plant Signal.

1171

Behav. 8 (2013) e25761-1 - e25761-5. 101.

B. Ye, J. Gressel, Transient, oxidant-induced antioxidant transcript and

M

1172

an

1170

enzyme levels correlate with greater oxidant-resistance in paraquat-resistant

1174

Conyza bonariensis, Planta 211 (2000) 50-61. 102.

Y. Shaaltiel, A. Glazer, P.F. Bocion, J. Gressel, Cross tolerance to herbicidal

te

1175

d

1173

and environmental oxidants of plant biotypes tolerant to paraquat, sulfur dioxide

1177

and ozone, Pest. Biochem. Physiol. 31 (1988) 13-23.

1178 1179 1180 1181 1182 1183 1184

Ac ce p

1176

103.

A. Altinkut, K. Kazan, Z. Ipekci, N. Gozukirmizi, Tolerance to paraquat is

correlated with the traits associated with water stress tolerance in segregating F2 populations of barley and wheat, Euphytica 121 (2001) 81-86.

104.

H.R. Lascano, M.N. Melchiorre, C.M. Luna, V.S. Trippi, Effect of photo-

oxidative stress induced by paraquat in two wheat cultivars with differential tolerance to water stress, Plant Sci. 164 (2003) 841-848.

105.

Y.H. Kim et al., Overexpression of sweet potato swpa4 peroxidase results in

1185

increased hydrogen peroxide production and enhances stress tolerance in tobacco,

1186

Planta 227 (2008) 867-881.

Page 48 of 65

49 1187

106.

T. Matsumura, N. Tabayashi, Y. Kamagata, C. Souma, H. Saruyama, Wheat

1188

catalase expressed in transgenic rice can improve tolerance against low

1189

temperature stress, Physiol. Plant. 1167 (2002) 317-327. 107.

Y.J. Im, M. Ji, A. Lee, R. Killens, A.M. Grunden, W.F. Boss, Expression of

ip t

1190

Pyrococcus furiosus superoxide reductase in Arabidopsis Enhances heat tolerance,

1192

Plant Physiol. 151 (2009) 893-904.

1193

108.

cr

1191

S.R. Prashanth, V. Sadhasivam, A. Parida, Over expression of cytosolic

copper/zinc superoxide dismutase from a mangrove plant Avicennia marina in

1195

indica Rice var Pusa Basmati-1 confers abiotic stress tolerance, Transgenic Res.

1196

17 (2008) 281-291.

an

109.

C. Huang, W. He, J. Guo, X. Chang, P. Su, L. Zhang, Increased sensitivity to

M

1197

us

1194

salt stress in an ascorbate-deficient Arabidopsis mutant, J. Exp. Bot. 56 (2005)

1199

3041-3049.

1202 1203 1204 1205 1206 1207

S. Echevarría-Zomeño et al., Regulation of translation initiation under biotic

te

1201

110.

and abiotic stresses, Int. J. Mol. Sci. 14 (2013) 4670-4683. 111.

J.X. Liu, S.H. Howell, Endoplasmic reticulum protein quality control and its

Ac ce p

1200

d

1198

relationship to environmental stress responses in plants, Plant Cell 22 (2010) 2930-2942.

112.

Y.L. Ruan, Sucrose metabolism: gateway to diverse carbon use and sugar

signaling, Annu. Rev. Plant Biol. 65 (2014) 33-67.

113.

N. Sreenivasulu, V.T. Harshavardhan, G. Govind, C. Seiler, A. Kohli,

1208

Contrapuntal role of ABA: does it mediate stress tolerance or plant growth

1209

retardation under long-term drought stress? Gene 506 (2012) 265-273.

Page 49 of 65

50 1210

114.

A. Wingler, T. Roitsch, Metabolic regulation of leaf senescence: interactions

1211

of sugar signalling with biotic and abiotic stress responses, Plant Biol. 10 Suppl 1

1212

(2008) 50-62.

1216 1217 1218 1219

ip t

networks controllng plant growth, Curr. Opin. Plant Biol. 13 (2010) 273-278. 116.

C. Jonak, L. Okrész , L. Bögre, H. Hirt, Complexity, cross talk and integration

cr

1215

S. Smeekens, J. Ma, J. Hanson, F. Rolland, Sugar signals and molecular

of plant MAP kinase signaling, Curr. Opin. Plant Biol. 5 (2002) 415-424. 117.

B. Wurzinger, A. Mair, B. Pfister, M. Teige, Cross-talk of calcium-dependent

us

1214

115.

protein kinase and MAP kinase signaling, Plant Signal. Behav. 6 (2011) 8-12. 118.

an

1213

V. Chinnusamy, K. Schumaker, J.K. Zhu, Molecular genetic perspectives on

cross-talk and specificity in abiotic stress signalling in plants, J. Exp. Bot. 55

1221

(2004) 225-236. 119.

H. Ji, J.M. Pardo, G. Batelli, M.J. Van Oosten, R.A. Bressan, X. Li, The Salt

d

1222

M

1220

Overly Sensitive (SOS) pathway: established and emerging roles, Mol. Plant 6

1224

(2013) 275-286.

1226 1227 1228 1229 1230 1231 1232 1233

120.

J. Martínez-Atienza et al., Conservation of the salt overly sensitive pathway in

Ac ce p

1225

te

1223

rice, Plant Physiol. 143 (2007) 1001-1012.

121.

E. Baena-Gonsalez, F. Rolland, J.M. Thevelein, J. Sheen, A central integrator

of transcription networks in plant stress and energy signaling, Nature 448 (2007) 938-943.

122.

J. Hanson, J. Smeekens, Sugar perception and signaling - an update, Curr.

Opin. Plant Biol. 12 (2009) 562-567. 123.

S.J. Hey, E. Byrne, N.G. Halford, The interface between metabolic and stress

signaling, Ann. Bot. 105 (2010) 197-203.

Page 50 of 65

51 1234

124.

E.H. Byrne et al., Overexpression of GCN2-type protein kinase in wheat has

1235

profound effects on free amino acid concentration and gene expression, Plant

1236

Biotech. J. 10 (2012) 328-340.

1240 1241

controlling cross-tolerance, Trends Plant Sci. 5 (2000) 241-246. 126.

R. Mittler, Abiotic stress, the field environment and stress combination,

Trends Plant Sci. 11 (2006) 15-19. 127.

ip t

1239

C. Bowler, R. Fluhr, The role of calcium and activated oxygens as signals for

cr

1238

125.

P.C. Larosa, A.K. Handa, P.M. Hasegawa, R.A. Bressan, Abscisic acid

us

1237

accelerates adaptation of cultured tobacco cells to salt, Plant Physiol. 79 (1985)

1243

138-142. 128.

A.J. Robertson, M. Ishikawa, L.V. Gusta, S.L. MacKenzie, Abscisic acid-

M

1244

an

1242

induced heat tolerance in Bromus inermis Leyss cell-suspension cultures. Heat-

1246

stable, abscisic acid-responsive polypeptides in combination with sucrose confer

1247

enhanced thermostability, Plant Physiol. 105 (1994) 181-190. S. Lu, W. Su, H. Li, Z. Guo, Abscisic acid improves drought tolerance of

triploid Bermuda grass and involves H2O2- and NO-induced antioxidant enzyme

1254

Ac ce p

1249

129.

te

1248

d

1245

1255

132.

1250 1251 1252 1253

1256

activities, Plant Physiol. Biochem. 47 (2009) 132-138.

130.

D.R. Kelley, M. Estelle, Ubiquitin-mediated control of plant hormone

signaling, Plant Physiol. 160 (2012) 47-55.

131.

P.L. Gregersen, A. Culetic, L. Boschian, K. Krupinska, Plant senescence and

crop productivity, Plant Mol. Biol. 82 (2013) 603-622. K. Harris et al., Sorghum stay-green QTL individually reduce post-flowering

drought-induced leaf senescence, J. Exp. Bot. 58 (2007) 327-338.

Page 51 of 65

52 1257

133.

K.S.V. Jagadish, J.E. Cairns, A. Kumar, I.M. Somayanda, P.Q. Craufurd,

1258

Does susceptibility to heat stress confound screening for drought tolerance in rice?

1259

Func. Plant Biol. 38 (2011) 261-269.

1262

P. Ahmad et al., Role of transgenic plants in agriculture and biopharming,

ip t

1261

134.

Biotechnol. Adv. 30 (2012) 524-530. 135.

J. Deikman, M. Petracek, J.E. Heard, Drought tolerance through

cr

1260

biotechnology: improving translation from the laboratory to farmers' fields, Curr.

1264

Opin. Biotechnol. 23 (2012) 243-250. 136.

Z. Peleg, M.P. Apse, E. Blumwald, Engineering salinity and water-stress

an

1265

us

1263

tolerance in crop plants: getting closer to the field, Adv. Bot. Res. 57 (2011) 405-

1267

432.

1268

137.

M

1266

M. Reguera, Z. Peleg, E. Blumwald, targeting metabolic pathways for genetic

engineering abiotic stress-tolerance in crops, Biochim. Biophys. Acta 1819 (2012)

1270

186-194.

S.J. Oh, C.W. Kwon, D.W. Choi, S.I. Song, J.K. Kim, Expression of barley

HvCBF4 enhances tolerance to abiotic stress in transgenic rice. Plant Biotech. J. 5

1277

Ac ce p

1272

138.

te

1271

d

1269

1278

140.

1273 1274 1275 1276

(2007) 646-656.

139.

M. Kasuga, S. Miura, K. Shinozaki, K. Yamaguchi-Shinozaki, A combination

of the Arabidopsis DREB1A gene and stress-inducible rd29A promoter improved drought- and low-temperature stress tolerance in tobacco by gene transfer. Plant Cell Physiol. 45 (2004) 346-350. M. Ayada et al., Functional analysis of the durum wheat gene TdPIP2;1 and its

1279

promoter region in response to abiotic stress in rice, Plant Physiol. Biochem. 79

1280

(2014) 98-108.

Page 52 of 65

53 1281 1282 1283

141.

S.J. Oh et al., Overexpression of the transcription factor AP37 in rice improves

grain yield under drought conditions. Plant Physiol. 150 (2009) 1368-1379. 142.

D.L. Lawlor, Genetic engineering to improve plant performance under

drought: physiological evaluation of achievements, limitations, and possibilities, J.

1285

Exp. Bot. 64 (2013) 83-108. 143.

Z. Peleg, M. Reguera, E. Tumimbang, H. Walia, E. Blumwald, Cytokinin-

cr

1286

ip t

1284

mediated source/sink modifications improve drought tolerance and increase grain

1288

yield in rice under water-stress, Plant Biotechnol J. 9 (2011) 747-758.

1292 1293

an

1291

A. Alpi et al., Plant neurobiology: no brain, no gain? Trends Plant Sci. 12

(2007) 135-136. 145.

K. Brackmann, T. Greb, Long- and short-distance signaling in the regulation

M

1290

144.

of lateral plant growth, Physiol. Plant. 151 (2013) 134-141. 146.

A. Endo et al., Drought induction of Arabidopsis 9-cis-epoxycarotenoid

d

1289

us

1287

dioxygenase occurs in vascular parenchyma cells, Plant Physiol. 147 (2008) 1984-

1295

1993.

1297 1298 1299 1300 1301 1302 1303 1304

147.

R. Spicer, Symplasmic networks in secondary vascular tissues: parenchyma

Ac ce p

1296

te

1294

distribution and activity supporting long-distance transport, J. Exp. Bot. 65 (2014) 1829-1848.

148.

S. Vanneste, J. Friml, Auxin: a trigger for change in plant development, Cell

136 (2009) 1005-1016.

149.

Y. Fridman, S. Savaldi-Goldstein, Brassinosteroids in growth control: how,

when and where, Plant Sci. 209 (2013) 24-31. 150.

P.B. Brewer, H. Koltai, C.A. Beveridge, Diverse roles of strigolactones in

plant development, Mol. Plant. 6 (2013) 18-28.

Page 53 of 65

54 1305 1306 1307

151.

Z.Q. Fu, X. Dong, Systemic acquired resistance: turning local infection into

global defense, Annu. Rev. Plant Biol. 64 (2013) 839-863. 152.

P. Landi et al., Root-ABA1 QTL effects root lodging, grain yield, and other

agronomic traits in maize grown under well-watered and water-stressed

1309

conditions, J. Exp. Bot. 58 (2006) 319-326.

1314 1315 1316 1317 1318

cr

R.W. Parish, H.S. Phan, S. Iacunone, S.F. Li, Tapetal development and abiotic

us

154.

stress: a centre of vulnerability, Func. Plant Biol. 39 (2012) 553-559. 155.

an

1313

(2007) 115-136.

S. Gepstein, B.R. Glick, Strategies to ameliorate abiotic stress-induced plant

senescence, Plant Mol. Biol. 82 (2013) 623-633. 156.

M

1312

P.O. Lim, H.J. Kim, H.G. Nam, Leaf senescence, Ann. Rev. Plant Biol. 58

A.N. Stepanova, J.M. Alonso, Ethylene signaling and response: where

different regulatory modules meet, Curr. Opin. Plant Biol. 12 (2009) 548-555. 157.

d

1311

153.

R. Pierik, D. Tholen, H. Poorter, E.J.W. Visser, L.A.C.J. Voesenek, The Janus

te

1310

ip t

1308

face of ethylene: growth inhibition and stimulation, Trends Plant Sci. 11 (2006)

1320

176-183.

1321 1322 1323 1324 1325

Ac ce p

1319

158.

K.A. Franklin, Shade avoidance, New Phytol. 179 (2008) 930-944.

159.

Y. Jiang, G. Liang, S. Yang, D. Yu, Arabidopsis WRKY57 functions as a node

of convergence for jasmonic acid- and auxin-mediated signaling in jasmonic acidinduced leaf senescence, Plant Cell 26 (2014) 230-245.

160.

Y. Hu, L. Jiang, F. Wang, D. Yua, Jasmonate regulates the INDUCER OF

1326

CBF

EXPRESSION-C-REPEAT

BINDING

FACTOR/DRE

BINDING

1327

FACTOR1 cascade and freezing tolerance in Arabidopsis, Plant Cell 25 (2013)

1328

2907-2924.

Page 54 of 65

55 1329 1330 1331

161.

H. Claeys, S. De Bodt, D. Inzé, Gibberellins and DELLAs: central nodes in

growth regulatory networks, Trends Plant Sci. 19 (2014) 231-239. 162.

A.R.G. Plackett et al., DELLA activity is required for successful pollen

development in the Columbia ecotype of Arabidopsis, New Phytol. 210 (2014)

1333

825-836.

ip t

1332

163.

P. Hedden, The genes of the Green Revolution, Trends Genet. 19 (2003) 5-9.

1335

164.

F. Alghabari, M. Lukac, H.E. Jones, M.J. Gooding, Effect of Rht alleles on the

cr

1334

tolerance of wheat grain set to high temperature and drought stress during booting

1337

and anthesis, J. Agron. Crop Sci. 200 (2014) 36-45.

1340

an

1339

165.

P. Achard et al., Integration of plant responses to environmentally activated

phytohormonal signals, Science 311 (2006) 91-94. 166.

M

1338

us

1336

E.H. Colebrook, S.G. Thomas, A.L. Phillips, P. Hedden, The role of

gibberellin signalling in plant responses to abiotic stress, J. Exp. Biol. 217 (2014)

1342

67-75.

1345 1346 1347 1348 1349 1350 1351

te

1344

167.

X. Fu, N.P. Harberd, Auxin promotes Arabidopsis root growth by modulating

gibberellin response, Nature 421 (2003) 740-743.

Ac ce p

1343

d

1341

168.

P. Achard et al., The cold-inducible CBF1 factor-dependent signalling

pathway modulates the accumulation of the growth-repressing DELLA proteins via its effect on gibberellin metabolism, Plant Cell 20 (2008) 2117-2129.

169.

L.V. Kurepin et al., Role of CBFs as integrators of chloroplast Redox,

phytochrome and plant hormone signalling during cold acclimation, Int. J. Mol. Sci. 14 (2013) 12729-12763.

170.

D. Golldack, C. Li, H. Mohan, N. Probst, Gibberellins and abscisic acid signal

1352

crosstalk: living and developing under unfavorable conditions, Plant Cell Rep. 32

1353

(2013) 1007-1016.

Page 55 of 65

56 1354

171.

P. Achard, J.P. Renou, R. Berthome, N.P. Harberd, P. Genschik, Plant

1355

DELLAs retrain growth and promote survival of adversity by reducing the levels

1356

of reactive oxygen species, Cur. Biol. 18, 656-660. 172.

Y. Jiang, G, Liang, S. Yang, D. Yu, Arabidopsis WRKY57 functions as a node

ip t

1357

of convergence for jasmonic acid- and auxin-mediated signaling in jasmonic acid-

1359

induced leaf senescence, Plant Cell 26 (2014) 230-245.

1360

173.

cr

1358

Q. Ma et al., Comprehensive insights on how 2,4-dichlorophenoxyacetic acid

retards senescence in post-harvest citrus fruits using transcriptomic and proteomic

1362

approaches, J. Exp. Bot. 65 (2014) 61-74. 174.

an

1363

us

1361

C. Böttcher, C.A. Burbidge, P.K. Boss, C. Davies, Interactions between

ethylene and auxin are crucial to the control of grape (Vitis vinifera L.) berry

1365

ripening, BMC Plant Biol. 13 (2013) 222.

1368 1369 1370 1371 1372 1373 1374 1375 1376

d

E.A. Dun, P.B. Brewer, C.A. Beveridge, Strigolactones: discovery of the

elusive shoot branching hormone, Trends Plant Sci. 14 (2009) 364-372. 176.

te

1367

175.

M.G. Mason, J.J. Ross, B.A. Babst, B.N. Wienclaw, C.A.Beveridge, Sugar

demand, not auxin, is the initial regulator of apical dominance, Proc. Natl Acad.

Ac ce p

1366

M

1364

Sci. USA. 111 (2014) 6092-6097.

177.

T. Sakata et al., Auxins reverse plant male sterility caused by high

temperatures, Proc. Natl Acad. Sci. USA 107 (2010) 8569-8574.

178.

V. Cecchetti, M.M. Altamura, G. Falasca, P. Costantino, M. Cardarelli, Auxin

regulates Arabidopsis anther dehiscence, pollen maturation, and filament elongation, Plant Cell 20 (2008) 1760-1774.

179.

K. Hirano et al., Comprehensive transcriptome analysis of phytohormone

1377

biosynthesis and signaling genes in microspore/pollen and tapetum of rice, Plant

1378

Cell Physiol. 49 (2008) 1429-1450.

Page 56 of 65

57

1380 1381 1382 1383

180.

F. Wang, X. Cui, Y. Sun, C.H. Dong, Ethylene signaling and regulation in

plant growth and stress responses, Plant Cell Rep. 32 (2013) 1099-1109. 181.

R.F. Carvalho, M.L. Campos, R.A. Azevedo, The role of phytochrome in

stress tolerance, J. Integr. Plant Biol. 53 (2011) 920-929. 182.

ip t

1379

A. Castillon, H. Shen, E. Huq, Phytochrome Interacting Factors: central

players in phytochrome-mediated light signaling networks, Trends Plant Sci. 12,

1385

514-521. 183.

P. Hornitschek et al., Phytochrome interacting factors 4 and 5 control seedling

us

1386

cr

1384

growth in changing light conditions by directly controlling auxin signaling, Plant

1388

J. 71 (2012) 699-711. 184.

P. Maibam et al., the influence of light quality, circadian rhythm, and

M

1389

an

1387

photoperiod on the CBF-mediated freezing tolerance, Int. J. Mol. Sci. 14 (2013)

1391

11527-11543. 185.

K.A. Franklin et al., Phytochrome-interacting factor 4 (PIF4) regulates auxin

te

1392

d

1390

biosynthesis at high temperature, Proc. Natl Acad. Sci. USA 108 (2011) 20231-

1394

2035.

1395 1396 1397 1398 1399 1400 1401 1402

Ac ce p

1393

186.

P. Leivar, P.H. Quail, PIFs: pivotal components in a cellular signaling hub,

Trends Plant Sci. 16 (2011) 19-28.

187.

S.L. Lau, X.W. Deng, Plant hormone signaling lights up: integrators of light

and hormones, Curr. Opin. Plant Biol. 13 (2010) 571-577.

188.

A. Sanchez, J. Shin, S.J. Davis, Abiotic stress and the circadian clock, Plant

Sign. Behav. 6 (2011) 223-231. 189.

I. Sairanen et al., Soluble carbohydrates regulate auxin biosynthesis via PIF

proteins in Arabidopsis, Plant Cell 24 (2012) 4907-4916.

Page 57 of 65

58 1403 1404 1405

190.

K.A. Franklin, G.C. Whitelam, Light-quality regulation of freezing tolerance

in Arabidopsis thaliana, Nat. Genet. 39 (2007) 1410-1413. 191.

Z. Bieniawska et al., Disruption of the Arabidopsis circadian clock is

responsible for extensive variation in the cold-responsive transcriptome, Plant

1407

Physiol. 147 (2008) 263-279. 192.

S. Kidokoro et al., The phytochrome-interacting factor PIF7 negatively

cr

1408

ip t

1406

regulates DREB1 expression under circadian control in Arabidopsis, Plant

1410

Physiol. 151 (2009) 2046-2057. 193.

D. Chen et al., Antagonistic basic helix-loop-helix/bZIP transcription factors

an

1411

us

1409

form transcriptional modules that integrate light and reactive oxygen species

1413

signaling in Arabidopsis, Plant Cell 25 (2013) 1657-1673.

M

1412

Ac ce p

te

d

1414

Page 58 of 65

59 1414 1415

Figure Legends Figure 1: Effect of domestication in rice. Oryza rufipogon (top panel) is

1417

considered to be the ancestor of cultivated rice (Oryza sativa; bottom panel). The

1418

changes in development and overall appearance between the two lines are very

1419

obvious.

us

cr

ip t

1416

1420

Figure 2: Proposed components of abiotic stress responses in plants.

1422

Early responses to abiotic stress are likely to be general stress responses: cellular

1423

protection of macromolecules and oxidative stress, metabolic adjustments and

1424

hormonal changes that lead to developmental responses. Genetic variation in crop

1425

species can lead to different responses in reaction to certain threshold levels of

1426

abiotic stresses. Some germplasm will initiate a senescence response and stop

1427

growth, with negative yield consequences. Other germplasm is more resilient and

1429 1430 1431 1432

M

d

te

Ac ce p

1428

an

1421

will maintain growth and productivity as much as possible. The latter phenotype will require successful interaction between the different early responses to establish stress-specific responses. Genetic variation can occur at different levels of the response pathway. QTL (stars with letter “Q”) in different parts of the general stress response will affect response to different abiotic stresses. QTL in

1433

the stress-specific responses are predicted to only affect response to a particular

1434

stress. Targeting the genetic variation in the general stress response could be more

1435

successful, but may require additional QTL in the stress-specific responses. QTL

1436

for stress-specific responses may have negative effects for other stresses.

Page 59 of 65

60 1437 Figure 3: Effect of drought stress at the reproductive stage in wheat. Drought

1439

stress leads to extensive leaf senescence in wheat (top left). Re-watering results in

1440

the development of new freshly green tillers that will flower and produce grains,

1441

while grain development in the older stressed tillers is either aborted or leads to

1442

spikes without grain (top right). The close-up pictures at the bottom show prolific

1443

initiation of new tillers in response to re-watering after drought treatment.

us

cr

ip t

1438

an

1444

Figure 4: Simplified schematic representation of the components involved in the

1446

environmental response module that controls plant growth, based on the shade

1447

avoidance response pathway in Arabidopsis. The PIF and DELLA proteins are

1448

central regulators of growth and environmental responses in plants.

d

te Ac ce p

1449

M

1445

Page 60 of 65

61  Abiotic stresses affect yield and productivity of crop plants.  Stress tolerance in crops correlates with maintenance of growth and productivity.  Early stress-responsive genes control growth responses  Selection strategies should use stress-induced traits and focus on early general stress response.  Plant hormone interactions control development and abiotic stress tolerance.

ip t

1449 1450 1451 1452 1453 1454 1455

Ac ce p

te

d

M

an

us

cr

1456

Page 61 of 65

Ac ce pt e

d

M

an

us

cr

ip t

Figure(s)

Figure 1

Page 62 of 65

ip t

Q

Q Q

Oxidative Stress Q

Metabolism

Q

Q

Development Q

Q

Q

Growth

an

Senescence

M

Light Stress

Q

Temperature Q Stress

Water Stress

Others

Q

Q

Ac ce pt e

d

Stress Specific Responses

us

General Stress Response

cr

Environmental Signals

Figure 2

Page 63 of 65

ip t

Ac ce pt e

d

M

an

us

cr

+H2O

Figure 3

Page 64 of 65

Phytochrome

Receptors

an

Circadian Clock

us

cr

Abiotic stresses

ip t

Light

?

PIF

ROS Signalling

M

Sugars

d

DELLA

Ac ce pt e

Ethylene

Senescence

CBF/DREB GA

Auxin Growth

Stress-specific response

Figure 4

Page 65 of 65

To grow or not to grow: a stressful decision for plants.

Progress in improving abiotic stress tolerance of crop plants using classic breeding and selection approaches has been slow. This has generally been b...
926KB Sizes 2 Downloads 8 Views