1 short title: transcriptomics of grape berry development · 3 50 abstract 51 grapevine berry...

62
1 Short title: Transcriptomics of grape berry development 1 2 Corresponding author details: Sara Zenoni, Department of Biotechnology, University of 3 Verona, Strada le Grazie 15, 37134, Verona, Italy, Tel. +39-045-8027941; Fax +39-045- 4 8027929 5 6 Transcriptomic differences in grapevine varieties correlate with berry anthocyanin skin 7 Mélanie Massonnet 1 , Marianna Fasoli 1 , Giovanni Battista Tornielli 1 , Mario Altieri 1 , Marco 8 Sandri 1 , Paola Zuccolotto 2 , Paola Paci 3,4 , Massimo Gardiman 5 , Sara Zenoni 1 *, Mario 9 Pezzotti 1 10 11 1 Department of Biotechnology, University of Verona, Strada le Grazie 15, 37134, Verona, 12 Italy 13 2 Big&Open Data Innovation Laboratory (BODaI-Lab), University of Brescia, Via Branze 45, 14 25123 Brescia, Italy 15 3 Institute for Systems Analysis and Computer Science "Antonio Ruberti," National Research 16 Council, 00185 Rome, Italy SysBio Centre for Systems Biology, 00185 Rome, Italy 17 4 SysBio Centre for Systems Biology, 00185 Rome, Italy 18 5 (CREA-VE), Viale XXVIII Aprile 26, Conegliano (TV) 31015, Italy 19 20 One sentence summary 21 Grapevine berry development features both common and skin color-dependent transcriptomic 22 changes. 23 24 List of author contributions 25 M.M performed the experiments, analyzed the data, interpreted the results and wrote the 26 manuscript; M.F interpreted the results and wrote the manuscript; G.B.T. interpreted the 27 results and was involved in the experimental design; M.A analyzed the sequence data; M.S 28 and P.Z. performed gene clustering analysis and analyzed PCA loadings; P.P analyzed the 29 Plant Physiology Preview. Published on June 26, 2017, as DOI:10.1104/pp.17.00311 Copyright 2017 by the American Society of Plant Biologists www.plantphysiol.org on March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

Upload: others

Post on 18-Mar-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

  • 1

    Short title: Transcriptomics of grape berry development 1

    2

    Corresponding author details: Sara Zenoni, Department of Biotechnology, University of 3 Verona, Strada le Grazie 15, 37134, Verona, Italy, Tel. +39-045-8027941; Fax +39-045-4

    8027929 5

    6

    Transcriptomic differences in grapevine varieties correlate with berry anthocyanin skin 7

    Mélanie Massonnet1, Marianna Fasoli1, Giovanni Battista Tornielli1, Mario Altieri1, Marco 8

    Sandri1, Paola Zuccolotto2, Paola Paci3,4, Massimo Gardiman5, Sara Zenoni1*, Mario 9

    Pezzotti1 10

    11

    1Department of Biotechnology, University of Verona, Strada le Grazie 15, 37134, Verona, 12

    Italy 13

    2Big&Open Data Innovation Laboratory (BODaI-Lab), University of Brescia, Via Branze 45, 14

    25123 Brescia, Italy 15

    3Institute for Systems Analysis and Computer Science "Antonio Ruberti," National Research 16

    Council, 00185 Rome, Italy SysBio Centre for Systems Biology, 00185 Rome, Italy 17

    4SysBio Centre for Systems Biology, 00185 Rome, Italy 18

    5 (CREA-VE), Viale XXVIII Aprile 26, Conegliano (TV) 31015, Italy 19

    20

    One sentence summary 21

    Grapevine berry development features both common and skin color-dependent transcriptomic 22

    changes. 23

    24

    List of author contributions 25

    M.M performed the experiments, analyzed the data, interpreted the results and wrote the 26

    manuscript; M.F interpreted the results and wrote the manuscript; G.B.T. interpreted the 27

    results and was involved in the experimental design; M.A analyzed the sequence data; M.S 28

    and P.Z. performed gene clustering analysis and analyzed PCA loadings; P.P analyzed the 29

    Plant Physiology Preview. Published on June 26, 2017, as DOI:10.1104/pp.17.00311

    Copyright 2017 by the American Society of Plant Biologists

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 2

    switch genes; M.G. sampled the berries and was involved in the experimental design; S.Z was 30

    involved in the experimental design, supervised the study and wrote the manuscript; M.P. 31

    conceived the study and reviewed the manuscript. 32

    All authors contributed to the discussion of the results, reviewed the manuscript, and 33

    approved the final article. 34

    35

    Funding information 36

    This work was supported by the Valorizzazione dei Principali Vitigni Autoctoni Italiani e dei 37

    loro Terroir (Vigneto) project funded by the Italian Ministry of Agricultural and Forestry 38

    Policies and by the European funded COST ACTION FA1106 Quality fruit. 39

    40

    Present addresses 41

    Mélanie Massonnet: Department of Viticulture and Enology, University of California Davis, 42

    Davis, CA 95616, USA 43

    Marianna Fasoli: E & J Gallo Winery Modesto, CA 95353, USA. 44

    45

    Corresponding author email 46

    [email protected] 47

    48

    49

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 3

    ABSTRACT 50

    Grapevine berry development involves a succession of physiological and biochemical 51

    changes reflecting the transcriptional modulation of thousands of genes. Although recent 52

    studies have investigated the dynamic transcriptome during berry development, most have 53

    focused on a single grapevine variety so there is a lack of comparative data representing 54

    different cultivars. Here we report the first genome-wide transcriptional analysis of 120 RNA 55

    samples corresponding to 10 Italian grapevine varieties collected at four growth stages. The 56

    10 varieties, representing five red-skinned and five white-skinned berries, were all cultivated 57

    in the same experimental vineyard, to reduce environmental variability. The comparison of 58

    transcriptional changes during berry formation and ripening allowed us to determine the 59

    transcriptomic traits common to all varieties, thus defining the core transcriptome of berry 60

    development, as well as the transcriptional dynamics underlying differences between red and 61

    white berry varieties. A greater variation among the red cultivars than between red and white 62

    cultivars at the transcriptome level was revealed, suggesting that anthocyanin accumulation 63

    during berry maturation has a direct impact on the transcriptomic regulation of multiple 64

    biological processes. The expression of genes related to phenylpropanoid/flavonoid 65

    biosynthesis clearly distinguished the behavior of red and white berry genotypes during 66

    ripening but also reflected the differential accumulation of anthocyanins in the red berries, 67

    indicating some form of cross talk between the activation of stilbene biosynthesis and the 68

    accumulation of anthocyanins in ripening berries. 69

    70

    71

    72

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 4

    INTRODUCTION 73

    Grapevine (Vitis vinifera L.) is one of the most widely cultivated fruit crops, accounting for 74

    7.5 million hectares of arable land in 2015 and producing 75.1 million metric tons of berries 75

    (http://www.oiv.int/). Grape berries are grown commercially for fresh fruit, juice and raisins, 76

    but mostly for fermentation into wine, with almost 274 million hectoliters produced in 2015. 77

    Although winemaking techniques and equipment have improved over recent decades, the 78

    attributes of wine still rely predominantly on berry quality, i.e. the metabolic composition at 79

    harvest (Bindon et al., 2014). Berry quality traits depend mainly on sugars, organic acids, and 80

    secondary metabolites such as tannins, flavonols, anthocyanins, aroma precursors, and 81

    volatile compounds that accumulate during berry development (Conde et al., 2007). The 82

    metabolic profile depends on the genotype and environment, but also on the specific 83

    ‘genotype x environment’ interaction (Jackson and Lombard, 1993; Duchêne and Schneider, 84

    2005; van Leeuwen et al., 2004). Viticulture practices also affect berry ripening and final 85

    quality, and strategies that can be used to promote maturation include irrigation, canopy 86

    management and cropping levels (Jackson, 2014). These strategies are often applied without a 87

    full understanding of the impact at the molecular level, with the exception of a few studies 88

    that have correlated biochemical and physiological outcomes with transcriptomic changes 89

    (Deluc et al., 2009; Leng et al., 2016; Pastore et al., 2011 and 2013; Savoi et al., 2016). 90

    Berry development is a complex process involving profound physiological and metabolic 91

    changes. The grape berry undergoes two distinct growth phases: the herbaceous (green) phase 92

    from flowering to ~60 days, and the ripening/maturation phase, during which the berry 93

    reaches full maturity (Coombe, 1992). During the herbaceous phase, the berry grows rapidly 94

    by cell division then cell expansion and the seeds emerge. Accordingly, the berry is hard, 95

    green due to the presence of chlorophyll, highly acidic due to the accumulation of organic 96

    acids, and bitter and astringent due to the high concentration of tannins in the skin (Conde et 97

    al, 2007). The end of the herbaceous phase is characterized by a slowing of berry growth 98

    during which malic acid accumulates in the flesh and the overall organic acid concentration 99

    reaches its highest level. The transition to the ripening phase is known as véraison, and this 100

    occurs when the seeds are mature. During the ripening phase, further metabolic changes make 101

    the fruit edible and attractive to promote seed dispersal, including a change of skin color, cell 102

    expansion coupled with water influx, berry softening, the accumulation of sugars in the pulp, 103

    the loss of organic acids and tannins, and the synthesis of volatile aromas (Conde et al, 2007). 104

    At harvest, fully ripe berries manifest cultivar-specific features defining their shape, weight 105

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 5

    and metabolic composition, and most of all the skin color, which results in red or white 106

    skinned varieties. Red berries accumulate anthocyanins in the skin cell vacuoles whereas 107

    white berries do not due to mutations affecting the transcription factors (MybA1 and MybA2) 108

    that activate the gene encoding UDP glucose:flavonoid 3-o-glucosyltransferase (UFGT), 109

    which catalyzes a rate-limiting step in anthocyanin synthesis (Kobayashi et al., 2004; Walker 110

    et al., 2007). The relative stability of the quantitative and qualitative anthocyanin profiles in 111

    each variety, together with the widespread distribution of anthocyanins among grape 112

    cultivars, means that the anthocyanin profile is a valuable chemotaxonomic parameter for the 113

    classification of red berry varieties (Wenzel et al., 1987; Mattivi et al., 2006). 114

    The publication of the grapevine reference genome (Jaillon et al., 2007) and the development 115

    of new tools for transcriptomics have facilitated recent advances in the genome-wide analysis 116

    of dynamic gene expression during berry development (Zamboni et al., 2010; Zenoni et al., 117

    2010; Fortes et al., 2011; Guillaumie et al., 2011; Fasoli et al., 2012; Lijavetzky et al., 2012; 118

    Sweetman et al., 2012; Dal Santo et al., 2013b; Cramer et al., 2014; Palumbo et al., 2014; Dal 119

    Santo et al., 2016; Wong et al., 2016). A deep transcriptome shift driving the berry into the 120

    maturation program was demonstrated in the red cultivar Corvina, showing that more genes 121

    are downregulated than upregulated during veraison (Fasoli et al., 2012). The application of 122

    different network methods to berry development transcriptome datasets identified so-called 123

    switch genes, which probably encode key regulators of the developmental transition (Palumbo 124

    et al., 2014). However, further molecular changes not strictly related to skin color have been 125

    observed during berry development in different genotypes, suggesting that the core 126

    transcriptomic changes may be supplemented by variety-dependent effects (Degu et al, 2014; 127

    Corso et al., 2015). The transcriptomic plasticity of berry ripening has been reported in both 128

    red and white cultivars growing in different areas, emphasizing the need to account for 129

    environmental variability when comparing berry ripening in different genotypes (Dal Santo et 130

    al., 2013a; 2016). 131

    In this study, we selected five red and five white Italian grapevine varieties, typical of 132

    different Regions, representing diverse agronomic traits and different environmental 133

    adaptations. Transcriptional changes during berry ripening were compared when the 10 134

    varieties were grown under the same environmental conditions, allowing us to distinguish 135

    core transcriptomic traits from those dependent on the berry skin color. Multivariate analysis 136

    revealed significant variation among the red cultivars, prompting us to focus on the 137

    transcriptional regulation of the general phenylpropanoid pathway and the influence of 138

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 6

    anthocyanin during berry maturation. 139

    140

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 7

    RESULTS 141

    Transcriptomic profiling of berry development in 10 grapevine varieties 142

    We analyzed the berry transcriptomes of 10 grapevine varieties (five red and five white) 143

    cultivated under the same conditions and collected during the same year (Fig. 1A). Berries 144 were sampled in triplicate at four developmental stages, two before and two after veraison 145

    (Fig. 1B): the pea-sized berry stage (Bbch 75) at 20 days after flowering (P), the ‘berries 146 beginning to touch’ stage (Bbch77) just prior to veraison (PV), the berry softening stage 147 (Bbch 85) at the end of veraison (EV), and the fully-ripe berry stage (Bbch 89) at harvest (H). 148 Brix values were measured in all varieties at the EV and H stages, revealing diverse sugar 149

    accumulation trends, with Refosco, then Barbera, Negro amaro and Moscato bianco showing 150

    the highest sugar concentrations at H, and Passerina the lowest (Fig. 1C). The total 151 anthocyanin levels were measured in the red varieties at the same stages, also revealing 152

    different trends, with Refosco accumulating the highest level of anthocyanins at H (Fig. 1C). 153

    The triplicate sampling of four development stages in 10 varieties yielded 120 RNA samples 154

    for RNA sequencing (RNA-Seq) analysis, and 28,607,535 ± 5,342,814 reads were mapped 155

    onto the 12x reference genome PN40024 (Jaillon et al., 2007) (Supplemental Table S1). The 156 mean normalized expression value per transcript (FPKM – fragments per kilobase of mapped 157

    reads) of the three biological replicates was calculated for each sample using the geometric 158

    normalization method and genes with a mean FPKM value >1 in at least one of the 40 159

    triplicate samples were considered to be expressed. The resulting dataset comprising 21,746 160

    transcripts was used for subsequent analysis (Supplementary Dataset S1). 161

    The entire dataset was analyzed first by identifying the number of expressed genes in each 162

    variety at each growth stage, revealing that the number of expressed genes decreased slightly 163

    after veraison in all varieties (Fig. 2 and Supplemental Figure S1). This decline seemed to 164 affect more particularly genes with high and medium expression levels (>100 and >10 FPKM, 165

    respectively) whereas genes with low and very low expression levels (< 10 and

  • 8

    distinction between the transcriptomes before and after veraison, regardless of the variety or 173

    skin color (Fig. 3A). A dendrogram generated by converting correlation coefficients into 174 distance values confirmed the strong clustering of the samples independent of the genotype 175

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 9

    and skin color during the herbaceous phase (Fig. 3B), whereas during maturation the Refosco, 176 Barbera and Negro amaro berry transcriptomes grouped separately from the other varieties. 177

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 10

    We investigated these relationships in more detail by principal component analysis (PCA) and 178

    orthogonal projection of latent structures discriminant analysis (O2PLS-DA), revealing 179

    biomarkers specific to the herbaceous or maturation phases, and biomarkers specific to each 180

    of the four stages. These results are presented in the Supplemental Data. 181

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 11

    182

    Definition of the core berry development transcriptome 183

    In order to define genes with similar expression profiles in all the 10 varieties during berry 184

    development, we first identified differentially expressed genes (DEGs) between each pair of 185

    consecutive developmental stages (i.e. P-PV, PV-EV, EV-H) using Cuffdiff v2.0.2 (Roberts 186

    et al., 2011). The number of DEGs ranged from 11,997 in Passerina to 16,709 in Barbera, 187

    indicating that berry development involves the transcriptional modulation of a large number 188

    of genes (Supplemental Dataset S2). A further inspection of the number of DEGs in each 189 pairwise comparison showed that the dynamics of gene expression modulation differed 190

    among the 10 varieties (Fig. 4A). Negro amaro, Garganega, Sangiovese and Passerina were 191 characterized by an enormous increase in the number of DEGs between PV and EV, followed 192

    by a marked decline in the number of DEGs from EV to H. Primitivo, Vermentino, Glera and 193

    Moscato bianco showed a similar trend, but with a smaller amplitude than the other varieties. 194

    Refosco showed a marked increase of the number of DEGs from PV to EV followed by slight 195

    further increase from EV to H, whereas Barbera was characterized by the progressive 196

    decrease of DEGs throughout development (Fig. 4A). These trends were mostly conserved 197 when applying a cut-off of gene expression of FPKM > 1 in at least one developmental stage 198

    for each variety, indicating that the analysis was not biased by any variety. In addition, the 199

    trends were similar with higher gene expression level cut-offs such as FPKM > 10 and FPKM 200

    > 100 (Supplemental Figure S2). 201

    By comparing DEGs in the 10 cultivars, we found that 4,613 genes were commonly 202

    upregulated or downregulated in at least one pairwise comparison in all 10 varieties, and 203

    exhibited an expression intensity >1 FPKM in at least one growth stage in all varieties (Fig. 204 4B and Supplemental Dataset S3). The most intense transcriptional modulation occurred 205 between PV and EV, underscoring the dramatic metabolic changes that characterize veraison. 206

    Moreover, we found that more genes were downregulated rather than upregulated during 207

    berry development, in particular during the transition from EV to H (Fig. 4B), confirming the 208 previous results reported for Corvina berries (Fasoli et al., 2012). 209

    In order to establish the core berry development transcriptome, clustering analysis was 210

    applied to the 4,613 genes commonly modulated during berry development, allowing the 211

    identification of 913 genes showing similar expression trends in all 10 varieties, including 858 212

    genes (94%) and 746 genes (82%) with FPKM > 5 and FPKM > 10 in at least one sample, 213

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 12

    respectively. These genes grouped into six clusters: the first three clusters contained genes 214

    with expression peaks within one herbaceous stage or throughout the herbaceous phase, 215

    whereas the last three clusters contained genes with expression peaks within one ripening 216

    stage or throughout the maturation phase (Fig. 4C and Supplemental Dataset S4). The 217

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 13

    VitisNet functional annotations were used to assign Gene Ontology (GO) categories to these 218

    913 genes (Grimplet et al., 2009) and enrichment analysis for the GO categories among each 219

    cluster was performed using the BiNGO tool with PlantGOslim categories (Maere et al., 220

    2005; Supplemental Dataset S4). 221

    Cluster 1 contained 542 genes whose expression declined throughout berry development, and 222

    the GO category “Cell cycle” was significantly overrepresented (adjusted P-value ≤ 0.01), 223

    including a cyclin B gene and five cyclin D genes that are essential parts of the plant cell 224

    cycle machinery (De Veylder et al., 2007). Another enriched GO category was “Cellular 225

    component organization”, including several cellulose synthase genes, some xyloglucan 226

    endotransglucosylase/hydrolase genes and genes involved in pectin metabolism (Desprez et 227

    al., 2007; Rose et al., 2002). Cluster 2 contained 94 genes whose expression peaked at the PV 228

    stage. A large proportion of these genes represented the GO categories “DNA/RNA metabolic 229

    process” and “Transcription factor activity” (Fig. 4C), including 6 genes encoding ribosomal 230 proteins and 10 encoding transcription factors. Cluster 3 comprised genes whose expression 231

    declined after veraison, and as expected the GO category “Photosynthesis” was significantly 232

    overrepresented, including two genes encoding photosystem I light harvesting chlorophyll a/b 233

    binding proteins (VIT_15s0024g00040; VIT_18s0001g10550) and one encoding a 234

    photosystem II oxygen-evolving enhancer protein (VIT_12s0028g01080). The same cluster 235

    also contained genes involved in auxin transport and gene regulation, including two genes 236

    encoding auxin efflux carriers and four encoding auxin response factors. Cluster 4 contained 237

    20 genes whose expression peaked at the EV stage, including the sugar transporter gene 238

    HT6/TMT2 (VIT_18s0122g00850) and a SWEET gene (VIT_17s0000g00830), a tonoplast 239

    intrinsic aquaporin (TIP) gene (VIT_13s0019g00330), a pectinesterase gene 240

    (VIT_09s0002g00330) and the expansin gene EXPA18 (VIT_17s0053g00990), the latter two 241

    being involved in cell wall loosening (Seymour and Gross, 1996; Dal Santo et al., 2013b). 242

    Cluster 5 contained 134 genes whose expression increased between EV and H. Interestingly, 243

    we noted the presence of the expansin gene EXPB4 (VIT_15s0021g02700), which is probably 244

    involved in cell wall expansion and modification during ripening (Dal Santo et al., 2013b), 245

    and the germacrene D synthase gene TPS07 (VIT_18s0001g04280), which is involved in 246

    sesquiterpene production (Martin et al., 2010 and 2012). Cluster 6 comprised only three genes 247

    whose expression increased after veraison and then remained constant until harvest. These 248

    genes encoded a xyloglucan endotransglucosylase/hydrolase (VIT_01s0011g06250), a 249

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 14

    predicted E3 ubiquitin-protein ligase (VIT_19s0015g00670, VIT_19s0015g00680) and an 250

    uncharacterized protein (VIT_00s0265g00130). 251

    The entire dataset was also searched to find genes with a constant expression level throughout 252

    berry development in all genotypes. Transcripts were classified according to their coefficient 253

    of variation across the 40 triplicate samples as described by Massa et al. (2011). We identified 254

    80 genes with a constant level of expression (coefficient of variation < 0.1) throughout berry 255

    development in all 10 varieties (Supplemental Dataset S5). The four genes with the lowest 256 coefficients of variation (Supplemental Figure S3) encoded a ubiquitin-specific protease 257 (VIT_13s0073g00610), actin (VIT_04s0044g00580), a β-adaptin (VIT_02s0025g00100) and 258

    a phosphoinositide 4-kinase (VIT_04s0044g01210), with expression intensities of 259

    approximately 58, 1295, 86, and 14 FPKM, respectively. Moreover, a comparison of the 80 260

    constitutively expressed genes with the 76 non-plastic and constitutive transcripts in Corvina 261

    berries from veraison to full maturity identified by Dal Santo et al. (2013a) revealed two 262

    genes common to both studies: one encoding a ubiquitin-specific protease 263

    (VIT_00s0174g00150) and the other encoding an oligo-uridylate-binding protein 264

    (VIT_04s0008g02930). These latter two genes are therefore strong candidate reference genes 265

    for the quantitative analysis of gene expression in the berry pericarp throughout ripening. 266

    267

    Red and white berry varieties express common and specific switch genes that may 268 regulate berry development 269

    We previously used integrated network analysis to compare the transcriptomic datasets of the 270

    five red-skinned berry varieties (20 samples), revealing the presence of switch genes 271

    potentially involved in the regulation of the developmental transition at veraison (Palumbo et 272

    al., 2014). The SWIM tool (Paci et al., 2017) used to unveil the switch genes in that 273

    investigation, which included the identification of fight-club hubs and the heat cartography 274

    approach (Palumbo et al., 2014; Paci et al., 2017), was therefore applied to the transcriptomic 275

    datasets of the white-skinned berries described herein. 276

    We compared the transcriptomic profiles of the white berry samples representing the 277

    herbaceous and maturation phases and identified 1,824 genes showing significant differential 278

    expression, including 1,464 downregulated and 360 upregulated after veraison 279

    (Supplemental Figure S4 and Supplemental Dataset S6). This confirms that the shift to the 280 maturation phase predominantly involves the suppression of gene expression and only a few 281

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 15

    genes are activated, as previously reported (Palumbo et al., 2014). A co-expression network 282

    comprising 1,752 nodes and 327,052 edges was generated with 1,824 DEGs using a predicted 283

    Pearson correlation threshold of 0.91 (Supplemental Figure S5A and B and Supplemental 284 Dataset S7). The specific topological proprieties of the co-expression network revealed the 285 tri-modal distribution of the average Pearson correlation coefficients (APCCs) and confirmed 286

    the identification of three types of hubs (Supplemental Figure S5C), as previously observed 287 for the red berries and for the grapevine atlas (Palumbo et al., 2014). Specifically, 1,318 genes 288

    were classified as party hubs, 55 as date hubs and 258 as fight-club hubs (Supplemental 289 Dataset S8) (Han et al., 2004). The relationship between structure and function within this 290 co-expression network was then investigated by k-means clustering analysis, allowing the 291

    identification of three modules in the network. By assigning a color scale proportional to the 292

    APCC values, we obtained a heat cartography map (Supplemental Figure S6) that reveals 293 the switch genes – i.e. genes that are more strongly linked to inversely correlated rather than 294

    positively correlated genes in the network – expressed at low levels during the herbaceous 295

    phase and at a significantly higher level during maturation. We identified 212 switch genes in 296

    the white berry varieties, representing diverse functions as observed among the switch genes 297

    in the red berry varieties (Supplemental Table S2). The comparison of switch genes in the 298 white and red varieties showed that 131 switch genes were shared by all varieties, including 299

    31 previously identified in the grapevine transcriptomic atlas (Supplemental Dataset S9). 300 The 131 common switch genes mainly belonged to the GO categories “Transcription factor 301

    activity”, “Carbohydrate metabolic process”, “Response to stress” and “Cell wall 302

    metabolism”, including several genes already known to play major roles during berry ripening 303

    (Palumbo et al., 2014). The common switch genes included ERF1 (VIT_05s0049g00510) and 304

    two other ethylene response factor genes, and an auxin-responsive gene 305

    (VIT_16s0098g01150), suggesting that ethylene and auxin play a key role in the berry 306

    developmental transition. The list of common switch genes also included 19 genes encoding 307

    transcription factors from various families (Table 1) that are likely to represent master 308 regulators of the developmental shift from the herbaceous phase to the maturation phase. 309

    Interestingly, the original list also included MYBA1, which encodes a transcription factor 310

    required for anthocyanin synthesis in red varieties. The gene is mutated and inactive in white 311

    varieties, hence the lack of anthocyanins in the berry skin. By reviewing the alignment of 312

    reads onto the reference genome, we found that MYBA2 reads were wrongly attributed to 313

    MYBA1 in the white-skinned berries, as explained in detail in the Supplemental Data. We 314 also identified genes encoding a basic helix-loop-helix (bHLH) protein, two proteins 315

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 16

    containing a lateral organ boundaries (LOB) domain that might be involved in organ 316

    development by integrating developmental changes in response to phytohormone signaling or 317

    environmental cues (Xu et al., 2016), four NAC-domain proteins and four proteins with zinc 318

    fingers. Interestingly, the list of switch genes also included AGL15a (Grimplet et al., 2016), 319

    encoding the MADS box protein Agamous-like 15a (VIT_13s0158g00100), and the 320

    transcription factor gene WRKY19 (VIT_07s0005g01710). 321

    The switch genes specific to white grape varieties included three belonging to the GO 322

    category “Transcription factor activity”, namely WRKY37 (VIT_12s0059g00880) and two 323

    zinc finger genes, and several belonging to the GO category “Response to hormone stimulus”, 324

    including the 1-aminocyclopropane-1-carboxylate oxidase (ACO) gene 325

    (VIT_11s0016g02380) which is involved in ethylene synthesis (Penrose and Glick, 1997). 326

    Further genes represented the categories “Lipid metabolic process”, e.g. the lipoxygenase 327

    gene LOXA (VIT_06s0004g01510) which is directly involved in the onset of ripening 328

    (Podolyan et al., 2010), and “Response to stress”, e.g. three thaumatin genes, the major latex-329

    like ripening protein gene grip61 (VIT_01s0011g05140) and a gene encoding an early-330

    responsive dehydration protein (VIT_02s0109g00230) (Supplemental Dataset S9). 331

    The switch genes specific to red grape varieties were mainly represented by the GO category 332

    “Secondary metabolic process”, including UFGT (VIT_16s0039g02230) and the glutathione 333

    S-transferase gene GST4 (VIT_04s0079g00690), whose absence among the white switch 334

    genes confirmed the misleading identification of MYBA1 (see above). The red switch genes 335

    also encompassed several in the category “Transcription factor activity”, including three zinc 336

    finger genes and WRKY52 (VIT_17s0000g01280) (Supplemental Dataset S9). 337

    338

    The greatest transcriptomic differences among genotypes occur during ripening 339

    The FPKM dataset representing 40 triplicate samples was investigated by PCA, revealing a 340

    clear separation among varieties at the H stage (Fig. 5A). The first two principal components 341 (PC1 and PC2) explained 44.3% of the total transcriptional variance, with PC1 explaining 342

    34.7% and PC2 the remaining 9.6%. PC1 can probably be attributed to the time course of 343

    berry development from P to H (from left to right), whereas PC2 appears to distinguish stages 344

    P and H from PV and EV. Although the berry transcriptomes from P to PV were clearly 345

    separated by PC2 regardless of the genotype, the degree of separation between transcriptomes 346

    at EV and H tended to be influenced by the genotype. This was particularly evident for the 347

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 17

    Barbera, Negro amaro and Refosco berry transcriptomes, which separated more than the other 348

    varieties during the maturation phase. 349

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 18

    This trend remained evident when PCA was applied separately to the five red-skinned 350

    varieties (Fig. 5B) and the five white-skinned varieties (Fig. 5C). The distribution of the 351 samples was almost identical during the herbaceous phase, with the P and PV samples 352

    separated by both principal components. During the maturation phase, the red and white 353

    varieties behaved differently: the red berry transcriptomes were separated by both components 354

    at the EV and H stages, with a wide distribution at H, whereas the white berry transcriptomes 355

    were tightly clustered in two groups corresponding to the EV and H stages, and were only 356

    slightly separated by PC1. 357

    Genes responsible for these color-dependent differences in behavior during the maturation 358

    phase were identified by comparing the five red and five white transcriptomes at the EV and 359

    H stages using a between-subjects t-test with an overall P-value threshold of 0.01 (Saeed et 360

    al., 2003). This revealed 443 DEGs between the red and white varieties at EV, corresponding 361

    to 382 genes more strongly expressed in the red berries and 61 more strongly expressed in the 362

    white berries (Supplemental Dataset S10). By applying a cut-off of FPKM>5 and of 363 FPKM>10 in at least one variety at EV, we found that the total number of DEGs at EV was 364

    reduced to 84% and 72%, respectively. 365

    Similarly, 837 DEGs were found at H, including 536 genes more strongly expressed in the red 366

    berries and 289 more strongly expressed in the white berries (Supplemental Dataset S10). 367 By applying a cut-off of FPKM>5 and of FPKM>10 in at least one variety at H, we found that 368

    the total number of DEGs at H was reduced to 79% and 62%, respectively. 369

    These results show that color-specific differences involve more genes specifically 370

    overexpressed in the red berries than in the white berries and that only a minor percentage of 371

    differentially expressed genes was characterized by a low expression level (FPKM ≤ 10 and 372

    >1). Furthermore, data reported in the Supplemental Dataset S10 show that genes 373 overexpressed in red berries reached much higher fold change (FC) at both maturation stages 374

    than genes overexpressed in white berries. 375

    By focusing on DEGs satisfying the condition FC > 4 at one or both stages, we identified 268 376

    genes in red berries and only 41 genes in white berries (Fig. 5D and Supplemental Dataset 377 S10). As expected, the 268 genes strongly overexpressed in red berries included several 378 structural and regulatory genes involved in the biosynthesis and transport of anthocyanins, 379

    such as GST4, UFGT, the transcriptional regulator gene MYBA1 and several genes encoding 380

    flavonoid 3',5'-hydroxylase (F3'5'H). Interestingly, we also found several genes encoding 381

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 19

    stilbene synthase (STS) and phenylalanine ammonia-lyase (PAL) among the DEGs at stage H, 382

    as well as several genes belonging to the GO categories “Transcriptional factor activity”, 383

    “Response to hormone stimulus”, “Carbohydrate metabolic process”, “Signal transduction” 384

    and “Transport” (Supplementary Dataset S10). The 41 genes strongly overexpressed in 385 white berries included several involved in photosynthesis, terpenoid biosynthesis, cell wall 386

    metabolism (e.g. three expansin genes and a cellulose synthase gene) and auxin metabolism, 387

    suggesting the developmental transition to maturation in white berries features a delay in the 388

    shutdown of vegetative metabolism. 389

    The influence of these major DEGs on the different transcriptomic behaviors observed in 390

    Figure 5A was investigated by applying PCA to the dataset of 40 triplicate samples having 391 first removed the DEGs (Fig. 5E). The removal of these genes from the dataset only 392 marginally affected the relationships among the berry transcriptomes during the maturation 393

    phase, suggesting that a larger number of genes must be involved. 394

    395

    Differences among berry transcriptomes at harvest correlate with anthocyanin 396 accumulation 397

    The relationship between samples was investigated further by applying PCA to the averaged 398

    10 berry samples at stage H. A unique principal component, describing the 27.4% of the 399

    variability, separated the Barbera, Negro amaro and Refosco berry transcriptomes from the 400

    other varieties (Supplemental Figure S7). The minimum number of genes driving the sample 401 separation in this PCA was determined by removing sets of genes step by step, each step 402

    reducing the loading value cutoff by 0.05 until the computation of a new principal component 403

    no longer provided enough variability between the transcriptomes. The threshold was 0.65/–404

    0.65, corresponding to 6,003 removed genes: 3,450 with positive loadings and 2,553 with 405

    negative loadings, i.e. positive and negative correlation to the principal component 406

    (Supplemental Dataset S11). The removal of these genes resulted in the tighter clustering of 407 berry transcriptomes during the maturation phase (Fig. 6A) and in the red berry data subset, 408 showing that the red berries at stages EV and H were slightly distinguished by PC1, and 409

    likewise the white berries (Supplemental Figure S8). On the other hand, the 6,003 genes did 410 not significantly contribute to the separation of the white berry samples during the maturation 411

    phase (Supplemental Figure S8C and D), strongly suggesting their involvement in 412 maturation programs specific to the red berry varieties, particularly Refosco, Negro amaro 413

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 20

    and Barbera. These three varieties, especially Refosco, are characterized by the rapid 414

    accumulation of anthocyanins during maturation and high anthocyanin levels at stage H (Fig. 415 1A), suggesting an association between anthocyanin accumulation and the divergence of 416 berry transcriptomes during ripening. 417

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 21

    The biological processes underlying these transcriptomic differences were determined by 418

    identifying the GO categories significantly overrepresented among the 6,003 selected genes 419

    (representing the most strongly and weakly expressed genes during maturation) in varieties 420

    Barbera, Negro amaro and Refosco (Fig. 6B and C). Among the 3,450 transcripts with a 421 loading value greater than 0.65, we found that the GO categories “Secondary metabolic 422

    process”, “Transcription” and “Catabolic process” were significantly overrepresented (Fig. 423 6B), whereas “Photosynthesis”, “Generation of precursor metabolites and energy” and 424 “Translation” were significantly overrepresented among the 2,553 genes with loading values 425

    below the threshold –0.65 (Fig. 6C). A detailed description of these results is provided in the 426 Supplemental Data. Interestingly, among the 3,450 genes strongly expressed in Barbera, 427 Negro amaro and Refosco, those with the greatest loading values belonged mainly to the 428

    categories “DNA/RNA metabolic process”, “Transcription factor activity”, “Transport” and 429

    “Carbohydrate metabolic process” (Supplemental Dataset S11). This indicates that Barbera, 430 Negro amaro and Refosco berry maturation involves a deep shift in the regulation of genes 431

    controlling many metabolic pathways in addition to anthocyanin biosynthesis, the latter being 432

    greatly enhanced compared to the varieties Sangiovese and Primitivo. 433

    In order to explore the extent to which differences in gene expression among the 10 varieties 434

    at stage H reflect the anthocyanin content regardless of the genotype, we consulted the lists of 435

    genes that are differentially expressed both when overexpressing the anthocyanin synthesis 436

    regulator MYBA1 in the white variety Chardonnay and when silencing its expression in the 437

    red variety Shiraz (Rinaldo et al., 2015). This recent study showed that the transgenic grape 438

    berries changed color either due to ectopic anthocyanin accumulation in Chardonnay or much 439

    lower anthocyanin levels in Shiraz, representing a suitable framework to investigate the effect 440

    of anthocyanin accumulation on the berry transcriptome during ripening in the same genetic 441

    background. We compared the 3,450 genes expressed most strongly in the three red varieties 442

    with the highest anthocyanin levels (loading value > 0.65) with the 787 genes upregulated in 443

    ripe Chardonnay berries overexpressing MYBA1 and the 850 genes downregulated in ripe 444

    Shiraz berries lacking MYBA1 expression (Rinaldo et al., 2015), representing the genes most 445

    likely to be regulated by anthocyanins. We found that 102 and 349 genes, representing 13% 446

    and 41% of the DEGs in Chardonnay and Shiraz, respectively, were common to our list of 447

    3,450 genes (Supplemental Dataset S11 and Supplemental Figure S9), including 26 DEGs 448 common to both Chardonnay and Shiraz. A similar proportion was found when we compared 449

    the 2,553 genes expressed most weakly in the three red varieties (loading values < –0.65) with 450

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 22

    the 478 genes downregulated in transgenic Chardonnay and the 135 genes upregulated in 451

    transgenic Shiraz. In this case, 43 Chardonnay genes and 47 Shiraz genes were shared with 452

    our list of 2,553 genes (Supplemental Figure S9), representing 9% and 35% of the DEGs in 453 Chardonnay and Shiraz, respectively, and including two DEGs common to both Chardonnay 454

    and Shiraz. The expression of these genes is probably affected by the loss of anthocyanins. 455

    The distribution of GO categories revealed that the 102 genes shared with Chardonnay berries 456

    overexpressing MYBA1 belonged mainly to the categories “Secondary metabolic process”, 457

    “Signal transduction”, “Transport”, “DNA/RNA metabolic process” and “Carbohydrate 458

    metabolic process”, whereas the 349 genes shared with Shiraz berries lacking MYBA1 459

    expression belonged predominantly to the categories “Transcription factor activity”, 460

    “DNA/RNA metabolic process”, “Transport”, “Signal transduction” and “Secondary 461

    metabolic process” (Supplemental Figure S9). These data strongly suggest that the 462 accumulation of anthocyanins may directly influence the expression of genes with a broad 463

    range of functions, especially transcriptional regulation, signal transduction and transport. 464

    465

    Genes in the phenylpropanoid/flavonoid pathway not directly controlled by MYBA 466 proteins are modulated in berries accumulating high levels of anthocyanins 467

    Variety-specific trends in the expression of phenylpropanoid pathway genes during berry 468

    development were investigated by preparing heat maps focusing separately on the herbaceous 469

    phase (Supplemental Figure S10) and the maturation phase (Figure 7A). Major differences 470 between the white and red berry varieties were clearly shown for the flavonoid structural 471

    genes and particularly for genes directly or indirectly under the control of MYBA1 and 472

    MYBA2, but differences were also evident in the early general phenylpropanoid pathway. 473

    During ripening, some genes were expressed solely in the red varieties, including most F3'5'H 474

    genes, UFGT and both anthocyanin O-methyltransferase (AOMT) genes (Figure 7A), the 475 anthocyanin acyltransferase gene 3AT (VIT_03s0017g00870), the anthoMATE3 (AM3) and 476

    GST4 genes (Supplemental Figure S10B). The expression of these genes generally peaked at 477 EV followed by a slight decline at stage H, confirming their involvement in anthocyanin 478

    biosynthesis, decoration and transport. Differences in expression pattern of these genes were 479

    found among red varieties not strictly mirroring the differences in the final anthocyanin 480

    accumulation. For example, one AOMT and the AM3 showed high expression level at EV 481

    berries in Primitivo, which is one of the low anthocyanin accumulating variety (Figure 7A 482

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 23

    and Supplemental Figure S10B). MYBA1 was strongly expressed in the maturing red berries 483 but a low level of expression was also observed in the white varieties due to the mapping 484

    errors described above and in the Supplemental Data. MYBA2 was similarly expressed in the 485 red and white varieties, although the gene is mutated in white varieties and the transcript 486

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 24

    encodes a nonfunctional MYBA2 protein (Walter et al., 2007). Genes encoding chalcone 487

    synthase (CHS), chalcone isomerase (CHI), flavonoid-3'-hydroxylase (F3'H), one favanone-3-488

    hydroxylase (F3H), dihydroflavonol reductase (DFR) and leucoanthocyanidin dioxygenase 489

    (LDOX) were expressed in both the red and white varieties albeit at higher levels in the reds 490

    (Fig. 7A). These data indicate that the expression of genes involved in flavonoid biosynthesis 491 are probably responsible for the differences between red and white varieties during berry 492

    maturation, but not for the differences in total anthocyanin levels at maturity among the five 493

    red varieties (Fig. 7B). 494

    Concerning the general phenylpropanoid pathway, several PAL, cinnamate-4-hydroxylase 495

    (C4H) and 4-coumaroyl:CoA-ligase (4CL) genes were expressed more strongly in the red 496

    varieties and this distinguished them from the white varieties during ripening (Fig. 7A). 497 Among the 15 members of the PAL gene family, only two clearly showed an expression 498

    profile similar to the anthocyanin-related genes, i.e. peaking at the EV stage. The other 13 499

    members reached a peak of expression at full ripening, mirroring the expression profile of 500

    other phenylpropanoid genes (three C4H and three 4CL genes) and in particular the 44 STS 501

    genes and their regulators MYB14 and MYB15 (Höll et al., 2013) (Fig. 7A). This behavior 502 was particularly pronounced in Refosco, characterized by the highest anthocyanin level at H 503

    (Fig. 7B). The expression of LDOX, UFGT, one PAL member (VIT_16s0039g01120) and the 504 STS27 (VIT_16s0100g00750), was validated in red varieties by RT-qPCR (Supplemental 505 Figure S11). Overall, these results highlight the coordinated strong activation of the stilbene 506 biosynthesis pathway only during the late ripening stage of red berry varieties. Interestingly, 507

    the expression of genes in this pathway reflected the differential accumulation of 508

    anthocyanins in the red berries, suggesting there may be some form of cross-talk between the 509

    activation of early phenylpropanoid genes and anthocyanin accumulation in berries (Fig. 7A 510 and B). 511

    During the herbaceous phase, the expression of phenylpropanoid-related genes could not 512

    distinguish the red and white varieties (Supplemental Figure S10A). Some of the genes 513 described above that are expressed more strongly in red grapes during ripening, are recruited 514

    in both red and white grapes during the herbaceous phase, probably for flavan-3-ol and 515

    pro-anthocyanidin biosynthesis (Boss et al., 1996a and b). Accordingly, genes encoding CHS, 516

    CHI, one F3H, DFR, LDOX, at least three PALs, one C4H, two 4CLs (Supplemental Figure 517 S10A), seven flavonol-synthases (FLS), the leuco-anthocyanidin reductases (LAR1 and 518 LAR2), the anthocyanidin reductase (ANR) (Supplemental Figure S10B) plus the regulator 519

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 25

    genes MybPA1/2 and Myb5a (Cavallini et al., 2014) are more strongly expressed at the P 520

    stage in all varieties. Interestingly, among the three PAL genes expressed at the P stage in all 521

    varieties, PAL2 and VIT_06s0004g02620 appear to be recruited for anthocyanin synthesis, 522

    whereas PAL1 appears to be related to stilbene biosynthesis in ripening red berries 523

    (Supplemental Figure S10A). 524

    525

    526

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 26

    DISCUSSION 527

    The 10 berry varieties showed diverse ripening parameters during maturation 528

    Previous investigations of berry development have focused on single varieties or a 529

    comparison of two varieties, and have considered a range of environments, which makes the 530

    genotype-specific effects more difficult to evaluate (Zenoni et al., 2010; Fasoli et al., 2012; 531

    Lijavetzky et al., 2012; Da Silva et al., 2013; Degu et al., 2014). Here we considered 10 532

    diverse berry varieties grown in the same environment, and carried out an exhaustive 533

    transcriptomic comparison at four developmental stages, focusing on traits relevant to the 534

    berry skin color (particularly the extent of anthocyanin accumulation). 535

    Five red-skinned berry varieties (Sangiovese, Barbera, Negro amaro, Refosco and Primitivo) 536

    and five white-skinned berry varieties (Vermentino, Garganega, Glera, Moscato bianco and 537

    Passerina) were selected from among several hundred Italian varieties for their agronomical 538

    and oenological diversity and their importance in the Italian wine production sector. These 539

    varieties are traditionally cultivated in different Italian regions, but in this experiment all 540

    varieties were grafted onto the same rootstock and cultivated in the same vineyard under 541

    identical conditions in order to focus on the genotype-specific effects while eliminating 542

    variation caused by the environment and agricultural practices, both of which have previously 543

    been shown to have a direct effect on berry metabolism during ripening as recently reported 544

    for the cultivars Corvina (Dal Santo et al., 2013a; Anesi et al., 2015) and Garganega (Dal 545

    Santo et al., 2016). 546

    Despite the uniform environmental conditions and agricultural practices, the 10 varieties 547

    differed substantially in terms of sugar accumulation and (in the red berry varieties) the 548

    accumulation and final content of anthocyanins. The total anthocyanin content at harvest 549

    under our experimental conditions supported previous results reported by Mattivi et al. 550

    (2006). Nevertheless, the Negro amaro berries did not achieve the color content expected for 551

    this cultivar at harvest when grown in its typical production area in Southern Italy (3.15 g/kg), 552

    indicating a slight ‘genotype x environment’ interaction effect (Mattivi et al., 2006). 553

    554

    Determination of a core berry development transcriptome 555

    The diverse ripening characteristics of the 10 varieties grown under the same environmental 556

    conditions provided a framework for the determination of common and genotype-dependent 557

    transcriptional changes during development. In all 10 varieties, the number of expressed 558

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 27

    genes declined after veraison, particularly genes with high and medium expression levels. 559

    Another common feature was that post-veraison development featured a greater number of 560

    downregulated genes than upregulated genes, with a peak of modulation between the PV and 561

    EV stages. The massive downregulation of genes during the transition from green to ripening 562

    berries has previously been reported for the variety Corvina (Palumbo et al., 2014). 563

    Clustering analysis identified 913 genes with one of six shared expression profiles among the 564

    10 varieties, probably representing the core transcriptome of berry development. Clusters 565

    showing downward expression trends during development encompassed the majority of 566

    common DEGs, reflecting the suppression of vegetative growth processes, whereas few genes 567

    representing secondary metabolism and stress responses were induced during maturation. 568

    Accordingly, we identified a much greater number of biomarkers representing the herbaceous 569

    development stages than the maturation stages (Supplemental Data), suggesting that the 570 general downregulation of genes after veraison was common to all varieties whereas the 571

    upregulation of metabolic genes during maturation tended to be more variety specific. 572

    Most of the genes in the six clusters represented physiological processes that are well known 573

    to occur during berry development. For example, many genes involved in photosynthesis, 574

    cellular component organization and the cell-cycle machinery were downregulated after stage 575

    P (cluster 1) or at the end of the herbaceous phase (cluster 3), reflecting the shutdown of 576

    photosynthesis and cell division typically observed after veraison (Dokoozlian, 2000). 577

    Likewise, the decline in berry acidity during maturation was indicated by the downregulation 578

    of genes encoding phosphoenolpyruvate carboxylase (PEPC) and L-idonate dehydrogenase, 579

    which are required for the synthesis of malic and tartaric acids, respectively (Famiani et al., 580

    2014; Steiner and McKinley, 2015). Three sucrose synthase genes (SUC1, SUS3 and SUS5) 581

    were downregulated after stage P (cluster 1) and three sugar transporter genes including the 582

    tonoplast monosaccharide transporter gene TMT3 were also downregulated during the green 583

    phase (clusters 1 and 3). Conversely, genes characterized by an expression peak at the end of 584

    veraison (cluster 4) included the hexose transporter gene HT6/TMT2, which may thus 585

    facilitate the vacuolar accumulation of hexose at the inception of ripening (Lecourieux et al., 586

    2014), and the aquaporin gene TIP1-3, which may promote berry enlargement by increasing 587

    the absorption of water. 588

    Gene modulation relating to cell wall metabolism was observed at several stages, showing 589

    that each growth event, i.e. rapid cell division and cell enlargement during the herbaceous 590

    phase, and berry expansion and softening during the maturation phase, involved distinct sets 591

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 28

    of genes. A large number of cell wall-related genes was expressed at stage P (cluster 1), 592

    including those involved in cytokinesis and xylem formation, the latter required for 593

    symplastic unloading (Jürgens, 2000; Zhang et al., 2006). Cluster 1 also contained several 594

    cellulose synthase genes, including A3 (VIT_08s0007g08380), which is required for primary 595

    cell wall synthesis (Desprez et al., 2007), and other members of subfamily A, which may be 596

    involved in secondary cell wall metabolism (Taylor et al., 1999 and 2000). 597

    The distribution of modulated expansin genes across multiple clusters indicates the 598

    involvement of expansin proteins in both vegetative growth (EXPA6, EXPA16 and EXPB2 in 599

    cluster 1) and berry expansion and softening during maturation (EXPA18 and EXPB4 in 600

    clusters 4 and 5, respectively), as previously described in Corvina berries (Dal Santo et al., 601

    2013a). Other common transcriptional modulations appear to regulate the xyloglucan 602

    components of the hemicellulose matrix, which may control morphogenesis (Xiao et al., 603

    2016). Four xyloglucan endotransglucosylase/hydrolase genes were commonly upregulated at 604

    stage P, suggesting a role in cell wall synthesis during cell division, whereas another was 605

    commonly upregulated during maturation (cluster 6), suggesting a role in cell wall loosening 606

    and tissue softening (Nishitani and Vissenberg, 2006). 607

    Berry softening is a crucial physiological event marking the onset of ripening, together with 608

    sugar accumulation and color development. A delay in berry softening was recently 609

    associated with the inhibition of sugar accumulation and anthocyanin biosynthesis 610

    (Castellarin et al., 2016). Softening and a loss of turgor are early events in the ripening 611

    process and may even commence before veraison (Coombe and McCarthy, 2000). We 612

    therefore focused on genes upregulated during the late stages of the herbaceous development 613

    phase, and identified two cellulose synthase-like E1 (CSLE1) genes, which synthesize the 614

    backbones of non-cellulosic polysaccharides (hemicelluloses), and a pectin methylesterase 615

    inhibitor (PMEI), which has previously been shown to control pectin methylesterase activity 616

    in tomato at the post-transcriptional level (Di Matteo et al., 2005). The grapevine PMEI1 617

    gene, which is strongly expressed during green berry growth, was recently proposed to inhibit 618

    pectin methylesterase activity and thus prevent premature berry softening related to pectin 619

    degradation (Lionetti et al., 2015). These genes may thus represent the triggers for ripening 620

    by regulating the initial events at veraison, such as softening and loss of turgor (Castellarin et 621

    al., 2016). 622

    The common berry development transcriptome also featured genes highlighting the interplay 623

    between auxin and ethylene signaling during ripening (Giovannoni et al., 2001 Böttcher and 624

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 29

    Davies, 2012). The substantial downregulation of 10 auxin-related genes after stage P (cluster 625

    1) and 7 genes after stage PV (cluster 3) correlated with the use of indole-3-acetic acid (IAA) 626

    before maturation to delay ripening (Böttcher et al., 2010; 2013; Cawthon and Morris, 1982). 627

    Although auxin biosynthesis may reach a peak at the initiation of ripening, the IAA is 628

    sequestered into IAA-Asp conjugates that accumulate rapidly at veraison and remain at the 629

    same level throughout ripening (Böttcher et al., 2013). The hypothesis of active auxin 630

    biosynthesis during berry development is supported by the upregulation of 10 genes involved 631

    in auxin metabolism (Supplemental Dataset S4) and an auxin efflux carrier protein at stage 632 PV (cluster 2), and of an auxin-independent growth promoter and an indole-3-acetate 633

    β-glucosyltransferase after stage PV (cluster 5). The upregulation of an indole-3-acetate 634

    β-glucosyltransferase, putatively involved in auxin inactivation by the IAA sugar conjugation, 635

    could take part to the mechanisms by which the endogenous IAA concentration decreased and 636

    was maintained at low levels in maturing fruit (Bottcher et al., 2010). No genes involved in 637

    IAA catabolism or amino acid conjugation were identified, possibly due to the incomplete 638

    functional annotation of the predicted grapevine genes or because a brief period of 639

    transcriptional modulation featuring these genes was not covered by our samples. 640

    Grape berries are non-climacteric fruits so there is no sharp increase in ethylene production 641

    and respiratory activity at the onset of ripening (Böttcher et al., 2013). Nevertheless, the 642

    ethylene concentration peaked before the initiation of ripening (Chervin et al., 2004), and 643

    ethylene-releasing compounds can influence the ripening time and the auxin concentration 644

    (Böttcher et al., 2013). The proposed cross-talk between ethylene and auxin signaling in fruit 645

    ripening has also been supported by the observation that auxin depletion in apple can induce 646

    the expression of genes required for ethylene biosynthesis (Shin et al. 2016) whereas the 647

    inhibition of ethylene signaling can induce the expression of genes related to auxin signaling 648

    (Tadiello et al., 2016). The downregulation of the ethylene-responsive transcription factor 649

    genes SHINE3 and ERF042 after stage P (cluster 1) and the peak induction of an ethylene 650

    responsive protein gene at stage PV (cluster 2) were consistent with a regulatory role for 651

    ethylene in berry development. 652

    Common genes in clusters 5 and 6 showed an increase in expression after veraison and the 653

    GO categories “Transcription factor activity” and “DNA/RNA metabolic process” were 654

    prevalent, suggesting a role in the transcriptomic reprograming that accompanies maturation. 655

    Three NAC genes were identified: NAC03, a close homologue of LeNOR, one of the master 656

    regulators of fruit ripening in tomato (Martel et al., 2011); NAC11, a berry switch gene 657

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 30

    probably involved in the transition from vegetative growth to maturation (Palumbo et al., 658

    2014); and NAC37, which may be involved in the regulation of secondary cell wall biogenesis 659

    (Wang et al., 2013). The upregulation of WRKY37, another switch gene in white berries and a 660

    homolog of AtWRKY6, suggests the induction of cell death-related processes (Robatzek and 661

    Somssich, 2001), including the loss of cell vitality observed during the advanced stages of 662

    grape berry maturation (Tilbrook and Tyerman, 2008). 663

    The core set of 19 switch genes encoding transcription factors among the genes commonly 664

    upregulated in all 10 varieties after veraison represents the putative regulators of the shift to 665

    maturity in the core berry development transcriptome. These genes are candidates for further 666

    functional studies aiming to dissect the molecular mechanisms that govern the important 667

    physiological changes that occur during the onset of ripening. 668

    669

    Transcriptomic behavior during maturation correlates with the anthocyanin content 670

    PCA applied to the entire dataset emphasized the deep shift in the berry transcriptome 671

    between the herbaceous and maturation phases, as well as the separation between white and 672

    red berry varieties after veraison, mostly evident at stage H. Barbera, Negro amaro and 673

    particularly Refosco transcriptomes featured a more advanced ripening state described by 674

    PC2, reflecting the particularly high levels of sugars and anthocyanins in these varieties (Fig. 675 4B). In contrast, the white berry genotypes showed no evidence of such enhanced 676 transcriptomic changes after veraison, and seemed to be more transcriptionally active during 677

    the herbaceous development phase (Fig. 4C). It is likely that the transcriptomic diversity 678 among the red berry varieties can be attributed to the generally more advanced ripening at 679

    stage H compared to the white berry varieties. However, the final Brix degrees (Fig. 1C) 680 indicate that Moscato bianco behaved more similarly to Barbera, Negro amaro and Refosco 681

    than to the other four white varieties, even though the Moscato bianco berry transcriptome at 682

    stage H was more similar to the white varieties than the red ones (Fig. 4A). These data 683 suggest that the sugar concentration, a very important technological ripening parameter for 684

    winemakers, only partially indicates the physiological stage of berry ripening revealed by 685

    transcriptome analysis. The extent of anthocyanin accumulation appeared to be more directly 686

    related to the transcriptomic program of fruit maturation. With this regard, the highest 687

    anthocyanin accumulation of Refosco at H fully typifies this behavior being associated with a 688

    distinctive transcriptomic trend and a marked induction of early phenypropanoid- and 689

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 31

    stilbenoid-related genes during maturation. 690

    It is well known that the red color of grapes reflects the biosynthesis and accumulation of 691

    anthocyanins (generally in the skin). The transcription factors MYBA1 and MYBA2 692

    additively regulate the structural genes for anthocyanin biosynthesis, and in white cultivars 693

    this is prevented by a double mutation that inactivates both proteins (Kobayashi et al., 2004, 694

    Walker et al., 2007). When comparing the two color groups at both maturation stages, we 695

    found that most of the DEGs were preferentially expressed in red berries, and represented the 696

    phenylpropanoid/flavonoid pathway, anthocyanin transport and anthocyanin modification. 697

    Nevertheless, several other DEGs were preferentially expressed in red berries, including 698

    genes involved in carbohydrate metabolism, such as the cell wall apoplastic invertase gene 699

    cwINV4, the neutralization of reactive oxygen species, such as the gene encoding ascorbate 700

    oxidase (Pilati et al., 2014), and abiotic stress responses, such as 10 WRKY genes (Liu et al., 701

    2011; Wang et al., 2014) and the early responsive dehydration protein 15 gene (ERD15). The 702

    DEGs preferentially expressed in white berries were mostly involved in photosynthesis, 703

    strongly suggesting the presence of residual photosynthetic activity at full maturity, but also 704

    included genes responsible for the aroma traits specific to white berries, such as the two 705

    terpene synthase genes TPS55 and TPS59 (Martin et al., 2010) (Fig. 8). Other DEGs 706 expressed in white berries encoded a multidrug and toxic compound extrusion (MATE) 707

    protein, which probably transports secondary metabolites to the vacuole (Zhao et al., 2011; 708

    Marinova et al., 2007), a nitrate transporter, and a major facilitator superfamily protein 709

    (MFS), possibly required for the production of volatile thiol precursors that characterize white 710

    berry varieties (Tominaga et al., 2000) (Fig. 8). Notably, when the DEGs distinguishing red 711 and white berry varieties were removed, the transcriptomic differences among berry samples 712

    were still evident, suggesting that genotypes are distinguished by profound transcriptomic 713

    differences affecting many processes in addition to anthocyanin accumulation and related 714

    MYBA functions, as previously reported in high-anthocyanin tomatoes (Povero et al., 2011). 715

    When PCA was applied to the entire transcriptome dataset, it was clear that the major 716

    contribution to the separation of samples at stage H was provided by the red berry genotypes. 717

    We therefore focused on those samples and found that 6,003 transcripts were needed to 718

    maintain the separation among the red varieties, confirming that a substantial proportion of 719

    the berry transcriptome changes during ripening, particularly in the varieties Barbera, Negro 720

    amaro and Refosco, and that the transcriptional rearrangement may be related to anthocyanin 721

    content. The 6,003 selected genes included many involved in carbohydrate metabolism, 722

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 32

    relating to the differences in sugar content among the genotypes. The HT5 and INV4 genes 723

    encoding a hexose transporter and a cell wall invertase, which were identified as DEGs when 724

    searching for differences between the red and white varieties, were particularly strongly 725

    expressed in Barbera, Negro amaro and Refosco. In contrast, the HT2, HT3/HT7 and 726

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 33

    HT6/TMT2 genes were expressed at higher levels in all the other grapes. Many stress response 727

    genes were also strongly expressed in Barbera, Negro amaro and Refosco berries, including 728

    37 STS genes, 7 jasmonate ZIM-domain (JAZ/TIFY) genes, and several genes encoding 729

    transcription factors or related to transport, signal transduction and secondary metabolism 730

    (Supplemental Dataset S11). 731

    In order to determine whether these diverse functions could be linked to the anthocyanin 732

    content indirectly (i.e. not through the direct activity of MYBA transcription factors) and in a 733

    variety-independent manner, we compared the 6,003 selected genes with two external gene 734

    sets: genes modulated by the ectopic expression of MYBA1 in the white variety Chardonnay, 735

    resulting in a red phenotype, and genes modulated by the silencing of MYBA1 in the red 736

    variety Shiraz, resulting in a white phenotype (Rinaldo et al., 2015). The ectopic expression 737

    of MYBA1 in Chardonnay driven by the CaMV 35S promoter caused this factor to be 738

    expressed also in pulp, which is not directly comparable with our experiment. In contrast, 739

    MYBA1 silencing in Shiraz represented a more suitable framework for our investigation: a 740

    large number of shared DEGs were identified during ripening in the transgenic Shiraz berries 741

    and the Barbera, Negro amaro and Refosco varieties. The commonly modulated genes were 742

    mostly related to carbohydrate metabolism, transcriptional activity and transport, suggesting 743

    that anthocyanin levels may influence many other processes (Fig. 8). 744

    The extent of anthocyanin accumulation in the berry skin might correlate with its opacity to 745

    sunlight, profoundly influencing the transcriptomic changes during ripening. This is supported 746

    by the preferential expression of several genes related to photosynthesis or the response to 747

    visible and UV light in the white berries (and in red berries with low levels of anthocyanins) 748

    such as HY5, encoding a bHLH transcription factor that mediates the promotion of flavonol 749

    accumulation in the berry in response to UV-B (Loyola et al., 2016), and TPS genes that 750

    promote terpenoid accumulation in response to light (Friedel et al., 2016). Anthocyanin levels 751

    could also influence other metabolic pathways, similar to the way in which sugars act as 752

    signals to promote anthocyanin accumulation (Dai et al., 2014). Anthocyanin biosynthesis 753

    could also be part of a more general maturation process controlled by unidentified master 754

    regulators that are more active in Refosco, Barbera and Negro amaro than in the other 755

    varieties. 756

    757

    Stilbenoid biosynthesis may influence anthocyanin accumulation during maturation 758

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 34

    Our survey provided a comprehensive profile of the phenylpropanoid/flavonoid biosynthesis 759

    pathway in 10 varieties. The quantity and composition of flavonoids differs among grape 760

    varieties and contributes to their distinctive traits (Mattivi et al., 2006). Indeed, differences in 761

    the expression pattern of genes involved in anthocyanin decoration were found among red 762

    varieties, suggesting differences in final anthocyanin profiles. The very high expression level 763

    of one isoform of AOMT in Primitivo berries matched the high ratio of methoxyl to hydroxyl 764

    anthocyanin forms reported in this variety; likewise, the low 3AT expression level in 765

    Sangiovese berries corroborated the low acylated anthocyanins content in this cultivar 766

    (Mattivi et al., 2006). 767

    As well as distinguishing white and red cultivars, the total anthocyanin content was much 768

    higher in Refosco, Barbera, and Negro amaro berries compared to Primitivo and especially 769

    Sangiovese. The expression profiles of genes involved in flavonoid biosynthesis mirrored 770

    these results: the entire phenylpropanoid pathway was more transcriptionally active in 771

    Refosco, Negro amaro, and Barbera berries than Sangiovese and Primitivo, even at the early 772

    steps catalyzed by PAL, C4H and 4CL (Figure 7A). In particular, we observed diverse 773 expression profiles among the PAL gene family, characterizing the P stage in both white and 774

    red berries (Supplemental Figure S10) and the maturation phase (and specifically stage H) in 775 the red varieties (Fig. 7A). 776

    These diverse expression profiles suggest that individual PAL genes serve phase-specific 777

    functions, as already shown for the PAL gene family in the grapevine expression atlas (Fasoli 778

    et al., 2012). Interestingly, the expression of PAL genes at stage H matched the expression 779

    profile of 44 STS genes and the related transcriptional regulators MYB14 and MYB15 (Höll 780

    et al., 2013), suggesting that the transcriptional activation of specific members of the PAL 781

    gene family diverts flux towards stilbene synthesis. Stilbenes accumulate strongly in response 782

    to a wide range of biotic and abiotic stresses (Versari et al., 2001; Vannozzi et al., 2011), and 783

    are also part of the late ripening program in healthy unstressed berries (Gatto et al., 2008). 784

    Red berries preferentially expressed genes involved in the neutralization of reactive oxygen 785

    species and in abiotic stress responses, which correlates with the transcriptional activation of 786

    the stilbenoid pathway in different varieties. Varietal-specific stilbene profiles have been 787

    reported previously (Gatto et al., 2008; Vincenzi et al., 2013; Zenoni et al., 2016) suggesting 788

    that the transcriptional activation of the stilbenoid pathway in Barbera, Negro amaro and 789

    Refosco may be specific to the cultivar. 790

    Although Refosco, Barbera, and Negro amaro accumulated more anthocyanins than the other 791

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 35

    varieties, this was only slightly reflected by the expression of genes directly involved in 792

    anthocyanin biosynthesis such as UFGT and MYBA1/MYBA2 (Fig. 7A and Supplemental 793 Figure S11). The stilbene and flavonoid pathways branch from the general phenylpropanoid 794 pathway and thus share the same precursors. Therefore, whereas the varietal-specific 795

    reinforcement of the early steps in the general phenylpropanoid pathway may channel 796

    precursors into the stilbene branch, the flavonoid branch may benefit indirectly, resulting in 797

    the higher accumulation of anthocyanins. Nevertheless, the possibility that differences in 798

    anthocyanin accumulation may also rely on different biosynthetic enzyme activities and/or 799

    anthocyanin turnover cannot be ruled out. It was recently proposed that the stability of the 800

    CHS, that catalyzes the first step of flavonoid biosynthesis, is controlled by a proteolytic 801

    regulator in response to developmental cues and environmental stimuli (Zhang et al., 2017), 802

    unveiling a possible post-translational mechanism controlling flavonoid biosynthesis. 803

    804

    CONCLUSION 805

    We have investigated the core transcriptome of grape berry development and have also 806

    identified variety-specific processes that lead to the diverse characteristics of different 807

    grapevine genotypes. The core berry development transcriptome, comprising the genes with 808

    similar transcriptional trends across all 10 varieties, emphasized that the transition from 809

    vegetative growth to maturation relies more on the downregulation of genes associated with 810

    vegetative growth than on the induction of ripening-specific traits. This core transcriptional 811

    program provides a solid basis for the scientific community, allowing the functional 812

    characterization of genes involved in the onset of ripening. The comparison of red and white 813

    berries revealed substantial transcriptomic variation among the red cultivars during ripening, 814

    correlating with differences in anthocyanin accumulation, suggesting that color development 815

    affects the maturation process itself by triggering the transcriptional reprogramming of 816

    several biological processes. This hypothesis helps to explain the differences between red and 817

    white cultivars but also the differences among red berry varieties. 818

    819

    MATERIALS AND METHODS 820

    Plant material 821

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 36

    Berries were collected during the 2011 season from 10 grapevine (Vitis vinifera L.) cultivars: 822

    Sangiovese, Barbera, Negro amaro, Refosco dal peduncolo rosso, Primitivo, Vermentino, 823

    Garganega, Glera, Moscato bianco and Passerina. The 10 varieties were cultivated in the same 824

    vineyard at the CREA-VE grapevine germplasm collection (Susegana, Veneto region, Italy). 825

    All vines were grafted onto SO4 rootstock, had the same age and were cultivated on 826

    homogeneous soil with the same training system and cultural practices (soil management, 827

    irrigation, fertilization, pruning, and disease control). 828

    Samples of berries were harvested at four developmental stages: pea-sized berry stage (Bbch 829

    75) at 20 days after flowering (P), ‘berries beginning to touch’ stage (Bbch77) just prior to 830

    veraison (PV), berry softening stage (Bbch 85) at the end of veraison (EV) and fully-ripe 831

    berry stage (Bbch 89) at harvest (H) (Supplemental Table S3). Each sample consisted of at 832 least 100 berries collected from five vines of each variety block, from both sides of the 833

    canopy and from different areas of the clusters. Sampling at stage H considered the 834

    commercial Brix degree value of each variety but also the sanitary status of the berries (i.e. 835

    clusters were harvested before the first signs of rotting). The berries were collected at the 836

    same time of day (9–10 am), immediately frozen in liquid nitrogen and stored at –80 ºC. From 837

    each berry sample, three biological replicates were prepared. The Brix degree value of the 838

    must was measured using a digital DBR35 refractometer (Giorgio Bormac, Carpi, Italy). 839

    840

    RNA extraction 841

    Total RNA was extracted from approximately 400 mg of berry pericarp tissue (entire berries 842

    without seeds) ground in liquid nitrogen, using the Spectrum™ Plant Total RNA kit (Sigma-843

    Aldrich, St. Louis, MO, USA) with some modifications (Fasoli et al., 2012). RNA quality and 844

    quantity were determined using a Nanodrop 2000 spectrophotometer (Thermo Fisher 845

    Scientific, Waltham, MA, USA) and a Bioanalyzer Chip RNA 7500 series II (Agilent 846

    Technologies, Santa Clara, CA, USA). 847

    848

    Library preparation and sequencing 849

    The 40 triplicate samples (10 varieties at four stages) yielded 120 non-directional cDNA 850

    libraries, which were prepared from 2.5 µg of total RNA using the Illumina TruSeq RNA 851

    Sample preparation protocol (Illumina Inc., San Diego, CA, USA). Library quality was 852

    determined using a Bioanalyzer Chip DNA 1000 (Agilent). Single-end reads of 100 853

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 37

    nucleotides were obtained using an Illumina Hiseq 1000 sequencer, and sequencing data were 854

    generated using the base-calling software Illumina Casava v1.8 (33,438,747 ± 5,610,087 855

    reads per sample). 856

    857

    Sequencing data analysis 858

    The reads were aligned onto the Vitis vinifera 12x reference genome PN40024 (Jaillon et al., 859

    2007) using TopHat v2.0.6 (Kim et al., 2013) with default parameters. An average of 85.55% 860

    of reads was mapped for each sample (Supplemental Table S1). Mapped reads were used to 861 reconstruct the transcripts in Cufflinks v2.0.2 (Roberts et al., 2011) and the reference genome 862

    annotation V1 (http://genomes.cribi.unipd.it/DATA). All reconstructed transcripts were 863

    merged into a single non-redundant list of 29,287 transcripts using Cuffmerge, which merges 864

    transcripts whose reads overlap and share a similar exon structure thus generating a longer 865

    chain of connected exons. The gene ID is named as a row of the merged transcripts. Using 866

    this novel list of transcripts, the normalized averaged expression of each transcript was 867

    calculated as an FPKM value for each triplicate using the geometric normalization method. In 868

    total, 26,807 genes were detected as expressed, i.e. the averaged FPKM value was > 0 in at 869

    least one sample. Displaying the FPKM distribution of these 26,807 genes across the 40 870

    samples (Supplemental Figure S12) showed a bimodal distribution as observed by 871 Hebenstreit et al. (2011) and Nagaraj et al. (2011). Separating FPKM values in two groups: 872

    FPKM < 1 and FPKM > 1 enables to fit the main distribution centered at a value of ∼4 log2 873

    FPKM, as described by Hebenstreit et al. (2011). Therefore, genes with an expression 874

    intensity < 1 FPKM across the 40 samples (roughly corresponding to 1 mRNA per average 875

    cell in the sample) were classified as minimally expressed genes and were removed from the 876

    gene expression dataset. 877

    The remaining dataset of 21,746 transcripts was used in the subsequent experiments. DEGs 878

    were identified by comparing each pair of consecutive developmental stages (i.e. P-PV, PV-879

    EV, and EV-H) using Cuffdiff v2.0.2 (Trapnell et al., 2012) with a false discovery rate 880

    defined by Benjamini-Hochberg multiple tests of 5%. 881

    882

    Correlation analysis 883

    A correlation matrix was prepared using R software and Pearson’s correlation coefficient as 884

    the statistical metric in order to compare the 40 transcriptomes. The analysis was performed 885

    www.plantphysiol.orgon March 24, 2020 - Published by Downloaded from Copyright © 2017 American Society of Plant Biologists. All rights reserved.

    http://www.plantphysiol.org

  • 38

    using the log2 (FPKM average + 1) of the 21,746 transcripts considered to be expressed 886

    (Vasudevan et al., 2015). Correlation values were converted into distance coefficients to 887

    define the height scale of the dendrogram. 888

    889

    PCA and loading normalization 890

    PCAs were carried out using SIMCA P+ v13.0 (Umetrics, Umeå, Sweden). The PCA applied 891

    to the ripe berry transcriptomes was carried out using the 10 stage H samples and the 892

    corresponding 21,746 gene expression values. Loadings, which represent new values of the 893

    variables (genes) in the model plane, were extracted and normalized into linear correlation 894

    coefficients between the genes and the principal component, by applying the following 895

    formula: 896 × × 897 with N as the total number of genes involved in the PCA, e.g. 21,231 genes in this case. In 898

    fact, 515 genes were excluded from the analysis due to the low number of samples (fewer 899

    than two) yielding a