Accepted Manuscript Regulation of photosynthesis during abiotic stress-induced photoinhibition Mayank Anand Gururani, Jelli Venkatesh, Lam-Son Phan Tran PII:

S1674-2052(15)00238-5

DOI:

10.1016/j.molp.2015.05.005

Reference:

MOLP 136

To appear in:

MOLECULAR PLANT

Received Date: 22 January 2015 Revised Date:

12 May 2015

Accepted Date: 12 May 2015

Please cite this article as: Gururani M.A., Venkatesh J., and Tran L.-S.P. (2015). Regulation of photosynthesis during abiotic stress-induced photoinhibition. Mol. Plant. doi: 10.1016/ j.molp.2015.05.005. 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.

ACCEPTED MANUSCRIPT

1

Regulation of photosynthesis during abiotic stress-induced photoinhibition

2

Mayank Anand Gururani1, Jelli Venkatesh2, Lam-Son Phan Tran3

3 1

School of Biotechnology, Yeungnam University, Gyeongsan, Gyeongbook 712-749, Korea

5

2

Department of Bioresource and Food Science, Konkuk University, Seoul, 143-701, Korea

6

3

7

Suehiro-cho, Tsurumi, Yokohama 230-0045, Japan

Running title: Photosynthesis under abiotic stress

10

SC

Signaling Pathway Research Unit, RIKEN Center for Sustainable Resource Science, 1-7-22,

8 9

RI PT

4

Corresponding author: Lam-Son Phan Tran

12

Signaling Pathway Research Unit, RIKEN Center for Sustainable Resource Science, Yokohama,

13

Kanagawa, Japan Tel. +81-45-503-9593

14

Fax. +81-45-503-9591

15

Email: [email protected]

16

M AN U

11

Short Summary

18

The current understanding of abiotic stress-induced photoinhibition in higher plants is reviewed.

19

Furthermore, the mechanism of PSII damage-repair cycle and the involvement of various factors,

20

such as phytohormones and transcription factors in regulation of photosynthetic machinery are

21

discussed.

AC C

EP

TE D

17

1

ACCEPTED MANUSCRIPT

22

ABSTRACT

23

Plants as sessile organisms are continuously exposed to abiotic stress conditions that impose

25

numerous detrimental effects on them, causing tremendous yield loss. Abiotic stresses, including

26

high sunlight, confer serious damage to the photosynthetic machinery of the plants. Photosystem

27

II (PSII) is one of the most susceptible components of the photosynthetic machinery that bears

28

the brunt of abiotic stress. In addition to the generation of reactive oxygen species (ROS) by

29

abiotic stress, ROS can also result from the absorption of excessive sunlight by the light-

30

harvesting complex (LHC). ROS can damage the photosynthetic apparatus, particularly PSII,

31

resulting in photoinhibition due to an imbalance in the photosynthetic redox signaling pathways

32

and the inhibition of PSII repair. Designing plants with improved abiotic stress tolerance will

33

require a comprehensive understanding of ROS signaling and regulatory functions of various

34

components, including protein kinases, transcription factors and phytohormones, in responses of

35

photosynthetic machinery to abiotic stress. Bioenergetics approaches, such as chlorophyll a

36

transient kinetics analysis, have facilitated our understanding of plant vitality and the assessment

37

of PSII efficiency under adverse environmental conditions. This review discusses the current

38

understanding and indicates potential areas of further studies on the regulation of the

39

photosynthetic machinery under abiotic stress.

SC

M AN U

TE D

40

RI PT

24

Key words: abiotic stress; chlorophyll a; fluorescence; hormones; light harvesting complex;

42

photosynthesis; photosystem; redox signaling; transcription factors

AC C

43

EP

41

2

ACCEPTED MANUSCRIPT

44

1. INTRODUCTION

45

Photosynthesis is a multistep process of successive redox reactions that occur when the light-

47

harvesting complexes (LHCs) absorb photonic energy and transfer it to photosystem (PS)

48

reaction centers via excitons (Baker, 2008) (Figure 1). Abiotic stress caused by adverse

49

environmental conditions, such as drought, heat, heavy metal toxicity and high light (HL),

50

resulted in an over-reduction of the electron transport chain (ETC) which in turn leads to

51

photooxidation (Foyer et al., 2012; Foyer and Noctor, 2005; Kangasjarvi et al., 2012; Nishiyama

52

et al., 2006; Nishiyama and Murata, 2014; Rochaix, 2011; Takahashi and Murata, 2008). Plants

53

have several mechanisms to overcome this problem, e.g. reducing the rate of electron transport

54

by converting the excessively absorbed light into thermal energy. The dissipation of excess

55

excitation energy as heat is known as non-photochemical quenching (NPQ) of chlorophyll

56

fluorescence (Nath et al., 2013a; Rochaix, 2011; Spetea et al., 2014; Tikkanen et al., 2011). NPQ

57

is a major photoprotective response as it reduces the concentration of chlorophyll excited states

58

(Chl*) in PSII by activating a heat dissipation channel (Cazzaniga et al., 2013). Changes in the

59

distribution and molecular orientation of chlorophyll proteins in the thylakoid membrane lead to

60

the NPQ during HL stress (Herbstova et al., 2012; Johnson et al., 2011). Earlier studies have

61

indicated that during abiotic stress, energized electrons are allocated to dioxygen (O2) (Baker,

62

2008; Toth et al., 2011). This O2 is used in two vital photosynthetic reactions, photorespiration

63

and the Mehler peroxidase (MP) reaction. In the MP reaction, O2 is reduced to superoxide (O2-)

64

and the O2- to hydrogen peroxide (H2O2), both of which are potent reactive oxygen species

65

(ROS) molecules. Consequently, less nicotinamide adenine dinucleotide phosphate (NADP+) is

66

reduced which slows down the rate of CO2 fixation.

67

2O2 + 2Fdred −→ 2O2•− + 2Fdox, where Fd is ferredoxin.

SC

M AN U

TE D

EP

AC C

68

RI PT

46

Plants perceive stress signals through receptors that trigger molecular cascades to

69

transmit the signals to regulatory systems via ion channels, signaling proteins, and secondary

70

messengers (Choudhary et al., 2012; Le et al., 2012; Schmutz et al., 2010). The regulatory

71

system is comprised of various components, including phytohormones, transcription factors

72

(TFs), mitogen-activated protein kinases, and photosynthetic protein kinases and phosphatases,

73

that regulate the expression of various stress-responsive genes (Foyer and Shigeoka, 2011;

74

Osakabe et al., 2014; Puranik et al., 2012). Although the generation of ROS by abiotic stress and 3

ACCEPTED MANUSCRIPT

their detoxification have been extensively studied, the precise mechanisms underlying the

76

distribution of ROS in specific cellular compartments are still unclear. Intracellular generation of

77

H2O2, the most stable ROS, has always been associated with chloroplasts but recent studies

78

suggest that peroxisomes may produce more H2O2 than other organelles (Noctor et al., 2014).

79

The production of ROS in leaf tissues is regulated by the harvesting and distribution of light

80

energy to the photosynthetic machinery (Tikkanen et al., 2014a). ROS are generated in

81

chloroplasts within the ETCs of PSII and PSI during the light reactions, and their production is

82

further increased during stress when carbon dioxide (CO2) is limited and ATP synthesis is

83

impaired (Nishiyama and Murata, 2014; Noctor et al., 2014; Takahashi and Murata, 2008;

84

Yamamoto et al., 2008). The damage to DNA, lipids and proteins by elevated levels of

85

intracellular ROS is known as a result of oxidative stress (Foyer and Shigeoka, 2011; Schmutz et

86

al., 2010). The detrimental effects of ROS accumulation in chloroplasts include the inhibition of

87

de novo synthesis of D1 protein (also known as photosystem b A or PsbA) which is needed for

88

PSII repair (Miyao, 1994; Nishiyama et al., 2011b; Tikkanen et al., 2011) (Figure 2),

89

suppression of ROS-responsive chloroplastic enzymes (Kato and Sakamoto, 2014; Yoshioka et

90

al., 2006), and the disarrangement of thylakoid architecture (Gratao et al., 2009). Recent studies

91

in various organisms suggest that ROS also play pivotal roles as signaling molecules in normal

92

biochemical and physiological responses (Foyer and Shigeoka, 2011; Schmutz et al., 2010;

93

Tikkanen et al., 2014a). The level of intracellular ROS can be regulated by modulating the

94

expression of genes that encode antioxidants, such as ascorbate or glutathione (Upadhyaya et al.,

95

2011), or ROS-scavenging enzymes, such as peroxidases or dismutases (Schmutz et al., 2010;

96

Takahashi and Murata, 2008). Even ROS-scavenging systems, however, do not completely

97

remove intracellular ROS. Numerous reports have documented an increase in abiotic stress

98

tolerance by the genetic manipulation of antioxidants and ROS-scavenging enzymes (Schmutz et

99

al., 2010; Toth et al., 2011; Upadhyaya et al., 2011). Associating such modifications with

101

SC

M AN U

TE D

EP

AC C

100

RI PT

75

cellular redox homeostasis and redox signaling, however, may be a more appropriate approach.

102

2. IMPACT OF CELLULAR ROS PRODUCTION AND PHOTOINHIBITION ON THE

103

PHOTOSYNTHETIC APPARATUS

104

4

ACCEPTED MANUSCRIPT

The PSII complex accomplishes the unique task of splitting water molecules and liberating

106

molecular oxygen through an oxygen-evolving complex (OEC) (Gururani et al., 2012; Tikkanen

107

and Aro, 2014; Yamamoto et al., 2008; Yi et al., 2005). The released electrons are then

108

transferred to the PSI complex via ETC from plastocyanin to ferredoxin (Nellaepalli et al., 2014)

109

(Figure 1). These reactions generate a high level of reactive oxygen radicals in PSII, which

110

cause photodamage to the photosynthetic machinery (Henmi et al., 2004; Murata et al., 2007;

111

Takahashi and Badger, 2011; Takahashi and Murata, 2008). One of the primary sources of PSII

112

photodamage is the light-dependent disruption of OEC, which results in the release of

113

manganese ions (Hakala et al., 2005; Henmi et al., 2004; Murata et al., 2007; Takahashi and

114

Badger, 2011). The inhibition of PSII activity under continuous exposure to HL is commonly

115

known as photoinhibition (Murata et al., 2007).

SC

M AN U

116

RI PT

105

Environmental stresses do not directly affect photoinhibition but rather facilitate the inhibition of PSII damage repair (Murata et al., 2007; Nishiyama et al., 2011b; Nishiyama and

118

Murata, 2014). Several studies in plants and cyanobacteria have suggested that extent of

119

photodamage to PSII is directly proportional to the intensity of incident light and that this

120

proportionality remains unaffected by various environmental stresses (Allakhverdiev and Murata,

121

2004; Gombos et al., 1994; Nishiyama et al., 2004; Takahashi and Murata, 2008). When light

122

energy absorbed by the LHCII pigments is higher than the energy consumed, photoinhibition

123

increases exponentially, thus causing severe damage to PSII (Nishiyama and Murata, 2014;

124

Tikkanen and Aro, 2014). Superoxide anion radical or singlet oxygen (·O2-) is produced when

125

more electrons are released in the ETC than the electron-consuming capacity of the Calvin cycle.

126

The singlet oxygen is then converted to H2O2 by the action of superoxide dismutase (SOD)

127

(Figure 1) (Nishiyama and Murata, 2014). According to the currently accepted scheme of PSII

128

photoinhibition, ROS, such as superoxide radicals and singlet oxygen produced as a result of HL,

129

high salinity (Allakhverdiev and Murata, 2004; Allakhverdiev et al., 2002; Ohnishi and Murata,

130

2006), high or low temperature (Allakhverdiev et al., 2007; Allakhverdiev and Murata, 2004;

131

Mohanty et al., 2007; Takahashi et al., 2009; Yang et al., 2007) and limited CO2 fixation

132

(Takahashi and Murata, 2005; Takahashi and Murata, 2006; Wang et al., 2014), can inhibit the

133

translation of PsbA mRNA, thus inactivating the PSII repair process (Figure 2H) (Nishiyama et

134

al., 2006; Nishiyama et al., 2011b; Nishiyama and Murata, 2014; Takahashi and Murata, 2008).

135

Accumulation of intracellular ROS depends on a balance between the generation and

AC C

EP

TE D

117

5

ACCEPTED MANUSCRIPT

detoxification of ROS via various ROS-scavenging enzymes (Figure 1). This postulation makes

137

the ROS-scavenging enzymes an obvious target to reduce the levels of ROS. Several studies

138

suggest that decreased levels of intracellular ROS and reduced rate of photoinhibition can be

139

achieved by engineering the production of ROS-scavenging enzymes, such as catalase (Al-

140

Taweel et al., 2007; Miyagawa et al., 2000; Shikanai et al., 1998) and APX (Kornyeyev et al.,

141

2003; Maruta et al., 2010; Murgia et al., 2004; Yabuta et al., 2002), in higher plants. Similarly,

142

increased levels of low molecular weight antioxidants, such as α-tocopherol (vitamin E)

143

(Demmig-Adams et al., 2013; Havaux et al., 2005) and carotenoids (Ramel et al., 2012; Stahl

144

and Sies, 2003; Triantaphylides and Havaux, 2009), have also been proposed to reduce the level

145

of singlet oxygen. On the basis of previous findings, it is reasonable to predict that increased

146

activity of ROS-scavenging enzymes and increased accumulation of antioxidants might reduce

147

the levels of intracellular ROS, thereby allowing the synthesis of D1 protein. PSI and PSII are

148

associated with their respective light-absorbing antenna systems, LHCI and LHCII (Klimmek et

149

al., 2006; Rochaix, 2014; Shapiguzov et al., 2010). In Arabidopsis thaliana, six and fifteen genes

150

encoding the components of LHCI and LHCII, respectively, have been identified (Klimmek et al.,

151

2006; Rochaix, 2014). The PSI complex is mainly comprised of 15 photosystem a (Psa) subunits,

152

(PsaA - L and PsaN-P) (Nellaepalli et al., 2014; Rochaix, 2014). Chlorophyll molecules bound to

153

PSI reaction centers act as light absorbing pigments, and the entire PSI complex is bound to its

154

corresponding LHCI proteins (Amunts et al., 2010). The compositions of PSII and PSI antenna

155

complexes are quite distinct, as LHCII mainly contains chlorophyll b while LHCI is enriched

156

with chlorophyll a (Rochaix, 2014; Xu et al., 2012). Additionally, PSII-LHCII supercomplex

157

contains a D1-D2 (or PsbA-PsbD) dimer coupled with two antennal proteins, the chlorophyll

158

proteins (CPs) CP43 and CP47, and three minor proteins CP24, CP26 and CP29 (Che et al.,

159

2013; de Bianchi et al., 2008). The entire PSII-LHCII supercomplex is embedded in the granum,

160

while PSI is localized in the stromal lamellae of the chloroplasts (Rochaix, 2014).

SC

M AN U

TE D

EP

AC C

161

RI PT

136

Recent studies on the acclimation of PSII-LHCII supercomplex during photoinhibition

162

suggest that moderate phosphorylation of PSII under normal light results in efficient LHCII

163

activity and slower PSII damage (Grieco et al., 2012; Nath et al., 2013a; Tikkanen and Aro,

164

2014; Tikkanen et al., 2010). Under these conditions, the PSII-LHCII supercomplex remains

165

intact and moderates the transfer of energy from LHCII to PSI. Moderate phosphorylation of

166

LHCII allows the PSI complexes to move towards the grana margins, which ensures the transfer 6

ACCEPTED MANUSCRIPT

of adequate amounts of excitation energy to PSI. However, PSII proteins under HL conditions

168

are phosphorylated at a much higher rate in order to negate the effect of photodamaged PSII

169

complexes. The PSII-LHCII supercomplex loses its structural integrity, and the transfer of excess

170

excitation energy from PSII-LHCII to PSI becomes unlimited (Grieco et al., 2012; Tikkanen and

171

Aro, 2014; Tikkanen et al., 2010). Under conditions of continuous HL, dephosphorylation of

172

LHCII inhibits any further transfer of energy to PSI (Tikkanen and Aro, 2014).

173

RI PT

167

Unlike PSII, PSI is not frequently damaged due to a very efficient mechanism that

protects it from photoinhibition. Photodamage to PSI is primarily caused when the supply of

175

electrons from PSII exceeds the electron-accepting capacity of PSI, and once the PSI is damaged,

176

the consequent recovery of PSI centers becomes very slow (Sonoike, 2011; Tikkanen et al.,

177

2014b). Slow recovery of PSI centers in chilling stress-treated plant leaves has been reported in

178

Arabidopsis (Zhang and Scheller, 2004), cucumber (Terashima et al., 1994; Kudoh and Sonoike,

179

2002; Zhou et al., 2004, Zhang et al., 2014) and sweet pepper (Li et al., 2004). Apparently,

180

photoinhibition in isolated thylakoids has been reported to be promptly induced under non-

181

stressed conditions, whereas specific environmental conditions and specific plant species are

182

required to induce photoinhibition under in vivo conditions (Sonoike, 2011). This is one of the

183

most important factors that have limited our understanding of the putative protective mechanism

184

of PSI photoinhibition. Nevertheless, several studies have indicated that the proton gradient-

185

dependent or Cytb6f-mediated slowdown of ETC and LHCII-mediated excitation of PSII and

186

PSI via NPQ and LHCII phosphorylation regulate the photoprotection of PSI in higher plants

187

(Sonoike et al., 1995; Joliot and Johnson, 2011; Suorsa et al., 2012; Grieco et al., 2012).

188

Recently, Tikkannen et al. (2014b) demonstrated that in addition to the above mentioned

189

mechanisms, controlled photoinhibition of PSII regulates the ETC and prevents the formation of

190

ROS and photodamage to PSI. Increased PSI-mediated cyclic electron flow was reported in

191

Secale cereale plants subjected to chilling and high light stress, indicating the role of

192

temperature/light-dependent acclimation in the induction of selective tolerance to PSI

193

photoinhibition (Ivanov et al. 1998). Similarly, it has been reported that cyclic electron flow

194

around PSI is required to produce a proton gradient that in turn leads to an efficient NPQ under

195

heat stress in Ficus concinna trees (Jin et al., 2009).

AC C

EP

TE D

M AN U

SC

174

196

7

ACCEPTED MANUSCRIPT

197

3. ROLE OF PSII PROTEIN PHOSPHORYLATION IN REGULATION OF D1

198

DEGRADATION

199

PSII, which is often referred to as the engine of life on earth, is the component of the

201

photosynthetic machinery that is most susceptible to abiotic stresses (Nath et al., 2013a). Plants

202

have evolved a mechanism for repairing PSII damage in order to attenuate photodamage and for

203

the efficient functioning of PSII (Yamamoto et al., 2008). Photodamage caused by HL in plants

204

can be determined by inhibiting the PSII repair process using chemicals like chloramphenicol or

205

lincomycin that can inhibit the protein synthesis in the exposed plant cells (Murata et al., 2007).

206

The PSII damage-repair cycle includes (i) phosphorylation and dephosphorylation of PSII

207

proteins, (ii) disassembling of the PSII complex, (iii) proteolysis and de novo synthesis of D1

208

protein, and (iv) reconstitution of the PSII complex (Bonardi et al., 2005; Nath et al., 2013a;

209

Nixon et al., 2010; Samol et al., 2012; Tikkanen et al., 2014b) (Figure 2).

SC

M AN U

210

RI PT

200

Synchronizing the photosynthetic apparatus requires a balance between the excitation energies of PSII and PSI. This balance has been attributed to the ability to regulate the level of

212

phosphorylation and dephosphorylation of LHCII and PSII proteins (Grieco et al., 2012;

213

Tikkanen et al., 2010; Tikkanen et al., 2014b). Phosphorylation of LHCII and PSII proteins in

214

Arabidopsis is facilitated by the state transition kinases, STN7 and STN8, respectively (Bonardi

215

et al., 2005), while their dephosphorylation is triggered by thylakoid associated phosphatase 38

216

(TAP38 or PPH1) (Pribil et al., 2010; Shapiguzov et al., 2010) and Psb core phosphatase (PBCP)

217

(Samol et al., 2012), respectively (Figure 2). A recent study has demonstrated that TAP38/PPH1

218

phosphatase was required to prevent the canonical state transition upon increase in light intensity

219

(Nageswara et al., 2015).

EP

AC C

220

TE D

211

221

4. PROTEASES INVOLVED IN THE DEGRADATION AND SYNTHESIS OF D1

222

PROTEIN

223 224

PSII phosphorylation regulates the functional folding and macroscopic structure of plant

225

thylakoid membranes (Fristedt et al., 2009; Nath et al., 2013a; Pribil et al., 2014). STN7-

226

dependent phosphorylation appears to provide adequate levels of excitation energy to PSI to

227

accomplish an efficient electron transfer from PSII to PSI (Grieco et al., 2012). Additionally, as a 8

ACCEPTED MANUSCRIPT

retrograde signaling kinase, STN7 is known to trigger the phosphorylation cascade, and thus

229

regulating the expression of photosynthesis-related genes and assembly of the photosynthetic

230

machinery (Tikkanen et al., 2012). The functional characterization of Arabidopsis stn8 single

231

and stn7 × stn8 double mutants indicated that STN7 can phosphorylate LHCII, as well as PSII

232

proteins to some extent, because complete inhibition of PSII phosphorylation was only observed

233

in the stn7 x stn8 double mutant (Bonardi et al., 2005; Fristedt et al., 2009). In contrast, studies

234

of a rice stn8 mutant revealed that Osstn8 alone can produce all of the phenotypes observed in

235

Arabidopsis stn7 x stn8 mutants, indicating that distinctly specific regulatory mechanisms

236

involving STN7 and STN8 exist in monocots vs. dicots (Nath et al., 2013b). Studies on

237

TAP38/PPH1 have indicated that PPH1 may have other unknown function(s) besides STN7-

238

mediated dephosphorylation of LHCII, as co-expression of PPH1 and STN7 was not detected

239

(Nath et al., 2013a; Obayashi et al., 2009). Recent studies using Arabidopsis mutants with

240

inactivated protein phosphatase 2C (PP2C)-type PBCP revealed that dephosphorylation of PSII

241

subunits is crucial for efficient degradation of D1 (Puthiyaveetil et al., 2014). Additionally,

242

PBCP has been found to regulate thylakoid stacking (Bonardi et al., 2005). Although the role of

243

LHCII phosphorylation in PSII repair is still not clear, the phosphorylation of CP29, a minor

244

subunit of LHCII, appears to be essential for the disassembly of the LHCII-PSII supercomplex

245

(Fristedt and Vener, 2011).

SC

M AN U

TE D

246

RI PT

228

Reaction-center D1 protein is a key player in the PSII repair cycle (Nath et al., 2013a; Nishiyama and Murata, 2014; Rochaix, 2014; Tikkanen and Aro, 2014). The proteolysis and de

248

novo synthesis of D1 protein during the PSII repair cycle are facilitated by chloroplastic

249

proteases in the lumen, stroma and the thylakoid envelope (Che et al., 2013; Kapri-Pardes et al.,

250

2007; Kato et al., 2009; Pribil et al., 2014; Schuhmann and Adamska, 2012; Sun et al., 2007).

251

Serine-, metallo- and putative cysteine- and aspartic acid-proteases, have been identified, some

252

of which require ATP (Pribil et al., 2014; Sakamoto, 2006; van der Hoorn, 2008). Presumably,

253

many other proteases remain to be discovered. Known protease functions include chloroplastic

254

biogenesis, degradation of signaling components, maintenance of chloroplastic homeostasis, and

255

degradation of damaged proteins (Che et al., 2013; Sun et al., 2010a; Sun et al., 2010b; Yin et al.,

256

2008). The current understanding of chloroplastic proteases that play pivotal roles in D1 protein

257

processing and degradation are summarized in the next section and in Supplementary Table 1.

AC C

EP

247

258

9

ACCEPTED MANUSCRIPT

259

4.1. Degradation (Deg) proteases

260

Chloroplastic Deg proteases have received significant attention primarily due to their role in the

262

degradation of photodamaged PSII proteins. The function of many of these proteases, however,

263

remains to be determined. In plants, periplasmic degradation (Deg) proteases are ATP-

264

independent serine endopeptidases which are involved in the degradation of photodamaged

265

proteins (Haussuhl et al., 2001; Sakamoto, 2006; Schuhmann and Adamska, 2012; Sun et al.,

266

2010a; van der Hoorn, 2008) (Figure 2). Although these proteases were first discovered in

267

Escherichia coli, they are present in most organisms (Schuhmann and Adamska, 2012).

268

Arabidopsis contains 16 Deg proteases (Deg1-16) that, with the exception of Deg5, possess a C-

269

terminal protease domain and an N-terminal PDZ domain (Haussuhl et al., 2001). Deg proteases

270

are present in chloroplasts, mitochondria, peroxisomes and nuclei (Schuhmann and Adamska,

271

2012). Of the five known chloroplastic Deg proteases, Deg1, Deg5 and Deg8 are attached to the

272

luminal side of thylakoid membranes, and Deg2 and Deg7 are in the stroma (Chi et al., 2012).

273

Arabidopsis mutants with reduced levels of Deg1 exhibit enhanced sensitivity to photoinhibition

274

and an increased accumulation of D1 protein; indicating that Deg1 plays a role in the degradation

275

of D1 protein (Kapri-Pardes et al., 2007). In addition, recent studies in Arabidopsis have

276

indicated the involvement of Deg1in the degradation of CP29 and CP26 proteins under

277

photoinhibition (Zienkiewicz et al., 2012), and of Deg2 in degradation of CP24 under various

278

abiotic stress conditions (Lucinski et al., 2011). The Deg2 protease has been associated with

279

abiotic stress responses in plants. Arabidopsis deg2 mutants showed impaired activity to degrade

280

CP24 in response to various abiotic stresses, such as high salinity, HL and wounding, indicating

281

that Deg2 protease is required for plant growth and development under optimal as well as

282

stressed conditions (Lucinski et al., 2011). Interestingly, the effect of different abiotic stresses on

283

the accumulation of proteases is distinctive. For instance, 2- to 4-fold increase in the amount of

284

Deg2 was observed in Arabidopsis leaves treated with high salinity, desiccation or high

285

irradiance for 2 h, while heat-treated leaves exhibited only trace amounts of Deg2 in the

286

thylakoid membranes (Haussuhl et al., 2001).

AC C

EP

TE D

M AN U

SC

RI PT

261

287 288

4.2. Filamentation temperature-sensitive H (FtsH) proteases

289

10

ACCEPTED MANUSCRIPT

FtsH proteases are ATP-dependent metalloproteases found in the thylakoids of higher-plant

291

chloroplasts. Twelve genes encoding FtsH proteins have been identified in the Arabidopsis

292

nuclear genome; nine of which (FtsH1, 2, 5-9, 11-12) are found in chloroplasts, three (FtsH3,4

293

and 10) in mitochondria and one (FtsH11) in both chloroplasts and mitochondria (Yoshioka-

294

Nishimura and Yamamoto, 2014). The exact location of the FtsH proteases in the thylakoid

295

membrane still remains to be elucidated. Given their large hexameric form (Kirchhoff et al.,

296

2011), however, it is most likely that they reside in the unstacked region of the thylakoid

297

membrane rather than in the stacked grana. Early studies of Arabidopsis proteases suggested that

298

Deg proteases initiated D1 degradation while further degradation was accomplished by FtsH

299

proteases (Kapri-Pardes et al., 2007; Sun et al., 2010a; Sun et al., 2007). In contrast, recent

300

studies, using Arabidopsis deg and ftsH mutants, have suggested that FtsH proteins act as the

301

initiating proteases and Deg as the assisting proteases in D1 protein degradation (Adam et al.,

302

2011; Kato et al., 2012) (Figure 2). A recent study using transmission electron microscopy and

303

immunogold labeling of FtsH in spinach leaves under HL stress revealed unstacking of

304

thylakoids on the marginal regions of grana stacks, indicating a prominent role of thylakoid for

305

efficient repair of photodamaged D1 proteins (Yoshioka-Nishimura et al., 2014). Interestingly,

306

several reports have corroborated D1 phosphorylation and protease activity with abiotic stress

307

tolerance. It has been proposed that protein phosphorylation of D1 plays an important role in

308

protection of plants against the oxidative damage (Chen et al., 2012). Superoxide anions, H2O2

309

and hydroxyl radical are reportedly involved in the D1degradation under HL as one of the

310

initiating factors (Miyao, 1994; Miyao et al., 1995). Increased levels of H2O2 were found in

311

Arabidopsis mutants with limited FtsH activity. Apparently, the protein instability caused by

312

partial cleavage of D1 proteins led to cytotoxicity that resulted in increased levels of ROS (Kato

313

and Sakamoto, 2014). Earlier, FtsH was reported to be involved in D1 turnover in response to

314

heat stress in spinach (Yoshioka et al., 2006). One of the 12 known FtsHs, FtsH11 was proposed

315

to have a direct role in thermo-tolerance in Arabidopsis plants (Chen et al., 2006) and that

316

FtsH11 protects the photosynthesis apparatus from heat stress at all stages of plant development

317

(Chen et al., 2006).

AC C

EP

TE D

M AN U

SC

RI PT

290

318 319

4.3. C-terminus-processing (Ctp) proteases

320

11

ACCEPTED MANUSCRIPT

Further processing of D1 protein by Ctp proteases is essential for the formation of fully

322

functional PSII complexes that are capable of splitting water (Chi et al., 2012; Yin et al., 2008)

323

(Figure 2). Unlike bacteria, in which many types of Ctp proteases exist, in plants, such as

324

Arabidopsis, only CtpA-type proteases have been identified (Che et al., 2013). Three genes

325

encoding CtpA-type proteases have been identified in Arabidopsis (Chi et al., 2012; Yin et al.,

326

2008). Characterization of CtpA1(CtpA/At3g57680) indicated that D1degradation significantly

327

increased in the Arabidopsis AtctpA1loss-of-function mutant. CtpA1 thus appears to be essential

328

for the efficient repair of PSII under HL (Yin et al., 2008). Additionally, functional analysis of

329

another CtpA/At4g17740 in Arabidopsis revealed that the loss of its activity led to a recovery of

330

D1 protein in its precursor form, indicating that CtpA/At4g17740 is essential for D1 C-terminal

331

processing, and thus PSII damage repair (Che et al., 2013). In order to determine the role of

332

cytokinin (CK) in HL-induced stress, expression of several Deg (Deg5 and Deg8) and FtsH

333

(FtsH1, FtsH2, FtsH5 and FtsH8) genes, as well as that of all three identified CtpA genes, were

334

analyzed in CK-deficient and wild-type (WT) Arabidopsis plants (Cortleven et al., 2014). No

335

significant difference in the expression of the examined FtsH and Deg genes were observed

336

between the WT and CK-deficient plants exposed to HL. While differences in the level of

337

expression of CtpA/At4g17740 and CtpA/At5g46390 in response to HL were non-significant in

338

the WT and CK-deficient plants, expression of AtCtpA1 was more strongly up-regulated by HL

339

in WT than in CK-deficient plants. These data suggest that HL-responsive expression of AtCtpA1

340

is CK-dependent. Additionally, AtCtpA1 mutants exhibited a WT-like response to HL (Cortleven

341

et al., 2014), indicating that AtCtpA1might act in concert with other proteins to regulate plant

342

response to HL.

SC

M AN U

TE D

EP

AC C

343

RI PT

321

344

5. ASSESSMENT OF PHOTOSYSTEM II EFFICIENCY: MEASUREMENT OF

345

CHLOROPHYLL A FLUORESCENCE

346 347

The chlorophyll a fluorescence induction reflects the variations in chlorophyll a fluorescence

348

intensity, which occur when a photosynthetic specimen is moved from darkness to light

349

(Papageorgiou et al., 2007). The fluorescence induction is divided into two phases: (i) the fast

350

induction phase or the OJIP phase, in which O is origin, P is peak and J-I are the intermediate

351

phases, and (ii) the slow induction phase or PSMT phase in which P is peak, S is steady, M is 12

ACCEPTED MANUSCRIPT

maximum and T is the terminal state (Figure 3). The fast induction kinetics expresses the

353

primary photochemistry of PSII, while the slow kinetics is a complex phase primarily related to

354

the interactions between processes in the thylakoids and in the reductive carbon cycle of the

355

stroma (Krause and Weis 1984). In other words, slow induction phase is the light that is emitted

356

from photosynthetic specimen in the red-infra-red region of the spectrum for a short time after

357

the fast fluorescence has decayed. The fast phase of the transient provides relevant information

358

of the events that take place during the reduction of electron acceptors during ETC. On the other

359

hand, the slow transient phase is very difficult to be interpreted, primarily because several

360

different processes like non-photochemical quenching, ATP synthesis and Calvin–Benson cycle

361

begin to be involved during this phase (Stirbet and Govindjee, 2011). Nevertheless, the slow

362

fluorescence emission of PSII is considered as a useful approach to quantitatively study the light-

363

induced electron transfer and related events, such as proton movement, which are not detectable

364

by conventional spectroscopic methods (Goltsev et al., 2009). Over the years, researchers have

365

analyzed the slow fluorescence method to assess the plant’s performance under low temperature

366

(Itoh, 1980), high salinity (Zhang and Xing 2008), heavy metal stress (Plekhanov and Chemeris

367

2003), heat stress (Goltsev et al., 1987; Bilger and Schreiber 1990), drought (Mladenova et al.

368

1998), light stress (Valikhanov et al., 2002) and UV irradiation (Zhang et al., 2007a). Zhang et

369

al. (2007b) described a technique for detecting plant senescence based on quantitative

370

measurements of slow fluorescence and proposed that the changes in slow fluorescence intensity

371

reflect the changes in photosynthetic capacity and chlorophyll content during age-dependent and

372

hormone-modulated senescence (Zhang et al., 2007b). PSII efficiency under normal and stress

373

conditions can also be determined by fast chlorophyll a transient kinetics (Baker, 2008; Gururani

374

et al., 2012; Gururani et al., 2013; Gururani et al., 2015) (Figure 4). Analysis of fast chlorophyll

375

a transients has the potential to reveal interesting details pertaining to the alteration and

376

adjustment of the photosynthetic machinery during stress conditions. The reactivity of the

377

photosynthetic apparatus is important to the physiological status and vitality of plants subjected

378

to environmental stress. Measuring alterations of fast chlorophyll a fluorescence transients has

379

become a widely applied technique for assessing reactivity. Fluorescence (F) increases sharply

380

from F0 (minimum) to Fm (maximum) in dark-adapted plants exposed to a strong saturating light

381

pulse [3000-12000 µmol photons m-2 s-1, 200-1000 milliseconds (ms)]. At high temporal

382

resolution (0 to 200 ms), this increase can be seen as a polyphasic OJIP transient in three phases:

AC C

EP

TE D

M AN U

SC

RI PT

352

13

ACCEPTED MANUSCRIPT

OJ (0-3 ms), JI (3-30 ms) and IP (30-200 ms). Recent findings indicate that the increase from F0

384

to Fm may indicate the reduction of quinone (QA), the primary electron acceptor of PSII

385

(Schansker et al., 2014). The OJIP (or JIP) test elaborated by Strasser et al. (Strasser and Strasser,

386

1995; Strasser et al., 2004) is the main explanatory model used for explaining OJIP transients

387

(Baker, 2008; Baker and Rosenqvist, 2004; Maxwell and Johnson, 2000). The model compares

388

the photosynthetic activities of stressed and control plants and the test is a good, non-invasive

389

tool for analyzing the effects of a variety of stress factors on plants (Goltsev et al., 2012; Ranjan

390

et al., 2014; Toth et al., 2011; Zivcak et al., 2014a; Zivcak et al., 2014b). An analysis of

391

transgenic potato lines with altered photosystem b O (PsbO) production indicated a possible role

392

for this protein in enhanced potato tuberization and abiotic stress tolerance (Gururani et al., 2012;

393

Gururani et al., 2013). The exact mechanism behind the biochemical changes triggered by altered

394

production of PsbO, however, requires further investigation. A hypothetical model is provided

395

for the putative changes in the photosynthetic apparatus based on OJIP analyses (Figure 4).

SC

M AN U

396

RI PT

383

Chlorophyll a fluorescence transients provide extremely relevant information about PSII photochemistry and the events in the ETC. OJIP analysis simplifies this information through a

398

series of theoretical assumptions and mathematical calculations using specialized software. The

399

validity of some OJIP parameters has been a subject of debate among biophysicists (Stirbet and

400

Govindjee, 2011), but the acceptance of this analytical assessment of plant photosynthetic

401

efficiency and vitality will likely increase.

403 404

6. ROLE OF EXTRINSIC PROTEINS OF PSII IN D1 TURNOVER

EP

402

TE D

397

It is also imperative to recognize the role of extrinsic PSII proteins in D1 turnover in order to

406

fully understand the PSII repair cycle. The extrinsic proteins of PSII in higher plants mainly

407

include Psb subunits, PsbO, PsbP and PsbQ (Bricker et al., 2012). In vitro experiments with

408

spinach indicate that PSII devoid of PsbO becomes more vulnerable to photoinhibition and

409

accumulates significant amounts of D1 and CP43 (Henmi et al., 2004). PsbO may prevent

410

unnecessary interactions between photodamaged D1 and CP43, and the extended structure of

411

PsbO might protect the surface of D1 from reactive oxygen species (Yamamoto et al., 2008).

412

Moreover, PsbO has been proposed to function as GTPase and to regulate the phosphorylation

413

state of the D1 protein; a process that is associated with the efficient turnover of D1 protein

AC C

405

14

ACCEPTED MANUSCRIPT

during PSII repair (Aro et al., 2005; Bricker and Frankel, 2011; Lundin et al., 2007; Lundin et al.,

415

2008). PsbO-deficient potato plants exhibited reduced D1 and CP43 gene expression under a

416

normal light regimen (Gururani et al., 2012; Gururani et al., 2013). Specific roles for various

417

PsbO isoforms in oxygen evolution, PSII stability and plant growth have been identified (Lundin

418

et al., 2008). The precise mechanism of PsbO turnover and its putative role, in addition to the

419

stabilization of the oxygen-evolving complex, remains contentious and yet to be determined

420

(Kangasjarvi et al., 2012). Recent studies showed that removal of two intrinsic PSII proteins, the

421

PsbQ and PsbR, although resulted in only minor changes in terms of oxygen evolution and plant

422

growth, but showed a significant reduction in PSII activity and in PSII–LHCII super-complex

423

accumulation (Allahverdiyeva et al., 2013). Ishihara et al. (2007) demonstrated that a PsbP-like

424

protein 1 (PPL1) is required for the efficient repair of photodamaged PSII although the

425

underlying mechanism is not fully understood. PsbP, PsbQ and PsbR can be phosphorylated in

426

the thylakoid lumen (Ifuku, 2014; Reiland et al., 2009), indicating that their phosphorylation

427

might affect the assembly of PSII. On the other hand, removal of PsbP and PsbQ can induce

428

conformational changes in the arrangement of PSII protein assembly which includes the D1and

429

D2 proteins (Boekema et al., 2000; Bricker et al., 2012; Tomita et al., 2009). The addition or

430

removal of a phosphate group can alter protein conformation. Hence, examining putative

431

conformational changes in these extrinsic proteins during the PSII repair cycle would help

432

determine the state of phosphorylation of these proteins.

SC

M AN U

TE D

433

RI PT

414

7. INVOLVEMENT OF HORMONAL REGULATORY NETWORKS IN THE

435

RESPONSE OF PHOTOSYNTHETIC MACHINERY TO ABIOTIC STRESS

AC C

436

EP

434

437

Increasing evidence suggests that the interaction between phytohormones and cellular redox is an

438

essential aspect of the response of the photosynthetic structures to various abiotic stresses (Kim

439

et al., 2012; Krumova et al., 2013; Mayzlish-Gati et al., 2010). Recent studies using Arabidopsis

440

mutants with impaired photosynthetic light-harvesting indicated a strong interaction between the

441

control of excitation energy transfer and hormonal regulation (Tikkanen et al., 2014a). The

442

regulation of hormone metabolism by ROS generation and the intricacies of the crosstalk among

443

various hormones in response to various abiotic stresses have been extensively studied (Figure

444

5; Supplementary Table 2). In addition, a number of TFs have been shown to regulate 15

ACCEPTED MANUSCRIPT

photosynthesis through major hormonal pathways (Figure 5). For instance, TFs such as BZR1

446

and AtWRKY, were found to influence cell-wall and photosynthesis/chloroplast-related genes,

447

while a few TFs, for example GhDREB and CRF6 were shown to regulate PSII efficiency and

448

chlorophyll accumulation (Nguyen et al., 2014; Toledo-Ortiz et al., 2014; Waters et al., 2009;

449

Zhang et al., 2008). Supplementary Table 3 summarizes information on the recently identified

450

TFs that were found to be associated with the regulation of photosynthetic machinery under

451

normal and abiotic stress conditions.

452

RI PT

445

Brassinosteroids (BRs) play important roles in plant growth, defense, abiotic stress

tolerance, and the maintenance of high PSII efficiency and photosynthetic carbon fixation in

454

higher plants (Choudhary et al., 2012; Hu et al., 2013; Krumova et al., 2013; Oh et al., 2010).

455

Studies of BR-treated plants and BR-deficient mutants indicate a connection between BRs and

456

genes involved in photosynthesis (Bai et al., 2012; Oh et al., 2011). An Arabidopsis

457

brassinosteroid-insensitive 1(bri1) mutant exhibited a down-regulation of genes associated with

458

the regulation of photosynthesis and was characterized by stunted growth, reduced

459

photosynthetic activity, and a disrupted PSII assembly (Kim et al., 2012). Extensive microscopic,

460

fluorescence spectroscopic, and polarographic analyses of Arabidopsis mutants with altered BR

461

responses revealed enlarged thylakoids, smaller PSII complexes, inhibited oxygen evolution, and

462

reduced PSII quantum yields (Krumova et al., 2013). Komatsu et al. (2010) demonstrated that

463

BR deficiency led to an increased accumulation of chlorophyll and photosynthetic proteins that

464

changed the leaf color from green to dark-green. A brassinazole-insensitive pale green2-1

465

(BPG2) gene was proposed to mediate light-regulated chloroplast protein translation under BR-

466

deficient conditions. Exogenous application of BRs in pepper (Capsicum annuum) plants appear

467

to mitigate the deleterious effects of drought on photosynthesis by maintaining or increasing the

468

efficient use of light and NPQ in PSII antennae (Hu et al., 2013). Several other studies have

469

demonstrated BR-induced changes in thylakoid structure and the regulation of PSII and ETC in

470

photosynthesis (Dobrikova et al., 2014; Rothova et al., 2014). Therefore, the role of BRs in PSII

471

damage repair and in modification of thylakoid structural dynamics deserves further

472

investigation.

473

AC C

EP

TE D

M AN U

SC

453

Strigolactones (SLs), another group of phytohormones, may play an important role in

474

plants in the positive regulation of genes associated with harvesting light (Mashiguchi et al.,

475

2009). Several genes encoding light-harvesting chlorophyll a/b binding (LHCB) precursors, 16

ACCEPTED MANUSCRIPT

RuBisCO, and components of PSI and PSII are induced by GR24, a synthetic SL compound

477

(Mayzlish-Gati et al., 2010). In the Arabidopsis SL-signaling max2 mutant, the expression of

478

many genes involved in photosynthesis that are repressed by drought, were up-regulated,

479

especially in response to dehydration, in comparison to WT plants. These data indicate a

480

correlation between the mis-regulation of these genes and the reduced drought tolerance

481

observed in the max2 plants (Ha et al., 2014). Max2 plants may be more sensitive to the high

482

energy demands of photosynthesis, and thus require more resources even when the supply of the

483

resource is inadequate.

RI PT

476

Gibberellins (GAs) have been shown to regulate photosynthesis in addition to their

485

promotion of cell division and seed germination (Huerta et al., 2008; Zhou et al., 2011). In

486

addition, GA and kinetin (a type of CK) were reported to promote PSI and PSII activities by

487

influencing development of the photosynthetic electron transport system in greening cucumber

488

cotyledons (Pedhadiya et al., 1987). Short-term application of GA3 has been found to enhance

489

the net photosynthetic rate as well as the photosynthetic oxygen evolution in isolated broad bean

490

protoplasts (Yuan and Xu, 2001). Transgenic Brassica napus plants with decreased GA

491

bioactivity exhibited a significant increase in photosynthetic capacity (Zhou et al., 2011). In

492

contrast, transgenic citrange (Citrus sinensis x Poncirus trifoliata) plants with higher levels of

493

endogenous GA exhibited significant up-regulation of many genes involved in photosynthesis

494

and water stress alleviation (Huerta et al., 2008). A study in Arabidopsis reported that GA3-

495

treated WT plants or transgenic plants overexpressing a GA-responsive gene showed improved

496

tolerance to salt, oxidative and heat stress (Alonso-Ramirez et al., 2009). The increase in stress

497

tolerance was associated with higher levels of salicylic acid (SA), however, it was unclear if

498

photosynthesis was altered by the GA3 treatment or the increased endogenous level of GA level,

499

which might contribute to the improvement in stress tolerance. Further studies are required to

500

understand the effect of GAs on the relationship of photosynthesis to abiotic stress responses.

M AN U

TE D

EP

AC C

501

SC

484

Abscisic acid (ABA) is the hormone most extensively studied in relation to plant

502

response to abiotic stress. ABA has been reported to directly influence the photosynthetic oxygen

503

evolution connected with the functioning of PS II centers by disrupting the granal chloroplast

504

structure in barley (Maslenkova et al., 1989). Thermostability of photosynthetic apparatus was

505

significantly increased in ABA-treated barley seedlings under heat stress. Light scattering and

506

fluorescence analyses showed a reduced heat damage of the chloroplast ultrastructure and a 17

ACCEPTED MANUSCRIPT

marked decline in heat-induced increase in initial fluorescence (Fo) (Ivanov et al., 1992).

508

Exogenous application of ABA increases the content of total carotenoids, xanthophylls, and

509

chlorophyll in leaves (Barickman et al., 2014), thereby ameliorating the impact of excessive

510

excitation energy on PSII. Application of ABA also protects PSII by inducing an increase in

511

NPQ through the enhancement of xanthophyll-cycle pools and the de-epoxidation of

512

xanthophyll-cycle components (Zhu et al., 2011). The LHCB protein family plays an important

513

role in adaptation to environmental stress (Liu et al., 2013; Voigt et al., 2010). Contrary to the

514

initial belief that ABA was a negative regulator of the expression of LHCB genes (Staneloni et

515

al., 2008), ABA is believed to be required for the full expression of various LHCB genes (Xu et

516

al., 2012). Down-regulation of LHCB genes decreases ABA signaling and response to drought,

517

perhaps partly by modulating ROS homeostasis (Xu et al., 2012). The discrepancy regarding the

518

role of ABA in the regulation of photosynthesis may be attributed to the different experimental

519

systems and methods used to determine photosynthetic output.

SC

M AN U

520

RI PT

507

SA is a common phenolic compound well studied for its biochemical, physiological and plant growth regulation activities (El-Tayeb, 2005; Arfan et al., 2007). However, very little is

522

known about the SA-related mechanisms that control the maintenance of photosynthesis under

523

abiotic stress. Increased accumulation of endogenous SA levels was associated with a sharp

524

decrease in maximum PSII quantum yield (Fv/Fm) in leaves of Phillyrea angustifolia plants

525

exposed to drought (Munné-Bosch and Peñuelas, 2003). Improved photosynthetic capacity was

526

observed in SA-treated wheat plants under high salinity stress (Arfan et al., 2007). Similarly,

527

exogenous application of SA was reported to induce pre-adaptive responses to salt stress,

528

consequently improving and retaining the integrity of cell membranes in barley plants (El-Tayeb,

529

2005). SA-treated grapevine leaves showed accelerated recovery of net photosynthetic rate and

530

donor and acceptor parameters of PSII under heat stress. The authors concluded that the effects

531

of SA pre-treatment might be related to the enhanced levels of chloroplastic heat shock proteins,

532

resulting in improved net photosynetic rate (Wang et al., 2010). In a similar study, SA treatment

533

of wheat leaves enhanced Fv/Fm, actual photochemical efficiency of PSII and photosynthetic

534

electron transport rate, as well as improved net photosynthetic rate and reduced damages of heat

535

and high light stresses on D1 protein and PSII (Zhao et al., 2011).

536 537

AC C

EP

TE D

521

A small number of studies have also indicated that auxin, CK, jasmonic acid (JA) and ethylene may also play important roles in the stability of PSI and PSII, and thus in improvement 18

ACCEPTED MANUSCRIPT

of photosynthesis in plants exposed to abiotic stress (Supplementary Table 2). Transgenic

539

tobacco plants with enhanced endogenous CK level were reported to have increased transcript

540

abundance of genes associated with PSI, PSII and Cytb6f complex under drought stress.

541

Enhanced endogenous level of CK in transgenic plants appeared to promote the protection of

542

photosynthetic apparatus under prolonged drought conditions (Rivero et al., 2007, 2010).

543

Exogenous application of indole acetic acid, a plant hormone of auxin class and GA3 was

544

reported to reduce the effects of excessive copper mainly by maintaining the Fv/Fm and net

545

photosynthetic rate in sunflower plants (Ouzounidou and Ilias, 2005). Similarly, exogenous

546

application of ethylene was reported to regulate the protection of photosynthesis against Ni- and

547

Zn-induced heavy metal stress in Brassica juncea plants (Khan and Khan, 2014). Monitoring the

548

effects of exogenously applied hormones on photosynthesis is a valuable strategy but the

549

transcriptomic analysis of plant response to hormones in photosynthetic mutants would provide a

550

more detailed and comprehensive picture of the role of hormones in photosynthesis under normal

551

and abiotic stress conditions.

SC

M AN U

552

RI PT

538

As described earlier, apart from inducing LHCB genes (Liu et al., 2013; Voigt et al., 2010), ABA has been reported to regulate carotenoid biosynthesis (Barickman et al., 2014), an

554

important LHC pigment. Since both ABA and SLs are carotenoid-derived hormones, potential

555

crosstalk may exist between SLs, ABA, and light harvesting pathways. A BR-dependent, GA-

556

regulated transcriptome was recently found to be enriched with light-regulated genes and genes

557

involved in cell-wall synthesis and photosynthesis. Similar lines of evidence indicate a strong

558

association between various hormones and light-harvesting pathways (Attaran et al., 2014;

559

Cortleven et al., 2014; Staneloni et al., 2008). While BRs have been reported to modulate PSII

560

efficiency and thylakoid architecture (Dobrikova et al., 2014; Krumova et al., 2013; Oh et al.,

561

2011), SLs were demonstrated to act as positive regulators of light harvesting (Mayzlish-Gati et

562

al., 2010). Coordinated crosstalk among SL-, ABA-, and CK-signaling networks regulates the

563

adaptive response of plants to adverse environmental conditions (Ha et al., 2014; Nishiyama et

564

al., 2011a). For instance, increased accumulation of ethylene in drought-treated plants not only

565

enhances senescence but also disrupts ABA-mediated regulation of photosynthesis and leaf

566

growth (Bartoli et al., 2013). Therefore, the ratio of ethylene and ABA determines the plant

567

responses to drought, which further warrants a better understanding of crosstalk between these

568

hormones. Future efforts to determine the points of intersection that are involved in the co-

AC C

EP

TE D

553

19

ACCEPTED MANUSCRIPT

569

regulation of hormones and light on photosynthetic components would help to enable the

570

identification of candidate genes that could be utilized to limit the level of photoinhibition in

571

chloroplasts when plants are subjected to abiotic stress.

572

8. FUTURE PERSPECTIVES

RI PT

573 574

Designing plants with increased levels of abiotic stress tolerance has been a major challenge for

576

researchers. This is partly due to a substantial lack of information on the intricacies of ROS

577

signaling, which limits the ability to determine its regulatory role in abiotic stress response.

578

Previous efforts have largely focused on gene expression studies, but RNA data alone is

579

insufficient. Transcriptomic, proteomic, and metabolomic analyses, novel cellular-imaging

580

techniques, and real-time detection tools would substantially enhance our understanding of ROS

581

signaling.

M AN U

582

SC

575

Increasing the level of chloroplastic antioxidants or the activity of ROS-scavenging enzymes is clearly not a productive approach to improve crops. Alternative approaches, such as

584

modifying LHC and PSII components, have also been suggested as an alternative approach

585

(Horton, 2012; Tikkanen and Aro, 2014). The complex events associated with phosphorylation,

586

dephosphorylation, and the migration of thylakoid membrane proteins and reaction centers,

587

however, remain largely speculative. For example, the roles of STN7 and STN8 are known, but

588

it remains unclear if crosstalk exists in the regulation of these proteins. Moreover, recent findings

589

have suggested that the interaction between STN7 and STN8 is different in monocots and dicots

590

(Nath et al., 2013b). These observations also hold true for the role of minor LHCII antennal

591

proteins, because the phosphorylation of CP29 was reported to be species-specific (Chen et al.,

592

2009). Currently, little information is available on the phosphorylation of other minor antennal

593

proteins of LHCII. Similar ambiguities exist for the proteolytic mechanism and involvement of

594

ATP-dependent proteases in the PSII repair cycle. Contrasting reports and insufficient

595

knowledge on various proteases and intrinsic PSII proteins raises many questions, which only

596

further studies can answer. A major challenge will be to elucidate the linkage between the

597

identified kinases and phosphatases and their influence on the redox state of the photosynthetic

598

ETC. Identification of the substrates of proteases and determining their specificity would help to

599

understand their mode of action. In order to assess damage to photosynthetic machinery at a

AC C

EP

TE D

583

20

ACCEPTED MANUSCRIPT

600

bioenergetic level, techniques such as chlorophyll a transient kinetics analysis are expected to

601

significantly contribute to our knowledge (Figure 4). The redox state of the photosynthetic ETC regulates the complex events during PSII

603

repair cycle as well as several other responses that involve the chloroplast and nuclear gene

604

expression. While the redox state of ETC is regulated by various environmental factors, the

605

chloroplast and nuclear gene expression can be transduced by changes in hormonal signaling

606

pathways. Therefore, changes in photosynthesis and levels of hormones in plants subjected to

607

environmental stresses should not be seen as isolated events, particularly given the emerging

608

evidence that indicates the involvement of phytohormones in PSII damage repair (Cortleven et

609

al., 2014). However, the extensive molecular connections among the signaling networks of

610

various hormones make it a daunting task to unravel the central roles of individual hormones in

611

coordinating the expression of photosynthetic genes and regulation of PSII damage repair.

612

Moreover, the physiological significance of changes in expression recorded in many abiotic

613

stress-responsive genes, particularly in those encoding photosynthetic proteins, is yet to be

614

understood. Genome-wide transcriptomic analyses of integrated hormonal regulatory networks

615

could provide a global view of their functions and molecular links with putative functions in the

616

regulation of photosynthesis. Additionally, it is important to use metabolomic and proteomic

617

approaches for understanding the metabolic responses of plants for photosynthetic acclimation to

618

various abiotic stresses. Such integration of different datasets would improve our understanding

619

of the physiological significance of hormones in regulation of photosynthesis performance under

620

abiotic stresses. Cellular responses to abiotic stresses also involve the transcriptional regulation

621

of photosynthetic metabolism. Several reports suggest that improved abiotic stress tolerance can

622

be achieved by engineering the expression of TFs (Puranik et al., 2012; Zhang et al., 2008)

623

(Supplementary Table 3). The involvement of these TFs in the regulation of genes associated

624

with photosynthesis, however, is still not clear. Taken together, designing plants with improved

625

abiotic stress tolerance and enhanced photosynthetic production will require a more

626

comprehensive understanding of the way photosynthetic machinery mediates the environmental

627

cues and the inherent metabolic signaling. Hence, it is crucial to focus on combining

628

interdisciplinary strategies from different research areas of plant sciences which could facilitate

629

the engineering of plant architecture in a sustainable way.

AC C

EP

TE D

M AN U

SC

RI PT

602

630

21

ACCEPTED MANUSCRIPT

631

ACKNOWLEDGEMENTS

632

This work was supported by the Scientific Research C Grant (no. 25440149) from Japan Society

633

for the Promotion of Science to L.-S. P. Tran.

634

REFERENCES

636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674

Adam, Z., Frottin, F., Espagne, C., Meinnel, T., and Giglione, C. (2011). Interplay between Nterminal methionine excision and FtsH protease is essential for normal chloroplast development and function in Arabidopsis. Plant Cell. 23, 3745-3760. Al-Taweel, K., Iwaki, T., Yabuta, Y., Shigeoka, S., Murata, N., and Wadano, A. (2007). A bacterial transgene for catalase protects translation of d1 protein during exposure of saltstressed tobacco leaves to strong light. Plant Physiol. 145, 258-265. Allahverdiyeva, Y., Suorsa, M., Rossi, F., Pavesi, A., Kater, M.M., Antonacci, A., Tadini, L., Pribil, M., Schneider, A., Wanner, G., et al. (2013). Arabidopsis plants lacking PsbQ and PsbR subunits of the oxygen-evolving complex show altered PSII super-complex organization and short-term adaptive mechanisms. Plant J. 75, 671-684. Allakhverdiev, S.I., Los, D.A., Mohanty, P., Nishiyama, Y., and Murata, N. (2007). Glycinebetaine alleviates the inhibitory effect of moderate heat stress on the repair of photosystem II during photoinhibition. Biochimica et biophysica acta 1767:1363-1371. Allakhverdiev, S.I., and Murata, N. (2004). Environmental stress inhibits the synthesis de novo of proteins involved in the photodamage-repair cycle of Photosystem II in Synechocystis sp. PCC 6803. Biochim. Biophys. Acta. 1657, 23-32. Allakhverdiev, S.I., Nishiyama, Y., Miyairi, S., Yamamoto, H., Inagaki, N., Kanesaki, Y., and Murata, N. (2002). Salt stress inhibits the repair of photodamaged photosystem II by suppressing the transcription and translation of psbA genes in synechocystis. Plant Physiol. 130, 1443-1453. Alonso-Ramirez, A., Rodriguez, D., Reyes, D., Jimenez, J.A., Nicolas, G., Lopez-Climent, M., Gomez-Cadenas, A., and Nicolas, C. (2009). Evidence for a role of gibberellins in salicylic acid-modulated early plant responses to abiotic stress in Arabidopsis seeds. Plant Physiol. 150, 1335-1344. Amunts, A., Toporik, H., Borovikova, A., and Nelson, N. (2010). Structure determination and improved model of plant photosystem I. J. Biol. Chem. 285, 3478-3486. Arfan, M., Athar, H. R., and Ashraf, M. (2007). Does exogenous application of salicylic acid through the rooting medium modulate growth and photosynthetic capacity in two differently adapted spring wheat cultivars under salt stress? J. Plant Physiol. 164, 685694. Aro, E.M., Suorsa, M., Rokka, A., Allahverdiyeva, Y., Paakkarinen, V., Saleem, A., Battchikova, N., and Rintamaki, E. (2005). Dynamics of photosystem II: a proteomic approach to thylakoid protein complexes. J. Exp. Bot. 56, 347-356. Attaran, E., Major, I.T., Cruz, J.A., Rosa, B.A., Koo, A.J., Chen, J., Kramer, D.M., He, S.Y., and Howe, G.A. (2014). Temporal dynamics of growth and photosynthesis suppression in response to jasmonate signaling. Plant Physiol. 165, 1302-1314. Bai, M.Y., Shang, J.X., Oh, E., Fan, M., Bai, Y., Zentella, R., Sun, T.P., and Wang, Z.Y. (2012). Brassinosteroid, gibberellin and phytochrome impinge on a common transcription module in Arabidopsis. Nature Cell Biol. 14, 810-817.

AC C

EP

TE D

M AN U

SC

RI PT

635

22

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Baker, N.R. (2008). Chlorophyll fluorescence: a probe of photosynthesis in vivo. Annu. Rev. Plant Biol. 59, 89-113. Baker, N.R., and Rosenqvist, E. (2004). Applications of chlorophyll fluorescence can improve crop production strategies: an examination of future possibilities. J. Exp. Bot. 55, 16071621. Barickman, T.C., Kopsell, D.A., and Sams, C.E. (2014). Abscisic acid increases carotenoid and chlorophyll concentrations in leaves and fruit of two tomato genotypes. J. Am. Soc. Hort. Sci. 139, 261-266. Bilger, W., and Schreiber, U. (1990) Chlorophyll luminescence as an indicator of stress-induced damage to the photosynthetic apparatus. Effects of heat-stress in isolated chloroplasts. Photosynth. Res. 25, 161–171. Boekema, E.J., van Breemen, J.F.L., van Roon, H., and Dekker, J.P. (2000). Conformational changes in photosystem II supercomplexes upon removal of extrinsic subunits. Biochemistry. 39, 12907-12915. Bonardi, V., Pesaresi, P., Becker, T., Schleiff, E., Wagner, R., Pfannschmidt, T., Jahns, P., and Leister, D. (2005). Photosystem II core phosphorylation and photosynthetic acclimation require two different protein kinases. Nature. 437, 1179-1182. Bricker, T.M., and Frankel, L.K. (2011). Auxiliary functions of the PsbO, PsbP and PsbQ proteins of higher plant Photosystem II: A critical analysis. J. Photochem. Photobiol. B. 104, 165-178. Bricker, T.M., Roose, J.L., Fagerlund, R.D., Frankel, L.K., and Eaton-Rye, J.J. (2012). The extrinsic proteins of Photosystem II. Biochim.Biophys. Acta. 1817, 121-142. Cazzaniga, S., Dall'Osto, L., Kong, S.G., Wada, M., and Bassi, R. (2013). Interaction between avoidance of photon absorption, excess energy dissipation and zeaxanthin synthesis against photooxidative stress in Arabidopsis. Plant J. 76, 568-579. Chao, Y.Y., Chen, C.Y., Huang, W.D., and Kao, C.H. (2010). Salicylic acid-mediated hydrogen peroxide accumulation and protection against Cd toxicity in rice leaves. Plant Soil. 329, 327-337. Che, Y., Fu, A., Hou, X., McDonald, K., Buchanan, B.B., Huang, W., and Luan, S. (2013). Cterminal processing of reaction center protein D1 is essential for the function and assembly of photosystem II in Arabidopsis. Proc. Natl. Acad. Sci. U S A. 110, 1624716252. Chen, J., Burke, J.J., Velten, J., and Xin, Z. (2006). FtsH11 protease plays a critical role in Arabidopsis thermotolerance. Plant J. 48, 73-84. Chen, L., Jia, H., Tian, Q., Du, L., Gao, Y., Miao, X., and Liu, Y. (2012). Protecting effect of phosphorylation on oxidative damage of D1 protein by down-regulating the production of superoxide anion in photosystem II membranes under high light. Photosyn. Res. 112, 141-148. Chen, Y.E., Yuan, S., Du, J.B., Xu, M.Y., Zhang, Z.W., and Lin, H.H. (2009). Phosphorylation of photosynthetic antenna protein CP29 and photosystem II structure changes in monocotyledonous plants under environmental stresses. Biochemistry. 48, 9757-9763. Chi, W., Sun, X., and Zhang, L. (2012). The roles of chloroplast proteases in the biogenesis and maintenance of photosystem II. Biochim. Biophys. Acta. 1817, 239-246. Choudhary, S.P., Yu, J.Q., Yamaguchi-Shinozaki, K., Shinozaki, K., and Tran, L.S. (2012). Benefits of brassinosteroid crosstalk. Trends Plant Sci. 17, 594-605.

AC C

675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719

23

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Cortleven, A., Nitschke, S., Klaumunzer, M., Abdelgawad, H., Asard, H., Grimm, B., Riefler, M., and Schmulling, T. (2014). A novel protective function for cytokinin in the light stress response is mediated by the Arabidopsis histidine kinase2 and Arabidopsis histidine kinase3 receptors. Plant Physiol. 164, 1470-1483. de Bianchi, S., Dall'Osto, L., Tognon, G., Morosinotto, T., and Bassi, R. (2008). Minor antenna proteins CP24 and CP26 affect the interactions between photosystem II subunits and the electron transport rate in grana membranes of Arabidopsis. Plant Cell. 20, 1012-1028. Demmig-Adams, B., Cohu, C.M., Amiard, V., Zadelhoff, G., Veldink, G.A., Muller, O., and Adams, W.W. (2013). Emerging trade-offs - impact of photoprotectants (PsbS, xanthophylls, and vitamin E) on oxylipins as regulators of development and defense. New Phytol. 197, 720-729. Dobrikova, A.G., Vladkova, R.S., Rashkov, G.D., Todinova, S.J., Krumova, S.B., and Apostolova, E.L. (2014). Effects of exogenous 24-epibrassinolide on the photosynthetic membranes under non-stress conditions. Plant Physiol. Biochem. 80, 75-82. Dombrecht, B., Xue, G.P., Sprague, S.J., Kirkegaard, J.A., Ross, J.J., Reid, J.B., Fitt, G.P., Sewelam, N., Schenk, P.M., Manners, J.M., et al. (2007). MYC2 differentially modulates diverse jasmonate-dependent functions in Arabidopsis. Plant Cell. 19, 2225-2245. Dubois, M., Skirycz, A., Claeys, H., Maleux, K., Dhondt, S., De Bodt, S., Vanden Bossche, R., De Milde, L., Yoshizumi, T., Matsui, M., et al. (2013). ETHYLENE RESPONSE FACTOR6 acts as a central regulator of leaf growth under water-limiting conditions in Arabidopsis. Plant Physiol. 162, 319-332. El-Tayeb, M.A. (2005). Response of barley grains to the interactive e. ect of salinity and salicylic acid. Plant Growth Regul. 45, 215-224. Foyer, C.H., Neukermans, J., Queval, G., Noctor, G., and Harbinson, J. (2012). Photosynthetic control of electron transport and the regulation of gene expression. J. Exp. Bot. 63, 16371661. Foyer, C.H., and Noctor, G. (2005). Redox homeostasis and antioxidant signaling: a metabolic interface between stress perception and physiological responses. Plant Cell. 17, 18661875. Foyer, C.H., and Shigeoka, S. (2011). Understanding oxidative stress and antioxidant functions to enhance photosynthesis. Plant Physiol. 155, 93-100. Fristedt, R., and Vener, A.V. (2011). High light induced disassembly of photosystem II supercomplexes in Arabidopsis requires STN7-dependent phosphorylation of CP29. PloS One. 6, e24565. Fristedt, R., Willig, A., Granath, P., Crevecoeur, M., Rochaix, J.D., and Vener, A.V. (2009). Phosphorylation of photosystem II controls functional macroscopic folding of photosynthetic membranes in Arabidopsis. Plant Cell. 21, 3950-3964. Goltsev, V., Yordanov, I., Stojanova, T., and Popov, O. (1987) High temperature damage and acclimation of the photosynthetic apparatus. II. The effect of mono- and divalent cations and pH on the temperature sensitivity of some functional characteristics of chloroplasts isolated from heat-acclimated and non-acclimated bean plants. Planta 170, 478–488. Goltsev, V., Zaharieva, I., Chernev, P., and Strasser, R.J. (2009). Delayed fluorescence in photosynthesis. Photosyn. Res. 101, 217-232. Goltsev, V., Zaharieva, I., Chernev, P., Kouzmanova, M., Kalaji, H.M., Yordanov, I., Krasteva, V., Alexandrov, V., Stefanov, D., Allakhverdiev, S.I., et al. (2012). Drought-induced modifications of photosynthetic electron transport in intact leaves: analysis and use of

AC C

720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765

24

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

neural networks as a tool for a rapid non-invasive estimation. Biochim. Biophys. Acta. 1817, 1490-1498. Gombos, Z., Wada, H., and Murata, N. (1994). The Recovery of Photosynthesis from LowTemperature Photoinhibition Is Accelerated by the Unsaturation of Membrane-Lipids - a Mechanism of Chilling Tolerance. Proc. Natl. Acad. Sci. U S A. 91, 8787-8791. Gratao, P.L., Monteiro, C.C., Rossi, M.L., Martinelli, A.P., Peres, L.E.P., Medici, L.O., Lea, P.J., and Azevedo, R.A. (2009). Differential ultrastructural changes in tomato hormonal mutants exposed to cadmium. Environ. Exp. Bot. 67, 387-394. Grieco, M., Tikkanen, M., Paakkarinen, V., Kangasjarvi, S., and Aro, E.M. (2012). Steady-state phosphorylation of light-harvesting complex II proteins preserves photosystem I under fluctuating white light. Plant Physiol. 160, 1896-1910. Gururani, M.A., Upadhyaya, C.P., Strasser, R.J., Woong, Y.J., and Park, S.W. (2012). Physiological and biochemical responses of transgenic potato plants with altered expression of PSII manganese stabilizing protein. Plant Physiol. Biochem. 58, 182-194. Gururani, M.A., Upadhyaya, C.P., Strasser, R.J., Yu, J.W., and Park, S.W. (2013). Evaluation of abiotic stress tolerance in transgenic potato plants with reduced expression of PSII manganese stabilizing protein. Plant Sci. 198, 7-16. Gururani, M.A., Venkatesh, J., Ganesan, M., Strasser, R.J., Han, Y.J., Kim, J.I., Lee, H.Y., and Song, P.S. (2015) In vivo assessment of cold tolerance through chlorophyll-a fluorescence in transgenic zoysiagrass expressing mutant phytochrome A. PLoS ONE (In Press) doi: 10.1371/journal.pone.0127200 Ha, C.V., Leyva-Gonzalez, M.A., Osakabe, Y., Tran, U.T., Nishiyama, R., Watanabe, Y., Tanaka, M., Seki, M., Yamaguchi, S., Dong, N.V., et al. (2014). Positive regulatory role of strigolactone in plant responses to drought and salt stress. Proc. Natl. Acad. Sci. U S A. 111, 851-856. Hakala, M., Tuominen, I., Keranen, M., Tyystjarvi, T., and Tyystjarvi, E. (2005). Evidence for the role of the oxygen-evolving manganese complex in photoinhibition of photosystem II. Biochim. Biophys. Acta. 1706, 68-80. Haussuhl, K., Andersson, B., and Adamska, I. (2001). A chloroplast DegP2 protease performs the primary cleavage of the photodamaged D1 protein in plant photosystem II. EMBO J. 20, 713-722. Havaux, M., Eymery, F., Porfirova, S., Rey, P., and Dormann, P. (2005). Vitamin E protects against photoinhibition and photooxidative stress in Arabidopsis thaliana. Plant Cell. 17, 3451-3469. Henmi, T., Miyao, M., and Yamamoto, Y. (2004). Release and reactive-oxygen-mediated damage of the oxygen-evolving complex subunits of PSII during photoinhibition. Plant Cell Physiol. 45, 243-250. Herbstova, M., Tietz, S., Kinzel, C., Turkina, M.V., and Kirchhoff, H. (2012). Architectural switch in plant photosynthetic membranes induced by light stress. Proc. Natl. Acad. Sci. U S A. 109, 20130-20135. Hichri, I., Muhovski, Y., Zizkova, E., Dobrev, P.I., Franco-Zorrilla, J.M., Solano, R., LopezVidriero, I., Motyka, V., and Lutts, S. (2014). The Solanum lycopersicum Zinc Finger2 cysteine-2/histidine-2 repressor-like transcription factor regulates development and tolerance to salinity in tomato and Arabidopsis. Plant Physiol. 164, 1967-1990. Horton, P. (2012). Optimization of light harvesting and photoprotection: molecular mechanisms and physiological consequences. Phil. Trans. Roy. Soc. Lond. B. 367, 3455-3465.

AC C

766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811

25

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Hu, W.H., Yan, X.H., Xiao, Y.A., Zeng, J.J., Qi, H.J., and Ogweno, J.O. (2013). 24epibrassinosteroid alleviate drought-induced inhibition of photosynthesis in Capsicum annuum. Sci. Hort. 150, 232-237. Huerta, L., Forment, J., Gadea, J., Fagoaga, C., Pena, L., Perez-Amador, M.A., and GarciaMartinez, J.L. (2008). Gene expression analysis in citrus reveals the role of gibberellins on photosynthesis and stress. Plant Cell Environ. 31, 1620-1633. Ifuku, K. (2014). The PsbP and PsbQ family proteins in the photosynthetic machinery of chloroplasts. Plant Physiol Biochem. 81, 108-114. Ishihara, S., Takabayashi, A., Ido, K., Endo, T., Ifuku, K., and Sato, F. (2007). Distinct functions for the two PsbP-like proteins PPL1 and PPL2 in the chloroplast thylakoid lumen of Arabidopsis. Plant Physiol. 145, 668-679. Itoh, S. (1980) Correlation between the time course of millisecond delayed fluorescence and that of prompt fluorescence at low temperature in uncoupled spinach chloroplasts. Plant Cell Physiol. 21, 873–884. Ivanov, A.G., Kitcheva, M.I., Christov, A.M., and Popova, L.P. (1992). Effects of abscisic acid treatment on the thermostability of the photosynthetic apparatus in barley chloroplasts. Plant Physiol. 98, 1228-1232. Ivanov, A.G., Morgan, R.M., Gray, G.R., Velitchkova, M.Y., and Huner, N.P.A. (1998). Temperature/light dependent development of selective resistance to photoinhibition of photosystem I. FEBS lett.430, 288-292. Jin, S.H., Li, X.Q., Hu, J.Y., and Wang, J.G. (2009). Cyclic electron flow around photosystem I is required for adaptation to high temperature in a subtropical forest tree, Ficus concinna. J. Zhejiang Univ. Sci. B 10, 784-790. Johnson, M.P., Goral, T.K., Duffy, C.D., Brain, A.P., Mullineaux, C.W., and Ruban, A.V. (2011). Photoprotective energy dissipation involves the reorganization of photosystem II lightharvesting complexes in the grana membranes of spinach chloroplasts. Plant Cell. 23, 1468-1479. Joliot, P., and Johnson, G.N. (2011). Regulation of cyclic and linear electron flow in higher plants. Proc. Natl. Acad. Sci. U.S.A. 108, 13317-13322. Kangasjarvi, S., Neukermans, J., Li, S., Aro, E.M., and Noctor, G. (2012). Photosynthesis, photorespiration, and light signalling in defence responses. J. Exp. Bot. 63, 1619-1636. Kapri-Pardes, E., Naveh, L., and Adam, Z. (2007). The thylakoid lumen protease Deg1 is involved in the repair of photosystem II from photoinhibition in Arabidopsis. Plant Cell. 19, 1039-1047. Kato, Y., Miura, E., Ido, K., Ifuku, K., and Sakamoto, W. (2009). The variegated mutants lacking chloroplastic FtsHs are defective in D1 degradation and accumulate reactive oxygen species. Plant Physiol. 151, 1790-1801. Kato, Y., and Sakamoto, W. (2014). Phosphorylation of photosystem II core proteins prevents undesirable cleavage of D1 and contributes to the fine-tuned repair of photosystem II. Plant J. 79, 312-321. Kato, Y., Sun, X.W., Zhang, L.X., and Sakamoto, W. (2012). Cooperative D1 degradation in the photosystem ii repair mediated by chloroplastic proteases in Arabidopsis. Plant Physiol. 159, 1428-1439. Khan, M.I., and Khan, N.A. (2014). Ethylene reverses photosynthetic inhibition by nickel and zinc in mustard through changes in PS II activity, photosynthetic nitrogen use efficiency, and antioxidant metabolism. Protoplasma. 251, 1007-1019.

AC C

812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857

26

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Kim, T.W., Michniewicz, M., Bergmann, D.C., and Wang, Z.Y. (2012). Brassinosteroid regulates stomatal development by GSK3-mediated inhibition of a MAPK pathway. Nature. 482, 419-422. Kirchhoff, H., Hall, C., Wood, M., Herbstova, M., Tsabari, O., Nevo, R., Charuvi, D., Shimoni, E., and Reich, Z. (2011). Dynamic control of protein diffusion within the granal thylakoid lumen. Proc. Natl. Acad. Sci. U S A. 108, 20248-20253. Klimmek, F., Sjodin, A., Noutsos, C., Leister, D., and Jansson, S. (2006). Abundantly and rarely expressed Lhc protein genes exhibit distinct regulation patterns in plants. Plant Physiol. 140, 793-804. Komatsu, T., Kawaide, H., Saito, C., Yamagami, A., Shimada, S., Nakazawa, M., Matsui, M., Nakano, A., Tsujimoto, M., Natsume, M., Abe, H., Asami, T., and Nakano, T. (2010). The chloroplast protein BPG2 functions in brassinosteroid-mediated post-transcriptional accumulation of chloroplast rRNA. Plant J. 61, 409–422. Kornyeyev, D., Logan, B.A., Allen, R.D., and Holaday, A.S. (2003). Effect of chloroplastic overproduction of ascorbate peroxidase on photosynthesis and photoprotection in cotton leaves subjected to low temperature photoinhibition. Plant Sci. 165, 1033-1041. Krause, G.H., and Weis, E. (1991) Chlorophyll fluorescence and photosynthesis: the basics. Annu. Rev. Plant Physiol. Plant Mol. Biol. 42, 313–349. Krumova, S., Zhiponova, M., Dankov, K., Velikova, V., Balashev, K., Andreeva, T., Russinova, E., and Taneva, S. (2013). Brassinosteroids regulate the thylakoid membrane architecture and the photosystem II function. J. Photochem. Photobiol. B. 126, 97-104. Kudoh, H., and Sonoike, K. (2002). Irreversible damage to photosystem I by chilling in the light: cause of the degradation of chlorophyll after returning to normal growth temperature. Planta 215, 541-548. Le, D.T., Nishiyama, R., Watanabe, Y., Vankova, R., Tanaka, M., Seki, M., Ham le, H., Yamaguchi-Shinozaki, K., Shinozaki, K., and Tran, L.S. (2012). Identification and Expression Analysis of Cytokinin Metabolic Genes in Soybean under Normal and Drought Conditions in Relation to Cytokinin Levels. PloS One 7, e42411. Li, X.G., Duan, W., Meng, Q.W., Zou, Q., and Zhao, S.J. (2004). The function of chloroplastic NAD (P) H dehydrogenase in tobacco during chilling stress under low irradiance. Plant Cell Physiol. 45, 103-108. Lim, P.O., Lee, I.C., Kim, J., Kim, H.J., Ryu, J.S., Woo, H.R., and Nam, H.G. (2010). Auxin response factor 2 (ARF2) plays a major role in regulating auxin-mediated leaf longevity. J. Exp. Bot. 61, 1419-1430. Liu, R., Xu, Y.H., Jiang, S.C., Lu, K., Lu, Y.F., Feng, X.J., Wu, Z., Liang, S., Yu, Y.T., Wang, X.F., et al. (2013). Light-harvesting chlorophyll a/b-binding proteins, positively involved in abscisic acid signalling, require a transcription repressor, WRKY40, to balance their function. J. Exp. Bot. 64, 5443-5456. Lucinski, R., Misztal, L., Samardakiewicz, S., and Jackowski, G. (2011). The thylakoid protease Deg2 is involved in stress-related degradation of the photosystem II light-harvesting protein Lhcb6 in Arabidopsis thaliana. New Phytol. 192, 74-86. Lundin, B., Hansson, M., Schoefs, B., Vener, A.V., and Spetea, C. (2007). The Arabidopsis PsbO2 protein regulates dephosphorylation and turnover of the photosystem II reaction centre D1 protein. Plant J. 49, 528-539. Lundin, B., Nurmi, M., Rojas-Stuetz, M., Aro, E.M., Adamska, I., and Spetea, C. (2008). Towards understanding the functional difference between the two PsbO isoforms in

AC C

858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903

27

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Arabidopsis thaliana-insights from phenotypic analyses of psbo knockout mutants. Photosyn. Res. 98, 405-414. Maruta, T., Tanouchi, A., Tamoi, M., Yabuta, Y., Yoshimura, K., Ishikawa, T., and Shigeoka, S. (2010). Arabidopsis chloroplastic ascorbate peroxidase isoenzymes play a dual role in photoprotection and gene regulation under photooxidative stress. Plant Cell Physiol. 51, 190-200. Mashiguchi, K., Sasaki, E., Shimada, Y., Nagae, M., Ueno, K., Nakano, T., Yoneyama, K., Suzuki, Y., and Asami, T. (2009). Feedback-regulation of strigolactone biosynthetic genes and strigolactone-regulated genes in Arabidopsis. Biosci. Biotechnol. Biochem. 73, 2460-2465. Maslenkova, L.T., Zanev, Y., and Popova, L.P. (1989). Effect of abscisic acid on the photosynthetic oxygen evolution in barley chloroplasts. Photosyn Res. 21, 45-50. Maxwell, K., and Johnson, G.N. (2000). Chlorophyll fluorescence - a practical guide. J. Exp. Bot. 51, 659-668. Mayzlish-Gati, E., LekKala, S.P., Resnick, N., Wininger, S., Bhattacharya, C., Lemcoff, J.H., Kapulnik, Y., and Koltai, H. (2010). Strigolactones are positive regulators of lightharvesting genes in tomato. J. Exp. Bot. 61, 3129-3136. Miyagawa, Y., Tamoi, M., and Shigeoka, S. (2000). Evaluation of the defense system in chloroplasts to photooxidative stress caused by paraquat using transgenic tobacco plants expressing catalase from Escherichia coli. Plant Cell Physiol. 41, 311-320. Miyao, M. (1994). Involvement of active oxygen species in degradation of the D1 protein under strong illumination in isolated subcomplexes of photosystem II. Biochemistry. 33, 97229730. Miyao, M., Ikeuchi, M., Yamamoto, N., and Ono, T. (1995). Specific degradation of the D1 protein of photosystem II by treatment with hydrogen peroxide in darkness: implications for the mechanism of degradation of the D1 protein under illumination. Biochemistry. 34, 10019-10026. Mladenova, I., Maini, C., Mallegni, P., Goltsev, V., Vladova, R., Vinarova, K., and Rocheva, S. (1998) Siapton—an amino-acid-based biostimulant reducing osmostress metabolic changes in maize. AgroFood-Industry Hi-Tech 9, 18–22. Mohanty, P., Allakhverdiev, S.I., and Murata, N. (2007). Application of low temperatures during photoinhibition allows characterization of individual steps in photodamage and the repair of photosystem II. Photosyn. Res. 94, 217-224. Munne-Bosch, S., and Penuelas, J. (2003). Photo-and antioxidative protection, and a role for salicylic acid during drought and recovery in field-grown Phillyrea angustifolia plants. Planta 217, 758-766. Murata, N., Takahashi, S., Nishiyama, Y., and Allakhverdiev, S.I. (2007). Photoinhibition of photosystem II under environmental stress. Biochim. Biophys. Acta. 1767, 414-421. Murgia, I., Tarantino, D., Vannini, C., Bracale, M., Carravieri, S., and Soave, C. (2004). Arabidopsis thaliana plants overexpressing thylakoidal ascorbate peroxidase show increased resistance to Paraquat-induced photooxidative stress and to nitric oxideinduced cell death. Plant J. 38, 940-953. Nageswara, R.M., Suorsa, M., Rantala, M., Aro, E.M., and Tikkanen, M. (2015). Plants actively avoid state-transitions upon changes in light intensity-role of light-harvesting complex II protein dephosphorylation in high light. Plant Physiol. pp-00488.

AC C

904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948

28

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Nath, K., Jajoo, A., Poudyal, R.S., Timilsina, R., Park, Y.S., Aro, E.M., Nam, H.G., and Lee, C.H. (2013a). Towards a critical understanding of the photosystem II repair mechanism and its regulation during stress conditions. FEBS Lett. 587, 3372-3381. Nath, K., Poudyal, R.S., Eom, J.S., Park, Y.S., Zulfugarov, I.S., Mishra, S.R., Tovuu, A., Ryoo, N., Yoon, H.S., Nam, H.G., et al. (2013b). Loss-of-function of OsSTN8 suppresses the photosystem II core protein phosphorylation and interferes with the photosystem II repair mechanism in rice (Oryza sativa). Plant J. 76, 675-686. Nellaepalli, S., Zsiros, O., Toth, T., Yadavalli, V., Garab, G., Subramanyam, R., and Kovacs, L. (2014). Heat- and light-induced detachment of the light harvesting complex from isolated photosystem I supercomplexes. J. Photochem. Photobiol. B. 137, 13-20. Nguyen, C.V., Vrebalov, J.T., Gapper, N.E., Zheng, Y., Zhong, S., Fei, Z., and Giovannoni, J.J. (2014). Tomato GOLDEN2-LIKE transcription factors reveal molecular gradients that function during fruit development and ripening. Plant Cell. 26, 585-601. Nishiyama, R., Watanabe, Y., Fujita, Y., Le, D.T., Kojima, M., Werner, T., Vankova, R., Yamaguchi-Shinozaki, K., Shinozaki, K., Kakimoto, T., et al. (2011a). Analysis of cytokinin mutants and regulation of cytokinin metabolic genes reveals important regulatory roles of cytokinins in drought, salt and abscisic acid responses, and abscisic acid biosynthesis. Plant Cell. 23, 2169-2183. Nishiyama, Y., Allakhverdiev, S.I., and Murata, N. (2006). A new paradigm for the action of reactive oxygen species in the photoinhibition of photosystem II. Biochim. Biophys. Acta. 1757, 742-749. Nishiyama, Y., Allakhverdiev, S.I., and Murata, N. (2011b). Protein synthesis is the primary target of reactive oxygen species in the photoinhibition of photosystem II. Physiol. Plant. 142, 35-46. Nishiyama, Y., Allakhverdiev, S.I., Yamamoto, H., Hayashi, H., and Murata, N. (2004). Singlet oxygen inhibits the repair of photosystem II by suppressing the translation elongation of the D1 protein in Synechocystis sp. PCC 6803. Biochemistry. 43, 11321-11330. Nishiyama, Y., and Murata, N. (2014). Revised scheme for the mechanism of photoinhibition and its application to enhance the abiotic stress tolerance of the photosynthetic machinery. Appl. Microbiol. Biotechnol. 98, 8777-8796. Nixon, P.J., Michoux, F., Yu, J., Boehm, M., and Komenda, J. (2010). Recent advances in understanding the assembly and repair of photosystem II. Annal. Bot. 106, 1-16. Noctor, G., Mhamdi, A., and Foyer, C.H. (2014). The roles of reactive oxygen metabolism in drought: not so cut and dried. Plant Physiol. 164, 1636-1648. Obayashi, T., Hayashi, S., Saeki, M., Ohta, H., and Kinoshita, K. (2009). ATTED-II provides coexpressed gene networks for Arabidopsis. Nucleic Acid. Res. 37, D987-991. Oh, M.H., Sun, J., Oh, D.H., Zielinski, R.E., Clouse, S.D., and Huber, S.C. (2011). Enhancing Arabidopsis leaf growth by engineering the BRASSINOSTEROID INSENSITIVE1 receptor kinase. Plant Physiol. 157, 120-131. Oh, M.H., Wang, X., Wu, X., Zhao, Y., Clouse, S.D., and Huber, S.C. (2010). Autophosphorylation of Tyr-610 in the receptor kinase BAK1 plays a role in brassinosteroid signaling and basal defense gene expression. Proc. Natl. Acad. Sci. U S A. 107, 17827-17832. Ohnishi, N., and Murata, N. (2006). Glycinebetaine counteracts the inhibitory effects of salt stress on the degradation and synthesis of D1 protein during photoinhibition in Synechococcus sp. PCC 7942. Plant Physiol. 141, 758-765.

AC C

949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994

29

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Osakabe, Y., Yamaguchi-Shinozaki, K., Shinozaki, K., and Tran, L.S. (2014). ABA control of plant macroelement membrane transport systems in response to water deficit and high salinity. New Phytol. 202, 35-49. Ouzounidou, G., and Ilias, I. (2005). Hormone-induced protection of sunflower photosynthetic apparatus against copper toxicity. Biol. Plant. 49, 223-228. Papageorgiou, G.C., Tsimilli-Michael, M., and Stamatakis, K. (2007). The fast and slow kinetics of chlorophyll a fluorescence induction in plants, algae and cyanobacteria: a viewpoint. Photosyn. Res. 94, 275-290. Pedhadiya, M.D., Vaishnav, P.P., and Singh, Y.D. (1987). Development of photosynthetic electron transport reactions under the influence of phytohormones and nitrate nutrition in greening cucumber cotyledons. Photosyn. Res. 13, 159-165. Plekhanov, S.E., and Chemeris, I.K. (2003) Early toxic effect of zinc, cobalt, and cadmium on photosynthetic activity of green alga Chlorella pyrenoidosa Chick S-39. Izvestiia Akademii nauk Ser Biol/RAN 5, 610–616. Pribil, M., Labs, M., and Leister, D. (2014). Structure and dynamics of thylakoids in land plants. J. Exp. Bot. 65, 1955-1972. Pribil, M., Pesaresi, P., Hertle, A., Barbato, R., and Leister, D. (2010). Role of plastid protein phosphatase TAP38 in LHCII dephosphorylation and thylakoid electron flow. PLoS Biol. 8, e1000288. Puranik, S., Sahu, P.P., Srivastava, P.S., and Prasad, M. (2012). NAC proteins: regulation and role in stress tolerance. Trends Plant Sci. 17, 369-381. Puthiyaveetil, S., Woodiwiss, T., Knoerdel, R., Zia, A., Wood, M., Hoehner, R., and Kirchhoff, H. (2014). Significance of the photosystem II core phosphatase PBCP for plant viability and protein repair in thylakoid membranes. Plant Cell Physiol. 55, 1245-1254. Ramel, F., Birtic, S., Cuine, S., Triantaphylides, C., Ravanat, J.L., and Havaux, M. (2012). Chemical quenching of singlet oxygen by carotenoids in plants. Plant Physiol. 158, 12671278. Ranjan, S., Singh, R., Singh, M., Pathre, U.V., and Shirke, P.A. (2014). Characterizing photoinhibition and photosynthesis in juvenile-red versus mature-green leaves of Jatropha curcas L. Plant Physiol. Biochem. 79, 48-59. Reiland, S., Messerli, G., Baerenfaller, K., Gerrits, B., Endler, A., Grossmann, J., Gruissem, W., and Baginsky, S. (2009). Large-scale Arabidopsis phosphoproteome profiling reveals novel chloroplast kinase substrates and phosphorylation networks. Plant Physiol. 150, 889-903. Rivero, R.M., Kojima, M., Gepstein, A., Sakakibara, H., Mittler, R., Gepstein, S., and Blumwald, E. (2007). Delayed leaf senescence induces extreme drought tolerance in a flowering plant. Proc. Natl. Acad. Sci. 104, 19631-19636. Rivero, R.M., Gimeno, J., Van Deynze, A., Walia, H., and Blumwald, E. (2010). Enhanced cytokinin synthesis in tobacco plants expressing PSARK:: IPT prevents the degradation of photosynthetic protein complexes during drought. Plant Cell Physiol. 51, 1929-1941. Rochaix, J.D. (2011). Regulation of photosynthetic electron transport. Biochim Biophys. Acta 1807, 375-383. Rochaix, J.D. (2014). Regulation and dynamics of the light-harvesting system. Annu. Rev. Plant Biol. 65, 287-309.

AC C

995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038

30

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Rothova, O., Hola, D., Kocova, M., Tumova, L., Hnilicka, F., Hnilickova, H., Kamlar, M., and Macek, T. (2014). 24-epibrassinolide and 20-hydroxyecdysone affect photosynthesis differently in maize and spinach. Steroids. 85, 44-57. Sakamoto, W. (2006). Protein degradation machineries in plastids. Annu. Rev.Plant Biol. 57, 599-621. Samol, I., Shapiguzov, A., Ingelsson, B., Fucile, G., Crevecoeur, M., Vener, A.V., Rochaix, J.D., and Goldschmidt-Clermont, M. (2012). Identification of a photosystem II phosphatase involved in light acclimation in Arabidopsis. Plant Cell. 24, 2596-2609. Schansker, G., Toth, S.Z., Holzwarth, A.R., and Garab, G. (2014). Chlorophyll a fluorescence: beyond the limits of the QA model. Photosyn. Res. 120, 43-58. Schmutz, J., Cannon, S.B., Schlueter, J., Ma, J., Mitros, T., Nelson, W., Hyten, D.L., Song, Q., Thelen, J.J., Cheng, J., et al. (2010). Genome sequence of the palaeopolyploid soybean. Nature. 463, 178-183. Schuhmann, H., and Adamska, I. (2012). Deg proteases and their role in protein quality control and processing in different subcellular compartments of the plant cell. Physiol Plant. 145, 224-234. Shan, D.P., Huang, J.G., Yang, Y.T., Guo, Y.H., Wu, C.A., Yang, G.D., Gao, Z., and Zheng, C.C. (2007). Cotton GhDREB1 increases plant tolerance to low temperature and is negatively regulated by gibberellic acid. New Phytol. 176, 70-81. Shapiguzov, A., Ingelsson, B., Samol, I., Andres, C., Kessler, F., Rochaix, J.D., Vener, A.V., and Goldschmidt-Clermont, M. (2010). The PPH1 phosphatase is specifically involved in LHCII dephosphorylation and state transitions in Arabidopsis. Proc. Natl. Acad. Sci. U S A. 107, 4782-4787. Shikanai, T., Takeda, T., Yamauchi, H., Sano, S., Tomizawa, K.I., Yokota, A., and Shigeoka, S. (1998). Inhibition of ascorbate peroxidase under oxidative stress in tobacco having bacterial catalase in chloroplasts. FEBS Lett. 428, 47-51. Sonoike, K. (1995). Selective photoinhibition of photosystem I in isolated thylakoid membranes from cucumber and spinach. Plant Cell Physiol. 36, 825-830. Sonoike, K. (2011). Photoinhibition of photosystem I. Physiol. Plant. 142, 56-64.Spetea, C., Rintamaki, E., and Schoefs, B. (2014). Changing the light environment: chloroplast signalling and response mechanisms. Phil. Trans. Roy. Soc. Lond. B. 369, 20130220. Stahl, W., and Sies, H. (2003). Antioxidant activity of carotenoids. Mol. Asp. Med. 24, 345-351. Staneloni, R.J., Rodriguez-Batiller, M.J., and Casal, J.J. (2008). Abscisic acid, high-light, and oxidative stress down-regulate a photosynthetic gene via a promoter motif not involved in phytochrome-mediated transcriptional regulation. Mol. Plant 1, 75-83. Stirbet, A., and Govindjee. (2011). On the relation between the Kautsky effect (chlorophyll a fluorescence induction) and photosystem II: Basics and applications of the OJIP fluorescence transient. J. Photochem. Photobiol. B. 104, 236-257. Stirbet, A. (2011). On the relation between the Kautsky effect (chlorophyll a fluorescence induction) and photosystem II: basics and applications of the OJIP fluorescence transient. J. Photochem. Photobiol. B 104, 236-257. Strasser, B.J., and Strasser, R.J. (1995). Measuring fast fluorescence transients to address environmental questions: the JIP test In: Photosynthesis: from Light to Biosphere--Mathis, P., ed. The Netherlands: Kluwer Academic Publisher. 977-980.

AC C

1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082

31

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Strasser, R.J., Tsimilli-Michael, M., and Srivastava, A. (2004). Analysis of the chlorophyll a fluorescence transient. In: Chlorophyll a fluorescence: a signature of photosynthesis. -Papageogiou, G.C., and Govindjee, eds. The Netherlands: Springer. 321-362. Sun, X., Fu, T., Chen, N., Guo, J., Ma, J., Zou, M., Lu, C., and Zhang, L. (2010a). The stromal chloroplast Deg7 protease participates in the repair of photosystem II after photoinhibition in Arabidopsis. Plant Physiol. 152, 1263-1273. Sun, X.W., Ouyang, M., Guo, J.K., Ma, J.F., Lu, C.M., Adam, Z., and Zhang, L.X. (2010b). The thylakoid protease Deg1 is involved in photosystem-II assembly in Arabidopsis thaliana. Plant J. 62, 240-249. Sun, X.W., Peng, L.W., Guo, J.K., Chi, W., Ma, J.F., Lu, C.M., and Zhang, L.X. (2007). Formation of DEG5 and DEG8 complexes and their involvement in the degradation of photodamaged photosystem II reaction center D1 protein in Arabidopsis. Plant Cell. 19, 1347-1361. Suorsa, M., Järvi, S., Grieco, M., Nurmi, M., Pietrzykowska, M., Rantala, M., Kangasjärvi S, Paakkarinen, V., Tikkanen, M., Jansson, S., and Aro, E.M. (2012). PROTON GRADIENT REGULATION5 is essential for proper acclimation of Arabidopsis photosystem I to naturally and artificially fluctuating light conditions. Plant Cell 24, 2934-2948. Takahashi, S., and Badger, M.R. (2011). Photoprotection in plants: a new light on photosystem II damage. Trends Plant Sci. 16, 53-60. Takahashi, S., and Murata, N. (2005). Interruption of the Calvin cycle inhibits the repair of Photosystem II from photodamage. Biochim. Biophys. Acta. 1708, 352-361. Takahashi, S., and Murata, N. (2006). Glycerate-3-phosphate, produced by CO2 fixation in the Calvin cycle, is critical for the synthesis of the D1 protein of photosystem II. Biochim. Biophys. Acta. 1757, 198-205. Takahashi, S., and Murata, N. (2008). How do environmental stresses accelerate photoinhibition? Trends Plant Sci. 13, 178-182. Takahashi, S., Whitney, S.M., and Badger, M.R. (2009). Different thermal sensitivity of the repair of photodamaged photosynthetic machinery in cultured Symbiodinium species. Proc. Natl. Acad. Sci. U S A. 106, 3237-3242. Terashima, I., Funayama, S., and Sonoike, K. (1994). The site of photoinhibition in leaves of Cucumis sativus L. at low temperatures is photosystem I, not photosystem II. Planta 193, 300-306. Tikkanen, M., and Aro, E.M. (2014). Integrative regulatory network of plant thylakoid energy transduction. Trends Plant Sci. 19, 10-17. Tikkanen, M., Gollan, P.J., Mekala, N.R., Isojarvi, J., and Aro, E.M. (2014a). Light-harvesting mutants show differential gene expression upon shift to high light as a consequence of photosynthetic redox and reactive oxygen species metabolism. Phil. Trans. Roy. Soc. Lond. B. 369, 20130229. Tikkanen, M., Gollan, P.J., Suorsa, M., Kangasjarvi, S., and Aro, E.M. (2012). STN7 operates in retrograde signaling through controlling redox balance in the electron transfer chain. Front. Plant Sci. 3, 277. Tikkanen, M., Grieco, M., and Aro, E.M. (2011). Novel insights into plant light-harvesting complex II phosphorylation and 'state transitions'. Trends Plant Sci. 16, 126-131.

AC C

1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126

32

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Tikkanen, M., Grieco, M., Kangasjarvi, S., and Aro, E.M. (2010). Thylakoid protein phosphorylation in higher plant chloroplasts optimizes electron transfer under fluctuating light. Plant Physiol. 152, 723-735. Tikkanen, M., Mekala, N.R., and Aro, E.M. (2014b). Photosystem II photoinhibition-repair cycle protects Photosystem I from irreversible damage. Biochim. Biophys. Acta. 1837, 210-215. Toledo-Ortiz, G., Johansson, H., Lee, K.P., Bou-Torrent, J., Stewart, K., Steel, G., RodriguezConcepcion, M., and Halliday, K.J. (2014). The HY5-PIF regulatory module coordinates light and temperature control of photosynthetic gene transcription. PLoS Genet. 10, e1004416. Tomita, M., Ifuku, K., Sato, F., and Noguchi, T. (2009). FTIR evidence that the PsbP extrinsic protein induces protein conformational changes around the oxygen-evolving Mn cluster in Photosystem II. Biochemistry 48:6318-6325. Toth, S.Z., Nagy, V., Puthur, J.T., Kovacs, L., and Garab, G. (2011). The physiological role of ascorbate as photosystem ii electron donor: protection against photoinactivation in heatstressed leaves. Plant Physiol. 156, 382-392. Triantaphylides, C., and Havaux, M. (2009). Singlet oxygen in plants: production, detoxification and signaling. Trends Plant Sci. 14, 219-228. Upadhyaya, C.P., Venkatesh, J., Gururani, M.A., Asnin, L., Sharma, K., Ajappala, H., and Park, S.W. (2011). Transgenic potato overproducing L-ascorbic acid resisted an increase in methylglyoxal under salinity stress via maintaining higher reduced glutathione level and glyoxalase enzyme activity. Biotechnol. Lett. 33, 2297-2307. Valikhanov, K.M., Zakhidov, E.A., Zakhidova, M.A., Kasymdzhanov, M.A., Kurbanov, S.S., Nematov, S.K., and Khabibullaev, P.K. (2002) Kinetics of photoinhibition and delayed fluorescence in the plant photosynthetic system. Doklady Biochem. Biophys. 387, 331– 334. van der Hoorn, R.A. (2008). Plant proteases: from phenotypes to molecular mechanisms. Annu. Rev. Plant Biol. 59, 191-223. Voigt, C., Oster, U., Bornke, F., Jahns, P., Dietz, K.J., Leister, D., and Kleine, T. (2010). Indepth analysis of the distinctive effects of norflurazon implies that tetrapyrrole biosynthesis, organellar gene expression and ABA cooperate in the GUN-type of plastid signalling. Physiol. Plant. 138, 503-519. Wang, L.J., Fan, L., Loescher, W., Duan, W., Liu, G.J., Cheng, J.S., Luo, H.B., and Li, S.H. (2010). Salicylic acid alleviates decreases in photosynthesis under heat stress and accelerates recovery in grapevine leaves. BMC Plant Biol. 10, 34. Wang, Y., Noguchi, K., Ono, N., Inoue, S., Terashima, I., and Kinoshita, T. (2014). Overexpression of plasma membrane H+-ATPase in guard cells promotes light-induced stomatal opening and enhances plant growth. Proc. Natl. Acad. Sci. U S A. 111, 533-538. Waters, M.T., Wang, P., Korkaric, M., Capper, R.G., Saunders, N.J., and Langdale, J.A. (2009). GLK transcription factors coordinate expression of the photosynthetic apparatus in Arabidopsis. Plant Cell. 21, 1109-1128. Xu, Y.H., Liu, R., Yan, L., Liu, Z.Q., Jiang, S.C., Shen, Y.Y., Wang, X.F., and Zhang, D.P. (2012). Light-harvesting chlorophyll a/b-binding proteins are required for stomatal response to abscisic acid in Arabidopsis. J. Exp. Bot. 63, 1095-1106. Yabuta, Y., Motoki, T., Yoshimura, K., Takeda, T., Ishikawa, T., and Shigeoka, S. (2002). Thylakoid membrane-bound ascorbate peroxidase is a limiting factor of antioxidative systems under photo-oxidative stress. Plant J. 32, 915-925.

AC C

1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172

33

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Yamamoto, Y., Aminaka, R., Yoshioka, M., Khatoon, M., Komayama, K., Takenaka, D., Yamashita, A., Nijo, N., Inagawa, K., Morita, N., et al. (2008). Quality control of photosystem II: impact of light and heat stresses. Photosyn. Res. 98, 589-608. Yang, X., Wen, X., Gong, H., Lu, Q., Yang, Z., Tang, Y., Liang, Z., and Lu, C. (2007). Genetic engineering of the biosynthesis of glycinebetaine enhances thermotolerance of photosystem II in tobacco plants. Planta. 225, 719-733. Yi, X.P., McChargue, M., Laborde, S., Frankel, L.K., and Bricker, T.M. (2005). The manganesestabilizing protein is required for photosystem II assembly/stability and photoautotrophy in higher plants. J. Biol. Chem. 280, 16170-16174. Yin, S.M., Sun, X.W., and Zhang, L.X. (2008). An Arabidopsis ctpA homologue is involved in the repair of photosystem II under high light. Chin. Sci. Bull. 53, 1021-1026. Yoshioka-Nishimura, M., Nanba, D., Takaki, T., Ohba, C., Tsumura, N., Morita, N., Sakamoto, H., Murata, K., and Yamamoto, Y. (2014). Quality control of photosystem II: direct imaging of the changes in the thylakoid structure and distribution of FtsH proteases in spinach chloroplasts under light stress. Plant Cell Physiol. 55, 1255-1265. Yoshioka-Nishimura, M., and Yamamoto, Y. (2014). Quality control of photosystem II: The molecular basis for the action of FtsH protease and the dynamics of the thylakoid membranes. J. Photochem. Photobiol. B.137, 100-106 . Yoshioka, M., Uchida, S., Mori, H., Komayama, K., Ohira, S., Morita, N., Nakanishi, T., and Yamamoto, Y. (2006). Quality control of photosystem II. Cleavage of reaction center D1 protein in spinach thylakoids by FtsH protease under moderate heat stress. J. Biol. Chem. 281, 21660-21669. Yuan, L., and Xu, D.Q. (2001). Stimulation effect of gibberellic acid short-term treatment on leaf photosynthesis related to the increase in Rubisco content in broad bean and soybean. Photosyn. Res. 68, 39-47. Zhang, X., Wollenweber, B., Jiang, D., Liu, F., and Zhao, J. (2008). Water deficits and heat shock effects on photosynthesis of a transgenic Arabidopsis thaliana constitutively expressing ABP9, a bZIP transcription factor. J. Exp. Bot. 59, 839-848. Zhang, L.R., Xing, D., Wang, J.S., Zeng, L.Z., and Li, Q. (2007a) Light induced delayed fluorescence as an indicator for UV-B radiation environment stress on plants. J Optoelectron Laser 18, 878–881. Zhang, L., Xing, D., Wang, J., and Li, L. (2007b) Rapid and non-invasive detection of plants senescence using a delayed fluorescence technique. Photochem. Photobiol. Sci. 6, 635– 641. Zhang, L., and Xing, D. (2008) Rapid determination of the damage to photosynthesis caused by salt and osmotic stresses using delayed fluorescence of chloroplasts. Photochem. Photobiol. Sci. 7, 352–360. Zhang, S., and Scheller, H.V. (2004). Photoinhibition of photosystem I at chilling temperature and subsequent recovery in Arabidopsis thaliana. Plant Cell Physiol. 45, 1595-1602. Zhang, Z.S., Yang, C., Gao, H.Y., Zhang, L.T., Fan, X.L., and Liu, M.J. (2014). The higher sensitivity of PSI to ROS results in lower chilling–light tolerance of photosystems in young leaves of cucumber. J. Photochem. Photobiol. B Biol. 137, 127-134. Zhao, H.-Z., Zhao, X.-Z., Ma, P.-F., Wang, Y.-X., Hu, W.-W., Li, L.-H., and Zhao, Z.-D. (2011). Effects of salicylic acid on protein kinase activity and chloroplast D1 protein degradation in wheat leaves subjected to heat and high light stress. Acta Ecol. Sin. 31, 259-263.

AC C

1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217

34

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Zhou, Y.H., Yu, J.Q., Huang, L.F., and Nogués, S. (2004). The relationship between CO2 assimilation, photosynthetic electron transport and water–water cycle in chill‐exposed cucumber leaves under low light and subsequent recovery. Plant Cell Environ. 27, 15031514. Zhou, B., Peng, D., Lin, J.Z., Huang, X.Q., Peng, W.S., He, R.Q., Guo, M., Tang, D.Y., Zhao, X.Y., and Liu, X.M. (2011). Heterologous expression of a gibberellin 2-oxidase gene from Arabidopsis thaliana enhanced the photosynthesis capacity in Brassica napus L. J. Plant Biol. 54, 23-32. Zhu, S.Q., Chen, M.W., Ji, B.H., Jiao, D.M., and Liang, J.S. (2011). Roles of xanthophylls and exogenous ABA in protection against NaCl-induced photodamage in rice (Oryza sativa L) and cabbage (Brassica campestris). J. Exp. Bot. 62, 4617-4625. Zienkiewicz, M., Ferenc, A., Wasilewska, W., and Romanowska, E. (2012). High light stimulates Deg1-dependent cleavage of the minor LHCII antenna proteins CP26 and CP29 and the PsbS protein in Arabidopsis thaliana. Planta. 235, 279-288. Zivcak, M., Brestic, M., Kalaji, H.M., and Govindjee. (2014a). Photosynthetic responses of sunand shade-grown barley leaves to high light: is the lower PSII connectivity in shade leaves associated with protection against excess of light? Photosyn. Res. 119, 339-354. Zivcak, M., Kalaji, H.M., Shao, H.B., Olsovska, K., and Brestic, M. (2014b). Photosynthetic proton and electron transport in wheat leaves under prolonged moderate drought stress. J. Photochem. Photobiol. B. 137, 107-115. Zwack, P.J., Robinson, B.R., Risley, M.G., and Rashotte, A.M. (2013). Cytokinin response factor 6 negatively regulates leaf senescence and is induced in response to cytokinin and numerous abiotic stresses. Plant Cell Physiol. 54, 971-981.

AC C

1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240

35

ACCEPTED MANUSCRIPT

1241

FIGURE LEGENDS

1242

Figure 1. Schematic Representation of Photosynthetic Redox Signals and the Detoxification

1244

of Reactive Oxygen Species (ROS). ROS, such as hydrogen peroxide (H2O2), generated in

1245

response to various stress factors, such as high irradiance (indicated by the red lightning bolt),

1246

and redox signals from the electron transport chain (ETC), initiate signaling cascades that

1247

eventually lead to nuclear-gene expression. In photosystem II (PSII), the splitting of water (H2O)

1248

at oxygen-evolving complex (OEC) produces H+ ions that are transferred across the membrane

1249

to ATP synthase (ATP syn), while electrons (e-) are transferred to PSI via the cytochrome b6f

1250

(Cytb6f) protein complex. Plastoquinone (Pq) accepts two protons (H+) from the stromal side

1251

along with two electrons (e-) from PSII and transfers the protons to the luminal side, while

1252

electrons are transferred to PSI via Cytb6f complex and plastocyanin (PC) pool. Electrons

1253

removed from water are transferred to the single-electron carrier ferredoxin (Fd). Ferredoxin

1254

NADP+ reductase (FNR) then transfers an electron from each of the two Fd molecules to a single

1255

molecule of the two-electron carrier nicotineamide adenine dinucleotide phosphate H (NADPH).

1256

Similarly, H+ ions from PSI are transferred to ATP synthase where proton gradients are used for

1257

producing ATP molecules from ADP. Electrons transferred from PSI via Fd are also converted

1258

into superoxide (O2-) in the Mehler peroxidase reaction (MP reaction). The O2- is detoxified by

1259

superoxide dismutase (SOD), a ROS-scavenging enzyme. Another ROS-scavenging enzyme,

1260

ascorbate peroxidase (APX), uses ascorbate (Asc) to reduce H2O2 to H2O. The scavenging of

1261

free radicals by APX increases the levels of monodehydroascorbate (MDA) and

1262

dehydroascorbate (DHA). High levels of MDA and DHA are reduced by MDA reductase

1263

(MDAR), DHA reductase (DHAR) and glutathione reductase (GR). During ascorbate-

1264

glutathione recycling, GR catalyzes the reduction of oxidized glutathione (GSSG) to GSH in

1265

order to maintain a higher ratio of GSH:GSSG in the cytosol. Initial reports indicated that

1266

residual H2O2 from incomplete detoxification can initiate mitogen-activated protein (MAP)

1267

kinases that in turn direct cellular responses such as nuclear-gene expression. Thioredoxins

1268

(TRXs) accept electrons from PSI and utilize them to phosphorylate the light-harvesting complex

1269

and possibly to induce nuclear-gene expression. Similarly, the plastoquinone (Pq) pool induces

1270

the expression of certain protein kinases and transcription factors (TFs) which can, in turn,

1271

trigger transcriptional-signal transduction in the nucleus.

AC C

EP

TE D

M AN U

SC

RI PT

1243

36

ACCEPTED MANUSCRIPT

1272

Figure 2. Schematic Illustrating the Successive Steps of the Photosystem II (PSII) Repair

1274

Cycle after High Light Stress-Induced Photodamage. The photosystem II-light harvesting

1275

supercomplex (PSII-LHCII) has two components: (i) LHCII consists of a protein trimer (denoted

1276

as II) and three less abundant minor antennal chloroplast proteins (CPs), CP24, CP26 and CP29

1277

(denoted as 24, 26 and 29), and (ii) PSII which mainly consists of a D1-D2 (or PsbA-PsbD

1278

where Psb is photosystem b) heterodimer coupled with chlorophyll a containing CP43, CP47,

1279

PsbP (P), PsbQ (Q) and PsbO (O). (A) Photodamage is caused when light-absorbing antennae

1280

receive high-intensity light that damages D1 (denoted by cross-hatching). (B) To counter this

1281

damage, LHC dissipates the excess excitation energy in the form of heat via non-photochemical

1282

quenching. (C) After photodamage has occurred, the protein trimer of LHCII is phosphorylated

1283

by STN7, and the PSII proteins D1, D2 and CP43 are phosphorylated by STN8. Recent studies

1284

indicate, however, that STN7 can also phosphorylate PSII proteins to some extent (Bonardi et al.,

1285

2005; Fristedt et al., 2009; Nath et al., 2013b). The phosphorylation of one of the minor proteins

1286

of LHCII, CP29 (denoted in gray) has been found to be responsible for the disassembly of the

1287

PSII-LHCII supercomplex (Fristedt and Vener, 2011). (D) Phosphorylated proteins of the PSII-

1288

LHCII supercomplex are then dephosphorylated. Thylakoid associated phosphatase 38 (TAP38,

1289

yellow boxes) dephosphorylates LHCII proteins, and Psb core phosphatase (PBCP, brown

1290

polygons) dephosphorylates PSII proteins. (E) Dephosphorylation results in the disassembling of

1291

the PSII-LHCII supercomplex into a PSII repair intermediate and a peripheral antennal complex.

1292

(F) A zinc metalloprotease FtsH then recognizes the N-terminal of the photodamaged D1and

1293

degrades it in a complex processive manner (Adam et al., 2011). (G) ATP-independent Deg5 and

1294

Deg8 endoproteases further assist the processive degradation of photodamaged D1 protein by

1295

FtsH (Kato et al., 2012). (H) The proteolytic degradation of D1 is followed by the synthesis of

1296

nascent copies of D1 protein, facilitated by a plant-specific chloroplastic protease ClpA, which is

1297

involved in processing the C-terminus of D1. Plants possess highly efficient ROS-scavenging

1298

mechanisms, but ROS pose a major obstacle in this step by inhibiting the de novo synthesis of

1299

new D1 proteins. (I) Once nascent copies of D1 protein are synthesized and the co-translational

1300

assembly of D1 is completed, they are inserted into PSII and are post-translationally modified.

1301

Lastly, after the reassembly of its core components, PSII migrates toward the grana in its

1302

functional form.

AC C

EP

TE D

M AN U

SC

RI PT

1273

37

ACCEPTED MANUSCRIPT

Figure 3. A General Representation of Chlorophyll Fluorescence Induction Curve in Plant

1304

Samples. When a dark-adapted plant sample is exposed to light, certain characteristic changes in

1305

chlorophyll fluorescence are observed. These changes are known as fluorescence induction or

1306

fluorescence transient, generally seen as two transient phases that are labeled by using the

1307

observed inflection points: (i) fast chlorophyll fluorescence induction (up to few hundred

1308

milliseconds) or the so-called OJIP phase where, O (origin) is the first measured minimum

1309

fluorescence level, P is the peak and J and I are intermediate inflections (ii) slow chlorophyll

1310

fluorescence induction or the PSMT phase where S stands for steady state, M for a maximum,

1311

and T for a terminal steady state level.

SC

RI PT

1303

1312

Figure 4. Hypothetical Scheme of a Multistep Electron Transport Chain, Based on the

1314

Emission Kinetics of Chlorophyll a. According to this scheme, the conversion of light energy

1315

to chemical energy begins with three consecutive basic steps, (i) ABS, absorption of light energy

1316

(photons) by chlorophyll molecules in the light-harvesting complex of photosystem II (PSII)

1317

(LHCII), (ii) TR, trapping of the excitation energy by the reaction center (RC) and (iii) ET,

1318

electron transport via PSII RCs (PSII-RC). When plant tissues are exposed to a short pulse of

1319

strong light, chlorophyll a (Chl a) absorbs the light energy and uses it in photosynthesis (Chl a*

1320

excited). At the same time, some light is emitted at a lower energy level as a function of time (Ft),

1321

a phenomenon called fluorescence emission. The fluorescence kinetics data is analyzed, and

1322

changes in the photosynthetic apparatus, such as the primary reaction of photochemistry, are

1323

derived through a series of equations, commonly known as the OJIP test (Strasser and Strasser,

1324

1995). Potato plants with altered expression of a PSII-related gene encoding the manganese

1325

stabilizing protein were analyzed with the OJIP test and a hypothesis for the putative changes in

1326

PSII was provided (Gururani et al., 2012; Gururani et al., 2013). The integrity of PSII after the

1327

primary reactions of photochemistry creates a primary redox (Red, reduction; Ox, oxidation)

1328

potential within and around PSII between a primary electron donor pigment P680+ and an electron

1329

acceptor, Quinone (Qa). Re-reduction of the primary electron donor pigment P680+ by internal

1330

electron donors occurs in in vitro systems using Hill reaction agents such as hydroxylamine (HA),

1331

Mn2+ and diphenylcarbazone (DPC). Re-reduction of the primary electron donor pigment (P680+)

1332

by internal electron donors also occurs in vivo. For example, tyrosine, ascorbate (Asc) and

1333

proline (Pro) are generally available as internal electron donors. The first chemical step occurs

AC C

EP

TE D

M AN U

1313

38

ACCEPTED MANUSCRIPT

when an excited donor pigment molecule with a light absorption peak of 680 nm (P680) donates

1335

an electron to pheophytin (Ph), producing oxidized P680 (P680+) and reduced Ph (Ph-) in PSII. The

1336

oxygen-evolving complex (OEC) competes with the internal electron donation, where the

1337

reaction constant for splitting water, kW, is much lower than that of non-water electron donors

1338

(kD>>kW). Under various abiotic stress conditions, PSII is damaged, and the OEC then

1339

presumably favors electron donation by non-water electron donors with a high rate constant, kD.

1340

The total electron transport then increases as OEC activity decreases, due to the easy

1341

accessibility of non-water electrons from molecules such as Asc and Pro. The fraction of

1342

electrons donated by water is lower in stressed samples. Under normal conditions, an intact

1343

manganese cluster (Mn4CaO5) at the OEC thus promotes electron donation from water to PSII-

1344

RC with a low kW and reduces the accessibility of non-water electron donation to the RC of PSII.

M AN U

SC

RI PT

1334

1345

Figure 5. Putative Involvement of Phytohormones and Representative Transcription

1347

Factors (TFs) in the Regulation of the Photosynthetic Machinery in Plants Under Abiotic

1348

Stress Conditions. Red dotted lines indicate crosstalk between two hormones that are inhibitory

1349

to each other. Green dotted lines indicate crosstalk between two hormones that up-regulate each

1350

other. ABA, abscisic acid; ARF2, auxin response factor 2 (Lim et al., 2010); AtNAC2,

1351

Arabidopsis thaliana NAC [NAM (no apical meristem), ATAF (Arabidopsis transcription

1352

activation factor) and CUC (cup-shaped cotyledon)] 2 TF (Ha et al., 2014); BR, brassinosteroid;

1353

BZR1, brassinazole resistant 1 TF (Bai et al., 2012); CRF6, cytokinin response factor 6 (Zwack

1354

et al., 2013); CK, cytokinin; Cytb6f, cytochrome b6f complex; ERF6, ethylene response factor 6

1355

(Dubois et al., 2013); GA, gibberellic acid; GhDREB1, Gossypium hirusitum dehydration-

1356

responsive element binding protein 1 (Shan et al., 2007); JA, jasmonic acid; MYC 2,

1357

myelocytomatosis 2 TF (Dombrecht et al., 2007); PSI, photosystem I; PSII, photosystem II; ROS,

1358

reactive oxygen species; SA, salicylic acid; SL, strigolactone; SlZF2, Solanum lycopersicum

1359

Zinc Finger2 TF (Hichri et al., 2014); OsWRKY45 (Chao et al., 2010), Oryza sativa WRKY45,

1360

a TF belonging to the WRKY TF family that derives its name from the protein sequence motif

1361

WRKYGQK.

AC C

EP

TE D

1346

1362 1363 1364 39

ACCEPTED MANUSCRIPT

RI PT

Chloroplast

ATP ADP

Fd

SOD

MDAR

PS I NADP+

APx

MDA

DHA PC

DHAR

e-

TRs

PS II

2H2O lumen

stroma

Thylakoid

H2O

MDA

DHA

GSSG

GR

GSH

Asc-glutathione recycling

MAP Kinase

EP

OEC

H2O

MDAR

TFs, Kinases H+

AC C

O2 + 4H+

Pq e-

Asc

TE D

Cytb6f

APx

Asc

Asc

NADPH

e-

H2O2

SC

FNR

MP O2reaction

M AN U

O2

Nuclear gene expression

DHAR

ACCEPTED MANUSCRIPT

A

B

II

II

II

24 26 29

P

II

24 26 29

SC

D1 D2 47 43 damaged D1 Q P O

M AN U

Reconstitution of PSII components

I

C-terminal processing

H

D1 processing

D1 43

AC C

D2 47

Co-translational assembly CtpA

24 26 29

G

D2 47 P Q P O

P D1

43

P

STN8 Dephosphorylation TAP38 II

II

D

24 26 29

D1 43

D2 47

PBCP FtsH

EP

Post-translational assembly of PSII proteins ROS

D1 43

TE D

D1 D2 47 43 P Q O

Membrane insertion

II

Phosphorylation

D1 D2 47 43 Q P O

Protein trafficking

P

II

Photodamage

lumen

C

STN7

Heat

RI PT

stroma

D2 47

D1 43

Disassembly

D2 47

Proteolysis

PSII repair intermediate

Deg D1 degradtion

F

Peripheral antenna

E

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Chl a*

SC

Chl a

RI PT

ABS

*

Mn4CaO5

Ox O2 + H+ OEC

-

-

-

kD

Tyrosine

Asc

-

HA

QA -

ET

QA

P680Ph

-

kD DPC Pro Non-water electron donors

Mn2+

Ph-

P680+

EP

Red kW

AC C

H2O

TE D

Ft

M AN U

TR

LHCII

PSII-RC

ACCEPTED MANUSCRIPT

PS II

RI PT

Stress PS I

Cytb6f

SC

Stress signal perception

GA

JA

SA

ABA

TE D

Hormonal regulation

M AN U

ROS

AC C

Structural and functional adaptations

ARF2

BR

SL

BZR1

AtNAC2

CK

CRF6

EP

Transcriptional GhDREB1 OsWRKY45 MYC2 SlZF2 control

Auxin

Regulation of photosynthetic machinery

Ethylene

ERF6

Regulation of Photosynthesis during Abiotic Stress-Induced Photoinhibition.

Plants as sessile organisms are continuously exposed to abiotic stress conditions that impose numerous detrimental effects and cause tremendous loss o...
2MB Sizes 2 Downloads 24 Views