mrf family genes are involved in protein translation ... · report that. 26. arabidopsis. thaliana....

80
1 RESEARCH ARTICLE 1 2 MRF Family Genes Are Involved in Translation Control, Especially Under 3 Energy-Deficient Conditions, and Their Expression and Functions Are 4 Modulated by the TOR Signaling Pathway 5 6 Du-Hwa Lee, Seung Jun Park, Chang Sook Ahn, and Hyun-Sook Pai* 7 Department of Systems Biology, Yonsei University, Seoul 120-749, Korea 8 *Corresponding author: E-mail: [email protected]; Fax: 82-2-312-5657 9 10 Short title: MRF family genes in plant translation control 11 12 One-sentence summary: MRF family genes encode translation regulatory factors, with functions that are 13 important under energy-deficient conditions, and the TOR signaling pathway modulates MRF expression 14 and functions. 15 16 The author responsible for distribution of materials integral to the findings presented in this article and in 17 accordance with the policy described in the Instructions for Authors (www.plantcell.org) is: Hyun-Sook 18 Pai ([email protected]). 19 20 ABSTRACT 21 Dynamic control of protein translation in response to the environment is essential for the survival of plant 22 cells. Target of rapamycin (TOR) coordinates protein synthesis with cellular energy/nutrient availability 23 through transcriptional modulation and phosphorylation of the translation machinery. However, 24 mechanisms of TOR-mediated translation control are poorly understood in plants. Here, we report that 25 Arabidopsis thaliana MRF (MA3 DOMAIN-CONTAINING TRANSLATION REGULATORY 26 FACTOR) family genes encode translation regulatory factors under TOR control, and their functions are 27 particularly important in energy-deficient conditions. Four MRF family genes (MRF1MRF4) are 28 transcriptionally induced by dark and starvation (DS). Silencing of multiple MRFs increases susceptibility 29 to DS and treatment with a TOR inhibitor, while MRF1 overexpression decreases susceptibility. MRF 30 proteins interact with eIF4A and co-fractionate with ribosomes. MRF silencing decreases translation 31 activity, while MRF1 overexpression increases it, accompanied by altered ribosome patterns, particularly 32 in DS. Furthermore, MRF deficiency in DS causes altered distribution of mRNAs in sucrose gradient 33 fractions, and accelerates rRNA degradation. MRF1 is phosphorylated in vivo, and phosphorylated by S6 34 kinases in vitro. MRF expression, and MRF1 ribosome association and phosphorylation are modulated by 35 cellular energy status and TOR activity. We discuss possible mechanisms of the function of MRF family 36 proteins under normal and energy-deficient conditions and their functional link with the TOR pathway. 37 38 INTRODUCTION 39 Translation, a fundamental cellular process that is highly conserved in eukaryotes, occurs in four 40 stages: initiation, elongation, termination, and ribosome recycling (Sonenberg and Hinnebusch, 41 Plant Cell Advance Publication. Published on October 30, 2017, doi:10.1105/tpc.17.00563 ©2017 American Society of Plant Biologists. All Rights Reserved

Upload: vanduong

Post on 31-Aug-2018

212 views

Category:

Documents


0 download

TRANSCRIPT

1

RESEARCH ARTICLE 1

2

MRF Family Genes Are Involved in Translation Control, Especially Under 3

Energy-Deficient Conditions, and Their Expression and Functions Are 4

Modulated by the TOR Signaling Pathway 5

6

Du-Hwa Lee, Seung Jun Park, Chang Sook Ahn, and Hyun-Sook Pai*7

Department of Systems Biology, Yonsei University, Seoul 120-749, Korea 8

*Corresponding author: E-mail: [email protected]; Fax: 82-2-312-56579

10

Short title: MRF family genes in plant translation control 11

12

One-sentence summary: MRF family genes encode translation regulatory factors, with functions that are 13

important under energy-deficient conditions, and the TOR signaling pathway modulates MRF expression 14

and functions. 15

16

The author responsible for distribution of materials integral to the findings presented in this article and in 17

accordance with the policy described in the Instructions for Authors (www.plantcell.org) is: Hyun-Sook 18

Pai ([email protected]). 19

20

ABSTRACT 21

Dynamic control of protein translation in response to the environment is essential for the survival of plant 22

cells. Target of rapamycin (TOR) coordinates protein synthesis with cellular energy/nutrient availability 23

through transcriptional modulation and phosphorylation of the translation machinery. However, 24

mechanisms of TOR-mediated translation control are poorly understood in plants. Here, we report that 25

Arabidopsis thaliana MRF (MA3 DOMAIN-CONTAINING TRANSLATION REGULATORY 26

FACTOR) family genes encode translation regulatory factors under TOR control, and their functions are 27

particularly important in energy-deficient conditions. Four MRF family genes (MRF1–MRF4) are 28

transcriptionally induced by dark and starvation (DS). Silencing of multiple MRFs increases susceptibility 29

to DS and treatment with a TOR inhibitor, while MRF1 overexpression decreases susceptibility. MRF 30

proteins interact with eIF4A and co-fractionate with ribosomes. MRF silencing decreases translation 31

activity, while MRF1 overexpression increases it, accompanied by altered ribosome patterns, particularly 32

in DS. Furthermore, MRF deficiency in DS causes altered distribution of mRNAs in sucrose gradient 33

fractions, and accelerates rRNA degradation. MRF1 is phosphorylated in vivo, and phosphorylated by S6 34

kinases in vitro. MRF expression, and MRF1 ribosome association and phosphorylation are modulated by 35

cellular energy status and TOR activity. We discuss possible mechanisms of the function of MRF family 36

proteins under normal and energy-deficient conditions and their functional link with the TOR pathway. 37

38

INTRODUCTION 39

Translation, a fundamental cellular process that is highly conserved in eukaryotes, occurs in four 40

stages: initiation, elongation, termination, and ribosome recycling (Sonenberg and Hinnebusch, 41

Plant Cell Advance Publication. Published on October 30, 2017, doi:10.1105/tpc.17.00563

©2017 American Society of Plant Biologists. All Rights Reserved

2

2009). Initiation is the rate-limiting step, and is controlled by eukaryotic translation initiation 42

factors (eIFs) and many other accessory proteins (Holcik and Sonenberg, 2005). During the 43

initiation step, the eIF2-GTP-Met-tRNAiMet

ternary complex binds to the eukaryotic small 44

ribosomal subunit (40S) to form the 43S pre-initiation complex (PIC). The 43S PIC attaches to 45

the 5′-end of mRNA via the eIF4F complex composed of eIF4E (5’ cap-binding protein) and 46

eIF4G (scaffold). The 5’ cap-bound eIF4F complex recruits eIF4A (DEAD-box RNA helicase), 47

eIF4B (eIF4A enhancer) and PABPs [poly (A)-binding proteins] (Muench et al., 2012; Browning 48

and Bailey-Serres, 2015; Merchante et al., 2017). A second form of eIF4F, eIFiso4F, exists only 49

in plants and is composed of eIFiso4G and eIFiso4E; the eIFiso4F form shows differential 50

translation-promoting activities on mRNAs (Allen et al., 1992; Patrick and Browning, 2012; 51

Browning and Bailey-Serres, 2015). The 43S PIC including eIF4F or eIFiso4F scans along the 52

5′-untranslated region of the mRNA to select the AUG codon, at which point it is joined with the 53

60S subunit via eIF5B to form a functional 80S ribosome (Jackson et al., 2010; Browning and 54

Bailey-Serres, 2015). 55

Control of global translation activity is critical for cellular adaptation to fluctuating growth 56

conditions and environmental stimuli (Sonenberg and Hinnebusch, 2009; Sengupta et al., 2010). 57

Translation initiation that determines the overall rate of translation is the primary target for 58

regulation under stress conditions; two key points of the regulation are ternary complex 59

formation and 5’-cap recognition (Jackson et al., 2010). Many stress conditions trigger 60

phosphorylation of eIF2α by eIF2α kinases, inhibiting ternary complex formation in mammals; 61

phosphorylation of eIF2α inhibits the eIF2B-catalyzed exchange of GDP for GTP, required for 62

regeneration of active eIF2-GTP (Jackson et al., 2010; Silvera et al., 2010). Arabidopsis and rice 63

(Oryza sativa) have a single eIF2α kinase (GCN2) that phosphorylates eIF2α in response to 64

nutrient starvation, UV, cold shock, and wounding (Browning, 2004; Lageix et al., 2008). A 65

recent study revealed that Arabidopsis GCN2 requires an interaction partner GCN1 for eIF2α 66

phosphorylation (Wang et al., 2017). Since plant eIF2 has much less affinity to GDP than its 67

mammalian counterparts, the eIF2B activity as a guanine nucleotide exchange factor may not be 68

absolutely necessary in plants (Shaikhin et al., 1992; Browning and Bailey-Serres, 2015). Further 69

studies will be required to determine the roles of eIF2α phosphorylation and eIF2B in global 70

translation control in plants. The 5’-cap recognition step is the major regulation point of 71

3

translation initiation in mammals and yeast, but how the step is regulated according to cell 72

environments is largely unknown in plants. Arabidopsis eIF4A is phosphorylated by cyclin-73

dependent kinase A (CDKA) on Thr-164 residue, and a phosphomimetic mutant (T164E) of 74

eIF4A-1 lacks ATPase and helicase activities and cannot support protein translation, suggesting 75

that CDKA may repress translation during cell proliferation through eIF4A phosphorylation 76

(Hutchins et al., 2004). The eIF4E-binding protein (4E-BP) is the main regulator of the 5’-cap 77

recognition step in mammals. Stress signaling activates 4E-BP, which suppresses translation 78

initiation by interfering with the interaction between eIF4E and eIF4G (Sonenberg and 79

Hinnebusch, 2009; Jackson et al., 2010). Plants lack 4E-BP, and it remains to be discovered 80

whether plants have evolved an equivalent system for global regulation of translation (Lageix et 81

al., 2008; Xiong et al., 2013). 82

The target of rapamycin (TOR) signaling pathway plays a vital role in sensing and 83

responding to external signals to control cell growth and metabolism in all eukaryotes. Such 84

signals include nutrient availability, energy status, growth factors, and environmental conditions 85

(Dobrenel et al., 2016a). The TOR pathway regulates myriad biological processes, including 86

transcription, translation, ribosome biogenesis, protein trafficking, and autophagy (Wullschleger 87

et al., 2006). The mammalian TOR complex 1 (mTORC1) regulates the activity of the translation 88

initiation machinery, including eIF4G, eIF4B, and 4E-BP, through direct or indirect 89

phosphorylation (Ma and Blenis, 2009). In plants, inactivation of TOR causes a significant 90

decrease in polysome accumulation (Deprost et al., 2007; Ahn et al., 2011). Schepetilnikov et al. 91

(2013) suggested that TOR and S6K1 stimulate translation re-initiation of upstream open reading 92

frame (uORF)-containing mRNAs via phosphorylation of eIF3h in Arabidopsis. Considering that 93

uORFs are found in more than 30% of Arabidopsis full-length mRNAs, the aforementioned 94

mechanism of translation control may play an important role in plant development (Zhou et al., 95

2010). As with yeast and mammalian S6Ks, Arabidopsis S6Ks phosphorylate the 40S ribosomal 96

protein S6 (RPS6) (Turck et al., 2004; Mahfouz et al., 2006; Dobrenel et al., 2016b). Additional 97

substrates of mammalian S6Ks have been identified, including eIF4B, elongation factor 2 kinase, 98

and cap-binding protein 80 (Magnuson et al., 2012). However, downstream targets of plant S6Ks 99

related to plant mRNA translation are largely unknown, apart from RPS6 and eIF3h. 100

Throughout their lifespan, plants are exposed to diverse stresses, which disrupt cellular 101

4

energy homeostasis. Protein synthesis demands a large amount of energy, and is tightly regulated 102

according to the cellular energy status in yeast and mammals (Ma and Blenis, 2009). Similarly, a 103

tight correlation between translation activity and cellular sugar levels has been observed in plants 104

(Pal et al., 2013; Lastdrager et al., 2014). The key players of plant energy signaling are sucrose 105

non-fermenting 1-related protein kinase 1 (SnRK1) and TOR, which act in an antagonistic 106

crosstalk in the plant’s response to energy deprivation (Lastdrager et al., 2014; Mair et al., 2015; 107

Nukarinen et al., 2016). TOR silencing mimics energy starvation conditions, and activates 108

catabolic processes and autophagy while repressing global translation (Deprost et al., 2007; 109

Moreau et al., 2012; Ren et al., 2012; Caldana et al., 2013; Xiong et al., 2013). However, the 110

detailed mechanisms of TOR’s control of stress responses, particularly regarding global mRNA 111

translation, are largely unclear in plants. 112

Programmed cell death 4 (PDCD4) is a tumor suppressor that has been implicated in the 113

development of multiple cancers (Lankat-Buttgereit and Goke, 2009). Human PDCD4 (hPDCD4) 114

binds to eIF4A through its two MA3 domains, inhibiting the eIF4A helicase activity and the 115

eIF4A-eIF4G interaction, leading to a decrease in translation initiation rates (Loh et al., 2009). 116

Homologs of hPDCD4 are found in animals, plants, and lower eukaryotes, but not in yeast. Only 117

the homologs of higher plants contain four MA3 domains in tandem, instead of two in the other 118

systems (Cheng et al., 2013). The Arabidopsis thaliana genome contains four genes encoding 119

PDCD4 homologs, and one of them was recently reported to interact with the ethylene signaling 120

protein EIN2, hence it was designated ECIP1 (EIN2 C-TERMINUS-INTERACTING PROTEIN 121

1; AT4G24800). Loss-of-function mutations in ECIP1 have been shown to result in ethylene 122

hypersensitivity (Lei et al., 2011). Apart from these findings, we lack evidence of the cellular 123

functions of these homologs. Here, we investigated protein characteristics and in planta 124

functions of four PDCD4 homologs in Arabidopsis. Our results suggested that these proteins 125

positively regulate protein translation in plants, particularly under dark and starvation conditions; 126

we thus designated them MA3-containing translation regulatory factor (MRF) 1 to 4. We also 127

found that the transcription of the MRF genes, and ribosome association and phosphorylation of 128

MRF1 are modulated by TOR activity, suggesting a functional link with the TOR signaling 129

pathway. 130

131

5

RESULTS 132

MRF Family Proteins Have Four MA3 Domains 133

The Arabidopsis MRF gene family consists of four genes, MRF1, MRF2, MRF3 (ECIP1), and 134

MRF4; each encodes a protein with four tandem MA3 domains (Figure 1A; Supplemental Table 135

1; Cheng et al. (2013). The MA3 domain functions as a protein-protein interaction module 136

(Yang et al., 2003; Yang et al., 2004). Notably, the N-terminal region of MRF4 is smaller than 137

that of the other MRFs. Arabidopsis MRF proteins exhibit less than 30% protein sequence 138

identity to human PDCD4. Arabidopsis MRF gene family is divided into two clades: one 139

including MRF1, MRF3, and MRF4, and the other including MRF2 (Cheng et al., 2013). 140

141

MRF Family Genes Are Transcriptionally Induced by Dark and Starvation 142

We examined expression patterns of each MRF gene using reverse transcription quantitative PCR 143

(RT-qPCR) with gene-specific primers (Supplemental Table 2). All MRF genes were expressed in 144

all major plant organs, but exhibited different expression profiles. MRF1 and MRF3 transcripts 145

were more abundant in vegetative tissues, such as rosette leaves, cauline leaves, and stems, than 146

in reproductive tissues, such as flower buds and flowers (Figure 1B). MRF4 transcript levels 147

were highest in rosette leaves and flower buds, although the overall mRNA level was much 148

lower for MRF4 than for the other MRF genes. MRF2, which belongs to a different subgroup 149

within the MRF gene family, displayed the highest transcript levels in the reproductive organs, in 150

contrast to the other MRF genes (Figure 1B). According to the Genevestigator database 151

(https://genevestigator.com), MRF1 and MRF3 transcript levels increased 2-fold in the dark, and 152

decreased 4-fold upon glucose feeding after starvation. Since MRF4 was not included in the 153

Genechip Affymetrix ATH1 genome array, we examined expression patterns of all MRF genes in 154

response to darkness and starvation using Arabidopsis seedlings grown in liquid culture for 12 155

days (Figure 1C, D). RT-qPCR revealed that transcript levels of all MRF genes dramatically 156

increased after 30 min of darkness. During the next 24 hours, MRF1 transcripts stayed at the high 157

level and MRF3 transcript levels increased further, while MRF2 and MRF4 transcript levels 158

decreased after reaching a peak (Figure 1C). 159

We next examined MRF gene expression upon glucose feeding following a period of 160

starvation (Figure 1D). All MRF genes were highly induced after 24 h starvation (no glucose in 161

6

medium). Subsequent glucose feeding in the following 4 h did not affect MRF2 expression, but 162

significantly decreased the transcript levels of MRF1, MRF3, and MRF4, at all glucose 163

concentrations (10-180 mM). Mannitol feeding (10-180 mM) for 4 h after starvation had little 164

effect on the transcript levels, suggesting that osmotic changes were not responsible for the 165

observed transcriptional repression by glucose (Supplemental Figure 1). These results suggest 166

that MRF family genes are transcriptionally induced by darkness and starvation, i.e. low-energy 167

conditions. Finally, transcript levels were compared between the four MRF genes after 24 h 168

starvation and subsequent 4 h glucose feeding (10-180 mM; Supplemental Figure 2). Although 169

the primer efficiency may differ between the MRF genes, the results clearly suggest that MRF1 is 170

a dominantly expressed gene among MRF family members after 24-h starvation. 171

172

MRF Family Proteins Predominantly Localize in the Cytosol 173

To investigate the subcellular localization of MRF family proteins, we generated green 174

fluorescent protein (GFP) fusion constructs of MRF1 to MRF4 under the control of the CaMV 175

35S promoter. The constructs were transiently expressed in Nicotiana benthamiana leaves via 176

agroinfiltration. Confocal laser scanning microscopy using protoplasts isolated from the 177

infiltrated leaves showed that MRF1 to MRF4 are mainly localized in the cytosol, but also found 178

near the nucleus (Figure 2A). The green fluorescence signals partially overlapped with the red 179

fluorescence of the nucleus marker (histone H2B:mRFP), but did not overlap with the 180

chloroplast autofluorescence signal. To verify the localization of the MRFs, we prepared total 181

(T), nuclear (N), and cytosolic (C) protein fractions from the infiltrated N. benthamiana leaves, 182

and performed immunoblotting using anti-GFP antibody and anti-histone H3 antibody as nuclear 183

markers (Figure 2B). All of the MRF:GFP proteins were predominantly associated with the 184

cytosolic fraction. Finally, we generated transgenic Arabidopsis lines that expressed GFP-fusion 185

proteins of each MRF under the control of the CaMV35S promoter. Confocal microscopy of the 186

leaf epidermal cells confirmed the cytosolic localization of all MRFs (Figure 2C). GFP signals 187

were rarely observed near the nucleus. 188

189

Altered MRF Expression Affects Flowering Time under Long-Day Conditions 190

7

Three T-DNA insertion mutant lines were available for MRF3 (Supplemental Figure 3A); mrf3-3 191

and mrf3-2 alleles had been previously identified and designated ecip1-1 and ecip1-2, 192

respectively (Lei et al., 2011). RT-PCR analyses showed that mrf3-2 (ecip1-1) is a null allele, 193

while mrf3-3 (ecip1-2) produces a truncated MRF3 mRNA (Supplemental Figure 3D, E; Lei et 194

al., 2011). mrf3-1 has a T-DNA insertion in the promoter region, resulting in a slight decrease in 195

MRF3 mRNA. The single T-DNA insertion mutant of MRF4, mrf4-1, has no full-length 196

transcripts based on RT-PCR (Supplemental Figure 3A, F, G). There were no T-DNA insertion 197

mutants available for MRF1. Homozygote mrf2 seeds could not be obtained from two T-DNA 198

insertion alleles, both carrying the insertion in the 3rd

exon. Since no usable mutant lines were 199

available, we generated artificial microRNA (amiRNA) lines for MRF1 and MRF2 under the 200

control of the CaMV 35S promoter in Arabidopsis (Col-0 ecotype). Three unique target sites 201

were designed for both MRF1 and MRF2 using an online designer tool (Schwab et al., 2006), 202

and more than 10 independent amiRNA transgenic lines were generated per target site (Figure 203

3A). RT-qPCR showed that significant gene silencing occurred in target a and c lines of MRF1, 204

and target d and f lines of MRF2 amiRNA lines (Supplemental Figure 3B, C). However, MRF1 205

target b and MRF2 target e sites were not effective in inducing target gene silencing. 206

To circumvent possible redundancy of MRF functions, we also generated multiple-target 207

amiRNA (Ami-m) lines against MRF1, MRF3, and MRF4, which belong to the same subgroup 208

and show similar expression patterns (Figure 3A; Cheng et al., 2013). The Ami-m target 209

sequence was designed to silence all three genes. In addition, we generated constitutive 210

overexpression (OE) lines of Flag tag-fused MRF1 (Flag:MRF1) under the control of a 211

CaMV35S promoter (Figure 3A), since MRF1 is most highly expressed among the MRF family 212

genes, particularly under starvation conditions (Supplemental Figure 2). Based on RT-qPCR, 213

MRF1 and MRF3 transcript levels in all three independent Ami-m lines (#3, #10, and #14) 214

decreased to 17-34% of the wild-type (WT) levels (Figure 3B). MRF4 transcript level was 215

reduced to at most ~60% of the WT levels under dark and starvation conditions, possibly due to 216

low target efficiency (Supplemental Figure 4A). Selected MRF1 OE lines, #1 and #2, produced 217

~34- and ~15.5-fold higher levels of MRF1 transcripts than the WT, respectively (Figure 3C). 218

We observed overall growth and development of amiRNA lines of MRF1 and MRF2, and 219

T-DNA insertion mutants of MRF3 and MRF4 (Supplemental Figure 4B, C). Under normal long-220

8

day conditions, these plants did not show any visible developmental abnormalities, except for 221

slightly early flowering in mrf3-2 and mrf3-3 T-DNA mutants, and MRF1 amiRNA lines, which 222

was measured by counting numbers of rosette and cauline leaves, consistent with previous 223

reports on ecip1 (mrf3) (Lei et al., 2011). MRF2 amiRNA lines and mrf4 T-DNA mutant did not 224

show early flowering phenotype (Supplemental Figure 4B, C). Both the Ami-m and MRF1 OE 225

lines exhibited no gross abnormalities during vegetative growth. However, the Ami-m lines 226

clearly showed early flowering phenotype under long-day conditions, while in MRF1 OE #1 and 227

#2 lines flowering was significantly delayed (Figure 3D, E). Thus, altered MRF gene expression 228

affects the transition from vegetative to reproductive growth. 229

230

MRF Modulates Plant Resistance to Darkness and Starvation Stress 231

Since expression of all MRF genes was induced by darkness and starvation, we investigated 232

plant phenotypes under these conditions and after subsequent re-illumination and glucose 233

feeding, using a liquid culture system (Figure 4). WT, MRF Ami-m (#3, #10, and #14), and 234

MRF1 OE (#1 and #2) seedlings at 12 days after germination (DAG), which were grown under 235

light and glucose (LG) conditions (control), were exposed to darkness and starvation (DS; no 236

sugar in the medium) for 5 days, and then transferred to LG conditions (ReLG) for 5 days to 237

observe re-greening. The seedlings had shown no differences in growth or total chlorophyll 238

content before treatment (Figure 4A, top; Supplemental Figure 5A). All seedlings started to lose 239

chlorophyll after 2 days of DS, progressing to complete chlorosis within 5 days (Figure 4A). 240

ReLG treatment after DS induced regeneration of green leaves at the shoot apex. MRF1 OE 241

seedlings re-greened somewhat earlier than WT, while Ami-m seedlings showed markedly 242

delayed re-greening after 5 days of ReLG treatment (Figure 4A). Correspondingly, total 243

chlorophyll contents in Ami-m and OE lines were lower and higher, respectively, than those in 244

the WT after 3 days of ReLG treatment (Supplemental Figure 5A, B). MRF1 amiRNA line (ami 245

c-#12) exhibited slightly delayed re-greening, but no visible DS-induced senescence and re-246

greening phenotypes were found in MRF2 amiRNA lines and T-DNA insertion mutants mrf3 and 247

mrf4 (Supplemental Figure 6). 248

Using RT-qPCR, we determined time-course expression profiles of CHLOROPHYLL A/B-249

BINDING PROTEIN 2 (CAB2), SENESCENCE 4 (SEN4), DARK-INDUCED 6 (DIN6), DIN10, 250

9

and PROLINE DEHYDROGENASE (PRODH) mRNAs during the DS and ReLG treatments 251

(Figure 4B, D; Supplemental Figure 7A, C, E). In addition, we plotted relative transcript levels 252

of the genes at specific time points when the samples showed the biggest differences (Figure 4C, 253

E; Supplemental Figure 7B, D, F). CAB2 is a representative marker for cellular photosynthetic 254

activity and SEN4 (encoding xyloglucan endotransglucosylase/hydrolase) is a marker for natural 255

and dark-induced senescence. Expression of DIN6 (encoding asparagine synthethase), DIN10 256

(encoding raffinose synthase), and ProDH are up-regulated in darkness. The MRF1 OE and Ami-257

m lines maintained higher and lower CAB2 mRNA levels than WT, respectively, after 3 days of 258

ReLG treatment (Figure 4B, C). In contrast, mRNA levels of SEN4, DIN6, DIN10, and ProDH 259

were higher than WT in the Ami-m lines and lower in the OE lines after 5 days of DS (Figure 4D, 260

E; Supplemental Figure 7). Detached leaf senescence assays using the 6-7th

leaf from soil-grown 261

plants resulted in similar findings: measured after 4 days in the dark, Ami-m leaves were more 262

susceptible to dark-induced senescence than WT with lower total chlorophyll levels, while MRF1 263

OE leaves exhibited delayed senescence with higher chlorophyll contents (Supplemental Figure 264

5C, D). Collectively, these results suggest that MRF deficiency and MRF1 overexpression lead 265

to decreased and increased resistance to DS-induced senescence, respectively. 266

267

MRF Proteins Interact with eIF4A in Vivo 268

MRF proteins have four tandem MA3 domains (Figure 1A). eIF4G and PDCD4 have one and 269

two MA3 domains, respectively, and these domains function in eIF4A binding (Yang et al., 2003; 270

Yang et al., 2004; Loh et al., 2009). There is moderate sequence similarity between the MA3 271

domains of human eIF4G and PDCD4, and between the MA3 domains of Arabidopsis eIF4G 272

and MRFs (Supplemental Figure 8). To determine whether MRF proteins interact with eIF4A in 273

planta, we performed bimolecular fluorescence complementation (BiFC). MRFs and eIF4A-1 of 274

Arabidopsis were expressed in combination as yellow fluorescent protein (YFP)N- and YFP

C-275

fusion proteins in N. benthamiana leaves by agroinfiltration. Confocal microscopy revealed 276

yellow fluorescence in the cytosol of the leaf cells in every combination, suggesting interaction 277

between all members of the MRF family and eIF4A-1, occurring mostly in the cytosol (Figure 278

5A; Supplemental Figure 9). However, BiFC of MRFs with eIF4E-1, the 5′ cap-binding protein, 279

10

did not result in yellow fluorescence in any combination, suggesting a lack of interaction, despite 280

the protein expression in leaf cells shown by immunoblotting with polyclonal anti-GFP antibody 281

(Figure 5A; Supplemental Figures 9, 10). As the control, no interaction was observed between 282

eIF4A-1 and Gle1 (mRNA export factor localized in both nuclear envelope and cytosol; 283

Supplemental Table 1; Lee et al. (2015)), or between the YFPN and YFP

C vector control. 284

Next, we performed co-immunoprecipitation assays (Figure 5B). Flag-tagged MRFs and 285

Myc-tagged eIF4A-1 were expressed in N. benthamiana leaves using agroinfiltration. eIF4A-286

1:Myc was immunoprecipitated from leaf extracts using anti-Myc antibody-conjugated resin, 287

followed by immunoblotting with anti-Flag antibody to detect co-immunoprecipitated Flag:MRF 288

proteins. All MRF proteins were detected in the immunoprecipitates, supporting the interactions 289

between MRF family proteins and eIF4A-1 in vivo (Figure 5B). Furthermore, after extensive 290

washing, Flag:MRF1 co-immunoprecipitated with eIF4A-1, but not with eIF4A-2 or eIF4A-3, 291

suggesting that MRF1 interacts most strongly with eIF4A-1 among eIF4A family proteins 292

(Supplemental Figure 11). 293

Finally, we performed yeast two-hybrid assays to measure the relative binding affinity of 294

each MRF for eIF4A (Figure 5C). GAL4 activating domain (AD)-fused eIF4A-1 and GAL4 295

promoter binding domain (BD)-fused MRFs were expressed together in yeast, and α-296

galactosidase assays were performed and the activity was statistically analyzed with biological 297

replications. Immunoblotting with anti-Myc and anti-HA antibodies showed that the expression 298

levels of BD- and AD-fusion proteins were similar in these yeast cells (Supplemental Figure 12). 299

The α-galactosidase activity of MRF1 and MRF3 appeared to be somewhat higher than that of 300

MRF2 and MRF4, suggesting higher binding affinity for eIF4A-1. 301

302

MRF1 Co-Sediments with Monosomes 303

Since MRF proteins interact with eIF4A, it is possible that they are involved in protein 304

translation control in plants. We first investigated the co-fractionation of MRFs with ribosomes 305

(Figure 6A). GFP-fused MRFs, and Myc-tagged eIF4A-1 and eIF4E-1 were expressed in leaves 306

of soil-grown N. benthamiana plants via agroinfiltration. The leaf extracts were then fractionated 307

on a 15-50% sucrose density gradient. After ultracentrifugation, fractions were collected, and 308

11

subjected to immunoblot analyses with anti-GFP and anti-Myc antibodies. As a control for 309

fractionation, immunoblotting was performed with an antibody against the 60S ribosomal protein 310

L10a (RPL10a), which is associated with 60S large subunits, 80S monosomes, and polysomes. 311

All MRF proteins were distributed up to the fractions in which 60S/monosomes occurred, similar 312

to the pattern of eIF4E-1, suggesting a possibility of ribosomal association of MRFs (Figure 6A). 313

eIF4A-1 was broadly distributed in the fractions including the polysomal fractions. 314

Next, we tested whether RNA is involved in MRF co-sedimentation with ribosomes 315

(Figure 6B). Cell extracts were prepared from Arabidopsis MRF1 OE seedlings grown in liquid 316

culture under LG conditions, and briefly treated with RNase A or RNase-free water (control) 317

before sucrose gradient sedimentation (15-50%). After ultracentrifugation, fractions were 318

collected for immunoblotting with anti-Flag and anti-L10a antibodies while performing 319

polysome analyses. Without RNase treatment, Flag:MRF1 was co-fractionated with ribosome 320

subunits and light polysomes (Figure 6B, left). The observed differences in MRF distribution 321

patterns between N. benthamiana and Arabidopsis samples may be caused by their differences in 322

plant developmental stages and cellular energy/nutrient status. A brief RNase treatment resulted 323

in a large increase in 60S/80S amounts and disruption of polysomes by digesting ribosome-324

associated mRNAs (Figure 6B, right). Accordingly, RPL10a mostly accumulated in the 60S/80S 325

region based on immunoblotting. Interestingly, the RNase treatment shifted MRF1 to less dense 326

fractions, demonstrating that MRF1 co-sedimentation with ribosomes partially depends on RNA. 327

To explore further the relationship between MRF function and RNA, we performed 328

complementation assays using an E. coli mutant lacking cold-shock proteins (Figure 6C). RNA 329

chaperones, such as bacterial cold-shock proteins (CSPs), are required for adaptation to low 330

temperatures, because cellular RNAs tend to form stable nonfunctional secondary structures 331

under low temperature conditions (Kang et al., 2013). The E. coli BX04 strain is a quadruple 332

mutant of cold-shock proteins (ΔcspA, ΔcspB, ΔcspE, and ΔcspG), and cannot grow at low 333

temperatures (Xia et al., 2001). Plasmids containing MRFs, E. coli CspA encoding a cold-shock 334

protein, and Arabidopsis LOW EXPRESSION OF OSMOTICALLY RESPONSIVE GENES 4 335

(LOS4) (Gong et al., 2005; Lee et al., 2015), which encodes an RNA helicase/mRNA export 336

factor, were introduced into the BX04 strain, followed by IPTG (isopropyl β-D-1-337

12

thiogalactopyranoside) treatment to induce protein expression. All of the transformants in the 338

BX04 background, including the vector control, grew well at 37 °C. However, at 18 °C, growth 339

of the vector control was defective, while CspA strongly complemented the cold-sensitive 340

growth defect of the BX04 strain. All MRF proteins partially rescued the cold-sensitive 341

phenotype of the mutant strain, similar to the activity of the RNA helicase LOS4 (Figure 6C). 342

Collectively, these results suggest that RNA interaction may play a role in MRF functions in 343

plants. 344

345

MRF Deficiency Causes Reduced Cellular Translation Activity under DS Conditions 346

To investigate the role of MRFs in protein translation, we first determined cellular translation 347

activity in WT, MRF Ami-m (#3, #10, and #14), and MRF1 OE (#1 and #2) seedlings using 35

S-348

methionine labeling (Figure 7A, B; Supplemental Figure 13). Seedlings were grown for 12 days 349

under long-day conditions in liquid culture. For pre-treatment, they were then transferred to fresh 350

1/2 MS medium with 30 mM glucose for 30-minute incubation in the light (LG), or to 1/2 MS 351

medium without glucose for 30-minute incubation in the dark (DS). After 30 min, 35

S-352

methionine was added to the medium, and the seedlings were incubated further for 2.5 h under 353

the same conditions. After labeling, protein extracts from the seedlings were subjected to SDS-354

PAGE and analyzed by a phosphorimager. Radioactive bands indicated newly synthesized 355

proteins after incorporation of 35

S-methionine. Based on the intensity of radioactivity, nascent 356

protein synthesis was lower in all of the lines under DS conditions than under LG, as previously 357

reported (Juntawong and Bailey-Serres, 2012). Comparing radioactive band intensity between 358

the lines under LG conditions, we found a slightly decreased signal from Ami-m (#3) lines, but 359

no other meaningful differences (Supplemental Figure 13B). However, under DS conditions, 360

nascent protein synthesis in all of the Ami-m lines decreased to 60~75% of the WT levels, 361

suggesting a significant reduction in global translation (Figure 7B). MRF1 OE lines showed 362

slightly higher cellular translation activity than that in the WT. These results suggest that MRFs 363

are required for efficient protein translation under low-energy conditions. 364

365

Polyribosome Profiles in MRF-Deficient and MRF1-OE Plants under Normal and Energy-366

13

Deficient Conditions 367

Next, we performed polysome analyses in WT, MRF Ami-m (#3), and MRF1 OE (#1) lines using 368

sucrose density gradient sedimentation (Figure 7C). The seedlings were grown for 12 days under 369

long-day conditions in liquid culture, and then transferred either to fresh 1/2 MS medium with 30 370

mM glucose, followed by a 2-hour incubation in the light (LG); or to 1/2 MS medium without 371

glucose, followed by a 1-day incubation in the dark (DS). The seedlings were treated with 372

cycloheximide (50 μg/ml) for 5 min before harvest to preserve the polyribosome profile. 373

Ribosome profiles at LG (blue line) and DS (red line) were superimposed, and the average ratio 374

of polysomes to nonpolysomes (P/NP) was calculated based on four independent sedimentation 375

experiments. Under LG conditions, the Ami-m line exhibited consistently lower 60S and 80S 376

peaks than the WT without significant difference in polysome profiles, while the overall profiles 377

of the MRF1 OE line were similar to those of WT (Figure 7C; blue line). The P/NP ratio in the 378

Ami-m lines (4.1 ± 0.3) appeared to be slightly higher than that of the WT (4.0 ± 0.1) and MRF1 379

OE lines (3.9 ± 0.3) in LG (2 h). 1-day DS treatment reduced the amount of polysomes and 380

increased the amount of 60S/80S in WT seedlings, implying repression of global translation 381

(Figure 7C; red line). Polysome peaks of the Ami-m and MRF1 OE lines were almost 382

indistinguishable from those of the WT. However, the 60S/80S peak was significantly lower in 383

the Ami-m line and slightly higher in the MRF1 OE line, respectively, than in the WT. Under DS 384

conditions, the P/NP ratio in the Ami-m (1.6 ± 0.1) and MRF1 OE lines (1.1 ± 0.1) was higher 385

and lower than that of the WT (1.3 ± 0.1), respectively. Thus, the P/NP ratio tended to differ 386

more significantly between the lines under DS than under LG, albeit not statistically significant. 387

These results suggest that MRF deficiency may cause decreased 60S/80S amounts under LG (2 h) 388

and DS (1 d) conditions, while MRF1 overexpression may cause increased 60S/80S amounts 389

only under DS (1 d) conditions. 390

We also performed polysome analyses without cycloheximide pretreatment after LG (2 h), 391

DS (2 h), and DS (1 d) treatments (Supplemental Figure 14). Removal of the pretreatment step 392

changed the shape of the ribosome profiles particularly in LG; it increased the 60/80S peaks and 393

decreased polysomes. Nevertheless, these experiments led to similar conclusions to those of 394

Figure 6C. The Ami-m lines showed decreased 60S/80S peaks under LG (2 h), DS (2 h), and DS 395

(1 d) conditions, while the MRF1 OE lines showed increased 60S/80S peaks only under DS (1 d) 396

14

conditions. 397

398

MRF Deficiency Decreases Cellular Ribosomal RNA Contents under DS Conditions 399

To determine whether MRF deficiency or MRF1 OE affects cellular rRNA levels, we examined 400

total RNA contents in the WT, Ami-m (#3, #10, and #14), and MRF1 OE (#1 and #2) lines after 401

prolonged DS and ReLG treatments, because total RNA is a useful proxy for rRNA (Figure 7D). 402

Seedlings grown under LG conditions (control) were subjected to 2 and 5 days of DS treatment 403

(DS+2 and DS+5), followed by 3 and 5 days of LG treatment (ReLG+3 and ReLG+5) as 404

described in Figure 4A. Total RNA contents in the seedlings of equal weight were measured at 405

each stage. There were no significant differences in the contents among these seedlings before 406

treatment (control). In all the lines, however, the RNA contents progressively decreased during 407

prolonged DS, and then increased upon subsequent supply of light and glucose (ReLG). Total 408

RNA contents were consistently lower in the Ami-m lines throughout DS and ReLG treatments, 409

suggesting that MRF deficiency accelerates rRNA degradation upon prolonged DS, considering 410

the stability of rRNAs, and delays rRNA recovery during ReLG. Notably, the Ami-m (#3) line 411

that showed the most severe phenotype during prolonged DS and ReLG treatments (Figure 4A) 412

also exhibited the most significant decrease in total rRNA contents during the treatments (Figure 413

7D). In contrast, the MRF1 OE lines contained higher RNA contents, particularly during ReLG 414

treatment, suggesting faster rRNA recovery. 415

416

MRF Deficiency Causes Altered Distribution of mRNAs in Sucrose Gradient Fractions 417

To analyze the effects of MRF deficiency on protein translation in detail, we examined the 418

distribution patterns of selected mRNAs in sucrose gradient fractions of WT and Ami-m (#3) 419

seedlings after LG (2 h) and DS (1 d) treatment (Figure 8A; Supplemental Figure 15A). Then we 420

quantified the abundance of the mRNAs in polysomal (P; fractions 9-15) and nonpolysomal 421

fractions (NP; fractions 1-8), as a percentage of their total amount in all fractions (Figure 8B; 422

Supplemental Figure 15B). The selected genes were previously used in translation studies 423

(Nicolai et al., 2006; Juntawong and Bailey-Serres, 2012; Perrella et al., 2013; Schepetilnikov et 424

al., 2013; Juntawong et al., 2014). After LG (2 h) or DS (1 d) treatment, cell extracts from the 425

seedlings were subjected to sucrose density gradient sedimentation using the same protocol as in 426

15

Figure 7C, and 15 fractions were collected from each tube. Total RNA was precipitated from 427

each fraction, followed by cDNA synthesis and RT-qPCR for calculation of the average mRNA 428

content in each fraction. 429

Distribution patterns of PP2AA3 (PP2A regulatory subunit A3), GAPC (cytosolic 430

glyceraldehyde-3-phosphate dehydrogenase), HDA19 (histone deacetylase), and HDC1 (histone 431

deacetylation complex1) mRNAs in sucrose gradients were very different between LG and DS 432

conditions in both WT and Ami-m lines (Figure 8A). While these mRNAs accumulated mostly in 433

heavy polysomes in LG, their accumulation shifted to less dense fractions including monosomes 434

in DS, suggesting a decrease in global translation under DS conditions. This result is consistent 435

with the report that HDA19 mRNA translation is repressed by sucrose starvation (Nicolai et al., 436

2006). For all four mRNAs, there was no significant difference between the WT and Ami-m lines 437

regarding the mRNA percentage in P and NP fractions under LG (Figure 8B). However, under 438

DS, the percentage of polysome-associated mRNAs were higher in the Ami-m lines than in the 439

WT samples for a significant portion of tested mRNAs, based on statistical analyses of three 440

biological replicates (Figure 8B; Supplemental Figure 15B). Those mRNAs include PP2AA3, 441

HDA19, HDC1, MCCA (3-methylcrotonyl-CoA carboxylase), CAB2, and RHIP1 (RGS1 and 442

HXK1 interacting protein 1), with particularly significant effects on HDA19 and HDC1. 443

However, there was no statistically significant difference between two samples for GAPC, bZIP1 444

(basic-leucine zipper transcription factor 1), and bZIP11 mRNAs. There is a possibility that the 445

higher polysomal loading of these particular mRNAs indicates more robust translation of the 446

mRNAs under DS. However, considering that Ami-m lines exhibited a decrease in 60S/80S 447

amounts, global translation activity, and cellular rRNA contents under DS conditions, it may 448

suggest that MRF deficiency delays/impairs normal progression of translation for the mRNAs. 449

Translation of some mRNAs such as HDA19 and HDC1 may be more sensitive to MRF 450

deficiency. Future studies including genome-wide translatome analyses would provide a clear 451

picture. 452

453

Phosphorylation and Ribosome Association of MRF1 Are Regulated by Cellular Energy 454

Availability 455

In animals and yeast, the rapid response to cellular nutrient availability is generally controlled by 456

16

phosphorylation of several regulatory proteins (Sonenberg and Hinnebusch, 2009; Jackson et al., 457

2010). Therefore, we examined whether MRF proteins are phosphorylated in vivo. Each MRF 458

protein was expressed as a Myc fusion in N. benthamiana leaves. Then the leaf extracts were 459

subjected to Zn2+

-Phos-tag SDS-PAGE, and also to normal SDS-PAGE, with or without 460

treatment with recombinant lambda phosphatase (λPP), a non-specific protein phosphatase. The 461

more common Mn2+

-Phos-tag SDS-PAGE system did not clearly separate phosphorylated MRF1 462

from its unphosphorylated form. Immunoblotting with anti-Myc antibody showed most visible 463

mobility shift in MRF1 after λPP treatment in Phos-tag SDS-PAGE; normal SDS-PAGE failed to 464

detect it (Figure 9A). Thus, Zn2+

-Phos-tag gels detected phosphorylation of MRF1 in vivo. 465

To test the possibility that MRF1 phosphorylation depends on cellular energy conditions, 466

we performed Phos-tag and normal SDS-PAGE using seedling extracts from MRF1 OE (#1) 467

lines, in which MRF1 is fused with the Flag-tag sequence (Figure 9B). We failed to make 468

specific antibodies against the endogenous MRF1. Twelve-day-old seedlings grown in liquid 469

culture were kept in DS conditions for 1-48 h, and then re-illuminated and fed with 30 mM 470

glucose (LG) for 0.5-2 h. Immunoblotting with anti-Flag antibody showed that the upper band, 471

representing phosphorylated MRF1, started to decrease after 2 h of DS treatment, almost 472

disappeared after 48 h, and reappeared after 30 min of LG treatment (Figure 9B). These results 473

suggest that MRF1 is de-phosphorylated under energy deprivation conditions, but becomes 474

rapidly phosphorylated in vivo when energy is reintroduced. 475

We examined MRF1 co-sedimentation with ribosomes under different energy conditions 476

(Figure 9C). Twelve-day-old MRF1 OE (#1) seedlings grown in liquid culture were harvested 477

without treatment (control), after 24-h DS treatment, or after 1-h LG treatment following DS, 478

and the extracts were loaded on a 10-55% sucrose gradient for ultracentrifugation. Inactive 479

monosomes were expected to accumulate in DS conditions, which decrease protein synthesis. As 480

expected, 24-h DS treatment led to a higher monosome peak and lower polysomal peaks than 481

observed for the control (Figure 9C, upper). After 1-h LG treatment following DS, the 482

monosome amount decreased and polysome amounts were partially restored. The ribosomal 483

fractions were then subjected to immunoblotting with anti-Flag antibody to detect Flag:MRF1, 484

and with anti-RPL10a antibody as control (Figure 9C, lower). In normal conditions, MRF1 was 485

broadly detected up to the 13th

fraction (light polysomes). After DS treatment, MRF1 was 486

17

distributed only up to the 9th

fraction (monosomes). Interestingly, despite the large increase in the 487

monosome peak upon DS, the amount of MRF1 found in the monosome fractions (fractions 6 488

and 7) did not increase proportionally (Figure 9C, center). This result suggests that MRF1 has 489

low affinity towards the inactive 80S ribosomes. 1-h LG treatment following the DS expanded 490

MRF1 distribution up to the 10th

fraction (light polysomes). These results suggest that MRF1-491

ribosome association is correlated with cellular translation activity, which depends on the energy 492

status of the cell. 493

To explore the mechanism of the condition-dependent MRF1-ribosome association, we 494

examined MRF1-eIF4A interactions under LG and DS conditions using co-immunoprecipitation 495

(Figure 9D). eIF4A-1 was broadly distributed in the ribosome fractions up to the polysome 496

fraction (Figure 6A). Leaf disks were prepared from N. benthamiana plants expressing 497

Flag:MRF1 and eIF4A-1:Myc, and then were floated for 3 h either in the light on MS medium 498

with 30 mM glucose (LG), or in the dark on MS medium without glucose (DS). Interestingly, 499

MRF1 was co-immunoprecipitated with eIF4A-1 more efficiently under LG, rather than DS, 500

conditions (Figure 9D). A stronger MRF1-eIF4A-1 interaction might have contributed to the 501

more robust MRF1-ribosome association under LG conditions. Collectively, these results suggest 502

that MRF1 phosphorylation and ribosome association are modulated by cellular energy 503

availability. 504

505

TOR Regulates MRF Gene Expression in Response to Starvation and Sugar Feeding 506

Our results suggest that MRF transcription is modulated by cellular energy status (Figure 1), and 507

that MRFs are involved in protein translation control (Figures 7, 8). TOR kinase is an important 508

regulator of protein translation and plays a role in linking nutrient signaling to cellular 509

adaptations (Dobrenel et al., 2016a). To explore a possible relationship between MRFs and the 510

TOR signaling pathway, we first tested whether the abundance of MRF transcripts is changed by 511

TOR activity, using estradiol-inducible TOR RNAi (es-tor1) lines (Figure 10A). Estradiol 512

treatment reduced the TOR mRNA levels (Supplemental Figure 16A). Nine-day-old seedlings 513

grown in liquid culture were treated for 24 h with either ethanol (-ES) or estradiol (10 µM; +ES). 514

Then they were transferred to fresh medium containing ethanol or estradiol, but lacking sugar, 515

and incubated for 24 h for starvation treatment. After 24 h of starvation, glucose, mannitol, or 516

18

sucrose was added to the medium (final concentration 30 mM) for further incubation for 4 h. RT-517

qPCR of (-)ES samples revealed that feeding with glucose or sucrose, but not with mannitol, led 518

to significantly lower transcript levels of MRF1, MRF3, and MRF4 than in starvation conditions 519

(Figure 10A), consistent with the previous results (Figure 1D, Supplemental Figure 1). However, 520

TOR silencing by estradiol treatment disrupted these transcript level changes of the three MRF 521

genes, with a particularly strong effect on MRF1 and MRF4. MRF2 gene expression remained 522

unchanged upon estradiol treatment. These results suggest that TOR regulates MRF gene 523

expression in response to the cellular energy status, except for MRF2. 524

525

Phosphorylation and Ribosome Association of MRF1 Are Regulated by TOR Activity 526

TOR inhibition reduces translation in eukaryotes, mainly through repression of both translation 527

initiation and new ribosome synthesis (Powers and Walter, 1999; Cherkasova and Hinnebusch, 528

2003). Recently, it was reported that TOR inhibition by rapamycin triggers a rapid decrease (40-529

60%) in ribosome content in yeast, through rapid cytoplasmic turnover of the existing ribosomes 530

(Pestov and Shcherbik, 2012). We peformed virus-induced gene silencing (VIGS) using the 531

tobacco rattle virus (TRV) system against TOR in MRF1 OE (#1) Arabidopsis plants, to analyze 532

phosphorylation status and ribosome association of Flag:MRF1 in TOR-silenced leaf cells 533

(Figure 10B, C). The 6-8th

leaves were collected from the TRV control and from the TOR VIGS 534

plants grown in soil at 10 days after infiltration (DAI), when morphological phenotypes were not 535

yet fully visible in the TOR VIGS plants. TOR VIGS caused defective plant growth and 536

excessive starch accumulation at 14 DAI, associated with reduced TOR mRNA levels 537

(Supplemental Figure 16B-D). At ~21 DAI, TOR VIGS plants exhibited growth arrest and severe 538

chlorosis. Immunoblotting of the Phos-tag gel with anti-Flag antibody revealed that the upper 539

phosphorylated form of Flag:MRF1 was significantly reduced in TOR-silenced cells, while the 540

lower unphosphorylated form was increased, suggesting that reduced TOR activity leads to 541

MRF1 dephosphorylation (Figure 10B). Ribosome fractionation was performed using the same 542

gram-fresh-weight of leaves from TRV control and TOR VIGS plants in a 10-55% sucrose 543

gradient, followed by immunoblotting with anti-Flag and anti-RPL10a antibodies (Figure 10C). 544

RPL10a distribution patterns in the ribosomal fractions suggested reduced translation activity in 545

TOR VIGS plants. Furthermore, MRF1 cofractionation with ribosomes was diminished in TOR-546

19

silenced cells, suggesting that MRF1 distribution is modulated by TOR activity either directly or 547

indirectly (Figure 10C). 548

Since TOR activity regulated MRF gene expression, MRF1 phosphorylation, and MRF1-549

ribosome association, we next examined the sensitivity of MRF Ami-m (#3, #10, and #14) and 550

MRF1 OE (#1 and #2) seedlings to Torin-1 (TOR inhibitor) compared with WT seedlings (Figure 551

10D, E). Since root growth is sensitive to TOR inhibition, we measured root length of the 552

seedlings after 3 and 5 days of Torin-1 (2 μM) or control DMSO treatment. Torin-1 caused 553

reduced root growth and premature senescence in all seedlings with respect to the DMSO 554

controls; the Ami-m and MRF1 OE seedlings appeared to be more susceptible and more resistant 555

to Torin-1 than WT seedlings, respectively (Figure 10D, E). Consistent with their senescence 556

symptoms, the Ami-m and OE seedlings showed the respective lower and higher CAB2 mRNA 557

levels than that of the WT at 3 days after Torin-1 treatment (Figure 10D; Supplemental Figure 558

17). Collectively, these results suggest that active TOR promotes MRF1 phosphorylation and 559

ribosome association, and MRF mutants show altered Torin-1 sensitivity. 560

561

MRF1 Is Phosphorylated by S6K1 and S6K2 in Vitro 562

Since MRF1 phosphorylation is regulated by the TOR signaling pathway, and S6 kinase (S6K) is 563

a major effector of TOR (Mahfouz et al., 2006; Schepetilnikov et al., 2013), we examined 564

whether MRF proteins are phosphorylated by S6K in vitro. MRF proteins were purified from E. 565

coli as MBP fusion proteins. S6K1:Myc and S6K2:Myc were expressed in N. benthamiana 566

leaves and immunoprecipitated using anti-Myc antibody-conjugated resin. In vitro protein kinase 567

assays showed that only MRF1 was phosphorylated by the immunoprecipitated S6K1 and S6K2 568

(Figure 11A, B), which is consistent with MRF1 being mainly phosphorylated in vivo (Figure 569

9A). MRF1 phosphorylation did not occur without the addition of S6K1 (Supplemental Figure 570

18A). Arabidopsis S6K1 has a conserved site, T449, which is phosphorylated by the TOR kinase 571

(Xiong and Sheen, 2012; Ahn et al., 2015). We determined MRF1 phosphorylation activities of 572

two mutant forms of S6K1, a phospho-null mutant (T449A) and a phospho-mimetic mutant 573

(T449D), and compared them to normal S6K1 (Figure 11C). The T449A mutation reduced 574

MRF1 phosphorylation activity of S6K1, while T449D had no effect. These results suggest that 575

phosphorylation at the T449 site in S6K1 is important for MRF1 phosphorylation in vitro (Figure 576

20

11C). The MBP control was not phosphorylated by either mutant form (Supplemental Figure 577

18B). 578

BiFC demonstrated that MRF1 interacts with both S6K1 and S6K2; co-expression of 579

YFPN-fused MRF1 and YFP

C-fused S6K1 or S6K2 in N. benthamiana leaves resulted in yellow 580

fluorescence in epidermal cells, observed by confocal microscopy (Figure 11D). MRF1 581

interactions with S6K1 and S6K2 were further confirmed by using co-immunoprecipitation 582

(Figure 11E, F). Flag:MRF1 and S6K1/2:Myc were co-expressed in N. benthamiana leaves. 583

S6K1:Myc and S6K2:Myc were immunoprecipitated from leaf extracts using anti-Myc antibody-584

conjugated resin. Immunoblotting with anti-Flag antibody detected Flag:MRF1 proteins co-585

immunoprecipitated with both S6K1:Myc and S6K2:Myc. Collectively, these results suggest that 586

MRF1 is a potential substrate of S6K1 and S6K2, and TOR-S6K1 signaling regulates MRF1 587

phosphorylation. 588

589

DISCUSSION 590

Low-energy stress induces massive transcriptional, translational, and metabolic reprogramming 591

in plants (Tome et al., 2014). The stress decreases global protein translation rates, but the 592

repression is rapidly reversible upon energy supply (Juntawong and Bailey-Serres, 2012). Recent 593

studies using global translation analyses suggested paradigms of plant translation control in 594

response to hypoxia, daily light-dark cycle, carbon deprivation, and extended darkness 595

(Juntawong and Bailey-Serres, 2012; Liu et al., 2012; Pal et al., 2013; Gamm et al., 2014; 596

Juntawong et al., 2014). Despite the general repression under unfavorable conditions, a basal 597

level of translation is essential for maintaining cell homeostasis during the nighttime and coping 598

with stress. However, the regulatory mechanisms of basal translation and the related control 599

factors are largely unknown in plants. Here, we identified MRF family proteins as translation 600

regulators in plants, with functions that are particularly important for translation under energy-601

deficient conditions. 602

We investigated the gene expression, protein characteristics, and cellular functions of four 603

MRF genes in Arabidopsis. Expression of all four MRF genes was induced in dark and starvation 604

(DS) conditions, but with different profiles (Figure 1C, D). MRF Ami-m lines designed to 605

simultaneously silence MRF1, MRF3, and MRF4 exhibited early flowering and early DS-606

21

induced senescence (Figures 3, 4). MRF1 OE lines exhibited the opposite phenotypes: late 607

flowering and delayed senescence in DS conditions. However, T-DNA insertion mutants and 608

amiRNA lines targeted for a single MRF gene did not show any significant phenotypic 609

differences from the WT, suggesting that three MRF genes, MRF1, MRF3, and MRF4, are 610

functionally redundant (Supplemental Figures 4, 6). It has been proposed that MRF family genes 611

divergently evolved through gene duplication events from the ancestral algal linage that 612

contained two MA3 domains (Cheng et al., 2013). It is possible that MRF proteins have 613

individual functions in other conditions through their differences in gene expression and 614

biochemical activity. 615

MA3 is known to be a protein-protein interaction domain, acting as an eIF4A-binding 616

module (Yang et al., 2003; Yang et al., 2004; Loh et al., 2009). eIF4A may be an archetypal 617

DEAD-box RNA helicase (Andreou and Klostermeier, 2013). Possessing RNA-dependent 618

ATPase activity and weak RNA helicase activity, eIF4A catalyzes the unwinding of mRNA 619

secondary structures at the 5′-UTR region to facilitate ribosome scanning through the region for 620

the initiation codon (Hemerly et al., 1995; Jackson et al., 2010). Moreover, eIF4A promotes the 621

dissociation of 5′-UTR-bound proteins and release of the 43S initiation complex from the 5′-cap, 622

using its helicase activity (Jankowsky et al., 2001; Andreou and Klostermeier, 2013). 623

Arabidopsis plants express three eIF4A proteins: eIF4A-1, eIF4A-2, and eIF4A-3. eIF4A-1 and 624

eIF4A-2 are involved in translation, but eIF4A-3, a nuclear protein, participates in RNA 625

processing by incorporating into the exon junction complex (Koroleva et al., 2009). Here, we 626

found that MRF proteins are associated with eIF4A and co-fractionated with ribosomes (Figures 627

5, 6A). MRF1 appears to interact with eIF4A-1 most strongly, and our data are consistent with 628

the possibility that MRF1 associates with ribosomes partially depending on RNA, although 629

additional experiments are required to confirm the finding (Figure 6B; Supplementary Figure 11). 630

MRFs can also rescue the BX04 mutant phenotype, albeit partially. Collectively, these results 631

suggest that MRFs are functionally linked to eIF4A and RNA. Interactions with both mRNA and 632

the translation apparatus may contribute to ribosome association of MRFs. 633

The human homolog of the MRFs is hPDCD4, which contains two MA3 domains. 634

hPDCD4 binding to eIF4A inhibits eIF4A helicase activity and interferes with eIF4A binding to 635

the MA3 domain of eIF4G, suggesting that hPDCD4 suppresses the translation initiation of 636

22

mRNAs with structured 5′-UTRs (Yang et al., 2003; Yang et al., 2004). Indeed, p53 mRNA with 637

a highly structured 5′-UTR has been identified as an endogenous target mRNA of hPDCD4 638

(Wedeken et al., 2011). hPDCD4 also appears to have an additional mechanism of translational 639

suppression; it binds to a secondary structure located in the coding region of c-myb mRNA and 640

blocks translation elongation through interaction with the poly(A)-binding protein (Fehler et al., 641

2014). Thus, the mechanisms of hPDCD4 action in translation control appear to be complex. 642

MRF silencing and MRF1 overexpression in DS conditions caused reduced and slightly 643

increased translation activity, respectively; these effects were not clear under LG conditions 644

(Figure 7; Supplemental Figure 13). Thus, normal MRF activity was required for efficient 645

translation in energy-deficient conditions. Interestingly, MRF silencing triggered a visible 646

decrease in 60S/80S amounts under DS conditions (Figure 7C; Supplemental Figure 14). In 647

contrast, 1-day DS slightly increased the 60S/80S peak in the MRF1 OE lines. Polysome profiles 648

were largely unaffected in all conditions. The average ratio of polysomes to nonpolysomes (P/NP) 649

in MRF deficient and MRF1 OE lines tended to be higher and lower than that of WT, 650

respectively, particularly in DS (Figure 7C). Furthermore, MRF deficiency increased the 651

percentage of polysome-associated mRNAs under DS conditions (for a significant portion of 652

mRNAs tested), despite the decrease in translation rates upon MRF silencing (Figure 8; 653

Supplemental Figure 15). 654

A high P/NP ratio usually represents efficient translation initiation, and mutations 655

impairing translation initiation generally decrease the polysome peak with a concomitant 656

increase in the monosome peak in eukaryotes (Kim et al., 2004; Jao and Chen, 2006; Bolger et 657

al., 2008; Saini et al., 2009; Juntawong and Bailey-Serres, 2012). If MRFs were repressors of 658

translation initiation, like hPDCD4, MRF deficiency would lead to a high P/NP ratio. However, 659

this hypothesis is not consistent with the reduction in 35

S-incorporation found after MRF-660

silencing in DS. On the other hand, specific defects in translation termination/ribosome recycling, 661

initiation, or elongation might have increased the P/NP ratio, simultaneously reducing the 662

translation rate. An MRF function in termination/ribosome recycling seems unlikely, because 663

MRFs were seldom detected in heavy polysomal fractions (Figures 6B, 9C), and were not co-664

immunoprecipitated with poly(A)-binding proteins, PAB2 and PAB8 (data not shown). 665

Furthermore, MRF interaction with eIF4A suggests MRF involvement in initiation control. 666

23

MRFs may act in concert with eIF4A to facilitate ribosome scanning through the 5′-UTR, and 667

MRF deficiency may lead to a reduction in active monosomes in DS. Finally, MRFs may 668

promote translation elongation. As described earlier, the Ami-m lines showed reduced protein 669

translation activity, accompanied by reduced total rRNA contents under DS. The Ami-m lines 670

also showed a higher P/NP ratio and an increase in polysome loading of specific mRNAs under 671

DS. Collectively, these results are most consistent with the hypothesis that MRF deficiency 672

delays translation elongation and subsequent ribosome run-off during DS. Interestingly, the 673

aforementioned molecular phenotypes of the Ami-m (#3) plants under DS mimicked those of the 674

yeast eIF5A mutants defective in translation elongation; disruption of eIF5A activity resulted in 675

reduced monosome amounts without affecting polysomes, and reduced 35

S-methionine 676

incorporation in yeast (Saini et al., 2009). Yeast eIF5A was broadly distributed in sucrose 677

gradients up to the polysome fractions in logarithmic phase cells, but less eIF5A was detected, 678

and only up to the monosome fractions, in stationary phase cells (Jao and Chen, 2006). These 679

ribosome association patterns are analogous with those of MRF1 under LG and DS conditions 680

(Figures 6B, 9C). Interestingly, eIF4A has been identified as a potential interacting partner of 681

pumpkin eIF5A isoforms (Ma et al., 2010). Taken together, our results suggest that MRFs play a 682

positive role in plant translation, possibly modulating translation initiation and/or elongation, and 683

are particularly important in low-energy conditions such as DS. Future studies will address 684

molecular mechanisms of MRF function in plant translation. 685

The dramatic decrease in rRNAs during prolonged DS in plants (Figure 7D) is reminiscent 686

of rapid rRNA degradation in E. coli cells during starvation or at the entry into the stationary 687

phase (Zundel et al., 2009; Piir et al., 2011). The authors proposed that ribosomes not engaged in 688

translation, and consequently present as subunits, are sensitive to endoribonuclease cleavage and 689

subsequent degradation (Zundel et al., 2009). Furthermore, TOR inactivation by rapamycin 690

rapidly decreased the cellular ribosome numbers by 40-60% in yeast, correlated with rRNA 691

degradation by cytoplasmic nucleases (Pestov and Shcherbik, 2012). These results suggest that 692

TOR controls both new ribosome biosynthesis and degradation of mature ribosomes, in order to 693

adjust the size of the translation machinery to changing environmental conditions. In this study, 694

MRF deficiency accelerated rRNA degradation during DS, while MRF1 OE accelerated rRNA 695

recovery during ReLG (Figure 7D). We speculate that the translation defect by MRF deficiency 696

24

under DS may lead to premature dissociation of the translating ribosomes or stalled ribosomes, 697

which then become a target for degradation via cytosolic ribonucleases, ribophagy, or both 698

(Zundel et al., 2009; Floyd et al., 2015). In contrast, MRF1 OE may boost mRNA translation 699

during ReLG, resulting in faster rRNA recovery. Alternatively, higher cell viability of MRF1 OE 700

seedlings during prolonged DS may be an underlying cause of faster rRNA recovery, 701

accomplished by higher rates of rDNA transcription and ribosome biogenesis. 702

Plant TOR kinase connects photosynthesis-driven nutrient availability to comprehensive 703

growth programs through signal transduction and transcription networks (Xiong et al., 2013). 704

TOR downregulation causes wide-reaching transcriptional changes for metabolic reprogramming, 705

accompanied by the accumulation of amino acids and organic acids, which mimics starvation 706

conditions (Moreau et al., 2012; Caldana et al., 2013). Here, we found that TOR modulates the 707

transcript abundance of MRF genes, except MRF2, in response to starvation and sugar feeding 708

(Figure 10A). Thus, MRF genes belong to the TOR-regulated transcriptional networks, which 709

fluctuate according to nutrient/energy availability. MRF silencing and MRF1 overexpression 710

caused hypersensitivity and resistance, respectively, to both DS and Torin-1 treatments (Figures 711

4, 10D, 10E). Furthermore, MRF1 phosphorylation was positively regulated by the TOR 712

pathway, possibly through the action of S6K1/2 (Figures 10B, 11). Mammalian PDCD4 is 713

phosphorylated by S6K1 following mTORC activation, but PDCD4 phosphorylation leads to its 714

dissociation from eIF4A and subsequent ubiquitylation and degradation (Dorrello et al., 2006; 715

Dennis et al., 2012). Transition from DS to LG conditions led to rapid phosphorylation and 716

incorporation of MRF1 into light polysomal fractions (Figure 9B, C), while TOR silencing 717

significantly inhibited MRF1 phosphorylation and decreased MRF1 co-fractionation with 718

ribosomes (Figure 10B, C). Combined with the fact that the affinity between MRF1 and eIF4A-1 719

is higher in LG than in DS (Figure 9D), these results suggest that MRF1 phosphorylation 720

positively correlates with active translation under LG conditions. Yet MRF1 becomes 721

dephosphorylated in DS, but is still co-sedimented with ribosomes. Thus, unphosphorylated 722

forms of MRF1 also appear to be active in translation, particularly under energy-deficient 723

conditions when their abundance increases. There is a possibility that MRFs may interact with 724

different helicases or initiation factors to promote translation in DS. Considering that MRF1 is 725

the most abundant MRF member under starvation conditions, phosphorylation of MRF1 by the 726

25

TOR-S6K signaling pathway during the transition from DS to LG may play a role in rapid 727

rebooting of active translation when the cell environment becomes favorable to growth. 728

Our data suggest that translation of some mRNAs is more sensitive to MRF deficiency: e.g. 729

HDA19 (histone deacetylase) and HDC1 (histone deacetylation complex1) (Figure 8; 730

Supplemental Figure 15). HDA19 and HDC1 are components of the histone deacetylase complex, 731

which epigenetically controls gene expression through repressive function of histone 732

deacetylation (Perrella et al., 2013). HDA19 and HDC1 are involved in plant’s response to ABA, 733

abiotic stresses, and in seed germination, among other functions (Perrella et al., 2013; Mehdi et 734

al., 2016). Altered translation of HDA19 and HDC1 mRNAs might have contributed to the MRF-735

deficient phenotypes in DS. Additionally, MRF functions may be connected to the ethylene 736

signaling pathway. MRF3 was initially identified as EIN2 C-terminus interacting protein 1 737

(ECIP1), and loss-of-function of ECIP1 caused an enhanced ethylene response (Lei et al., 2011). 738

The EIN2 C-terminus after cleavage was found to suppress the translation of EIN3-BINDING F-739

BOX 1 (EBF1) and EBF2 mRNA, both encoding negative regulators of the ethylene pathway, 740

through direct binding to the multiple poly-uridylate motifs in their 3′-UTR and forming 741

processing bodies with ETHYLENE INSENSITIVE 5 (EIN5) and poly(A)-binding proteins (Li 742

et al., 2015; Merchante et al., 2015). In this scenario, MRF3 may inhibit the translation 743

repression activity of EIN2 via a protein-protein interaction near the 3′-UTR of EBF1/2 mRNA. 744

It is unclear whether the EIN2-interacting activity is confined to MRF3 among the four MRF 745

members. Further studies are required to uncover the detailed mechanisms of MRF action in 746

translation and their effects on mRNA translation at a global scale. 747

748

METHODS 749

750

Plant Materials and Growth Conditions 751

Arabidopsis thaliana (ecotype Columbia-0) plants were grown in a growth chamber [23 °C, 100-752

120 μmol m–2

s–1

light intensity using light bulbs (Philips TLD36W/865/FL40SS/36/EX-D), and 753

16 h light:8 h dark cycle]. For the liquid culture, Arabidopsis seeds were surface sterilized and 754

sown in six-well plates containing 1 mL of liquid medium (0.5× Murashige-Skoog [MS] medium 755

[Duchefa], pH 5.7 adjusted with KOH). After germination, seedlings were grown in 0.5× MS 756

26

medium containing 30 mM glucose, changed every other day. 757

758

Generation of Arabidopsis Transgenic Lines 759

AmiRNAs targeting MRF genes were designed by using the Web MicroRNA Designer 760

(http://wmd3.weigelworld.org). Seven different amiRNA PCR products were generated using 761

primer pairs (Supplemental Table 2) and the amiRNA cloning vector pRS300, containing the 762

miR319a backbone as a template, as previously described (Schwab et al., 2006). PCR products 763

were cloned into the pGEM-T-easy vector (Promega), and then transferred to the 764

pCAMBIA1390 vector using the SalI and BamHI sites. To generate MRF1 overexpression lines, 765

the MRF1 protein-coding sequence was cloned into the pCAMBIA-Flag vector using the SalI 766

and EcoRI sites. Arabidopsis (Col-0) plants were transformed by the floral dip method (Clough 767

and Bent, 1998), using Agrobacterium GV3101 strain. More than 30 independent T1 lines were 768

generated for each construct, from which 5-7 T2 lines were selected for T3 propagation, based on 769

gene expression levels. The seed batch that showed 100% hygromycin resistance was selected as 770

the homozygous T3 generation. T3 and T4 homozygous seeds were used for the analyses. 771

772

Virus-Induced Gene Silencing (VIGS) 773

VIGS was performed in Arabidopsis as described previously (Burch-Smith et al., 2006), using 774

soil-grown seedlings at two- to four-leaf stages. A 649-bp cDNA fragment of Arabidopsis TOR 775

was PCR-amplified using TOR-specific primers (Supplemental Table 2). The cDNA fragment 776

was then cloned into the pTRV2 vector (Burch-Smith et al., 2006) using the EcoRI and XhoI 777

sites. pTRV1 containing the viral RNA-dependent RNA polymerase, pTRV2-TOR, and pTRV2 778

empty vector (control) were introduced into Agrobacterium tumefaciens GV3101 strain. The 779

recombinant Agrobacterium strains were cultured overnight in Luria-Bertani medium containing 780

10 mM MES, 20 μM acetosyringone, Kanamycin (50 μg/mL), and Rifampicin (50 μg/mL), and 781

then harvested and resuspended in infiltration medium (10 mM MgCl2, 10 mM MES, and 200 782

μM acetosyringone) to OD600=1.5. After incubation at 23°C for 4 h, the Agrobacterium culture 783

was infiltrated into the largest true leaf using a needless syringe. 784

785

Agrobacterium tumefaciens-Mediated Transient Expression 786

27

Agrobacterium-mediated transient expression was performed using Agrobacterium C58C1 strain 787

as described (Voinnet et al., 2003), except for BiFC experiments that used the GV3101 strain. 788

Overnight-grown Agrobacterium culture was resuspended in infiltration medium (10 mM MES-789

KOH, pH 5.7, 10 mM MgSO4, and 500 μM acetosyringone) to different OD600 depending on 790

experiments (described below), and incubated for 2 h at room temperature, before infiltration 791

into N. benthamiana leaves. In all experiments, Agrobacterium strain carrying the 35S:p19 792

construct (Voinnet et al., 2003) was co-infiltrated at different OD600 ratio depending on 793

experiments (described below), in order to achieve maximum levels of protein expression. 794

Expressed proteins were analyzed at 48-72 h post-infiltration. 795

796

Analysis of the Re-Greening Phenotypes 797

The re-greening assay was performed using 12-day-old seedlings grown in liquid culture. After 798

washing five times with the 0.5X MS medium without glucose, seedlings were incubated for 5 799

days in the dark in the medium lacking glucose for dark/starvation (DS) treatment. After DS, the 800

seedlings were supplied with fresh medium containing 30 mM glucose and incubated in the light 801

for 5 days under the long-day conditions. 802

803

Detached Leaf Senescence Assay 804

The 5th

and 6th

leaves from 3-week-old Arabidopsis plants grown in soil were used for the assay. 805

Detached leaves were floated adaxial side up on the surface of sterilized water in petri dishes. 806

Plates were placed at 23oC in the dark for the indicated times. 807

808

Chlorophyll Measurement 809

Chlorophyll was extracted from four seedlings or individual leaves in aqueous 80% acetone. 810

Absorbance of the extract was measured at 663.6 and 646.6 nm using VersaMax Absorbance 811

Microplate Reader (Molecular Devices). The total chlorophyll contents were calculated based on 812

the absorbance as previously described (Porra and Scheer, 2000), and normalized by fresh weight. 813

814

RT-qPCR 815

Total RNA was extracted using the IQeasy Plus Plant RNA Extraction Mini Kit (iNtRON 816

28

Biotechnology, Korea) according to the manufacturer’s instructions. 2 μg of total RNA was used 817

for cDNA synthesis using the RevertAid First Strand cDNA Synthesis Kit (Thermo Fisher 818

Scientific) with oligo-dT primers according to the manufacturer’s instructions. Real-time 819

quantitative RT-PCR was performed with diluted cDNAs (1:100) in 96-well plates using the 820

SYBR Premix Ex Taq (TAKARA) and the StepOnePlus Real-Time PCR System (Applied 821

Biosystems), as previously described (Ahn et al., 2011). PP2AA3 (protein phosphatase 2A 822

subunit A3) mRNA was used as control for normalization. To determine relative expression 823

levels of four MRF genes, the Ct (threshold cycle) value of each MRF gene was compared with 824

that of PP2AA3 mRNA as control. 825

826

Subcellular Localization 827

Protein coding sequences of the MRF genes were PCR-amplified using primers listed in 828

Supplemental Table 2, and then cloned into the pCAMBIA1390-sGFP vector using SalI and 829

SmaI sites for MRF1, MRF2, and MRF3, and SalI and EcoRI sites for MRF4. These GFP 830

constructs were agro-infiltrated into N. benthamiana leaves. Agrobacterium strains containing 831

the GFP construct and 35S:p19 construct were co-infiltrated into N. benthamiana leaves at the 832

OD600 ratio of 1.5:0.8. Expression of the GFP fusion proteins was monitored 48 h post-833

infiltration in protoplasts prepared from the infiltrated leaves by a confocal laser scanning 834

microscope (Carl Zeiss LSM 510). 835

836

Subcellular Fractionation 837

Nuclear and cytosolic fractionation was performed using the CelLytic PN Plant Nuclei 838

Isolation/Extraction Kit (Sigma-Aldrich) according to the manufacturer’s protocol. We followed 839

the protocol for semi-pure nuclei preparation for Arabidopsis plants. After fractionation, each 840

fraction was mixed with 2X SDS sample buffer, and subjected to SDS-PAGE using 10-15% 841

gradient gel and immunoblotting with the mouse monoclonal anti-GFP antibody (Clontech, cat. 842

no. 632381, lot no. A0042539; 1:10,000) and the rabbit polyclonal anti-Histone H3 antibody 843

(Santa Cruz, cat. no. Sc10809, lot no. G0110; 1:2000). Signals were detected by Imagequant 844

LAS 4000 (GE Healthcare Life Sciences). 845

846

29

BiFC 847

BiFC was performed with the pSPYNE vector containing the N-terminal region of YFP (amino 848

acid residues 1-155) and the pSPYCE vector containing the C-terminal region of YFP (residues 849

156-239) (Walter et al., 2004). MRF1 and MRF4 coding sequences (CDS) were cloned into 850

pSPYNE using SalI and KpnI sites, while MRF2 and MRF3 CDSs were cloned into pSPYNE 851

using SalI and SmaI sites. Arabidopsis eIF4A-1 and eIF4E-1 CDSs were cloned into pSPYCE 852

using SpeI/SmaI and XbaI/XhoI sites, respectively. Construction of pSPYCE-S6K1 and 853

pSPYCE-S6K2 was previously described (Ahn et al., 2015). Agrobacterium strains containing 854

the pSPYNE, pSPYCE, and 35S:p19 construct was co-infiltrated at the OD600 ratio of 1:1:1.5 855

into leaves of 3 week-old N. benthamiana plants. BiFC signals were monitored 48 h post-856

infiltration in the abaxial side of leaf epidermis using a confocal laser scanning microscope 857

(Zeiss LSM510). To detect protein expression, 50 μg of protein extract was subjected into SDS-858

PAGE using 10-15% gradient gel and immunoblotting was performed using the goat polyclonal 859

anti-GFP antibody (ABM, cat. no. G095; 1:5,000). Signals were detected by Imagequant LAS 860

4000 (GE Healthcare Life Sciences). 861

862

Yeast Two Hybrid Assay 863

The Matchmarker Gold Yeast Two-Hybrid System (Clontech) was used for the analysis. 864

Arabidopsis eIF4A-1 CDS was cloned into the pGADT7 vector (Clontech) containing the GAL4 865

activation domain using EcoRI and SmaI sites. CDSs of the MRF genes were cloned into the 866

pGBKT7 vector (Clontech) containing the GAL4 DNA-binding domain using EcoRI and SalI 867

sites. Yeast two hybrid assays were performed according to the manufacturer’s manual 868

(Clontech). Alpha-galactosidase activity was measured by using the 200-μL scale assay with 48-869

h incubation time, according to the manufacturer’s instruction (Clontech). The activity was 870

statistically analyzed with three biological replications. To examine protein expression in yeast 871

cells, proteins were extracted from 800 μL of yeast cell cultures (OD600 = 1.5) using the NaOH/β-872

mercaptoethanol extraction method. Then the protein extracts (5 μL) were subjected to 8% SDS-873

PAGE and immunoblotting using the anti-c-Myc-Peroxidase-conjugated antibody (Sigma-874

Aldrich, cat. no. A5598, lot no. 045M4854; 1:5,000) and the anti-HA-Peroxidase-conjugated 875

antibody (Roche, cat. no. 12013819001; 1:10,000). Signals were detected by Imagequant LAS 876

30

4000 (GE Healthcare Life Sciences). 877

878

Complementation of E. coli BX04 Strains 879

Bacterial complementation using the pINIII vector was performed as previously described 880

(Nakaminami et al., 2006). The pINIII-CspA plamid was obtained from Dr. Hunseung Kang 881

(Chonnam National University, Korea). Protein coding sequences of the MRF genes and 882

Arabidopsis LOS4 were cloned into the pINIII vector using EcoRI and EcoRI/BamHI sites, 883

respectively. The recombinant plasmids were transformed into E. coil BX04 strain (ΔcspA, 884

ΔcspB, ΔCspE, and ΔcspG) (Xia et al., 2001). The transformed BX04 cell culture (OD600=1) was 885

diluted and spotted onto LB plates containing ampicillin (50 μg/mL), kanamycin (50 μg/mL), 886

and IPTG (0.1 mM) for induction of protein expression. Spotted plates were incubated at 37°C 887

for 1 day or at 18°C for 5 days. 888

889

Co-Immunoprecipitation 890

Arabidopsis eIF4A-1 CDS was cloned into the pCAMBIA1390-6xMyc vector using PstI and 891

EcoRI sites. CDSs of the MRF genes were cloned into the pCAMBIA1390-3xFlag vector using 892

SalI and EcoRI sites. Construction of pCAMBIA1390-S6K1:Myc and pCAMBIA1390-893

S6K2:Myc was previously described (Ahn et al., 2015). Agrobacterium strains containing each 894

pCAMBIA construct and 35S:p19 construct was co-infiltrated into 4-week-old N. benthamiana 895

leaves at the OD600 ratio of 1.0:1.0:1.5. At 72 h post-infiltration, leaves were ground and mixed 896

with an equal volume of ice-cold immunoprecipitation buffer [50 mM sodium phosphate buffer, 897

pH 7.4, 150 mM NaCl, 10% glycerol, 5 mM EDTA, 1 mM DTT, 1% Triton X-100, 50 mM 898

MG132, 2 mM Na3VO4, 2 mM NaF, 20 mM β-glycerophosphate, and cOmplete protease 899

inhibitor cocktail (Roche)]. After brief centrifugation to remove cell debris, the supernatant 900

containing 1 mg of total protein was incubated with EZview Red anti-c-Myc affinity gel (10 μL 901

gel per 1 mg of total proteins; Sigma-Aldrich) at 4oC for 4 h. After incubation, the affinity gel 902

was washed four times with IP washing buffer (50 mM sodium phosphate buffer, pH 7.4, 150 903

mM NaCl, 10% glycerol, 5 mM EDTA, 1 mM DTT, and 0.1% Triton X-100), and then 904

resuspended in 50 μL 2X SDS sample buffer. 10 μg of INPUT protein (1%) and 15 μL of IP elute 905

were subjected into 10 % SDS-PAGE. Immunoblotting was performed using the mouse 906

31

monoclonal anti-D tag antibody (ABM, cat. no. G191; 1:5,000) that recognizes Flag epitope, and 907

the mouse monoclonal anti-Myc antibody (ABM, cat. no. G019; 1:5,000). Signals were detected 908

by Imagequant LAS 4000 (GE Healthcare Life Sciences). 909

910

35S-Methionine Labeling 911

35S-methionine labeling was performed as described previously (Ahn et al., 2011). Twelve-day-912

old seedlings grown in liquid culture were pre-treated in LG or DS conditions for 30 min, and 913

then treated with 50 µCi of 35

S-methionine for 2.5 h under the same conditions. After two washes 914

with the culture medium, total proteins were extracted, normalized by Bradford assay, and 915

subjected to 8% Bis-Tris NuPAGE (Thermo Fisher Scientific). The gel was stained with Sun-gel 916

staining solution (LPS Solution, Korea) and dried using a gel drier. Radioactive signal was 917

detected by a phosphorimager (BAS-2500; Fujifilm). Intensity of Coomassie Brilliant Blue 918

staining and radioactive signal was measured by ImageJ software from same area (25-100 kDa) 919

of the gel (https://imagej.nih.gov/ij/). 920

921

Sucrose Density Gradient Sedimentation and Polysome Analysis 922

Frozen tissues (0.2 g) were mixed with 1 mL of polysome isolation buffer (200 mM Tris-HCl, 923

pH 8.4, 50 mM KCl, 1% sodium deoxycholate, 25 mM MgCl2, 2% polyethylene [10] tridecyl 924

ether, 400 U/mL RNasin [Promega], and 50 μg/mL cycloheximide). Cell debris was removed by 925

brief centrifugation. Cell extracts (500 μL) were loaded onto an 11.5 mL sucrose gradient and 926

spun in a Beckman SW41Ti rotor at 38,000 rpm for 3.5 h at 4 °C. Fifteen fractions of 0.8 mL 927

were collected using the BioLogic low-pressure liquid chromatography system (Bio-Rad) with a 928

fraction collector. Absorbance was automatically detected at 254 nm for polysome analysis. One-929

half volume of all fractions was precipitated using the methanol precipitation method. After 930

precipitation, the pellet was resuspended in 50 μL of 1X protein sample buffer (200 mM Tris-931

HCl, pH 6.8, 3% SDS, 15% β-mercaptoethanol, 30% glycerol, 0.09% bromophenol blue, and 932

100 mM DTT) and incubated in 70℃ for 10 min. 10-20 μL of protein samples were loaded onto 933

8% Bis-Tris NuPAGE gels. Immunoblotting was performed with mouse monoclonal antibodies 934

against GFP (Clontech; cat. no. 632381, lot no. A0042539; 1:10,000), Myc (ABM, cat. no. G019; 935

1:5,000), and RPL10a (Santa Cruz, cat. no. Sc-100827, lot no. H2212; 1:4,000). 936

32

937

RNase A Treatment 938

Cell extracts (500 μL) in polysome isolation buffer (except RNase inhibitor) were incubated with 939

RNAse A at the final concentration of 1 mg/mL (Sigma-Aldrich) for 5 min on ice. After 5 min, 940

200 U of RNase inhibitor (RNasin) was added into the reaction mix to stop the RNAse action. 941

Then the reaction mix was loaded onto sucrose density gradients for polysome analysis and 942

fractionation as described above. 943

944

Analyses of mRNA Distribution Patterns and Quantification 945

To purify RNA, each sucrose gradient fraction (800 μL) was mixed with 300 μL of QIAzol Lysis 946

Reagent (QIAGEN) and 300 μL of chloroform, followed by vortexing and brief centrifugation. 947

To precipitate RNA, the upper aqueous layer was mixed with 700 μL of isopropanol and 70 μL 948

of 3 M sodium acetate (pH 5.7), and then centrifuged for 20 min at 4℃ at the maximum speed. 949

The pellet was washed with 75% ethanol and resuspended in 20 μL of RNase-free water. 7.5 μL 950

of RNA was used for cDNA synthesis using the RevertAid First Strand cDNA Synthesis Kit 951

(Thermo Fisher Scientific) with oligo-dT primers according to the manufacturer’s instructions. 952

The cDNA was diluted 20 times in water and analyzed by real-time quantitative PCR using 953

SYBR Premix Ex Taq (TAKARA) and the StepOnePlus Real-Time PCR System (Applied 954

Bioscience). After PCR, the obtained Ct values were converted to transcript amounts. The 955

abundance of mRNA in each fraction was calculated as a percentage of their total amount in all 956

fractions. 957

958

Zn2+

-Phos-Tag SDS-PAGE and Lambda Phosphatase Treatment 959

Zn2+

-Phos-tag SDS-PAGE was carried out under neutral pH conditions as previously described 960

(Kinoshita and Kinoshita-Kikuta, 2011), with some modifications. Homogenized samples were 961

resuspended in the same volume of extraction buffer (50 mM Tris-HCl, pH 7.5, 100 mM NaCl, 1 962

mM DTT, and cOmplete protease inhibitor cocktail [Roche]). For the λ-phosphatase treatment, 963

10 mM MnCl2 and either 400 U (1 μL) λ-phosphatase (New England Biolabs), or enzyme 964

storage buffer (100 mM NaCl, 50 mM HEPES, pH 7.5, 0.1 mM MnCl2, 0.1 mM EGTA, 2 mM 965

DTT, 0.01% Brij35, and 50% glycerol) was added to aliquots (21.5 μL) of protein extracts for 966

33

incubation at 30 °C for 1 h. The reaction was stopped by adding 3× protein sample buffer (200 967

mM Tris-HCl, pH 6.8, 3% SDS, 15% β-mercaptoethanol, 30% glycerol, 0.09% bromophenol 968

blue, and 100 mM DTT), followed by incubation at 70°C for 10 min. Then the samples were 969

subjected to 8% Zn2+

-Phos-tag Bis-Tris NuPAGE with 50 μM ZnCl2 and 50 μM Phos-taq system 970

(Wako). After electrophoresis, the Phos-tag gels were washed twice for 10 min each with transfer 971

buffer (10 mM CAPS, pH 11) containing 10 mM EDTA, followed by 10 min washing with 972

transfer buffer without EDTA. Immunoblotting was performed with the rabbit polyclonal anti-973

Myc antibody (ABM, cat. no. G019; 1:5,000) and anti-Flag M2-HRP-conjugated antibody 974

(Sigma-Aldrich, cat. no. A8592, lot no. 059K6059; 1:10,000), according to the manufacturer’s 975

instructions. 976

977

Purification of Recombinant Proteins 978

CDSs of MRF genes were cloned into pMal-C2X vector (New England Biolabs) for MBP fusion, 979

using EcoRI and SalI sites. The constructs were transformed into E. coli BL21 (DE3) strain. 980

Cells were grown in 1% glucose-NZCYM medium containing ampicillin (50 µg/mL) at 37°C to 981

an OD600 of 0.4, and induced by 0.25 mM IPTG for 16 h. MBP fusion proteins were purified 982

using Amylose Resin (New England Biolabs), following the manufacturer’s instruction with 983

minor modification. We used a single buffer (20 mM Tris-HCl, pH 7.5, and 200 mM NaCl) 984

throughout the purification procedure, and added 10 mM maltose into the buffer for elution. 985

After purification, proteins were concentrated using Amicon Ultracel 30K (Millipore) according 986

to the manufacturer’s instruction. Purified proteins were stored at -80oC until use. 987

988

In Vitro Kinase Assay Using Immunoprecipitated S6K1 and S6K2 989

Kinase assay was performed as previously described by Xiong et al. (2013) with minor 990

modifications. S6K1:Myc or S6K2:Myc proteins were transiently expressed in N. benthamiana 991

leaves via Agrobacterium infiltration. Infiltrated leaves were ground on liquid nitrogen and 992

mixed with extraction buffer (50 mM Sodium phosphate buffer, pH 7.4, 150 mM NaCl, 10% 993

Glycerol, 5 mM EDTA, 1 mM DTT, 1% Triton X-100, 50 mM MG132, 2 mM Na3VO4, 2 mM 994

NaF, 20 mM β-glycerophosphate, and cOmplete protease inhibitor cocktail [Roche]). After brief 995

centrifugation, the supernatant was mixed with EZview Red Anti-c-Myc affinity gel (10 μL gel 996

34

per 1 mg of total proteins; Sigma-Aldrich) for incubation at 4 °C for 4 h. After incubation, the 997

affinity gel was washed twice with low-salt buffer 1 (40 mM HEPES, pH 7.4, 150 mM NaCl, 2 998

mM EDTA, 10 mM pyrophosphate, 10 mM β-glycerophosphate, and 0.3% CHAPS), followed 999

by washing twice with low-salt buffer 2 (25 mM HEPES, pH 7.4, and 20 mM KCl). Kinase 1000

reaction was carried out with 2 μCi [γ-32

P]-ATP, immunoprecipitated S6K1/2, and 10 μg1001

recombinant substrate proteins in 20 μL kinase buffer (20 mM HEPES, pH 7.4, 125 mM NaCl, 1 1002

mM DTT, 10 mM MgCl2, 5 mM MnCl2, and 10 μM ATP) for 30 min at 30 °C. The reaction was 1003

stopped by adding 2× SDS sample buffer and boiling for 1 min. After SDS-PAGE, the gel was 1004

stained with Sun-Gel staining solution (LPS Solution, Korea), and dried using a gel drier. 1005

Radioactivity within the gel was detected using a phosphorimager (BAS-2500; Fujifilm). 1006

Immunoblotting of input samples was performed with the mouse monoclonal anti-Myc antibody 1007

(ABM, cat. no. G019; 1:5,000). 1008

1009

Accession Numbers 1010

All accession numbers can be found in Supplemental Table 1. 1011

1012

1013

SUPPLEMENTAL DATA 1014

Supplemental Figure 1. RT-PCR Analyses of MRF Expression in Response to Mannitol 1015

Feeding after Starvation (Related to Figure 1D). 1016

Supplemental Figure 2. Relative Transcript Levels of the MRF Genes (Related to Figure 1D). 1017

Supplemental Figure 3. Analyses of T-DNA Insertion Mutants of MRF3 and MRF4, and 1018

AmiRNA Lines of MRF1 and MRF2 (Related to Figure 3). 1019

Supplemental Figure 4. MRF4 mRNA Levels in Ami-m Lines and Flowering Phenotypes in T-1020

DNA Insertion Mutants and AmiRNA Lines of the MRF Genes (Related to Figure 3). 1021

Supplemental Figure 5. Measurement of Chlorophyll Contents and Detached Leaf Senescence 1022

Assay (Related to Figure 4A). 1023

Supplemental Figure 6. Re-greening Phenotypes of AmiRNA Lines and T-DNA Insertion 1024

Mutants (Related to Figure 4A). 1025

Supplemental Figure 7. Gene Expression of Ami-m and MRF1 OE Lines (Related to Figure 4B, 1026

35

C). 1027

Supplemental Figure 8. Protein Sequence Alignment of MA3 Domains in Diverse Proteins 1028

(Related to Figure 5). 1029

Supplemental Figure 9. BiFC Assays for Interactions between MRFs and eIF4A (Related to 1030

Figure 5A). 1031

Supplemental Figure 10. Immunoblotting to Determine Protein Expression in BiFC 1032

Experiments Shown in Figure 5A. 1033

Supplemental Figure 11. Co-immunoprecipitation of MRFs with the eIF4A Family Proteins 1034

(Related to Figure 5B). 1035

Supplemental Figure 12. Immunoblotting to Determine Protein Expression in Yeast Two-1036

Hybrid Experiments Shown in Figure 5C. 1037

Supplemental Figure 13. 35

S-Methionine Labeling under Light/Glucose Conditions (Related to 1038

Figure 7A, B). 1039

(Related to Figure 7A, B). 1040

Supplemental Figure 14. Polysome Analyses without a Brief Cycloheximide Treatment Step 1041

before Sample Harvest (Related to Figure 7C). 1042

Supplemental Figure 15. Distribution Patterns and Quantifications of Specific mRNAs in 1043

Sucrose Gradient Fractions (Related to Figure 8). 1044

Supplemental Figure 16 Gene Silencing Phenotypes and Reduced mRNA Levels of TOR in 1045

TOR RNAi and VIGS Plants (Related to Figure 10A-C). 1046

Supplemental Figure 17. CAB2 Transcript Levels in Seedlings after Torin-1 Treatment (Related 1047

to Figure 10D). 1048

Supplemental Figure 18. Control Experiments for in Vitro Kinase Assay Shown in Figure 11. 1049

Supplemental Table 1. Information on the Genes Used in This Study. 1050

Supplemental Table 2. Primers Used in This Study. 1051

1052

ACKNOWLEDGEMENTS 1053

1054

The authors wish to thank Drs. Masayori Inouye (Rutgers University, USA) and Sangita Phadtare 1055

(Rowan University, USA) for providing the E. coli BX04 mutant strain and pINIII vector, Dr. Jen 1056

36

Sheen (Harvard Medical School, USA) for providing seeds of the estradiol-inducible TOR RNAi 1057

lines, Dr. Detlef Weigel (Max Planck Institute for Developmental Biology, Germany) for the 1058

pRS300 vector, and Dr. Hunseung Kang (Chonnam National University, Korea) for helpful 1059

discussions. This research was supported by the Cooperative Research Program for Agriculture 1060

Science & Technology Development (Project numbers PJ01114701 [PMBC] and PJ01118901 1061

[SSAC]) from the Rural Development Administration, and the Mid-Career Researcher Program 1062

(NRF-2016R1A2B4013180) from the National Research Foundation (NRF) of the Republic of 1063

Korea. 1064

1065

AUTHOR CONTRIBUTIONS 1066

1067

D.-H.L. performed most of the experiments and analyzed the results together with S.J.P. and 1068

C.S.A. D.-H.L. and H.-S.P. designed the experiments and wrote the manuscript. All authors 1069

discussed the results and commented on the manuscript. 1070

1071

FIGURE LEGENDS 1072

1073

Figure 1. Predicted Protein Structure and Expression Patterns of Four MRF Genes. 1074

(A) Schematic representation of four Arabidopsis MRF proteins with four MA3 domains 1075

arranged in tandem. Residue numbers are marked. aa, amino acids. 1076

(B) RT-qPCR analyses of MRF gene expression in plant organs. Three organ pieces were 1077

collected from three different 6-week-old Arabidopsis plants: the 7th

and 8th

rosette leaves (RL), 1078

the 1st cauline leaves (CL), stems (~1 cm from the bottom; St), and the primary roots (R). Ten 1079

pieces of buds (stages 11-12; B) and open flowers (F) were also collected from the three plants. 1080

The collected tissues were combined for RNA extraction and RT-qPCR. Transcript levels are 1081

expressed relative to those in rosette leaf (RL). 1082

(C) RT-qPCR analyses of MRF gene expression in response to darkness. Twelve seedlings grown 1083

in three different sets in liquid culture were incubated in the dark for the indicated times. 1084

Transcript levels are expressed relative to those at 0 h. 1085

37

(D) RT-qPCR analyses of MRF gene expression in response to starvation and glucose feeding. 1086

Twelve seedlings grown in three different sets in liquid culture were incubated in glucose-free 1087

medium for 24 h (starvation; S), and then fed with the indicated concentrations of glucose for 4 h. 1088

Transcript levels are expressed relative to those before starvation (BS). 1089

For (B) to (D), transcript levels are normalized by PP2AA3 mRNA, and error bars represent 1090

standard errors (SE) calculated from triplicate technical replications. 1091

1092

Figure 2. Cytosolic Localization of MRF Proteins. 1093

(A) Subcellular localization of MRF:GFP fusion proteins in leaf protoplasts. Each MRF:GFP1094

was transiently expressed together with histone H2B:mRFP as a nuclear marker in N. 1095

benthamiana leaves via agro-infiltration. Protoplasts were prepared from the infiltrated leaves, 1096

and observed by confocal microscopy. Chlorophyll autofluorescence was pseudo-colored blue. 1097

More than 20 cells showing green fluorescence were observed for each construct. Scale bars = 1098

10 µm. 1099

(B) Subcellular fractionation. N. benthamiana leaf extracts expressing MRF:GFP proteins were1100

fractionated and subjected to SDS-PAGE using 10-15% gradient gel, followed by 1101

immunoblotting with anti-GFP antibody. Total (T), nuclear (N), and cytosolic (C) fractions were 1102

indicated. Histone H3 was detected as a nuclear marker protein using anti-H3 antibody. Two 1103

independent experiments yielded similar results. 1104

(C) Confocal microscopy of GFP fluorescence in epidermal cells of the Arabidopsis transgenic1105

plants that express each MRF gene under the CaMV35S promoter. Multiple independent 1106

transgenic lines were analyzed for each MRF gene, which similarly suggested cytosolic 1107

localization of MRF proteins. More than three independent observations were made for each 1108

transgenic line. Scale bars = 10 µm. 1109

1110

Figure 3. Generation of MRF Artificial miRNA and MRF1 Overexpression Lines, and Analysis 1111

of their Flowering Phenotypes. 1112

(A) Description of MRF artificial miRNA (amiRNA) and MRF1 overexpression (OE) lines (left),1113

and target sites for the amiRNA lines (right). The target sites (arrowheads) were designed for 1114

38

silencing of MRF1, MRF2, or multiple genes (MRF1, MRF3, and MRF4). “Ami-m” represents 1115

amiRNA lines with multiple targets. 1116

(B) RT-qPCR to determine MRF1 and MRF3 mRNA levels in the Ami-m lines. Transcript levels 1117

in the Ami-m lines are expressed relative to those in the WT. Values represent the means ± S.E. 1118

of N = three biological replicates of 10-day-old seedlings grown in different sets in liquid culture. 1119

Asterisks denote statistical significance of the differences between the WT and the transgenic 1120

lines, calculated using Student’s t-test (***, P ≤ 0.001). 1121

(C) RT-qPCR to determine MRF1 mRNA levels in the MRF1 OE lines, compared with those in 1122

WT. Error bars represent SE from triplicate biological replications using 10-day-old seedlings 1123

grown in different sets of liquid culture (***, P ≤ 0.001). 1124

(D) Flowering phenotypes of the Ami-m and MRF1 OE lines. Plants were grown for 4 weeks 1125

under long-day conditions. 1126

(E) Quantification of rosette and cauline leaf numbers at the bolting stage with the first open 1127

flower. Values represent means SE of 40 plants per sample (***, P ≤ 0.001). 1128

1129

Figure 4. Phenotypes and Gene Expression of the MRF Ami-m and MRF1 OE Lines after DS 1130

and ReLG treatments. 1131

(A) Seedlings at 12 days after germination were incubated in the dark/starvation (DS) for 5 days, 1132

and then re-illuminated and fed with 30 mM glucose (ReLG) for 5 days. Photos were taken 1133

periodically during the process. d, days. 1134

(B, C) Time-course RT-qPCR analyses of CAB2 mRNA levels. Seedlings grown under LG 1135

conditions were subjected to 2 and 5 days of DS treatment (DS+2 and DS+5), followed by 3 and 1136

5 days of LG treatment (ReLG+3 and ReLG+5). RT-qPCR was performed for CAB2 mRNAs at 1137

the indicated time points (B). The relative CAB2 transcript levels in different lines at ReLG+3 1138

[boxed with dotted line in (B)] were plotted (C). The transcript level is normalized by PP2AA3 1139

mRNA, and expressed relative to those in WT. Values represent the means ± S.E. of N = three 1140

biological replicates of seedlings grown in different sets in liquid culture. Asterisks denote 1141

statistical significance of the differences between WT and the transgenic lines, calculated using 1142

Student’s t-test (*, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001). 1143

39

(D, E) Time-course RT-qPCR analyses of SEN4 mRNA levels. The SEN4 transcript levels at the 1144

indicated time points are shown (D), and the values at DS+5 [boxed with dotted line in (D)] were 1145

plotted (E). 1146

1147

Figure 5. Interactions between MRFs and Eukaryotic Translation Initiation Factor 4A. 1148

(A) Bimolecular fluorescence complementation (BiFC). YFPN- and YFP

C-fusion proteins were 1149

co-expressed in N. benthamiana leaves by agroinfiltration. Leaf epidermal cells were observed 1150

by confocal microscopy. More than 20 leaf cells showing yellow fluorescence were observed for 1151

each BiFC experiment. Bars = 20 µm. 1152

(B) Co-immunoprecipitation. Each MRF protein in Flag fusion (Flag:MRF) was expressed alone 1153

or together with eIF4A-1:Myc in N. benthamiana leaves. Total leaf proteins were 1154

immunoprecipitated with anti-Myc antibody-conjugated resin, and the co-immunoprecipitate was 1155

detected using the anti-Flag antibody. 1156

(C) Yeast two-hybrid assay. GAL4 activation domain (AD)-fused to eIF4A-1 and GAL4 DNA 1157

binding domain (BD)-fused MRF proteins were co-expressed in yeast. Alpha-galactosidase 1158

activity indicates protein-protein interaction affinity. Error bars represent SE from triplicate 1159

biological replications using three individual colonies. Asterisks denote the statistical 1160

significance of the differences between the control (AD:eIF4A-1/BD vector) and other samples 1161

(*, P ≤ 0.05; **, P ≤ 0.01). 1162

1163

Figure 6. Ribosome Association of MRF Proteins and BX04 Complementation Assays. 1164

(A) Co-fractionation of MRFs with ribosome subunits and translation initiation factors. 1165

MRFs:GFP, eIF4E-1:Myc, and eIF4A-1:Myc were expressed in N. benthamiana leaves. After 1166

sucrose density gradient sedimentation, the fractions were subjected to immunoblotting with 1167

anti-GFP, anti-Myc, and anti-60S ribosomal protein L10a (RPL10a) antibodies. Lanes indicated 1168

the fractions from top (15%) to bottom (50%). 1169

(B) Distribution of MRFs in sucrose gradient fractions after RNAse A treatment. Total cell 1170

extract was prepared from Flag-MRF1 OE seedlings (#1) grown under light/glucose conditions. 1171

The cell extract was treated with 1 mg/ml of RNase A on ice for 15 min (+RNAse A) or with 1172

RNase-free water (control), prior to sucrose density gradient sedimentation (15%-50%). The UV 1173

40

absorbance at 254 nm was monitored for gradient fractions to produce the absorbance profiles 1174

(top). The collected fractions were subjected to immunoblotting with anti-Flag and anti-L10a 1175

antibodies (bottom). 1176

(C) BX04 complementation assays. The E. coli BX04 strain is a quadruple mutant of cold-shock 1177

proteins, which cannot grow at low temperature. The BX04 strain was transformed with 1178

plasmids carrying MRFs, E. coli CspA (cold-shock protein; positive control), Arabidopsis LOS4 1179

(RNA helicase), and vector control. The transformants were grown overnight, and then serially 1180

diluted and spotted onto media plates. The plates were incubated at 37 °C (left) and 18 °C (right). 1181

1182

Figure 7. 35

S-Methionine Labeling under Dark/Starvation Conditions, Polysome Analyses, and 1183

Total RNA Contents 1184

(A) Autoradiography images of 35

S-Met incorporation. Seedlings grown in liquid culture were 1185

pre-incubated without glucose in the dark for 30 min, followed by 35

S-Met labeling for 2.5 h 1186

under the same conditions. After SDS-PAGE of protein extracts from the labeled seedlings, the 1187

gel was stained with Coomassie brilliant blue (CBB) and dried. The radioactive signal within the 1188

gel was detected by a phosphorimager. Four independent experiments yielded similar results, and 1189

representative images are shown. 1190

(B) Relative band intensity. The radioactive intensity of 35

S-Met-labeled proteins was normalized 1191

by CBB band intensity, and the ratio was expressed relative to the WT. Error bars represent SE 1192

from four biological replications based on four independent experiments (*, P ≤ 0.05; **, P ≤ 1193

0.01). 1194

(C) Polysome analyses. Seedlings were incubated with LG for 2 h or DS for 1 day (1 d). The 1195

seedlings were treated with cycloheximide (50 μg/ml) for 5 min before harvest, and total cell 1196

extracts from the seedlings were subjected to sucrose density gradient sedimentation (15%-50%). 1197

The UV absorbance at 254 nm was monitored for gradient fractions to produce the absorbance 1198

profiles. The absorbance profiles of LG (blue lines) and DS (red lines) samples were 1199

superimposed for comparison. An average ratio (P/NP) of polysomes to 60S/80S ribosomes was 1200

calculated for each sample using Image J program, from four biological replications based on 1201

four independent experiments. 1202

41

(D) Total RNA contents after prolonged DS and ReLG treatments. Seedlings grown under LG 1203

conditions (control; CTL) were subjected to 2 and 5 days of DS treatment (DS+2 and DS+5), 1204

followed by 3 and 5 days of LG treatment (ReLG+3 and ReLG+5) as described in Figure 3A. 1205

Total RNA was extracted from an equal weight of the seedlings at each stage, and measured by 1206

absorbance at 260 nm using a spectrophotometer. Error bars represent SE from triplicate 1207

biological replications using seedlings grown in different sets of liquid culture. Asterisks denote 1208

statistical significance of the differences between WT and the transgenic lines, calculated using 1209

Student’s t-test (*, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001). 1210

1211

Figure 8. Distribution Patterns and Quantifications of Specific mRNAs in Sucrose Gradient 1212

Fractions. 1213

(A) Distribution of PP2AA3, GAPC, HDA19, and HDC1 mRNAs in sucrose gradient fractions.1214

WT and Ami-m (#3) seedlings were incubated under LG conditions for 2 h (left) or under DS 1215

conditions for 1 d (right). Total cell extracts prepared from the seedlings were subjected to 1216

sucrose density gradient sedimentation (15%-50%), and total 15 fractions were collected from 1217

each tube. Total RNA was extracted from each fraction, followed by cDNA synthesis and RT-1218

qPCR using gene-specific primers. The abundance of mRNA in each fraction was quantified as a 1219

percentage of their total amount in all fractions. Similar results were obtained in three 1220

independent experiments, and a representative result is shown. Error bars represent SE from 1221

three technical replications. 1222

(B) The abundance of mRNA in polysomal (P; fractions 9-15) and nonpolysomal fractions (NP;1223

fractions 1-8), quantified as a percentage of their total amount. Ami3 represents Ami-m (#3). 1224

Error bars represent SE from three biological replications based on three independent 1225

experiments (*, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001). 1226

1227

Figure 9. Phosphorylation and Ribosome Association of MRF1 According to Cellular Energy 1228

Availability. 1229

(A) Phosphorylation of MRF1 in vivo. Total protein extracts from N. benthamiana leaves, which1230

express MRF:Myc proteins, were treated with the lambda phosphatase (λPP). After treatment, 1231

the samples were subjected to Zn2+

-Phostag SDS-PAGE (top) and to normal SDS-PAGE (bottom)1232

42

for immunoblotting with anti-Myc antibody. The phosphorylated form of MRF1 was marked 1233

with the asterisk. 1234

(B) Phosphorylation of MRF1 under different energy conditions. Flag:MRF1 OE (#1) seedlings1235

were incubated under dark/starvation conditions for 1-48 h, and then re-illuminated and fed with 1236

30 mM glucose (light/glucose) for 0.5-2 h. Protein extracts from the seedlings harvested at 1237

different time points were separated by Zn2+

-Phostag SDS-PAGE (top) and by normal SDS-1238

PAGE (middle), followed by immunoblotting with anti-Flag antibody. The Rubisco large subunit 1239

was stained with CBB as loading control (bottom). 1240

(C) Ribosome association of Flag:MRF1 under different energy conditions. Flag:MRF1 OE (#1)1241

seedlings were incubated in the dark/starvation for 24 h, followed by re-illumination and 1242

glucose-feeding for 1 h. Polysome analysis was performed by ultracentrifugation through a 10-1243

55% sucrose gradient. Then the fractions were precipitated and analyzed by immunoblotting with 1244

anti-Flag and anti-RPL10a antibodies. Lanes indicate the fractions from top (10%) to bottom 1245

(55%). Arrowheads indicate the final positions of MRF1 detection. 1246

(D) Co-immunoprecipitation. Flag:MRF1 was expressed alone or together with eIF4A-1:Myc in1247

N. benthamiana leaves. Leaf disks were prepared for treatment with light/glucose (LG) or1248

dark/starvation (DS) for 3 h. Total leaf proteins were immunoprecipitated with anti-Myc 1249

antibody-conjugated resin, and the co-immunoprecipitate was detected using the anti-Flag 1250

antibody. 1251

1252

Figure 10. TOR-Modulated MRF Gene Expression and MRF1 Phosphorylation, and Seedling 1253

Phenotypes upon Torin-1 Treatment. 1254

(A) Altered MRF gene expression in TOR-silenced seedlings in response to starvation and sugar1255

feeding. Estradiol-inducible TOR RNAi seedlings (es-tor1) were treated with ethanol (-ES) or 10 1256

μM estradiol (+ES) for gene silencing. Twelve seedlings grown in three different sets of liquid 1257

culture were incubated in glucose-free medium for 24 h (Stv), and then fed with 30 mM glucose 1258

(Glc), mannitol (Man), and sucrose (Suc) for 4 h. RT-qPCR was performed with gene-specific 1259

primers. Transcript levels are normalized by PP2AA3 mRNA, and expressed relative to those of 1260

Stv samples. Error bars represent SE from triplicate technical replications. 1261

43

(B) MRF1 phosphorylation in TOR-silenced plants. TOR VIGS was performed in Flag:MRF1 1262

OE (#1) lines. Protein extracts from TRV control or TOR VIGS leaves (10 DAI) were separated 1263

by Phostag SDS-PAGE (top) and by normal SDS-PAGE (middle), followed by immunoblotting 1264

with anti-Flag antibody. The Rubisco large subunit was stained with CBB as loading control 1265

(bottom). 1266

(C) Ribosome association of Flag:MRF1 in TOR-silenced plants. TOR VIGS was performed in1267

Flag:MRF1 OE (#1) lines. Protein extracts from TRV control and TOR VIGS leaves (10 DAI) 1268

were fractionated by sucrose density gradient sedimentation (10-55%). The fractions were 1269

precipitated and analyzed by immunoblotting with anti-Flag and anti-RPL10a antibodies. Lanes 1270

indicate the fractions from top (10%) to bottom (55%). Arrowheads indicate the final positions of 1271

MRF1 detection. 1272

(D) Phenotypes of the Ami-m and MRF1 OE seedlings after Torin-1 treatment. Seven-day-old1273

seedlings grown in liquid culture were treated with Torin-1 (2 μM) or control DMSO for 3 days. 1274

(E) Root length of the seedlings was measured after treatment with Torin-1 (2 μM) or control1275

DMSO for 3 and 5 days. Each data point represents the mean SE (n > 14 seedlings). Asterisks 1276

denote statistical significance of the differences between Torin-1-treated samples and DMSO-1277

treated samples, calculated using Student’s t-test (*, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001). 1278

1279

Figure 11. In Vitro Phosphorylation of MRF1 by S6K kinases. 1280

(A) In vitro kinase assay of immunoprecipitated S6K1:Myc with the recombinant MBP:MRF1281

proteins as substrates. After the kinase assay with [γ-32

P]-ATP, SDS-PAGE was performed. 1282

Phosphorylated MBP:MRF proteins were detected by a phosphorimager (top); the MBP:MRF 1283

protein in the reaction was detected by CBB staining (middle); immunoprecipitated S6K1:Myc 1284

was detected by immunoblotting with anti-Myc antibody (bottom). 1285

(B) In vitro kinase assay of immunoprecipitated S6K2:Myc with the recombinant MBP:MRF1286

proteins as substrates. 1287

(C) In vitro kinase assay with S6K1 mutant forms that carry a mutation in the TOR1288

phosphorylation site T449. In vitro kinase assay was performed with S6K1:Myc, 1289

S6K1(T449A):Myc, and S6K1(T449D):Myc proteins. 1290

44

(D) BiFC analyses for MRF1 interactions with S6K1 and S6K2. MRF1:YFPN was expressed 1291

together with S6K1:YFPC or S6K2:YFP

C in N. benthamiana leaves using agroinfiltration. Leaf 1292

epidermal cells were observed by confocal microscopy. More than 20 leaf cells showing yellow 1293

fluorescence were observed for each BiFC experiment. As a negative control, MRF1:YFPN and 1294

eIF4E-1:YFPC were co-expressed in N. benthamiana leaves, which resulted in little yellow 1295

fluorescence. Bars = 20 µm. 1296

(E), (F) Co-immunoprecipitation of MRF1 with S6K1 and S6K2. Flag:MRF1 was expressed 1297

alone or together with S6K1:Myc (E) or S6K2:Myc (F) in N. benthamiana leaves. Total leaf 1298

proteins were immunoprecipitated with anti-Myc antibody-conjugated resin, and the co-1299

immunoprecipitate was detected using the anti-Flag antibody. 1300

1301

45

REFERENCES 1302

1303

Ahn, C.S., Ahn, H.K., and Pai, H.S. (2015). Overexpression of the PP2A regulatory subunit 1304

Tap46 leads to enhanced plant growth through stimulation of the TOR signalling pathway. 1305

J Exp Bot 66, 827-840. 1306

Ahn, C.S., Han, J.A., Lee, H.S., Lee, S., and Pai, H.S. (2011). The PP2A Regulatory Subunit 1307

Tap46, a Component of the TOR Signaling Pathway, Modulates Growth and Metabolism 1308

in Plants. Plant Cell 23, 185-209. 1309

Allen, M.L., Metz, A.M., Timmer, R.T., Rhoads, R.E., and Browning, K.S. (1992). Isolation 1310

and sequence of the cDNAs encoding the subunits of the isozyme form of wheat protein 1311

synthesis initiation factor 4F. J Biol Chem 267, 23232-23236. 1312

Andreou, A.Z., and Klostermeier, D. (2013). The DEAD-box helicase eIF4A: paradigm or the 1313

odd one out? RNA Biol 10, 19-32. 1314

Bolger, T.A., Folkmann, A.W., Tran, E.J., and Wente, S.R. (2008). The mRNA export factor 1315

Gle1 and inositol hexakisphosphate regulate distinct stages of translation. Cell 134, 624-1316

633. 1317

Browning, K.S. (2004). Plant translation initiation factors: it is not easy to be green. Biochem 1318

Soc Trans 32, 589-591. 1319

Browning, K.S., and Bailey-Serres, J. (2015). Mechanism of cytoplasmic mRNA translation. 1320

Arabidopsis Book 13, e0176. 1321

Burch-Smith, T.M., Schiff, M., Liu, Y., and Dinesh-Kumar, S.P. (2006). Efficient virus-1322

induced gene silencing in Arabidopsis. Plant Physiol 142, 21-27. 1323

Caldana, C., Li, Y., Leisse, A., Zhang, Y., Bartholomaeus, L., Fernie, A.R., Willmitzer, L., 1324

and Giavalisco, P. (2013). Systemic analysis of inducible target of rapamycin mutants 1325

reveal a general metabolic switch controlling growth in Arabidopsis thaliana. Plant J 73, 1326

897-909. 1327

Cheng, S., Liu, R., and Gallie, D.R. (2013). The unique evolution of the programmed cell death 1328

4 protein in plants. BMC Evol Biol 13, 199. 1329

Cherkasova, V.A., and Hinnebusch, A.G. (2003). Translational control by TOR and TAP42 1330

through dephosphorylation of eIF2alpha kinase GCN2. Genes Dev 17, 859-872. 1331

Clough, S.J., and Bent, A.F. (1998). Floral dip: a simplified method for Agrobacterium-1332

mediated transformation of Arabidopsis thaliana. Plant J 16, 735-743. 1333

Dennis, M.D., Jefferson, L.S., and Kimball, S.R. (2012). Role of p70S6K1-mediated 1334

phosphorylation of eIF4B and PDCD4 proteins in the regulation of protein synthesis. J 1335

Biol Chem 287, 42890-42899. 1336

Deprost, D., Yao, L., Sormani, R., Moreau, M., Leterreux, G., Nicolai, M., Bedu, M., 1337

Robaglia, C., and Meyer, C. (2007). The Arabidopsis TOR kinase links plant growth, 1338

yield, stress resistance and mRNA translation. EMBO Rep 8, 864-870. 1339

Dobrenel, T., Caldana, C., Hanson, J., Robaglia, C., Vincentz, M., Veit, B., and Meyer, C. 1340

(2016a). TOR Signaling and Nutrient Sensing. Annu Rev Plant Biol 67, 261-285. 1341

Dobrenel, T., Mancera-Martinez, E., Forzani, C., Azzopardi, M., Davanture, M., Moreau, 1342

M., Schepetilnikov, M., Chicher, J., Langella, O., Zivy, M., Robaglia, C., Ryabova, 1343

L.A., Hanson, J., and Meyer, C. (2016b). The Arabidopsis TOR Kinase Specifically 1344

Regulates the Expression of Nuclear Genes Coding for Plastidic Ribosomal Proteins and 1345

46

the Phosphorylation of the Cytosolic Ribosomal Protein S6. Front Plant Sci 7, 1611. 1346

Dorrello, N.V., Peschiaroli, A., Guardavaccaro, D., Colburn, N.H., Sherman, N.E., and 1347

Pagano, M. (2006). S6K1- and betaTRCP-mediated degradation of PDCD4 promotes 1348

protein translation and cell growth. Science 314, 467-471. 1349

Fehler, O., Singh, P., Haas, A., Ulrich, D., Muller, J.P., Ohnheiser, J., and Klempnauer, K.H. 1350

(2014). An evolutionarily conserved interaction of tumor suppressor protein Pdcd4 with 1351

the poly(A)-binding protein contributes to translation suppression by Pdcd4. Nucleic 1352

Acids Res 42, 11107-11118. 1353

Floyd, B.E., Morriss, S.C., MacIntosh, G.C., and Bassham, D.C. (2015). Evidence for 1354

autophagy-dependent pathways of rRNA turnover in Arabidopsis. Autophagy 11, 2199-1355

2212. 1356

Gamm, M., Peviani, A., Honsel, A., Snel, B., Smeekens, S., and Hanson, J. (2014). Increased 1357

sucrose levels mediate selective mRNA translation in Arabidopsis. BMC Plant Biol 14, 1358

306. 1359

Gong, Z., Dong, C.H., Lee, H., Zhu, J., Xiong, L., Gong, D., Stevenson, B., and Zhu, J.K. 1360

(2005). A DEAD box RNA helicase is essential for mRNA export and important for 1361

development and stress responses in Arabidopsis. Plant Cell 17, 256-267. 1362

Hemerly, A., Engler Jde, A., Bergounioux, C., Van Montagu, M., Engler, G., Inze, D., and 1363

Ferreira, P. (1995). Dominant negative mutants of the Cdc2 kinase uncouple cell 1364

division from iterative plant development. EMBO J 14, 3925-3936. 1365

Holcik, M., and Sonenberg, N. (2005). Translational control in stress and apoptosis. Nat Rev 1366

Mol Cell Biol 6, 318-327. 1367

Hutchins, A.P., Roberts, G.R., Lloyd, C.W., and Doonan, J.H. (2004). In vivo interaction 1368

between CDKA and eIF4A: a possible mechanism linking translation and cell 1369

proliferation. FEBS Lett 556, 91-94. 1370

Jackson, R.J., Hellen, C.U., and Pestova, T.V. (2010). The mechanism of eukaryotic translation 1371

initiation and principles of its regulation. Nat Rev Mol Cell Biol 11, 113-127. 1372

Jankowsky, E., Gross, C.H., Shuman, S., and Pyle, A.M. (2001). Active disruption of an 1373

RNA-protein interaction by a DExH/D RNA helicase. Science 291, 121-125. 1374

Jao, D.L., and Chen, K.Y. (2006). Tandem affinity purification revealed the hypusine-1375

dependent binding of eukaryotic initiation factor 5A to the translating 80S ribosomal 1376

complex. J Cell Biochem 97, 583-598. 1377

Juntawong, P., and Bailey-Serres, J. (2012). Dynamic Light Regulation of Translation Status in 1378

Arabidopsis thaliana. Front Plant Sci 3, 66. 1379

Juntawong, P., Girke, T., Bazin, J., and Bailey-Serres, J. (2014). Translational dynamics 1380

revealed by genome-wide profiling of ribosome footprints in Arabidopsis. Proc Natl Acad 1381

Sci U S A 111, E203-212. 1382

Kang, H., Park, S.J., and Kwak, K.J. (2013). Plant RNA chaperones in stress response. Trends 1383

Plant Sci 18, 100-106. 1384

Kim, T.H., Kim, B.H., Yahalom, A., Chamovitz, D.A., and von Arnim, A.G. (2004). 1385

Translational regulation via 5' mRNA leader sequences revealed by mutational analysis of 1386

the Arabidopsis translation initiation factor subunit eIF3h. Plant Cell 16, 3341-3356. 1387

Kinoshita, E., and Kinoshita-Kikuta, E. (2011). Improved Phos-tag SDS-PAGE under neutral 1388

pH conditions for advanced protein phosphorylation profiling. Proteomics 11, 319-323. 1389

Koroleva, O.A., Calder, G., Pendle, A.F., Kim, S.H., Lewandowska, D., Simpson, C.G., 1390

47

Jones, I.M., Brown, J.W., and Shaw, P.J. (2009). Dynamic behavior of Arabidopsis 1391

eIF4A-III, putative core protein of exon junction complex: fast relocation to nucleolus 1392

and splicing speckles under hypoxia. Plant Cell 21, 1592-1606. 1393

Lageix, S., Lanet, E., Pouch-Pelissier, M.N., Espagnol, M.C., Robaglia, C., Deragon, J.M., 1394

and Pelissier, T. (2008). Arabidopsis eIF2alpha kinase GCN2 is essential for growth in 1395

stress conditions and is activated by wounding. BMC Plant Biol 8, 134. 1396

Lankat-Buttgereit, B., and Goke, R. (2009). The tumour suppressor Pdcd4: recent advances in 1397

the elucidation of function and regulation. Biol Cell 101, 309-317. 1398

Lastdrager, J., Hanson, J., and Smeekens, S. (2014). Sugar signals and the control of plant 1399

growth and development. J Exp Bot 65, 799-807. 1400

Lee, H.S., Lee, D.H., Cho, H.K., Kim, S.H., Auh, J.H., and Pai, H.S. (2015). InsP6-Sensitive 1401

Variants of the Gle1 mRNA Export Factor Rescue Growth and Fertility Defects of the 1402

ipk1 Low-Phytic-Acid Mutation in Arabidopsis. Plant Cell. 1403

Lei, G., Shen, M., Li, Z.G., Zhang, B., Duan, K.X., Wang, N., Cao, Y.R., Zhang, W.K., Ma, 1404

B., Ling, H.Q., Chen, S.Y., and Zhang, J.S. (2011). EIN2 regulates salt stress response 1405

and interacts with a MA3 domain-containing protein ECIP1 in Arabidopsis. Plant Cell 1406

Environ 34, 1678-1692. 1407

Li, W., Ma, M., Feng, Y., Li, H., Wang, Y., Ma, Y., Li, M., An, F., and Guo, H. (2015). EIN2-1408

directed translational regulation of ethylene signaling in Arabidopsis. Cell 163, 670-683. 1409

Liu, M.J., Wu, S.H., Chen, H.M., and Wu, S.H. (2012). Widespread translational control 1410

contributes to the regulation of Arabidopsis photomorphogenesis. Mol Syst Biol 8, 566. 1411

Loh, P.G., Yang, H.S., Walsh, M.A., Wang, Q., Wang, X., Cheng, Z., Liu, D., and Song, H. 1412

(2009). Structural basis for translational inhibition by the tumour suppressor Pdcd4. 1413

EMBO J 28, 274-285. 1414

Ma, X.M., and Blenis, J. (2009). Molecular mechanisms of mTOR-mediated translational 1415

control. Nat Rev Mol Cell Biol 10, 307-318. 1416

Ma, Y., Miura, E., Ham, B.K., Cheng, H.W., Lee, Y.J., and Lucas, W.J. (2010). Pumpkin 1417

eIF5A isoforms interact with components of the translational machinery in the cucurbit 1418

sieve tube system. Plant J 64, 536-550. 1419

Magnuson, B., Ekim, B., and Fingar, D.C. (2012). Regulation and function of ribosomal 1420

protein S6 kinase (S6K) within mTOR signalling networks. Biochem J 441, 1-21. 1421

Mahfouz, M.M., Kim, S., Delauney, A.J., and Verma, D.P. (2006). Arabidopsis TARGET OF 1422

RAPAMYCIN interacts with RAPTOR, which regulates the activity of S6 kinase in 1423

response to osmotic stress signals. Plant Cell 18, 477-490. 1424

Mair, A., Pedrotti, L., Wurzinger, B., Anrather, D., Simeunovic, A., Weiste, C., Valerio, C., 1425

Dietrich, K., Kirchler, T., Nagele, T., Vicente Carbajosa, J., Hanson, J., Baena-1426

Gonzalez, E., Chaban, C., Weckwerth, W., Droge-Laser, W., and Teige, M. (2015). 1427

SnRK1-triggered switch of bZIP63 dimerization mediates the low-energy response in 1428

plants. Elife 4. 1429

Mehdi, S., Derkacheva, M., Ramstrom, M., Kralemann, L., Bergquist, J., and Hennig, L. 1430

(2016). The WD40 Domain Protein MSI1 Functions in a Histone Deacetylase Complex 1431

to Fine-Tune Abscisic Acid Signaling. Plant Cell 28, 42-54. 1432

Merchante, C., Stepanova, A.N., and Alonso, J.M. (2017). Translation regulation in plants: an 1433

interesting past, an exciting present and a promising future. Plant J 90, 628-653. 1434

Merchante, C., Brumos, J., Yun, J., Hu, Q., Spencer, K.R., Enriquez, P., Binder, B.M., 1435

48

Heber, S., Stepanova, A.N., and Alonso, J.M. (2015). Gene-specific translation 1436

regulation mediated by the hormone-signaling molecule EIN2. Cell 163, 684-697. 1437

Moreau, M., Azzopardi, M., Clement, G., Dobrenel, T., Marchive, C., Renne, C., Martin-1438

Magniette, M.L., Taconnat, L., Renou, J.P., Robaglia, C., and Meyer, C. (2012). 1439

Mutations in the Arabidopsis homolog of LST8/GbetaL, a partner of the target of 1440

Rapamycin kinase, impair plant growth, flowering, and metabolic adaptation to long days. 1441

Plant Cell 24, 463-481. 1442

Muench, D.G., Zhang, C., and Dahodwala, M. (2012). Control of cytoplasmic translation in 1443

plants. Wiley Interdiscip Rev RNA 3, 178-194. 1444

Nakaminami, K., Karlson, D.T., and Imai, R. (2006). Functional conservation of cold shock 1445

domains in bacteria and higher plants. Proc Natl Acad Sci U S A 103, 10122-10127. 1446

Nicolai, M., Roncato, M.A., Canoy, A.S., Rouquie, D., Sarda, X., Freyssinet, G., and 1447

Robaglia, C. (2006). Large-scale analysis of mRNA translation states during sucrose 1448

starvation in arabidopsis cells identifies cell proliferation and chromatin structure as 1449

targets of translational control. Plant Physiol 141, 663-673. 1450

Nukarinen, E., Nagele, T., Pedrotti, L., Wurzinger, B., Mair, A., Landgraf, R., Bornke, F., 1451

Hanson, J., Teige, M., Baena-Gonzalez, E., Droge-Laser, W., and Weckwerth, W. 1452

(2016). Quantitative phosphoproteomics reveals the role of the AMPK plant ortholog 1453

SnRK1 as a metabolic master regulator under energy deprivation. Sci Rep 6, 31697. 1454

Pal, S.K., Liput, M., Piques, M., Ishihara, H., Obata, T., Martins, M.C., Sulpice, R., van 1455

Dongen, J.T., Fernie, A.R., Yadav, U.P., Lunn, J.E., Usadel, B., and Stitt, M. (2013). 1456

Diurnal changes of polysome loading track sucrose content in the rosette of wild-type 1457

arabidopsis and the starchless pgm mutant. Plant Physiol 162, 1246-1265. 1458

Patrick, R.M., and Browning, K.S. (2012). The eIF4F and eIFiso4F Complexes of Plants: An 1459

Evolutionary Perspective. Comp Funct Genomics 2012, 287814. 1460

Perrella, G., Lopez-Vernaza, M.A., Carr, C., Sani, E., Gossele, V., Verduyn, C., Kellermeier, 1461

F., Hannah, M.A., and Amtmann, A. (2013). Histone deacetylase complex1 expression 1462

level titrates plant growth and abscisic acid sensitivity in Arabidopsis. Plant Cell 25, 1463

3491-3505. 1464

Pestov, D.G., and Shcherbik, N. (2012). Rapid cytoplasmic turnover of yeast ribosomes in 1465

response to rapamycin inhibition of TOR. Mol Cell Biol 32, 2135-2144. 1466

Piir, K., Paier, A., Liiv, A., Tenson, T., and Maivali, U. (2011). Ribosome degradation in 1467

growing bacteria. EMBO Rep 12, 458-462. 1468

Porra, R.J., and Scheer, H. (2000). (18)O and mass spectrometry in chlorophyll research: 1469

Derivation and loss of oxygen atoms at the periphery of the chlorophyll macrocycle 1470

during biosynthesis, degradation and adaptation. Photosynth Res 66, 159-175. 1471

Powers, T., and Walter, P. (1999). Regulation of ribosome biogenesis by the rapamycin-1472

sensitive TOR-signaling pathway in Saccharomyces cerevisiae. Mol Biol Cell 10, 987-1473

1000. 1474

Ren, M., Venglat, P., Qiu, S., Feng, L., Cao, Y., Wang, E., Xiang, D., Wang, J., Alexander, D., 1475

Chalivendra, S., Logan, D., Mattoo, A., Selvaraj, G., and Datla, R. (2012). Target of 1476

rapamycin signaling regulates metabolism, growth, and life span in Arabidopsis. Plant 1477

Cell 24, 4850-4874. 1478

Saini, P., Eyler, D.E., Green, R., and Dever, T.E. (2009). Hypusine-containing protein eIF5A 1479

promotes translation elongation. Nature 459, 118-121. 1480

49

Schepetilnikov, M., Dimitrova, M., Mancera-Martinez, E., Geldreich, A., Keller, M., and 1481

Ryabova, L.A. (2013). TOR and S6K1 promote translation reinitiation of uORF-1482

containing mRNAs via phosphorylation of eIF3h. EMBO J 32, 1087-1102. 1483

Schwab, R., Ossowski, S., Riester, M., Warthmann, N., and Weigel, D. (2006). Highly 1484

specific gene silencing by artificial microRNAs in Arabidopsis. Plant Cell 18, 1121-1133. 1485

Sengupta, S., Peterson, T.R., and Sabatini, D.M. (2010). Regulation of the mTOR complex 1 1486

pathway by nutrients, growth factors, and stress. Mol Cell 40, 310-322. 1487

Shaikhin, S.M., Smailov, S.K., Lee, A.V., Kozhanov, E.V., and Iskakov, B.K. (1992). 1488

Interaction of wheat germ translation initiation factor 2 with GDP and GTP. Biochimie 74, 1489

447-454. 1490

Silvera, D., Formenti, S.C., and Schneider, R.J. (2010). Translational control in cancer. Nat 1491

Rev Cancer 10, 254-266. 1492

Sonenberg, N., and Hinnebusch, A.G. (2009). Regulation of translation initiation in eukaryotes: 1493

mechanisms and biological targets. Cell 136, 731-745. 1494

Tome, F., Nagele, T., Adamo, M., Garg, A., Marco-Llorca, C., Nukarinen, E., Pedrotti, L., 1495

Peviani, A., Simeunovic, A., Tatkiewicz, A., Tomar, M., and Gamm, M. (2014). The 1496

low energy signaling network. Front Plant Sci 5, 353. 1497

Turck, F., Zilbermann, F., Kozma, S.C., Thomas, G., and Nagy, F. (2004). Phytohormones 1498

participate in an S6 kinase signal transduction pathway in Arabidopsis. Plant Physiol 134, 1499

1527-1535. 1500

Voinnet, O., Rivas, S., Mestre, P., and Baulcombe, D. (2003). An enhanced transient 1501

expression system in plants based on suppression of gene silencing by the p19 protein of 1502

tomato bushy stunt virus. Plant J 33, 949-956. 1503

Walter, M., Chaban, C., Schutze, K., Batistic, O., Weckermann, K., Nake, C., Blazevic, D., 1504

Grefen, C., Schumacher, K., Oecking, C., Harter, K., and Kudla, J. (2004). 1505

Visualization of protein interactions in living plant cells using bimolecular fluorescence 1506

complementation. Plant J 40, 428-438. 1507

Wang, L., Li, H., Zhao, C., Li, S., Kong, L., Wu, W., Kong, W., Liu, Y., Wei, Y., Zhu, J.K., 1508

and Zhang, H. (2017). The inhibition of protein translation mediated by AtGCN1 is 1509

essential for cold tolerance in Arabidopsis thaliana. Plant Cell Environ 40, 56-68. 1510

Wedeken, L., Singh, P., and Klempnauer, K.H. (2011). Tumor suppressor protein Pdcd4 1511

inhibits translation of p53 mRNA. J Biol Chem 286, 42855-42862. 1512

Wullschleger, S., Loewith, R., and Hall, M.N. (2006). TOR signaling in growth and 1513

metabolism. Cell 124, 471-484. 1514

Xia, B., Ke, H., and Inouye, M. (2001). Acquirement of cold sensitivity by quadruple deletion 1515

of the cspA family and its suppression by PNPase S1 domain in Escherichia coli. Mol 1516

Microbiol 40, 179-188. 1517

Xiong, Y., and Sheen, J. (2012). Rapamycin and glucose-target of rapamycin (TOR) protein 1518

signaling in plants. J Biol Chem 287, 2836-2842. 1519

Xiong, Y., McCormack, M., Li, L., Hall, Q., Xiang, C., and Sheen, J. (2013). Glucose-TOR 1520

signalling reprograms the transcriptome and activates meristems. Nature 496, 181-186. 1521

Yang, H.S., Cho, M.H., Zakowicz, H., Hegamyer, G., Sonenberg, N., and Colburn, N.H. 1522

(2004). A novel function of the MA-3 domains in transformation and translation 1523

suppressor Pdcd4 is essential for its binding to eukaryotic translation initiation factor 4A. 1524

Mol Cell Biol 24, 3894-3906. 1525

50

Yang, H.S., Jansen, A.P., Komar, A.A., Zheng, X., Merrick, W.C., Costes, S., Lockett, S.J., 1526

Sonenberg, N., and Colburn, N.H. (2003). The transformation suppressor Pdcd4 is a 1527

novel eukaryotic translation initiation factor 4A binding protein that inhibits translation. 1528

Mol Cell Biol 23, 26-37. 1529

Zhou, F., Roy, B., and von Arnim, A.G. (2010). Translation reinitiation and development are 1530

compromised in similar ways by mutations in translation initiation factor eIF3h and the 1531

ribosomal protein RPL24. BMC Plant Biol 10, 193. 1532

Zundel, M.A., Basturea, G.N., and Deutscher, M.P. (2009). Initiation of ribosome degradation 1533

during starvation in Escherichia coli. RNA 15, 977-983. 1534

1535

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

RL CL St R B F

Rela

tive

mR

NA

le

vels

0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

RL CL St R B F

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

RL CL St R B F

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

RL CL St R B F

B MRF1 MRF2 MRF3 MRF4

Figure 1

0

5

10

15

20

25

0 0.5 2 4 24

Re

lative m

RN

A le

vels

0

1

2

3

4

5

0 0.5 2 4 24

0

2

4

6

8

10

12

14

16

0 0.5 2 4 24

0

1

2

3

4

5

6

0 0.5 2 4 24

Dark (h)

C MRF1 MRF2 MRF3 MRF4

Dark (h) Dark (h) Dark (h)

A MA3 MA3 MA3 MA3

MA3 MA3 MA3 MA3

MA3 MA3 MA3 MA3

MA3 MA3 MA3 MA3

57 168 222 332 352 462 624

633 aa

516

702 aa

693 aa

702 aa

117 228 281 392 415 525 579 688

91 202 255 366 390 500 561 672

123 234 288 398 421 531 585 692

MRF1

MRF2

MRF3

MRF4

D

Glucose (mM)

MRF1 MRF2 MRF3 MRF4

Glucose (mM) Glucose (mM) Glucose (mM)

0

2

4

6

8

10

12

14

BS S 10 30 60 180

Rela

tive m

RN

A levle

ls

0

2

4

6

8

10

12

14

BS S 10 30 60 180

0

5

10

15

20

25

30

35

BS S 10 30 60 180

0

5

10

15

20

25

BS S 10 30 60 180

Figure 1. Predicted Protein Structure and Expression Patterns of Four MRF Genes.

(A) Schematic representation of four Arabidopsis MRF proteins with four MA3 domains arranged

in tandem. Residue numbers are marked. aa, amino acids.

(B) RT-qPCR analyses of MRF gene expression in plant organs. Three organ pieces were collected

from three different 6-week-old Arabidopsis plants: the 7th and 8th rosette leaves (RL), the 1st cauline

leaves (CL), stems (~1 cm from the bottom; St), and the primary roots (R). Ten pieces of buds

(stages 11-12; B) and open flowers (F) were also collected from the three plants. The collected

tissues were combined for RNA extraction and RT-qPCR. Transcript levels are expressed relative to

those in rosette leaf (RL). For (B) to (D), transcript levels are normalized by PP2AA3 mRNA, and

error bars represent standard errors (SE) calculated from triplicate technical replications.

(C) RT-qPCR analyses of MRF gene expression in response to darkness. Twelve seedlings grown in

three different sets of liquid culture were incubated in the dark for the indicated times. Transcript

levels are expressed relative to those at 0 h.

(D) RT-qPCR analyses of MRF gene expression in response to starvation and glucose feeding.

Twelve seedlings grown in three different sets of liquid culture were incubated in glucose-free

medium for 24 h (starvation; S), and then fed with the indicated concentrations of glucose for 4 h.

Transcript levels are expressed relative to those before starvation (BS).

100

140

10

15

kDa

MRF1:GFP

MRF2:GFP

MRF3:GFP

MRF4:GFP

GFP Histone

H2B:mRFP Chlorophyll Merged

A

B

C

α-GFP

α-H3

Total C N

MRF1:GFP

Total C N

MRF2:GFP

Total C N

MRF3:GFP

Total C N

MRF4:GFP

MRF1:GFP MRF2:GFP MRF3:GFP MRF4:GFP

Tra

nsg

enic

lin

es

Figure 2

Figure 2. Cytosolic Localization of MRF Proteins.

(A) Subcellular localization of MRF:GFP fusion proteins in leaf protoplasts. Each MRF:GFP was

transiently expressed together with histone H2B:mRFP as a nuclear marker in N. benthamiana

leaves via agro-infiltration. Protoplasts were prepared from the infiltrated leaves, and observed by

confocal microscopy. Chlorophyll autofluorescence was pseudo-colored blue. More than 20 cells

showing green fluorescence were observed for each construct. Scale bars = 10 µm.

(B) Subcellular fractionation. N. benthamiana leaf extracts expressing MRF:GFP proteins were

fractionated and subjected to SDS-PAGE using 10-15% gradient gel, followed by immunoblotting

with anti-GFP antibody. Total (T), nuclear (N), and cytosolic (C) fractions were indicated. Histone

H3 was detected as a nuclear marker protein using anti-H3 antibody. Two independent experiments

yielded similar results.

(C) Confocal microscopy of GFP fluorescence in epidermal cells of the Arabidopsis transgenic

plants that express each MRF gene under the CaMV35S promoter. Multiple independent transgenic

lines were analyzed for each MRF gene, which similarly suggested cytosolic localization of MRF

proteins. More than three independent observations were made for each transgenic line. Scale bars =

10 µm.

D WT

Ami-m

#3

Ami-m

#10

MRF1

OE #2

MRF1

OE #1

A

MRF1 amiRNA lines

(a, b, c)

MRF2 amiRNA lines

(d, e, f)

Flag:MRF1

overexpression lines

(OE #1, #2)

0

0.2

0.4

0.6

0.8

1.0

1.2

WT Ami-m#3

Ami-m#10

Ami-m#14

Rela

tive

mR

NA

le

vels

(M

RF

1 o

r M

RF

3/P

P2

AA

3) MRF1 MRF3

B

2256 2915 3009 2451

m

MRF1 (At5G63190.1)

1733 2278 2651

MRF2 (At1G22730.1)

MRF4 (At3G48390.1)

2398

MRF3 (At4G24800.1)

m

1547

Figure 3

Transgenic lines

MRF1, MRF3, and MRF4

multi-target amiRNA lines

(Ami-m #3, #10, #14)

m

Target sites for artificial miRNAs

m b a c

d e f

0

2

4

6

8

10

12

14

16

18

WT Ami-m#3

Ami-m#10

MRF1OE#1

MRF1OE#2

Leaf num

ber

(Rosett

e+

Caulin

e) Rosette CaulineE

0

5

10

15

20

25

30

35

40

WT MRF1OE #1

MRF1OE #2

Rela

tive m

RN

A levels

(M

RF

1/P

P2A

A3

)

C

Figure 3. Generation of MRF Artificial miRNA and MRF1 Overexpression Lines, and Analysis of

their Flowering Phenotypes.

(A) Description of MRF artificial miRNA (amiRNA) and MRF1 overexpression (OE) lines (left),

and target sites for the amiRNA lines (right). The target sites (arrowheads) were designed for

silencing of MRF1, MRF2, or multiple genes (MRF1, MRF3, and MRF4). “Ami-m” represents

amiRNA lines with multiple targets.

(B) RT-qPCR to determine MRF1 and MRF3 mRNA levels in the Ami-m lines. Transcript levels in

the Ami-m lines are expressed relative to those in the WT. Values represent the means ± S.E. of N =

three biological replicates of 10-day-old seedlings grown in different sets of liquid culture. Asterisks

denote statistical significance of the differences between the WT and the transgenic lines, calculated

using Student’s t-test (***, P ≤ 0.001).

(C) RT-qPCR to determine MRF1 mRNA levels in the MRF1 OE lines, compared with those in WT.

Error bars represent SE from triplicate biological replications using 10-day-old seedlings grown in

different sets of liquid culture (***, P ≤ 0.001).

(D) Flowering phenotypes of the Ami-m and MRF1 OE lines. Plants were grown for 4 weeks under

long-day conditions.

(E) Quantification of rosette and cauline leaf numbers at the bolting stage with the first open flower. Values represent

means SE of 40 plants per sample (***, P ≤ 0.001).

WT Ami-m (#3) Ami-m (#10) MRF1 OE (#1) MRF1 OE (#2) Ami-m (#14)

Da

rk+

Sta

rvation

(DS

)

+2d

Lig

ht+

Glu

cose

(Re

LG

)

+3d

+3d

+2d

A Figure 4

Seedlings

(CTL)

CAB2 (at ReLG+3)

SEN4 (at DS+5) E

C B

D Ami-m

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

CTL DS+2 DS+5 ReLG+3 ReLG+5

Rela

tive m

RN

A levels

(C

AB

2/P

P2A

A3)

WT #3 #10 #14 OE1 OE2

Ami-m

0

5

10

15

20

25

30

35

40

45

CTL DS+2 DS+5 ReLG+3 ReLG+5

Rela

tive m

RN

A levels

(S

EN

4/P

P2A

A3)

WT #3 #10 #14 OE1 OE2

Ami-m

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

WT #3 #10 #14 OE1 OE2

Rela

tive m

RN

A levels

(C

AB

2/P

P2A

A3)

Ami-m

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

WT #3 #10 #14 OE1 OE2

Rela

tive m

RN

A levels

(S

EN

4/P

P2A

A3)

CAB2

SEN4

Figure 4. Phenotypes and Gene Expression of the MRF Ami-m and MRF1 OE Lines after DS and

ReLG treatments.

(A) Seedlings at 12 days after germination were incubated in the dark/starvation (DS) for 5 days,

and then re-illuminated and fed with 30 mM glucose (ReLG) for 5 days. Photos were taken

periodically during the process. d, days.

(B, C) Time-course RT-qPCR analyses of CAB2 mRNA levels. Seedlings grown under LG

conditions were subjected to 2 and 5 days of DS treatment (DS+2 and DS+5), followed by 3 and 5

days of LG treatment (ReLG+3 and ReLG+5). RT-qPCR was performed for CAB2 mRNAs at the

indicated time points (B). The relative CAB2 transcript levels in different lines at ReLG+3 [boxed

with dotted line in (B)] were plotted (C). The transcript level is normalized by PP2AA3 mRNA, and

expressed relative to those in WT. Values represent the means ± S.E. of N = three biological

replicates of seedlings grown in different sets of liquid culture. Asterisks denote statistical

significance of the differences between WT and the transgenic lines, calculated using Student’s t-

test (*, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001).

(D, E) Time-course RT-qPCR analyses of SEN4 mRNA levels. The SEN4 transcript levels at the

indicated time points are shown (D), and the values at DS+5 [boxed with dotted line in (D)] were

plotted (E).

A B Flag:MRF1

eIF4A-1:Myc

IB : α-Flag

IB : α-Myc

INPUT IP : α-Myc

Flag:MRF3

eIF4A-1:Myc

IB : α-Flag

IB : α-Myc

INPUT IP : α-Myc

Flag:MRF2

eIF4A-1:Myc

IB : α-Flag

IB : α-Myc

INPUT IP : α-Myc

C

Figure 5

eIF4A-1:YFPC

YF

PN:G

le1

eIF4E-1:YFPC

MR

F1:Y

FP

NM

RF

2:Y

FP

NM

RF

3:Y

FP

NM

RF

4:Y

FP

N

YFPN / YFPC

0

10

20

30

40

50

60

70

80

90

100

AD:eIF4A-1BD

AD:eIF4A-1BD:MRF1

AD:eIF4A-1BD:MRF2

AD:eIF4A-1BD:MRF3

AD:eIF4A-1BD:MRF4

α-G

ala

cto

sid

ase a

ctivity (

un

it)

Flag:MRF4

eIF4A-1:Myc

IB : α-Flag

IB : α-Myc

INPUT IP : α-Myc

kDa

75

100

kDa

75

75

60

75

60

100

kDa

75

75

60

100

kDa

75

75

60

+ +

− ++ +

− +

+ +

− ++ +

− +

+ +

− ++ +

− +

+ +

− ++ +

− +

Figure 5. Interactions between MRFs and Eukaryotic Translation Initiation Factor 4A.

(A) Bimolecular fluorescence complementation (BiFC). YFPN- and YFPC-fusion proteins were co-

expressed in N. benthamiana leaves by agroinfiltration. Leaf epidermal cells were observed by

confocal microscopy. More than 20 leaf cells showing yellow fluorescence were observed for each

BiFC experiment. Bars = 20 µm.

(B) Co-immunoprecipitation. Each MRF protein in Flag fusion (Flag:MRF) was expressed alone or

together with eIF4A-1:Myc in N. benthamiana leaves. Total leaf proteins were immunoprecipitated

with anti-Myc antibody-conjugated resin, and the co-immunoprecipitate was detected using the anti-

Flag antibody.

(C) Yeast two-hybrid assay. GAL4 activation domain (AD)-fused to eIF4A-1 and GAL4 DNA

binding domain (BD)-fused MRF proteins were co-expressed in yeast. Alpha-galactosidase activity

indicates protein-protein interaction affinity. Error bars represent SE from triplicate biological

replications using three individual colonies. Asterisks denote the statistical significance of the

differences between the control (AD:eIF4A-1/BD vector) and other samples (*, P ≤ 0.05; **, P ≤

0.01).

Figure 6

15% 50%

A Sucrose gradient

MRF4:GFP

eIF4A-1:Myc

eIF4E-1:Myc

MRF3:GFP

MRF1:GFP

MRF2:GFP α-GFP

α-Myc

RPL10a α-L10a

60S/80S Polysomes

C

CspA

Vector

MRF1

MRF2

MRF3

MRF4

LOS4

37oC / 0.1 mM IPTG

10-1 10-2 10-3 10-4 10-5

10-1 10-2 10-3 10-4 10-5

18oC / 0.1 mM IPTG

E. coli BX04 mutant strain

CspA

Vector

MRF1

MRF2

MRF3

MRF4

LOS4

B

Flag:MRF1

(α-Flag)

α-L10a α-L10a

Flag:MRF1

(α-Flag)

2 3 4 5 6 7 8 9 10 11 12 13 14 2 3 4 5 6 7 8 9 10 11 12 13 14

0.080

0.100

0.120

0.140

0.160

0.080

0.100

0.120

0.140

0.160

15% 50% 15% 50%

Control (+) RNase A

OD

25

4

OD

25

4

polysome

60S 80S

60S/80S

100

kDa

100

100

100

75

45

35

75

60

35

25

100

75

kDa kDa

35

25

35

25

100

75

Figure 6. Co-Sedimentation of MRFs with Ribosomes and BX04 Complementation Assays.

(A) Co-fractionation of MRFs with ribosome subunits and translation initiation factors. MRFs:GFP,

eIF4E-1:Myc, and eIF4A-1:Myc were expressed in N. benthamiana leaves. After sucrose density

gradient sedimentation, the fractions were subjected to immunoblotting with anti-GFP, anti-Myc,

and anti-60S ribosomal protein L10a (RPL10a) antibodies. Lanes indicated the fractions from top

(15%) to bottom (50%).

(B) Distribution of MRFs in sucrose gradient fractions after RNAse A treatment. Total cell extract

was prepared from Flag-MRF1 OE seedlings (#1) grown under light/glucose conditions. The cell

extract was treated with 1 mg/ml of RNase A on ice for 15 min (+RNAse A) or with RNase-free

water (control), prior to sucrose density gradient sedimentation (15%-50%). The UV absorbance at

254 nm was monitored for gradient fractions to produce the absorbance profiles (top). The collected

fractions were subjected to immunoblotting with anti-Flag and anti-L10a antibodies (bottom).

(C) BX04 complementation assays. The E.coli BX04 strain is a quadruple mutant of cold-shock

proteins, which cannot grow at low temperature. The BX04 strain was transformed with plasmids

carrying MRFs, E. coli CspA (cold-shock protein; positive control), Arabidopsis LOS4 (RNA

helicase), and vector control. The transformants were grown overnight, and then serially diluted and

spotted onto media plates. The plates were incubated at 37 °C (left) and 18 °C (right).

Figure 7

Ami-m

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

WT #3 #10 #14 OE1 OE2

Rela

tive b

and

in

ten

sity

(Auto

rad

iogra

ph/C

BB

)

B

C

0.075

0.090

0.105

0.120

0.135

LG (2 h) DS (1 d)

40S

60S/80S

60S 80S

P/NP = 1.3±0.1

P/NP = 4.0±0.1

WT

0.075

0.090

0.105

0.120

0.135

LG (2 h) DS (1 d)

P/NP = 4.1±0.3

P/NP = 1.6±0.1

MRF Ami-m (#3)

0.075

0.090

0.105

0.120

0.135

LG (2 h) DS (1 d)

P/NP = 1.1±0.1

P/NP = 3.9±0.3

MRF1 OE (#1)

OD

25

4

polysome

15% 50% 15% 50% 15% 50%

A Dark / Starvation (3 h)

D

0

20

40

60

80

100

120

140

160

CTL DS+2 DS+5 ReLG+3 ReLG+5

Tota

l R

NA

co

nte

nts

g/g

FW

)

WT Ami-m #3 Ami-m #10 Ami-m #14 OE #1 OE #2

***

**

* *

*

*

*

*

** **

** **

Dark / Starvation (3 h)

35S

-Meth

ionin

e

Auto

rad

iogra

ph

C

BB

25

35

45

60

75

100

25

35

45

60

75

100

kDa

Figure 7. 35S-Methionine Labeling under Dark/Starvation Conditions, Polysome Analyses, and

Total RNA Contents

(A) Autoradiography images of 35S-Met incorporation. Seedlings grown in liquid culture were pre-

incubated without glucose in the dark for 30 min, followed by 35S-Met labeling for 2.5 h under the

same conditions. After SDS-PAGE of protein extracts from the labeled seedlings, the gel was

stained with Coomassie brilliant blue (CBB) and dried. The radioactive signal within the gel was

detected by a phosphorimager.

(B) Relative band intensity. The radioactive intensity of 35S-Met-labeled proteins was normalized by

CBB band intensity, and the ratio was expressed relative to the WT. Error bars represent SE from

four biological replications based on four independent experiments (*, P ≤ 0.05; **, P ≤ 0.01).

(C) Polysome analyses. Seedlings were incubated with LG for 2 h or DS for 1 day (1 d). The

seedlings were treated with cycloheximide (50 μg/ml) for 5 min before harvest, and total cell

extracts from the seedlings were subjected to sucrose density gradient sedimentation (15%-50%).

The UV absorbance at 254 nm was monitored for gradient fractions to produce the absorbance

profiles. The absorbance profiles of LG (blue lines) and DS (red lines) samples were superimposed

for comparison. An average ratio (P/NP) of polysomes to 60S/80S ribosomes was calculated for

each sample using Image J program, from four biological replications based on four independent

experiments.

(D) Total RNA contents after prolonged DS and ReLG treatments. Seedlings grown under LG

conditions (control; CTL) were subjected to 2 and 5 days of DS treatment (DS+2 and DS+5),

followed by 3 and 5 days of LG treatment (ReLG+3 and ReLG+5) as described in Figure 3A. Total

RNA was extracted from an equal weight of the seedlings at each stage, and measured by

absorbance at 260 nm using a spectrophotometer. Error bars represent SE from triplicate biological

replications using seedlings grown in different sets of liquid culture. Asterisks denote statistical

significance of the differences between WT and the transgenic lines, calculated using Student’s t-

test (*, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001).

PP2AA3 GAPC HDA19 HDC1

WT(LG)

Ami3(LG)

WT(DS)

Ami3(DS)

WT(LG)

Ami3(LG)

WT(DS)

Ami3(DS)

WT(LG)

Ami3(LG)

WT(DS)

Ami3(DS)

WT(LG)

Ami3(LG)

WT(DS)

Ami3(DS)

B

WT

Dark / Starvation (1 d) Light / Glucose (2 h)

WT Ami-m (#3) PP2AA3

1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

GAPC

1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

HDA19

HDC1 1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

1 3 5 7 9 11 13 15 1 3 5 7 9 11 13 15

A Figure 8

Ami-m (#3)

Polysome (P) Nonpolysome (NP)

*** **

*

% m

RN

A

% m

RN

A

% m

RN

A

% m

RN

A

% o

f to

tal

25

20

15

10

5

0

25

20

15

10

5

0

25

20

15

10

5

0

25

20

15

10

5

0

25

20

15

10

5

0

25

20

15

10

5

0

20

15

10

5

0

20

15

10

5

0

20

15

10

5

0

20

15

10

5

0

20

15

10

5

0

20

15

10

5

0

25

20

15

10

5

0

25

20

15

10

5

0

30

25

20

15

10

5

0

30

25

20

15

10

5

0

100

75

50

25

0

100

75

50

25

0

100

75

50

25

0

100

75

50

25

0

Figure 8. Distribution Patterns and Quantifications of Specific mRNAs in Sucrose Gradient

Fractions.

(A) Distribution of PP2AA3, GAPC, HDA19, and HDC1 mRNAs in sucrose gradient fractions. WT

and Ami-m (#3) seedlings were incubated under LG conditions for 2 h (left) or under DS conditions

for 1 d (right). Total cell extracts prepared from the seedlings were subjected to sucrose density

gradient sedimentation (15%-50%), and total 15 fractions were collected from each tube. Total RNA

was extracted from each fraction, followed by cDNA synthesis and real-time qRT-PCR using gene-

specific primers. The abundance of mRNA in each fraction was quantified as a percentage of their

total amount in all fractions. Similar results were obtained in three independent experiments, and a

representative result is shown. Error bars represent SE from three technical replications.

(B) The abundance of mRNA in polysomal (P; fractions 9-15) and nonpolysomal fractions (NP;

fractions 1-8), quantified as a percentage of their total amount. Ami3 represents Ami-m (#3). Error

bars represent SE from three biological replications based on three independent experiments (*, P ≤

0.05; **, P ≤ 0.01; ***, P ≤ 0.001).

Figure 9

B CTL 1 2 4 24 48 0.5 1 2 (h)

Phostag

α-Flag

α-Flag

CBB

Dark + Starvation Light + Glucose

Flag:MRF1

MRF4 :Myc

A

λPP

MRF1 :Myc

MRF2 :Myc

MRF3 :Myc

Phostag

α-Myc

α-Myc

*

C Dark + Starvation (24 h) Control Light + Glucose (1 h)

10% 55% 10% 55% 10% 55%

D

0.075

0.085

0.095

0.105

0.115

0.125

0.135

0.075

0.085

0.095

0.105

0.115

0.125

0.135

0.075

0.085

0.095

0.105

0.115

0.125

0.135

INPUT IP : α-Myc

α-Flag

α-Myc

Flag:MRF1

eIF4A-1:Myc

LG DS LG DS LG DS LG DS

OD

25

4

2 3 4 5 6 7 8 9 10 11 12 13 14

Flag:

MRF1

(α-Flag)

RPL10a

2 3 4 5 6 7 8 9 10 11 12 13 14 2 3 4 5 6 7 8 9 10 11 12 13 14

100

kDa

100

75

kDa

60

45

kDa

35

25

100

75

kDa

35

25

100

75

kDa

35

25

100

75

100

kDa

75

75

60

+ + + +

− − + +

+ + + +

− − + +

− + − + − + − +

Figure 9. MRF1 Phosphorylation and Ribosome Association According to Cellular Energy

Availability.

(A) Phosphorylation of MRF1 in vivo. Total protein extracts from N. benthamiana leaves, which

express MRF:Myc proteins, were treated with the lambda phosphatase (λPP). After treatment, the

samples were subjected to Zn2+-Phostag SDS-PAGE (top) and to normal SDS-PAGE (bottom) for

immunoblotting with anti-Myc antibody. The phosphorylated form of MRF1 was marked with the

asterisk.

(B) Phosphorylation of MRF1 under different energy conditions. Flag:MRF1 OE (#1) seedlings

were incubated under dark/starvation conditions for 1-48 h, and then re-illuminated and fed with 30

mM glucose (light/glucose) for 0.5-2 h. Protein extracts from the seedlings harvested at different

time points were separated by Zn2+-Phostag SDS-PAGE (top) and by normal SDS-PAGE (middle),

followed by immunoblotting with anti-Flag antibody. The Rubisco large subunit was stained with

CBB as loading control (bottom).

(C) Distribution of Flag:MRF1 in sucrose gradient fractions of Flag:MRF1 under different energy

conditions. Flag:MRF1 OE (#1) seedlings were incubated in the dark/starvation for 24 h, followed

by re-illumination and glucose-feeding for 1 h. Polysome analysis was performed by

ultracentrifugation through a 10-55% sucrose gradient. Then the fractions were precipitated and

analyzed by immunoblotting with anti-Flag and anti-RPL10a antibodies. Lanes indicate the

fractions from top (10%) to bottom (55%). Arrowheads indicate the final positions of MRF1

detection.

(D) Co-immunoprecipitation. Flag:MRF1 was expressed alone or together with eIF4A-1:Myc in N.

benthamiana leaves. Leaf disks were prepared for treatment with light/glucose (LG) or

dark/starvation (DS) for 3 h. Total leaf proteins were immunoprecipitated with anti-Myc antibody-

conjugated resin, and the co-immunoprecipitate was detected using the anti-Flag antibody.

TRV TOR VIGS

Phostag

α-Flag

α-Flag

CBB

B

Flag:MRF1

D

DM

SO

WT #3 #10 #14 OE#1 OE#2

Tori

n-1

(2 µ

M)

E Ami-m

C

TRV

control

TOR

VIGS

2 3 4 5 6 7 8 9 10 11 12 13 14

α-Flag

10% 55%

0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

DMSO(3 days)

DMSO(5 days)

Torin-1(3 days)

Torin-1(5 days)

Ro

ot le

ng

th (

cm

)

WT #3 #10 #14 OE1 OE2

Ami-m

100

75

kDa

60

45

100

kDa

75

100

75

A MRF1 MRF2 MRF3 MRF4

0

0.5

1.0

1.5

2.0

Stv

Glc

Ma

nS

uc

Stv

Glc

Ma

nS

uc

0

0.5

1.0

1.5

2.0

Stv

Glc

Ma

nS

uc

Stv

Glc

Ma

nS

uc

0

0.5

1.0

1.5

2.0

Stv

Glc

Ma

nS

uc

Stv

Glc

Ma

nS

uc

es-tor1

(RNAi) :

0

0.5

1.0

1.5

2.0

Stv

Glc

Man

Su

c

Stv

Glc

Man

Su

c

Rela

tive

mR

NA

le

vels

−ES +ES −ES +ES −ES +ES −ES +ES

Figure 10

25

25

α-Flag

α-RPL10a

α-RPL10a

Flag:MRF1

Figure 10. TOR-Modulated MRF Gene Expression and MRF1 Phosphorylation, and Seedling

Phenotypes upon Torin-1 Treatment.

(A) Altered MRF gene expression in TOR-silenced seedlings in response to starvation and sugar

feeding. Estradiol-inducible TOR RNAi seedlings (es-tor1) were treated with ethanol (-ES) or 10

μM estradiol (+ES) for gene silencing. Twelve seedlings grown in three different sets of liquid

culture were incubated in glucose-free medium for 24 h (Stv), and then fed with 30 mM glucose

(Glc), mannitol (Man), and sucrose (Suc) for 4 h. RT-qPCR was performed with gene-specific

primers. Transcript levels are normalized by PP2AA3 mRNA, and expressed relative to those of Stv

samples. Error bars represent SE from triplicate technical replications.

(B) MRF1 phosphorylation in TOR-silenced plants. TOR VIGS was performed in Flag:MRF1 OE

(#1) lines. Protein extracts from TRV control or TOR VIGS leaves (10 DAI) were separated by

Phostag SDS-PAGE (top) and by normal SDS-PAGE (middle), followed by immunoblotting with

anti-Flag antibody. The Rubisco large subunit was stained with CBB as loading control (bottom).

(C) Ribosome association of Flag:MRF1 in TOR-silenced plants. TOR VIGS was performed in

Flag:MRF1 OE (#1) lines. Protein extracts from TRV control and TOR VIGS leaves (10 DAI) were

fractionated by sucrose density gradient sedimentation (10-55%). The fractions were precipitated

and analyzed by immunoblotting with anti-Flag and anti-RPL10a antibodies. Lanes indicate the

fractions from top (10%) to bottom (55%). Arrowheads indicate the final positions of MRF1

detection.

(D) Phenotypes of the Ami-m and MRF1 OE seedlings after Torin-1 treatment. Seven-day-old

seedlings grown in liquid culture were treated with Torin-1 (2 μM) or control DMSO for 3 days.

(E) Root length of the seedlings was measured after treatment with Torin-1 (2 μM) or control

DMSO for 3 and 5 days. Each data point represents the mean SE (n > 14 seedlings). Asterisks

denote statistical significance of the differences between Torin-1-treated samples and DMSO-treated

samples, calculated using Student’s t-test (*, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001).

B

CBB

Kinase assay

α-Myc

MBP

MBP:MRF1

MBP:MRF2

MBP:MRF3

MBP:MRF4

[IP-Myc] S6K2:Myc

+

+

+

+

+

+

+

+

+

+

MBP:MRF1

[IP-Myc] S6K1:Myc

[IP-Myc] S6K1(T449A):Myc

[IP-Myc] S6K1(T449D):Myc

+

+

+

+

+

+

+

Kinase assay

CBB

α-Myc

C

A MBP

MBP:MRF1

MBP:MRF2

MBP:MRF3

MBP:MRF4

[IP-Myc] S6K1:Myc

+ − − − − +

+

+

+

+

+

+

+

+

CBB

Kinase assay

α-Myc

Figure 11

D

S6K

1:Y

FP

C

MRF1:YFPN MRF1:YFPN

S6K

2:Y

FP

C

INPUT IP : α-Myc

IB : Myc

IB : Flag

Flag:MTR1

S6K2:Myc

F

IB : Myc

INPUT IP : Myc

IB : Flag

Flag:MTR1

S6K1:Myc

E

140

100

kDa

140

100 100

75

140

100

kDa

140

100 100

75

140

100

kDa

140

100

100

75

100

75

kDa

100

75

100

75

kDa

100

75

+ +

− +

+ +

− +

+ +

− +

+ +

− +

Figure 11. In Vitro Phosphorylation of MRF1 by S6K kinases.

(A) In vitro kinase assay of immunoprecipitated S6K1:Myc with the recombinant MBP:MRF

proteins as substrates. After the kinase assay with [γ-32P]-ATP, SDS-PAGE was performed.

Phosphorylated MBP:MRF proteins were detected by a phosphorimager (top); the MBP:MRF

protein in the reaction was detected by CBB staining (middle); immunoprecipitated S6K1:Myc was

detected by immunoblotting with anti-Myc antibody (bottom).

(B) In vitro kinase assay of immunoprecipitated S6K2:Myc with the recombinant MBP:MRF

proteins as substrates.

(C) In vitro kinase assay with S6K1 mutant forms that carry a mutation in the TOR phosphorylation

site T449. In vitro kinase assay was performed with S6K1:Myc, S6K1(T449A):Myc, and

S6K1(T449D):Myc proteins.

(D) BiFC analyses for MRF1 interactions with S6K1 and S6K2. MRF1:YFPN was expressed

together with S6K1:YFPC or S6K2:YFPC in N. benthamiana leaves using agroinfiltration. Leaf

epidermal cells were observed by confocal microscopy. More than 20 leaf cells showing yellow

fluorescence were observed for each BiFC experiment. As a negative control, MRF1:YFPN and

eIF4E-1:YFPC were co-expressed in N. benthamiana leaves, which resulted in little yellow

fluorescence. Bars = 20 µm.

(E), (F) Co-immunoprecipitation of MRF1 with S6K1 and S6K2. Flag:MRF1 was expressed alone

or together with S6K1:Myc (E) or S6K2:Myc (F) in N. benthamiana leaves. Total leaf proteins were

immunoprecipitated with anti-Myc antibody-conjugated resin, and the co-immunoprecipitate was

detected using the anti-Flag antibody.

Parsed CitationsAhn, C.S., Ahn, H.K., and Pai, H.S. (2015). Overexpression of the PP2A regulatory subunit Tap46 leads to enhanced plant growththrough stimulation of the TOR signalling pathway. J Exp Bot 66, 827-840.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Ahn, C.S., Han, J.A., Lee, H.S., Lee, S., and Pai, H.S. (2011). The PP2A Regulatory Subunit Tap46, a Component of the TOR SignalingPathway, Modulates Growth and Metabolism in Plants. Plant Cell 23, 185-209.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Allen, M.L., Metz, A.M., Timmer, R.T., Rhoads, R.E., and Browning, K.S. (1992). Isolation and sequence of the cDNAs encoding thesubunits of the isozyme form of wheat protein synthesis initiation factor 4F. J Biol Chem 267, 23232-23236.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Andreou, A.Z., and Klostermeier, D. (2013). The DEAD-box helicase eIF4A: paradigm or the odd one out? RNA Biol 10, 19-32.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Bolger, T.A., Folkmann, A.W., Tran, E.J., and Wente, S.R. (2008). The mRNA export factor Gle1 and inositol hexakisphosphate regulatedistinct stages of translation. Cell 134, 624-633.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Browning, K.S. (2004). Plant translation initiation factors: it is not easy to be green. Biochem Soc Trans 32, 589-591.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Browning, K.S., and Bailey-Serres, J. (2015). Mechanism of cytoplasmic mRNA translation. Arabidopsis Book 13, e0176.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Burch-Smith, T.M., Schiff, M., Liu, Y., and Dinesh-Kumar, S.P. (2006). Efficient virus-induced gene silencing in Arabidopsis. PlantPhysiol 142, 21-27.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Caldana, C., Li, Y., Leisse, A., Zhang, Y., Bartholomaeus, L., Fernie, A.R., Willmitzer, L., and Giavalisco, P. (2013). Systemic analysis ofinducible target of rapamycin mutants reveal a general metabolic switch controlling growth in Arabidopsis thaliana. Plant J 73, 897-909.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Cheng, S., Liu, R., and Gallie, D.R. (2013). The unique evolution of the programmed cell death 4 protein in plants. BMC Evol Biol 13,199.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Cherkasova, V.A., and Hinnebusch, A.G. (2003). Translational control by TOR and TAP42 through dephosphorylation of eIF2alphakinase GCN2. Genes Dev 17, 859-872.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Clough, S.J., and Bent, A.F. (1998). Floral dip: a simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana.Plant J 16, 735-743.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Dennis, M.D., Jefferson, L.S., and Kimball, S.R. (2012). Role of p70S6K1-mediated phosphorylation of eIF4B and PDCD4 proteins in theregulation of protein synthesis. J Biol Chem 287, 42890-42899.

Pubmed: Author and TitleCrossRef: Author and Title

Google Scholar: Author Only Title Only Author and Title

Deprost, D., Yao, L., Sormani, R., Moreau, M., Leterreux, G., Nicolai, M., Bedu, M., Robaglia, C., and Meyer, C. (2007). The ArabidopsisTOR kinase links plant growth, yield, stress resistance and mRNA translation. EMBO Rep 8, 864-870.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Dobrenel, T., Caldana, C., Hanson, J., Robaglia, C., Vincentz, M., Veit, B., and Meyer, C. (2016a). TOR Signaling and Nutrient Sensing.Annu Rev Plant Biol 67, 261-285.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Dobrenel, T., Mancera-Martinez, E., Forzani, C., Azzopardi, M., Davanture, M., Moreau, M., Schepetilnikov, M., Chicher, J., Langella, O.,Zivy, M., Robaglia, C., Ryabova, L.A., Hanson, J., and Meyer, C. (2016b). The Arabidopsis TOR Kinase Specifically Regulates theExpression of Nuclear Genes Coding for Plastidic Ribosomal Proteins and the Phosphorylation of the Cytosolic Ribosomal Protein S6.Front Plant Sci 7, 1611.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Dorrello, N.V., Peschiaroli, A., Guardavaccaro, D., Colburn, N.H., Sherman, N.E., and Pagano, M. (2006). S6K1- and betaTRCP-mediateddegradation of PDCD4 promotes protein translation and cell growth. Science 314, 467-471.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Fehler, O., Singh, P., Haas, A., Ulrich, D., Muller, J.P., Ohnheiser, J., and Klempnauer, K.H. (2014). An evolutionarily conservedinteraction of tumor suppressor protein Pdcd4 with the poly(A)-binding protein contributes to translation suppression by Pdcd4.Nucleic Acids Res 42, 11107-11118.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Floyd, B.E., Morriss, S.C., MacIntosh, G.C., and Bassham, D.C. (2015). Evidence for autophagy-dependent pathways of rRNA turnoverin Arabidopsis. Autophagy 11, 2199-2212.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Gamm, M., Peviani, A., Honsel, A., Snel, B., Smeekens, S., and Hanson, J. (2014). Increased sucrose levels mediate selective mRNAtranslation in Arabidopsis. BMC Plant Biol 14, 306.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Gong, Z., Dong, C.H., Lee, H., Zhu, J., Xiong, L., Gong, D., Stevenson, B., and Zhu, J.K. (2005). A DEAD box RNA helicase is essential formRNA export and important for development and stress responses in Arabidopsis. Plant Cell 17, 256-267.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Hemerly, A., Engler Jde, A., Bergounioux, C., Van Montagu, M., Engler, G., Inze, D., and Ferreira, P. (1995). Dominant negative mutantsof the Cdc2 kinase uncouple cell division from iterative plant development. EMBO J 14, 3925-3936.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Holcik, M., and Sonenberg, N. (2005). Translational control in stress and apoptosis. Nat Rev Mol Cell Biol 6, 318-327.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Hutchins, A.P., Roberts, G.R., Lloyd, C.W., and Doonan, J.H. (2004). In vivo interaction between CDKA and eIF4A: a possible mechanismlinking translation and cell proliferation. FEBS Lett 556, 91-94.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Jackson, R.J., Hellen, C.U., and Pestova, T.V. (2010). The mechanism of eukaryotic translation initiation and principles of its regulation.Nat Rev Mol Cell Biol 11, 113-127.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Jankowsky, E., Gross, C.H., Shuman, S., and Pyle, A.M. (2001). Active disruption of an RNA-protein interaction by a DExH/D RNAhelicase. Science 291, 121-125.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Jao, D.L., and Chen, K.Y. (2006). Tandem affinity purification revealed the hypusine-dependent binding of eukaryotic initiation factor 5Ato the translating 80S ribosomal complex. J Cell Biochem 97, 583-598.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Juntawong, P., and Bailey-Serres, J. (2012). Dynamic Light Regulation of Translation Status in Arabidopsis thaliana. Front Plant Sci 3,66.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Juntawong, P., Girke, T., Bazin, J., and Bailey-Serres, J. (2014). Translational dynamics revealed by genome-wide profiling of ribosomefootprints in Arabidopsis. Proc Natl Acad Sci U S A 111, E203-212.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Kang, H., Park, S.J., and Kwak, K.J. (2013). Plant RNA chaperones in stress response. Trends Plant Sci 18, 100-106.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Kim, T.H., Kim, B.H., Yahalom, A., Chamovitz, D.A., and von Arnim, A.G. (2004). Translational regulation via 5' mRNA leader sequencesrevealed by mutational analysis of the Arabidopsis translation initiation factor subunit eIF3h. Plant Cell 16, 3341-3356.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Kinoshita, E., and Kinoshita-Kikuta, E. (2011). Improved Phos-tag SDS-PAGE under neutral pH conditions for advanced proteinphosphorylation profiling. Proteomics 11, 319-323.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Koroleva, O.A., Calder, G., Pendle, A.F., Kim, S.H., Lewandowska, D., Simpson, C.G., Jones, I.M., Brown, J.W., and Shaw, P.J. (2009).Dynamic behavior of Arabidopsis eIF4A-III, putative core protein of exon junction complex: fast relocation to nucleolus and splicingspeckles under hypoxia. Plant Cell 21, 1592-1606.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Lageix, S., Lanet, E., Pouch-Pelissier, M.N., Espagnol, M.C., Robaglia, C., Deragon, J.M., and Pelissier, T. (2008). Arabidopsis eIF2alphakinase GCN2 is essential for growth in stress conditions and is activated by wounding. BMC Plant Biol 8, 134.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Lankat-Buttgereit, B., and Goke, R. (2009). The tumour suppressor Pdcd4: recent advances in the elucidation of function andregulation. Biol Cell 101, 309-317.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Lastdrager, J., Hanson, J., and Smeekens, S. (2014). Sugar signals and the control of plant growth and development. J Exp Bot 65, 799-807.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Lee, H.S., Lee, D.H., Cho, H.K., Kim, S.H., Auh, J.H., and Pai, H.S. (2015). InsP6-Sensitive Variants of the Gle1 mRNA Export FactorRescue Growth and Fertility Defects of the ipk1 Low-Phytic-Acid Mutation in Arabidopsis. Plant Cell.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Lei, G., Shen, M., Li, Z.G., Zhang, B., Duan, K.X., Wang, N., Cao, Y.R., Zhang, W.K., Ma, B., Ling, H.Q., Chen, S.Y., and Zhang, J.S. (2011).EIN2 regulates salt stress response and interacts with a MA3 domain-containing protein ECIP1 in Arabidopsis. Plant Cell Environ 34,1678-1692.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Li, W., Ma, M., Feng, Y., Li, H., Wang, Y., Ma, Y., Li, M., An, F., and Guo, H. (2015). EIN2-directed translational regulation of ethylenesignaling in Arabidopsis. Cell 163, 670-683.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Liu, M.J., Wu, S.H., Chen, H.M., and Wu, S.H. (2012). Widespread translational control contributes to the regulation of Arabidopsisphotomorphogenesis. Mol Syst Biol 8, 566.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Loh, P.G., Yang, H.S., Walsh, M.A., Wang, Q., Wang, X., Cheng, Z., Liu, D., and Song, H. (2009). Structural basis for translationalinhibition by the tumour suppressor Pdcd4. EMBO J 28, 274-285.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Ma, X.M., and Blenis, J. (2009). Molecular mechanisms of mTOR-mediated translational control. Nat Rev Mol Cell Biol 10, 307-318.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Ma, Y., Miura, E., Ham, B.K., Cheng, H.W., Lee, Y.J., and Lucas, W.J. (2010). Pumpkin eIF5A isoforms interact with components of thetranslational machinery in the cucurbit sieve tube system. Plant J 64, 536-550.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Magnuson, B., Ekim, B., and Fingar, D.C. (2012). Regulation and function of ribosomal protein S6 kinase (S6K) within mTOR signallingnetworks. Biochem J 441, 1-21.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Mahfouz, M.M., Kim, S., Delauney, A.J., and Verma, D.P. (2006). Arabidopsis TARGET OF RAPAMYCIN interacts with RAPTOR, whichregulates the activity of S6 kinase in response to osmotic stress signals. Plant Cell 18, 477-490.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Mair, A., Pedrotti, L., Wurzinger, B., Anrather, D., Simeunovic, A., Weiste, C., Valerio, C., Dietrich, K., Kirchler, T., Nagele, T., VicenteCarbajosa, J., Hanson, J., Baena-Gonzalez, E., Chaban, C., Weckwerth, W., Droge-Laser, W., and Teige, M. (2015). SnRK1-triggeredswitch of bZIP63 dimerization mediates the low-energy response in plants. Elife 4.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Mehdi, S., Derkacheva, M., Ramstrom, M., Kralemann, L., Bergquist, J., and Hennig, L. (2016). The WD40 Domain Protein MSI1Functions in a Histone Deacetylase Complex to Fine-Tune Abscisic Acid Signaling. Plant Cell 28, 42-54.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Merchante, C., Stepanova, A.N., and Alonso, J.M. (2017). Translation regulation in plants: an interesting past, an exciting present and apromising future. Plant J 90, 628-653.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Merchante, C., Brumos, J., Yun, J., Hu, Q., Spencer, K.R., Enriquez, P., Binder, B.M., Heber, S., Stepanova, A.N., and Alonso, J.M.(2015). Gene-specific translation regulation mediated by the hormone-signaling molecule EIN2. Cell 163, 684-697.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Moreau, M., Azzopardi, M., Clement, G., Dobrenel, T., Marchive, C., Renne, C., Martin-Magniette, M.L., Taconnat, L., Renou, J.P.,Robaglia, C., and Meyer, C. (2012). Mutations in the Arabidopsis homolog of LST8/GbetaL, a partner of the target of Rapamycin kinase,impair plant growth, flowering, and metabolic adaptation to long days. Plant Cell 24, 463-481.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Google Scholar: Author Only Title Only Author and Title

Muench, D.G., Zhang, C., and Dahodwala, M. (2012). Control of cytoplasmic translation in plants. Wiley Interdiscip Rev RNA 3, 178-194.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Nakaminami, K., Karlson, D.T., and Imai, R. (2006). Functional conservation of cold shock domains in bacteria and higher plants. ProcNatl Acad Sci U S A 103, 10122-10127.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Nicolai, M., Roncato, M.A., Canoy, A.S., Rouquie, D., Sarda, X., Freyssinet, G., and Robaglia, C. (2006). Large-scale analysis of mRNAtranslation states during sucrose starvation in arabidopsis cells identifies cell proliferation and chromatin structure as targets oftranslational control. Plant Physiol 141, 663-673.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Nukarinen, E., Nagele, T., Pedrotti, L., Wurzinger, B., Mair, A., Landgraf, R., Bornke, F., Hanson, J., Teige, M., Baena-Gonzalez, E.,Droge-Laser, W., and Weckwerth, W. (2016). Quantitative phosphoproteomics reveals the role of the AMPK plant ortholog SnRK1 as ametabolic master regulator under energy deprivation. Sci Rep 6, 31697.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Pal, S.K., Liput, M., Piques, M., Ishihara, H., Obata, T., Martins, M.C., Sulpice, R., van Dongen, J.T., Fernie, A.R., Yadav, U.P., Lunn, J.E.,Usadel, B., and Stitt, M. (2013). Diurnal changes of polysome loading track sucrose content in the rosette of wild-type arabidopsis andthe starchless pgm mutant. Plant Physiol 162, 1246-1265.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Patrick, R.M., and Browning, K.S. (2012). The eIF4F and eIFiso4F Complexes of Plants: An Evolutionary Perspective. Comp FunctGenomics 2012, 287814.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Perrella, G., Lopez-Vernaza, M.A., Carr, C., Sani, E., Gossele, V., Verduyn, C., Kellermeier, F., Hannah, M.A., and Amtmann, A. (2013).Histone deacetylase complex1 expression level titrates plant growth and abscisic acid sensitivity in Arabidopsis. Plant Cell 25, 3491-3505.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Pestov, D.G., and Shcherbik, N. (2012). Rapid cytoplasmic turnover of yeast ribosomes in response to rapamycin inhibition of TOR. MolCell Biol 32, 2135-2144.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Piir, K., Paier, A., Liiv, A., Tenson, T., and Maivali, U. (2011). Ribosome degradation in growing bacteria. EMBO Rep 12, 458-462.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Porra, R.J., and Scheer, H. (2000). (18)O and mass spectrometry in chlorophyll research: Derivation and loss of oxygen atoms at theperiphery of the chlorophyll macrocycle during biosynthesis, degradation and adaptation. Photosynth Res 66, 159-175.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Powers, T., and Walter, P. (1999). Regulation of ribosome biogenesis by the rapamycin-sensitive TOR-signaling pathway inSaccharomyces cerevisiae. Mol Biol Cell 10, 987-1000.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Ren, M., Venglat, P., Qiu, S., Feng, L., Cao, Y., Wang, E., Xiang, D., Wang, J., Alexander, D., Chalivendra, S., Logan, D., Mattoo, A.,Selvaraj, G., and Datla, R. (2012). Target of rapamycin signaling regulates metabolism, growth, and life span in Arabidopsis. Plant Cell24, 4850-4874.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Saini, P., Eyler, D.E., Green, R., and Dever, T.E. (2009). Hypusine-containing protein eIF5A promotes translation elongation. Nature 459,118-121.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Schepetilnikov, M., Dimitrova, M., Mancera-Martinez, E., Geldreich, A., Keller, M., and Ryabova, L.A. (2013). TOR and S6K1 promotetranslation reinitiation of uORF-containing mRNAs via phosphorylation of eIF3h. EMBO J 32, 1087-1102.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Schwab, R., Ossowski, S., Riester, M., Warthmann, N., and Weigel, D. (2006). Highly specific gene silencing by artificial microRNAs inArabidopsis. Plant Cell 18, 1121-1133.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Sengupta, S., Peterson, T.R., and Sabatini, D.M. (2010). Regulation of the mTOR complex 1 pathway by nutrients, growth factors, andstress. Mol Cell 40, 310-322.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Shaikhin, S.M., Smailov, S.K., Lee, A.V., Kozhanov, E.V., and Iskakov, B.K. (1992). Interaction of wheat germ translation initiation factor2 with GDP and GTP. Biochimie 74, 447-454.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Silvera, D., Formenti, S.C., and Schneider, R.J. (2010). Translational control in cancer. Nat Rev Cancer 10, 254-266.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Sonenberg, N., and Hinnebusch, A.G. (2009). Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell136, 731-745.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Tome, F., Nagele, T., Adamo, M., Garg, A., Marco-Llorca, C., Nukarinen, E., Pedrotti, L., Peviani, A., Simeunovic, A., Tatkiewicz, A.,Tomar, M., and Gamm, M. (2014). The low energy signaling network. Front Plant Sci 5, 353.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Turck, F., Zilbermann, F., Kozma, S.C., Thomas, G., and Nagy, F. (2004). Phytohormones participate in an S6 kinase signal transductionpathway in Arabidopsis. Plant Physiol 134, 1527-1535.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Voinnet, O., Rivas, S., Mestre, P., and Baulcombe, D. (2003). An enhanced transient expression system in plants based on suppressionof gene silencing by the p19 protein of tomato bushy stunt virus. Plant J 33, 949-956.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Walter, M., Chaban, C., Schutze, K., Batistic, O., Weckermann, K., Nake, C., Blazevic, D., Grefen, C., Schumacher, K., Oecking, C.,Harter, K., and Kudla, J. (2004). Visualization of protein interactions in living plant cells using bimolecular fluorescencecomplementation. Plant J 40, 428-438.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Wang, L., Li, H., Zhao, C., Li, S., Kong, L., Wu, W., Kong, W., Liu, Y., Wei, Y., Zhu, J.K., and Zhang, H. (2017). The inhibition of proteintranslation mediated by AtGCN1 is essential for cold tolerance in Arabidopsis thaliana. Plant Cell Environ 40, 56-68.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Wedeken, L., Singh, P., and Klempnauer, K.H. (2011). Tumor suppressor protein Pdcd4 inhibits translation of p53 mRNA. J Biol Chem286, 42855-42862.

Pubmed: Author and Title

CrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Wullschleger, S., Loewith, R., and Hall, M.N. (2006). TOR signaling in growth and metabolism. Cell 124, 471-484.Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Xia, B., Ke, H., and Inouye, M. (2001). Acquirement of cold sensitivity by quadruple deletion of the cspA family and its suppression byPNPase S1 domain in Escherichia coli. Mol Microbiol 40, 179-188.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Xiong, Y., and Sheen, J. (2012). Rapamycin and glucose-target of rapamycin (TOR) protein signaling in plants. J Biol Chem 287, 2836-2842.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Xiong, Y., McCormack, M., Li, L., Hall, Q., Xiang, C., and Sheen, J. (2013). Glucose-TOR signalling reprograms the transcriptome andactivates meristems. Nature 496, 181-186.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Yang, H.S., Cho, M.H., Zakowicz, H., Hegamyer, G., Sonenberg, N., and Colburn, N.H. (2004). A novel function of the MA-3 domains intransformation and translation suppressor Pdcd4 is essential for its binding to eukaryotic translation initiation factor 4A. Mol Cell Biol24, 3894-3906.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Yang, H.S., Jansen, A.P., Komar, A.A., Zheng, X., Merrick, W.C., Costes, S., Lockett, S.J., Sonenberg, N., and Colburn, N.H. (2003). Thetransformation suppressor Pdcd4 is a novel eukaryotic translation initiation factor 4A binding protein that inhibits translation. Mol CellBiol 23, 26-37.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Zhou, F., Roy, B., and von Arnim, A.G. (2010). Translation reinitiation and development are compromised in similar ways by mutations intranslation initiation factor eIF3h and the ribosomal protein RPL24. BMC Plant Biol 10, 193.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

Zundel, M.A., Basturea, G.N., and Deutscher, M.P. (2009). Initiation of ribosome degradation during starvation in Escherichia coli. RNA15, 977-983.

Pubmed: Author and TitleCrossRef: Author and TitleGoogle Scholar: Author Only Title Only Author and Title

DOI 10.1105/tpc.17.00563; originally published online October 30, 2017;Plant Cell

Du-Hwa Lee, Seung Jun Park, Chang Sook Ahn and Hyun-Sook PaiConditions, and Their Expression and Functions Are Modulated by the TOR Signaling Pathway

MRF Family Genes Are Involved in Protein Translation Control, Especially under Energy-Deficient

 This information is current as of August 31, 2018

 

Supplemental Data /content/suppl/2017/10/30/tpc.17.00563.DC1.html

Permissions https://www.copyright.com/ccc/openurl.do?sid=pd_hw1532298X&issn=1532298X&WT.mc_id=pd_hw1532298X

eTOCs http://www.plantcell.org/cgi/alerts/ctmain

Sign up for eTOCs at:

CiteTrack Alerts http://www.plantcell.org/cgi/alerts/ctmain

Sign up for CiteTrack Alerts at:

Subscription Information http://www.aspb.org/publications/subscriptions.cfm

is available at:Plant Physiology and The Plant CellSubscription Information for

ADVANCING THE SCIENCE OF PLANT BIOLOGY © American Society of Plant Biologists