• Non ci sono risultati.

Dissecting molecular and physiological response mechanisms to high solar radiation in cyanic and acyanic leaves: a cse study on red and green basil

N/A
N/A
Protected

Academic year: 2021

Condividi "Dissecting molecular and physiological response mechanisms to high solar radiation in cyanic and acyanic leaves: a cse study on red and green basil"

Copied!
36
0
0

Testo completo

(1)

Dissecting molecular and physiological response mechanisms to high solar

radiation in cyanic and acyanic leaves: a case study on red and green basil

Massimiliano Tattini1* (massimiliano.tattini@ipsp.cnr.it) Federico Sebastiani1 (federico.sebastiani@ipsp.cnr.it) Cecilia Brunetti2,3 (cbrunetti@ivalsa.cnr.it) Alessio Fini2 (alessio.fini@unifi.it) Sara Torre1 (sara.torre@ipsp.cnr.it) Antonella Gori2 (antonella.gori@unifi.it) Mauro Centritto3 (mauro.centritto@cnr.it) Francesco Ferrini2 (francesco.ferrini@unifi.it) Marco Landi4 (marco.landi@forunipi.it) Lucia Guidi4 (lucia.guidi@unipi.it)

1 National Research Council of Italy, Dept. of Biology, Agriculture and Food Sciences, Institute for Sustainable Plant Protection, I- 50019 Sesto Fiorentino, Florence, Italy,

2 Department of Agri-Food Production and Environmental Sciences, University of Florence, Viale delle Idee 30, I-50019, Sesto Fiorentino, Florence, Italy

3 National Research Council of Italy, Dept. of Biology, Agriculture and Food Sciences, Trees and Timber Institute, I- 50019 Sesto Fiorentino, Florence, Italy

4 Department of Agriculture, Food and Environment, University of Pisa, I-56124 Pisa, Italy *Corresponding author: Massimiliano Tattini; Phone +39 055 4574038

Highlight

The attenuation of blue-green light by epidermal anthocyanins evokes shade-avoidance responses, affects leaf development and gas exchange performance, the biosynthesis and catabolism of MEP-derived products, and the biosynthesis of flavonols

Running title: Response mechanisms to high sunlight in red and green leaves Submission date: February 22 2017

Total words: 6267 /5852 3 Tables; 7 Figures

4 Tables and 1 Figure in Supplementary Material 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27

(2)

Abstract

Photosynthetic performance, the expression of genes involved in light signaling, and in the biosynthesis of isoprenoids and phenylpropanoids were analyzed in green (TIG) and red (RR) basil to dissect physiological and molecular response mechanisms to high sunlight. We show that attenuation of blue-green light by epidermal anthocyanins evokes shade-avoidance responses with consequential effects on leaf morpho-anatomical traits and gas exchange performance: RR had lower mesophyll conductance, partially compensated by less effective control of stomata movements in comparison to TIG. Photosynthesis declined more in TIG than in RR facing high sunlight, because of larger stomatal limitations and transient impairment of PSII photochemistry. The MEP pathway mostly promoted the synthesis and de-epoxidation of violaxanthin-cycle pigments in TIG and of neoxanthin and lutein in RR. This allowed green leaves to process effectively the excess of radiant energy, and red leaves to optimize light harvesting and photoprotection. Enhanced GE de-glucosylation and reduced ABA-oxidation, but not superior de-novo ABA synthesis, sustained the greater stomata closure observed in TIG than in RR. Our study reveals competition between anthocyanin and flavonol biosynthesis. This raises the possibility that red individuals, while better protected against supernumerary photons over the visible-waveband, might suffer more than green morphs from excessive UV-radiation.

Key words: ABA-oxidation; ABA-GE de-glucosylation; anthocyanin vs. flavonol biosynthesis; light-responsive transcription factors; MEP and phenylpropanoid pathways; mesophyll and stomatal conductance; RNA-Seq and qT-RT PCR; shade avoidance responses.

Introduction

After more than three decades of extensive research, whether red individuals perform better than green individuals when exposed to excessive solar radiation, remains a highly debated matter (Hughes, 2011; Landi et al., 2015). There are several reasons responsible for these ‘apparently’ unsuccessfully experiments. Firstly, red colouring of leaves is a transient or a permanent condition. Leaves of some species are red during the juvenile phase (Liakopoulos et al., 2006; Zeliou et al., 2009) or become red during autumn-winter (i.e., ‘winter reddening’, Kytridis et al., 2008; Hughes, 2011). In other species, genotypes display leaves that are red throughout the entire life cycle (Kyparissis et al., 2007; Tattini et

al., 2014; Logan et al., 2015). Secondly, anthocyanins may be located in different leaf tissues.

Anthocyanins occur in both the adaxial and the abaxial epidermis (Hughes and Smith, 2007; Tattini et 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56

(3)

al., 2014), in just one of the two epidermises (Hughes et al., 2005; Hughes and Smith, 2007), or may be

distributed in the mesophyll (Gould and Quinn, 1999; Kytridis and Manetas, 2006). Thirdly, “not all anthocyanins are born equal” (Kovinich et al., 2014) and, consequently, have different abilities to absorb the visible wavelengths of the solar spectrum (Merzlyak et al., 2008; Jordheim et al., 2016).

Strikingly different experimental set-ups, which include both plant material and growth conditions, add further complexity in providing conclusive evidence about the photoprotective role of anthocyanins. Usually, the photosynthetic machinery is better protected in cyanic than in acyanic leaves under transient exposure to severe light excess (Feild, et al., 2001; Hughes and Smith, 2007). Photosystem II (PSII) photochemistry also recovers faster in red than in green leaves during relief from photoinhibitory light (Landi et al., 2014; Logan et al., 2015). More contradictory observations stem from studies conducted on long-term basis in the field (Steyn et al., 2002; Gould, 2004). Photosynthesis, at saturating light, did not vary in a range of co-occurring winter-red and winter-green angiosperms (Hughes and Smith, 2007; Hughes, 2011), as also observed in high light-grown red and green basil (Tattini et al., 2014). In contrast, red individuals of Cistus creticus and Pistacia lentiscus displayed inherent photosynthetic inferiority, and incurred greater risk of photoinhibition than did green individuals (Kytridis et al., 2008; Zeliou et al., 2009; Nikoforou et al., 2011).

The ability of anthocyanins to absorb over the blue-green region of the solar spectrum (Feild et

al., 2001) may profoundly alter cryptochrome-dependent photomorphogenesis in red morphs and evoke

shade avoidance responses (Keller et al., 2011; Keuskamp et al., 2011; Zhang and Folta, 2012; Pedmale et al., 2016). Red individuals are less plastic indeed than green individuals to changes in sunlight availability, sometimes referred as to the ‘shade-syndrome’ displayed red leaves (Manetas et

al., 2003). Consistently, red morphs display leaves that are thinner and with less compact mesophyll

than the leaves of green morphs when fully exposed to solar radiation (Kyparissis et al., 2007; Lan et

al., 2011; Tattini et al., 2014). The relative abilities of green and red individuals to modify leaf

morpho-anatomical traits (e.g., leaf thickness and mesophyll compactness) depending on light availability (not only to transmit blue-green light at different rates in the leaf), may largely affect their gas exchange performance (Terashima et al., 2011). The issue has not received attention, but compelling evidence emphasizes the role of mesophyll conductance to CO2 (gm), which strongly

depends on the mesophyll anatomy (Flexas et al., 2016), in limiting photosynthesis in shade- or sun-adapted plants, particularly when challenged by an excess of solar irradiance (Fini et al., 2016; Peguero-Pina et al., 2017). 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87

(4)

The attenuation of solar radiation by epidermal anthocyanins may also affect the biosynthesis of photosynthetic and non-photosynthetic pigments. While the concentration of chlorophyll is usually higher, the biosynthesis and de-epoxidation of violaxanthin-cycle (VAZ) pigments is frequently lower in red than in green leaves (Steyn et al., 2002; Hughes, 2011; Landi et al., 2015). This apparently reduces the capacity of red leaves to process excess energy (e.g., through nonphotochemical quenching, NPQ, Demmig-Adams et al., 2012) and limit photo-oxidative stress in the chloroplast (Esteban et al., 2015) in comparison to green leaves. On the contrary, there is contrasting evidence about the phenylpropanoid metabolism in cyanic and acyanic leaves (Stich et al., 1992; Hughes et al., 2010; Luo

et al., 2016). In a few instances, red morphs display more active phenylpropanoid biosynthesis than

green morphs (Gould et al., 2002; Neill et al., 2002; Hada et al., 2003). Nonetheless, red basil invested considerably smaller amounts of fresh assimilated carbon to phenylpropanoid, particularly towards colourless flavonoid biosynthesis during shade-to-sun acclimation as compared to green basil (Tattini

et al., 2014).

In our study, analyses were conducted of photosynthetic performance, targeted transcriptomics and metabolomics to dissect molecular and physiological response mechanisms to high solar irradiance in green (‘Tigullio’, TG) and red (‘Red Rubin’, RR) basil growing during Mediterranean summer. We performed measurements of gas exchange, including response curves of CO2 assimilation to changes in photosynthetic photon flux density (AN/PPFD), intercellular (AN/Ci) and chloroplast CO2 concentration (AN/Cc), and of chlorophyll a fluorescence kinetics. Expression analysis of genes regulating photo-morphogenesis, as well as the biosynthesis of isoprenoids and phenylpropanoids was conducted using RNA-seq and qRT-PCR. Individual carotenoids and phenylpropanoids, free-ABA and its glucoside ester (ABA-GE) were quantified using HPLC-DAD-MS analysis. To the best of our knowledge, this is the first report dissecting the response mechanisms, from the molecular to the whole-leaf level, of cyanic and acyanic individuals to high solar irradiance.

Materials and methods

Plant material and growth conditions

Three-week-old potted green (‘Tigullio’, TIG) and red basil (‘Red Rubin’, RR), were grown under natural sunlight irradiance in Florence, Italy (43°46’N, 11°15’E) under minimum/maximum temperatures of 18.5 ± 1.6/32.5 ± 2.4 °C (mean ± SD, n = 28) over a four-week-period in July. Photosynthetic photon flux density (estimated over the 400–700 nm solar waveband) at midday was 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117

(5)

1850 ± 155 µmol quanta m-2 s-1 (mean ± SD, n = 28). Four-week-old leaves, newly developed in full sunlight were used for measurements.

Gas exchange and chlorophyll a fluorescence kinetics

Diurnal changes in CO2 assimilation (AN) and stomatal conductance (gs) were recorded in situ from

08:30 to 17:30 h with a LI-6400 portable photosynthesis system (LI-Cor, Lincoln, NE, USA) operating at 400 µmol mol−1 ambient CO

2 concentration. Daily carbon gain was calculated as described in Tattini

et al. (2004). Response curves of CO2 assimilation to changes in photosynthetic photon flux density

(AN/PPFD) and substomatal CO2 concentration (AN/Ci) were obtained as reported recently in Fini et al.

(2016). AN/Ci curves were performed in leaves exposed at saturating PPFD (1200 or 900 µmol quanta

m-2 s-1 for RR or TIG, respectively). In the cuvette, leaves were kept at a relative humidity of 55-60% and at a temperature of 30 °C. The apparent maximum rate of carboxylation by Rubisco (Vc,max (Ci))

was calculated from AN/Ci curves as reported in Fini et al. (2016).

Mesophyll conductance to CO2 (gm) was estimated using the variable J method (Harley et al., 1992)

according to the equation:

gm =

AN

C i−Γ∗[J f +8( A+R day)]

J f −4 ( A+R day )

(1)

where Γ* is the apparent CO2 compensation point to photorespiration, Jf is the rate of linear electron transport estimated from chlorophyll fluorescence, and (AN + Rday) is the gross photosynthetic rate (including non-photorespiratory respiration). Γ* was calculated using the Rubisco specific factor (τ) estimated for annual herbs (91.3 at 25 °C, see Galmés et al., 2005) as:

Γ* = 0.5 O/τ (2)

where O is the oxygen concentration.

Jf was calculated as:

Jf = ΦPSII ×·α × 0.5 × PPFD (3)

where α, the leaf absorptance, measured with a spectroradiometer equipped with an integrating sphere (LI-Cor 1800-12S), was 0.9 in ‘Red Rubin’ and 0.85 in ‘Tigullio’; the factor 0.5 was chosen assuming an equal distribution of absorbed photons between PSI and PSII. After every measurement, it was 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149

(6)

verified that Jf corresponded to electron transport rate under non-photorespiratory conditions (i.e. lowering O2 concentration at 1%) estimated by gas exchange (Loreto et al., 1992; Laisk and Loreto, 1996). The CO2 concentration in the chloroplasts (Cc) was then calculated as:

Cc = Ci  AN/gm (4)

Mesophyll conductance (gm) measured in each specific step of the curve, allowed calculate Cc from the

corresponding Ci value, and draw response curves of CO2 assimilation to chloroplastic CO2 concentration (AN/Cc curves), from which actual maximum rate of carboxylation by Rubisco (Vc,max,

(Cc)) was calculated according to Sharkey et al. (2007).

A modulated Chl a fluorescence analysis was conducted on dark-adapted (over a 40-min period) leaves sampled at different hours of the day using a PAM-2000 fluorometer (Walz, Effeltrich, Germany) connected to a Walz 2030-B leaf-clip holder through a Walz 2010-F fiber optic cable. The maximum efficiency of photosystem II (PSII) photochemistry was calculated as Fv/Fm = (Fm − F0)/Fm,

where Fv is the variable fluorescence and Fm is the maximum fluorescence of dark-adapted leaves. The

minimal fluorescence, F0, was measured using a modulated light pulse < 1 μmol m-2 s-1, to avoid

appreciable variable fluorescence. Fm and F′m were determined at 20 kHz using a 0.8-s saturating light

pulse of white light at 8000 μmol m-2 s-1. F′

m, the maximum fluorescence in light conditions, was

estimated at 900 or 1200 μmol quanta m-2 s-1 in TIG and RR, respectively (based on PPFDs at which saturation of photosynthesis occurred in TIG and RR). PSII quantum yield in the light (ΦPSII) and nonphotochemical quenching [NPQ = (Fm/F′m) − 1] were estimated using the saturation pulse method of Schreiber et al. (1986), and calculated according to Genty et al. (1989) and Bilger and Björkman (1990), respectively.

RNA-Seq analysis

Gene expression levels were calculated using the RPKM (Reads per Kilobase per Million Mapped reads) method (Mortazavi et al., 2008), by mapping raw reads of green and red basil to the basil reference transcriptome. The reference transcriptome of sweet basil was assembled using the Illumina platform as detailed in Torre et al. (2016). Sequence reads have been deposited at the Sequence Read Archive (SRA) of the National Centre of Biotechnology Information (NCBI, Bethesda, MD, USA) under the accession number SRA313233.

Gene expression analysis by quantitative Real-Time PCR (qRT-PCR)

150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180

(7)

Total RNA (2 μg), treated with RNAse-free DNaseI (Invitrogen Ambion, Carlsbad, CA, USA), was used as the template for cDNA synthesis, performed using SuperScript® VILO cDNA Synthesis Kit (Invitrogen), as previously reported (Torre et al., 2016). qRT-PCR analysis was carried out in a StepOnePlus real-time PCR system (Applied Biosystems, Foster City, CA, USA) using SYBR green technology (Power SYBR green PCR Master Mix; Applied Biosystems) according to the manufacturer’s instruction, under the following cycling conditions: 10 min at 95°C, 40 cycles of 15 s at 95°C, 1 min at 60°C. Amplification specificity was checked by the analysis of dissociation curve. Transcripts abundance, expressed as normalized relative quantities (NRQs), was calculated by geometric normalization against the mean of three basil housekeeping genes, namely actin (ACT), tubulin (TUB) and glyceraldehyde 3-phosphate dehydrogenase (GAPDH) (Vandesompele et al., 2002). The primer sequences used for qRT-PCR are listed in Supplementary Table S1.

Analysis of photosynthetic pigments, abscisic acid and phenylpropanoids

Individual carotenoids were identified and quantified as reported in Tattini et al. (2014). Freeze dried leaf material (200 mg) was extracted with 2 × 5 mL acetone (added with 0.5 g L-1 CaCO

3) and 15 µL of supernatant injected into a Perkin Elmer Flexar liquid chromatograph equipped with a quaternary 200Q/410 pump and an LC 200 diode array detector (DAD) (all from Perkin Elmer, Bradford, CT, USA). Photosynthetic pigments were separated in a 250 × 4.6 mm Agilent Zorbax SB-C18 (5 µm) column operating at 30°C, eluted for 18 min with a linear gradient solvent system, at a flow rate of 1 mL min-1, from 100% CH

3CN/MeOH (95/5 with 0.05% of triethylamine) to 100% MeOH/ethyl acetate (6.8/3.2). Violaxanthin cycle pigments (violaxanthin, antheraxanthin, zeaxanthin, collectively named VAZ), neoxanthin, lutein, and β-carotene, were identified using visible spectral characteristics and retention times. Individual carotenoids and chlorophylls were calibrated using authentic standards from Extrasynthese (Lyon-Nord, Genay, France) and Sigma Aldrich (Sigma-Aldrich Italia, Milan, Italy), respectively.

The concentrations of free-ABA and ABA-glucoside (ABA-GE) were determined in 200 mg of freeze-dried leaf tissue, ground in liquid nitrogen, to which 50 ng of deuterated abscisic acid (d6-ABA) and 50 ng of deuterated ABA-GE (d5-ABA-GE, from National Research Council of Canada, Saskatoon, SK, Canada) were added. The tissue was extracted with 3 mL CH3OH/H2O (50/50 adjusted to pH 2.5 with HCOOH) for 30 min at 4°C, and the supernatant partitioned with 3 × 3 mL n-hexane. The CH3OH/H2O fraction was loaded onto Sep-Pak C18 cartridges (Waters, Massachusetts, USA), 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210

(8)

which were washed with 2 mL pH 2.5 H2O, and then eluted with 1.2 mL ethyl acetate. The eluate, dried under nitrogen, was rinsed with 500 µL pH 2.5 CH3OH/H2O. Aliquots of 3 µL were injected in a LC-ESI-MS/MS equipment, consisting of an Agilent LC1200 chromatograph coupled with an Agilent 6410 triple quadrupole MS detector equipped with an ESI source (all from Agilent Technologies, Santa Clara, CA, USA), operating in negative ion mode. Free-ABA (ABA) and ABA-glucoside (ABA-GE) were separated in an Agilent Poroshell C18 column (3.0 × 100 mm, 2.7 µm), eluted with a linear gradient solvent, at a flow rate of 0.3 mL min-1, from 95% H

2O (with 0.1 % of HCOOH, solvent A) to 100% CH3CN/MeOH (50/50, with 0.1 % of HCOOH, solvent B) over a 30-min run. Quantification of free-ABA and ABA-GE was conducted in multiple reaction mode (MRM) as reported in López-Carbonell et al. (2009).

Identification and quantification of phenylpropanoids, namely caffeic acid derivatives, both colorless flavonoids and anthocyanins was carried out following the protocol reported in Tattini et al. (2014) with few modifications. Freeze dried leaf tissue (300 mg) was extracted with 3 × 6 mL 70% EtOH adjusted to pH 2 with formic acid (HCOOH). The supernatant was partitioned with 3 × 5 mL n-hexane, reduced to dryness under vacuum and rinsed with CH3OH/H2O (50/50, pH 2). Aliquots of 5µL were injected into the Flexar liquid chromatograph unit reported above. Metabolites were separated in a 4.6 × 150 mm Poroshell 120 SB-C18 column (2.7 µm) (Agilent Technologies, Milan, Italy), operating at 30°C, and eluted at a flow rate of 0.6 mL min-1 with a linear gradient solvent system consisting of H

2O (with 3% HCOOH, solvent A), CH3OH (with 3% HCOOH, solvent B) and CH3CN (with 3% HCOOH, solvent C). The chromatographic run lasted 45 min: 0-5 min 74% A, 8% B, 8% C; 5-35 min 56% A, 22% B, 22% C; 35-45 min 20% A, 40% B, 40% C. Individual phenylpropanoids were identified based on their retention times, UV-spectral characteristics and mass-spectrometric data. Quantification of anthocyanins was performed at 530 nm using calibration curves of cyanidin 3-O-glucoside and pelargonidin 3-O-glucoside (Extrasynthese). Quantification of caffeic acid derivatives (caftaric, chicoric and rosmarinic acids) was performed at 330 nm using the calibration curve of rosmarinic acid (Extrasynthese). Quantification of a luteolin derivative was performed at 345 nm using the calibration curve of luteolin 7-O-glucoside (Extrasynthese). HPMS analysis was carried out in a LC-MS-8030 triple quadrupole mass spectrometer operating in the electrospray ionization (ESI) mode and a Nexera HPLC system (all from Shimadzu, Kyoto, Japan). The mass spectrometer operated in negative ion scan mode for caffeic acid derivatives and colorless flavonoid detection, and in positive ion scan mode for anthocyanin detection. Product ion spectra were obtained using Argon as collision gas at a 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241

(9)

pressure of 230 kPa. Identification and quantification of quercetin 3-O-rutinoside was performed by HPLC-MS-MS in MRM, since the flavonol concentration was at detection limits in HPLC-DAD. Flavonoids located on adaxial and abaxial surfaces and in the epidermises of leaves (referred as to Epidermal flavonoids throughout the paper) were optically estimated in vivo using the Multiplex 2 (FORCE-A, Orsay, France) portable fluorimetric sensor, as detailed in Agati et al. (2011). The Chl fluorescence signals under red light excitation (λexc = 625 nm, FRFR) and UV-excitation (λexc = 375 nm, FRFUV) were used to calculate the content of epidermal flavonoids, as:

Epidermal flavonoids = FRFR (FRFUV)-1 (5)

Experimental design and statistics

The experiment was performed using a completely randomized design with 32 plants per cultivar. Diurnal variations in gas exchange, chlorophyll fluorescence kinetics, the concentrations of isoprenoids and phenylpropanoids was performed on four replicate plants, sampled at 08:30, 12:00, 14:30 and 17:30 h. Each replicate for metabolite analysis consisted of three leaves. Five replicate plants per cultivar, sampled between 10:00-14:30 h on two consecutive days, were used for drawing AN/Ci, AN/Cc,

and AN/PPFD curves. Data of diurnal measurements were analyzed with repeated measures ANOVA

with cultivar considered to be the between-subjects effect and daily hour as the within-subjects effect (SPSS v.20, IBM, NY, USA). Parameters derived from AN/Ci and AN/Cc curves were subjected to a

one-way ANOVA. Since the concentrations of individual phenylpropanoids did not significantly vary during the day, data were pooled together and subjected to a one-way ANOVA. Epidermal flavonoids were measured of six replicate plants (two leaves per plant) at different hours of the day and data pooled together prior to statistical analysis. Significant differences among means were estimated at the 5% (P < 0.05) level, using the Tukey’s test. RNA-seq was performed on five plants (two leaves per plant) per cultivar sampled at each individual daily hour and pooled together prior to analysis. Statistical treatment of data was performed with the Z-test (Kal et al., 1999) with LOG2 fold change > ±1 and adjusted P-value < 0.05, to identify genes with significantly different expression between green and red basil morphs (CLC Genomic Workbench, Qiagen, Hilden, Germany). Quantitative Real-Time-PCR analysis was performed in triplicate, and data are reported as mean ± SD.

Results

242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270

(10)

Gas exchange and PSII photochemistry

Photosynthesis did not differ between green (TIG) and red (RR) basil on diurnal basis, since total assimilated CO2, calculated by integrating measurements of AN from 08:30 to 17:30 h (Fig. 1A), was

308 ± 24 (mean ± SD, n = 4) in RR and 288 ± 28 mmol m-2 d-1 in TIG, respectively. TIG had either superior capacity in early morning and late afternoon or lower ability during the central daily hours to assimilate carbon as compared to RR (Fig. 1B). This conforms to the observation that saturation of photosynthesis occurred at 900 and 1200 µmol quanta m-2 s-1 PPFD in TIG and RR, respectively, and photoinhibition of photosynthesis was apparent in TIG at PPFD > 1300 µmol quanta m-2 s-1 (Fig. 1A). Stomatal and mesophyll CO2 conductances also differed between TIG and RR. While stomatal conductance (gs) was higher (+65% or +50% when calculated from AN/Ci curves or from daily in situ

measurements of gas exchange, respectively; Table1 and Fig. 1C), mesophyll conductance to CO2 (gm)

was lower (46%) in RR than in TIG (Table 1). Consistently, apparent (Vc,max(Ci)) was lower, whereas actual (Vc,max (Cc)) carboxylation efficiency was higher in RR than in TIG (Table 1). Stomatal conductance varied to greater extent in TIG than in RR on daily basis (Fig. 1C). Intrinsic water use efficiency (iWUE, AN/gs) was much higher in TIG (on average iWUE = 0.071) than in RR

leaves (on average iWUE = 0.042). The morning- to-midday increase in iWUE was markedly higher in TIG (+72%) than in RR (+18%).

Sunlight irradiance affected PSII photochemistry to a greater degree in TIG than in RR leaves, as revealed by diurnal variations in Chla fluorescence-derived parameters (Fig. 2). Maximal efficiency of PSII photochemistry, Fv/Fm, declined more in TIG than in RR, from early morning to midday (Fig. 2A). Fv/Fm quickly recovered in TIG in the early afternoon, but remained significantly lower than in RR. The actual efficiency of PSII photochemistry, ΦPSII, was on average 20% lower and declined more in TIG than in RR from early morning to late afternoon (Fig. 2B). Mechanisms involved in the thermal dissipation of excess radiant energy, as estimated by NPQ (Fig. 2C), did not operate much in RR, even in leaves suffering from a severe excess of sunlight, in contrast to that observed in TIG.

Genes regulating photomorphogenesis

A set of genes involved in light signaling were differentially expressed in red and green basil (Fig. 3, Supplementary Table S2). Three members of the HOMODOMAIN-leucine zipper (HD-ZIP) family of transcription factors, ATHB1, ATHB2 and ATHB12, were overexpressed in RR, as also observed for the expressions of some relevant members of the basic helix-loop-helix (BHLH) family of TFs 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300

(11)

(bHLH112, bHLH130-Flowering BHLH4 (FBH4); bHLH135-ACTIVATION-TAGGED BRI1

(BRASSINOSTEROID-INSENSITIVE 1)-SUPPRESSOR 1 (ATBS1); bHLH144-SUPPRESSOR OF ACAULIS5 LIKE3 (SACL3). On the contrary, the expressions of ATHB10 (GLABRA 2), BHLH6

(MYC2) and bHLH150 (ATBS1 INTERACTING FACTOR, AIF1) were significantly higher in TIG.

Genes regulating isoprenoid and phenylpropanoid biosynthesis

The expressions of genes involved in early (D-xylulose-5-phosphate synthase, DXS; 1-deoxy-D-xylulose-5-phosphate reductoisomerase, DXR; geranylgeranyl pyrophosphate synthase; GGPS) and late (phytoene synthase, PSY; violaxanthin de-epoxidase, VDE) steps of the MEP pathway were significantly higher in TIG as compared to RR (Fig. 4, Supplementary Table S3). The expression of zeaxanthin epoxidase (ZEP) was instead lower in TIG than in RR. We did not observe significant differences in the expression of genes involved in de-novo ABA biosynthesis, such as neoxanthin synthase (NSY) and 9-cis-epoxycarotenoid dioxygenase (NCED). On the contrary, the expression of a gene coding for a member of the subfamily1 of β-glucosidases (BGL1, which catalyzes the de-glucosylation of ABA-GE, Lee et al., 2006) was higher in TIG than in RR, the reverse being observed for the expression of two genes regulating the oxidative degradation of ABA (ABA 8’-hydroxylase,

CYP707A1).

Green and red basil also largely differed in the expression of genes involved in the biosynthesis of phenylpropanoids (Figure 5, Supplementary Table S4). In detail, the expression levels of early genes of the general phenylpropanoid metabolism (phenylalanine ammonia-lyase, PAL; 4-coumarate-CoA ligase, 4CL), and most genes involved in flavonoid biosynthesis were higher in RR than in TIG. This included early (chalcone synthase CHS; chalcone-flavonone isomerase CHI; flavanone 3-dioxygenase

FHT; flavonoid 3’-monooxygenase, F3’H) and late genes of the flavonoid pathway (dihydroflavonol

4-reducase DFR; leucoanthocianidin dioxygenase, LDOX; anthocyanidin 3-O-glucosyltransferas, 3GT; anthocyanidin 3-O-glucoside 5-O-glucosyltransferase, 5GT; anthocyanidin 3-O-glucoside 6”-acyltransferase 3AT; anthocyanin 5-O-glucoside-6”’-O-malonyltransferase, 5MAT1). On the contrary, the expression of genes involved in flavonol biosynthesis, i.e. FLS, was higher in TIG than in RR.

Photosynthetic pigments, abscisic acid and phenylpropanoids

Green basil leaves had significantly lower (37%) Chltot concentration, higher both Chla to Chlb (+80%) ratio and total carotenoid concentration (Cartot +17%, Table 2) in comparison to red basil leaves. In 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329

(12)

detail, the concentrations of β-carotene (Fig. 6B) and VAZ (Fig. 6, C-E) were higher, whereas the concentrations of lutein (Fig. 3A) and neoxanthin (Fig. 6F) were lower in TIG than in RR. The pool of VAZ was steeply higher (+118%) in TIG than in RR when expressed on chlorophyll basis (Fig. 6G). While the concentration of carotenoids and the VAZ pool did not vary much in RR, the concentration of β-carotene and VAZ, particularly of the de-epoxided forms antheraxanthin and zeaxanthin varied considerably during the day in TIG.

The concentration of abscisic acid also differed between red and green basil. Leaves of TIG had greater concentration of both free-ABA (+103%, Fig. 7A) and ABA-GE (+24% Fig. 7B) than RR leaves. The daily trend of foliar free-ABA concentration was consistent with that of stomatal conductance, with the lowest or the highest concentrations observed in early morning or late afternoon, respectively. Morning to midday increases in free-ABA concentration (+136% in TIG, +61% in RR) and in the ratio of free-ABA to ABA-GE (+84% in TIG, +20% in RR) were markedly higher in TIG than in RR.

Irrespective of genotype, we did not observe diurnal changes in the whole-leaf concentration of phenylpropanoids as well as in the content of epidermal flavonoids. The data of the time course experiments have been pooled together prior to statistical analysis (Table 3). The concentration of colourless phenylpropanoids was 20% higher in TIG than in RR leaves, as the result of major increases in the concentrations of colourless flavonoids (+118%). Similarly, the level of epidermal flavonoids, as monitored by in vivo measurements of absorbance at 375 nm by our chlorophyll fluorescence-based methodology, was higher (+49%) in TIG than in RR. Anthocyanins detected in red basil were mostly cyanidin derivatives, both coumaroyl and malonyl derivatives of cyanidin 3-O-glucoside. Pelargonidin 3-O-derivatives accounted for 9% of the total anthocyanin pool.

Discussion

The data from our study offers a comprehensive picture of the effects of a constitutive epidermal cyanic filter on a suite of molecular and physiological traits that are responsible for the differential acclimation mechanisms to high solar irradiance in red and green individuals.

Blue light attenuation by epidermal anthocyanins evoke shade avoidance response and sustain higher photosynthetic performance in cyanic than in acyanic basil leaves exposed to excess light

330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357

(13)

Attenuation of blue light may simulate the shade avoidance response usually displayed by plants growing in low R/FR light (Keuskamp et al., 2011; Pedmale et al., 2016; Supplementary Fig. S1). Markers of shade avoidance responses, such as the three members of the HD-ZIP transcription factor family, ATHB1, ATHB2 and ATHB12, and the two members of the BHLH family of TFs, bHLH 135 and bHLH 144 (Ciolfi et al., 2013), were highly expressed indeed in RR. ATHB1 and ATHB2, whose expressions decrease in high green light (Zhang and Folta, 2012) or increase in low FR/R light (Steindler et al., 1999; Clitwood et al., 2015) may alter leaf cell fate, causing defects in palisade parenchyma development (Ayoama et al., 1995). Overexpression of ATHB-12 also regulates leaf growth by promoting cell expansion, and reducing the number of palisade cells (Hur et al., 2015).

BHLH 135 (SACL3) and BHLH 144 (ATBS1) genes also negatively affect cell proliferation, and Arabidopsis lines overexpressing ATBS1 display light hyposensitive phenotypes (Castelain et al.,

2012). Consistent with the light hyposensitivity usually displayed by red individuals, the expression of

BHLH 150, which antagonizes the regulatory effects of ATBS1 (Wang et al., 2009) was poorly

expressed indeed in RR. The expression of the BHLH6 gene (MYC2), a key component of cryptochrome-mediated responses (Gupta et al., 2014), was also low in shade-type RR leaves (Chico et

al., 2014). MYC2 has been shown to tightly regulate the whole-plant morphology, as plants

overexpressing MYC2 display shorter internode length and more lateral shoot branches (Yadav et al., 2005; Gupta et al., 2014), as usually observed in sun-adapted green plants, including sweet basil (Tattini et al., 2014). Therefore, data of our study confirm that blue light-induced shade avoidance response is inherently morpho-anatomical in its nature (Keller et al., 2011, Supplementary Fig. S1), and add conclusive evidence to previous observation that RR is almost unresponsive to changes in solar irradiance at the level of the leaf and the whole-plant (Tattini et al., 2014).

The low morpho-anatomical plasticity usually displayed by red leaves in response to light irradiance (Tattini et al., 2014; Landi et al., 2015) is consistent with the significantly lower mesophyll conductance to CO2 observed in RR than in TIG, since gm is usually low in shaded leaves (Flexas et al.,

2008; Peguero-Pina et al., 2016). Red basil partly compensated for the inherently lower gm through a

lower control of gs, since gtot did not differ much between TIG and RR (gtot was 46.9 in RR and 55.5 in

TIG, data not shown). The ability of anthocyanins to absorb more efficiently the green than the blue solar wavelengths, may help to explain the higher gs observed in RR, as green light opposes blue

light-induced stomata opening (Talbott et al., 2002; Kim et al., 2004). A high gm to gs ratio has been

previously observed in thick leaves with densely packed mesophyll (Niinemets et al., 2009; Peguero-358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388

(14)

Pina et al., 2017), to compensate for high stomatal resistance to CO2 diffusion. This may represent an effective physiological adjustment to increase iWUE (Flexas et al., 2016; Peguero-Pina et al., 2017), particularly during the hottest hours of the day. This is consistent with the green-leafed TIG in our study, which displayed markedly greater gm to gs ratio and iWUE as compared to red-leafed RR.

The absorption of solar blue wavelengths by epidermal anthocyanins was responsible for the greater capacity of RR to limit morning-to-midday inhibition in PSII photochemistry as compared to RR, as blue light may severely damage the oxygen-evolving complex in PSII (Takahashi and Murata, 2008). This, when coupled with the lower midday depression in gs, is consistent with the higher

photosynthesis of RR than of TIG under photoinhibitory light conditions. On the contrary, absorption of blue-green light by epidermal anthocyanins makes RR leaves less able than TIG leaves to use radiant energy for CO2 assimilation at relatively low solar irradiance (Gould et al., 2002; Logan et al., 2015). Therefore, our study adds strong support to previous suggestions that epidermal anthocyanins are effective photoprotective pigments (Feild et al., 2001; Hughes, 2011; Landi et al., 2015).

Light attenuation by epidermal anthocyanins affects the biosynthesis and catabolism of MEP-derived products

The higher expression of early (DXS and DXR) and late (GGPS and PSY) genes of the MEP pathway in TIG as compared to RR conforms to the notion that light is one of the primary environmental drivers for isoprenoid biosynthesis (Cordoba et al., 2009; Pokhilko et al., 2015). DXS, which is a rate limiting step for the MEP pathway flow (Wright et al., 2014), is among the most responsive MEP genes to light irradiance (Cordoba et al., 2009). An increased expression of PSY may also trigger an increased supply of MEP-derived precursors via post-transcriptional up-regulation of DXS (Rodriguez-Villalon et al., 2009), further promoting the biosynthesis of late MEP-derived products (Cozzonelli and Pogson, 2010). This is consistent with the significantly higher concentration of carotenoids, particularly of VAZ pigments detected in TIG than in RR leaves. As the ratio of β-carotene to lutein was markedly higher in TIG (1.34) than in RR (0.79), it is suggested that the light-induced activation of the MEP pathway in green basil specifically promoted VAZ biosynthesis and de-epoxidation (Lichtenthaler, 2007), thus sustaining NPQ (Demmig-Adams et al., 2012; Esteban et al., 2015). We hypothesize that zeaxanthin contributed only in part to NPQ in TIG leaves under photoinhibitory light conditions, probably performing additional antioxidant functions in the chloroplast (Havaux et al., 2007; Peguero-Pina et al., 2013; Tattini et al., 2015), since the VAZ pool exceeded the potential binding sites in antenna proteins 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418

(15)

(Esteban et al., 2015). In contrast, the MEP pathway mostly sustained the biosynthesis of chlorophyll and PSII light-harvesting complexes (LHCII) in shade-type RR leaves and, consequently, promoted mostly the biosynthesis of neoxanthin and lutein to optimize light harvesting and photoprotection (Mozzo et al., 2008; Horton, 2012).

In our study, the superior stomatal regulation in TIG correlates well with the levels of foliar free-ABA, as expected for plants that grew under optimal irrigation. The higher levels of foliar ABA observed in TIG than in RR, unlikely originated from a superior de-novo ABA synthesis through the plastidial MEP pathway: The expression of 9-cis epoxycarotenoid oxygenase (NCED) and neoxanthin synthase (NSY) did not vary between the two basil genotypes (Nambara and Marion-Poll, 2005; Galvez-Valdivieso et al., 2009; Finkelstein, 2013). It is conceivable that in leaves challenged by a severe light excess, violaxanthin primarily supports zeaxanthin biosynthesis, rather than constituting a substrate for neoxanthin and/or ABA biosynthesis (Lichtenthaler, 2007; Ruiz-Sola and Rodriguez-Concepcion, 2012). Data of our study indicate that de-glucosylation of ABA-GE (Lee et al., 2006) likely sustained the high accumulation of free-ABA in green basil leaves. The expression of β-glucosidase1 (BGL1) and the ratio of free-ABA to ABA-GE were markedly higher indeed in TIG as compared to RR. Since the expression of genes involved in the first committed (oxidation) step of the ABA catabolic pathway (ABAH, CYP707A1) was largely lower in TIG in comparison to RR, it is suggested that reduced ABA-oxidation also contributed to enhance the levels of free-ABA in green basil leaves. Our suggestion is corroborated by the notion that high light affects ABA degradation more than de-novo ABA synthesis (Kraepiel et al., 1994; Talmann, 2004), and that high blue light represses the expression of CYP707A1 (Barrero et al., 2014).

Light attenuation by epidermal anthocyanins represses the biosynthesis of colourless flavonoids

We have found that early genes of the general phenylpropanoid metabolism (PAL and 4CL), as well as early (CHS, CHI, FHT, F3’H) and late (DFR, LDOX) genes of the flavonoid branch pathway are largely over-expressed in RR. However, the concentrations of colourless flavonoids, particularly of luteolin 7-O and quercetin 3-O-derivatives were higher in TIG than in RR, adding further support to recent findings that blue light is particularly effective in promoting the biosynthesis of flavonols (Siipola et al., 2015; Barnes et al., 2016).

419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447

(16)

The results of our study offer additional evidence of a negative relation between flavonol and anthocyanin biosynthesis (Noda et al., 2004; Lotkowska et al., 2015; Luo et al., 2016). The synthesis of flavonols and anthocyanins originates indeed from the very same intermediate substrates, i.e. dihydro-flavonols (Davies et al., 2003; Yuan et al., 2016), and FLS or DFR synthesize flavonols or anthocyanins, respectively. In this study, the negligible transcript abundance of DFR in TIG likely favored quercetin 3-O-rutinoside biosynthesis as compared to RR (Tian et al. 2015). The higher expression of the HD-ZIP transcription factor GLABRA 2, which exclusively regulates the expression of ‘late’ genes of anthocyanin biosynthesis, suggests that the competition between flavonol and anthocyanin biosynthesis occurs at the level of the C2-C3 bond in the dihydro-flavonoid skeleton (Tian

et al., 2015; Wang et al., 2015; Yuan et al., 2016).

Conclusions

The attenuation of blue light by epidermal anthocyanins profoundly alters the expression of a suite of light responsive genes, and evokes shade avoidance response. This experiment suggests that the shade avoidance syndrome in red individuals primarily involve genes regulating leaf cell development (Supplementary Fig. S1), and reveals the consequential effects on gas exchange performance. We have shown here that the shade avoidance syndrome in red basil is responsible for the lower mesophyll conductance to CO2. The effects of blue-green light attenuation by epidermal anthocyanins on the leaf functional anatomy has received scarce attention, but the issue is of the greatest significance to reconcile contrasting results about the photosynthetic performance of red vs. green individuals growing in full sunlight.

Our study gives new insights on the molecular events that regulate the biosynthesis of MEP-derived products in red and green individuals. We confirm that in green leaves the MEP pathway mostly sustains VAZ synthesis and de-epoxidation to process effectively the excess of radiant energy. Consequently, green leaves activate routes alternative to de-novo ABA synthesis (through the plastidial MEP pathway) to enhance the levels of free-ABA and tightly regulate stomatal conductance under photoinhibitory light conditions. Data of our study are consistent with, and may help to explain previous observations that high light promotes de-glucosylation of ABA-GE and depress ABA oxidative degradation, but does not affect de-novo ABA synthesis.

We offer evidence of a strong competition between flavonol and anthocyanin biosynthesis. This raises the possibility that constitutively red individuals, while better protected against supernumerary 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477

(17)

photons over the visible portion of the solar spectrum, might suffer more than green morphs from an excess of the shortest solar wavelengths, because of a reduced capacity to absorb UV-B radiation and repair UV-B-induced DNA damage. The issue has ecological significance for the successful acclimation of red morphs to high doses of UV-radiation, and open possibilities for new experimentation in respect of the changing UV climate caused by the depletion of stratospheric ozone and rapid climate change.

478 479 480 481 482 483 484

(18)

Supplementary Data

Table S1. Primers used in this research

Table S2. Transcript abundance of light-responsive genes Table S3. Transcript abundance of MEP genes

Table S4. Transcript abundance of phenylpropanoid genes Figure S1. Light microscopy analysis of red and green basil leaf

Acknowledgments

We thank Dr Matthew Haworth for the critical reading of the manuscript, and Dr Cristina Giordano for performing light microscopy analysis. Work in the authors’ lab. was supported by Ministero dell’Istruzione, dell’Università e della Ricerca of Italy: PRIN 2010-2011 projects “PRO-ROOT” and “TreeCity”, and by Uniser Consortium, Pistoia.

References

Abe H, Yamaguchi-Shinozaki K, Urao T, Iwasaki T, Hosokawa D, Shinozaki K. 1997. Role of Arabidopsis MYC and MYB homologs in drought- and abscisic acid-regulated gene expression. The Plant Cell 9, 1859–1868.

Agati G, Cerovic ZG, Pinelli P, Tattini M. 2011. Light-induced accumulation of ortho-dihydroxylated flavonoids as non-destructively monitored by chlorophyll fluorescence excitation techniques. Environmental and Experimental Botany 73, 3-9.

Ayoama T, Dong C-H, Wu Y, Carabelli M, Sessa G, Rubert I, Morelli G, Chua N-H. 1995. Ectopic expression of the Arabidospsis transcriptional activator Athb-1 alters leaf cell fate in tobacco. The Plant Cell 7, 1773-1785.

Barnes PW, Flint SD, Ryel RJ, Tobler MA, Barkely AE, Wargent JJ. 2015. Rediscovering leaf optical properties: New insights into plant acclimation to solar UV radiation. Plant Physiology and Biochemistry 93, 94-100.

Barrero JM, Downie AB, Xu Q, Gubler F. 2014. A role for barley CRYPTOCHROME1 in light regulation of grain dormancy and germination. The Plant Cell 26, 1094-1104.

485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510

(19)

Bilger W, Björkman O. 1990. Role of the xanthophyll cycle in photoprotection elucidated by measurements of light-induced absorbance changes, fluorescence and photosynthesis in leaves of Hedera canariensis. Photosynthesis Research 25, 173–185.

Castelain M, Le Hir R, Bellini C. 2012. The non-DNA-binding bHLH transcription factor PRE3/bHLH135/ATBS1/TMO7 is involved in the regulation of light signaling pathway in Arabidopsis. Physiologia Plantarum 145, 450-460.

Chico JM, Fernández-Barbero G, Chini A, Fernández-Calvo P, Díez-Díaz M, Solano R. 2014. Repression of jasmonate-dependent defenses by shade involves differential regulation of protein stability of MYC transcription factors and their JAZ repressors in Arabidopsis. The Plant Cell 26, 1967-1980.

Ciolfi A, Sessa G, Sassi M, Possenti M, Salvucci S, Carabelli M, Morelli G, Rubert I. 2013. Dynamics of shade-avoidance response in Arabidopsis. Plant Physiology 163, 331-353.

Clitwood DH, Kumar R, Ranjan A, et al. 2015. Light-induced indeterminacy alters shade-avoiding tomato leaf morphology. Plant Physiology 169, 230-247

Cordoba E, Salmi M, León P. 2009. Unraveling the regulatory mechanisms that modulate the MEP pathway in higher plants. Journal of Experimental Botany 60, 2933-2943.

Cozzonelli CI, Pogson BJ. 2010. Source to sink: regulation of carotenoid biosynthesis in plants. Trends in Plant Science 15, 266-274.

Davies KM, Schwinn KE, Deroles SC, Manson DG, Lewis DH, Bloor SJ, Bradley JM. 2003. Enhancing anthocyanin production by altering competition for substrate between flavonol synthase and dihydroflavonol 4-reductase. Euphytica 131, 259–268.

Demmig-Adams B, Cohu CM, Müller O, Adams WW III. 2012. Modulation of photosynthetic energy conversion efficiency in nature: from seconds to seasons. Photosynthesis Research 113, 75-88.

Esteban R, Barrutia G, Artetxe U, Fernández-Marín B, Hernandez A, Garciá-Plazaola JI. 2015. Internal and external factors affecting photosynthetic pigment composition in plants: a meta-analytic approach. New Phytologist 206, 268-280.

Feild TS, Lee DW, Holbrook NM. 2001. Why leaves turn red in autumn. The role of anthocyanins in senescing leaves of red-osier dogwood. Plant Physiology 127: 566–574.

Finkelstein R. 2013. Abscisic acid synthesis and response. The Arabidopsis Book 11, e0166. 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538

(20)

Fini A, Loreto F, Tattini M, Giordano C, Ferrini F, Brunetti C, Centritto M. 2016. Mesophyll conductance plays a central role in leaf functioning of Oleaceae species exposed to contrasting sunlight irradiance. Physiologia Plantarum 157, 54-68.

Flexas J, Diaz-Espejo A, Conessa MA et al. 2016. Mesophyll conductance to CO2 and Rubisco as targets for improving intrinsic water use efficiency in C3 plants. Plant Cell and Environment 39, 965-982.

Flexas J, Ribas-Carbó M, Díaz-Espejo A, Galmés J, Medrano H. 2008. Mesophyll conductance to CO2: current knowledge and future prospects. Plant Cell and Environment 31, 602-621.

Galmés J, Flexas J, Keys AJ, Cifre J, Mitchell RAC, Madgwick PJ, Haslam RP, Medrano H, Parry MAJ. 2005. Rubisco specificity factor tends to be larger in plant species from dryer habitats and in species with persistent leaves. Plant Cell and Environment 28, 571-579.

Galvez-Valdivieso G, Fryer MJ, Lawson T et al. 2009. The high light response in Arabidopsis involves ABA signaling between vascular and bundle sheath cells. The Plant Cell 21, 2143-2162.

Genty B, Briantais JM, Baker NR. 1989. The relationship between the quantum yield of photosynthetic electron transport and quenching of chlorophyll fluorescence. Biochimica Biophysica Acta 990, 87–92.

Gould KS. 2004. Nature's Swiss army knife: The diverse protective roles of anthocyanins in leaves. Journal of Biomedicine and Biotechnology 2004, 314-320.

Gould KS, McKelvie J, Markham KR. 2002. Do anthocyanins function as antioxidants in leaves? Imaging of H2O2 in red and green leaves after mechanical injury. Plant Cell and Environment 25, 1261-1269.

Gould KS, Quinn BD. 1999. Do anthocyanins protect leaves of New Zealand native species from UV-B? New Zealand Journal of Botany 37, 175-178.

Gupta N, Prasad VBR, Chattopadhyay. 2014. LeMYC2 acts as a negative regulator of blue light mediated photomorphogenic growth, and promotes the growth of adult tomato plants. BMC Plant Biology 14, 38. Hada H, Hidema J, Maekawa M, Kumagai T. 2003. Higher amounts of anthocyanins and UV-absorbing

compounds effectively lowered CPD photorepair in purple rice (Oryza sativa L.). Plant Cell and Environment 26, 1691-1701.

Harley PC, Loreto F, Di Marco G, Sharkey TD. 1992. Theoretical considerations when estimating the mesophyll conductance to CO2 flux by analysis of the response of photosynthesis to CO2. Plant Physiology 98, 1429-1436. 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566

(21)

Havaux M, Dall’ Osto L, Bassi R. 2007. Zeaxanthin has enhanced antioxidant capacity with respect to all other xanthophylls in Arabidopsis leaves and functions independent of binding to PSII antennae. Plant Physiology 145, 1506– 1520.

Horton P. 2012. Optimization of light harvesting and photoprotection: molecular mechanisms and physiological consequences. Philosophical Transactions of the Royal Society B 367, 3455-3465.

Hughes NM. 2011. Winter leaf reddening in ‘evergreen’ species. New Phytologist 190, 573-581.

Hughes NM, Neufeld HS, Burkey KO. 2005. Functional role of anthocyanins in high-light winter leaves of the evergreen herb Galax urceolata. New Phytologist 168, 575-587.

Hughes NM, Smith WK. 2007. Attenuation of incident light in Galax urceolata (Diapensiaceae): concerted action of adaxial and abaxial anthocyanic layer on photoprotection. American Journal of Botany 94, 784-790.

Hughes NM, Smith WK, Gould KS. 2010. Red (anthocyanic) leaf margins do not correspond to increased phenolic content in New Zealand Veronica spp. Annals of Botany 105, 647-654.

Hur Y-S, Um J-H, Kim S, et al. 2015. Arabidopsis thaliana homeobox 12 (ATHB12), a homeodomain-leucine zipper protein, regulates leaf growth by promoting cell expansion and endoreduplication. New Phytologist 205, 316-328.

Jordheim M, Calcott K, Gould KS, Davies KM, Schwinn KE, and Andwersen OM. 2016. High concentrations of aromatic acylated anthocyanins found in cauline hairs in Plectranthus ciliatus. Phytochemistry 128, 27-34.

Kal A.J., van Zonneveld A.J., Benes V. et al. 1999. Dynamics of gene expression revealed by comparison of serial analysis of gene expression transcript profiles from yeast grown on two different carbon sources. Molecular Biology of the Cell 10, 1859-1872.

Keller MM, Jaillais Y, Pedmale UV, Moreno JE, Chory J, Ballaré CL. 2011. Cryptochrome 1 and phytochrome B control shade-avoidance responses in Arabidopsis via partially independent hormonal cascades. The Plant Journal 67, 195-207.

Keuskamp DH, Sasidharan R, Vos I, Peeters AJM, Voesenek LACJ, Pierik M. 2011. Blue-light-mediated shade avoidance requires combined auxin and brassinosteroid action in Arabidopsis seedlings. The Plant Journal 67, 208-217.

Kim H-H, Goins G, Wheeler R, Sager J. 2004. Stomatal conductance of lettuce grown under or exposed to different light quality. Annals of Botany 94, 691-697.

567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596

(22)

Kovinich N., Kayaja G., Chanoca A., Riedl K., Otegui M.S. & Grotewold E. 2014. Not all anthocyanins are born equal: distinct patterns induced by stress in Arabidopsis. Planta 240, 931-940.

Kraepiel Y. Rousselen P, Sotta B, Kerhoas L, Einhorn J, Caboche M, Miginiac E. 1994. Analysis of phytochrome- and ABA-deficient mutants suggests that ABA degradation is controlled by light in Nicotiana plumbaginifolia. The Plant Journal 6, 665-672.

Kyparissis A, Grammatikopoulos G, Manetas Y. 2007. Leaf morphological and physiological adjustments to the spectrally selective shade imposed by anthocyanins in Prunus cerasifera. Tree Physiology 27, 849-857.

Kytridis V-P, Karageorgou P, Levizou E, Manetas Y. 2008. Intra-specific variation in transient accumulation of leaf anthocyanins in Cistus creticus during winter: Evidence that anthocyanins may compensate for an inherent photosynthetic and photoprotective inferiority of the red-leaf phenotype. Journal of Plant Physiology 165, 952-959.

Kytridis V-P, Manetas Y. 2006. Mesophyll versus epidermal anthocyanins as potential in vivo antioxidants: Evidence linking the putative antioxidant role to the proximity of oxy-radical source. Journal of Experimental Botany 57, 2203-2210.

Lan J-X, Li A-L, Chen C-X. 2011. Effect of transient accumulation of anthocyanin on leaf development and photoprotection of Fagopyrum dibotrys mutant. Biologia Plantarum 54, 766-770.

Landi M, Guidi L, Pardossi A, Tattini M, Gould KS. 2014. Photoprotection by foliar anthocyanins mitigates effects of boron toxicity in sweet basil (Ocimum basilicum). Planta 240, 941-953.

Landi M, Tattini M, Gould KS. 2015. Multiple functional roles of anthocyanins in plan-environment interactions. Environmental and Experimental Botany 119, 54-62.

Laisk A, Loreto F. 1996. Determining photosynthetic parameters from leaf CO2 exchange and chlorophyll fluorescence. Plant Physiology 110, 903-912.

Lee KH, Piao HL, Kim H-Y, Choi SM, Jiang F, Hartung W, Hwang I, Kwak JM, Lee IJ, Hwang I. 2006. Activation of glucosidase via stress-induced polymerization rapidly increases active pools of abscisic acid. Cell 126, 1109-1120.

Liakopoulos G, Nikopoulos D, Klouvatou A, Vekkos KA, Manetas Y, Karabourniotis G. 2006. The photoprotective role of epidermal anthocyanins and surface pubescence in young leaves of grapevine (Vitis vinifera). Annals of Botany 98, 257-265.

597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625

(23)

Lichtenthaler HK. 2007. Biosynthesis, accumulation and emission of carotenoids, α-tocopherol, plastoquinone, and isoprene in leaves under high photosynthetic irradiance. Photosynthesis Research 92, 163-179. Logan BA, Stafstrom WC, Walsh MJ, Reblin JS, Gould KS. 2015. Examining the photoprotection

hypothesis for adaxial foliar anthocyanin accumulation by revisiting comparisons of green and red-leafed varieties of coleus (Solenostemon scutellarioides). Photosynthesis Research 124, 267-274.

López-Carbonell M, Gabasa M, Jáurequi O. 2009. Enhanced determination of abscisic acid (ABA) and abscisic acid glucose ester (ABA-GE) in Cistus albidus plants by liquid chromatography-mass spectrometry in tandem mode. Plant Physiology and Biochemistry 47, 256-261.

Loreto F, Harley PC, Di Marco G, Sharkey TD. 1992. Estimation of mesophyll conductance to CO2 flux by three different methods. Plant Physiology 98, 1437-1443.

Lotkowska ME, Tohge T, Fernie AR, Xue G-P, Balazadeh S, Mueller-Roeber B. 2015. The Arabidopsis transcription factor MYB112 promotes anthocyanin formation during salinity and under high light stress. Plant Physiology 169, 1862-1880.

Luo P, Guogui N, Wang Z, Shen Y, Huanan J, Li P, Huang S, Zhao J, Manzhu B. 2016. Disequilibrium of flavonol synthase and dihydroflavonol-4-reductase expression associated tightly to white vs. red color flower formation in plants. Frontiers in Plant Science 6, 1257.

Manetas Y, Petropoulou Y, Psara GK, Drinia A. 2003. Exposed red (anthocyanic) leaves of Quercus coccifera display shade characteristics. Functional Plant Biology 30, 265-270.

Merzlyak MN, Chivkunova OB, Solovchenko AE, Naqvi KR. 2008. Light absorption by anthocyanins in juvenile, stressed, and senescing leaves. Journal of Experimental Botany 59, 3903-3911.

Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B. 2008. Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nature Methods 5, 621-628.

Mozzo M, Dall' Osto L, Hienerwadel R, Bassi R, Croce R. 2008. Photoprotection in the antenna complexes of

Photosystem II. Role of individual xanthophylls in chlorophyll triplet quenching. The Journal of Biological Chemistry 283, 6184-6192.

Nambara E, Marion-Poll A. 2005. Abscisic acid synthesis and catabolism. Annual Review of Plant Biology 56, 165-1685.

Neill SO, Gould KS, Kilmartin PA, Mitchell KA, Markham KR. 2002. Antioxidant capacities of green and cyanic leaves in the sun species Quintinia serrta. Functional Plant Biology 29, 1437-1443.

626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654

(24)

Niinemets Ü, Díaz-Espejo A, Flexas J, Galmés J, Warren CR. 2009. Role of mesophyll diffusion conductance in constraining potential photosynthetic productivity in the field. Journal of Experimental Botany 60, 2249-2270.

Nikoforou C, Nikopoulos D, Manetas Y. 2011. The winter-red-leaf syndrome in Pistacia lentiscus: Evidence that the anthocyanic phenotype suffers from nitrogen deficiency, low carboxylation efficiency and high risk of photoinhibition. Journal of Plant Physiology 168, 2184-2187.

Noda N, Kanno Y, Kato N, Kazuma K, Suzuki M. 2004. Regulation of gene expression in flavonol and anthocyanin biosynthesis during petal development in lisianthus (Eustona grandiflorum). Physiologia Plantarum 122, 305-313.

Pedmale UV, Carol Huang S-s, Zander M, et al. 2016. Cryptochromes interact directly with PIFs to control plant growth under limiting light. Cell 164, 233-245.

Peguero-Pina JJ, Gil-Pelegrín E, Morales F. 2013. Three pools of zeaxanthin in Quercus coccifera leaves during light transitions with different roles in rapidly reversible photoprotective energy dissipation and photoprotection. Journal of Experimental Botany 64, 1649-1661.

Peguero-Pina JJ, Sancho-Knapik D, Flexas J, Galmés J, Niinemets U, Gil-Pelegrin E. 2016. Light acclimation of photosynthesis in two closely related firs (Abies pinsapo Boiss. and Abies alaba Mill.): the role of leaf anatomy and mesophyll conductance to CO2. Tree Physiology 36, 300-310.

Peguero-Pina JJ, Sisó S, Flexas J, Galmés J, Garciá-Nogales A, Niinemets U, Sanch-Knapik D, Saz MÁ, Gil-Pelegrin E. 2017. Cell-level anatomical characteristics explain high mesophyll conductance and photosynthetic capacity in sclerophyllous Mediterranean oaks. New Phytologist DOI: 10.1111/nph.14406.

Pokhilko A. Bou-Torrent J, Pulido P, Rodríguez-Concepción M, Ebenhöh O. 2015. Mathematical modelling of the diurnal regulation of the MEP pathway in Arabidopsis. New Phytologist 206, 1075-1085.

Rodriguez-Villalón A, Gas E, Rodríguez-Concepcíon M. 2009. Phytoene synthase activity controls the biosynthesis of carotenoids and the supply of their metabolic precursors in dark-grown Arabidopsis seedlings. Plant Journal 60, 424-435.

Ruiz-Sola MA, Rodríguez-Concepcíon M. 2012. Carotenoid biosynthesis in Arabidopsis: a colorful pathway. The Arabidopsis Book 10, e0158.

655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682

(25)

Schreiber U, Schliva U, Bilger B. 1986. Continuous recording of photochemical and non-photochemical chlorophyll fluorescence quenching with a new type of modulation fluorometer. Photosynthesis Research 10, 51–62.

Sharkey TD, Bernacchi CJ, Farquhar GD, Singsaas EL. 2007. Fitting photosynthetic carbon dioxide response curves for C3 leaves. Plant Cell & Environment 30, 1035-1040.

Siipola SM, Kotilainen T. Sipari N, Morales LO, Lindfors AV, Robson TM, Aphalo P. 2015. Epidermal UV-A absorbance and whole-leaf flavonoid composition in pea respond more to solar blue light than to solar UV radiation. Plant Cell & Environment 38, 941-952.

Steindler C, Matteucci A, Sessa G, Weimar T, Ohgishi M, Ayoama T, Morelli G, Ruberti I. 1999. Shade avoidance responses are mediated by the ATHB-2 HD-zip protein, a negative regulator of gene expression. Development 126, 4235-4245.

Steyn WJ, Wand SJE, Holcroft DM, Jacobs G. 2002. Anthocyanins in vegetative tissues: a proposed unified function in photoprotection. New Phytologist 155, 349-361.

Stich K, Eidenberger T, Wurst F. 1992. Flavonol synthase activity and the regulation of flavonol and anthocyanin biosynthesis during flower development in Dianthus caryophyllus L. (carnation). Zeitschrift für Naturforschung 47c, 553-560.

How do environmental stresses accelerate photoinhibition? Trends in Plant Science 13, 178–182.

Talbott LD, Nikolova G, Ortiz A, Shmayevich IJ Zeiger E. 2002. Green light reversal of blue light stomatal opening is found in a diversity of plant species. American Journal of Botany 89, 366-368.

Tallman G. 2004. Are diurnal patterns of stomatal movement the result of altering metabolism of endogenous guard cell ABA and accumulation of ABA delivered from the apoplast around guard cells by transpiration? Journal of Experimental Botany 55, 1963-1976.

Tattini M, Galardi C, Pinelli P, Massai R, Remorini D, Agati G. 2004. Differential accumulation of flavonoids and hydroxycinnamates in Ligustrum vulgare leaves under excess light and drought stress. New Phytologist 163, 547-561.

Tattini M Landi M, Brunetti C, Remorini D, Gould KS, Guidi L. 2014. Epidermal coumaroyl anthocyanins protect sweet basil against excess light stress: multiple consequences of light attenuation. Physiologia Plantarum 152, 585-598. 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710

(26)

Tattini M, Loreto F, Fini A, Guidi L, Brunetti C, Velikova V, Gori A, Ferrini F. 2015. Isoprenoids and phenylpropanoids are part of the antioxidant defense orchestrated daily by drought-stressed Platanus x acerifolia plants during Mediterranean summers. New Phytologist 37, 1950-1964.

Terashima I, Hanba YT, Tholen D, Niinemets Ü. 2011. Leaf functional anatomy in relation to photosynthesis. Plant Physiology 155, 108–116.

Tian JI, Han Z-Y, Zhang J, Hu Y, Song T, Yao Y. 2015. The balance of expression of dihydroflavonol 4-reductase and flavonol synthase regulates flavonoid biosynthesis and red foliage coloration in crabapples. Scientific Reports 5, 12228.

Torre S, Tattini M, Brunetti C, Guidi L, Gori A, Marzano C, Landi M, Sebastiani F. 2016. De novo assembly and comparative transcriptome analyses of red and green morphs of sweet basil grown in full sunlight. PLoS ONE 11, e0160370.

Vandesompele J, De Preter K, Pattin F, Poppe B, Van Roy N, De Paepe A, Speleman F. 2002. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome Biology 3, 7

Wang H, Zhu Y, Fujioka S, Asami T, Li J, Li J. 2009. Regulation of Arabidopsis brassinosteroid signaling by atypical basic helix-loop-helix proteins. The Plant Cell 21, 3781-3791.

Wang X, Wang X, Hu Q, Dai X, Tian H, Zheng K, Wang X, Mao T, Chen GJ, Wang S. 2015.

Characterization of an activation-tagged mutant uncovers a role of GLABRA2 in anthocyanin biosynthesis in Arabidopsis. The Plant Journal 83, 300-311.

Wright LP, Rohwer JM, Ghirardo A, Hammerbacher A, Ortiz-Alcade M, Raguschke B, Schnitzler J-P, Gershenzon J, Phillips MA. 2014. Deoxyxylulose 5-phosphate synthase controls flux through the methylerythritol 4-phosphate pathway in Arabidopsis. Plant Physiology 165, 1488-1504.

Yadav V, Mallappa C, Gangappa SN, Bhatia S, Chattopadhyay S. 2005. A basic helix-loop-helix transcription factor in Arabidopsis, MYC2, acts as a repressor of blue light–mediated photomorphogenic growth. The Plant Cell 17, 1953-1966.

Yuan Y-W, Rebocho AB, Sagawa JM, Stanley LE, Bradshaw Jr. HD. 2016. Competition between anthocyanin and flavonol biosynthesis produces spatial pattern variation of floral pigments between Mimulus species. Proceedings of the National Academy of Science USA 113, 2448-2453.

711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738

(27)

Zeliou K, Manetas Y, Petropoulou Y. 2009. Transient winter leaf reddening in Cistus creticus characterizes weak (stress-sensitive) individuals, yet anthocyanins cannot alleviate the adverse effects on photosynthesis. Journal of Experimental Botany 60, 3031-3042.

Zhang T, Folta KM. 2012. Green light signaling and adaptive response. Plant Signaling & Behavior 7, 1-4 739

740 741 742 743

(28)

Table 1. Carboxylation efficiency (Vc,max) based on intercellular (Ci) or

chloroplast (Cc) CO2 concentration, stomatal (gs) and mesophyll

conductance (gm) to CO2, intercellular (Ci) and chloroplast (Cc) CO2

concentration in red (Red Rubin) and green (Tigullio) basil leaves. Measurements were conducted on four-week-old leaves developed at full sunlight. Data are the mean ± SD (n = 5), and those in a row not accompanied by the same letter differ significantly at P < 0.05, using Tukey’s test.

Parameter Red Rubin Tigullio

Vc,max (Ci) (µmol m-2 s-1) 63.2 ± 5.4 b 77.9 ± 7.0 a Vc,max (Cc) (µmol m-2 s-1) 153.9 ± 11.1 a 122.7 ± 12.5 b gs (mmol m-2 s-1) 178.5 ± 11.6 a 104.5 ± 9.2 b gm (mmol m-2 s-1) 62.8 ± 8.8 b 115.8 ± 10.6 a Ci (µmol mol-1) 263.5 ± 11.0 a 218.7 ± 17.4 b Cc (µmol mol-1) 67.6 ± 5.8 b 111.3 ± 10.3 a 744 745 746 747 748 749 750 751 752

(29)

Table 2. The concentrations (µmol g-1 DW) of total chlorophyll (Chltot) and total carotenoids (Cartot), and the ratio of Chla to Chl b (Chla/Chlb) in red (Red Rubin, RR) and green (Tigullio, TIG) leaves at the end of a four-week period of exposure to full solar irradiance. Leaves were sampled at 08:30, 12:00, 14:30, and 17:30 h, and data have been pooled together prior to statistical analysis. Data are the mean ± SD (n = 16) and those in a row not accompanied by the same letter differ significantly at the 5% level, using Tukey’s test.

Red Rubin Tigullio

Chltot 5.84 ± 0.83 a 3.66 ± 0.52 b Chla/Chlb 2.08 ± 0.12 b 2.94 ± 0.14 a Cartot 1.14 ± 0.06 b 1.32 ± 0.08 a 753 754 755 756 757 758 759 760 761 762 763

Riferimenti

Documenti correlati

We present three methods to quantify the wavelength-dependent diffuse transmittance, relying on (1) measurement of radiance exiting the phantom by a detector far from the

● Lo stato degli studi sulla Romagna estense ​ , in “Atti e memorie della Deputazione di storia patria per le antiche provincie modenesi”, s. ● La legazione di Ferrara

Il complesso più attivo risulta essere quello con il legante Dit e non si osserva una variazione di selettività nei confronti delle linee tumorali al variare

At the time T b when bankruptcy occurs, we assume that the borrower pays a fixed price B, while lenders recover a fraction θ(x(T b )) ∈ [0, 1] of their outstanding capital..

[A stampa in Incastella mento, popolamento e signoria rurale tra Piemonte meridionale e Liguria. 95-102 © dell’autore – Distribuito in formato digitale da “Reti

Col processo d’interramento subìto dal bacino di Porto Pisano nel Cinquecento, e col favore della politica medicea, Livorno diventa il porto principale della Toscana (e

In a multivariable Cox model, allocation to tight BP control reduced the risk of cardiovascular events to a similar extent in patients with or without overt cardiovascular disease

These two members, named CHHip (CHH Immune-related Procambarus) and CHHop (CHH homologous Procambarus) share the typical features of the CHH (the presence of a