comparative genetic and metabolic characterization between

168
HAL Id: tel-01809532 https://tel.archives-ouvertes.fr/tel-01809532 Submitted on 6 Jun 2018 HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers. L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés. Comparative genetic and metabolic characterization between two table grape varieties with contrasted color berry skin : red Globe and Chimenti Globe Claudia Santibanez To cite this version: Claudia Santibanez. Comparative genetic and metabolic characterization between two table grape varieties with contrasted color berry skin : red Globe and Chimenti Globe. Agricultural sciences. Université de Bordeaux; Pontificia universidad católica de Chile (Santiago de Chile), 2017. English. NNT : 2017BORD0918. tel-01809532

Upload: others

Post on 20-Jul-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Comparative genetic and metabolic characterization between

HAL Id: tel-01809532https://tel.archives-ouvertes.fr/tel-01809532

Submitted on 6 Jun 2018

HAL is a multi-disciplinary open accessarchive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come fromteaching and research institutions in France orabroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, estdestinée au dépôt et à la diffusion de documentsscientifiques de niveau recherche, publiés ou non,émanant des établissements d’enseignement et derecherche français ou étrangers, des laboratoirespublics ou privés.

Comparative genetic and metabolic characterizationbetween two table grape varieties with contrasted color

berry skin : red Globe and Chimenti GlobeClaudia Santibanez

To cite this version:Claudia Santibanez. Comparative genetic and metabolic characterization between two table grapevarieties with contrasted color berry skin : red Globe and Chimenti Globe. Agricultural sciences.Université de Bordeaux; Pontificia universidad católica de Chile (Santiago de Chile), 2017. English.�NNT : 2017BORD0918�. �tel-01809532�

Page 2: Comparative genetic and metabolic characterization between

0

THÈSE EN COTUTELLE PRÉSENTÉE

POUR OBTENIR LE GRADE DE

DOCTEUR DE

L’UNIVERSITÉ DE BORDEAUX

ET DE

LA PONTIFICIA UNIVERSIDAD CATÓLICA DE CHILE

ÉCOLE DOCTORALE DE SCIENCES DE LA VIE ET LA SANTÉ

SPECIALITÉ : BIOLOGIE VÉGÉTALE

ET DU

DOCTORAT EN SCIENCES BIOLOGIQUES

SPECIALITÉ : GENETIQUE MOLECULAIRE ET MICROBIOLOGIE

Par Claudia SANTIBAÑEZ ORELLANA

COMPARATIVE GENETIC AND METABOLIC CHARACTERIZATION BETWEEN

TWO TABLE GRAPE VARIETIES WITH CONTRASTED COLOR BERRY SKIN:

RED GLOBE AND CHIMENTI GLOBE.

Sous la direction de Eric GOMÈS

et de Patricio ARCE JOHNSON

Soutenue le 18 de décembre 2017

Membres du jury : M. GUTIERREZ, Rodrigo, Professeur à La Pontificia Universidad de Chile Président

M. HANFORD, Michael, Professeur à L’Universidad de Chile Rapporteur

M. INOSTROZA, Claudio, Chargé de Recherche à L’ Universidad Catolica de Temuco Rapporteur

M. GOMÈS, Eric, Professeur de l’Université de Bordeaux Examinateur

M. ARCE, Patricio, Professeur à La Pontificia Universidad de Chile Examinateur

M. DELROT, Serge, Professeur de l’Université de Bordeaux Examinateur

Mme MEISEL, Lee, Professeur à L’ Universidad de Chile Invité

Page 3: Comparative genetic and metabolic characterization between

1

Comparative genetic and metabolic characterization between two table grape

varieties with contrasted color berry skin: Red Globe and Chimenti Globe

Abstract Grape berry development is a dynamic process exhibiting a double sigmoid curve separated by

a lag phase, characterized by coordinated biosynthesis of primary and secondary metabolites.

At the end of lag phase, a phenomenon called veraison occurs, where the fruit begins to take

color and maturation process is initiated. Anthocyanins are the responsible of berry skin

coloration and its regulation has been widely studied. However, few studies have focused on

metabolic and genetic characterization using varieties with contrasted color berry skin. Using

RNA-seq technology and metabolic analysis, we performed a new comparative characterization

of two table grapes, Chimenti Globe (CG) and Red Globe (RG), with contrasted color berry

skin: CG has a bright light red color and RG has a dark purple color. The originality of this

model is that CG was generated from a spontaneous event in field from a branch in a vine of

RG plant. Thus, the genetic background responsible for the color change is practically the same.

Berry metabolic content analysis showed the importance of veraison and ripening stages, in both

varieties in the study. Particularly, striking difference in metabolites concentration of

phenylpropanoid pathway: shikimate, UDP-glucose, phenylalanine and trehalose-6-phosphate,

among others. Also, differences between the varieties were due to changes in metabolites related

to the biosynthesis of sucrose and anthocyanin. CG only contained dihydroxylated

anthocyanins, peonidin and cyanidin, and not the trihydroxylated ones, malvidin, delphinidin

and petunidin, which was consistent with the observed color phenotype. From transcriptomic

analysis, we generated a heat map with 109 genes differentially expressed in CG in comparison

to RG, being many of these related to the flavonoid pathway. In addition, 11 gene copies of

flavonoid 3’5’-hydroxylase, a key enzyme for biosynthesis of trihydroxylated anthocyanins,

were not induced in CG. From this analysis, we selected a candidate gene to contribute in the

study of the anthocyanin biosynthetic pathway: Cytochrome b5 (Cytb5) encoding a key electron

donor protein not characterized in grapevine. Cytb5 overexpression in V. berlandieri x V.

Rupestris 110R embryos strongly suggested its participation in the pathway, since transgenic

embryos exhibited reddish color and in addition, an accelerated development compared to

control. With these results, we were able to provide new insights in anthocyanins regulation in

grapevines responsible for berry skin coloration, opening new fields in the search for molecular

regulators of the flavonoid pathway.

Keywords : Chimenti Globe, Red Globe, RNA-seq, skin, anthocyanin, cytb5.

Page 4: Comparative genetic and metabolic characterization between

2

Analyse génétique et métabolique comparée de deux variétés de raisins de

table présentant des pigmentations contrastées au niveau des pellicules des

baies : Red Globe et Chimenti Globe.

Résumé Le développement de la baie de raisin est un processus dynamique présentant une courbe de

croissance sigmoïde avec deux phases de croissance séparée par une phase de latence. Il se

caractérise par une biosynthèse coordonnée de métabolites primaires et secondaires. À la fin de

la phase de latence, un phénomène appelé véraison se produit, au cours duquel le fruit

commence à prendre de la couleur et le processus de maturation est initié. Les anthocyanes sont

responsables de la coloration des baies et la régulation de leur biosynthèse été largement étudiée.

Cependant, peu d'études ont porté sur la caractérisation métabolique et génétique de variétés de

baies de couleurs fortement contrastées, dans un même fond génétique. En combinant la

technologie RNAseq et des analyses métabolomiques, nous avons effectué une caractérisation

comparée de deux variétés de raisin de table : Chimenti Globe (CG, rose pâle) et Red Globe

(RG, rouge foncé). L'originalité de ce modèle est que CG a été généré à partir d'un événement

de mutation spontanée de RG, dans un vignoble de production, permettant ainsi l’étude de la

régulation de la biosynthèse des anthocyanes dans un même fond génétique.

L'analyse du contenu métabolique des baies a démontré l'importance des stades de

développement, de la véraison à la maturation, dans les deux variétés de raisin. En particulier,

des différences marquées dans les concentrations de certains métabolites de la voie des

phénylpropanoïdes (shikimate, UDP-glucose, phénylalanine) et du tréhalose-6-phosphate ont

été mises en évidence en post-véraison. De plus, les différences entre les variétés étaient dues à

des changements dans les métabolites liés à la biosynthèse du saccharose et de l'anthocyanine.

Les baies de CG ne contenaient que des anthocyanines dihydroxylées (péonidine et cyanidine),

et aucune quantité détectable d’anthocyanes trihydroxylées (malvidine, delphinidine et

pétunidine), qui sont abondamment présentes dans les pellicules des baies de RG. Ceci explique

le phénotype de couleur rose pâle des baies de CG. Une analyse transcriptomique globale par

RNAseq montre que 109 gènes sont exprimés de façon différentielle chez CG, en comparaison

avec RG, y compris de nombreux gènes liés au métabolisme des flavonoïdes. Notamment, 11

gènes codant pour des 3'5'-hydroxylases flavonoïde, une enzyme-clé pour la biosynthèse des

anthocyanes trihydroxylées, sont réprimés dans CG. A partir de cette analyse, un gène candidat

pour la régulation de la voie de biosynthèse des anthocyanes a été sélectionné : le gène Cytb5,

qui code pour un cytochrome b5, non caractérisé à ce jour chez la Vigne. La surexpression de

Cytb5 par transgénèse dans des vignes hybrides V. berlandieri x V. rupestris cv. 110R suggère

fortement un rôle clé pour ce gène dans la régulation de la biosynthèse des anthocyanes chez la

vigne : les embryons obtenus présentaient une forte coloration rouge, indiquant la présence

d’anthocyanes en quantités importantes dans les tissus végétatifs. De plus les embryons

transgéniques ont un développement accéléré par rapport aux embryons sauvages.

Ces travaux ont permis de mieux comprendre la régulation de l’accumulation des anthocyanes

responsables de la coloration des baies de raisin, et plus largement, la régulation du métabolisme

des flavonoïdes.

Mots clés : Chimenti Globe, Red Globe, RNA-seq, pellicule, anthocyane, cytb5.

Page 5: Comparative genetic and metabolic characterization between

3

Caracterización genética y metabólica comparativa entre dos tipos de uva

de mesa con color de piel contrastante: Red Globe y Chimenti Globe

Resumen El desarrollo de la uva es un proceso dinámico caracterizado por una curva de crecimiento doble

sigmoidea, separada por una fase lag, en donde ocurre una biosíntesis coordinada de metabolitos

primarios y secundarios. Al final de la fase lag, ocurre un fenómeno llamado pinta

correspondiente al comienzo de la coloración de la baya e iniciándose también el proceso de

maduración. Las antocianinas son las responsables de la coloración de la piel de las bayas y su

regulación ha sido ampliamente estudiada. Sin embargo, pocos estudios se han enfocado en la

caracterización genética y metabólica, utilizando variedades contrastantes de color de piel.

Utilizando análisis metabólico y la tecnología de RNA-seq, se realizó una nueva caracterización

comparativa de dos uvas de mesa, Chimenti Globe (CG) y Red Globe (RG), que poseen un color

de piel de la baya contrastante: CG tiene un color rojizo claro y RG posee un color morado. La

originalidad de este modelo es que CG fue generada en un evento espontáneo de campo desde

una rama de una planta RG. Por lo tanto, el background genético responsable del cambio de

color es el mismo.

El análisis del contenido metabólico de las pieles de las bayas reveló la importancia de las etapas

de desarrollo, pinta y maduración, en ambas variedades en estudio. En particular, la diferencia

en la concentración de metabolitos de la ruta fenilpropanoide, tales como shikimato y

fenilalanina y otras moléculas como UDP-glucosa y trehalosa-6-fosfato, entre otros. Asimismo,

las diferencias entre las variedades estuvieron dadas por cambios relacionados con la biosíntesis

de sacarosa y antocianinas. CG solo contenía antocianinas dihidroxiladas, cianidina y peonidina,

y no las del tipo trihidroxiladas, malvidina, delfinidina y petunidina, lo cual fue consistente con

el fenotipo del color de piel observado. A partir del análisis transcriptómico, generamos un

heatmap con 109 genes expresado diferencialmente en CG en comparación con RG, siendo

muchos de estos asociados a la ruta biosíntesis de flavonoides. Además, observamos que 11

copias del gen flavonoide 3'5'-hidroxilasa, que codifica una enzima clave para la biosíntesis de

antocianinas trihidroxiladas, no estaban inducidas en CG. A partir de este análisis, se seleccionó

un gen candidato para contribuir en el estudio de la ruta de biosíntesis de antocianinas:

citocromo b5 (Cytb5) que codifica una proteína clave donadora de electrones no caracterizada

en vides. La sobreexpresión de Cytb5 en embriones V. berlandieri x V. rupestris cv. 110R sugirió

fuertemente la participación en la ruta, ya que los embriones transgénicos exhibieron un color

rojizo e incluso, un desarrollo acelerado en comparación con el control. Con estos resultados,

hemos sido capaces de proporcionar información sobre la regulación de las antocianinas en

vides responsables de la coloración de la piel de las bayas, abriendo nuevos terrenos en la

búsqueda de reguladores moleculares de la vía flavonoide.

Palabras clave: Chimenti Globe, Red Globe, RNA-seq, piel, antocianinas, cytb5.

Page 6: Comparative genetic and metabolic characterization between

4

Unité de recherche dans Université de Bordeaux UMR 1287 Ecophysiologie et Génomique Fonctionnelle de la Vigne

Institut des Sciences de la Vigne et du Vin

210 Chemin de Leysotte

33140 Villenave d’Ornon, France

Research Unit in Pontificia Universidad Católica de Chile Laboratorio de Biología Molecular y Biotecnología Vegetal

Facultad de Ciencias Biológicas

Departamento de Genética Molecular y Microbiología

Avenida Libertador Bernardo Bernardo O’Higgins 340

8331150 Región Metropolitana Santiago de Chile

Page 7: Comparative genetic and metabolic characterization between

5

"Happiness can be found, even in the darkest of times,

if one only remembers to turn on the light."

Albus Dumbledore, Harry Potter and the Prisoner of Azkaban.

Page 8: Comparative genetic and metabolic characterization between

6

This work was supported by:

Scholarship of Comisión Nacional de Investigación (CONICYT) Nº 21120432.

Operational Expenses Scolarship of CONICYT year 2014 and 2015, Nº 21120432.

Millennium Nucleus of Plant Systems and Synthetic Biology NC130030.

Chimenti S.A.

Page 9: Comparative genetic and metabolic characterization between

7

ACKNOWLEDGEMENT

This thesis has been performed under co-supervision between the Pontificia Universidad

Católica de Chile and the University of Bordeaux, and thus, most of my activities were carried

out at the Laboratorio de Biología Molecular y Biotecnología Vegetal (Facultad de Ciencias

Biológicas, Departamento de Genética Molecular y Microbiología from PUC) and at the

Ecophysiologie et Génomique Fonctionnelle de la Vigne (EGFV), Institut des Sciences de la

Vigne et du Vin (ISVV) (INRA and Université de Bordeaux).

My thesis supervisors, Dr. Patricio Arce and Dr. Eric Gomes, for their help in the supervision

of this thesis. My most sincere thanks for all the time that this thesis was carried out.

Listy Martínez, Valeria Borjas, Don Julio, Toño, Ghislaine Hilbert and Christel Renaud for their

technical assistance and interest in the experimental part of my thesis.

The researchers of the institutions I have been working in: Dr. Zhanwu Dai, Dr. John Lunn and

Dr. José Tomás Matus for their contribution in key thesis experiments and analysis of them.

Dr. Michael Hanford (Raporteur), Dr. Claudio Inostroza (Raporteur), Serge Delrot

(Examinateur), Lee Meissel and Rodrigo Gutiérrez for accepting the hard work of being the jury

of this private and public doctoral defense.

Dr. Sarah Frusciante, Dr. María José Clemente, Dr. Roberta Triolo, Dr. Noé Cococh and Dr.

Hugo Campbell, for their role in my adaptation in my stays at the ISVV.

Dr. Rodrigo Loyola, Dr. Alejandra Serrano, Brenda Riquelme, Francisca Parada, Dr. Carmen

Espinoza, Dr. Grace Armijo, Dr. Mario Agurto, Aníbal Arce, Carlos Meyer and Tomás Moyano,

for all your comments and corrections to my thesis that helped to shape what is finally. Special

thanks to Consuelo Medina for accompanying me during my stay in Bordeaux, France and for

being a key person in the end of this thesis.

Dr. Daniela Muñoz, Evelyn Poblete, Rowena Cortés, Camila Almendra, Yessica Arcos,

Elizabeth Torres, Susan Hitschfeld, Daniela Herrera and Amparo Rodríguez, for being the best

friends and for being by my side in this long process.

Paola Cárdenas and Marisol Carrasco for being with me in the difficult moments and know how

to get a smile on me. My Harry Potter’s friends!

And most importantly, my family, Joaquín, Mary, Alejandra, Álvaro, Nicolás, Adolfo,

Francisca, Victoria, Lucía, Emilia, Joaquín junior, Alejandro, Rebeca, for allowing me to fulfill

my dreams and every single lab mate who have been there for me during this challenging period.

Page 10: Comparative genetic and metabolic characterization between

8

TABLE OF CONTENTS

CHAPTER I 9

GENERAL INTRODUCTION 9

1. Vitis vinifera 10

2. Metabolic changes during grapevine development 11

2.1 Phase I 11

2.2 Phase II 12

2.3 Phase III 13

3. Anthocyanins, responsible of grape berry coloration 13

4. Anthocyanins biosynthesis 16

5. Transportation of anthocyanins to the vacuole 21

6. Regulation of anthocyanin biosynthesis 22

6.1 Transcriptional regulation 22

6.2 Environmental factors that can affect anthocyanin accumulation 24

6.3 Hormonal factors that can affect anthocyanin accumulation 25

7. New advances in comparative studies between contrasted berry skin varieties

27

REFERENCES FROM CHAPTER I 31

CHAPTER II 47

PROBLEM STATEMENT: Hypothesis and Objectives 47

CHAPTER III 49

TRANSCRIPTOMIC AND METABOLIC ANALYSIS REVEALS NEW GENETIC

DIFFERENCES BETWEEN TWO TABLE GRAPES WITH CONTRASTED COLOR

BERRY SKIN

CHAPTER IV 125

OBJETIVE 3: Evaluate the function of at least one gene of interest, selected form the

above objetives, related with the branching point F3’-H and F3’5’-H of the anthocyanin

biosynthesis pathway

INTRODUCTION 126

MATERIAL AND METHODS 128

RESULTS 132

DISCUSSION AND CONCLUSION 138

CHAPTER V 141

GENERAL DISCUSSION 141

CHAPTER VI 156

CONCLUSION 156

REFERENCES FROM CHAPTER IV and V 159

Page 11: Comparative genetic and metabolic characterization between

9

CHAPTER I

GENERAL INTRODUCTION

Page 12: Comparative genetic and metabolic characterization between

10

In the present chapter the literature and accumulated knowledge of grapevine development,

anthocyanins biosynthesis and regulation of anthocyanin pathway are reviewed.

1. Vitis vinifera

Grapevine (Vitis vinifera) is an economically important perennial fruit crop whose main

destinations are fresh fruits as table grape, raisins and wine making (Coombe, 1992; Coombe &

Mccarthy, 2000). Vitis vinifera L. plants have been propagated asexually for a long time, as

cuttings grafted into rootstocks. As grapevine berries develop, they display modifications in

composition, texture, size, flavour, color and in pathogen susceptibility. The grape berry is

composed of seeds and 3 major types of tissue called exocarp, corresponding to skin; mesocarp

corresponding to flesh; and endocarp which correspond to surrounding seed tissue, the latter

being very difficult to differentiate from the mesocarp (FIGURE 1) (J. Kennedy, 2002). Skin

cells are smaller than flesh cells, but are much thicker, allowing less water loss. The outermost

layer of cells of the exocarp corresponding to epidermis is covered by cuticle on their external

faces, a wax that acts as the primary protective barrier against environmental stress (Hardie,

O’Brien, & Jaudzems, 1996; Pensec et al., 2014).

Grape berry is water and nutrient supplied through the pedicel by a vascular system composed

of phloem elements and xylem. The phloem is the vasculature involved in photosynthate

(hexoses) transport from the canopy to the vine. The phloem has a reduced function in early

berry development but, becomes the primary source of ingress after veraison. The xylem is the

vasculature responsible for transporting nutrients, minerals, water and growth regulators from

the root system to the rest of the plant. Unlike the phloem, the xylem does have a fundamental

Page 13: Comparative genetic and metabolic characterization between

11

role in the first stages of berry developmental, but less after veraison (Greenspan, Shackel, &

Matthews, 1994).

FIGURE 1: Grape berry structure. A simple diagram to describe the main tissues of the ripe

grapevine berry. Diagram from (Conde et al., 2007).

2. Metabolic changes during grapevine development

2.1 Phase I

Grape berry development consists of two successive sigmoidal growth periods separated by a

lag phase (FIGURE 2) (Coombe, 1992; Coombe & Mccarthy, 2000). Phase I correspond to the

first growth period and is initiated with fruit set. Also called herbaceous growth, it lasts 60 days,

depending on the variety and environmental conditions. It is characterized by important by an

early cellular divisions phase, followed later on by cellular expansion, which results in an

accumulation of water in the vacuoles as solutes accumulate. During this phase, seed embryos

Page 14: Comparative genetic and metabolic characterization between

12

are also produced. In metabolic terms, phase I is characterized by the accumulation of organic

acids, mostly tartaric and malic acid, stored in the vacuoles. Tartaric acid accumulates mainly

at the beginning of development and malic acid towards to the end of the herbaceous phase,

with a maximum concentration reached just before veraison (Conde et al., 2007; Downey,

Dokoozlian, & Krstic, 2006). These compounds provide acidity to the wine and therefore are

critical to its quality. Other important metabolites that accumulate in Phase I are

hydroxycinnamic acids, tannins, methoxypyrazines, citric acid, succinic acid, all strongly

involved in the organoleptic qualities of wine (J. A. Kennedy, Hayasaka, Waters, & Jones, 2001;

Serrano et al., 2017). Conversely, the concentration of sugars remains low throughout this phase

and chlorophyll is then the major pigment of berries. To support itself, the grape berry produces

a few sugars via photosynthesis in situ and massively imports carbon from the leaves (Conde et

al., 2007).

2.2 Phase II

Also called lag phase, Phase II is marked by the slowing down of cell growth (FIGURE 2). It

lasts between 8 and 12 days and during this phase, the seeds change color after desiccation of

their integument (Ristic & Iland, 2005) and also the increase in berry volume is exclusively due

to cell growth (Ebadi, Sedgley, May, & Coombe, 1996). At the end of this phase the

phenomenon called veraison occurs, in which the grape berry begins to gain color, thus third

phase of growth is initiated.

Page 15: Comparative genetic and metabolic characterization between

13

2.3 Phase III

This last phase in berry development, starts with veraison, which corresponds to the onset of

ripening and as its name says, it is known by the name of ripening stage. Grape berries switch

to a status where they are larger, softer, sweeter, less acidic, and strongly flavored and colored

(Conde et al., 2007; Coombe & Mccarthy, 2000) (FIGURE 2). Organic acids concentration

decreases and sucrose produced from photosynthesis is imported into the grape berry during

ripening stage as well as water, doubling its size (Dai et al., 2013). Once transported into the

berries, the sucrose is hydrolyzed into glucose and fructose. Beyond sugar accumulation, it

begins to synthesize secondary metabolites, contributing to the organoleptic qualities of wine.

For colored cultivars, skin tissue changes from green to purple, due to the accumulation of

anthocyanins. Near the end of the third phase of berry development, other secondary metabolites

are also accumulated, such as flavors and their precursors as well as flavonols. Ripening stage

ends around 120 days after flowering (Costantini, Battilana, Lamaj, Fanizza, & Grando, 2008;

J. Kennedy, 2002; Kuhn et al., 2014) (FIGURE 2).

3. Anthocyanins, responsible of grape berry coloration

Anthocyanins are the main pigments in flowers and fruits and are responsible for the red color

of pigmented grape berries. Their localization is in the berry skin, although in some cultivars

they are found in flesh, i.e. in the so-called “teinturier” varieties (Castillo-Muñoz, Fernández-

González, Gómez-Alonso, García-Romero, & Hermosín-Gutiérrez, 2009; Guan et al., 2012).

Principal anthocyanin functions are attraction of the animals that allow seed dispersion (Jaakola,

Page 16: Comparative genetic and metabolic characterization between

14

2013), protection of berry against solar exposure and UV-B radiation, reduction of oxidative

stress (Adams, 2006; Dixon, Xie, & Sharma, 2005), defense against many different pathogens,

among others (Chalker-scott, 1999).

The basic structure of anthocyanins is the flavylium cation, that consists of the typical flavonoid

carbon skeleton of C6-C3-C6, but has two double bonds on C ring, thus getting a positive charge

(FIGURE 3A) (He et al., 2010). The aglycone part of anthocyanin is called anthocyanidin and

the structure of the different anthocyanidins varies according to the position and the numbers of

hydroxyl and methyl groups on the B ring (FIGURE 3B and 3C) (Jimenez-garcia et al., 2012;

Tanaka, Sasaki, & Ohmiya, 2008).

The common anthocyanidins found in grapes are cyanidin, delphinidin, peonidin, petunidin and

malvidin, being malvidin the predominant anthocyanidin in most pigmented berry grape (Holton

& Cornish, 1995; Mazza, 1995). Blueness is increased with free hydroxyl groups, whereas

redness intensifies with the raising of the methylation of the hydroxyl groups. Thus, malvidin is

the reddest individual anthocyanidin. Sugars are bound to the anthocyanidins through a

glycosidic bond at the C3 and C5 positions, monoglycosylated anthocyanins in C3 being the

most common found in Vitis vinifera (Yamazaki et al., 2015). Methylation takes place after

glycosylation, after the anthocyanins are already formed. Both process contribute to the

solubility and stabilization of anthocyanidins by protecting them from oxidation. In Vitis

vinifera anthocyanins of 3-O-monoglucoside, acetylated and p-coumaroyl forms have been

identified (Acevedo et al., 2012; Haibo Wang, Race, & Shrikhande, 2003).

Page 17: Comparative genetic and metabolic characterization between

15

FIGURE 2: Graphical representation of grape berry development. The diagram shows

berry size and coloration during development and major developmental events. Also, it shows

an inflow rate of phloem and xylem into the berry and accumulation periods of different

compounds. Illustration from (J. Kennedy, 2002)

Page 18: Comparative genetic and metabolic characterization between

16

FIGURE 3: Anthocyanin structure. (A) Basic structure of an anthocyanidin showing the

numbering of the different carbon groups. The common modification positions are R1, R2, R3,

R4, R5 and R6. (B) Hydroxylated anthocyanins in which hydroxyl group is show in purple color.

(C) Methoxylated anthocyanins in which methoxyle group is show in green color. Modified

from (Y. Zhang, Butelli, & Martin, 2014).

4. Anthocyanins biosynthesis

Anthocyanins belong to the flavonoid class of compounds and are synthesized along the general

phenylpropanoid pathway by the activity of a cytosolic multi-enzymatic complex, associated to

the cytoplasmic face of the endoplasmic reticulum (Boss, Davies, & Robinson, 1996). Once

produced anthocynins are later on transported and stored in the skin cell vacuoles (Winkel-

Shirley, 2001). Most of the genes encoding enzymes of the anthocyanin biosynthesis pathway

have been well identified in grapevine (FIGURE 4).

Phenylpropanoid biosynthesis begins with the amino acid phenylalanine, which is a product of

the shikimate pathway (Maeda & Dudareva, 2012). The shikimate pathway begins with

erythrose-4-phosphate and phosphoenol pyruvate and is responsible for producing the other

aromatic amino acids: tyrosine and tryptophan. The first enzyme responsible for the phenolic

biosynthesis is phenylalanine ammonia-lyase (PAL), which converts phenylalanine into

Page 19: Comparative genetic and metabolic characterization between

17

cinnamic acid. PAL is encoded by a multigene family in the grapevine (Sparvoli, Martin,

Scienza, Gavazzi, & Tonelli, 1994) and is transcriptionally regulated in response to biotic and

abiotic factors (Boss et al., 1996). Cinnamic acid suffers a series of transformations resulting in

the synthesis of precursors of several simple phenolics, such as phenolic acids, lignin precursors,

etc. The incorporation of 3 molecules of malonyl-CoA, produced via the acetate pathway, with

the 4-coumaroyl-CoA starts the phenylpropanoid pathway. These last compounds are condensed

by chalcone synthase (CHS), the first committed enzyme of flavonoid biosynthesis. There are

at least in grapevines three genes encoding CHS, CHS1, CHS2 and CHS3, each one may act in

three different pathways to produce different secondary metabolites (Castillo-Muñoz et al.,

2009; Goto-Yamamoto, Wan, Masaki, & Kobayashi, 2002; Huiling Wang et al., 2016).

Then, CHS and chalcone isomerase (CHI) are the enzymes involved in a two-step condensation,

producing a colourless flavanone named naringenin. In Vitis vinifera, there is a single gene

encoding a putative CHI, that is expressed strongly at the onset of veraison (Goes et al., 2005).

Later, naringenin is oxidized by flavanone 3-hydroxylase (F3-H) giving rise the formation of

dihydrokaempferol (colourless dihydroflavonol). This compound is also the substrate for

flavonoid 3’-hydroxylase (F3'-H) and flavonoid 3’5’-hydroxylase (F3'5'-H), which allows

hydroxylation to dihydroquercetin and dihydromyricetin, respectively (FIGURE 4). In

grapevines, there are two genes encoding F3-H, two coding F3’-H and sixteen copies of the

gene coding for F3’5’-H, of which only eight encode full-length proteins (Falginella et al.,

2010). It has been showed that these genes are differentially expressed in the different tissues of

the berry, also according to the stage of development. The high copy number of these genes

would increase the complexity and diversity among pigmented varieties and F3′-H and F3’5’-

H expression directly affects the anthocyanin composition (Ali et al., 2011; Jochen Bogs, Ebadi,

Page 20: Comparative genetic and metabolic characterization between

18

McDavid, & Robinson, 2006; Falginella et al., 2010). A recent study showed that in Cabernet

Sauvignon and Shiraz, F3’5’H is as a key molecular player driving the phenylpropanoid

metabolic flux towards bluish anthocyanins production (Degu et al., 2014). These two genes,

F3′-H and F3’5’-H are members of P450 cytochrome family (de Vetten et al., 1999) and in

several studies it had been showed another member of P450 cytochrome family, cytochrome b5

(Cytb5) has previously been involved in F3’5’-H activity modulation in Petunia flowers, by

catalyzing the transfer of electrons to their prosthetic heme group (de Vetten et al., 1999). In

2006, (Jochen Bogs et al., 2006) demonstrated that a putative Cytb5 from grapevine cv Shiraz

modulated both F3’5’-H and F3’-H activities, even though the exact mechanism remains to be

deciphered.

Continuing with the following steps of the pathway, dihydroquercetin and dihydromyricetin are

converted to anthocyanidins (coloured but unstable pigments) by two reactions catalyzed by

dihydroflavonol reductase (DFR) and anthocyanidin synthase (ANS or also called LDOX). The

DFR converts dihydroquercetin and dihydromyricetin into leucocyanidin and leucodelphinidin

(colourless), respectively; and LDOX catalyzes the oxidation of leucocyanidin into cyanidin

(red-magenta anthocyanidin) and leucodelphinidin into delphinidin (purple-blue anthocyanidin)

(Abrahams et al., 2003; Sparvoli et al., 1994). In grapevines, DFR is coded by a single gene (He

et al., 2010; Winkel-Shirley, 2001), which it first expresses in the grape berry development

during phase I, decreased at veraison and it increases again during the phase III (Jochen Bogs

et al., 2005). Anthocyanidins formed by the reaction catalyzed by LDOX are immediately

modified by glycosylation, methylation and acylation, strongly necessary for their stabilization

as vacuolar anthocyanins (He et al., 2010; Springob, Nakajima, Yamazaki, & Saito, 2003). For

Page 21: Comparative genetic and metabolic characterization between

19

example, it had been reported that red grapes mutant lacking gene expression for the initial

glycosylation of anthocyanidins, did not generate anthocyanins (Boss et al., 1996).

Anthocyanidins are glycosylated by the action of UDP-glucose:flavonoid 3-O-

glucosyltransferase (UFGT), which catalyze the O-glycosylation in C3 position (Vitis vinifera

sp), using UDP-glucose as sugar donor (Ford, Boss, & Hoj, 1998). Other sugar donors can be

also being used occasionally by UFGT, such as UDP-xylose, UDP-galactose, UDP-arabinose,

UDP-rhamnose and UDP-glucuronic acid (Springob et al., 2003). In general, UFGT gene

expression is only detected in berry skin after veraison (Boss et al., 1996) and it is expressed

only in red cultivars (Kobayashi, Goto-Yamamoto, & Hirochika, 2004). In most non-V. vinifera

grape species the O-glycosylation can also occur at both of the C3 and C5 positions to form

anthocyanidin-3,5-O-diglucoside (Lamikanra, 1989).

With the participation of S-adenosyl-L-methionine (SAM), a O-methyltransferase (OMT)

catalyzes the methylation of the hydroxyl groups at the C3’ positions or both the C3’ and C5’

positions on the B rings of the anthocyanins (FIGURE 3A) (Ibrahim, Bruneau, & Bantignies,

1998). Acylation is also one of the most common modifications of anthocyanins catalized by an

anthocyanin acyltransferase (ACT) which produce 3-O-acetyl-, 3-O- coumaroyl-, and 3-O-

caffeoyl-monoglucosides by attaching acyl groups to the C6’’ position of the Glucose moiety

(Mazza, 1995; C.-L. Zhao et al., 2017). A recent study showed that Vv3ACT has novel activity

found in grapevine because it has a broad anthocyanin substrate specificity and is able to use

both aliphatic and aromatic acyl donors (Rinaldo et al., 2015).

Page 22: Comparative genetic and metabolic characterization between

20

FIGURE 4: Flavonoid biosynthesis pathway in grapevine. Enzymes that constitue the

pathway are shown, as well as the main intermediates and the transcription factors that regulate

the pathway (in light blue color). Blue boxes contain delphinidin-type anthocyanins and red

boxes cyaniding-type anthocyanins. CHS, chalcone synthase; CHI, chalcone isomerase; F3H,

flavanone 3-hydroxylase; F3’H, flavonoid 3’-hydroxylase; F3’5’H, flavonoid 3’5’-hydroxylase;

DFR, dihydroflavonol 4-reductase; LDOX, leucoanthocyanin dioxygenase; UFGT,

UDPglucose:flavonoid 3-O-glucosyltransferase; OMT, O-methyltransferase; FLS, flavonol

synthase; GST glutathione S-transferase; AM1/AM3, anthocyanin multidrug and toxic

extrusion transporters; ABCC1, ATP-binding cassette transporter. Adapted from (Kuhn et al.,

2014).

Page 23: Comparative genetic and metabolic characterization between

21

5. Transportation of anthocyanins to the vacuole

Anthocyanins are synthesized in the cytosol and stored into the vacuole (Chen, Wen, Kong, &

Pan, 2006). Three distinct, but potentially nonexclusive mechanisms for flavonoid transport in

plant cells have been proposed: vesicle trafficking, membrane transporter, and GST mediated

transport (J. Zhao, Pang, & Dixon, 2010). Recent evidence suggests that the three mechanisms

collaborate in vacuolar sequestration of flavonoids or flavonoid secretion through the plasma

membrane to the apoplast (J. Zhao et al., 2010).

Vesicular transport is mostly supported by microscopy observations (Gomez et al., 2009, 2011;

H. Zhang, Wang, Deroles, Bennett, & Davies, 2006). Concerning the membrane transporter

mechanism, molecular and genetic evidences support the involvement of both ATP Binding

Cassette (ABC) proteins and Multidrug and Toxic Extrusion (MATE) in the transport of

anthocyanins (Gomez et al., 2009; Goodman, Casati, & Walbot, 2004; J. Zhao et al., 2010).

MATE transporters include a large family of membrane transporters. The first gene identified

was Transparent Testa 12 (TT12) involved in flavonoid biosynthesis in Arabidopsis thaliana

seed testa (Debeaujon, Peeters, Léon-kloosterziel, & Koornneef, 2001) and was able to transport

cyanidin 3-O-glucoside, when expressed in heterologous form in yeast (Marinova et al., 2007).

The grapevine VvGST genes are also involved in anthocyanin transport (Conn, Curtis, Bézier,

Franco, & Zhang, 2008), and in VvGST-knockdown hairy roots, anthocyanins accumulated only

in small vesicles but did not make it to the vacuole. Also, hairy roots with knockdown of AM1

and AM3 had no anthocyanin-filled small vesicles (Gomez et al., 2011). This suggests that AM1

and AM3 proteins sequestrate anthocyanins from the cytoplasm to the small vesicles before

Page 24: Comparative genetic and metabolic characterization between

22

VvGST-aided transport to the central vacuole occurs via vesicle trafficking (Gomez et al.,

2011).

Genetic studies have shown that multidrug resistance-associated protein (MRP)/C-type of ABC

(ABCC) transporters are also involved in anthocyanin accumulation with the assumption that

they transport flavonoid conjugates with glutathione (GSH) (Banasiak, Biała, Staszków,

Swarcewicz, & Ewa, 2017; J. Zhao et al., 2010). In grapevine, it was shown that free GSH is

strictly required for transport of malvidin 3-O-glucoside into yeast vacuoles with ABCC1 (Goes

et al., 2005). Therefore, grapevine ABCC1 is a GSH-dependent anthocyanin transporter, rather

than an anthocyanin–GSH transporter, but the actual mechanism remains unclear.

6. Regulation of anthocyanin biosynthesis

6.1 Transcriptional regulation

Anthocyanin biosynthesis pathway is under the control of several different families of

regulatory genes. These correspond to transcription factor type of R2R3-MYB proteins, bHLH

(basic helix–loop-helix) proteins and WDR (tryptophan–aspartic acid repeat) proteins, which

also play crucial roles in the regulation of other flavonoid products, such as proanthocyanidins

and flavonols (He et al., 2010; Tian, Pang, & Dixon, 2008). FIGURE 4 from (Kuhn et al., 2014)

in addition to the structural genes of the pathway, also shows the transcription factors

characterized so far. These transcription factors could interact as ternary complexes MYB-

bHLH-WDR to regulate the expression of genes encoding enzymes involved in the final steps

of flavonoid biosynthetic pathway (Imène Hichri et al., 2011).

Page 25: Comparative genetic and metabolic characterization between

23

In grapevine, 108 members of R2R3-MYB gene subfamily have been classified and described

(José Tomás Matus, Aquea, & Arce-Johnson, 2008). MYB transcription factors are positive

regulators of the flavonoid pathway such as MYBPA1 for tannin production (J Bogs, Jaffe,

Takos, Walker, & Robinson, 2007); MYBF1 for flavonol production (Czemmel et al., 2009);

MYB5a and MYB5b in anthocyanin and proanthocyanidin production (Cavallini et al., 2017;

L. Deluc et al., 2008); and MYBA1 and MYBA2 for direct control of anthocyanin biosynthesis

(Kobayashi et al., 2004). These two genes are located in tandem in a single locus on

chromosome 2, which is called the “berry color locus” (Kobayashi et al., 2004; Walker, Lee, &

Robinson, 2006). MYBA1 transcriptionally regulates UFGT, which glycosylates

anthocyanidins, stabilizing them and therefore responsible for the anthocyanins accumulation

in grape berry skins of pigmented varieties (Ford et al., 1998). White skinned grape contains a

mutated allele of MYBA1 due to the presence of the retrotransposon Gret1 in its promoter

region, resulting in a white berry allele (Kobayashi et al., 2004). Pigmented cultivars possess at

least one allele of the MybA1 locus that does not contain Gret1 and is thus called a red berry

allele (This, Lacombe, Cadle-Davidson, & Owens, 2007). Also, a 33bp insertion was found in

the second intron of MYBA1, generating low gene expression levels in grape pink berry skin

varieties (Shimazaki, Fujita, Kobayashi, & Suzuki, 2011) and a deletion of a large region of

chromosome 2 including MYBA1 and MYBA2 genes produced the bud sports of Cabernet

Sauvignon called “Malian” with red-grey berry skin (Walker et al., 2006).

The others transcription factors WDR and bHLH have been characterized to a lesser extend in

grapevine. WDR1 promotes anthocyanin accumulation when ectopically expressed in

Arabidopsis thaliana (J. T. Matus et al., 2010). MYC1 (a bHLH transcription factor) is involved

in the anthocyanin and tannin biosynthesis regulation and trans-activates gene promoter of CHI,

Page 26: Comparative genetic and metabolic characterization between

24

of ANR (anthocyanidin reductase which catalyze the synthesis of flavan-3-ols such as catechin

and epicatechin) and of UFGT. MYC1 also interacts with different MYB partners to active

transcription in yeast (Imne Hichri et al., 2010)

Concerning negative regulators of anthocyanin biosynthesis, only a single MYB transcription

factor, MYB4, was identified and it could repress UFGT gene expression (José Tomás Matus et

al., 2009)

6.2 Environmental factors that can affect anthocyanin accumulation

In recent years, our knowledge about control of environment in berry ripening has greatly

increased. Several studies have shown that light exposure can increase anthocyanins

concentration, especially in skin cells, and that shading can have the opposite effect (Azuma &

Yakushiji, 2012).

In Cabernet Sauvignon it was shown that light exclusion or leaf shading significantly decreases

berry anthocyanin content and expression of genes associated directly with anthocyanin

biosynthesis such as MYBA1, MYBA2 and UFGT (Koyama & Goto-yamamoto, 2008; José

Tomás Matus et al., 2009). A recent study showed that several MYB transcription factors related

to flavonoid biosynthesis in grapevine respond differently to light and temperature conditions

(Greer & Weston, 2015; Jeong, Goto-yamamoto, Kobayashi, & Esaka, 2004). It was also

observed that gene expression of MYBA1-2, MYBA1-3 and MYBA2 was different in grapevine

berry skin in darkness or light at low (15ºC) or high (35ºC) temperature conditions, whereas

Page 27: Comparative genetic and metabolic characterization between

25

other studies concluded MYB5a and MYB5b did not respond to these environment stimuli

(Azuma & Yakushiji, 2012).

Water deficit (Deis, Cavagnaro, Bottini, Wuilloud, & Silva, 2011; L. G. Deluc et al., 2009) can

also affect anthocyanins content and other polyphenols like stilbenes and sugars, increasing their

concentrations, affecting final berry weight and up-regulating LDOX, UFGT and MYBA1

(Castellarin et al., 2007).

Also, light quality can affect anthocyanins biosynthesis, especially UV light. UV-B significantly

affects grape berry ripening, increasing anthocyanin and flavonols content and decreasing berry

weight (Berli et al., 2010; Berli, Fanzone, & Piccoli, 2011). Another environmental factor can

affect anthocyanin content is the infection by pathogens such as viruses (Singh, Singh, Swinny,

& Cameron, 2008; Vega, Gutierrez, Peña-Neira, Cramer, & Arce-Johnson, 2011).

Some of these environmental signals are also involved in active hormone contents regulation,

affecting berry ripening. For example, ABA levels increase in response to water deficit and low

temperatures exposure (L. G. Deluc et al., 2009; Yamane, Jeong, Goto-yamamoto, Koshita, &

Kobayashi, 2006).

6.3 Hormonal factors that can affect anthocyanin accumulation

Several studies have suggest abscissic acid (ABA) as one of the key signal triggering grape

berry ripening, due to the studies that show a significantly increase in ABA berry content at the

beginning of phase III (Lacampagne, Gagné, & Gény, 2010; Sun et al., 2010). In Cabernet

Sauvignon, a rapid increase in ABA content coincides with the rise in total anthocyanin content,

however, anthocyanins keep increasing during ripening stage, while ABA levels start to

Page 28: Comparative genetic and metabolic characterization between

26

decrease at the same time, suggesting that ABA only is necessary to trigger anthocyanin

biosynthesis (Wheeler, Loveys, Ford, & Davies, 2009). In addition, exogenous application of

ABA also increases anthocyanin content, berry weight and decrease acidity (Berli et al., 2010;

Gambetta, Matthews, Shaghasi, McElrone, & Castellarin, 2010; Wheeler et al., 2009). Increased

sugar concentrations in grape berry is associates with an increase in ABA content, but ABA

treatment does not affect sugar content (Peppi & Fidelibus, 2008). Regarding gene regulation

of ABA, it had been showed that expression of PAL, CHS, F3-H, GST, VvMYBA1, among others,

is increased in Cabernet Sauvignon skin treated with ABA at veraison stage (Koyama &

Sadamatsu, 2010). Another interesting fact is that ABA treatments repress putative auxin

response genes and at the same time increase the expression of genes related to ethylene

biosynthesis, suggesting ABA effects are mediated in conjunction with other hormones.

Auxin level, which is high in the green phase of rapid berry growth, strongly decreases levels

during veraison and ripening stages (Böttcher, Keyzers, Boss, & Davies, 2010). It had been

reported that grape berries treated with naphthalene acetic acid (NAA) 1 week before veraison

shows a delay in anthocyanins and sugar accumulation, parallel to changes in transcripts

involved in up-regulation of genes related to ethylene biosynthesis and down-regulation of

genes related to ABA biosynthesis (Ziliotto et al., 2012).

When berries are treated at veraison stage with an ethylene-releasing compound (2-CEPA), an

increase occurs in anthocyanin content (El-Kereamy et al., 2003), and on the other hand, when

an inhibitor of ethylene receptors is used (1-MCP) produce an decreased in anthocyanin content

accompanied by an increase in acidity in grape berries (Chervin et al., 2004). 1-MPC use also

suggest a relationship of ethylene with ABA, because its use produces a decrease in ABA

content in grape berry (Sun et al., 2010)

Page 29: Comparative genetic and metabolic characterization between

27

It had been showed that cytokinins increase berry weight but decrease sugar and anthocyanin

content (Peppi & Fidelibus, 2008) Finally, jasmonic acid increases anthocyanin production in

grape cell suspensions (Belhadj et al., 2008).

7. New advances in comparative studies between contrasted berry skin

varieties

In the last decade, several studies demonstrated that transcriptomic remodeling drives the

transition from the vegetative to the ripening stage, a process which is highly coordinated at the

transcriptional level (Fasoli et al., 2012). There has been advances in deep-sequencing

technologies that, integrated with the availability of grapevine genomic and transcriptomic data,

allow to carry out RNA-transcript expression analysis on a global scale at key points during

grapevine development (Serrano et al., 2017).

Many transcriptomic studies focused on berry development on a specific variety. A study in

Cabernet Sauvignon using Affymetrix technique, had reported that approximately 60% of the

detected transcripts correspond to differentially expressed genes between at least two out of the

seven stages of berry development, demonstrating the dynamic nature of the developmental

process in grapevine. These genes were related with many development processes in which 21

Unigenes encoding biosynthetic enzymes of the general phenylpropanoid and flavonoid

pathways were found to exhibit differential mRNA expression patterns across berry

development (L. G. Deluc et al., 2007). Studies in other pigmented varieties such as Shiraz also

showed that gene clusters for ripe berry contain genes related with phenylpropanoid metabolism

and secondary metabolism (Sweetman, Wong, Ford, & Drew, 2012). Another similar study but,

Page 30: Comparative genetic and metabolic characterization between

28

using a white-skinned berry variety called Garganega, realized its metabolomic and

transcriptomic characterization. It was shown that Garganega was highly plastic in different

environments and indeed more plastic than ripening Corvina berries, predominantly concerning

phenolic compounds accumulation (Dal Santo et al., 2016).

In relation with comparative studies between varieties, there are several examples that compare

varieties with the same berry skin color. Using berry skin of Norton variety, an important North

American wine grape, compared with Cabernet Sauvignon, Ali et al. 2011, showed an up-

regulation of flavonoid metabolism. Higher transcript levels in F3’-H, F3’5’-H, LDOX, UFGT,

among others, were reported in Norton variety compared to Cabernet Sauvignon (Ali et al.,

2011). Another study in the table grape ‘Kyoho’ variety versus its bud mutant ‘Fengzao’ indicate

that the mutant reached ripening stage 30 days earlier than ‘Kyoho’ (Guo et al., 2016). Another

comparative transcriptomic study concluded that differentially expressed genes may suggest

that gene activities related to ROS and pathogenesis are involved in contributing to the

phenotype observed, due to flavonoid metabolism differences (Guo et al., 2016). In addition,

light-red-skinned ‘Ruby Okuyama’ and more intense and uniform rosy-skinned ‘Benitaka’

correspond to bud sports of white-skinned ‘Italia’ (Azuma et al., 2009). ‘Ruby Okuyama’ was

caused by the recovery of VvmybA1 expression and ‘Benitaka’ due to presence of a novel

VvmybA1BEN allele that restored VvmybA1 transcripts (Azuma et al., 2009). Much less

information exists about studies that take into account the full range of berry skin color shades

that is present in V. vinifera cultivars. That is why we were interested in contributing to fill this

gap by considering more types of color berry skin, in near isogenic backgrounds.

Page 31: Comparative genetic and metabolic characterization between

29

We have two grapes that present contrasted differences in berry skin. One of them is Red Globe

(RG), a table grape whose berries are characterized by be seeded, of a large size, thick and

consistent skin of dark purple color. And the second one corresponds to Chimenti Globe (CG)

that has similar characteristics but differs in color skin that is bright light red. It originates from

a spontaneous mutation event that occurred in the vineyard from a branch of a vine of RG plant

in Talagante, Metropolitan Region of Chile, and was subsequent propagated through clonal

multiplication (FIGURE 5). As a first approximation, we asked if the color change was due to

a change in the key regulator known in grapevines, MYBA1, which it has been associated as

responsible for the majority color of berry skin found in grapes. We think that something similar

to what was observed in the work of Walker et al., 2006 with the variety Cabernet Sauvignon

and his bud sport was happening. We performed an analysis gene expression of the white and

red allele of MYBA1 by RT-PCR. We found surprisingly that both varieties, CG and RG, are

heterozygous for color berry locus since they express both alleles at the same time (FIGURE

6). With this observation, we discard MYBA1 like a possible explanation and not directly related

with the color change observed in this new variety. We are wondering about the relationship in

specific berry color and the type of anthocyanins that could have, which led us directly to think

of some change or modification in the so-called branching point in anthocyanin biosynthesis,

shaped by the enzymes F3’5’-H and F3’-H. Anyway, we do not have the certainty to which level

of anthocyanin biosynthesis was affected. For that reason, we consider CG and RG varieties an

interesting study model due to the genetic background responsible for the color change is the

almost identical (i.e. near isogenic), with differences affecting the color of the berry, therefore

allowing us to study the process of accumulation of anthocyanins using two near-isogenic

varieties with contrasting berry skin color.

Page 32: Comparative genetic and metabolic characterization between

30

FIGURE 5: Chimenti Globe and Red Globe table grapes. In the picture is show varieties

selected for this thesis as model of study. CG corresponds to the one on the left of the picture

and RG on the right. This picture was taken in ripening stage on season 2013. Both varieties are

located in Talagante, Metropolitan Region of Chile

FIGURE 6: Analysis of MybA1 gene structure in CG and RG. (A) PCR analysis of white

allele in CG, RG varieties and control Merlot (Me), Cabernet Sauvignon (CS), Chardonnay (Ch)

and Sauvignon Blanc (SB). (B) PCR analysis of red allele in CG and RG varieties and control

described above. (C) Integrity control of actin gene. Numbers on the left indicate the positions

of the molecular size marker 1 Kb Plus DNA Ladder and on the right the exact size of white and

red allele.

Page 33: Comparative genetic and metabolic characterization between

31

REFERENCES FROM CHAPTER I

Abrahams, S., Lee, E., Walker, A. R., Tanner, G. J., Larkin, P. J., & Ashton, A. R. (2003). The

Arabidopsis TDS4 gene encodes leucoanthocyanidin dioxygenase ( LDOX ) and is

essential for proanthocyanidin synthesis and vacuole development, 624–636.

http://doi.org/10.1046/j.1365-313X.2003.01834.x

Acevedo, A., Cruz, D., Hilbert, G., Riviere, C., Mengin, V., Ollat, N., … Richard, T. (2012).

Analytica Chimica Acta Anthocyanin identification and composition of wild Vitis spp .

accessions by using LC – MS and LC – NMR. Analytica Chimica Acta, 732, 145–152.

http://doi.org/10.1016/j.aca.2011.11.060

Adams, D. O. (2006). Phenolics and ripening in grape berries. American Journal of Enology

and Viticulture, 57(3), 249–256.

Ali, M. B., Howard, S., Chen, S., Wang, Y., Yu, O., Kovacs, L. G., & Qiu, W. (2011). Berry

skin development in Norton grape: distinct patterns of transcriptional regulation and

flavonoid biosynthesis. BMC Plant Biology, 11(1), 7. http://doi.org/10.1186/1471-2229-

11-7

Azuma, A., Kobayashi, S., Goto-Yamamoto, N., Shiraishi, M., Mitani, N., Yakushiji, H., &

Koshita, Y. (2009). Color recovery in berries of grape (Vitis vinifera L.) “Benitaka”, a bud

sport of “Italia”, is caused by a novel allele at the VvmybA1 locus. Plant Science, 176(4),

470–478. http://doi.org/10.1016/j.plantsci.2008.12.015

Azuma, A., & Yakushiji, H. (2012). Flavonoid biosynthesis-related genes in grape skin are

differentially regulated by temperature and light conditions, 1067–1080.

http://doi.org/10.1007/s00425-012-1650-x

Banasiak, J., Biała, W., Staszków, A., Swarcewicz, B., & Ewa, K. (2017). A Medicago

Page 34: Comparative genetic and metabolic characterization between

32

truncatula ABC transporter belonging to subfamily G modulates the level of isoflavonoids,

64(4), 1005–1015. http://doi.org/10.1093/jxb/ers380

Belhadj, A., Telef, N., Saigne, C., Cluzet, S., Barrieu, F., Hamdi, S., & Mérillon, J.-M. (2008).

Effect of methyl jasmonate in combination with carbohydrates on gene expression of PR

proteins, stilbene and anthocyanin accumulation in grapevine cell cultures. Plant

Physiology and Biochemistry, 46(4), 493–499.

http://doi.org/https://doi.org/10.1016/j.plaphy.2007.12.001

Berli, F. J., Fanzone, M., & Piccoli, P. (2011). Solar UV-B and ABA Are Involved in Phenol

Metabolism of Vitis vinifera L . Increasing Biosynthesis of Berry Skin Polyphenols, 4874–

4884.

Berli, F. J., Moreno, D., Piccoli, P., Hespanhol-viana, L., Silva, M. F., Bressan-smith, R., …

Bottini, R. (2010). Abscisic acid is involved in the response of grape ( Vitis vinifera L .)

cv . Malbec leaf tissues to ultraviolet-B radiation by enhancing ultraviolet-absorbing

compounds , antioxidant enzymes and membrane sterols, 1–10.

http://doi.org/10.1111/j.1365-3040.2009.02044.x

Bogs, J., Downey, M. O., Harvey, J. S., Ashton, A. R., Tanner, G. J., & Robinson, S. P. (2005).

Proanthocyanidin Synthesis and Expression of Genes Encoding Leucoanthocyanidin

Reductase and Anthocyanidin Reductase in Developing Grape Berries and Grapevine

Leaves 1 [ w ], 139(October), 652–663. http://doi.org/10.1104/pp.105.064238.disease

Bogs, J., Ebadi, A., McDavid, D., & Robinson, S. P. (2006). Identification of the flavonoid

hydroxylases from grapevine and their regulation during fruit development. Plant

Physiology, 140(1), 279–91. http://doi.org/10.1104/pp.105.073262

Bogs, J., Jaffe, F. W., Takos, A. M., Walker, A. R., & Robinson, S. P. (2007). The grapevine

Page 35: Comparative genetic and metabolic characterization between

33

transcription factor VvMYBPA1 regulates proanthocyanidin synthesis during fruit

development. Plant Physiol, 143. http://doi.org/10.1104/pp.106.093203

Boss, P. K., Davies, C., & Robinson, S. P. (1996). Analysis of the Expression of Anthocyanin

Pathway Genes. Plant Physiology, 111(1 996), 1059–1066.

http://doi.org/10.1104/pp.111.4.1059

Böttcher, C., Keyzers, R. A., Boss, P. K., & Davies, C. (2010). Sequestration of auxin by the

indole-3-acetic acid-amido synthetase GH3-1 in grape berry (Vitis vinifera L.) and the

proposed role of auxin conjugation during ripening. Journal of Experimental Botany,

61(13), 3615–3625. http://doi.org/10.1093/jxb/erq174

Castellarin, S. D., Pfeiffer, A., Sivilotti, P., Degan, M., Peterlunger, E., & Gaspero, G. D. I.

(2007). Transcriptional regulation of anthocyanin biosynthesis in ripening fruits of

grapevine under seasonal water deficit, 1381–1399. http://doi.org/10.1111/j.1365-

3040.2007.01716.x

Castillo-Muñoz, N., Fernández-González, M., Gómez-Alonso, S., García-Romero, E., &

Hermosín-Gutiérrez, I. (2009). Red-Color Related Phenolic Composition of Garnacha

Tintorera (Vitis vinifera L.) Grapes and Red Wines. Journal of Agricultural and Food

Chemistry, 57(17), 7883–7891. http://doi.org/10.1021/jf9002736

Cavallini, E., Zenoni, S., Finezzo, L., Guzzo, F., Zamboni, A., Avesani, L., & Tornielli, G. B.

(2017). Functional Diversification of Grapevine MYB5a and MYB5b in the Control of

Flavonoid Biosynthesis in a Petunia Anthocyanin Regulatory Mutant, 55(November), 517–

534. http://doi.org/10.1093/pcp/pct190

Chalker-scott, L. (1999). Invited Review Environmental Significance of Anthocyanins in Plant

Stress Responses, 70(1), 1–9. http://doi.org/10.1111/j.1365-313X.2008.03447.x.

Page 36: Comparative genetic and metabolic characterization between

34

Chen, J., Wen, P., Kong, W., & Pan, Q. (2006). Changes and subcellular localizations of the

enzymes involved in phenylpropanoid metabolism during grape berry development, 163.

http://doi.org/10.1016/j.jplph.2005.07.006

Chervin, C., El-Kereamy, A., Roustan, J. P., Latché, A., Lamon, J., & Bouzayen, M. (2004).

Ethylene seems required for the berry development and ripening in grape, a non-climacteric

fruit. Plant Science, 167(6), 1301–1305. http://doi.org/10.1016/j.plantsci.2004.06.026

Conde, C., Silva, P., Fontes, N., Dias, A. C. P., Tavares, R. M., Sousa, M. J., … Gerós, H.

(2007). Biochemical changes throughout grape berry development and fruit and wine

quality. Food, 1, 1–22. http://doi.org/10.1093/jxb/ert395

Conn, S., Curtis, C., Bézier, A., Franco, C., & Zhang, W. (2008). Purification , molecular

cloning , and characterization of glutathione S -transferases ( GSTs ) from pigmented Vitis

vinifera L . cell suspension cultures as putative anthocyanin transport proteins. Journal of

Experimental Botany, 59(13), 3621–3634. http://doi.org/10.1093/jxb/ern217

Coombe, B. G. (1992). Research on Development and Ripening of the Grape Berry. American

Journal of Enology and Viticulture, 43(1), 101 LP-110. Retrieved from

http://www.ajevonline.org/content/43/1/101.abstract

Coombe, B. G., & Mccarthy, M. G. (2000). Dynamics of grape berry growth and physiology of

ripening. Australian Journal of Grape and Wine Research, 6(2), 131–135.

http://doi.org/10.1111/j.1755-0238.2000.tb00171.x

Costantini, L., Battilana, J., Lamaj, F., Fanizza, G., & Grando, M. S. (2008). Berry and

phenology-related traits in grapevine (Vitis vinifera L.): from quantitative trait loci to

underlying genes. BMC Plant Biology, 8, 38. http://doi.org/10.1186/1471-2229-8-38

Czemmel, S., Stracke, R., Weisshaar, B., Cordon, N., Harris, N. N., Walker, A. R., … Germany,

Page 37: Comparative genetic and metabolic characterization between

35

S. C. (2009). The Grapevine R2R3-MYB Transcription Factor VvMYBF1 Regulates

Flavonol Synthesis in Developing, 151(November), 1513–1530.

http://doi.org/10.1104/pp.109.142059

Dai, Z. W., Meddar, M., Renaud, C., Merlin, I., Hilbert, G., Delrot, S., & Gomès, E. (2013).

Long-term in vitro culture of grape berries and its application to assess the effects of sugar

supply on anthocyanin accumulation. Journal of Experimental Botany, 65(16), 4665–4677.

http://doi.org/10.1093/jxb/ert489

Dal Santo, S., Fasoli, M., Negri, S., D’Incà, E., Vicenzi, N., Guzzo, F., … Zenoni, S. (2016).

Plasticity of the Berry Ripening Program in a White Grape Variety. Frontiers in Plant

Science, 7(July), 970. http://doi.org/10.3389/fpls.2016.00970

de Vetten, N., ter Horst, J., van Schaik, H.-P., de Boer, A., Mol, J., & Koes, R. (1999). A

cytochrome b 5 is required for full activity of flavonoid 3’,5’ -hydroxylase , a cytochrome

P450 involved in the formation of blue flower colors, 96(January), 778–783.

Debeaujon, I., Peeters, A. J. M., Léon-kloosterziel, K. M., & Koornneef, M. (2001). The

TRANSPARENT TESTA12 Gene of Arabidopsis Encodes a Multidrug Secondary

Transporter-like Protein Required for Flavonoid Sequestration in Vacuoles of the Seed

Coat Endothelium, 13(April), 853–871.

Degu, A., Hochberg, U., Sikron, N., Venturini, L., Buson, G., Ghan, R., … Fait, A. (2014).

Metabolite and transcript profiling of berry skin during fruit development elucidates

differential regulation between Cabernet Sauvignon and Shiraz cultivars at branching

points in the polyphenol pathway. BMC Plant Biology, 14(1), 1–20.

http://doi.org/10.1186/s12870-014-0188-4

Deis, L., Cavagnaro, B., Bottini, R., Wuilloud, R., & Silva, M. F. (2011). Water deficit and

Page 38: Comparative genetic and metabolic characterization between

36

exogenous ABA significantly affect grape and wine phenolic composition under in field

and in-vitro conditions, 11–21. http://doi.org/10.1007/s10725-011-9570-5

Deluc, L., Bogs, J., Walker, A. R., Ferrier, T., Decendit, A., Merillon, J.-M., … Barrieu, F.

(2008). The transcription factor VvMYB5b contributes to the regulation of anthocyanin

and proanthocyanidin biosynthesis in developing grape berries. Plant Physiology, 147(4),

2041–2053. http://doi.org/10.1104/pp.108.118919

Deluc, L. G., Grimplet, J., Wheatley, M. D., Tillett, R. L., Quilici, D. R., Osborne, C., …

Cramer, G. R. (2007). Transcriptomic and metabolite analyses of Cabernet Sauvignon

grape berry development. BMC Genomics, 8, 429. http://doi.org/10.1186/1471-2164-8-429

Deluc, L. G., Quilici, D. R., Decendit, A., Grimplet, J., Wheatley, M. D., Schlauch, K. a, …

Cramer, G. R. (2009). Water deficit alters differentially metabolic pathways affecting

important flavor and quality traits in grape berries of Cabernet Sauvignon and Chardonnay.

BMC Genomics, 10, 212. http://doi.org/10.1186/1471-2164-10-212

Dixon, R. A., Xie, D. Y., & Sharma, S. B. (2005). Proanthocyanidins - A final frontier in

flavonoid research? New Phytologist, 165(1), 9–28. http://doi.org/10.1111/j.1469-

8137.2004.01217.x

Downey, M. O., Dokoozlian, N. K., & Krstic, M. P. (2006). Cultural Practice and Environmental

Impacts on the Flavonoid Composition of Grapes and Wine: A Review of Recent Research.

American Journal of Enology and Viticulture, 57(3), 257 LP-268. Retrieved from

http://www.ajevonline.org/content/57/3/257.abstract

Ebadi, A., Sedgley, M., May, P., & Coombe, B. g. (1996). Seed development and abortion in

Vitis vinifera L., cv Chardonnay. International Journal of Plant Science, 157(6), 703–712.

http://doi.org/10.1086/297392

Page 39: Comparative genetic and metabolic characterization between

37

El-Kereamy, A., Chervin, C., Roustan, J., Cheynier, V., Souquet, J., Moutounet, M., …

Bouzayen, M. (2003). Exogenous ethylene stimulates the long-term expression of genes

related to anthocyanin biosynthesis in grape berries. Physiologia Plantarum, 119(2), 175–

182. http://doi.org/10.1034/j.1399-3054.2003.00165.x

Falginella, L., Castellarin, S. D., Testolin, R., Gambetta, G. A., Morgante, M., & Di Gaspero,

G. (2010). Expansion and subfunctionalisation of flavonoid 3’,5’-hydroxylases in the

grapevine lineage. BMC Genomics, 11(1), 562. http://doi.org/10.1186/1471-2164-11-562

Fasoli, M., Dal Santo, S., Zenoni, S., Tornielli, G. B., Farina, L., Zamboni, a., … Pezzotti, M.

(2012). The Grapevine Expression Atlas Reveals a Deep Transcriptome Shift Driving the

Entire Plant into a Maturation Program. The Plant Cell, 24(9), 3489–3505.

http://doi.org/10.1105/tpc.112.100230

Ford, C. M., Boss, P. K., & Hoj, P. B. (1998). Cloning and characterization of Vitis vinifera

UDP-glucose:flavonoid 3-O-glucosyltransferase, a homolog of the enzyme encoded by the

maize bronze-1 locus that may primarily serve to glucosylate anthocyanidins in vivo.

Journal of Biological Chemistry, 273(15), 9224–9233.

http://doi.org/10.1074/jbc.273.15.9224

Gambetta, G. A., Matthews, M. A., Shaghasi, T. H., McElrone, A. J., & Castellarin, S. D. (2010).

Sugar and abscisic acid signaling orthologs are activated at the onset of ripening in grape.

Planta, 232(1), 219–234. http://doi.org/10.1007/s00425-010-1165-2

Goes, F., Iandolino, A., Al-kayal, F., Bohlmann, M. C., Cushman, M. A., Lim, H., … Cook, D.

R. (2005). Characterizing the Grape Transcriptome . Analysis of Expressed Sequence Tags

from Multiple Vitis Species and Development of a Compendium of Gene Expression

during Berry Development 1 [ w ], 139(October), 574–597.

Page 40: Comparative genetic and metabolic characterization between

38

http://doi.org/10.1104/pp.105.065748.574

Gomez, C., Conejero, G., Torregrosa, L., Cheynier, V., Terrier, N., & Ageorges, A. (2011). In

vivo grapevine anthocyanin transport involves vesicle- mediated trafficking and the

contribution of anthoMATE transporters and GST. The Plant Journal, 67(6), 960–970.

http://doi.org/10.1111/j.1365-313X.2011.04648.x

Gomez, C., Terrier, N., Torregrosa, L., Vialet, S., Fournier-Level, A., Verries, C., … Ageorges,

A. (2009). Grapevine MATE-Type Proteins Act as Vacuolar H<sup>+</sup>-

Dependent Acylated Anthocyanin Transporters. Plant Physiology, 150(1), 402 LP-415.

Retrieved from http://www.plantphysiol.org/content/150/1/402.abstract

Goodman, C. D., Casati, P., & Walbot, V. (2004). A Multidrug Resistance – Associated Protein

Involved in Anthocyanin Transport in Zea mays, 16(July), 1812–1826.

http://doi.org/10.1105/tpc.022574.weed

Goto-Yamamoto, N., Wan, G. H., Masaki, K., & Kobayashi, S. (2002). Structure and

transcription of three chalcone synthase genes of grapevine (Vitis vinifera). Plant Science,

162(6), 867–872. http://doi.org/https://doi.org/10.1016/S0168-9452(02)00042-0

Greenspan, M. D., Shackel, K. A., & Matthews, M. A. (1994). Developmental changes in the

diurnal water budget of the grape berry exposed to water deficits. Plant, Cell and

Environment, 17(7), 811–820. http://doi.org/10.1111/j.1365-3040.1994.tb00175.x

Greer, D. H., & Weston, C. (2015). Heat stress affects flowering , berry growth , sugar

accumulation and photosynthesis of Vitis vinifera cv . Semillon ... cv . Semillon grapevines

grown in a controlled environment. Functional Plant Biology, 37(January 2010).

http://doi.org/10.1071/FP09209

Guan, L., Li, J., Fan, P., Chen, S., Fang, J., Li, S., & Wu, B. (2012). Anthocyanin Accumulation

Page 41: Comparative genetic and metabolic characterization between

39

in Various Organs of a Teinturier Cultivar ( Vitis vinifera L .) during the Growing Season,

2, 177–184. http://doi.org/10.5344/ajev.2011.11063

Guo, D.-L., Xi, F.-F., Yu, Y.-H., Zhang, X.-Y., Zhang, G.-H., & Zhong, G.-Y. (2016).

Comparative RNA-Seq profiling of berry development between table grape “Kyoho” and

its early-ripening mutant ’Fengzao’. BMC Genomics, 17(1), 795.

http://doi.org/10.1186/s12864-016-3051-1

Hardie, W. J., O’Brien, T. P., & Jaudzems, V. G. (1996). Morphology, anatomy and

development of the pericarp after anthesis in grape, Vitis vinifera L. Australian Journal of

Grape and Wine Research, 2(2), 97–142. http://doi.org/10.1111/j.1755-

0238.1996.tb00101.x

He, F., Mu, L., Yan, G. L., Liang, N. N., Pan, Q. H., Wang, J., … Duan, C. Q. (2010).

Biosynthesis of anthocyanins and their regulation in colored grapes. Molecules, 15(12),

9057–9091. http://doi.org/10.3390/molecules15129057

Hichri, I., Barrieu, F., Bogs, J., Kappel, C., Delrot, S., & Lauvergeat, V. (2011). Recent advances

in the transcriptional regulation of the flavonoid biosynthetic pathway. Journal of

Experimental Botany, 62(8), 2465–2483. http://doi.org/10.1093/jxb/erq442

Hichri, I., Heppel, S. C., Pillet, J., Léon, C., Czemmel, S., Delrot, S., … Bogs, J. (2010). The

basic helix-loop-helix transcription factor MYC1 is involved in the regulation of the

flavonoid biosynthesis pathway in grapevine. Molecular Plant, 3(3), 509–523.

http://doi.org/10.1093/mp/ssp118

Holton, T. A., & Cornish, E. C. (1995). Genetics and Biochemistry of Ant hocyanin

Biosynthesis, 7(July), 1071–1083.

Ibrahim, R. K., Bruneau, A., & Bantignies, B. (1998). Plant O -methyltransferases : molecular

Page 42: Comparative genetic and metabolic characterization between

40

analysis , common signature and classification, 1–10.

Jaakola, L. (2013). New insights into the regulation of anthocyanin biosynthesis in fruits. Trends

in Plant Science, 18(9), 477–483. http://doi.org/10.1016/j.tplants.2013.06.003

Jeong, S. T., Goto-yamamoto, N., Kobayashi, S., & Esaka, M. (2004). Effects of plant hormones

and shading on the accumulation of anthocyanins and the expression of anthocyanin

biosynthetic genes in grape berry skins, 167, 247–252.

http://doi.org/10.1016/j.plantsci.2004.03.021

Jimenez-garcia, S. N., Guevara-gonzalez, R. G., Miranda-lopez, R., Feregrino-perez, A. A.,

Torres-pacheco, I., & Vazquez-cruz, M. A. (2012). Functional properties and quality

characteristics of bioactive compounds in berries : Biochemistry , biotechnology , and

genomics. FRIN. http://doi.org/10.1016/j.foodres.2012.11.004

Kennedy, J. (2002). Understanding grape berry development. Winery and Vineyard, (August).

Kennedy, J. A., Hayasaka, Y., Waters, E. J., & Jones, G. P. (2001). Composition of Grape Skin

Proanthocyanidins at Different Stages of Berry Development, 5348–5355.

Kobayashi, S., Goto-Yamamoto, N., & Hirochika, H. (2004). Retrotransposon-induced

mutations in grape skin color. Science (New York, N.Y.), 304(5673), 982.

http://doi.org/10.1126/science.1095011

Koyama, K., & Goto-yamamoto, N. (2008). Bunch Shading During Different Developmental

Stages Affects the Phenolic Biosynthesis in Berry Skins of “ Cabernet Sauvignon ” Grapes,

133(6), 743–753.

Koyama, K., & Sadamatsu, K. (2010). Abscisic acid stimulated ripening and gene expression in

berry skins of the Cabernet Sauvignon grape, 367–381. http://doi.org/10.1007/s10142-009-

0145-8

Page 43: Comparative genetic and metabolic characterization between

41

Kuhn, N., Guan, L., Dai, Z. W., Wu, B. H., Lauvergeat, V., Gomes, E., … Delrot, S. (2014).

Berry ripening: Recently heard through the grapevine. Journal of Experimental Botany,

65(16), 4543–4559. http://doi.org/10.1093/jxb/ert395

Lacampagne, S., Gagné, S., & Gény, L. (2010). Involvement of Abscisic Acid in Controlling

the Proanthocyanidin Biosynthesis Pathway in Grape Skin: New Elements Regarding the

Regulation of Tannin Composition and Leucoanthocyanidin Reductase (LAR) and

Anthocyanidin Reductase (ANR) Activities and Expres. Journal of Plant Growth

Regulation, 29(1), 81–90. http://doi.org/10.1007/s00344-009-9115-6

Lamikanra, O. (1989). Anthocyanins of Vitis rotundifolia hybrid grapes. Food Chemistry, 33(3),

225–237. http://doi.org/https://doi.org/10.1016/0308-8146(89)90016-2

Maeda, H., & Dudareva, N. (2012). The shikimate pathway and aromatic amino Acid

biosynthesis in plants. Annual Review of Plant Biology, 63, 73–105.

http://doi.org/10.1146/annurev-arplant-042811-105439

Marinova, K., Pourcel, L., Weder, B., Schwarz, M., Barron, D., Routaboul, J., … Klein, M.

(2007). The Arabidopsis MATE Transporter TT12 Acts as a Vacuolar Flavonoid / H 1 -

Antiporter Active in Proanthocyanidin-Accumulating Cells of the Seed Coat, 19(June),

2023–2038. http://doi.org/10.1105/tpc.106.046029

Matus, J. T., Aquea, F., & Arce-Johnson, P. (2008). Analysis of the grape MYB R2R3 subfamily

reveals expanded wine quality-related clades and conserved gene structure organization

across Vitis and Arabidopsis genomes. BMC Plant Biology, 8(1), 83.

http://doi.org/10.1186/1471-2229-8-83

Matus, J. T., Loyola, R., Vega, A., Pen, A., Bordeu, E., & Alcalde, A. (2009). Post-veraison

sunlight exposure induces MYB-mediated transcriptional regulation of anthocyanin and

Page 44: Comparative genetic and metabolic characterization between

42

flavonol synthesis in berry skins of Vitis vinifera. Journal of Experimental Botany, 60(3),

853–867. http://doi.org/10.1093/jxb/ern336

Matus, J. T., Poupin, M. J., Cañón, P., Bordeu, E., Alcalde, J. A., & Arce-Johnson, P. (2010).

Isolation of WDR and bHLH genes related to flavonoid synthesis in grapevine (Vitis

vinifera L.). Plant Molecular Biology, 72(6), 607–620. http://doi.org/10.1007/s11103-010-

9597-4

Mazza, G. (1995). Anthocyanins in grapes and grape products. Critical Reviews in Food Science

and Nutrition, 35(4), 341–371. http://doi.org/10.1097/NT.0b013e31823db374

Pensec, F., Pączkowski, C., Grabarczyk, M., Woźniak, A., Bénard-Gellon, M., Bertsch, C., …

Szakiel, A. (2014). Changes in the Triterpenoid Content of Cuticular Waxes during Fruit

Ripening of Eight Grape ( Vitis vinifera ) Cultivars Grown in the Upper Rhine Valley.

Journal of Agricultural and Food Chemistry, 62(32), 7998–8007.

http://doi.org/10.1021/jf502033s

Peppi, M. C., & Fidelibus, M. W. (2008). Effects of Forchlorfenuron and Abscisic Acid on the

Quality of “ Flame Seedless ” Grapes, 43(1), 173–176.

Rinaldo, A. R., Cavallini, E., Jia, Y., Moss, S. M. A., Mcdavid, D. A. J., Hooper, L. C., …

Walker, A. R. (2015). A Grapevine Anthocyanin Acyltransferase , Transcriptionally

Regulated by VvMYBA , Can Produce Most Acylated Anthocyanins Present in Grape

Skins 1, 169(November), 1897–1916. http://doi.org/10.1104/pp.15.01255

Ristic, R., & Iland, P. G. (2005). Relationships between seed and berry development of Vitis

Vinifera L. cv Shiraz: Developmental changes in seed morphology and phenolic

composition. Australian Journal of Grape and Wine Research, 11(1), 43–58.

http://doi.org/10.1111/j.1755-0238.2005.tb00278.x

Page 45: Comparative genetic and metabolic characterization between

43

Serrano, A., Espinoza, C., Armijo, G., Inostroza-Blancheteau, C., Poblete, E., Meyer-Regueiro,

C., … Arce-Johnson, P. (2017). Omics Approaches for Understanding Grapevine Berry

Development: Regulatory Networks Associated with Endogenous Processes and

Environmental Responses. Frontiers in Plant Science, 8(September), 1–15.

http://doi.org/10.3389/fpls.2017.01486

Shimazaki, M., Fujita, K., Kobayashi, H., & Suzuki, S. (2011). Pink-colored grape berry is the

result of short insertion in intron of color regulatory gene. PLoS ONE, 6(6), 1–8.

http://doi.org/10.1371/journal.pone.0021308

Singh, H., Singh, Z., Swinny, E., & Cameron, I. (2008). Plant Science Girdling and grapevine

leafroll associated viruses affect berry weight , colour development and accumulation of

anthocyanins in “ Crimson Seedless ” grapes during maturation and ripening, 175, 885–

897. http://doi.org/10.1016/j.plantsci.2008.09.005

Sparvoli, F., Martin, C., Scienza, A., Gavazzi, G., & Tonelli, C. (1994). Cloning and molecular

analysis of structural genes involved in flavonoid and stilbene biosynthesis in grape (Vitis

vinifera L.). Plant Molecular Biology, 24(5), 743–755. http://doi.org/10.1007/BF00029856

Springob, K., Nakajima, J., Yamazaki, M., & Saito, K. (2003). Recent advances in the

biosynthesis and accumulation of anthocyanins, 288–303.

http://doi.org/10.1039/b109542k

Sun, L., Zhang, M., Ren, J., Qi, J., Zhang, G., & Leng, P. (2010). Reciprocity between abscisic

acid and ethylene at the onset of berry ripening and after harvest. BMC Plant Biology,

10(1), 257. http://doi.org/10.1186/1471-2229-10-257

Sweetman, C., Wong, D. C., Ford, C. M., & Drew, D. P. (2012). Transcriptome analysis at four

developmental stages of grape berry (Vitis vinifera cv. Shiraz) provides insights into

Page 46: Comparative genetic and metabolic characterization between

44

regulated and coordinated gene expression. BMC Genomics, 13(1), 691.

http://doi.org/10.1186/1471-2164-13-691

Tanaka, Y., Sasaki, N., & Ohmiya, A. (2008). Biosynthesis of plant pigments : anthocyanins ,

betalains and carotenoids. The Plant Journal, 54(4), 733–749.

http://doi.org/10.1111/j.1365-313X.2008.03447.x

This, P., Lacombe, T., Cadle-Davidson, M., & Owens, C. L. (2007). Wine grape (Vitis vinifera

L.) color associates with allelic variation in the domestication gene VvmybA1. Theoretical

and Applied Genetics, 114(4), 723–730. http://doi.org/10.1007/s00122-006-0472-2

Tian, L., Pang, Æ. Y., & Dixon, Æ. R. A. (2008). Biosynthesis and genetic engineering of

proanthocyanidins and ( iso ) flavonoids, 445–465. http://doi.org/10.1007/s11101-007-

9076-y

Vega, A., Gutierrez, R., Peña-Neira, A., Cramer, G., & Arce-Johnson, P. (2011). Water Deficit

Increases Stilbene Metabolism in Cabernet Sauvignon Berries. Plant Molecular Biology,

77, 289–297. http://doi.org/10.1021/jf1024888

Walker, A. R., Lee, E., & Robinson, S. P. (2006). Two new grape cultivars, bud sports of

Cabernet Sauvignon bearing pale-coloured berries, are the result of deletion of two

regulatory genes of the berry colour locus. Plant Molecular Biology, 62(4–5), 623–635.

http://doi.org/10.1007/s11103-006-9043-9

Wang, H., Race, E. J., & Shrikhande, A. J. (2003). Characterization of Anthocyanins in Grape

Juices by Ion Trap Liquid Chromatography−Mass Spectrometry. Journal of Agricultural

and Food Chemistry, 51(7), 1839–1844. http://doi.org/10.1021/jf0260747

Wang, H., Wang, W., Zhan, J., Yan, A., Sun, L., Zhang, G., … Xu, H. (2016). The accumulation

and localization of chalcone synthase in grapevine (Vitis vinifera L.). Plant Physiology and

Page 47: Comparative genetic and metabolic characterization between

45

Biochemistry. Elsevier Ltd. http://doi.org/10.1016/j.plaphy.2016.04.042

Wheeler, S., Loveys, B., Ford, C., & Davies, C. (2009). The relationship between the expression

of abscisic acid biosynthesis genes , accumulation of abscisic acid and the promotion of

Vitis vinifera L . berry ripening by abscisic acid. http://doi.org/10.1111/j.1755-

0238.2008.00045.x

Winkel-Shirley, B. (2001). Flavonoid Biosynthesis. A Colorful Model for Genetics,

Biochemistry, Cell Biology, and Biotechnology. Plant Physiology, 126(2), 485 LP-493.

Retrieved from http://www.plantphysiol.org/content/126/2/485.abstract

Yamane, T., Jeong, S. T., Goto-yamamoto, N., Koshita, Y., & Kobayashi, S. (2006). Effects of

Temperature on Anthocyanin Biosynthesis in Grape Berry Skins, 1, 54–59.

Yamazaki, M., Yamagishi, E., Gong, Z., Fukuchi-Mizutani, M., Fukui, Y., Tanaka, Y., … Saito,

K. (2015). Two flavonoid glucosyltransferases from Petunia hybrida : Molecular cloning ,

biochemical properties and ... Plant Molecular Biology, 48(February).

http://doi.org/10.1023/A:1014043214943

Zhang, H., Wang, L., Deroles, S., Bennett, R., & Davies, K. (2006). New insight into the

structures and formation of anthocyanic vacuolar inclusions in flower petals, 14, 1–14.

http://doi.org/10.1186/1471-2229-6-29

Zhang, Y., Butelli, E., & Martin, C. (2014). ScienceDirect Engineering anthocyanin

biosynthesis in plants. Current Opinion in Plant Biology, 19, 81–90.

http://doi.org/10.1016/j.pbi.2014.05.011

Zhao, C.-L., Yu, Y.-Q., Chen, Z.-J., Wen, G.-S., Wei, F.-G., Zheng, Q., … Xiao, X.-L. (2017).

Stability-increasing effects of anthocyanin glycosyl acylation. Food Chemistry,

214(Supplement C), 119–128.

Page 48: Comparative genetic and metabolic characterization between

46

http://doi.org/https://doi.org/10.1016/j.foodchem.2016.07.073

Zhao, J., Pang, Y., & Dixon, R. A. (2010). The Mysteries of Proanthocyanidin Transport and

Polymerization. Plant Physiology, 153(2), 437 LP-443. Retrieved from

http://www.plantphysiol.org/content/153/2/437.abstract

Ziliotto, F., Corso, M., Rizzini, F. M., Rasori, A., Botton, A., & Bonghi, C. (2012). Grape berry

ripening delay induced by a pre-véraison NAA treatment is paralleled by a shift in the

expression pattern of auxin- and ethylene-related genes. BMC Plant Biology, 12(1), 185.

http://doi.org/10.1186/1471-2229-12-185

Page 49: Comparative genetic and metabolic characterization between

47

CHAPTER II

PROBLEM STATMENT: Hypothesis and Objectives

Page 50: Comparative genetic and metabolic characterization between

48

According to all the previously mentioned in the General Introduction, I propose the following

hypothesis of my thesis:

Color differences in berry skin between Chimenti Globe and Red Globe are due to

metabolic and transcriptomic changes related with the branching point F3’-H and F3’5’-

H of the anthocyanin biosynthesis pathway

The main objective of this thesis was:

To determine transcriptomic and metabolic changes associated with the branching point F3’-H

and F3’5’-H of the anthocyanin biosynthesis pathway, responsible for the color differences in

berry skin between Chimenti Globe (CG) and Red Globe (RG).

The general objective was divided into the following partial objectives:

1. Characterize metabolic changes related to differences in anthocyanin accumulation

between CG and RG.

2. Analyze transcriptomic changes involved in differences in anthocyanin accumulation

between CG and RG.

3. Evaluate the function of at least one gene of interest, selected from the above objectives,

related with the branching point F3’-H and F3’5’-H of the anthocyanin biosynthesis

pathway.

The objectives 1 and 2 have been developed in the Chapter 3 and objective 3 in Chapter 4.

Page 51: Comparative genetic and metabolic characterization between

49

CHAPTER III

TRANSCRIPTOMIC AND METABOLIC ANALYSIS REVEALS NEW

GENETIC DIFFERENCES BETWEEN TWO TABLE GRAPES WITH CONTRASTED

COLOR BERRY SKIN

This article will be submitted to Frontiers in Plant Science

Page 52: Comparative genetic and metabolic characterization between

50

TITLE:

TRANSCRIPTOMIC AND METABOLIC ANALYSIS REVEAL NEW

GENETIC DIFFERENCES BETWEEN TWO TABLE GRAPES WITH CONTRASTED

COLOR BERRY SKIN

Claudia Santibáñez1,2, Carlos Meyer1, Litsy Martínez1, Tomás Moyano3, John Lunn4, Regina

Feil4, Zhanwu Dai2, Ghislaine Hilbert2, Christel Renaud2, Serge Delrot2, Patricio Arce1* and

Eric Gomès2*

1 Laboratorio de Biología Molecular y Biotecnología Vegetal, Departamento de Genética

Molecular y Microbiología, Facultad de Ciencias Biológicas, Pontificia Universidad Católica

de Chile, Santiago 8331150, Chile.

2 Université de Bordeaux, ISVV, INRA, UMR 1287 Ecophysiology and Functional Genomic of

Grapevine, Villenave d’Ornon, France.

3 FONDAP Center for Genome Regulation, Millennium Nucleus Center for Plant Systems and

Synthetic Biology, Departamento de Genética Molecular y Microbiologia, Facultad de Ciencias

Biológicas, Ponticia Universidad Católica de Chile, Santiago 8331150, Chile

4 Max Plant Institute of Molecular Plant Physiology, Wissenschaftspark Golm, Am Mühlenberg

1, 14476 Potsdam (OT) Golm, Germany.

*Correspondence: [email protected] , [email protected]

Page 53: Comparative genetic and metabolic characterization between

51

ABSTRACT

Anthocyanins are pigments responsible of grape berry skin coloration and its biosynthesis

regulation has been widely studied. However, few studies focus on metabolic and genetic

characterization of such metabolism using varieties of the same genetic background with

contrasting berry skins color. Using metabolic profiling and RNA-seq technology, we

performed a comparative characterization of two table grapes, Chimenti Globe (CG) and Red

Globe (RG), with contrasting color berry skin: CG has a bright light red color and RG has a

purple color. The originality of this model is that CG was generated from a spontaneous

mutation event in field from a branch in a vine of RG plant, allowing us to study anthocyanin

metabolism regulation in near isogenic backgrounds. Berry metabolic content analysis showed

the importance of developmental stages, veraison and ripening, in both varieties of the study.

Striking differences in metabolite concentrations of the phenylpropanoid pathway, such as

shikimate, phenylalanine and other molecules such as UDP-glucose and trehalose-6-phosphate,

among others, the latter being a signalling molecule with major implications in plant growth,

development, and metabolism regulation. Also, in contrast to RG, CG only contained

dihydroxylated anthocyanins, peonidin and cyanidin, and not the trihydroxylated ones,

malvidin, delphinidin and petunidin, which was consistent with the observed color phenotype.

From transcriptomic analysis using Illumina Hiseq200, we generated a heat map with 109 genes

differentially expressed in CG in comparison to RG, many of these being related to the flavonoid

pathway. In addition, 11 gene copies of flavonoid 3’5’-hydroxylase, a key enzyme for the

biosynthesis of trihydroxylated anthocyanins, were repressed in CG. These results, provide new

Page 54: Comparative genetic and metabolic characterization between

52

insights in anthocyanins biosynthesis regulation in grape, offering new clues in the search for

molecular regulators of the flavonoid pathway.

Keywords: RNA-seq, skin, anthocyanin, grape, Red Globe, Chimenti Globe.

Page 55: Comparative genetic and metabolic characterization between

53

INTRODUCTION

Grapevine (Vitis vinifera L.) is the most important non-climacteric fruit crop worldwide, used

for fresh consumption as table grapes, dried into raisins and processed into wine and liquors.

Berry development is a process displaying a double sigmoid curve with two growth phases

separated by a lag phase. In the first phase a fruit growth is observed, during which organic

acids, tannins, hydroxycinnamates and several phenolic compound precursors are synthesized

and the number of fruit cells is defined (Adams, 2006). Then, there is a lag phase where there is

no significant growth of the fruit. At the end of this phase a phenomenon called veraison take

place, where the fruit initiates its ripening and begins to take color (COOMBE, 1995). In the

third phase, a second growth of the berry occurs, organic acids levels decrease (mostly due to

malate breakdown), sugars and secondary metabolites (aroma, phenolics) are synthetized and

accumulated, to eventually end up with a fruit exhibiting desirable quality traits (Conde et al.,

2007).

Anthocyanins are secondary metabolites and one of the main pigments present in red grape berry

skin (Souquet, Cheynier, Brossaud, & Moutounet, 1996), which are associated to organoleptic

properties in red wine such as color, bitterness, astringency, and are also antioxidants with

beneficial effects on human health (Dixon, Xie, & Sharma, 2005). The biosynthesis of

anthocyanins occurs via the flavonoid pathway at the cytosolic surface of the endoplasmic

reticulum, in grape cells (P. K. Boss, Davies, & Robinson, 1996). Most of the structural genes

encoding the enzymes of the anthocyanin biosynthesis pathway have been identified and

characterized (Kuhn et al., 2014). The regulation of these genes is achieved by ternary

Page 56: Comparative genetic and metabolic characterization between

54

complexes, formed by transcription factors from the R2R3-MYB (L. Deluc et al., 2006, 2008;

Kobayashi, Goto-Yamamoto, & Hirochika, 2004; This, Lacombe, Cadle-Davidson, & Owens,

2007) and bHLH (basic helix–loop–helix) families and WDR (tryptophan–aspartic acid repeat)

proteins (Hichri et al., 2010; Matus et al., 2010).

In the last decade, several studies demonstrated that transcriptomic remodeling drives the

transition from the vegetative to the ripening stage, a process which is highly coordinated at the

transcriptional level (Fasoli et al., 2012). There has been advances in deep-sequencing

technologies that, integrated with the availability of grapevine genomic and transcriptomic data,

allow to carry out RNA-transcript expression analysis on a global scale at key points during

grapevine development (Serrano et al., 2017). However, most of these studies focus on

characterization of one single cultivar such as Cabernet Sauvignon (L. G. Deluc et al., 2007);

comparison between red cultivars from contrasted genotypes (Guo et al., 2016; Sweetman,

Wong, Ford, & Drew, 2012); or between red and white cultivars (Yakushiji et al., 2006). They

do not, however, take into account the full range of berry color shades that is present in V.

vinifera cultivars, nor do they consider the comparison or closely related, near-isogenic

cultivars.

Here, we performed a metabolic and transcriptomic characterization of berry development and

ripening using as a model two closely related varieties with different berry color skin. The first

one corresponds to Red Globe (RG) variety, which is a table grape whose berries are of a large

size, thick and consistent skin of dark purple color. The second one, Chimenti Globe (CG), has

similar characteristics but differs in color skin that is bright light red. It originates from a

Page 57: Comparative genetic and metabolic characterization between

55

spontaneous mutation event in field from a branch in a vine of RG plant in Talagante,

Metropolitan Region of Chile that was propagated through clonal multiplication (FIGURE 1C).

The interest of this model is the genetic background responsible for the color change is the

almost identical, with differences affecting the color of the berry therefore allowing us to study

the process of accumulation of anthocyanins using two near-isogenic varieties with contrasting

berry skin color.

Page 58: Comparative genetic and metabolic characterization between

56

MATERIALS AND METHODS

Plant material and experimental design

Plant material used in this study corresponds to six plants of commercial table grape cv. ‘Red

Globe’ (RG) and six plants of cv. ‘Chimenti Globe’ (CG), located in Camino Loreto, Parcela

8-9, Talagante, Región Metropolitana, Santiago of Chile. All plants were in the same plot, one

plant next to the other (FIGURE 1). Sampling were performed during seasons 2013 and 2014.

Veraison stage was classically determined as the time at which clusters were 30–50% colored,

and ripening stage when berry sugar content reached 22-23°Brix (22-23% w/w soluble solids).

A random cluster for each sampling stage representing biological unit was selected for each

plant in both cultivars, stage and season. Berries were immediately peeled and frozen in liquid

nitrogen, and stored at –80°C until required for further analysis.

Metabolites extraction and analysis

For metabolite analysis, berry skin from 2013 and 2014 season were used and each sample was

analyzed separately, except for anthocyanin analysis where were used only five samples for

each variety.

Anthocyanins were extracted from 300 mg freeze-dried ground powder from skin of RG and

CG using 1 ml methanol containing 0.1% HCl (v/v). Extracts were filtered across 0.45 μm

polypropylene syringe filter (Pall Gelman Corp., Ann Harbor, MI, USA) for HPLC analysis.

Each individual anthocyanin was analyzed as described in (Dai et al., 2013) using HPLC. For

anthocyanin quantification was integrated peak area at 520 nm and using Malvidin 3-glucoside

as standard (Extrasynthese, Lyon, France).

Page 59: Comparative genetic and metabolic characterization between

57

Primary metabolites such as Tre6P, glycolytic and TCA cycle intermediates were extracted from

25 mg of frozen powder with chloroform/methanol solution, and measured by high-performance

anion-exchange liquid chromatography coupled to tandem mass spectrometry (LC-MS/MS) as

described by (John E. Lunn et al., 2006). Amino-acids were extracted and analyzed by high

performance liquid chromatography as described by (Martínez-Lüscher et al., 2014).

For sugars and organic acids, an aliquot of 150-200 mg of skin fine powder was extracted

sequentially with ethanol 80% and 50%, dried in a SpeedVac and re-dissolved in 1 ml of sterile

ultrapure water. Hexose content (glucose and fructose) were measured enzymatically with an

automated micro-plate reader (Elx800UV, Biotek Instruments Inc., Winooski, VT, USA)

according to (Gomez, Bancel, Rubio, & Vercambre, 2007). Tartaric acid content was measured

by using the colorimetric method based on ammonium vanadate reactions (Pereira et al., 2006).

Malic acid content was determined using an enzyme-coupled spectrophotometric method that

measures the change in absorbance at 340 nm from the reduction of NAD+ to NADH (Pereira

et al., 2006).

RNA extraction

Total RNA were isolated from berry skins according to the procedure of (Reid, Olsson,

Schlosser, Peng, & Lund, 2006), using a CTAB-spermidine extraction buffer. This material was

used for transcriptome sequencing by RNA-seq technology and validation of data by

quantitative real-time RT-PCR (qRT-PCR). For RNA sequencing, only berry skin from 2013

season was used and to carry out this analysis was necessary to reduce all material into 3 pools

per cultivar (CG and RG) and stage (veraison and ripening), reaching a total number of 12

Page 60: Comparative genetic and metabolic characterization between

58

samples for this study. Each pool was made by mixing 2 individual RNA extractions (1 µg each

one) and then mixing them to obtain a concentration of 2 µg per pool (TABLE S1).

Total RNA were then sent to Macrogen (Macrogen Inc. Seoul, Korea) according to their delivery

protocol consisting in ethanol precipitation condition. Briefly, it was added 0.1 volume of 3 M

Sodium Acetate pH 7-8 to total RNA solution, and mixed gently; then it was added 2 volumes

of 100% ethanol, leaving the samples reading to be sent.

RNA Sequencing and Reads Mapping

mRNA were isolated from 10 μg of total RNA, fragmented, converted to cDNA, and amplified

by PCR to Illumina ® TruSeqTM RNA Sample Preparation Kit (Illumina, Inc., USA). Pair-end

100bp sequence reads were generated using the Illumina Genome Analyzer II (Illumina) and

Illumina HiSeq 2000 (Illumina) at Macrogen Inc. (https://dna.macrogen.com/) according to the

manufacturer’s recommendations. Trimmomatic (Bolger, Lohse, & Usadel, 2014) was used to

clean the reads prior to mapping and to remove the Illumina adapter sequences as well as the

low-quality sequences from the ends of the reads. All the distinct clean reads were aligned to

the Grape Genome Database hosted at CRIBI V2 (http://genomes.cribi.unipd.it/grape/) (Vitulo

et al., 2014). Uniquely mapped reads were counted by HITSAT2 software (Kim, Langmead, &

Salzberg, 2015) and featureCounts in the Rsubread package (Liao, Smyth, & Shi, 2013).

Screening of DEGs

Analyses for discovering differentially expressed genes (DEGs) were performed with DEseq2

using false discovery rate (FDR) with a threshold of 0.05 and absolute value log2 ratio ≥ 1 (Love,

Huber, & Anders, 2014). MultiExperiment Viewer (MeV) was used for gene clustering analysis

Page 61: Comparative genetic and metabolic characterization between

59

that was performed by the k-means method with Pearson’s correlation distance. MeV also was

used for graphical representation of DEGs and it was performed a heatmap using Fold Change

≥ 1 (Howe, Sinha, Schlauch, & Quackenbush, 2011).

PANTHER™ Version 12.0 program (Mi et al., 2017) with default parameters was used to

perform the GO functional enrichment analysis with the Bonferroni correction for multiple

testing (P ≤ 0.05). Corrected P-values < 0.05 were used to obtain significant GO enrichment.

Validation of RNA-seq data by quantitative real-time RT-PCR (qRT-PCR)

Two ug of total RNA were treated with TURBO DNA-freeTM DNase (Ambion®) and were

reverse transcribed with random primers using SuperScript II RT (InvitrogenTM Co., Carlsbad,

CA, USA), according to the manufacturer’s instructions. Relative transcript quantification of

DEGs was performed by real-time RT-PCR (qRT-PCR) using the BRILLIANT II SYBR®

GREEN QPCR Master Mix and the Mx3000P qPCR system (Stratagene, Agilent Technologies

Inc., Santa Clara, CA, USA) according to the manufacturer’s protocol. Expression levels of all

evaluated genes were calculated from six biological replicates, relative to Vv60SRP

housekeeping control gene. We used the 2-∆∆Ct method for the statistical analysis (Schefe,

Lehmann, Buschmann, Unger, & Funke-Kaiser, 2006).

Pimers used in qRT-PCR for DEGs correspond to F3’5’H multicopies (IDs:

VIT_206s0009g02805, VIT_206s0009g02810, VIT_206s0009g02840, VIT_206s0009g02860,

VIT_206s0009g02873, VIT_206s0009g02880, VIT_206s0009g02920, VIT_206s0009g02970

and VIT_206s0009g03010): F5’-TGGATAAAGTGATTGGAAGGA-3’ and R5’-

GGATGGGTGCTTCCGGAAGCTT-3’; CYTB5 (ID: VIT_218s0001g09400): F5’-

GGCTCCCCAAAAAGTTTTCA-3’ and R5’-TGCCATGAATGACGAACCAG-3’;

UFGT2230 (ID: VIT_216s0039g02230): F5’-GCTTTGATCAAATTCTCTCACA-3’ and R5’-

Page 62: Comparative genetic and metabolic characterization between

60

ACACTAAATCCACCAGGGTTTT-3’; CCOAMT (IDs: VIT_201s0010g03490 and

VIT_201s0010g03510): F5’-TTGATCCAGATAAAGAAGCGTA-3’ and R5’-

CGGCAATGAGATCATTTAGAAC-3’; ANP-like (IDs: VIT_216s0050g00900 and

VIT_216s0050g00910): F5’-ATTCGGGTCTCCAATGAGCT-3’ and R5’-

GTGAGAAACAGCCTCTTGCA-3’; GST13 (ID: VIT_207s0104g01800): F5’-

AATTCCCATGTCCACATGCACC-3’ and R5’-GGGATTCTTGGCAAGAAAAGGA -3’;

F3H (ID: VIT_204s0023g03370): F5’-AGGCAACCGTGTATCCTCTG-3’ and R5’-

TTCTCCACGTCTTGCAACTG-3’; CHS (ID: VIT_205s0136g00260): F5’-

CCCATGTTCGAATTGGTTTC-3’ and R5’- TGTGCAATCCAGAAAATGGA-3’; MYB24

(ID: VIT_214s0066g01090): F5’-ACCCTGCAATTCTCAGGATG-3’ and R5’-

CAGGCCTCAGGTAGTTCAGC-3’; COMT-like (IDs: VIT_215s0048g02480.1,

VIT_215s0048g02490.1 and VIT_215s0048g02490.2): F5’-

TACCCTGGCAAGGACCCCAGAT-3’ and R5’-AGGTGCTCAAAACCCTTGTAGG-3’;

VvABCTRANSPORTER (ID: VIT_214s0060g00720): F5’-AATGTGAAGAGCGAGGTGCT-

3’ and R5’-CGACTGCCTTTCGTTCTTTC-3’; VvSRP60 (ID: VIT_05s0077g02060): F5’-

ATCTACCTCAAGCTCCTAGTC-3’ and 5’-CAATCTTGTCCTCCTTTCCT-3’. Also were

analyzed 6 gene copies of F3’5’H: VIT_206s0009g02810, F5’-

AGCATGGTGGTGGCCTCAACTC-3’ and R5’-TAGCTTTCTCAGTAGCTTCCAC-3’;

VIT_206s0009g02830, F5’- GACCGGTTATTAACAAAGATGA-3’ and R5’-

TGTCTGTTCCAGGAGGAGTGCT-3’; VIT_206s0009g02860, F5’-

CACTAAGCCATGGCCATAGATA-3’ and R5’- GCAACATGAGGCATGTTACCTA-3’;

VIT_206s0009g02880, F5’-AGTCAAATGAGTTCAAGGACATGG-3’ and R5’-

AGGAGCACACGGCGTCAGCCCA-3’; VIT_206s0009g02920, F5’-

Page 63: Comparative genetic and metabolic characterization between

61

TTTGGATCCTTCAAATATTC-3’ and R5’-GAAAAGATGGAGATGGCTCT-3’;

VIT_206s0009g03010, F5’-TAACGGTATAAGATGTATTAGC-3’and R5’-

GAAAAATAGAAACTATAACCAA-3’.

Statistical data analysis

Data were analyzed with multivariate analysis methods using the R statistics environment (R

Core Team, 2010). In order to evaluate alterations in metabolite levels during stages of

grapevine development, principal component analysis (PCA) was performed on mean-centered

and scaled data using the ade4 package in R (Dray & Dufour, 2007). Differences between

development stages and seasons were analyzed by a two-way ANOVA followed with a Tukey

multiple comparison post-hoc test at P < 0.05, for example, in the case of PCA analysis,

component 1. For differences between varieties were analyzed using unpaired T-test, for

example in the case of PCA analysis, component 2 and qRT-PCR analysis. GraphPad Prism

version 6.00 for Windows (GraphPad Software, La Jolla California USA, www.graphpad.com)

was used for graphics representation and analysis. For RNA-seq analysis, Trimmomatic (Bolger

et al., 2014), HISAT2 (Kim et al., 2015) and Rsubread package (Liao et al., 2013) were used

with default parameters. DEGs analysis were performed with DEseq2 using FDR with a

threshold of 0.05 and absolute value log2 ratio ≥ 1 (Love et al., 2014). (MeV) was used for

cluster analysis by the k-means method with Pearson’s correlation distance and graphical

representation of DEGs with a heatmap using FC ≥ 1 (Howe et al., 2011). PANTHER™ Version

12.0 program (Mi et al., 2017) with default parameters was used to perform the GO analysis

with the Bonferroni correction for multiple testing (P ≤ 0.05). We used corrected P-values <

0.05 to obtain a significant enrichment of the genes.

Page 64: Comparative genetic and metabolic characterization between

62

RESULTS

Metabolic characterization of Chimenti Globe and Red Globe

Primary metabolite profiling of RG and CG

Principal component analysis (PCA) was conducted to obtain a comprehensive view of

distribution of metabolites during veraison and ripening stages, between both cultivars on study,

CG and RG, in two seasons evaluated, 2013 and 2014. The first two principal components (PC1

and PC2) explained about 64% of the total variance and allowed to discriminate developmental

stages between CG and RG (FIGURE 2A). PC1 (43,75%) separated developmental stages,

veraison and ripening in both seasons evaluated. This discrimination was mainly determined by

the concentrations of metabolites related with phenylpropanoid metabolism such as shikimate,

UDP-glucose and phenylalanine. In addition, it was observed that the molecular regulator

trehalose-6-phosphate and TCA cycle intermediates such as citrate, isocitrate and other several

amino-acids also contributed to separated veraison from ripening (FIGURE 2B and FIGURE

S1). PC2 (20,26%) resolved cultivar type, CG and RG, even in 2013 and 2014 years, but this

discrimination only was observed at the ripening stage. These results were explained in changes

observed in metabolism related with biosynthesis of sucrose: glyceraldehyde 3-phosphate,

fructose 6-phosphate, glucose 6-phosphate, fructose 1,6-biophosphate, glucose 1,6-

biophosphate, glycerol 3-phosphate, fructose, glucose and phosphoenolpyruvate; and also, by

anthocyanins compounds (FIGURE 2B and FIGURE S2).

Page 65: Comparative genetic and metabolic characterization between

63

Anthocyanin contents of RG and CG berry skins

We analyzed total anthocyanin levels present in berries skins from CG and RG cultivars during

same developmental stages and seasons studied before. During veraison stage, CG showed

almost null concentration of any types of anthocyanins, and in ripening we observed significant

differences in total anthocyanin concentration between CG and RG. RG contains about 7 times

more of total anthocyanins than CG, in samples from both seasons (FIGURE 3A and 3C,

respectively). For example, in ripening stage 2013, RG had 14.39 mg.g-1 DW and CG 1.93

mg.g-1 DW of total of anthocyanin content (TABLE 1). Detailed analysis of individual

anthocyanins revealed that CG mainly contains only two types of anthocyanins: peonidin (in the

form of peonidin 3-glucoside and peonidin 3-(p-coumaryl)-glucoside) and cyanidin (in the form

of cyanidin 3-glucoside and cyanidin 3-(p-coumaryl)-glucoside) which confers the most reddish

tonality to grapevines berries (FIGURE 3B and 3D). The others 3 types of anthocyanins were

present in minute amounts (or even no detectable at all) at both developmental stages and for

both seasons analyzed. These corresponded to malvidin (in the form of malvidin 3-glucoside),

petunidin (in the form of petunidin 3-glucoside) and delphinidin (in the form of delphinidin 3-

glucoside and delphinidin 3-(acetyl)-glucoside), which are bluish anthocyanins (TABLE 1).

Conversely, in RG we observed all 5 types of anthocyanins mentioned above, in higher

concentrations than in the CG cultivar, both at veraison and ripening stages in 2013 and 2014

samples (FIGURE 3B and 3D, respectively).

Page 66: Comparative genetic and metabolic characterization between

64

Transcriptomic characterization between Chimenti Globe and Red Globe

RNA-Seq Data mapping and alignment to Grapevine Genome Database

Based on the model study used in this work, we proposed that the gene expression program

responsible for changes in berry skin color for CG could have been differentially altered

compared to RG. To investigate this hypothesis, we performed a genome-wide transcriptome

comparison using Illumina Hiseq2000, and NGS technology. Twelve RNA samples were

analyzed corresponding to pools of biological replicates of CG and RG, from veraison and

ripening stages during 2013 season (TABLE S1). From these samples, we obtained an average

of ~4.8 – 8.1 million total sequencing reads pairs, for each sample (TABLE S3). These total

reads were cleaned from adapters and removed from low-quality bases from the ends of the

reads using Trimmomatic (Bolger et al., 2014). and were aligned to Grapevine Reference

Genome hosted at CRIBI V2 (http://genomes.cribi.unipd.it/grape/) (Vitulo et al., 2014) using

HISAT2 (Kim et al., 2015) and Rsubread package (Liao et al., 2013) (TABLE S3).

Approximately ~3.3 – 4.9 million reads aligned 1 time onto the genome, ~0.5 – 0.8 million reads

aligned 0 times and ~0.7 – 1.1 million reads aligned more than 1 time (72%, 11-20% and 15-

17% of the total reads after quality analysis, respectively) (TABLE S2). The reads that aligned

1 time to genome were used for following analysis.

Chimenti Globe exhibited altered flavonoid metabolism compared to Red Globe

To further explore the expression profiles of genes between CG and RG cultivars, all clean reads

were aligned to the grape genome database V2 (Vitulo et al., 2014) and then a comparative

transcriptomic analysis was performed to screen for DEGs using the DEseq2 program (Love et

Page 67: Comparative genetic and metabolic characterization between

65

al., 2014). As shown in FIGURE 4, a heat map was generated where a total of 109 genes were

differentially expressed in CG in comparison to RG, during veraison and ripening stages, in the

2013 samples. At veraison a total of 64 genes were down-regulated and 45 genes were up-

regulated in CG cultivar; and at ripening stage, 66 genes were down-regulated and 43 genes

were up-regulated in CG cultivar, compared to RG.

Many of the 109 DEGs were related to flavonoid metabolism and were categorized into six

clusters: cluster A (1 gene), cluster B (23 genes) and cluster C (31 genes) that were not induced

in CG at both developmental stages. Flavonoid metabolism genes such as flavonoid 3,5-

hydroxylases (VIT_206s0009g02805-2810-2830-2840-2846-2860-2873-2880-2970-2920-

3010) flavanone 3-hydroxylase (VIT_204s0023g03370) glutathione transferases

(VIT_204s0079g00690-710), chalcone synthase (VIT_205s0136g00260), cytochrome b5

(VIT_218s0001g09400), caffeoyl-CoA O-methyltransferase (VIT_201s0010g03490-3510) and

UDP-glucose:flavonoid 3-O-glucosyltransferase (VIT_216s0039g02230) fall in that category

(TABLE 2). Cluster E (34 genes) encompass genes that were induced at both developmental

stages in CG such as cytochrome p450 (VIT_201s0137g00540), catechol-O-methyltransferases

(VIT_215s0048g02480-2490), MYB24 (VIT_214s0066g01090) and ABC TRANSPORTER

(VIT_214s0060g00720) (TABLE 3). Finally, Cluster D (11 genes) and cluster F (9 genes)

combined genes that were induced and not induced at veraison and ripening, such as for example

stilbene synthases (VIT_210s0042g00840-880-920) (TABLE 4).

Page 68: Comparative genetic and metabolic characterization between

66

GO Analysis of DEGs

We performed a Gene Ontology (GO) enrichment analysis for enlight functional categories

represented in the DEGs. In order to be able to use the available PHANTER Classification

System tool (Mi et al., 2017), we had to convert the gene ID references of 109 DEGs to the V1

version of the grapevine reference genome (http://genomes.cribi.unipd.it/DATA/V1/), which

caused 7 genes to be removed for the GO analysis for they were absent from this older

annotation of the grape genome. Based on these 102 DEGs, 46 DEGs were found to be enriched

for the term “molecular function” (FIGURE 5A): the DEGs were enriched in the categories

“binding” (GO:0005488, 7 genes), “signal transducer activity” (GO:0004871, 2 genes),

“catalytic activity” (GO:0003824, 30 genes) and “transporter activity” (GO:0005215, 7 genes).

A total of 59 genes were enriched into the term “biological process” (FIGURE 5B). The DEGs

were enriched in the categories “response to stimulus” (GO:0050896, 7 genes), “developmental

process” (GO:0032502, 1 gene), “cellular process” (GO:0009987, 17 genes), “transporter

activity” (GO:0005215, 7 genes), “metabolic process” (GO:0008152, 26 genes), “biological

regulation” (GO:0065007, 4 genes), “cellular component organization or biogenesis”

(GO:0071840, gene) and “localization” (GO:0051179, 3 genes). Finally, 29 genes were enriched

into the term “cellular component” (FIGURE 5C): the DEGs were enriched in the categories

“cell junction” (GO:0030054, 1 gene), “membrane” (GO:0016020, gene), “macromolecular

complex” (GO:0032991, gene, “cell part” (GO:0044464, 15 genes) and “organelle”

(GO:0043226, 11 genes).

Page 69: Comparative genetic and metabolic characterization between

67

Confirmation of the RNA-Seq Results

To validate the RNA-seq results, 12 of DEGs were selected and qRT-PCR analysis was

performed (FIGURE 6). From clusters of not induced genes (clusters A, B and C) 9 genes were

selected: F3’5’-H (Flavonoid 3’5’-hydroxylase) whose expression signal corresponds to 9

copies of this gene, CYTB5 (Cytochrome b5), UFGT (UDP-glucose:flavonoid 3-O-

glucosyltransferase), CCOAOMT (caffeoyl-CoA O-methyltransferase), ANP-like (anthocyanin

permease like), GST13 (glutathione S-transferase 13), F3H (Flavanone 3-hydroxylase), CHS

(Chalcone synthase). Seven of these genes (F3’5’-H, CYTB5, UFGT, CCOAOMT, GST13, F3H

and CHS) showed significant differences between CG and RG at both veraison and ripening

stages, fully supporting the RNA-seq results; ANP-like gene only showed differences during

ripening stage.

To gain more insight in gene expression 9 copies of F3’5’-H we designed specific primers for

6 of them: 2810, 2830, 2860, 2880, 2920 and 3010. In general, we observed significant

differences in each gene copies evaluated, in which there was not induction in CG compared to

RG, in veraison and ripening stages (FIGURE S4); also, this observation was correlated with

specific reads extracted from RNA-seq results for each gene copy of F3’5-H (TABLE S5).

From cluster of induced genes (cluster E) 3 genes were selected: MYB24, ABC transporter,

COMT-like (Catechol-O-methyltransferase) (FIGURE 6). It was observed that MYB24 was

significantly induced in veraison and ripening stages in CG and ABC transporter and COMT-

like genes were induced in CG only during veraison. Finally, the STS gene (Stilbene synthase)

was selected from combined clusters (cluster D and F). It exhibits a higher expression at ripening

and a lower one at veraison in CG compared to RG, but not with significant differences.

Page 70: Comparative genetic and metabolic characterization between

68

Eleven copies of F3’5’-H genes that were closely located in the same physical region of

chromosome 6 were repressed in Chimenti Globe

We observed that within the list of genes belonging to the repression cluster there were 11 copies

of the gene F3’5’-H repressed in CG: VIT_206s0009g02805, VIT_206s0009g02810,

VIT_206s0009g02840, VIT_206s0009g02860, VIT_206s0009g02873, VIT_206s0009g02880,

VIT_206s0009g02920, VIT_206s0009g02970 and VIT_206s0009g03010. This prompted us to

search for their physical location on the grape genome assembly, and this led to the conclusion

that all these genes were located closely in the same region of chromosome 6 (FIGURE 7). This

corresponded to a segment of the chromosome that span from approximately 15,660 to 16,060

MB.

Page 71: Comparative genetic and metabolic characterization between

69

DISCUSSION

Metabolic changes during veraison and ripening stages in CG and RG

Berry ripening is an active process that involves regulation of global gene expression (Fasoli et

al., 2012) and changes in the accumulation of primary and secondary metabolites (P. K. Boss et

al., 1996; Conde et al., 2007). Berry organoleptic quality depends mainly on sugars, organic

acids and secondary metabolites such as tannins, volatile compounds, flavonols, aroma

precursors and anthocyanins, that accumulate during fruit development. Anthocyanins are the

responsible for the coloration process of berry skin. Many studies have focused on this process

and the vast majority centered on varieties with same skin color (Degu et al., 2014; Fraige,

González-Fernández, Carrilho, & Jorrín-Novo, 2015; Guo et al., 2016; Wu et al., 2014). Fewer

have focused on comparatives studies using varieties with red color skin and white color skin

(no presence of anthocyanins) (Ghan et al., 2015; Massonnet et al., 2017; Negri, Prinsi, Failla,

Scienza, & Espen, 2015; Niu, Wu, Yang, & Li, 2013) and even less includes pink-colored

skinned berries (Cardoso, Lau, Dias, Fevereiro, & Maniatis, 2012; Shimazaki, Fujita,

Kobayashi, & Suzuki, 2011; Walker, Lee, & Robinson, 2006).

To understand in depth the molecular changes related to coloration of grape berry skin, we used

RG and CG varieties in this study, two varieties of the near isogenic genetic background, which

differ in berry color skin from purple to bright light red. First, we performed a comparative

metabolic analysis of skin at two developmental stages, veraison and ripening, using LC-

MS/MS, HPLC, colorimetric and enzymatic assays for various types of metabolites. We chose

these developmental stages for their relevance and importance during berry development and

Page 72: Comparative genetic and metabolic characterization between

70

determination of crucial characteristics in the grape fruit: veraison marks the beginning of

ripening period and includes the start of coloration of berry skin, and the ripe stage represents

the culmination of berry developmental process (He et al., 2010; Kuhn et al., 2014). We

observed with PCA analysis that developmental stages were the primary discriminant

parameters in such comparative study, despite the use of different varieties, when PC1 was

analyzed (FIGURE 2A). This observation is concordant with previous published studies (Guo

et al., 2016; Massonnet et al., 2017; Wong, Lopez Gutierrez, Dimopoulos, Gambetta, &

Castellarin, 2016; Wu et al., 2014) in which, despite the use of different varieties, the importance

of developmental stages appeared to overrule variety origin. Another interesting result was that

T6P, the main influential metabolite in PC1 that explained 43,75% of data variability and

separated veraison from ripening in both CG and RG. T6P is a sugar phosphate considered as a

metabolic signaling molecule in plants; which regulates developmental growth (John Edward

Lunn, Delorge, Figueroa, Van Dijck, & Stitt, 2014) and was demonstrated to be involved in

sugar signaling in Arabidopsis thaliana (John E. Lunn et al., 2006). In Vitis vinifera there is no

clear evidence of control of T6P during berry development, although several transcriptomic

studies have shown that orthologue genes encoding T6P synthase, T6P synthase/phosphatase

and SnRK1 (Sucrose-non-fermentative Related Kinase 1, another sugar signaling pathway

component) are tightly regulated during berry development (L. G. Deluc et al., 2007; Gambetta,

Matthews, Shaghasi, McElrone, & Castellarin, 2010). Our data showed an increase in T6P

concentration in both CG and RG berry skin, from veraison to ripe stage (FIGURE S1). This

constitute a new finding in grape berry development, because previous studies analyzing T6P

in grapevine berries used total berries (Dai et al., 2013), and did not separate skin and flesh,

which probably lead to dilution of T6P signal from the skin. Due to the fact that sugar stimulates

Page 73: Comparative genetic and metabolic characterization between

71

anthocyanin biosynthesis in grape berry and grape cell culture (Dai et al., 2013; Larronde et al.,

1998), it would be interesting to study the expression of genes related to sugar accumulation

and metabolism in our RNA-seq data, to check whether there are differences between our

contrasted color skin varieties CG and RG.

Another important metabolite distinguishing veraison and ripe stages in PC1 (FIGURE 2A and

FIGURE S1) was shikimate, an intermediary in shimikate pathway that mediates the carbon

flow from carbohydrate metabolism to phenylpropanoid and aromatic compound biosynthesis

(Maeda & Dudareva, 2012; Zhang et al., 2012). This strongly suggest the importance of this

metabolite in species such as vitis vinifera, in which berry ripening involved phenylalanine

consumption (one of the main destination of chorismate, the final product of shikimate pathway)

for biosynthesis of flavonoid compounds such as anthocyanins.

According to PC2 which discriminated varieties, CG from RG, we observed many metabolites

directly related with early products of sucrose biosynthesis such as glyceraldehyde 3-phosphate,

fructose 1,6-biphosphate, fructose 6-phosphate and glucose 6-phosphate; and other metabolites

less related to this process (FIGURE S2). Also, another important metabolite that discriminated

varieties were anthocyanins (FIGURE 3A and 3C) where we observed that only traces of

malvidin, petunidin and delphinidin anthocyanins were detected in CG (FIGURE 3B and 3C),

indicating a disruption in the tri-hydroxylated branch of the anthocyanin biosynthesis pathway,

linked to flavonoid 3’5’-hydroxylase enzyme (F3’5’H) activity, an observation latter confirmed

by RNA-seq results (discussed below). Accordingly, an increase in F3’5’H activity have been

previously reported to explain higher concentration of tri-hydroxylated anthocyanins in

Page 74: Comparative genetic and metabolic characterization between

72

Cabernet Sauvignon and Shiraz, pointing F3’5’H plays as a key molecular player driving the

phenylpropanoid metabolic flux towards bluish anthocyanins production (Degu et al., 2014).

If we take both results together, we can link them because there is a relationship between sugars

and anthocyanin accumulation described in previous studies. For example, studies in Shiraz

grape berries showed that soluble solids (measured as Brixº and represent glucose and fructose

in grape berry) began to increase 8 weeks post flowering and continue to increase, and

anthocyanins are then synthetized 10 weeks post flowering (Paul K. Boss, Davies, & Robinson,

1996). Also, in Vitis vinifera cell cultures was found that sucrose modulated anthocyanins

content, and this was observed also with fructose and glucose (Larronde et al., 1998). In our

study the anthocyanin biosynthesis pathway was affected in last steps in CG, and maybe this

observation can be linked to modifications in biosynthesis of sucrose in CG or otherwise,

changes in anthocyanins in CG modified intermediates of sucrose. That is an interesting point

of discussion.

Transcriptomic changes during veraison and ripening in CG an RG varieties

RNA-seq analysis had been often used for searching and characterizing new gene expression

patterns in pathway related to grape berry development (Degu et al., 2014; Fortes et al., 2011;

Ghan et al., 2015, 2017; Guo et al., 2016; Li et al., 2014; Lijavetzky et al., 2012; Massonnet et

al., 2017; Sweetman et al., 2012; Wong et al., 2016; Wu et al., 2014; Zenoni et al., 2010).

However, few studies focused on the global gene expression in contrasted color berry skin from

near isogenic backgrounds, and even less considered pink varieties. In our study, besides

metabolic studies, skin transcriptome was analyzed at two developmental stages using Illumina

Hiseq2000 in two near isogenic varieties, RG and CG. According to DEGs analysis, we reported

Page 75: Comparative genetic and metabolic characterization between

73

that several genes related to phenylpropanoid pathway, which grouped in three main clusters of

repression, induction or a combination of both, were differentially expressed in CG compared

to RG (TABLE 2, 3 and 4, respectively). Among them, we found that 11 copies of F3’5’-H,

encoding a key enzyme in anthocyanin biosynthesis pathway were repressed in CG. This

enzyme corresponds branching point of the flavonoid pathway, where dihydrokaempferol is the

substrate shared in common (Kuhn et al., 2014). Dihydrokaempferol is converted via F3’5’-H

to dihydromyricetin, producing the glucoside, acetyl, coumaroyl and caffeoyl forms of

delphinidin, petunidin and malvidin, that correspond to the delphinidin-type (tri-hydroxylated)

anthocyanins of bluish color. Dihydrokaempferol is also used by F3’-H to produce

dihydroquercetin and gives rise to the biosynthesis of glucoside, acetyl and coumaroyl forms of

cyanidin and peonidin, that correspond to cyanidin-type (di-hydroxylated) anthocyanins or

reddish color (Degu et al., 2014). Indeed, our results unambiguously showed that CG berry skins

did not contain delphindin-type anthocyanins (FIGURE 3B and 3D) and therefore, is consistent

with the repression of 11 gene copies of F3’5’-H in CG. Additionally, another cunning

observation was the adjacent location of each of these copies in a single cluster within

chromosome 6 (FIGURE 7). It is known that there are many copies of F3’5’-H in Vitis vinifera,

and it was shown that F3’5’-H in grapevine are highly redundant and predominately expressed

in berry skins (Castellarin et al., 2006; Falginella et al., 2010). However, to the best of our

knowledge, there was no evidence of the repression of such high number of gene copies at the

same time in a single variety. This raises the question of how the not induction of the F3’5’-H

genes occur in CG. At least two mechanisms can be foreseen, among others. Firstly, since all

these genes are physically close together in a single region of chromosome 6, an epigenetic can

be postulated. Although left uninvestigated until recently, the field of epigenetic regulation in

Page 76: Comparative genetic and metabolic characterization between

74

grapevine has gain considerable attention in the last years; and epigenetic regulation were shown

to be potentially part of the so-called “terroir” effect in Shiraz grapes from Australia (Xie et al.,

2017). Secondly, a still to be discovered master regulator (i.e a transcription factor) might be

defective in the CG. For example, an identical transcription factor biding motif might be present

in the promoter regions of all 11 gene copies, which would be in agreement with the hypothesis

of a master regulator for these F3’5’-H genes. Further work is necessary to decide between these

two hypotheses.

In parallel, several genes putatively involved in anthocyanin transport to the vacuoles were

differentially expressed in CG compared to RG at both veraison and ripening stages, such as

GST13 and ABC transporter (FIGURE 6). GST13 (VIT_207s0104g01800) expression was

down-regulated at both developmental stages in CG. Previous studies in model plants showed

that members of the multigenic GST family have been described to be necessary for anthocyanin

accumulation such as AN9 in petunia (Alfenito et al., 1998), TT19 in Arabidopsis (Kitamura,

Shikazono, & Tanaka, 2004) and BZ2 in maize (Goodman, Casati, & Walbot, 2004). A recent

study in vitis vinifera performed molecular dynamics and docking analysis together with gene

analysis expression for GST1 and GST4 in different organs during berry development and also.

The ability of GST1 and GST4 to rescue the phenotype of the Arabidopsis tt19-1 mutant was

also tested. The results showed that at least GST4 behaved like Arabidopsis TT19 in the

transport of anthocyanins and proanthocyanidins, revealing a novel function for his gene in

flavonoid transport (Pérez-Díaz, Madrid-Espinoza, Salinas-Cornejo, González-Villanueva, &

Ruiz-Lara, 2016). In our RNA-seq data, we noticed that an ABC (ATP-binding cassette)

transporter (VIT_214s0060g00720) was induced only at veraison in CG. ABC transporters have

Page 77: Comparative genetic and metabolic characterization between

75

also been involved in vacuolar sequestration of flavonoids in maize (Goodman et al., 2004) and

in the transport of malvidin 3-O-glucoside into vacuoles, in a strictly GSH-dependent fashion

in grapevine (Zarrouk et al., 2013). In addition, ANP-like genes differentially expressed in our

RNA-seq study (VIT_216s0050g00900 and VIT_216s0050g00910) encodes anthocyanin

vacuolar transporters that were repressed in CG, and have been suggested to participate in

anthocyanin vacuolar sequestration in grapes (Mathews et al., 2003; Wu et al., 2014). These

findings reported in our study related to GST13 and ANP-like, repressed in CG and on the other

hand ABC transporter induced in CG, might suggest a dynamic process in transport and

sequestration of anthocyanin into vacuoles occurring in this variety context. Anyways, due to

these genes have not been characterized in grapevines, they become in new possible candidates

to study for better understanding anthocyanin transport.

In addition, several genes related to anthocyanin O-methylation, such as CCoAOMT and

COMT-like, were also found to be differentially expressed between RG and CG in our study.

CcoAOMT is a multifunctional O-methyltransferase that might contribute to anthocyanin

methylation activity in grapevine, but this was only observed under drought stress conditions

(Giordano et al., 2016). Our data showed that CcoAOMT was repressed in veraison and ripening

stage in CG, suggesting that maybe abiotic stress regulation in CG was modified for some

unknown reason. Stress in grapevine had been generally associated with biosynthesis of tri-

hydroxylated anthocyanins (delphinidin-type), which would make sense according to the

phenotype observed in CG. Finally, COMT-like is an OMT using catechol as substrate and

characterized in S. lycopersicum (Mageroy et al., 2012) but not in grapevine so far. It was found

to be induced in CG during veraison and at ripening stage. Further studies will be necessary to

understand the function of this enzyme in CG.

Page 78: Comparative genetic and metabolic characterization between

76

Another key gene found to be down-regulated in CG was CHS (FIGURE 6), that encodes the

first enzyme of the flavonoid pathway (Petrussa et al., 2013; Wang et al., 2016) and this would

be coherent with the observed metabolic profile of CG, which contains less total anthocyanins

than RG. Also, STS gene, encoding another early enzyme in flavonoid metabolism and involved

in resveratrol biosynthesis (Dubrovina & Kiselev, 2017; Kiselev, Aleynova, Grigorchuk, &

Dubrovina, 2016) was induced in CG at veraison and ripening stages (FIGURE 6). This

possibly reveals a possible compensation point in flavonoid metabolism: CHS utilizes the same

substrates as STS but, is responsible for flavonoid-type compound formation. Thus, STS would

act as an escape valve to absorb the excess of substrate flux generated in the upper part of the

flavonoid metabolism because of the decrease in CHS activity. Taken together, this suggests a

multilevel alteration of the flavonoid pathway, and not only of its tri-hydroxylated branch in CG

variety.

Most studies investigating the mechanisms of changes in berry skin color explain them through

the regulatory effect of MybA transcription factors, whose genes are closely clustered in a single

locus on chromosome 2, referred to as the “berry color locus” (Kobayashi et al., 2004; Walker

et al., 2006). One of these is the transcription factor MybA1, a major determinant of grape berry

anthocyanin biosynthesis pathway (Kobayashi et al., 2004; Lijavetzky et al., 2006). MybA1

transcriptionally regulates UFGT gene, which glycosylates anthocyanidins, stabilizing them and

therefore responsible for the anthocyanins accumulation in grape berry skins (Ford, Boss, &

Hoj, 1998). White skinned grape varieties that do not produce anthocyanins, contain a mutated

allele of MYBA1 in which the presence of the retrotransposon Gret1 in promoter region

interrupts the normal transcription process, resulting in a white berry allele (Kobayashi et al.,

2004). Pigmented cultivars possess at least one allele of the MybA1 locus that does not contain

Page 79: Comparative genetic and metabolic characterization between

77

Gret1 and is called red berry allele (This et al., 2007). In the case of pink skin berry varieties of

oriental origin, a different situation was reported: a 33bp insertion was found in the second

intron of MybA1, generating low expression levels of MybA1 in grape from these cultivars

(Shimazaki et al., 2011). Another different example is a bud sports of Cabernet Sauvignon called

“Malian” with red-grey berry skin, associated to the deletion of a large region of chromosome

2 that includes both MybA1 and MybA2 genes (MybA2 is located adjacent to MybA1 in the color

berry locus) (Walker et al., 2006). We reported that UFGT gene (VIT_216s0039g02230) is

repressed in CG at veraison and ripening stages (FIGURE 6), which could suggest an alteration

in MybA1 expression, according to literature describe above. MybA1, however, was not deemed

significantly affected in DEGs analysis (FIGURE 4). Nevertheless, we decided to check for the

presence of white and red allele in CG and RG, using the same set of primer used in (Shimazaki

et al., 2011). We observed the presence of white and red allele of MybA1 in all CG and RG

samples tested (FIGURE S5A and S5B, respectively), demonstrating that both varieties are

heterozygous for this gene and that our results cannot be explained by structural changes in

MybA1 DNA sequence.

Another MYB gene, MYB24 (VIT_214s0066g01090) was found to be induced in CG (TABLE

3, FIGURE 6). Recently, it was reported that MYB24 was associated with terpene biosynthesis

in berries in response to drought (Savoi et al., 2016); also was highly co-expressed with flower

and fruit specific terpene synthase (TPS) (Wong et al., 2016); and induced by solar UV radiation

in Tempranillo berry skins, in conjunction with terpenoid metabolism induction (Carbonell-

Bejerano et al., 2014). Therefore, it would seem that this transcription factor would be in charge

of leading the responses towards the activation of biosynthesis of terpenes. So, how could this

be related to berry color? With these antecedents we hypothesized that this transcription factor

Page 80: Comparative genetic and metabolic characterization between

78

could be linked by activating the terpene biosynthesis pathway in CG, which would therefore

affect the flow towards anthocyanin biosynthesis, or vice versa, a change or block in the

anthocyanin biosynthesis pathway, takes the flow to the terpenes, which would be mediated via

MYB24. How to know what is happening in CG? Future studies using this transcription factor

to evaluate berry color in CG would be necessary.

Another gene of particular interest reported in our study was Cytb5, which was repressed in CG

at veraison and ripening stages (FIGURE 6). Cytb5 has previously been involved in F3’5’-H

activity modulation in Petunia flowers, catalyzing the transfer of electrons to their prosthetic

heme group (de Vetten et al., 1999). In 2006, (Bogs, Ebadi, McDavid, & Robinson, 2006)

demonstrated that a putative Cytb5 from grapevine cv Shiraz) modulated both F3’5’-H and F3’-

H activites even though the exact mechanism remains to be deciphered. This is of interest

because this suggests a possible link between the drastic down-regulation of 11 gene copies of

F3’5’H and the repression of the Cytb5 gene observed in CG. Additional work will be required

to test that hypothesis.

Page 81: Comparative genetic and metabolic characterization between

79

CONCLUSION

This comparative study revealed important differences at the metabolic and transcriptomic level

between two contrasted color berry skin, but nevertheless near isogenic, table grapes. Using

RNA-seq we showed that flavonoid metabolism was affected at different levels in CG compared

to RG at both veraison and ripening stage. While the nature of the mutation responsible for color

changes in berry skin of CG from its progenitor RG remains unknown, the comparative RNA

profiling of the berry development between CG and RG showed the deregulation of genes

directly related to anthocyanin biosynthesis: CHS, STS, CYTB5, UFGT, MYB24, F3’5’-H, F3’-

H, CCoAOMT, COMT, GST13, ABC transporter and ANP-like. A prominent result was the

drastic down-regulation of 11 gene copies of F3’5’H in CG positioned in a single cluster in

chromosome 6, which was totally consistent with the anthocyanin content observed in CG, with

the absence of delphinidin-type (tri-hydroxylated) anthocyanins. This left in evidence an

arrangement occurring in this metabolic process in CG compared to RG and open new types of

regulation undescribed in anthocyanin biosynthesis in grapevines.

This study provides new insights in the genetic and metabolic background that may be

responsible of color change in CG compared to RG, that constitute a relevant biological material

to study in depth the mechanisms underlying berry skin color intensity.

Page 82: Comparative genetic and metabolic characterization between

80

AUTHOR CONTRIBUTIONS

The manuscript was written by Claudia Santibáñez. Tomás Moyano and Carlos Meyer made the

bioinformatics data analysis and datamining. John Lunn and Regina Feil carried out metabolic

analysis by LC-MS/MS. Ghislaine Hilbert and Christel Renaud helped in hexose, anthocyanin

and organic acids experiments. Litsy Martínez took care of biological material. Claudia

Santibáñez performed vegetal material sampling, RNA extraction, qRT-PCR and data analysis

of all study. Eric Gomès, Patricio and Serge Delrot contribute in supervision, design of

experiments and revision of the manuscript.

Page 83: Comparative genetic and metabolic characterization between

81

ACKNOWLEDGMENTS

We thank to CONICYT scholarship doctorate 2012 Nº 21120432, Operational Expenses

Scolarship of CONICYT Nº 21120432, Millennium Nucleus of Plant Systems and Synthetic

Biology NC130030 and CHIMENTI S.A.

Page 84: Comparative genetic and metabolic characterization between

82

FIGURE 1: Field photographs of cultivars RG and CG in the different stages during 2013 season sampling. (A) RG cluster during

veraison stage. (B) CG cluster during veraison stage. (C) RG and CG clusters during ripening stage. (D and E) RG and CG plants used

in this study. RG correspond to left side of photograph and CG to right side, during veraison and ripening stage, respectively.

Page 85: Comparative genetic and metabolic characterization between

83

FIGURE 2: Principal component analysis of metabolites present in berry skin of RG and CG. (A) Discriminations of cultivars RG

and CG, veraison and ripening stages for the 2013 and 2014 seasons. (B) Loading plots of metabolites analyzed for the first two

components (PC1 and PC2).

Page 86: Comparative genetic and metabolic characterization between

84

0%

20%

40%

60%

80%

100%

VeraisonRipening

VeraisonRipening

% A

nth

ocyan

ins c

on

ten

tm

g g

-1 s

kin

DW

Axis Title

Malvidin

Peonidin

Petunidin

Cyanidin

Delphinidin

Chimenti Red Globe

0%

20%

40%

60%

80%

100%

VeraisonRipening

VeraisonRipening

% A

nth

ocyan

ins c

on

ten

tm

g g

-1 s

kin

DW

Axis Title

Chimenti Globe Red Globe

A B

C D

Verai

son

Rip

enin

g

0

5

10

15

Stages

To

tal a

nth

oc

ya

nin

s

mg

g-1

sk

in D

W

Chimenti Globe

Red Globe

****

**

Stages

To

tal a

nth

oc

ya

nin

s

mg

g-1

sk

in D

W

Verai

son

Rip

enin

g

0

5

10

15Chimenti Globe

Red Globe

***

**

Page 87: Comparative genetic and metabolic characterization between

85

FIGURE 3: Content of anthocyanins in Chimenti globe and Red Globe berry skin, at

veraison and ripening stages (mg g-1 DW). (A) and (C) Total anthocyanins content in 2013

and 2014 season, respectively. The results are the mean of five biological replicates, error

bars represent standard error of mean. Asterisks indicate the result of unpaired T-test. **, P

< 0.0081; ***, P = 0.0001; ****, P < 0.0001. (B) and (D) Percentages of main anthocyanins

in 2013 and 2014 season, respectively. Malvidin corresponds to malvidin 3-glucoside

(purple); Peonidin to peonidin 3-glucoside and peonidin 3-(p-coumaryl)-glucoside

(magenta); Petunidin to petunidin 3-glucoside (indigo); Cyanidin to cyanidin 3-glucoside and

cyanidin 3-(p-coumaryl)-glucoside (red); and Delphinidin to delphinidin 3-glucoside and

delphinidin 3-(acetyl)-glucoside (blue). DW, Dry Weight.

Page 88: Comparative genetic and metabolic characterization between

86

TABLE 1: Average content of different types of anthocyanins in Chimenti Globe and Red Globe berry skin (mg g-1 DW).

CHIMENTI GLOBE RED GLOBE

Veraison Ripening Veraison Ripening

2013 2014 2013 2014 2013 2014 2013 2014

Delphinidin 0.00 ± 0.00a 0.00 ± 0.00a 0.00 ± 0.00a 0.01 ± 0.00a 0.07 ± 0.02e 0.09 ± 0.09b 0.52 ± 0.08e 0.53 ± 0.03c

Petunidin 0.00 ± 0.00a 0.00 ± 0.00a 0.01 ± 0.01a 0.01 ± 0.00a 0.06 ± 0.02e 0.09 ± 0.09a 0.60 ± 0.11e 0.48 ± 0.23c

Malvidin 0.00 ± 0.00a 0.00 ± 0.00a 0.01 ± 0.01a 0.03 ± 0.01a 0.21 ± 0.06e 0.32 ± 0.37a 2.07 ± 0.57e 1.38 ± 0.64c

Cianidin 0.11 ± 0.03a 0.04 ± 0.02a 0.69 ± 0.49a 0.65 ± 0.14a 0.39 ± 0.19c 0.65 ± 0.26d 2.17 ± 0.24d 2.43 ± 0.65d

Peonidin 0.16 ± 0.04a 0.06 ± 0.02a 1.21 ± 0.40a 1.13 ± 0.32a 0.75 ± 0.33c 1.41 ± 0.86c 9.04 ± 1.22e 7.90 ± 2.08e

Total 0.27 ± 0.05a 0.11 ± 0.04a 1.93 ± 0.85a 1.84 ± 0.44a 1.49 ± 0.57c 2.57 ±1.57c 14.39 ± 1.21e 12.71 ± 3.48d

The letters represent the result of an unpaired T-test between Chimenti Globe and Red Globe in each developmental stage during 2013

and 2014 season: a represent no significant differences; b represent *, P = 0.0497; c represent **, P < 0.0093; d represent ***, P < 0.0009; e represent ****, P < 0.0001.

Page 89: Comparative genetic and metabolic characterization between

87

Page 90: Comparative genetic and metabolic characterization between

88

FIGURE 4: Hierarchical clustering heat map and cluster tree from 109 DEGs in Chimenti

Globe and Red Globe. Each row represents a gene and each column represents a developmental

stage, veraison (left panel) and ripening (right panel). The color key indicates logFC values ≥ 1,

ranging from bright yellow for most up-regulated to bright blue for the most down-regulated

genes, considering a FDR of 0.05. Genes that showed a similar developmental profile were

clustered together. Six clusters were generated.

Page 91: Comparative genetic and metabolic characterization between

89

TABLE 2: The DEGs ID of repression cluster A, B and C during veraison and ripening stages

Nº Cluster Genoscope ID Log2FCV Log2FCR Blast NCBI

1 A VIT_204s0079g00710 -4,66 -5.44 glutathione s-transferase

2 B VIT_206s0009g02830 -2.04 -5.20 flavonoid 3', 5'- hydroxylase

3 B VIT_206s0009g02840 -1.27 -4.88 flavonoid 3', 5'- hydroxylase

4 B VIT_207s0104g01650 -1.64 -4.30 potassium transporter 5

5 B VIT_206s0009g02810 -1.73 -4.42 flavonoid 3', 5'- hydroxylase

6 B VIT_206s0009g02920 -2.04 -4.23 flavonoid 3', 5'- hydroxylase

7 B VIT_206s0009g02970 -2.15 -4.38 flavonoid 3', 5'- hydroxylase

8 B VIT_206s0009g02805 -2.41 -3.72 flavonoid 3', 5'- hydroxylase

9 B VIT_206s0009g02880 -2.22 -3.50 flavonoid 3', 5'- hydroxylase

10 B VIT_207s0104g01800 -1.99 -3.53 glutathione s-transferase f13

11 B VIT_206s0009g02846 -1.54 -3.98 flavonoid 3', 5'- hydroxylase

12 B VIT_206s0009g02873 -1.70 -3.64 flavonoid 3', 5'- hydroxylase

13 B VIT_200s0256g00120 -1.22 -3.30 protein

14 B VIT_214s0036g00330 -1.02 -2.68 disease resistance protein at4g27190-like

15 B VIT_205s0077g01730 -1.03 -2.67 uncharacterized protein loc100243551

16 B VIT_214s0036g00100 -1.05 -2.39 disease resistance protein at4g27190-like

17 B VIT_207s0031g01040 -1.11 -2.34 l-ascorbate oxidase

18 B VIT_207s0197g00230 -1.15 -2.35 probable disease resistance protein at5g66900-like

19 B VIT_216s0050g00900 -1.25 -2.46 multidrug resistance

20 B VIT_206s0009g02860 -1.75 -2.36 flavonoid 3', 5'- hydroxylase

21 B VIT_219s0027g01070 -1.75 -2.48 disease resistance protein rga4-like

22 B VIT_203s0088g00260 -1.93 -2.45 serine carboxypeptidase-like 18-like

23 B VIT_216s0050g00910 -2.08 -2.67 multidrug resistance

Page 92: Comparative genetic and metabolic characterization between

90

24 B VIT_204s0023g03370 -1.99 -2.90 flavanone 3- hydroxylase

25 C VIT_216s0022g00670 -1.30 -1.04 vacuolar invertase

26 C VIT_200s0282g00040 -1.05 -1.01 protein spinster homolog 1-like

27 C VIT_203s0091g00410 -1.03 -1.07 uncharacterized protein

28 C VIT_214s0030g01160 -1.06 -1.16 disease resistance protein at4g27190-like

29 C VIT_215s0046g01560 -1.10 -1.15 bcl-2-associated athanogene 6

30 C VIT_207s0197g00140 -1.12 -1.28 probable disease resistance protein at5g66900-like

31 C VIT_208s0040g01140 -1.14 -1.27 serine carboxypeptidase-like 45-like

32 C VIT_205s0077g01720 -1.08 -1.29 uncharacterized protein loc100256873

33 C VIT_216s0013g01500 -1.08 -1.35 leucine-rich repeat receptor protein kinase exs-like

34 C VIT_204s0044g00170 -1.03 -1.34 dna mismatch repair protein mlh3

35 C VIT_210s0003g02000 -1.06 -1.42 protein

36 C VIT_208s0007g03560 -1.23 -1.63 solute carrier family 35 member f1-like

37 C VIT_200s0160g00100 -1.11 -1.61 tmv resistance protein n-like

38 C VIT_203s0017g00870 -1.13 -1.65 protein

39 C VIT_210s0003g01995 -1.07 -1.71 atp binding

40 C VIT_216s0100g00670 -1.10 -1.84 homeobox-leucine zipper protein anthocyaninless 2-like

41 C VIT_206s0061g01230 -1.11 -1.87 glucomannan 4-beta-mannosyltransferase 2

42 C VIT_214s0036g00310 -1.05 -2.17 disease resistance protein at4g27190-like

43 C VIT_217s0000g07375 -1.02 -2.06 enhanced disease susceptibility 1

44 C VIT_208s0007g03570 -1.13 -2.04 solute carrier family 35 member f1-like

45 C VIT_205s0136g00260 -1.54 -2.19 chalcone synthase

46 C VIT_200s0245g00030 -1.49 -1.86 uncharacterized protein loc100852932

47 C VIT_206s0009g03010 -1.94 -1.78 flavonoid 3', 5'- hydroxylase

48 C VIT_214s0066g00900 -1.77 -1.41 cysteine-rich receptor-like protein kinase 42

Page 93: Comparative genetic and metabolic characterization between

91

49 C VIT_205s0049g00480 -2.00 -1.38 at1g04360

50 C VIT_216s0039g02220 -2.18 -1.45 exportin-1-like isoform 1

51 C VIT_201s0010g03510 -2.20 -1.41 caffeoyl- o-methyltransferase

52 C VIT_201s0010g03490 -2.36 -1.65 caffeoyl- o-methyltransferase

53 C VIT_216s0039g02230 -3.15 -1.35 udp-glucose:flavonoid 3-o-glucosyltransferase

54 C VIT_218s0001g09400 -3.82 -1.99 cytochrome b5

55 C VIT_204s0079g00690 -3.96 -2.53 glutathione s-transferase

Page 94: Comparative genetic and metabolic characterization between

92

TABLE 3: The DEGs ID of induction cluster E during veraison and ripening stages

Nº Cluster Genoscope ID Log2FCV Log2FCR Blast NCBI

67 E VIT_202s0087g00380 2.49 1.32 hypothetical protein

68 E VIT_209s0002g03460 1.73 1.26 clavaminate synthase-like protein

69 E VIT_204s0044g01230 1.56 1.38 hypothetical protein

70 E VIT_213s0067g02690 1.07 1.39 dimethylguanosine trna methyltransferase

71 E VIT_206s0004g03910 1.11 1.38 hypothetical protein

72 E VIT_202s0234g00110 1.13 1.41 oleosin low molecular weight isoform

73 E VIT_209s0002g03450 1.13 1.33 clavaminate synthase-like protein

74 E VIT_201s0137g00540 1.11 1.29 cytochrome p450

75 E VIT_203s0038g01380 1.01 1.23 calcium sensor

76 E VIT_208s0007g01470 1.06 1.21 copper transporter

77 E VIT_216s0098g00890 1.08 1.18 harpin-induced family protein

78 E VIT_219s0090g01350 1.03 1.16 hypothetical protein

79 E VIT_212s0028g03620 1.12 1.15 uncharacterized protein

80 E VIT_219s0014g04790 1.16 1.15 solute carrier family 22 member 15-like

81 E VIT_202s0025g02560 1.17 1.15 o-succinylhomoserine sulfhydrylase

82 E VIT_210s0003g03270 1.29 1.43 potassium channel

83 E VIT_205s0077g00635 1.32 1.40 hypothetical protein

84 E VIT_200s0684g00030 1.25 1.52 senescence-associated protein

85 E VIT_214s0060g00720 1.29 1.56 abc transporter family protein

86 E VIT_208s0058g00930 1.21 1.70 alanine--glyoxylate aminotransferase 2 homolog mitochondrial

87 E VIT_207s0005g00870 1.34 1.74 hypothetical protein

88 E VIT_215s0048g02490 1.45 1.76 catechol o-methyltransferase

89 E VIT_214s0066g01090 1.53 1.69 myb transcription factor 24

Page 95: Comparative genetic and metabolic characterization between

93

90 E VIT_214s0108g00450 1.05 1.93 hypothetical protein

91 E VIT_219s0014g05080 1.10 1.90 oxygen-evolving enhancer protein 3-1

92 E VIT_205s0102g00790 1.20 1.88 hypothetical protein

93 E VIT_201s0137g00680 1.21 1.99 hypothetical protein

94 E VIT_207s0129g00510 1.40 2.12 uncharacterized protein

95 E VIT_205s0020g04110 1.12 2.25 early light-inducible protein

96 E VIT_216s0050g01820 1.35 2.48 hypothetical protein

97 E VIT_201s0011g05440 1.92 2.24 hypothetical protein

98 E VIT_202s0025g01710 1.89 2.51 glutaredoxin family protein

99 E VIT_215s0048g02480 1.74 2.88 catechol o-methyltransferase

100 E VIT_218s0001g04780 1.64 3.76 sesquiterpene synthase 10 (TPS10)

Page 96: Comparative genetic and metabolic characterization between

94

TABLE 4: The DEGs ID of combined induction/repression clusters D and F during veraison and ripening stages

Nº Cluster Genoscope ID Log2FCV Log2FCR Blast NCBI

56 D VIT_215s0048g02420 1.03 -3.27 hypothetical protein

57 D VIT_210s0042g00880 1.31 -3.05 stilbene synthase

58 D VIT_216s0022g00650 1.22 -2.40 hypothetical protein

59 D VIT_210s0042g00840 1.28 -2.16 stilbene synthase

60 D VIT_209s0002g03560 1.05 -2.07 atp-binding cassette

61 D VIT_200s0483g00020 1.12 -2.03 suppressor-like protein

62 D VIT_218s0041g02020 1.11 -1.82 12-oxophytodienoate reductase

63 D VIT_202s0025g04860 1.31 -1.59 hypothetical protein

64 D VIT_210s0042g00920 1.30 -1.59 stilbene synthase

65 D VIT_201s0011g05890 1.42 -1.52 hypothetical protein

66 D VIT_201s0182g00130 1.54 -1.78 phosphate transporter pho1 homolog 3-like

101 F VIT_212s0028g01130 -1.00 1.86 hypothetical protein

102 F VIT_217s0000g09800 -1.02 1.62 hypothetical protein

103 F VIT_200s0301g00070 -1.32 1.59 hypothetical protein

104 F VIT_212s0142g00160 -1.23 1.51 hypothetical protein

105 F VIT_215s0046g00080 -1.17 1.39 hypothetical protein

106 F VIT_200s0400g00030 -1.41 1.25 leucine-rich repeat receptor protein kinase exs-like

107 F VIT_205s0020g02210 -1.46 1.12 histidine phosphotransfer protein

108 F VIT_203s0038g01310 -1.28 1.10 saur family protein

109 F VIT_204s0023g03230 -1.03 1.03 saur family protein

Page 97: Comparative genetic and metabolic characterization between

95

FIGURE 5: Gene ontology classification of DEGs. GO annotations for differentially

expressed genes during veraison and ripening stages in CG and RG varieties. (A) Molecular

Function, (B) Biological Process and (C) Cellular Component.

Page 98: Comparative genetic and metabolic characterization between

96

Page 99: Comparative genetic and metabolic characterization between

97

FIGURE 6: Real-time quantitative PCR (qRT-PCR) validation of 12 selected DEGs

identified by RNA-seq in berry skin of CG and RG. Samples were collected at veraison

and ripening stages in 2013, and genes were selected from repression, induction and

combined repression/induction clusters. The results are the mean of six biological replicates,

error bars represent standard error of mean. Asterisks indicate the result of unpaired T-test.

*, P < 0.0465; **, P < 0.0053; ***, P < 0.0009; ****, P < 0.0001. FW, Fresh Weight.

Page 100: Comparative genetic and metabolic characterization between

98

FIGURE 7: Localization of 11 gene copies of flavonoid 3’5’-hydroxylases on

chromosome 6 in Vitis vinifera. Diagram representation showing the position on Ch6 of

F3’5’-H genes. Ch6 has a size of 21.5MB and F3’5’-H gene copies from results are located

in an area 0.5MB

Page 101: Comparative genetic and metabolic characterization between

99

SUPPLEMENTARY MATERIAL

Page 102: Comparative genetic and metabolic characterization between

100

FIGURE S1: Metabolites whose percentage of contribution are higher in PC1 from

PCA (nmol g-1 FW), at veraison and ripening stages. (A, C, E, G, I) Metabolites

corresponding to the season 2013, and (B, D, F, H, J) corresponding to the season 2014. The

results are the mean of six biological replicates, error bars represent standard deviations.

Asterisks indicate the result of two-way ANOVA, followed by Tukey’s test. ****, P <

0.0001; ***, P < 0.0005; **, P < 0.0052; *, P < 0,0282. FW, Fresh Weight.

Page 103: Comparative genetic and metabolic characterization between

101

Page 104: Comparative genetic and metabolic characterization between

102

Page 105: Comparative genetic and metabolic characterization between

103

FIGURE S2: Metabolites whose percentage of contribution are higher in PC2 from

PCA (nmol g-1 FW). (A, C, E, G, I, K, M, O, Q) Metabolites corresponding to the season

2013; (B, D, F, H, J, L, N, P, R) metabolites corresponding to the season 2014. The results

are the mean of six biological replicates, error bars represent standard error of mean.

Asterisks indicate the result of unpaired T-test. *, P < 0.0443; **, P < 0.0085; ***, P <

0.0003.; ****. FW, Fresh Weight.

Page 106: Comparative genetic and metabolic characterization between

104

FIGURE S3: Analysis of MybA1 gene structure in CG and RG. (A) PCR analysis of

white allele in CG, RG varieties and control Merlot (Me), Cabernet Sauvignon (CS),

Chardonnay (Ch) and Sauvignon Blanc (SB). (B) PCR analysis of red allele in CG and RG

varieties and control described above. (C) Integrity control of actin gene. Numbers on the

left indicate the positions of the molecular size marker from KBplus and on the right the

exact size of white and red allele.

Page 107: Comparative genetic and metabolic characterization between

105

Page 108: Comparative genetic and metabolic characterization between

106

FIGURE S4: Real-time quantitative PCR (qRT-PCR) validation of 6 gene copies of

flavonoid 3’-5’ hydroxylases identified by RNA-seq in berry skin of CG and RG.

Samples were collected at veraison and ripening stages in 2013. The gene copies of F3’5-H

corresponded to 2810, 2830, 2860, 2880, 2920 and 3010. The results are the mean of six

biological replicates, error bars represent standard error of mean. Asterisks indicate the result

of unpaired T-test. **, P < 0.0085; ***, P < 0.0003; ****, P < 0.0001. FW, Fresh Weight.

Page 109: Comparative genetic and metabolic characterization between

107

TABLE S1: Pool designation of CG and RG plants for RNA-seq analysis.

SKIN SAMPLE ID POOL NAME

Veraison Chimenti Globe 1

POOL-1A Veraison Chimenti Globe 2

Veraison Chimenti Globe 3

POOL-1B Veraison Chimenti Globe 4

Veraison Chimenti Globe 5

POOL-1C Veraison Chimenti Globe 6

Veraison Red Globe 1

POOL-2A Veraison Red Globe 2

Veraison Red Globe 3

POOL-2B Veraison Red Globe 4

Veraison Red Globe 5

POOL-2C Veraison Red Globe 6

Ripening Chimenti Globe 1

POOL-3A Ripening Chimenti Globe 2

Ripening Chimenti Globe 3

POOL-3B Ripening Chimenti Globe 4

Ripening Chimenti Globe 5

POOL-3C Ripening Chimenti Globe 6

Ripening Red Globe 1

POOL-4A Ripening Red Globe 2

Ripening Red Globe 3

POOL-4B Ripening Red Globe 4

Ripening Red Globe 5

POOL-4C Ripening Red Globe 6

Page 110: Comparative genetic and metabolic characterization between

108

TABLE S2: Alignment of RNA-seq reads to the Grapevine Reference Genome V2

using HISAT2

Sample

ID

Total reads

paired

Reads aligned

0 times

Reads aligned

1 time

Reads aligned

>1 time

Pool-1A 4575605

(100.00%)

508068

(11.1%)

3301327

(72.15%)

766210

(16.75%)

Pool-1B 4871122

(100.00%)

766210

(11.6%)

3508905

(72.03%)

797273

(16.37%)

Pool-1C 5350777

(100.00%)

642477

(12.01%)

3808111

(71.07%)

900189

(16.82%)

Pool-2A 6169343

(100.00%)

670998

(10.88%)

4482441

(72.66%)

1015904

(16.47%)

Pool-2B 6180772

(100.00%)

654089

(10.58%)

4493751

(72.71%)

1032932

(16.71%)

Pool-2C 5010252

(100.00%)

868444

(17.33%)

3387979

(67.62%)

753829

(15.05%)

Pool-3A 7690225

(100.00%)

1588674

(20.66%)

4964146

(64.55%)

1137405

(14.79%)

Pool-3B 5624673

(100.00%)

883057

(15.70%)

3820294

(67.92%)

921322

(16.38%)

Pool-3C 5360934

(100.00%)

715581

(13.35%)

3748929

(69.93%)

896424

(16.72%)

Pool-4A 6895816

(100.00%)

791137

(11.47%)

4928442

(71.47%)

1176237

(17.06%)

Pool-4B 7147979

(100.00%)

767762

(10.74%)

5160674

(72.20%)

1219543

(17.06%)

Pool-4C 6229341

(100.00%)

729131

(11.7%)

4436080

(71.21%)

1064130

(17.08%)

Page 111: Comparative genetic and metabolic characterization between

109

TABLE S3: Cleaning raw reads using Trimmomatic

Sample

ID

Input total

reads pairs

Both

Surviving

Forward Only

Surviving

Reverse Only

Surviving

Dropped

Pool-1A 4852418

(100.00%)

4575605

(94.30%)

178246

(3.67%)

60885

(1.25%)

37682

(0.78%)

Pool-1B 5173857

(100.00%)

4871122

(94.15%)

201139

(3.89%)

60598

(1.17%)

40998

(0.79%)

Pool-1C 5681319

(100.00%)

5350777

(94.18%)

213111

(3.75%)

72260

(1.27%)

45171

(0.80%)

Pool-2A 6553961

(100.00%)

6169343

(94.13%)

246950

(3.77%)

85694

(1.31%)

51974

(0.79%)

Pool-2B 6570611

(100.00%)

6180772

(94.07%)

258552

(3.93%)

79087

(1.20%)

52200

(0.79%)

Pool-2C 5311185

(100.00%)

5010252

(94.33%)

197744

(3.72%)

64427

(1.21%)

38762

(0.73%)

Pool-3A 8183865

(100.00%)

7690225

(93.97%)

309652

(3.78%)

111186

(1.36%)

72802

(0.89%)

Pool-3B 5982329

(100.00%)

5624673

(94.02%)

232378

(3.88%)

76608

(1.28%)

48670

(0.81%)

Pool-3C 5694271

(100.00%)

5360934

(94.15%)

213180

(3.74%)

74459

(1.31%)

45698

(0.80%)

Pool-4A 7331098

(100.00%)

6895816

(94.06%)

286456

(3.91%)

90148

(1.23%)

58678

(0.80%)

Pool-4B 7604337

(100.00%)

7147979

(94.00%)

298053

(3.92%)

95992

(1.26%)

62313

(0.82%)

Pool-4C 6606565

(100.00%)

6229341

(94.29%)

241663

(3.66%)

84617

(1.28%)

50944

(0.77%)

Page 112: Comparative genetic and metabolic characterization between

110

TABLE S4: Gene mapping analysis from aligned reads to Grapevine Reference Genome V2

Status Pool-1A Pool-1B Pool-1C Pool-2A Pool-2B Pool-2C Pool-3A Pool-3B Pool-3C Pool-4A Pool-4B Pool-4C

Assigned 4166463 4429083 4850826 5613766 5621724 4087521 6382908 4990397 4860962 6323211 6573172 5691548

Unassigned_Ambiguity 0 0 0 0 0 0 0 0 0 0 0 0

Unassigned_MultiMapping 310015 321664 355914 403352 433391 299855 316418 363941 369011 545779 528626 499145

Unassigned_NoFeatures 129072 141636 140843 185174 179281 379406 408854 121440 92241 131729 139122 125051

Unassigned_Unmapped 154182 170911 214395 205819 204309 420061 770711 369383 260927 230418 225009 218959

Unassigned_MappingQuality 0 0 0 0 0 0 0 0 0 0 0 0

Unassigned_FragmentLength 0 0 0 0 0 0 0 0 0 0 0 0

Unassigned_Chimera 0 0 0 0 0 0 0 0 0 0 0 0

Unassigned_Secondary 0 0 0 0 0 0 0 0 0 0 0 0

Unassigned_Nonjunction 0 0 0 0 0 0 0 0 0 0 0 0

Unassigned_Duplicate 0 0 0 0 0 0 0 0 0 0 0 0

Page 113: Comparative genetic and metabolic characterization between

111

TABLE S5: Reads from RNA-seq results for each 11 gene copies of F3’5-H.

Flavonoid 3’5’ hydroxylase

CG

Veraison

RG

Veraison

CG

Ripening

RG

Ripening

Gene 2805 2 50 0 38

Gene 2810 15 91 0 65

Gene 2830 32 214 1 150

Gene 2840 18 186 1 135

Gene 2846 11 59 0 38

Gene 2860 9 68 15 153

Gene 2873 4 38 2 70

Gene 2880 2 50 1 38

Gene 2920 20 154 2 104

Gene 2970 12 106 1 81

Gene 3010 30 182 61 354

Page 114: Comparative genetic and metabolic characterization between

112

REFERENCES FROM CHAPTER III

Adams, D. O. (2006). Phenolics and ripening in grape berries. American Journal of Enology

and Viticulture, 57(3), 249–256.

Alfenito, M. R., Souer, E., Goodman, C. D., Buell, R., Mol, J., Koes, R., & Walbot, V. (1998).

Functional Complementation of Anthocyanin Sequestration in the Vacuole by Widely

Divergent Glutathione S -Transferases, 10(July), 1135–1149.

Bogs, J., Ebadi, A., McDavid, D., & Robinson, S. P. (2006). Identification of the flavonoid

hydroxylases from grapevine and their regulation during fruit development. Plant

Physiology, 140(1), 279–91. http://doi.org/10.1104/pp.105.073262

Bolger, A. M., Lohse, M., & Usadel, B. (2014). Trimmomatic: A flexible trimmer for Illumina

sequence data. Bioinformatics, 30(15), 2114–2120.

http://doi.org/10.1093/bioinformatics/btu170

Boss, P. K., Davies, C., & Robinson, S. P. (1996). Analysis of the Expression of Anthocyanin

Pathway Genes. Plant Physiology, 111(1 996), 1059–1066.

http://doi.org/10.1104/pp.111.4.1059

Boss, P. K., Davies, C., & Robinson, S. P. (1996). Anthocyanin composition and anthocyanin

pathway gene expression in grapevine sports differing in berry skin colour. Australian

Journal of Grape and Wine Research, 2(3), 163–170. http://doi.org/10.1111/j.1755-

0238.1996.tb00104.x

Carbonell-Bejerano, P., Diago, M.-P., Martínez-Abaigar, J., Martínez-Zapater, J. M.,

Tardáguila, J., & Núñez-Olivera, E. (2014). Solar ultraviolet radiation is necessary to

enhance grapevine fruit ripening transcriptional and phenolic responses. BMC Plant

Page 115: Comparative genetic and metabolic characterization between

113

Biology, 14(1), 183. http://doi.org/10.1186/1471-2229-14-183

Cardoso, S., Lau, W., Dias, J. E., Fevereiro, P., & Maniatis, N. (2012). A Candidate-Gene

Association Study for Berry Colour and Anthocyanin Content in Vitis vinifera L. PLoS

ONE, 7(9). http://doi.org/10.1371/journal.pone.0046021

Castellarin, S. D., Di Gaspero, G., Marconi, R., Nonis, A., Peterlunger, E., Paillard, S., …

Testolin, R. (2006). Colour variation in red grapevines (Vitis vinifera L.): genomic

organisation, expression of flavonoid 3’-hydroxylase, flavonoid 3’,5’-hydroxylase genes

and related metabolite profiling of red cyanidin-/blue delphinidin-based anthocyanins in

berry skin. BMC Genomics, 7, 12. http://doi.org/10.1186/1471-2164-7-12

Conde, C., Silva, P., Fontes, N., Dias, A. C. P., Tavares, R. M., Sousa, M. J., … Gerós, H.

(2007). Biochemical changes throughout grape berry development and fruit and wine

quality. Food, 1, 1–22. http://doi.org/10.1093/jxb/ert395

COOMBE, B. G. (1995). Growth Stages of the Grapevine: Adoption of a system for identifying

grapevine growth stages. Australian Journal of Grape and Wine Research, 1(2), 104–110.

http://doi.org/10.1111/j.1755-0238.1995.tb00086.x

Dai, Z. W., Meddar, M., Renaud, C., Merlin, I., Hilbert, G., Delrot, S., & Gomès, E. (2013).

Long-term in vitro culture of grape berries and its application to assess the effects of sugar

supply on anthocyanin accumulation. Journal of Experimental Botany, 65(16), 4665–4677.

http://doi.org/10.1093/jxb/ert489

de Vetten, N., ter Horst, J., van Schaik, H.-P., de Boer, A., Mol, J., & Koes, R. (1999). A

cytochrome b 5 is required for full activity of flavonoid 3’,5’ -hydroxylase , a cytochrome

P450 involved in the formation of blue flower colors, 96(January), 778–783.

Degu, A., Hochberg, U., Sikron, N., Venturini, L., Buson, G., Ghan, R., … Fait, A. (2014).

Page 116: Comparative genetic and metabolic characterization between

114

Metabolite and transcript profiling of berry skin during fruit development elucidates

differential regulation between Cabernet Sauvignon and Shiraz cultivars at branching

points in the polyphenol pathway. BMC Plant Biology, 14(1), 1–20.

http://doi.org/10.1186/s12870-014-0188-4

Deluc, L., Barrieu, F. C., Marchive, C., Lauvergeat, V., Decendit, A., Richard, T., … Hamdi, S.

(2006). Characterization of a grapevine R2R3-MYB transcription factor that regulates the

phenylpropanoid pathway. Plant Physiology, 140(2), 499.

http://doi.org/10.1104/pp.105.067231

Deluc, L., Bogs, J., Walker, A. R., Ferrier, T., Decendit, A., Merillon, J.-M., … Barrieu, F.

(2008). The transcription factor VvMYB5b contributes to the regulation of anthocyanin

and proanthocyanidin biosynthesis in developing grape berries. Plant Physiology, 147(4),

2041–2053. http://doi.org/10.1104/pp.108.118919

Deluc, L. G., Grimplet, J., Wheatley, M. D., Tillett, R. L., Quilici, D. R., Osborne, C., …

Cramer, G. R. (2007). Transcriptomic and metabolite analyses of Cabernet Sauvignon

grape berry development. BMC Genomics, 8, 429. http://doi.org/10.1186/1471-2164-8-429

Dixon, R. A., Xie, D. Y., & Sharma, S. B. (2005). Proanthocyanidins - A final frontier in

flavonoid research? New Phytologist, 165(1), 9–28. http://doi.org/10.1111/j.1469-

8137.2004.01217.x

Dray, S., & Dufour, A.-B. (2007). The ade4 Package: Implementing the Duality Diagram for

Ecologists. Journal of Statistical Software, 22(4). http://doi.org/10.18637/jss.v022.i04

Dubrovina, A., & Kiselev, K. (2017). Regulation of stilbene biosynthesis in plants. Planta.

http://doi.org/10.1007/s00425-017-2730-8

Falginella, L., Castellarin, S. D., Testolin, R., Gambetta, G. A., Morgante, M., & Di Gaspero,

Page 117: Comparative genetic and metabolic characterization between

115

G. (2010). Expansion and subfunctionalisation of flavonoid 3’,5’-hydroxylases in the

grapevine lineage. BMC Genomics, 11(1), 562. http://doi.org/10.1186/1471-2164-11-562

Fasoli, M., Dal Santo, S., Zenoni, S., Tornielli, G. B., Farina, L., Zamboni, a., … Pezzotti, M.

(2012). The Grapevine Expression Atlas Reveals a Deep Transcriptome Shift Driving the

Entire Plant into a Maturation Program. The Plant Cell, 24(9), 3489–3505.

http://doi.org/10.1105/tpc.112.100230

Ford, C. M., Boss, P. K., & Hoj, P. B. (1998). Cloning and characterization of Vitis vinifera

UDP-glucose:flavonoid 3-O-glucosyltransferase, a homolog of the enzyme encoded by the

maize bronze-1 locus that may primarily serve to glucosylate anthocyanidins in vivo.

Journal of Biological Chemistry, 273(15), 9224–9233.

http://doi.org/10.1074/jbc.273.15.9224

Fortes, A. M., Agudelo-Romero, P., Silva, M. S., Ali, K., Sousa, L., Maltese, F., … Pais, M. S.

(2011). Transcript and metabolite analysis in Trincadeira cultivar reveals novel information

regarding the dynamics of grape ripening. BMC Plant Biology, 11(1), 149.

http://doi.org/10.1186/1471-2229-11-149

Fraige, K., González-Fernández, R., Carrilho, E., & Jorrín-Novo, J. V. (2015). Metabolite and

proteome changes during the ripening of Syrah and Cabernet Sauvignon grape varieties

cultured in a nontraditional wine region in Brazil. Journal of Proteomics, 113, 206–225.

http://doi.org/10.1016/j.jprot.2014.09.021

Gambetta, G. A., Matthews, M. A., Shaghasi, T. H., McElrone, A. J., & Castellarin, S. D. (2010).

Sugar and abscisic acid signaling orthologs are activated at the onset of ripening in grape.

Planta, 232(1), 219–234. http://doi.org/10.1007/s00425-010-1165-2

Ghan, R., Petereit, J., Tillett, R. L., Schlauch, K. A., Toubiana, D., Fait, A., & Cramer, G. R.

Page 118: Comparative genetic and metabolic characterization between

116

(2017). The common transcriptional subnetworks of the grape berry skin in the late stages

of ripening. BMC Plant Biology, 17(1), 94. http://doi.org/10.1186/s12870-017-1043-1

Ghan, R., Van Sluyter, S. C., Hochberg, U., Degu, A., Hopper, D. W., Tillet, R. L., … Cramer,

G. R. (2015). Five omic technologies are concordant in differentiating the biochemical

characteristics of the berries of five grapevine (Vitis vinifera L.) cultivars. BMC Genomics,

16(1), 946. http://doi.org/10.1186/s12864-015-2115-y

Giordano, D., Provenzano, S., Ferrandino, A., Vitali, M., Pagliarani, C., Roman, F., … Schubert,

A. (2016). Characterization of a multifunctional caffeoyl-CoA O-methyltransferase

activated in grape berries upon drought stress. Plant Physiology and Biochemistry, 101,

23–32. http://doi.org/10.1016/j.plaphy.2016.01.015

Gomez, L., Bancel, D., Rubio, E., & Vercambre, G. (2007). The microplate reader: an efficient

tool for the separate enzymatic analysis of sugars in plant tissues—validation of a micro-

method. Journal of the Science of Food and Agriculture, 87(10), 1893–1905.

http://doi.org/10.1002/jsfa.2924

Goodman, C. D., Casati, P., & Walbot, V. (2004). A Multidrug Resistance – Associated Protein

Involved in Anthocyanin Transport in Zea mays, 16(July), 1812–1826.

http://doi.org/10.1105/tpc.022574.weed

Guo, D.-L., Xi, F.-F., Yu, Y.-H., Zhang, X.-Y., Zhang, G.-H., & Zhong, G.-Y. (2016).

Comparative RNA-Seq profiling of berry development between table grape “Kyoho” and

its early-ripening mutant ’Fengzao’. BMC Genomics, 17(1), 795.

http://doi.org/10.1186/s12864-016-3051-1

He, F., Mu, L., Yan, G. L., Liang, N. N., Pan, Q. H., Wang, J., … Duan, C. Q. (2010).

Biosynthesis of anthocyanins and their regulation in colored grapes. Molecules, 15(12),

Page 119: Comparative genetic and metabolic characterization between

117

9057–9091. http://doi.org/10.3390/molecules15129057

Hichri, I., Heppel, S. C., Pillet, J., Léon, C., Czemmel, S., Delrot, S., … Bogs, J. (2010). The

basic helix-loop-helix transcription factor MYC1 is involved in the regulation of the

flavonoid biosynthesis pathway in grapevine. Molecular Plant, 3(3), 509–523.

http://doi.org/10.1093/mp/ssp118

Howe, E. A., Sinha, R., Schlauch, D., & Quackenbush, J. (2011). RNA-Seq analysis in MeV,

27(22), 3209–3210. http://doi.org/10.1093/bioinformatics/btr490

Kim, D., Langmead, B., & Salzberg, S. L. (2015). HISAT: a fast spliced aligner with low

memory requirements. Nature Methods, 12(4), 357–360.

http://doi.org/10.1038/nmeth.3317

Kiselev, K. V, Aleynova, O. A., Grigorchuk, V. P., & Dubrovina, A. S. (2016). Stilbene

accumulation and expression of stilbene biosynthesis pathway genes in wild grapevine

Vitis amurensis Rupr . http://doi.org/10.1007/s00425-016-2598-z

Kitamura, S., Shikazono, N., & Tanaka, A. (2004). TRANSPARENT TESTA 19 is involved in

the accumulation of both anthocyanins and proanthocyanidins in Arabidopsis. The Plant

Journal, 37(1), 104–114. http://doi.org/10.1046/j.1365-313X.2003.01943.x

Kobayashi, S., Goto-Yamamoto, N., & Hirochika, H. (2004). Retrotransposon-induced

mutations in grape skin color. Science (New York, N.Y.), 304(5673), 982.

http://doi.org/10.1126/science.1095011

Kuhn, N., Guan, L., Dai, Z. W., Wu, B. H., Lauvergeat, V., Gomes, E., … Delrot, S. (2014).

Berry ripening: Recently heard through the grapevine. Journal of Experimental Botany,

65(16), 4543–4559. http://doi.org/10.1093/jxb/ert395

Larronde, F., Krisa, S., Decendit, A., Chèze, C., Deffieux, G., & Mérillon, J. M. (1998).

Page 120: Comparative genetic and metabolic characterization between

118

Regulation of polyphenol production in Vitis vinifera cell suspension cultures by sugars.

Plant Cell Reports, 17(12), 946–950. http://doi.org/10.1007/s002990050515

Li, Q., He, F., Zhu, B.-Q., Liu, B., Sun, R.-Z., Duan, C.-Q., … Wang, J. (2014). Comparison of

distinct transcriptional expression patterns of flavonoid biosynthesis in Cabernet

Sauvignon grapes from east and west China. Plant Physiology and Biochemistry : PPB /

Societe Francaise de Physiologie Vegetale, 84(17), 45–56.

http://doi.org/10.1016/j.plaphy.2014.08.026

Liao, Y., Smyth, G. K., & Shi, W. (2013). The Subread aligner: Fast, accurate and scalable read

mapping by seed-and-vote. Nucleic Acids Research, 41(10).

http://doi.org/10.1093/nar/gkt214

Lijavetzky, D., Carbonell-Bejerano, P., Grimplet, J., Bravo, G., Flores, P., Fenoll, J., …

Martínez-Zapater, J. M. (2012). Berry flesh and skin ripening features in Vitis vinifera as

assessed by transcriptional profiling. PLoS ONE, 7(6).

http://doi.org/10.1371/journal.pone.0039547

Lijavetzky, D., Ruiz-Garcia, L., Cabezas, J. A., De Andrés, M. T., Bravo, G., Ibáñez, A., …

Martínez-Zapater, J. M. (2006). Molecular genetics of berry colour variation in table grape.

Molecular Genetics and Genomics, 276(5), 427–435. http://doi.org/10.1007/s00438-006-

0149-1

Love, M. I., Huber, W., & Anders, S. (2014). Moderated estimation of fold change and

dispersion for RNA-seq data with DESeq2. Genome Biology, 15(12), 550.

http://doi.org/10.1186/s13059-014-0550-8

Lunn, J. E., Delorge, I., Figueroa, C. M., Van Dijck, P., & Stitt, M. (2014). Trehalose

metabolism in plants. Plant Journal, 79(4), 544–567. http://doi.org/10.1111/tpj.12509

Page 121: Comparative genetic and metabolic characterization between

119

Lunn, J. E., Feil, R., Hendriks, J. H. M., Gibon, Y., Morcuende, R., Osuna, D., … Stitt, M.

(2006). Sugar-induced increases in trehalose 6-phosphate are correlated with redox

activation of ADPglucose pyrophosphorylase and higher rates of starch synthesis in

Arabidopsis thaliana. Biochemical Journal, 397(1), 139–148.

http://doi.org/10.1042/BJ20060083

Maeda, H., & Dudareva, N. (2012). The shikimate pathway and aromatic amino Acid

biosynthesis in plants. Annual Review of Plant Biology, 63, 73–105.

http://doi.org/10.1146/annurev-arplant-042811-105439

Mageroy, M. H., Tieman, D. M., Floystad, A., Taylor, M. G., Klee, H. J., Molecular, P., …

Sciences, H. (2012). A Solanum lycopersicum catechol- O -methyltransferase involved in

synthesis of the flavor molecule guaiacol, 1043–1051. http://doi.org/10.1111/j.1365-

313X.2011.04854.x

Martínez-Lüscher, J., Sánchez-Díaz, M., Delrot, S., Aguirreolea, J., Pascual, I., & Gomès, E.

(2014). Ultraviolet-B Radiation and Water Deficit Interact to Alter Flavonol and

Anthocyanin Profiles in Grapevine Berries through Transcriptomic Regulation. Plant and

Cell Physiology, 55(September 2014), 1925–1936. http://doi.org/10.1093/pcp/pcu121

Massonnet, M., Fasoli, M., Tornielli, G. B., Altieri, M., Sandri, M., Zuccolotto, P., … Pezzotti,

M. (2017). Ripening Transcriptomic Program in Red and White Grapevine Varieties

Correlates with Berry Skin Anthocyanin Accumulation. Plant Physiology, 174(4), 2376–

2396. http://doi.org/10.1104/pp.17.00311

Mathews, H., Clendennen, S. K., Caldwell, C. G., Liu, X. L., Connors, K., Matheis, N., …

Wagner, D. R. (2003). Activation Tagging in Tomato Identifies a Transcriptional Regulator

of Anthocyanin Biosynthesis , Modification , and Transport, 15(August), 1689–1703.

Page 122: Comparative genetic and metabolic characterization between

120

http://doi.org/10.1105/tpc.012963.1

Matus, J. T., Poupin, M. J., Cañón, P., Bordeu, E., Alcalde, J. A., & Arce-Johnson, P. (2010).

Isolation of WDR and bHLH genes related to flavonoid synthesis in grapevine (Vitis

vinifera L.). Plant Molecular Biology, 72(6), 607–620. http://doi.org/10.1007/s11103-010-

9597-4

Mi, H., Huang, X., Muruganujan, A., Tang, H., Mills, C., Kang, D., & Thomas, P. D. (2017).

PANTHER version 11: Expanded annotation data from Gene Ontology and Reactome

pathways, and data analysis tool enhancements. Nucleic Acids Research, 45(D1), D183–

D189. http://doi.org/10.1093/nar/gkw1138

Negri, A. S., Prinsi, B., Failla, O., Scienza, A., & Espen, L. (2015). Proteomic and metabolic

traits of grape exocarp to explain different anthocyanin concentrations of the cultivars.

Frontiers in Plant Science, 6(August), 603. http://doi.org/10.3389/fpls.2015.00603

Niu, N., Wu, B., Yang, P., & Li, S. (2013). Comparative analysis of the dynamic proteomic

profiles in berry skin between red and white grapes (Vitis vinifera L.) during fruit

coloration. Scientia Horticulturae, 164, 238–248.

http://doi.org/10.1016/j.scienta.2013.09.046

Pereira, G. E., Gaudillere, J. P., Pieri, P., Hilbert, G., Maucourt, M., Deborde, C., … Rolin, D.

(2006). Microclimate influence on mineral and metabolic profiles of grape berries. Journal

of Agricultural and Food Chemistry, 54(18), 6765–6775. http://doi.org/10.1021/jf061013k

Pérez-Díaz, R., Madrid-Espinoza, J., Salinas-Cornejo, J., González-Villanueva, E., & Ruiz-

Lara, S. (2016). Differential Roles for VviGST1, VviGST3, and VviGST4 in

Proanthocyanidin and Anthocyanin Transport in Vitis vinífera. Frontiers in Plant Science,

7(August), 1166. http://doi.org/10.3389/fpls.2016.01166

Page 123: Comparative genetic and metabolic characterization between

121

Petrussa, E., Braidot, E., Zancani, M., Peresson, C., Bertolini, A., Patui, S., & Vianello, A.

(2013). Plant Flavonoids — Biosynthesis , Transport and Involvement in Stress Responses,

14950–14973. http://doi.org/10.3390/ijms140714950

R Core Team. (2010). No Title. Vienna: R Foundation for Statistical Computing.

Reid, K. E., Olsson, N., Schlosser, J., Peng, F., & Lund, S. T. (2006). An optimized grapevine

RNA isolation procedure and statistical determination of reference genes for real-time RT-

PCR during berry development. BMC Plant Biology, 6, 27. http://doi.org/10.1186/1471-

2229-6-27

Savoi, S., Wong, D. C. J., Arapitsas, P., Miculan, M., Bucchetti, B., Peterlunger, E., …

Castellarin, S. D. (2016). Transcriptome and metabolite profiling reveals that prolonged

drought modulates the phenylpropanoid and terpenoid pathway in white grapes (Vitis

vinifera L.). BMC Plant Biology, 16(1), 67. http://doi.org/10.1186/s12870-016-0760-1

Schefe, J. H., Lehmann, K. E., Buschmann, I. R., Unger, T., & Funke-Kaiser, H. (2006).

Quantitative real-time RT-PCR data analysis: Current concepts and the novel “gene

expression’s C T difference” formula. Journal of Molecular Medicine, 84(11), 901–910.

http://doi.org/10.1007/s00109-006-0097-6

Serrano, A., Espinoza, C., Armijo, G., Inostroza-Blancheteau, C., Poblete, E., Meyer-Regueiro,

C., … Arce-Johnson, P. (2017). Omics Approaches for Understanding Grapevine Berry

Development: Regulatory Networks Associated with Endogenous Processes and

Environmental Responses. Frontiers in Plant Science, 8(September), 1–15.

http://doi.org/10.3389/fpls.2017.01486

Shimazaki, M., Fujita, K., Kobayashi, H., & Suzuki, S. (2011). Pink-colored grape berry is the

result of short insertion in intron of color regulatory gene. PLoS ONE, 6(6), 1–8.

Page 124: Comparative genetic and metabolic characterization between

122

http://doi.org/10.1371/journal.pone.0021308

Souquet, J. M., Cheynier, V., Brossaud, F., & Moutounet, M. (1996). Polymeric

proanthocyanidins from grape skins. Phytochemistry, 43(2), 509–512.

http://doi.org/10.1016/0031-9422(96)00301-9

Sweetman, C., Wong, D. C., Ford, C. M., & Drew, D. P. (2012). Transcriptome analysis at four

developmental stages of grape berry (Vitis vinifera cv. Shiraz) provides insights into

regulated and coordinated gene expression. BMC Genomics, 13(1), 691.

http://doi.org/10.1186/1471-2164-13-691

This, P., Lacombe, T., Cadle-Davidson, M., & Owens, C. L. (2007). Wine grape (Vitis vinifera

L.) color associates with allelic variation in the domestication gene VvmybA1. Theoretical

and Applied Genetics, 114(4), 723–730. http://doi.org/10.1007/s00122-006-0472-2

Vitulo, N., Forcato, C., Carpinelli, E., Telatin, A., Campagna, D., D’Angelo, M., … Valle, G.

(2014). A deep survey of alternative splicing in grape reveals changes in the splicing

machinery related to tissue, stress condition and genotype. BMC Plant Biology, 14(1), 99.

http://doi.org/10.1186/1471-2229-14-99

Walker, A. R., Lee, E., & Robinson, S. P. (2006). Two new grape cultivars, bud sports of

Cabernet Sauvignon bearing pale-coloured berries, are the result of deletion of two

regulatory genes of the berry colour locus. Plant Molecular Biology, 62(4–5), 623–635.

http://doi.org/10.1007/s11103-006-9043-9

Wang, H., Wang, W., Zhan, J., Yan, A., Sun, L., Zhang, G., … Xu, H. (2016). The accumulation

and localization of chalcone synthase in grapevine (Vitis vinifera L.). Plant Physiology and

Biochemistry. Elsevier Ltd. http://doi.org/10.1016/j.plaphy.2016.04.042

Wong, D. C. J., Lopez Gutierrez, R., Dimopoulos, N., Gambetta, G. A., & Castellarin, S. D.

Page 125: Comparative genetic and metabolic characterization between

123

(2016). Combined physiological, transcriptome, and cis-regulatory element analyses

indicate that key aspects of ripening, metabolism, and transcriptional program in grapes

(Vitis vinifera L.) are differentially modulated accordingly to fruit size. BMC Genomics,

17(1), 416. http://doi.org/10.1186/s12864-016-2660-z

Wu, B. H., Cao, Y. G., Guan, L., Xin, H. P., Li, J. H., & Li, S. H. (2014). Genome-wide

transcriptional profiles of the berry skin of two red grape cultivars (Vitis vinifera) in which

anthocyanin synthesis is sunlight-dependent or - Independent. PLoS ONE, 9(8).

http://doi.org/10.1371/journal.pone.0105959

Xie, H., Konate, M., Sai, N., Tesfamicael, K. G., Cavagnaro, T., Gilliham, M., … Lopez, C. M.

R. (2017). Global DNA Methylation Patterns Can Play a Role in Defining Terroir in

Grapevine (Vitis vinifera cv. Shiraz). Frontiers in Plant Science, 8(October), 1–16.

http://doi.org/10.3389/fpls.2017.01860

Yakushiji, H., Kobayashi, S., Goto-Yamamoto, N., Tae Jeong, S., Sueta, T., MI, N., & Azuma,

A. (2006). A Skin Color Mutation of Grapevine, from Black-Skinned Pinot Noir to White-

Skinned Pinot Blanc, Is Caused by Deletion of the Functional VvmybA1 Allele.

Bioscience, Biotechnology, and Biochemistry, 70(6), 1506–1508.

http://doi.org/10.1271/bbb.50647

Zarrouk, O., Vialet, S., Marlin, T., Chaves, M., Martinoia, E., & Nagy, R. (2013). ABCC1 , an

ATP Binding Cassette Protein from Grape Berry , Transports Anthocyanidin 3- O -

Glucosides, 25(May), 1840–1854. http://doi.org/10.1105/tpc.112.102152

Zenoni, S., Ferrarini, A., Giacomelli, E., Xumerle, L., Fasoli, M., Malerba, G., … Delledonne,

M. (2010). Characterization of transcriptional complexity during berry development in

Vitis vinifera using RNA-Seq. Plant Physiology, 152(4), 1787–1795.

Page 126: Comparative genetic and metabolic characterization between

124

http://doi.org/10.1104/pp.109.149716

Zhang, Z. Z., Li, X. X., Chu, Y. N., Zhang, M. X., Wen, Y. Q., Duan, C. Q., & Pan, Q. H.

(2012). Three types of ultraviolet irradiation differentially promote expression of shikimate

pathway genes and production of anthocyanins in grape berries. Plant Physiology and

Biochemistry, 57, 74–83. http://doi.org/10.1016/j.plaphy.2012.05.005

Page 127: Comparative genetic and metabolic characterization between

125

CHAPTER IV

OBJECTIVE 3: Evaluate the function of at least one gen of interest, selected

from the above objectives, related with the branching point F3’-H and F3’5’-

H of the anthocyanin biosynthesis pathway.

Page 128: Comparative genetic and metabolic characterization between

126

INTRODUCTION

In plant species including grapevine, the biosynthesis of cyanidin and delphinidin type

anthocyanins is determined by the activity of two enzymes, flavonoid 3’- and 3’,5’-

hydroxylases (F3’-H and F3’,5’-H) that add either a single hydroxyl group at the 3’ position or

two hydroxyl groups at the 3’ and 5’ positions to dihydrokaempferol and/or dihydroquercetin,

respectively. Once converted into 3’-dihydroquercetin (cyaniding type) or 3’,5’-

dihydromyricetin (delphinidin type), these intermediates flow through common downstream

enzymes to form di-substituted and tri-substituted anthocyanins and also other polyphenols

(flavonols, leucoanthocyanins, catechins, proanthocyanins, tannins) at different developmental

stages (He et al., 2010; Petroni & Tonelli, 2011; Tanaka, 2006). F3’-H and F3’,5’-H genes code

for CYP75 cytochrome P450 hydroxylases that have been well characterized in the ornamental

plants such as Petunia hybrid (Brugliera et al., 1994), Antirrhinum majus (Martin, Prescott,

Mackay, Bartlett, & Vrijlandt, 1991) among others. In ornamental species, F3’H and F3’,5’H

are found as low/medium copy number genes and have been genetically manipulated to produce

commercial new flower pigmentation which naturally do not produce certain types of

anthocyanins. In grapevine, gene copy number of F3’5’-H has increased through recurrent

cycles of duplication, give rise to a single copy gene on chromosome 8 and the others multiplies

copy gene in tandem on chromosome 6 (Falginella et al., 2010).

Cytochrome b5 proteins are components of electron transport systems found in various species

such as DIF-F in petunia flowers, which has been implicated as an alternative electron donor for

F3’5’-H (Hf1) that contributes to efficient accumulation of delphinidin-based anthocyanins (de

Vetten et al., 1999). Ectopic expression of these two genes of petunia, F3’5’H and cytochrome

Page 129: Comparative genetic and metabolic characterization between

127

b5, in a carnation cultivar that lack F3’5’H activity, resulted in efficient production of

delphinidin-based anthocyanins and subsequent petal color change (Tanaka & Aida, 2009). In

grapevine, it was observed in Syrah cultivar, that expression of the genes encoding these two

enzymes is increase in veraison stage related to the beginning of anthocyanin accumulation in

berry skin and were only identified in pigmented cultivars (Bogs, Ebadi, McDavid, & Robinson,

2006). Taking into account all these observations has been proposed to cytochrome b5 such a

possible a putative modulator of F3’5’-H and may regulate anthocyanin accumulation.

However, its relationship in the anthocyanins accumulation in grapevines has not been

demonstrated. Due to the fact that in RNA-seq results from Objetive 2 we identified 11 gene

copies of F3’5’-H repressed in CG compared to RG, and at the same time we observed Cytb5

repressed, we focus our interest in this gene to demonstrate its relationship anthocyanin

accumulation in grapevines We isolated Cytb5 coding sequence from CG and RG and we

generated a construction to later transform V. berlandieri x V. Rupestris 110R embryos that

overexpressed it. Our results showed that transgenic embryos overexpressing Cytb5 exhibited a

rapid growth and also colored areas where they were placed into light condition, not observed

in transgenic control line. The knowledge of their specific regulation could be important for

future attempts to engineer flavonoid composition in fruit.

Page 130: Comparative genetic and metabolic characterization between

128

MATERIAL AND METHODS

Binary vector constructions

From the same RNA extraction samples and subsequent cDNA synthesis from Objective 2 used

for qRT-PCR analysis (Santibanez et al., in preparation), were now used for generating binary

vector constructions containing CYTB5 gene. We amplified the coding sequence (CDS) of

CYTB5 (444bp) by PCR from CG and RG during veraison and ripening stages of season 2013.

Pimers used correspond to CYTB5 (ID: VIT_218s0001g09400): F5’-

CACCATGGCTCCCCAAAAAGTTT-3’ and R5’-TTAGGTGGTGAGGTGGGCAGCT-3’.

Also, was used primers of a housekeeping gene actin that amplified 666pb: F5’-

ACCAGAATCCAGCACAATC-3’ and R5’-ATGGCCGATACTGAAGATAT-3’.

PCR reactions were carried out using a T100™ Thermal Cycler (Bio-Rad Laboratories, Inc.) in

a 25 µl reaction mixture (1 µl template cDNA, 2.5 µl PCR Buffer 10X, 1 µl MgCl2 50mM, 0.3

µl DMSO, 0.5 µl each primer 10mM, 0.5 µl dNTPs 10mM, 0.13 µl Taq DNA polymerase

(Invitrogen)). Thermocycling conditions were as follows: one cycle at 95ºC for 7 min; 35 cycles

at 95ªC for 30 secs, 60ºC for 30 secs and 72ºC for 30 secs followed by one cycle at 72ºC for 7

min. Then, were cloned into pENTR™ Directional / SD/D-TOPO® (Invitrogen) using

chemically competent E. coli. Later, positive colonies transformed were used for plasmid

extraction using GeneJET Plasmid Miniprep Kit (Promega) according to manufacturer's

instructions, and then were sequenced. CYTB5-containing vector were recombined into

pK7FWG2 (Karimi, Inzé, & Depicker, 2002) using GatewayTM LR ClonaseTM II Enzyme

Mix (Invitrogen), using chemically competent E. coli and subsequent positive colonies

transformed were used for plasmid extraction using GeneJET Plasmid Miniprep Kit (Promega)

Page 131: Comparative genetic and metabolic characterization between

129

according to manufacturer's instructions. Two expression vector were generating,

P35S::VvCYTB5::T35S and control of empty pK7FWG2. Each expression vector was

introduced in A. tumefaciens strain GV3101 using electroporation.

Preparation of Agrobacterium culture

An overnight suspension culture of Agrobacterium already transformed with each binary vector

was prepared by inoculating 25 ml of LB (Luria-Bertani) liquid medium supplemented with 50

mg/l kanamycin and 100 mg/l rifampicin in a 125 ml flask with single colony-derived bacteria.

The culture was maintained at 28ºC on an orbital shaker (185 rpm) in dark. When the optical

density at 600 nm (OD600) reached a value of 0.8–1.0, bacterial suspension was transferred into

a sterile 50 ml centrifuge tube and spun down at 6000 rpm for 10 min at room temperature. The

supernatant was removed and 100 ml of liquid G1SCA medium (2.1 g of NITSCH and NITSCH

medium, 0.1 g of myo-inositol, 60 g of sucrose, 2.5 g of activated carbon, 5 g of Gelrite, 1 ml

of Gamborg 1000X vitamins, 1 mL of NOA (stock: 2 mg/ml), 1.66 ml of AIA (stock: 1.66

mg/ml), 0.2 ml of BAP (stock: 2 mg/ml), adjusted to a pH of 5.8 prior to add agar and activated

carbon), but without antibiotics and activated carbon was added to resuspend the cells by gentle

swirling agitation in a flask.

Inoculation of V. berlandieri x V. Rupestris 110R embryos with Agrobacterium tumefaciens

and subsequent maintenance in culture media

From flask containing 100 ml bacterial suspension, 10 ml was added in a sterile 15 ml centrifuge

tube for each binary vector construction, together with 0.5 g embryogenic tissue V. berlandieri

x V. Rupestris 110R and was maintained at 28ºC on an orbital shaker (185 rpm) for 10 min.

Page 132: Comparative genetic and metabolic characterization between

130

Bacterial suspension was withdrawn using a transfer pipette and any excess moisture was

removed by blotting with sterilized Whatman 3MM filter paper. The transformed embyros were

transferred in to GISCA medium with acetosyringone 200 mM and also with Whatman 3MM

filter paper and were cultivated for two days at 28ºC in dark. Subsequently, these embryos were

transferred into a flask containing G1SCA liquid medium with Timentin (300 mg/ml), to

eliminate the overgrowth of A. tumefaciens, during 3 days at 24ºC in dark. Again, embryos were

withdrawn using a transfer pipette and any excess moisture was removed by blotting with

sterilized Whatman 3MM filter paper. Dried embyros were transferred into GISCA medium

containing Timentin (300 mg/ml) and Kanamycin (50 mg/ml) to prevent the bacteria

overgrowth and transgenic embryo selection, respectively, for 1 month at 24ºC in dark. After

this month, embryos in torpedo stage were transferred into DE medium (2.21 g of MS basal

medium with vitamins, 0.05 g of micronutrients MS, 1 ml of vitamins MS 1000X, 0.1 g of myo-

inositol, 30 g of sucrose, 2.5 g of activated carbon and 5 g of Gelrite, adjusted to a pH of 5.8

prior to add agar and activated carbon) containing Timentin (300 mg/ml) and Kanamycin (50

mg/ml). After 2 months elapsed, the embryos were transferred during approximately 3 months

in GE medium (similar to DE medium but with additional 583μl of IAA hormone (stock: 3

mg/ml) and 350μl of GA3 hormone (stock: 1 mg/ml). When the germinated embryos will be

seedlings, it will be transferred to MS/2 medium (2.21 g of MS basal medium with vitamins,

0.1 g of myo-inositol, 30 g of sucrose, 2.5 g of activated carbon, 8.5 g of agar-agar and 3 ml

PPM).

Photographic record of phenotype of transgenic and control embryos

Page 133: Comparative genetic and metabolic characterization between

131

Each month of embryo subculture a photographic record was made using Camera Powershot

G16 Canon.

Page 134: Comparative genetic and metabolic characterization between

132

RESULTS

Generating constructions overexpressing CYTB5

For CYTB5 cloning in a binary destination vector, first was done PCR amplification from

cDNAs form CG and RG during veraison and ripening stages (FIGURE 7). From all samples

was obtained fragments of 444pb, which were cloned into pENTR™ Directional / SD/D-

TOPO®, and each reaction was used for chemically competent E. coli transformation and

subsequent plasmid extraction. A new fragment of 444pb were amplified by PCR from this

extraction (FIGURE 8) confirming that were cloned. CYTB5 cloned in TOPO-SD were

analyzed by sequencing and we observed that CG and RG presented the same sequence

described in database (data no show), suggesting that repression in gene expression of CYTB5

observed in RNA-seq and qRT-PCR results from objective 2, were not due to point mutations

in the CYTB5 coding sequence. Understanding the importance of demonstrate the relationship

of CYTB5 in the anthocyanin biosynthesis pathway in grapevine, an RG sequence was selected

for the following experiments. Then, a recombination reaction was done using GatewayTM LR

ClonaseTM II Enzyme Mix for obtain final expression vector pK7FWG2 (Karimi et al., 2002).

Recombination was checked with, a new CTYB5 PCR amplification was done (FIGURE 9A)

confirming a correct reaction. The next step was transform A. tumefaciens with final vector

P35S::VvCYTB5::T35S and control of empty pK7FWG2 using electroporation method. PCR

amplification of CYTB5 was performed in bacterial suspension for check positive colonies

(FIGURE 9B) and then it was done the plasmid extraction. Finally, a co-culture between A.

tumefaciens and V. berlandieri x V. Rupestris 110R embryos was performed.

Page 135: Comparative genetic and metabolic characterization between

133

FIGURE 7: CYTB5 amplification from cDNA using of CG and RG berries skin during

veraison and ripening stages. PCR reaction was done using specific forward primer containing

the sequence 5’CACC. Housekeeping corresponded to auxin gene and molecular marker

corresponded to 1 Kb Plus DNA Ladder.

Transgenic embryos overexpressing CTYB5

Transformed embryos were transferred into G1SCA medium for one month in dark at 24ºC,

containing kanamycin 50 mg/ml and timentin 300 mg/ml for transgenic embryos selection and

avoid overgrowth of A. tumefaciens used for transformation. It was observed a rapid grown in

transgenic embryos overexpressing CTYB5 compared to transgenic control condition (FIGURA

10A and 10B, respectively). Embryos that were in torpedo state were transferred into DE

medium for generate differentiation in them, and also DE medium contained antibiotic selection

for transgenic embryos and overgrowth of bacteria.

FIGURE 8: CYTB5 amplification from TOPO-SD of CG and RG berries skin during

veraison and ripening stages. PCR reaction was done using specific forward primer containing

the sequence 5’CACC. Molecular marker corresponded to 1 Kb Plus DNA Ladder.

Page 136: Comparative genetic and metabolic characterization between

134

After two more months in this medium (during each month the culture medium was renewed)

fulfilling 3 months in total of cultivation, another rapid growth was observed in transgenic

embryos compared to control (FIGURA 10C and 10D, respectively), where embryos

overexpressing CTYB5 were more differentiated. Also, in these embryos was observed

developed structures such the cotyledons. These differentiated embryos therefore, were

transferred into GE medium, similar to DE medium but containing growth hormones, auxin and

gibberellic acid, and were placed in light at 24ºC for generate germination in them. After one

month in GE medium, fulfilling 4 months in total of cultivation, transgenic embryos

overexpressing CYTB5 exhibited pigmented areas of red color indicated in yellow arrows and

also cotyledons more developed compared with control embryos that still remained in dark in

DE medium in that month (FIGURE 10E, 10F and 10G, respectively).

Finally, after another month in GE medium, fulfilling 5 months in total of cultivation, these

transgenic embryos overexpressing CYTB5 exhibited the same red color demonstrated that was

stable for 2 months, and also covering more areas along the embryo (FIGURE 10H). In

addition, it was observed that cotyledons started to produce chlorophyll due to green color

observed in them. The transgenic control embryos were passed to light conditions only in this

fifth month, showing a much slower growth (FIGURE 10I).

Page 137: Comparative genetic and metabolic characterization between

135

FIGURE 9: CYTB5 amplification of TOPO-SD recombination with binary destination

vector pK7FWG2 and final vector construction. (A) PCR amplification from recombination

reaction in E. coli. (B) PCR amplification in final expression vector in A. tu

Page 138: Comparative genetic and metabolic characterization between

136

Page 139: Comparative genetic and metabolic characterization between

137

FIGURE 10: Phenotype observed in transformed embryos during 5 months in media

culture. (A) and (B) Transgenic embryos overexpressing CYTB5 gene and transgenic control

embryos transformed with empty pK7FWG2, respectively. Correspond to the first month in

GISCA medium in dark at 24ºC. (C) and (D) Transgenic embryos overexpressing CYTB5 gene

and transgenic control embryos transformed with empty pK7FWG2, respectively. Correspond

to third month in DE medium in dark at 24ºC. (E) and (F) Transgenic embryos overexpressing

CYTB5 gene placed in light for one month in GE medium. Yellow arrows indicate embryos

areas generating reddish coloration. (G) transgenic control embryos transformed with empty

pK7FWG2 in dark condition in DE medium. (H) Transgenic embryos overexpressing CYTB5

gene placed in light for two months in GE medium. Yellow arrows indicate embryos areas

generating reddish coloration. (I) Transgenic control embryos transformed with empty

pK7FWG2 in dark in DE medium. All embryos transformed correspond to V. berlandieri x V.

Rupestris 110R.

Page 140: Comparative genetic and metabolic characterization between

138

DISCUSSION AND CONCLUSION

F3’-H and F3’5’-H genes have been intensively studied in flowers of various plant species (e.g.

carnation, petunia, or rose) due to the anthocyanin hydroxylation importance in determining

flower color shades. In petunia, it had been demonstrated that Cytb5 is required for full activity

of F3’5’-H, in order to produce the anthocyanins of the delphinidin type (de Vetten et al., 1999).

In grapevines, only it had been demonstrated that Cytb5 from Syrah variety, was expressing

after veraison stage at the same time with F3’5’-H and both genes were presented in red berry

skin varieties (Bogs et al., 2006). However, there is no evidence to suggest the participation of

this gene in the accumulation of anthocyanins in grapevines. Knowing that grapevine

embryogenic cultures have been for a long time as the targets of choice for inserting genes of

interest, since single cells on their surface can be prompted to develop into complete plants; and

in addition, that gene insertion into such totipotent cells results in grape plants that stably express

the required characteristic (Dhekney, Li, Grant, & Gray, 2016), we were interested in correlate

the expression of this gene with anthocyanins accumulation. The aim of this work was to isolate

Cytb5 CDS from RG and observed its participation in this process in grapevines. Here we

provide evidence suggestin that Cytb5 from RG (ID: VIT_218s0001g09400) can produced

coloration in transgenic embryos that normally do not produce it (FIGURE 10E, 10F and 10H).

Reporter genes have long proved useful for distinguishing transformed from untransformed cell

tissues or embryos and optimizing transformation protocols through expression quantification

(Matthews, Saunders, Gebhardt, Lin, & Koehler, 1995). Typical examples of reporter genes

correspond to the uidA gene (GUS) as well as fluorescent proteins (GFP or RFP) (Vidal et al.,

2010). In addition, expression of MYBA1 in somatic embryos in Thompson seedless variety

Page 141: Comparative genetic and metabolic characterization between

139

allowed to observed visible anthocyanin accumulation and it was demonstrated the potential of

this gene as a reporter gene for transient expression assays (Li, Dhekney, & Gray, 2011). A

similar result was observed in our experiment, suggesting that Cytb5 could be used such a new

reporter gene, due to the capacity to generate color in V. berlandieri x V. Rupestris 110R

embryos. However, we need more experiments to evaluate the gene stability during plant

development, if it produces normal plants and also considering that the use of these genes that

produced coloration, is restricted to experiments for which this phenomenon does not interfere

with the expression of investigated genes or pathways.

Overexpression experiments are very helpful approaches in learning about a gene of unknown

function. In our case, we fusion Cytb5 coding sequence with the strong constitutive promoter

cauliflower mosaic virus 35S (CaMV35S), a common promoter used for plant transformation

(Hull et al., 2017). However, the main disadvantage of this approach is that the gene product is

synthesized in excessive amounts, and this can be in tissues where it is not usually present (Jelly,

Valat, Walter, & Maillot, 2014). That is a possible explanation for our case with Cytb5, since

normally grapevine embryos do not produce pigments during their development to plants, and

this gene altered that process, a similar effect produced by Myba1.

It was observed that transgenic embryos overexpressing Cytb5 exhibited a rapid growth

compared with transgenic control (FIGURE 10H compare to 10I for example). F3’-H and

F3’5’-H are cytochrome P450 proteins whose activities are dependent on associated proteins,

such as cytochrome b5, catalyzing the transfer of electrons to their prosthetic heme group

(Vergeres and Waskell, 1995). It could be possible that the overexpression of it in grapevine

embryos would be improving the transfer of electrons in both cytochromes P450, as well as in

other reactions in which they could participate that still remains unknown. However, we need

Page 142: Comparative genetic and metabolic characterization between

140

more evidence for understand this process and wait for complete developed plants to compare

to control.

In conclusion, with these previous results in evaluation of gene expression of Cytb5, we may

suggest that for a first time this cytochrome from grapevine was involved in its participation in

anthocyanin accumulation and also may contribute in accelerated growth metabolism, but it is

only part of a speculation and therefore, several tests are needed to prove this.

Page 143: Comparative genetic and metabolic characterization between

141

CHAPTER V

GENERAL DISCUSSION

Page 144: Comparative genetic and metabolic characterization between

142

In this work grapevine ripening were studied at anthocyanin accumulation level related to

flavonoid biosynthesis, according to understand changes in transcripts and also metabolites in a

contrasted model of grapevine varieties, CG and RG. In first instance, study model could not be

explained in the traditional way related to changes in color berry skin, which is through genetic

changes in the MYBA1 transcription factor. We showed that anthocyanin biosynthesis was

altered at different level of structural genes codifying enzymes of this pathway according with

RNA-seq results, with particular interest in down-regulation of 11 multicopies of F3’5’-H in

CG variety during veraison and ripening stages of berry development. F3’-H together with

F3’5’-H are part of the branching point in flavonoid biosynthesis where both use as a substrate

dihydrokaempferol to produce 3’-dihydroquercetin (cyaniding type anthocyanin) or 3’,5’-

dihydromyricetin (delphinidin type anthocyanin), becoming a key step in the pathway to

produce the type of anthocyanins in the grapevines.

Metabolic changes between contrasted CG and RG varieties

The spontaneous generation of this new CG variety in the field, from a branch of the RG variety,

led us to wonder ourselves what was happening at metabolic level and then relate this to the

transcriptomic changes to be perform later. First, we performed a comparative metabolic

analysis of skin at two developmental stages, veraison and ripening, between CG and RG in two

seasons, 2013 and 2014. We evaluated primary metabolites, phosphorylated intermediates,

organic acids, trehalose 6-phosphate, hexoses and anthocyanins. The selection of veraison and

ripening stages was done due their relevance in the establishment of beginning and end of

anthocyanin accumulation in grape (Kennedy, 2002). We observed with PCA analysis that

developmental stages were the primary discriminant parameters in the comparative study when

Page 145: Comparative genetic and metabolic characterization between

143

PC1 was analyzed (Chapter 3). This observation is concordant with previous published studies

(Guo et al., 2016; Massonnet et al., 2017; Wong, Lopez Gutierrez, Dimopoulos, Gambetta, &

Castellarin, 2016; Wu et al., 2014) in which, despite the use of different varieties, the importance

of developmental stages appeared to overrule variety origin. The metabolites that explained el

43.75% of data variance due to veraison and ripening stages, some of importance appeared in

this process as T6P, a sugar phosphate considered as a metabolic signaling molecule in plants;

which regulates developmental growth (Lunn, Delorge, Figueroa, Van Dijck, & Stitt, 2014). In

vitis vinifera there is no clear evidence of control of T6P during berry development, although

several transcriptomic studies have shown that orthologue genes encoding T6P synthase, T6P

synthase/phosphatase and SnRK1 (Sucrose-non-fermentative Related Kinase 1, another sugar

signaling pathway component) are tightly regulated during berry development (Deluc et al.,

2007; Gambetta, Matthews, Shaghasi, McElrone, & Castellarin, 2010). Anyway, find T6P in

our results, may suggest its participation in grape berry development, focused on skin, an

approach that had not been observed (Dai et al., 2013).

Another metabolites involved in explained developmental stages between CG and RG were

shikimate and phenylalanine (Chapter 3) which have relation with general phenylpropanoid

pathway where is part flavonoid biosynthesis (Petrussa et al., 2013). Shikimate correspond to

an intermediary in shimikate pathway that mediates the carbon flow from carbohydrate

metabolism to phenylpropanoid and aromatic compound biosynthesis (Maeda & Dudareva,

2012; Zhang et al., 2012). On the other hand, phenylpropanoid biosynthesis begins with the

amino acid phenylalanine, which is a product of the shikimate pathway (Maeda & Dudareva,

2012), therefore, these two metabolites are directly linked (Tohge, Watanabe, Hoefgen, &

Fernie, 2013). In addition, phenylalanine content was higher in CG compared to RG during

Page 146: Comparative genetic and metabolic characterization between

144

veraison and ripening stages, maybe involved in alteration that caused color change in our study

model.

According to anthocyanin content, it was observed extreme differences between CG and RG

associated to total content and type of anthocyanins accumulated (Chapter 3). We observed that

CG contained less content of total anthocyanin compared to RG and also did not contain

delphinid-type anthocyanins, catalyzed by one of the two enzymes belong to branching point in

anthocyanin biosynthesis pathway, F3’5’-H (He et al., 2010). Taken together, these observations

suggest that according to our results, the anthocyanin biosynthesis pathway is affected.

Accordingly, an increase in F3’5’H activity have been previously reported to explain higher

concentration of tri-hydroxylated anthocyanins in Cabernet Sauvignon and Shiraz, pointing

F3’5’H plays as a key molecular player driving the phenylpropanoid metabolic flux towards

bluish anthocyanins production (Degu et al., 2014), but these will be discussed latter in RNA-

seq results.

In addition, there is another important point to analyze, that could also explain the difference in

anthocyanins between varieties. This refers to results in the PCA analysis, specifically related

to Component 2 that explained the differences between CG and RG. The main metabolites

involved were related to the sucrose biosynthesis pathway. It is known from literature that

increasing concentrations of glucose and fructose increase the accumulation of anthocyanins in

grapevine cell cultures (Larronde et al., 1998). Therefore, according to our results, we realize

that the lower number of intermediaries in CG towards the sucrose biosynthesis, could have an

implication at the final level of anthocyanins in CG. Or it could even be a consequence of the

change or blockage that is occurring in the pathway of anthocyanin biosynthesis in CG through

Page 147: Comparative genetic and metabolic characterization between

145

F3’5’-H. At least the question of what was happening is raised, and more studies would be

needed, for example, with applications with sugars in the CG variety and observe the results.

Transcriptomic changes between contrasted CG and RG varieties

Color mutations in grape berry skin are relatively frequent events and can be easily seen in the

vineyard. Most studies investigating the mechanisms of changes in berry skin color explain them

through the regulatory effect of MybA transcription factors, whose genes are closely clustered

in a single locus on chromosome 2, referred to as the “berry color locus” (Kobayashi, Goto-

Yamamoto, & Hirochika, 2004; Walker, Lee, & Robinson, 2006). One of these is the

transcription factor MybA1, a major determinant of grape berry anthocyanin biosynthesis

pathway (Kobayashi et al., 2004; Lijavetzky et al., 2006). MybA1 transcriptionally regulates

UFGT gene, which glycosylates anthocyanidins, stabilizing them and therefore responsible for

the anthocyanins accumulation in grape berry skins (Ford, Boss, & Hoj, 1998). White skinned

grape varieties that do not produce anthocyanins, contain a mutated allele of MYBA1 in which

the presence of the retrotransposon Gret1 in promoter region interrupts the normal transcription

process, resulting in a white berry allele (Kobayashi et al., 2004). Pigmented cultivars possess

at least one allele of the MybA1 locus that does not contain Gret1 and is called red berry allele

(This, Lacombe, Cadle-Davidson, & Owens, 2007). In the case of pink skin berry varieties of

oriental origin, a different situation was reported: a 33bp insertion was found in the second

intron of MybA1, generating low expression levels of MybA1 in grape from these cultivars

(Shimazaki, Fujita, Kobayashi, & Suzuki, 2011). Another different example is a bud sports of

Cabernet Sauvignon called “Malian” with red-grey berry skin, associated to the deletion of a

Page 148: Comparative genetic and metabolic characterization between

146

large region of chromosome 2 that includes both MybA1 and MybA2 genes (MybA2 is located

adjacent to MybA1 in the color berry locus) (Walker et al., 2006). In addition, light-red-skinned

‘Ruby Okuyama’ and more intense and uniform rosy-skinned ‘Benitaka’ correspond to bud

sports of white-skinned ‘Italia’ (Azuma et al., 2009). ‘Ruby Okuyama’ was caused by the

recovery of VvmybA1 expression and ‘Benitaka’ due to presence of a novel VvmybA1BEN allele

that restored VvmybA1 transcripts (Azuma et al., 2009). In our results, we reported that UFGT

gene (VIT_216s0039g02230) was repressed in CG at veraison and ripening stages, which could

suggest an alteration in MybA1 expression, according to literature describe above presented.

MybA1, however, was not deemed significantly affected in DEGs analysis. Nevertheless, we

decided to check for the presence of white and red allele in CG and RG, using the same set of

primer used in (Shimazaki et al., 2011), observing that both varieties are heterozygous for this

gene and that our results cannot be explained by structural changes in MybA1 DNA sequence.

This observation therefore led us to ask ourselves what was going on in CG variety and to be

able to answer this question we decided to carry out an analysis of global expression to in some

way “take a picture” of the genes that are being expressed in CG to be then compared with RG

and thus be able to arrive at a possible explanation of the change of coloration in berry skin.

We performed a skin transcriptome analysis using the same samples for metabolic analysis, at

the same two developmental stages using Illumina Hiseq2000 in these two near isogenic

varieties, RG and CG, in 2013 season. According to DEGs analysis, we reported that several

genes related to flavonoid pathway, which grouped in three main clusters of repression,

induction or a combination of both, were differentially expressed in CG compared to RG

(Chapter 3). The most interesting observation was that we found that 11 copies of F3’5’-H,

encoding a key enzyme in anthocyanin biosynthesis pathway were repressed in CG. This

Page 149: Comparative genetic and metabolic characterization between

147

enzyme corresponds branching point of the flavonoid pathway, where dihydrokaempferol is the

substrate shared in common (Tanaka, 2006). Dihydrokaempferol is converted via F3’5’-H to

dihydromyricetin, producing the glucoside, acetyl, coumaroyl and caffeoyl forms of

delphinidin, petunidin and malvidin, that correspond to the delphinidin-type (tri-hydroxylated)

anthocyanins of bluish color. Dihydrokaempferol is also used by F3’-H to produce

dihydroquercetin and gives rise to the biosynthesis of glucoside, acetyl and coumaroyl forms of

cyanidin and peonidin, that correspond to cyanidin-type (di-hydroxylated) anthocyanins or

reddish color (Degu et al., 2014). Indeed, this result allowed us to connect directly previously

metabolic results related to anthocyanin content, where CG berry skins did not contain

delphindin-type anthocyanins and therefore, is consistent with the repression of 11 gene copies

of F3’5’-H. In addition, another important observation was the adjacent location of each of these

copies in a single cluster within chromosome 6. It is known that there are many copies of F3’5’-

H in Vitis vinifera, and it was shown that F3’5’-H in grapevine are highly redundant and

predominately expressed in berry skins (Castellarin et al., 2006; Falginella et al., 2010).

However, to the best of our knowledge, there was no evidence of the repression of such high

number of gene copies at the same time in a single variety. This raises the question of how the

repression of the F3’5’-H genes occur in CG. At least two mechanisms can be predicted, among

others. Firstly, since all these genes are physically close together in a single region of

chromosome 6, an epigenetic can be postulated. Although left uninvestigated until recently, the

field of epigenetic regulation in grapevine has gain considerable attention in the last years; and

epigenetic regulation were shown to be potentially part of the so-called “terroir” effect in Shiraz

grapes from Australia (Xie et al., 2017). Secondly, a still to be discovered master regulator (i.e

a transcription factor) might be defective in the CG. For example, an identical transcription

Page 150: Comparative genetic and metabolic characterization between

148

factor binding motif might be present in the promoter regions of all 11 gene copies, which would

be in agreement with the hypothesis of a master regulator for these F3’5’-H genes. One possible

finding that could validate this theory is that all the hydroxylases under study are integrated into

the cytoplasmic membrane (data not showed, using UniProt) (Bateman et al., 2017), which

makes sense because anthocyanin are synthesized in the cytosol of the plant cell, and then they

are finally translocated in the vacuole for storage. And therefore, the regulation of all could be

given at the level of a master regulator since all hydroxylases would be fulfilling the same

function, in the same place in the cell. However, further work is necessary to decide between

these two hypotheses.

In parallel, several genes putatively involved in anthocyanin transport to the vacuoles were

differentially expressed in CG compared to RG at both veraison and ripening stages, such as

GST13 and ABC transporter proteins, which expression was down-regulated at both

developmental stages in CG. Several studies in Arabidopsis had been reported the ability of

GST1 and GST4 to rescue the phenotype of the Arabidopsis tt19-1 mutant, resulting in a

restoration of transport of anthocyanins and proanthocyanidins at least with GST4 (Pérez-Díaz,

Madrid-Espinoza, Salinas-Cornejo, González-Villanueva, & Ruiz-Lara, 2016). ABC

transporters have also been involved in vacuolar sequestration of flavonoids in maize

(Goodman, Casati, & Walbot, 2004) and in the transport of malvidin 3-O-glucoside into

vacuoles, in a strictly GSH-dependent fashion in grapevine (Zarrouk et al., 2013). These

findings reported in our study might suggest a dynamic process in transport and sequestration

of anthocyanin into vacuoles occurring in this variety context. Also, it would be interesting to

related if this alteration in anthocyanin transportation was caused by previous variation in

branching point related to F3’5’-H downregulated in CG variety. Anyways, due to these genes

Page 151: Comparative genetic and metabolic characterization between

149

have not been characterized in grapevines, they become in new possible candidates to study for

better understanding anthocyanin transport.

Another key genes found to be down-regulated and up-regulated in CG were CHS that encodes

the first enzyme of the flavonoid pathway (Petrussa et al., 2013; Wang et al., 2016) and STS,

the early enzyme in flavonoid metabolism and involved in resveratrol biosynthesis (Dubrovina

& Kiselev, 2017; Kiselev, Aleynova, Grigorchuk, & Dubrovina, 2016).

This possibly reveals a possible compensation point in flavonoid metabolism: CHS utilizes the

same substrates as STS but, is responsible for flavonoid-type compound formation. Thus, STS

would act as an escape valve to absorb the excess of substrate flux generated in the upper part

of the flavonoid metabolism because of the decrease in CHS activity. Taken together, this

suggests a multilevel alteration of the flavonoid pathway, and not only of its tri-hydroxylated

branch in CG variety.

The only transcription factor found in our DEGs results was MYB24 (VIT_214s0066g01090)

which was induced in CG. Recently, it was reported that MYB24 was induced by solar UV

radiation in Tempranillo berry skins, in conjunction with terpenoid metabolism induction

(Carbonell-Bejerano et al., 2014), and also was associated with terpene biosynthesis in berries

in response to drought (Savoi et al., 2016) and highly co-expressed with flower and fruit specific

terpene synthase (TPS) (Wong et al., 2016). So maybe MYB24 in CG activated terpene

biosynthesis pathway, theory that could be valid due to the induction of the sesquiterpene

synthase 10 (TPS10) in CG, observed in DEGS results (Chapter 3). In addition, in personal

communication with Tomás Matus, a recognized specialist in the MYB type transcription

Page 152: Comparative genetic and metabolic characterization between

150

factors in grapevine, told us about his work not yet published with MYB24. In a novel grape

mutant that had half red and half white berries, MYB24 was induced in half white part, and also

MYB24 can activate a terpene synthase, passing terpenes to fulfill a more photoprotective

function (such as anthocyanins) than aromatic function. Therefore, we think that what could be

happening in CG, is a change in the flow of the types of secondary metabolites synthesized,

where MYB24 would be taking this flow towards the synthesis of terpenes in CG and not

towards anthocyanins. Consequently, the terpenes would be supplying the photoprotective

function in CG, making this variety successful despite having a more attenuated color compared

to RG.

Finally, a gene that had not been previously functionally characterized in grapevines, and only

had been extrapolated its function in the anthocyanin biosynthesis according to observations in

ornamental plants, corresponds to Cytb5, another gene repressed in CG during veraison and

ripening stage (Chapter 3). Cytb5 has previously been involved in F3’5’-H activity modulation

in Petunia flowers, catalyzing the transfer of electrons to their prosthetic heme group (de Vetten

et al., 1999). In 2006, Bogs et al., demonstrated that a putative Cytb5 from grapevine cv Shiraz

modulated both F3’5’-H and F3’-H activities even though the exact mechanism remains to be

deciphered. Due to its participation in modulate F3’5’-H activity, we suspect that can exist a

possible link between the drastic down-regulation of 11 gene copies of F3’5’H and the

repression of the Cytb5 observed in CG. To demonstrate the relationship between Cybt5 in

anthocyanin accumulation we decided to overexpressed in V. berlandieri x V. Rupestris 110R

embryos, due to the fact that grapevine embryogenic cultures are very useful in study a particular

gene of interest (Dhekney et al., 2016). We observed that transgenic embryos overexpressing

Page 153: Comparative genetic and metabolic characterization between

151

Cytb5 exhibited coloration since they were placed into light conditions and in addition, was

observed a rapid growth of them compared to transgenic embryo control (FIGURE 10). We

proposed that Cytb5 could be used as reporter gene as MYBA1, but it need a longer follow-up

of the plants to really postulate it as a new candidate in genetic engineering in grapevines.

As summary, in FIGURE 11, the main events at the genetic level in CG are presented in a

simplified overview. In general, there is a decrease in the induction of several genes in the

anthocyanin biosynthesis pathway and also, activation of genes that drive the carbon flux to

other branches of the phenylpropanoid pathway, such as the induction of the COMT, goes

towards the biosynthesis of lignin; TPS10, which goes towards the biosynthesis of

sesquiterpenes; STS, with combined induction and repression dependent on the development

stage; and MYB24, previously characterized in induction works towards the biosynthesis of

terpenes . Therefore, we believe that in CG what could be happening is a switch in the carbon

flux from the anthocyanins to the biosynthesis of other secondary metabolites, thus supplying

in some way the photoprotective function of these compounds, since by containing less amount

of anthocyanins, and at the same time only contains dihydroxylated anthocyanins, could be a

variety without evolutionary success, but it not the case. And this change in the flow could be

led by MYB24, however, we know that it could also be an opposite effect, in which the

activation of these other secondary metabolites could have repressed the biosynthesis of

trihydroxylated anthocyanins, generating the phenotype observed in CG. Both theories could be

correct, and to test them, several experiments are needed to prove which one is appropriate to

the phenotype in CG.

Page 154: Comparative genetic and metabolic characterization between

152

Page 155: Comparative genetic and metabolic characterization between

153

FIGURE 11: General scheme proposed for Chimenti Globe variety. Simplified overview of

biosynthesis of flavonols, lignins, stilbenes, anthocyanins that belong to phenylpropanoid

pathway and terpenes, and also show the dynamics that would be occurring in CG at molecular

level. According to DEGS results, many genes were not induced (cyan color circles) and induced

(yellow color circles) in CG during veraison and ripening stages. In the case of STS, this gene

was not induced in veraison and induced during ripening stage. The final result was that

trihydroxylated anthocyanins such as delphinidin 3-glucoside, petunidin 3-glucoside and

malvidin 3-glucoside were not detected in CG, pathway branch leading by F3'5'H enzyme.

Therefore, the phenotype observed in CG would be given only by dihydroxylated anthocyanins,

cyanidin 3-glucoside and peonidin 3-glucoside leading by F3’H enzyme (showed in the scheme

enclosed with a pink rectangle). Terpene pathway is an additional route showed here, because

in CG were induced MYB24 and TPS10, related to activating terpene biosynthesis.

Abbreviations: PAL, Phenylalanine Ammonia Lyase; C4H, Cinnamate 4-Hydroxylase; 4CL, 4-

Coumaryol CoA Ligase; CHS, Chalcone Synthase; CHI, Chalcone Isomerase; F3H, Flavanone

3-Hydroxylase; F3’H, Flavonoid 3’-Hydroxylase; F3’5’H, Flavonoid 3’5’-Hydroxylase;

CYTB5, Cytochrome b5; FLS, Flavonol Synthase; DFR, Dihydroflavonol-4-reductase; LDOX,

Leucoanthocyanidin dioxygenase; UFGT, UDP glucose:flavonoid-3-O-glucosyltransferase;

OMT, O-Methyltransferase; GST13, Glutathione S-Transferase 13, ABC TRANSPORTER,

ATP-binding cassette transporter; TPS10, Sesquiterpene Synthase 10; DMAPP, Dimethylallyl

Pyrophosphate; IPP, Isopentenyl Pyrophosphate; COMT, Catechol-O-Methyltransferase; STS,

Stilbene Sythase.

Page 156: Comparative genetic and metabolic characterization between

154

Impact of the study model in agriculture

In general, berry color is a fundamental attribute for the quality of wine and in particular in table

grapes, for the quality of the fruit to be consumed. In table grapes, most of the commercial

varieties that now exist are white or black, only a few correspond to pink varieties, as is the case

of the Chimenti Globe characterized in this thesis.

The red berry trait is a very desired attribute in the fruit industry. It is common that the grape

berry is harvested before reaching maturity in order to have this desired characteristic, despite

not having all the organoleptic properties, suitable for consumption. From this point of view, it

is of great relevance to contribute with the pertinent scientific information to explain the

generation of color in grapevines and to suggest to genetic improvement programs the type of

genes that should be repressed or induced, if it want to obtain varieties with these characteristics.

In this sense a rapid procedure, without the need to wait for successive fructifications in the field

(from 3 to 4 years approximately) until the color stabilizes, could be to perform an RT-PCR for

the expression of F3'-H and F3'5 'H and based on its expression predict the type of anthocyanins

that will accumulate and the color that the grapes would have. Possibly the incorporation of this

variety found in the field, Chimenti Globe, in genetic improvement programs allows to transfer

these characteristics to new varieties of table grapes to achieve the desired colors. This would

be a very relevant attribute for genetic improvement programs in the world and especially for

breeding programs in Chile, given the potential of Chile as a main exporter of table grapes.

Page 157: Comparative genetic and metabolic characterization between

155

In addition, the finding that the type and concentration of anthocyanins in the grape berry could

be modulated through the differential expression of F3'-H and F3'5'-H, led us to believe that it

could be extrapolated to other fruits characterized by accumulating anthocyanins as is the case

of cherry (Prunus avium), raspberry (Rubus idaeus), strawberry (Fragaria x ananassa),

blackberry (Rubus spp.) and blueberry (Vaccinium corymbosum) and even native Chilean

berries such as Maqui (Aristotelia chilensis), calafate (Berberis microphylla) or murta (Ugni

molinae). And this would allow to modify these existing varieties to generate new colors, being

a great scientific contribution towards the biotechnological development of new commercial

fruit varieties

Page 158: Comparative genetic and metabolic characterization between

156

CHAPTER VI

CONCLUSION

Page 159: Comparative genetic and metabolic characterization between

157

This comparative study revealed important differences at the metabolic and transcriptomic level

between two contrasted color berry skin, but nevertheless near isogenic, table grapes. It was

possible to link the phenotype observed to an alteration in branching point of flavonoid pathway

that involved two cytochrome P450, F3’-H and F3’5’-H. Using RNA-seq we showed that

flavonoid metabolism was affected at different levels in CG compared to RG at both veraison

and ripening stage. While the nature of the mutation responsible for color changes in berry skin

of CG from its progenitor RG remains unknown, the comparative RNA profiling of the berry

development between CG and RG showed the deregulation of genes directly related to

anthocyanin biosynthesis: CHS, STS, Cytb5, UFGT, MYB24, F3’5’-H, F3’-H, CCoAOMT,

COMT, GST13, ABC transporter and ANP-like. A prominent result was the drastic down-

regulation of 11 gene copies of F3’5’H in CG positioned in a single cluster in chromosome 6,

which was totally consistent with the anthocyanin content observed in CG, with the absence of

delphinidin-type (tri-hydroxylated) anthocyanins. This left in evidence an arrangement

occurring in this metabolic process in CG compared to RG and open new types of regulation

undescribed in anthocyanin biosynthesis in grapevines, independent to MYBA1 alterations

previously described in other changes of color berry skin in grapevines. In addition, with

preliminary results observed in coloration in grapevine embryos due to overexpression of Cytb5,

we may suggest that for a first time this cytochrome from grapevine could be involved in its

participation in anthocyanin accumulation and regulating F3’5’-H activity as previously

demonstrated in ornamental flowers.

This study provides new insights in the genetic and metabolic background that may be

responsible of color change in CG compared to RG, that constitute a relevant biological material

Page 160: Comparative genetic and metabolic characterization between

158

to study in depth the mechanisms underlying berry skin color intensity. In addition, provides

new possible candidates to study in depth, as CYTB5 and MYB24, directly related with

anthocyanin accumulation in berry skin, being able to use them in generate changes in color

berry skin of new varieties, or even extrapolate their function in other species.

Page 161: Comparative genetic and metabolic characterization between

159

REFERENCES FROM CHAPTER IV and V

Azuma, A., Kobayashi, S., Goto-Yamamoto, N., Shiraishi, M., Mitani, N., Yakushiji, H., &

Koshita, Y. (2009). Color recovery in berries of grape (Vitis vinifera L.) “Benitaka”, a bud

sport of “Italia”, is caused by a novel allele at the VvmybA1 locus. Plant Science, 176(4),

470–478. http://doi.org/10.1016/j.plantsci.2008.12.015

Bateman, A., Martin, M. J., O’Donovan, C., Magrane, M., Alpi, E., Antunes, R., … Zhang, J.

(2017). UniProt: The universal protein knowledgebase. Nucleic Acids Research, 45(D1),

D158–D169. http://doi.org/10.1093/nar/gkw1099

Bogs, J., Ebadi, A., McDavid, D., & Robinson, S. P. (2006). Identification of the flavonoid

hydroxylases from grapevine and their regulation during fruit development. Plant

Physiology, 140(1), 279–91. http://doi.org/10.1104/pp.105.073262

Brugliera, F., Holton, T. A., Stevenson, T. W., Farcy, E., Lu, C.-Y., & Cornish, E. C. (1994).

Isolation and characterization of a cDNA clone corresponding to the Rt locus of Petunia

hybrida. The Plant Journal, 5(1), 81–92. http://doi.org/10.1046/j.1365-

313X.1994.5010081.x

Carbonell-Bejerano, P., Diago, M.-P., Martínez-Abaigar, J., Martínez-Zapater, J. M.,

Tardáguila, J., & Núñez-Olivera, E. (2014). Solar ultraviolet radiation is necessary to

enhance grapevine fruit ripening transcriptional and phenolic responses. BMC Plant

Biology, 14(1), 183. http://doi.org/10.1186/1471-2229-14-183

Castellarin, S. D., Di Gaspero, G., Marconi, R., Nonis, A., Peterlunger, E., Paillard, S., …

Testolin, R. (2006). Colour variation in red grapevines (Vitis vinifera L.): genomic

organisation, expression of flavonoid 3’-hydroxylase, flavonoid 3’,5’-hydroxylase genes

Page 162: Comparative genetic and metabolic characterization between

160

and related metabolite profiling of red cyanidin-/blue delphinidin-based anthocyanins in

berry skin. BMC Genomics, 7, 12. http://doi.org/10.1186/1471-2164-7-12

de Vetten, N., ter Horst, J., van Schaik, H.-P., de Boer, A., Mol, J., & Koes, R. (1999). A

cytochrome b 5 is required for full activity of flavonoid 3’,5’ -hydroxylase , a cytochrome

P450 involved in the formation of blue flower colors, 96(January), 778–783.

Degu, A., Hochberg, U., Sikron, N., Venturini, L., Buson, G., Ghan, R., … Fait, A. (2014).

Metabolite and transcript profiling of berry skin during fruit development elucidates

differential regulation between Cabernet Sauvignon and Shiraz cultivars at branching

points in the polyphenol pathway. BMC Plant Biology, 14(1), 1–20.

http://doi.org/10.1186/s12870-014-0188-4

Deluc, L. G., Grimplet, J., Wheatley, M. D., Tillett, R. L., Quilici, D. R., Osborne, C., …

Cramer, G. R. (2007). Transcriptomic and metabolite analyses of Cabernet Sauvignon

grape berry development. BMC Genomics, 8, 429. http://doi.org/10.1186/1471-2164-8-429

Dhekney, S. A., Li, Z. T., Grant, T. N. L., & Gray, D. J. (2016). Somatic Embryogenesis and

Genetic Modification of Vitis. In M. A. Germana & M. Lambardi (Eds.), In Vitro

Embryogenesis in Higher Plants (pp. 263–277). New York, NY: Springer New York.

http://doi.org/10.1007/978-1-4939-3061-6_11

Dubrovina, A., & Kiselev, K. (2017). Regulation of stilbene biosynthesis in plants. Planta.

http://doi.org/10.1007/s00425-017-2730-8

Falginella, L., Castellarin, S. D., Testolin, R., Gambetta, G. A., Morgante, M., & Di Gaspero,

G. (2010). Expansion and subfunctionalisation of flavonoid 3’,5’-hydroxylases in the

grapevine lineage. BMC Genomics, 11(1), 562. http://doi.org/10.1186/1471-2164-11-562

Ford, C. M., Boss, P. K., & Hoj, P. B. (1998). Cloning and characterization of Vitis vinifera

Page 163: Comparative genetic and metabolic characterization between

161

UDP-glucose:flavonoid 3-O-glucosyltransferase, a homolog of the enzyme encoded by the

maize bronze-1 locus that may primarily serve to glucosylate anthocyanidins in vivo.

Journal of Biological Chemistry, 273(15), 9224–9233.

http://doi.org/10.1074/jbc.273.15.9224

Gambetta, G. A., Matthews, M. A., Shaghasi, T. H., McElrone, A. J., & Castellarin, S. D. (2010).

Sugar and abscisic acid signaling orthologs are activated at the onset of ripening in grape.

Planta, 232(1), 219–234. http://doi.org/10.1007/s00425-010-1165-2

Goodman, C. D., Casati, P., & Walbot, V. (2004). A Multidrug Resistance – Associated Protein

Involved in Anthocyanin Transport in Zea mays, 16(July), 1812–1826.

http://doi.org/10.1105/tpc.022574.weed

Guo, D.-L., Xi, F.-F., Yu, Y.-H., Zhang, X.-Y., Zhang, G.-H., & Zhong, G.-Y. (2016).

Comparative RNA-Seq profiling of berry development between table grape “Kyoho” and

its early-ripening mutant ’Fengzao’. BMC Genomics, 17(1), 795.

http://doi.org/10.1186/s12864-016-3051-1

He, F., Mu, L., Yan, G. L., Liang, N. N., Pan, Q. H., Wang, J., … Duan, C. Q. (2010).

Biosynthesis of anthocyanins and their regulation in colored grapes. Molecules, 15(12),

9057–9091. http://doi.org/10.3390/molecules15129057

Hull, R., Covey, S. N., Dale, P., Hull, R., Covey, S. N., & Dale, P. (2017). Microbial Ecology

in Health and Disease Genetically modified plants and the 35S promoter : assessing the

risks and enhancing the debate Genetically modified plants and the 35S promoter :

assessing the risks and enhancing the debate, 2235(June), 0–5.

Jelly, N. S., Valat, L., Walter, B., & Maillot, P. (2014). Transient expression assays in grapevine:

A step towards genetic improvement. Plant Biotechnology Journal, 12(9), 1231–1245.

Page 164: Comparative genetic and metabolic characterization between

162

http://doi.org/10.1111/pbi.12294

Karimi, M., Inzé, D., & Depicker, A. (2002). GATEWAYTM vectors for Agrobacterium-

mediated plant transformation. Trends in Plant Science, 7(5), 193–195.

http://doi.org/10.1016/S1360-1385(02)02251-3

Kennedy, J. (2002). Understanding grape berry development. Winery and Vineyard, (August).

Kiselev, K. V, Aleynova, O. A., Grigorchuk, V. P., & Dubrovina, A. S. (2016). Stilbene

accumulation and expression of stilbene biosynthesis pathway genes in wild grapevine

Vitis amurensis Rupr . http://doi.org/10.1007/s00425-016-2598-z

Kobayashi, S., Goto-Yamamoto, N., & Hirochika, H. (2004). Retrotransposon-induced

mutations in grape skin color. Science (New York, N.Y.), 304(5673), 982.

http://doi.org/10.1126/science.1095011

Larronde, F., Krisa, S., Decendit, A., Chèze, C., Deffieux, G., & Mérillon, J. M. (1998).

Regulation of polyphenol production in Vitis vinifera cell suspension cultures by sugars.

Plant Cell Reports, 17(12), 946–950. http://doi.org/10.1007/s002990050515

Li, Z. T., Dhekney, S. A., & Gray, D. J. (2011). Use of the VvMybA1 gene for non-destructive

quantification of promoter activity via color histogram analysis in grapevine (Vitis

vinifera) and tobacco. Transgenic Research, 20(5), 1087–1097.

http://doi.org/10.1007/s11248-010-9482-6

Lijavetzky, D., Ruiz-García, L., Cabezas, J. A., De Andrés, M. T., Bravo, G., Ibáñez, A., …

Martínez-Zapater, J. M. (2006). Molecular genetics of berry colour variation in table grape.

Molecular Genetics and Genomics, 276(5), 427–435. http://doi.org/10.1007/s00438-006-

0149-1

Lunn, J. E., Delorge, I., Figueroa, C. M., Van Dijck, P., & Stitt, M. (2014). Trehalose

Page 165: Comparative genetic and metabolic characterization between

163

metabolism in plants. Plant Journal, 79(4), 544–567. http://doi.org/10.1111/tpj.12509

Maeda, H., & Dudareva, N. (2012). The shikimate pathway and aromatic amino Acid

biosynthesis in plants. Annual Review of Plant Biology, 63, 73–105.

http://doi.org/10.1146/annurev-arplant-042811-105439

Martin, C., Prescott, A., Mackay, S., Bartlett, J., & Vrijlandt, E. (1991). Control of Anthocyanin

Biosynthesis in Flowers of Antirrhinum-Majus. Plant Journal, 1(1), 37–49.

http://doi.org/DOI 10.1111/j.1365-313X.1991.00037.x

Massonnet, M., Fasoli, M., Tornielli, G. B., Altieri, M., Sandri, M., Zuccolotto, P., … Pezzotti,

M. (2017). Ripening Transcriptomic Program in Red and White Grapevine Varieties

Correlates with Berry Skin Anthocyanin Accumulation. Plant Physiology, 174(4), 2376–

2396. http://doi.org/10.1104/pp.17.00311

Matthews, B. F., Saunders, J. A., Gebhardt, J. S., Lin, J.-J., & Koehler, S. M. (1995). Reporter

Genes and Transient Assays for Plants. In J. A. Nickoloff (Ed.), Plant Cell Electroporation

and Electrofusion Protocols (pp. 147–162). Totowa, NJ: Springer New York.

http://doi.org/10.1385/0-89603-328-7:147

Pérez-Díaz, R., Madrid-Espinoza, J., Salinas-Cornejo, J., González-Villanueva, E., & Ruiz-

Lara, S. (2016). Differential Roles for VviGST1, VviGST3, and VviGST4 in

Proanthocyanidin and Anthocyanin Transport in Vitis vinífera. Frontiers in Plant Science,

7(August), 1166. http://doi.org/10.3389/fpls.2016.01166

Petroni, K., & Tonelli, C. (2011). Recent advances on the regulation of anthocyanin synthesis

in reproductive organs. Plant Science, 181(3), 219–229.

http://doi.org/10.1016/j.plantsci.2011.05.009

Petrussa, E., Braidot, E., Zancani, M., Peresson, C., Bertolini, A., Patui, S., & Vianello, A.

Page 166: Comparative genetic and metabolic characterization between

164

(2013). Plant Flavonoids—Biosynthesis, Transport and Involvement in Stress Responses.

International Journal of Molecular Sciences, 14(7), 14950–14973.

http://doi.org/10.3390/ijms140714950

Savoi, S., Wong, D. C. J., Arapitsas, P., Miculan, M., Bucchetti, B., Peterlunger, E., …

Castellarin, S. D. (2016). Transcriptome and metabolite profiling reveals that prolonged

drought modulates the phenylpropanoid and terpenoid pathway in white grapes (Vitis

vinifera L.). BMC Plant Biology, 16(1), 67. http://doi.org/10.1186/s12870-016-0760-1

Shimazaki, M., Fujita, K., Kobayashi, H., & Suzuki, S. (2011). Pink-colored grape berry is the

result of short insertion in intron of color regulatory gene. PLoS ONE, 6(6), 1–8.

http://doi.org/10.1371/journal.pone.0021308

Tanaka, Y. (2006). Flower colour and cytochromes P450. Phytochemistry Reviews, 5(2–3),

283–291. http://doi.org/10.1007/s11101-006-9003-7

Tanaka, Y., & Aida, R. (2009). Genetic engineering in floriculture. Molecular Techniques in

Crop Improvement: 2nd Edition, 695–717. http://doi.org/10.1007/978-90-481-2967-6_30

This, P., Lacombe, T., Cadle-Davidson, M., & Owens, C. L. (2007). Wine grape (Vitis vinifera

L.) color associates with allelic variation in the domestication gene VvmybA1. Theoretical

and Applied Genetics, 114(4), 723–730. http://doi.org/10.1007/s00122-006-0472-2

Tohge, T., Watanabe, M., Hoefgen, R., & Fernie, A. R. (2013). Shikimate and Phenylalanine

Biosynthesis in the Green Lineage. Frontiers in Plant Science, 4(March), 1–13.

http://doi.org/10.3389/fpls.2013.00062

Vidal, J. R., Gomez, C., Cutanda, M. C., Shrestha, B. R., Bouquet, A., Thomas, M. R., &

Torregrosa, L. (2010). Use of gene transfer technology for functional studies in grapevine.

Australian Journal of Grape and Wine Research, 16(SUPPL. 1), 138–151.

Page 167: Comparative genetic and metabolic characterization between

165

http://doi.org/10.1111/j.1755-0238.2009.00086.x

Walker, A. R., Lee, E., & Robinson, S. P. (2006). Two new grape cultivars, bud sports of

Cabernet Sauvignon bearing pale-coloured berries, are the result of deletion of two

regulatory genes of the berry colour locus. Plant Molecular Biology, 62(4–5), 623–635.

http://doi.org/10.1007/s11103-006-9043-9

Wang, H., Wang, W., Zhan, J., Yan, A., Sun, L., Zhang, G., … Xu, H. (2016). The accumulation

and localization of chalcone synthase in grapevine (Vitis vinifera L.). Plant Physiology and

Biochemistry. Elsevier Ltd. http://doi.org/10.1016/j.plaphy.2016.04.042

Wong, D. C. J., Lopez Gutierrez, R., Dimopoulos, N., Gambetta, G. A., & Castellarin, S. D.

(2016). Combined physiological, transcriptome, and cis-regulatory element analyses

indicate that key aspects of ripening, metabolism, and transcriptional program in grapes

(Vitis vinifera L.) are differentially modulated accordingly to fruit size. BMC Genomics,

17(1), 416. http://doi.org/10.1186/s12864-016-2660-z

Wu, B. H., Cao, Y. G., Guan, L., Xin, H. P., Li, J. H., & Li, S. H. (2014). Genome-wide

transcriptional profiles of the berry skin of two red grape cultivars (Vitis vinifera) in which

anthocyanin synthesis is sunlight-dependent or - Independent. PLoS ONE, 9(8).

http://doi.org/10.1371/journal.pone.0105959

Xie, H., Konate, M., Sai, N., Tesfamicael, K. G., Cavagnaro, T., Gilliham, M., … Lopez, C. M.

R. (2017). Global DNA Methylation Patterns Can Play a Role in Defining Terroir in

Grapevine (Vitis vinifera cv. Shiraz). Frontiers in Plant Science, 8(October), 1–16.

http://doi.org/10.3389/fpls.2017.01860

Zarrouk, O., Vialet, S., Marlin, T., Chaves, M., Martinoia, E., & Nagy, R. (2013). ABCC1 , an

ATP Binding Cassette Protein from Grape Berry , Transports Anthocyanidin 3- O -

Page 168: Comparative genetic and metabolic characterization between

166

Glucosides, 25(May), 1840–1854. http://doi.org/10.1105/tpc.112.102152

Zhang, Z. Z., Li, X. X., Chu, Y. N., Zhang, M. X., Wen, Y. Q., Duan, C. Q., & Pan, Q. H.

(2012). Three types of ultraviolet irradiation differentially promote expression of shikimate

pathway genes and production of anthocyanins in grape berries. Plant Physiology and

Biochemistry, 57, 74–83. http://doi.org/10.1016/j.plaphy.2012.05.005