A 2-year multisite study of viticultural and environmental ...

14
A 2-year multisite study of viticultural and environmental factors affecting rotundone concentration in Duras red wine Olivier Geffroy 1,2* , Juliette Descôtes 1 , Cécile Levasseur-Garcia 2 , Christian Debord 3 , Jean-Philippe Denux 2 and Thierry Dufourcq 4 1 Institut Français de la Vigne et du Vin Pôle Sud-Ouest, V’innopôle, BP 22 Brames-Aïgues, 81310 Lisle-sur-Tarn, France 2 Université de Toulouse, INP – École d’Ingénieurs de Purpan, 31076 Toulouse Cedex 3, France 3 Institut Français de la Vigne et du Vin Pôle Bordeaux Aquitaine, 39 rue Michel Montaigne, 33290 Blanquefort, France 4 Institut Français de la Vigne et Vin Pôle Sud-Ouest, Domaine de Mons, 32100 Caussens, France Corresponding author : [email protected] Aim: A study was carried out in 2013 and 2014 to determine the key environmental and viticultural variables affecting the concentration of rotundone, the black pepper aroma compound, in Vitis vinifera L. cv. Duras red wines at 10 different vineyard blocks. Methods and results: For each block, data for fruit quality attributes, as well as climatic and agronomical variables, were collected. Rotundone was quantified in wines prepared by microvinification techniques (in a 1-L Erlenmeyer flask). Rotundone concentration varied across blocks from 63 ng/L to 239 ng/L in 2013 and from 25 ng/L to 115 ng/L in 2014. Three separate partial least squares regression models were constructed to predict rotundone concentration in wines in 2013, in 2014, and in both vintages. Gluconic acid, a secondary metabolite of Botrytis cinerea, had a substantial contribution to the 2013 and multivintage models, with a negative regression coefficient with rotundone concentration. Other predictors were associated with abiotic factors such as cumulative rainfall, thermal index, hours of sunshine and mean daily irradiation. Conclusions: Our results indicate that mesoscale climatic variables are the key factors determining rotundone concentration, and also suggest that Botrytis cinerea may be involved in rotundone degradation. Significance and impact of the study: Our findings may assist grape growers producing Duras red wines to select specific vineyard blocks with the aim of producing wines with a desired rotundone concentration. They also open up new fields of investigation into mechanisms involved in possible rotundone degradation by Botrytis cinerea. abiotic factors, Botrytis cinerea, mesoscale climate, partial least squares regression model, rotundone ABSTRACT KEYWORDS Received: 29 October 2018 y Accepted: 23 Juin 2019 y Published: 24 July 2019 DOI: 10.20870/oeno-one.2019.53.3.2341 VINE AND WINE OPEN ACCESS JOURNAL 457 OENO One 2019, 3, 457-470 © 2019 International Viticulture and Enology Society - IVES Additional tables can be downloaded from https://oeno-one.eu/article/view/2341

Transcript of A 2-year multisite study of viticultural and environmental ...

Page 1: A 2-year multisite study of viticultural and environmental ...

A 2-year multisite study of viticultural and environmental factors affecting rotundone concentrationin Duras red wine

Olivier Geffroy1,2*, Juliette Descôtes1, Cécile Levasseur-Garcia2, Christian Debord3, Jean-Philippe Denux2 and Thierry Dufourcq4

1Institut Français de la Vigne et du Vin Pôle Sud-Ouest, V’innopôle, BP 22Brames-Aïgues, 81310 Lisle-sur-Tarn, France2Université de Toulouse, INP – École d’Ingénieurs de Purpan, 31076 Toulouse Cedex 3, France3Institut Français de la Vigne et du Vin Pôle Bordeaux Aquitaine, 39 rue Michel Montaigne, 33290 Blanquefort, France4Institut Français de la Vigne et Vin Pôle Sud-Ouest, Domaine de Mons, 32100 Caussens, France

Corresponding author : [email protected]

Aim: A study was carried out in 2013 and 2014 to determine the key environmental and viticultural variablesaffecting the concentration of rotundone, the black pepper aroma compound, in Vitis vinifera L. cv. Duras red winesat 10 different vineyard blocks.Methods and results: For each block, data for fruit quality attributes, as well as climatic and agronomical variables,were collected. Rotundone was quantified in wines prepared by microvinification techniques (in a 1-L Erlenmeyerflask). Rotundone concentration varied across blocks from 63 ng/L to 239 ng/L in 2013 and from 25 ng/L to115 ng/L in 2014. Three separate partial least squares regression models were constructed to predict rotundoneconcentration in wines in 2013, in 2014, and in both vintages. Gluconic acid, a secondary metabolite of Botrytiscinerea, had a substantial contribution to the 2013 and multivintage models, with a negative regression coefficientwith rotundone concentration. Other predictors were associated with abiotic factors such as cumulative rainfall,thermal index, hours of sunshine and mean daily irradiation.Conclusions: Our results indicate that mesoscale climatic variables are the key factors determining rotundoneconcentration, and also suggest that Botrytis cinerea may be involved in rotundone degradation.Significance and impact of the study: Our findings may assist grape growers producing Duras red wines to selectspecific vineyard blocks with the aim of producing wines with a desired rotundone concentration. They also open upnew fields of investigation into mechanisms involved in possible rotundone degradation by Botrytis cinerea.

abiotic factors, Botrytis cinerea, mesoscale climate, partial least squares regression model, rotundone

AB S TRACT

KEYWORD S

Received: 29 October 2018 y Accepted: 23 Juin 2019 y Published: 24 July 2019DOI: 10.20870/oeno-one.2019.53.3.2341

V I N E A N D W I N EOPEN ACCESS JOURNAL

457OENO One 2019, 3, 457-470 © 2019 International Viticulture and Enology Society - IVES

Additional tables can be downloaded from https://oeno-one.eu/article/view/2341

Page 2: A 2-year multisite study of viticultural and environmental ...

INTRODUCTION

Wine is a complex matrix that contains morethan 800 aroma compounds (Robinson et al.,2014). Most of the contributors to the varietalaroma of white wines, such as the monoterpenolsresponsible for floral notes in Muscat varieties(Williams et al., 1981) and the varietal thiolsimparting grapefruit and tropical fruit notes inSauvignon blanc wines (Tominaga et al., 1998),have been widely studied. Knowledge ofodorants accounting for the varietal character ofred wines, especially free aroma compoundsdirectly extracted from grapes without beingreleased from a precursor, had been limited togreen and undesirable aroma methoxypyrazinecompounds (Sala et al., 2000) until the discoveryof the sesquiterpene rotundone, which isresponsible for the black pepper aroma (Wood etal., 2008). Specific anosmia to this potentodorant, which has a detection threshold of16 ng/L in red wine, has been reported (Wood etal., 2008). In most cases, rotundone is perceivedby consumers either positively or in a neutralmanner (Geffroy et al., 2018a), and those whoappreciate peppery wines are generally wineconnoisseurs who pay more than the averageconsumer for a bottle of wine (Geffroy et al.,2016a).

Aside from Syrah/Shiraz wines, rotundone hasso far been identified at very high concentrations(above 100 ng/L) in red wines made fromSchioppettino (Caputi et al., 2011), Duras,Pineau d’Aunis (Geffroy et al., 2014), Gamay(Geffroy et al., 2016a) and Maturana tinta(Cullere et al., 2016). Differences in rotundoneconcentration have been identified betweencertified clones of Duras, and the grapevinedefense response to powdery mildew mightresult in enhanced rotundone production(Geffroy et al., 2015a). Zhang et al. (2016a)have shown that rotundone might not only beproduced as a response to herbivore attacks andproposed the existence of other biosyntheticpathways. Rotundone does not appear to betranslocated from vegetative tissues (Geffroy etal., 2016b; Zhang et al., 2016a), and inaccordance with the hypothesis of local synthesisin grape berries, crop load reduction throughgrape thinning has no effect on rotundoneconcentration in wine (Geffroy et al., 2014). Thecompound is produced by simple aerial orenzymatic oxidation of α-guaiene (Huang et al.,2014; Takase et al., 2016). Although rotundonehas been identified at early stages of

development in vegetative tissues (Zhang et al.,2016a), it is known to accumulate late in theseason in berries, with the concentrationincreasing from veraison to harvest (Caputi etal., 2011; Geffroy et al., 2014).

Cool and wet vintages promote the production ofpeppery red wines (Caputi et al., 2011). Soilcharacteristics (Scarlett et al., 2014), bunch zoneair and bunch surface temperatures (Zhang et al.,2015a), vineyard water balance (Zhang et al.,2015b), and vine water status over the veraison-to-harvest period (Geffroy et al., 2014; Geffroyet al., 2015b) have been identified as keyvariables that explain differences in rotundoneconcentration between vintages and within asingle vineyard block. Thus, the concentration ofrotundone has been reported to be enhanced bythe water supply through irrigation (Geffroy etal., 2014; Geffroy et al., 2016b). Conflictingresults have been reported regarding the effect ofbunch exposure, because leaf removal canreduce (Geffroy et al., 2014), increase, or haveno effect on rotundone accumulation (Homich etal., 2017; Geffroy et al., 2018b).

The hierarchy between viticultural andenvironmental factors driving rotundoneproduction, especially temperature and waterstatus, remains unclear. Although several studieshave been conducted on rotundone at an intra-block scale (Scarlett et al., 2014; Geffroy et al.,2015b; Zhang et al., 2015a; Bramley et al.,2017), no data are presently available for studiescarried out at a mesoscale. To study these aspectsand to identify possible correlations betweenrotundone and vine and climatic variables, a trialwas carried out in 2013 and 2014 on 10 Vitisvinifera L. cv. Duras vineyard blocks over theGaillac protected designation of origin in thesouth-west of France. For each plot, data for thevariables were collected or calculated mostly atveraison or between veraison and harvest, due tothe late accumulation of rotundone in berries.

MATERIALS AND METHODS

1. Vineyard blocks and vine panels

The characteristics of the 10 commercial drylandvineyard blocks used for this study are shown inTable 1. The plots were selected to capture awide range of soil and climatic conditions, withthe aim of maximizing the range of rotundoneconcentrations found in the blocks. The10 blocks were located within the Gaillac

Olivier Geffroy et al.

© 2019 International Viticulture and Enology Society - IVES OENO One 2019, 3, 457-470458

Page 3: A 2-year multisite study of viticultural and environmental ...

protected designation of origin, which covers asurface area of approximately 1000 km2.

At each block, one 30-vine panel (comprising 10consecutive vines from three adjacent rows) wasselected for data collection and sampling. Vinemanagement was performed by growers, exceptfor fruit-zone leaf removal, which is a practicethat is likely to affect rotundone concentration(Geffroy et al., 2014; Homich et al., 2017;Geffroy et al., 2018b). To avoid bias due todifferences in the timing and extent ofmechanical leaf removal between blocks whenthe practice had to be performed by growers, allthe leaves were removed manually from thebasal node to the first node above the top bunchon the morning-sun side of the canopy at mid-veraison, when rotundone starts to accumulate inberries (Caputi et al., 2011). Row orientationwas determined using Google Earth Pro softwareversion 7.3.0.3832 (Google Inc., Mountain View,CA, USA). Elevation and slope data wereextracted from the RGE ALTI 5 m device,provided by the French National Institute ofGeographic and Forest Information.

2. Phenology and vine measurements

For each block, the percentage of veraison wasdetermined by randomly collecting 100 berriesfrom both sides of the row and also several partsof the bunch, and by counting the number ofcolored berries every third day from the end ofJuly to the end of August. Linear regression testswere performed using Excel software (MicrosoftCorp., Redmond, WA, USA) to calculate theonset and end of veraison, the mid-veraison date,and the duration of veraison.

Exposed leaf area (m2/m2 soil), which is the leafarea that can be exposed to direct solar radiation,was estimated at mid-veraison using the protocolproposed by Murisier (1996). Measurementswere made for six consecutive vines on each ofthe three rows constituting the 30-vine panel.The ratio between canopy height and rowspacing, which is an indicator of the shade ofone vine row provided by an adjacent row(Smart, 1985), was also calculated.

For each sampled vine, the number of bunchesper vine was counted and the yield at harvest(kg/vine) was monitored by weighing individualcrop loads, using a Precia Molen C20 K balance(Precia SA, Privas, France). Mean bunchweights were also calculated for each block.

Leaf area to crop weight ratio, which is anindicator of vine balance and ability to producehigh-quality fruit and wines (Kliewer andDokoozlian, 2005) was calculated as follows:

leaf area to crop weight ratio = [exposed leafarea (m2/m2 soil) ´ 10 000] / [yield at harvest(kg/vine) ´ plant density (vines/ha)]

In 2014 only, powdery mildew (Erysiphenecator) infections were detected on bunchesfrom vines of one of the blocks. For this vintage,the severity of the infection was assessed atharvest by visually estimating the percentage ofthe bunch area colonized by the pathogen on 100grape bunches (50 on each side of the vine). In2013 and 2014, the severity of bunch rot(Botrytis cinerea) was assessed by determinationof the gluconic acid content of the grapes (seethe section Berry sampling and analyses), which

© 2019 International Viticulture and Enology Society - IVESOENO One 2019, 3, 457-470 459

TABLE 1. Characteristics of the 10 vineyard blocks used in the study

E, east; m a.s.l., meters above sea level; MGt, Millardet et de Grasset; MS, massal selection; N, north; S, south; W, west.

Block no. Clone no. Rootstock Plant density (vines/ha)

Training system

Year of plantation

Row orientation

Elevation(m a.s.l.)

Slope (%)

1 554 Gravesac 4545 Guyot 1999 NW to SE 159 1.62 554 101-14 MGt 4132 Guyot 2002 NE to SW 118 0.53 555 3309C 4545 Guyot 2005 NW to SE 195 4.14 555 140R 4545 Guyot 2004 N to S 226 4.05 554 3309C 6172 Cordon 2003 N to S 201 4.96 554 3309C 4545 Cordon 2005 NE to SW 280 5.57 MS SO4 4761 Goblet 1999 NW to SE 146 2.78 554 3309C 4545 Cordon 2001 NW to SE 129 1.29 654 SO4 4000 Cordon 1992 NW to SE 174 1.710 555 Fercal 4545 Guyot 1996 N to S 271 2.8

Page 4: A 2-year multisite study of viticultural and environmental ...

correlates with the visually estimated percentageof rotten grapes (Bocquet et al., 1995).

Pruning wood weight (kg/vine) was estimated byweighing 1-year-old shoots of 10 out of 30 vinesin January 2014 and 2015, using a Precia MolenC20 K balance.

3. Weather measurements and water-balancemodel

Daily high-resolution data provided by MétéoFrance (Toulouse, France), the French weatherservice, were obtained using ANTILOPE(Champeaux et al., 2011) for rainfall andAROME (Brousseau et al., 2011) for airtemperature and evapotranspiration. Resolutionswere 1 km2 and 4 km2 for ANTILOPE andAROME, respectively. These data were used tocalculate the average minimum, maximum andmean daily air temperature for each blockbetween veraison and harvest; to calculate theHuglin index over the period from 1 April to30 September (Tonietto and Carbonneau, 2002);to determine cumulative rainfall over differentperiods; and to run a water-balance model.

The percentage of soil covered by grass differedgreatly between the blocks, and the waterbalance for intercropped systems (WaLIS) modelwas used to simulate the stem water potential(ψstem) of the vine (Celette et al., 2010; Dufourcqet al., 2013) at different time points between15 days before mid-veraison and harvest foreach block. The minimal and maximal values ofψstem over the mid-veraison to harvest periodwere also identified for each block. Calibrationof the model was carried out in 2013 and 2014based on three measurements of ψstem that wereobtained between the beginning of July and theend of August for each block. Each measurementwas taken from 3 of the 30 vines and from twomature exposed leaves per vine, according to theprotocol proposed by Choné et al. (2001). Thesethree vines were marked, and ψstem was alwaysmeasured on these selected plants. On cloudlessdays, leaf blades were initially enclosed inplastic bags between 11:00 and 12:00 hours, andmeasurements were made between 14:00 and15:00 hours.

The hours of sunshine, solar irradiation, andhygrometry data over the mid-veraison toharvest period, with a 9-km2 spatial resolution,were obtained for each block from HelioClim-3,a satellite-based surface solar irradiationdatabase (Blanc et al., 2011; Eissa et al., 2015).

In April 2015, just before budburst, trunkcircumference, which is an index of vine vigorthat correlates well with the rotundoneconcentration in vines within the same block(Geffroy et al., 2015b), was determined for everyvine within each block. We used the average ofmeasurements obtained at three different heights:just above the graft, 10 cm under the trellis wire,and halfway between the two first points ofmeasurement.

4. Berry sampling and analyses

Previous modeling of the kinetics of rotundoneaccumulation in berries has shown that theconcentration increases from veraison andreaches a plateau after a certain time (Zhang etal., 2015b). According to Zhang et al., thisplateau is higher and is reached earlier duringcool and wet seasons. In 2011 and 2012, two hotand dry vintages that did not promote highrotundone concentration in the grapes of vinesgrown in the south-west of France, this plateauwas achieved for Duras wines 44 days after mid-veraison (Geffroy et al., 2014).

To collect grapes with the maximum rotundoneconcentration, each block was harvested in 2013and 2014 exactly 44 days after mid-veraison. Foreach block and for each year, two samples,consisting of 200 berries from both sides of therow and several parts of the bunch (four berriesper vine on each side of the row), were collected.

Before the start of the analysis, the mass of the200 berries in each sample was measured. Theberries were then crushed gently, and the mustsobtained were centrifuged (14,000 g for 6 min)to enable conventional laboratory and δ13Canalyses. Sugar concentration (°Brix) wasdetermined with a PAL digital handheld pocketrefractometer (Atago, Tokyo, Japan). Titratableacidity was measured according to the methodsof the Organisation Internationale de la Vigne etdu Vin (2009), and pH was measured with anHI 3221 pH meter (Hanna Instruments,Woonsocket, RI, USA).

A Konelab Arena 20 sequential analyzer(Thermo Fisher Scientific, Waltham, MA, USA),in conjunction with enzyme kits provided byseveral suppliers, was used to determine theconcentration of amino acids, ammonium,gluconic acid (Megazyme, Bray, Ireland) andmalic acid (Thermo Fisher Scientific). Potassiumconcentration was determined by flamephotometry (Bio Arrow, France) according to the

Olivier Geffroy et al.

© 2019 International Viticulture and Enology Society - IVES OENO One 2019, 3, 457-470460

Page 5: A 2-year multisite study of viticultural and environmental ...

methods of the Organisation Internationale de laVigne et du Vin (2009), and the concentration oftartaric acid was determined by colorimetrictitration (Hill and Caputi, 1970). δ13C, a proxyfor vine water status (van Leeuwen et al., 2009)was determined at harvest according to theprotocol described by Geffroy et al. (2014). Allthe analytical determinations were carried outtwice.

The berry skins that had not been used for theprevious analyses were crushed for 2 min at low

speed (600 rpm) using a Faciclic food blender(Moulinex, Ecully, France), in preparation forphenolic analyses. The concentration ofanthocyanins and the total phenolic index of thegrapes were quantified according to the methoddescribed by Cayla et al. (2002). In 2014 only,samples of 240 berries were also collected fromthe 10 blocks at mid-veraison to determine theweight of 200 berries, δ13C, sugar concentration,titratable acidity, pH, concentration of malic acidand amino acids, and ammonium content,according to previously described methods.

© 2019 International Viticulture and Enology Society - IVESOENO One 2019, 3, 457-470 461

TABLE 2. Variables for which data were collected

aDetermined only in 2014 and excluded from the multivintage model. Data included in the partial least squares regression modelsto predict rotundone concentration in wine

Block PhenologyVariable Variable Variable Period of measurement

Age of the plot 15 days before veraison Pruning wood weight WinterSlope Onset of veraison Trunk circumference WinterClone Mid-veraison Exposed leaf area Veraison

Rootstock End of veraison Height-to-row spacing ratio VeraisonTraining system Duration of veraison Leaf area to crop weight ratio Veraison/harvestRow orientation Stem water potential (!stem) 15 days before veraison

Plant density Stem water potential (!stem) Onset of veraisonElevation Stem water potential (!stem) Mid-veraison

Stem water potential (!stem) End of veraisonStem water potential (!stem) HarvestStem water potential (!stem) Minimum veraison – harvestStem water potential (!stem) Maximum veraison – harvest

Variable Period of measurement Variable Period of measurementYield per vine Harvest Mean air temperature Veraison – harvest

No. of bunches per vine Harvest Minimum air temperature Veraison – harvestBunch weight Harvest Maximum air temperature Veraison – harvest

Sugar concentration Veraison,a harvest Huglin index 1 April – 30 SeptemberMass of 200 berries Harvest Mean air hygrometry Veraison – harvest

Titratable acidity Veraison,a harvest Mean daily irradiation Veraison – harvestpH Veraison,a harvest Hours of sunshine Veraison – harvest

Malic acid concentration Veraison,a harvest Cumulative rainfall 1 January – 30 DecemberTartaric acid concentration Harvest Cumulative rainfall 1 April – 30 SeptemberAmino acids concentration Veraison,a harvest Cumulative rainfall 15 days before veraison – veraisonAmmonium concentration Veraison,a harvest Cumulative rainfall Veraison – harvest

Total phenolic index Harvest Evapotranspiration Veraison – harvestAnthocyanins content Harvest

"13C Veraison,a harvestGluconic acid concentration HarvestSeverity of powdery mildew Harvesta

Skin-to-juice ratio Harvest

Vine

Crop Climate

Page 6: A 2-year multisite study of viticultural and environmental ...

5. Microvinification and rotundone analysis

Rotundone content was analyzed indirectly inwines fermented under microvinificationconditions. For each block and for each year,two grape samples, comprising 800 g of berriesfrom both sides of the row and several parts ofthe bunch, were collected at harvest 44 daysafter mid-veraison. Microscale fermentation wascarried out in a 1-L Erlenmeyer flask, accordingto the protocol described by Geffroy et al.(2014). After 8 days of fermentation, the amountof wine and of skins was measured to determinethe skin-to-juice ratio at pressing (%). The wineswere centrifuged (14,000 g for 6 min), andsulfite (80 mg/L) was added before bottling andstorage at 4°C or less until the rotundoneanalysis.

The rotundone concentration of the wines wasdetermined by solid-phase microextraction-multidimensional gas chromatography–massspectrometry, according to the protocoldescribed by Geffroy et al. (2014). The analyseswere performed in two different years but duringthe same period of each year, to reduce potentialvariation associated with different post-bottlingtimes.

6. Data treatment

Details of the variables for which data werecollected are shown in Table 2. Rotundoneconcentration and data for the other variables,collected for the 10 blocks over the 2 years ofthe study, were subjected to analysis of variance,with the vintage as a factor and blocks asrepetitions, using Xlstat software version 19.547062 (Addinsoft, Paris, France). Fisher’s leastsignificant difference test was used as a post-hoccomparison of means at P £ 0.05.

Based on previous work conducted with the aimof modeling 3-isobutyl-2-methoxypyrazineconcentration in Cabernet Franc grapes(Scheiner et al., 2011), a multivariate analysisusing partial least squares regression (PLSR)was carried out to identify key factors affectingrotundone concentration in wine. The data wereaveraged over the replicates before PLSR wasperformed. For the multi-vintage model,variables that had been determined only in 2014were discarded. A first set of PLSR for which thenumber of latent variables was determined bythe lowest predicted residual sum of squares wasconducted, using Xlstat software to modelrotundone concentration in 2013, in 2014, and in

both vintages (Wold et al., 1984). Variables witha variable importance in projection score below1 were removed manually and the model wasregenerated. Once the number of variablesremaining in the model was lower than thenumber of observations (10 for 2013 and 2014,and 20 for the multi-vintage data set), PLSR wasconducted with Unscrambler X softwareversion 10.4 (Camo, Oslo, Norway).

The results were validated by a full cross-validation (leave-one-out) procedure. Variablesthat did not contribute sufficiently to the model(absolute values of the loading weights below0.1) were removed manually and the model wasregenerated. For the multivintage model, oneoutlier (block no. 6 in 2013) detected usingprincipal component analysis was removed.

RESULTS AND DISCUSSION

The data collected for each vineyard block in2013 and 2014 are presented as supplementarymaterial (Supplementary tables 1 and 2).

1. Main differences between the vintages

The 2013 vintage was characterized by a coldspring and regular and heavy rainfall events overthe vine vegetative cycle that resulted in severebunch rot (Botrytis cinerea) damage, ripeningdifficulties and delay of harvest. The meanharvest date for the 10 blocks was 6 October ±4 days. In 2014, the winter was rainy and warmand summer was extremely rainy between mid-July and the end of August, whereas conditionsduring ripening were dry and hot. The meanharvest date was 22 September ± 3 days.Compared with 2014, the 2013 vintage wassignificantly wetter as well as cooler over theentire winegrowing and ripening periods, asreflected by the cumulative rainfall between 1April and 30 September (423 ± 22 mm in 2013,358 ± 19 mm in 2014) and between veraison andharvest (115 ± 18 mm in 2013, 63 ± 13 mm in2014), the Huglin index (1862 ± 51 in 2013,1969 ± 48 in 2014), and the mean airtemperature over the veraison-to-harvest period(18.50 ± 0.56°C in 2013, 19.48 ± 0.42°C in2014). The values of ψstem were significantlylower at mid-veraison during the first year of thestudy (–0.87 ± 0.20 MPa in 2013,–0.60 ± 0.12 MPa in 2014). Owing to the largerquantities of rainfall received in 2013 over theveraison-to-harvest period, the level of waterdeficit was higher at harvest in 2014

Olivier Geffroy et al.

© 2019 International Viticulture and Enology Society - IVES OENO One 2019, 3, 457-470462

Page 7: A 2-year multisite study of viticultural and environmental ...

(–0.51 ± 0.17 MPa in 2013, –1.06 ± 0.14 MPa in2014).

Vintage had a weak effect on crop and grapeanalytical characteristics, because no significantdifferences in yield (1.75 ± 1.01 kg/vine in 2013,2.65 ± 1.22 kg/vine in 2014), sugar concentration(21.8 ± 0.9°Brix in 2013, 21.8 ± 0.6°Brix in2014), pH (3.06 ± 0.12 in 2013, 3.05 ± 0.07 in2014), titratable acidity (9.20 ± 1.13 g/L astartaric acid in 2013, 9.10 ± 1.22 g/L as tartaricacid in 2014), amino acid concentration(120 ± 33 mg/L in 2013, 136 ± 48 mg/L in 2014)or total phenolic index (79.6 ± 10.8 in 2013,88.5 ± 15.2 in 2014) were observed between2013 and 2014. Concentrations of anthocyanins(872 ± 143 mg/L in 2013, 1048 ± 194 mg/L in2014) and ammonium (40.9 ± 20.3 mg/L in2013, 63.8 ± 20.4 mg/L in 2014) were the onlyfruit compound variables that differedsignificantly between the two vintages.

Rotundone concentration was significantlyhigher (P < 0.001) in the cooler and wetter 2013growing season (Figure 1). This result isconsistent with those of previous studies (Caputiet al., 2011). Surprisingly, some blocks that hadhigh rotundone concentration in 2013 (i.e. blocksno. 2 and no. 8), with values among the highestever reported for Duras, had low to moderateconcentrations in 2014. In the same way, someblocks that had high concentrations in 2014 (i.e.blocks no. 3 and no. 5) had low to moderaterotundone concentrations in 2013. This

observation runs counter to the belief that somevineyards within a given wine region mayalways be prone to yielding wines withsubstantial amounts of rotundone. This led us tothe notion that qualitative or quantitative fixedvariables such as year of plantation, elevation,training system, clone or rootstock do not makea substantial contribution to the rotundonemodel.

2. Partial least squares regression rotundonemodels

Basic statistics of the models, regressioncoefficients, means and standard deviations ofthe variables included in the best models, areshown in Table 3. Regression plots of predictedrotundone concentration versus observedrotundone concentration in 2013, in 2014, and inboth vintages are presented in Figure 2. Theperformance of the 2013 vintage model,reflected by its values of R2 and root meansquare error, was better than that of the othermodels. However, all the models appear to beappropriate for predicting rotundoneconcentration using a limited number of inputvariables for blocks with low or high rotundoneconcentrations.

The loadings of the latent variables used in theregression models are shown in Table 4. Thefirst latent variable in the 2013 model wasmainly characterized by gluconic acidconcentration. For 2013, this variable was

© 2019 International Viticulture and Enology Society - IVESOENO One 2019, 3, 457-470 463

FIGURE 1. Rotundone concentration in wines made from grapes from the 10 vineyard blocks of Duras Nin 2013 and 2014.

Page 8: A 2-year multisite study of viticultural and environmental ...

significantly negatively correlated withrotundone concentration (P < 0.01, R2 = 0.69)(Figure 3). It also contributed to the multi-vintage model. This observation was somewhatunexpected, because laccase, the polyphenoloxidase produced by Botrytis cinerea, has the

ability to convert α-guaiene to rotundone via anon-specific oxidation reaction in the presence ofchemical mediators (Schilling et al., 2013).However, it has also been documented thatBotrytis cinerea can withstand the toxic effectsof plant compounds such as sesquiterpenes bylaccase-mediated oxidation and detoxification(Mayer and Staples, 2002). Further studies willbe necessary to confirm this hypothesis and tobetter understand the mechanisms involved inrotundone degradation by the fungus.

Other variables contributing to the latentvariables of the models were associated withabiotic factors, which indicates that mesoscaleclimate is the key factor driving rotundoneproduction. Notably, the Huglin index, a thermalindex based on a degree-day approach from 1April to 30 September, was a strong contributorto the 2014 model. This is consistent with thefinding that cool vintages promote rotundoneaccumulation in grapevine berries (Caputi et al.,2011). A contribution to the models from meanair temperature between veraison and harvestwould have been expected, because post-veraison berry surface temperature has beenidentified as the main determinant of rotundoneconcentration in grape berries (Zhang et al.,2015a).

Bunch temperature does not fully reflect airtemperature, because it is also determined byabsorbed radiation and convective heat loss (Dry,2009). The extent to which bunch temperatureexceeds air temperature depends on degree ofexposure, radiation load, wind velocity, berry orbunch size, berry color and bunch compactness.With the exception of wind velocity, which wasnot assessed, most of the data for these variableswere collected directly or depend on datacollected for other variables, such as bunchweight, mass of 200 berries, training system,rootstock, row orientation and exposed leaf area.If bunch temperature were the key variabledriving rotundone concentration, the bestrotundone model would include air temperatureand other variables affecting bunch exposure andpenetration of solar radiation through the canopy.It has recently been reported that the percentageof degree hours above 25°C (Dh25) in the fruitzone from veraison to harvest correlates withrotundone concentration in berries (Zhang et al.,2015a). Our study was designed and conducted2 years before publication of these findings. Dh25could not be calculated a posteriori, becausecollection of the data for this variable would

Olivier Geffroy et al.

© 2019 International Viticulture and Enology Society - IVES OENO One 2019, 3, 457-470464

FIGURE 2. Predicted versus observed rotundoneconcentration.Predicted versus observed rotundone concentration, usingdata from the models described in Table 3: A, 2013(R2 = 0.95); B, 2014 (R2 = 0.78); and C, multi-vintage(R2 = 0.81).

0

50

100

150

200

250

0 50 100 150 200 250 300

Pre

dict

ed ro

tund

one

(ng/

L)

Observed rotundone (ng/L)

0

20

40

60

80

100

120

0 20 40 60 80 100 120 140

Pre

dict

ed ro

tund

one

(ng/

L)

Observed rotundone (ng/L)

0

50

100

150

200

250

0 50 100 150 200 250 300

Pre

dict

ed ro

tund

one

(ng/

L)

Observed rotundone (ng/L)

C

!

"

250

50

100

1 (

200

) L/g(n150

enodnutro

dectid

0 0 50

dP

re

80

100

120

) L/

100 150 200

g(nenodnutrodrvesebO

250 300

) L/g

0

20

40

60

80

0 20

g(n

enodnutro

dectid

Pre

40 60 80 100

100 120 140

0 20

1 (

200

250

) L/g(n150

eno

40 60 80 100

g(nenodnutrodrvesebO

100 120 140

) L/g

0

50

100

0 50

dnutro

dectid

Pre

100 150 200

g(nenodnutrodrvesebO

250 300

) L/g

Page 9: A 2-year multisite study of viticultural and environmental ...

© 2019 International Viticulture and Enology Society - IVESOENO One 2019, 3, 457-470 465

TABLE 3. Basic statistics of the partial least squares regression models

NA, not applicable; NI, variable not included in the model. *Significantly higher of the means for 2013 and 2014 (leastsignificant difference, P £ 0.05). aRegression coefficient between the variable and rotundone concentrationRegression coefficients, means and standard deviations of variables for the best rotundone models

2013 2014 2013–2014 2013 2014Cumulative rainfall, veraison – harvest (mm) –0.90a NI NI 115 ± 18* 63 ± 13Hours of sunshine 0.72a NI 0.61a 273 ± 17 362 ± 5*Gluconic acid concentration (mg/L) –0.83a NI –0.001a 72.4 ± 69.1* 14.3 ± 21.2Cumulative rainfall, 1 April – 30 September (mm) NI 0.85a 0.77a 423 ± 22* 358 ± 19Huglin index NI –0.63a NI 1862 ± 51 1969 ± 48*Cumulative rainfall, 1 January – 31 December (mm) NI NI 0.66a 19.1 ± 12.6 50.1 ± 7.9*

Mean daily irradiation, veraison – harvest (W/m2) NI NI 0.59a 1505 ± 81 1863 ± 50*No. of latent variables 2 2 2 NA NARoot mean square error 11.4 14.1 28.5 NA NA

R2 (calibration) 0.95 0.78 0.81 NA NARoot mean square error of cross-validation 14.9 20.8 41.3 NA NA

R2 (validation) 0.92 0.53 0.59 NA NA

VariableModel Mean and standard

deviation of model

TABLE 4. Loadings of the variables contributing to the best rotundone models

LV, latent variable; NI, variable not included in the model.

LV 1 LV 2 LV 1 LV 2 LV 1 LV 2Gluconic acid concentration –0.99 0.32 NI NI 0.17 –0.84Hours of sunshine, veraison – harvest 0.11 0.78 NI NI –0.21 0.05Cumulative rainfall, veraison – harvest –0.18 –0.55 NI NI NI NICumulative rainfall, 1 April – 30 September NI NI 0.26 0.88 0.14 0.20Huglin index NI NI –1.00 0.48 NI NICumulative rainfall, 1 January – 31 December NI NI NI NI 0.43 0.42Mean daily irradiation, veraison – harvest NI NI NI NI –0.88 0.29Explained variance (%) 76.1 19.1 51.9 25.7 40.8 39.9

Variable2013 2014 2013–2014

0

50

100

150

200

250

300

0 50 100 150 200 250

Rot

undo

ne (n

g/L)

Gluconic acid (mg/L)

300

) L/g(n

enodnutoR

100

150

200

250

) L/g(n

enodnutoR

0

50

0 50

100 150

) L/(mgdciac incoulG

200 250

FIGURE 3. Relation between gluconic acid concentration and rotundone concentration.Relation between gluconic acid content and rotundone concentration in wines in 2013 (n = 10). Linear regression model:y = –0.6667x + 204.47, P < 0.01, R2 = 0.69.

Page 10: A 2-year multisite study of viticultural and environmental ...

have required the use of temperature loggerslocated within the canopy. We cannot completelyexclude the possibility that Dh25 would havebeen included in the models.

Mean daily irradiation between veraison andharvest had a substantial contribution to themulti-vintage model. Hours of sunshine was alsoincluded in the best rotundone models for 2013and 2013–2014. These two variables, reflectingthe quantity of ultraviolet radiation received bythe blocks, were positively correlated withrotundone concentration. Several studies haveinvestigated the light dependency ofsesquiterpene emission, with evidence thatemission of some sesquiterpenes is solelytemperature-controlled whereas emission ofothers is also affected by light (Duhl et al.,2008). Although additional studies would beneeded to confirm this hypothesis, our datasuggest that ultraviolet rays may stimulate theproduction of rotundone. This conclusion isconsistent with recent research conducted incool-climate vineyards on Noiret and Fercultivars, indicating that grape exposure by leafremoval could promote rotundone accumulation(Homich et al., 2017; Geffroy et al., 2018b).

The amount of rainfall between 1 January and 31December and during the period of vinevegetative growth (between 1 April and 30September) had a positive contribution to the2014 and multi-vintage models, whereas rainfallduring the late stage of development (i.e.between veraison and harvest) was negativelycorrelated with rotundone concentration in 2013.These results are somewhat unexpected, becauseirrigation around veraison, when rotundone startsto accumulate in berries, stimulates itsproduction (Geffroy et al., 2014; Geffroy et al.,2016b). For the 2013 rainy vintage, even if skin-to-juice ratio and mass of 200 berries were notincluded in the model, we cannot exclude thepossibility that the large excess of water mayhave resulted in a dilution effect through anincrease in berry size.

The contribution to the models from cumulativerainfall between 1 April and 30 September andbetween 1 January and 31 December deservesfurther comment. Rotundone is found at elevatedconcentrations in the early berry stages, notablyin flower caps (Zhang et al., 2016b). Rotundoneconcentration then decreases to reach a minimalconcentration at veraison, before increasingagain from veraison to harvest (Zhang et al.,

2016b). It could be assumed that theprecipitation regimen over the entire berrydevelopment period, not only during the laststage of maturation, may have affected the finalconcentration of rotundone through degradationor emission before veraison. Another hypothesisis that increased water availability over theperiod of vegetative growth and between 1January and 31 December would have led tohigher-vigor vines, resulting in greater bunchshading and lower bunch-zone air temperature.However, this hypothesis is unlikely, because thelevel of water deficit experienced by the vines atmid-veraison, when shoot growth stops, wasminor (ψstem = –0.87 ± 0.20 MPa in 2013,ψstem = –0.60 ± 0.12 MPa in 2014) andinsufficient to affect vine vegetative expression(van Leeuwen et al., 2009). Additionally, nosignificant correlation at P < 0.05 could beestablished for the whole data set betweencumulative rainfall and either pruning woodweight or exposed leaf area.

Consistent with the fact that the hierarchy ofblocks for rotundone concentration can differgreatly from one vintage to the other, none of therotundone models included any fixed variables.This is somewhat unexpected for the clone,because higher rotundone concentrations havebeen reported in Duras wines from clones 554and 654 than from clone 555 (Geffroy et al.,2015a). In the same way, the rootstock couldhave affected the concentration of rotundone inwine. The rootstock is likely to affect some keyvariables driving rotundone production, such asscion water status and bunch surface, through itseffect on vine vigor, leaf area and penetration ofsolar radiation (Koundouras et al., 2008).However, it may be difficult to draw firmconclusions, due to the small number ofobservations for each level of these two factorsand the fact that not all clone–rootstockcombinations were available.

It has been suggested that the grapevine defenseresponse to powdery mildew could enhancerotundone production in berries (Geffroy et al.,2015a). This hypothesis could not be confirmedunder in the present study, because severity ofpowdery mildew infection was not included inthe 2014 model. Period of infection by Erysiphenecator was not assessed in that previous studynor in the present study. On the one hand, asrotundone accumulates around veraison,differences in timing of infection, which usuallyoccurs when bunches are receptive (from

Olivier Geffroy et al.

© 2019 International Viticulture and Enology Society - IVES OENO One 2019, 3, 457-470466

Page 11: A 2-year multisite study of viticultural and environmental ...

blooming to veraison), may play a role indefense mechanisms leading to production ofrotundone. On the other hand, powdery mildewdamage was noticed only in 2014 at just oneblock, and it remains difficult to draw firmconclusions regarding the contribution of thisbiotic factor.

Consistent with the findings that rotundone is nottranslocated from vegetative tissues (Geffroy etal., 2016b; Zhang et al., 2016a) and that itsaccumulation does not depend on source–sinkrelations (Geffroy et al., 2014), it should bepointed out that the leaf area to crop weight ratiowas not included in any of the rotundone models.

The absence of contribution to the models fromsugar concentration at harvest is somewhatunexpected. Although recent studies have shownthat an increased alcohol content of wineimproves rotundone extraction rate from grapesinto wine (Zhang et al., 2017), our data suggestthat ethanol concentration in the resulting wineshad no effect on rotundone concentration. Thismay be explained by the relatively smallamplitude of difference between the lowest andhighest sugar concentrations (20.1°Brix and22.8°Brix, respectively) recorded between blocksfor the 2 years of the study.

We cannot rule out the possibility that someblocks were harvested in 2013 and 2014 beforerotundone concentration reached its maximum.This would mean that the proposed modelswould be valid for prediction of rotundoneconcentration only in wines from grapesharvested exactly 44 days after mid-veraison andmay not be applicable to grapes harvestedbeyond that time. However, this hypothesis isunlikely, because 2013 and 2014 werecharacterized by substantial concentrations ofrotundone in Duras wines (Geffroy et al.,2016b). Therefore, considering previous findings(Zhang et al., 2016a), the rotundone plateau andthe maximum rotundone concentration wereexpected to have been reached earlier than in2011 and 2012, and before 44 days after mid-veraison.

With only 2 years of data, we cannot completelydisregard the possibility that our model may notbe generalizable. This is particularly truebecause the two vintages of the study wererelatively cool and wet for the area, as illustratedby the data for Huglin index and cumulativerainfall between 1 April and 30 September. The

2015 vintage, which was hotter and dryer, wascharacterized by a different climatic pattern thanthat of 2013 and 2014. The equation of themulti-vintage model was used to predictrotundone concentration in wine fermentedunder the same microvinification conditionsfrom grapes collected in 2015 on a Duras plot44 days after mid-veraison. Based onmeasurements made on this block for mean dailyirradiation between veraison and harvest(1930 W/m2), hours of sunshine betweenveraison and harvest (419), cumulative rainfallbetween 1 April and 30 September (348 mm),cumulative rainfall between 1 January and 31December (516 mm) and gluconic acidconcentration (12 mg/L), the predictedrotundone concentration was 30.8 ng/L; theobserved value was 26.0 ng/L. In light of thisprediction, we believe that the model may besufficiently robust to also allow estimation ofrotundone concentration during hot vintages.

CONCLUSION

Our study allowed us to model and assess theeffect of environmental and viticultural factorson rotundone concentration in wine. Gluconicacid, a secondary metabolite of Botrytis cinereathat is negatively correlated with rotundoneconcentration, was included in the bestrotundone model for the 2013 vintage and multi-vintage (2013–2014) models. This observationsuggests that the fungus may be involved inrotundone degradation mechanisms, probablythrough its laccase activity. Other variablesincluded in the PLSR models were abioticfactors. This indicates that mesoscale climate isthe key factor driving rotundone production. Inmost cases, cumulative rainfall and thermalindex for the entire vine vegetative period seemto be better predictors than variables calculatedduring the maturation period only. The positivecontribution of mean daily irradiation betweenveraison and harvest and hours of sunshinesuggest that rotundone production may bestimulated by ultraviolet exposure. Our findingsmay assist grape growers producing Vitisvinifera L. cv. Duras red wines from vineyardsscattered across the studied viticultural area, toselect specific blocks with the aim of producingwines with a specific rotundone concentration.They also open new fields of investigation intothe mechanisms involved in possible rotundonedegradation by Botrytis cinerea.

© 2019 International Viticulture and Enology Society - IVESOENO One 2019, 3, 457-470 467

Page 12: A 2-year multisite study of viticultural and environmental ...

Acknowledgments: This study received financialsupport from FranceAgriMer and the Midi-Pyrénéesregion. We are grateful to Xavier Delpuech andBrigitte Mille, Institut Français de la Vin et du Vin,for their technical assistance in running the WaLISwater-balance model and the analyses, and we thankthe winegrowers for allowing us to use theirvineyards.

REFERENCESBlanc P., Gschwind B., Lefèvre M. and Wald L.,2011. The HelioClim project: surface solar irradiancedata for climate applications. Remote Sensing, 3,343–361. doi: 10.3390/rs3020343Bocquet F., Moncomble D. and Valade M., 1995.Etat sanitaire de la vendange et qualité des vins. LeVigneron Champenois, 7/8, 15–23.Bramley R.G.V., Siebert T.E., Herderich M.J. andKrstic M.P., 2017. Patterns of within-vineyard spatialvariation in the ‘pepper’ compound rotundone aretemporally stable from year to year. AustralianJournal of Grape and Wine Research, 23, 42–47. doi:10.1111/ajgw.12245Brousseau P., Berre L., Bouttier F. and Desroziers G.,2011. Background-error covariances for a convective-scale data-assimilation system: AROME–France 3D-Var. Quarterly Journal of the Royal MeteorologicalSociety, 137, 409–422. doi: 10.1002/qj.750Caputi L., Carlin S., Ghiglieno I., Stefanini M.,Valenti L., Vrhovsek U. and Mattivi F., 2011.Relationship of changes in rotundone content duringgrape ripening and winemaking to manipulation ofthe ‘peppery’ character of wine. Journal ofAgricultural and Food Chemistry, 59, 5565–5571.doi: 10.1021/jf200786uCayla L., Cottereau P. and Renard R., 2002.Estimation de la maturité phénolique des raisinsrouges par la méthode ITV standard. Revue Françaised’Oenologie, 193, 10–16.Celette F., Ripoche A. and Gary C., 2010. WaLIS – asimple model to simulate water partitioning in a cropassociation: the example of an intercropped vineyard.Agricultural Water Management, 97, 1749–1759. doi:10.1016/j.agwat.2010.06.008Champeaux J.L., Laurantin O., Mercier B., MounierF., Lassegues P. and Tabary P., 2011. Quantitativeprecipitation estimations using rain gauges and radarnetworks: inventory and prospects at Meteo-France.WMO Joint Meeting of CGS Expert Team onSurface-based Remotely-sensed Observations andCIMO Expert Team on Operational Remote Sensing,5–9 December 2011, Geneva, Switzerland.Choné X., van Leeuwen C., Dubourdieu D. andGaudillère J.P., 2001. Stem water potential is asensitive indicator of grapevine water status. Annalsof Botany, 87, 477–483. doi: 10.1006/anbo.2000.1361

Cullere L., Ontanon I., Escudero A. and Ferreira V.,2016. Straightforward strategy for quantifyingrotundone in wine at ng L–1 level using solid-phaseextraction and gas chromatography–quadrupole massspectrometry. Occurrence in different varieties ofspicy wines. Food Chemistry, 206, 267–273. doi:10.1016/j.foodchem.2016.03.039Dry P., 2009. Bunch Exposure Management.Technical booklet, Grape and Wine Research andDevelopment Corporation, Adelaide, Australia.Dufourcq T., Barraud G., Delpuech X. and Gaudin R.,2013. Use of the water balance model WaLIS topredict stem water potential on Colombard variety inGascony vineyard. Ciência e Técnica Vitivinícola, 28,75–79.Duhl T.R., Helmig D. and Guenther A., 2008.Sesquiterpene emissions from vegetation: a review.Biogeosciences, 5, 761–777. doi: 10.5194/bgd-4-3987-2007Eissa Y., Korany M., Aoun Y., Boraiy M., AbdelWahab M.M., Alfaro S.C., Blanc P., El-Metwally M.,Ghedira H., Hungershoefer K. and Wald L., 2015.Validation of the surface downwelling solarirradiance estimates of the HelioClim-3 database inEgypt. Remote Sensing , 7, 9269–9291. doi:10.3390/rs70709269Geffroy O., Dufourcq T., Carcenac D., Siebert T.,Herderich M. and Serrano E., 2014. Effect of ripenessand viticultural techniques on the rotundoneconcentration in red wine made from Vitis vinifera L.cv. Duras. Australian Journal of Grape and WineResearch, 20, 401–408. doi: 10.1111/ajgw.12084Geffroy O., Yobrégat O., Dufourcq T., Siebert T. andSerrano E., 2015a. Certified clone and powderymildew impact rotundone in red wine from Vitisvinifera L. cv. Duras N. Journal International desSciences de la Vigne et du Vin, 49, 231–240.Geffroy O., Scholasch T., Dufourcq T. and SerranoE., 2015b. Understanding and mapping rotundonespatial variability in Vitis vinifera L. c.v Duras. InProceedings of the 19th International MeetingGiESCO, 31 May – 5 June 2015, Gruissan, France,pp. 589–592.Geffroy O., Buissière C., Lempereur V. and ChateletB., 2016a. A sensory, chemical and consumer studyof the peppery typicality of French Gamay winesfrom cool-climate vineyards. Journal Internationaldes Sciences de la Vigne et du Vin, 50, 35–47. doi:10.20870/oeno-one.2016.50.1.53Geffroy O., Siebert T., Herderich M., Mille B. andSerrano E., 2016b. On-vine grape drying combinedwith irrigation allows to produce red wines withenhanced phenolic and rotundone concentrations.Scientia Horticulturae, 207, 208–217. doi:10.1016/j.scienta.2016.05.031Geffroy O., Descôtes J., Serrano E., Li Calzi M.,Dagan L. and Schneider R., 2018a. Can a certain

Olivier Geffroy et al.

© 2019 International Viticulture and Enology Society - IVES OENO One 2019, 3, 457-470468

Page 13: A 2-year multisite study of viticultural and environmental ...

concentration of rotundone be undesirable in Durasred wine? A study to estimate a consumer rejectionthreshold for the pepper aroma compound. AustralianJournal of Grape and Wine Research, 24, 88–95. doi:10.1111/ajgw.12299Geffroy O., Ibfelt K., Feilhès C. and Dufourcq T.,2018b. Manipulating the rotundone and 2-methoxy-3-isobutylpyrazine composition of Vitis vinifera L. cv.Fer red wines from cool-climate vineyards. InProceedings of the 12th International TerroirCongress, 18–22 June 2018, Saragossa, Spain.Hill G. and Caputi A., 1970. Colorimetricdetermination of tartaric acid in wine. AmericanJournal of Enology and Viticulture, 21, 153–161.Homich L.J., Elias R.J., Heuvel J.E.V. andCentinari M., 2017. Impact of fruit-zone leaf removalon rotundone concentration in Noiret. AmericanJournal of Enology and Viticulture, 68, 447–457. doi:10.5344/ajev.2017.16106Huang A.C., Burrett S., Sefton M.A. and Taylor D.K.,2014. Production of the pepper aroma compound, (–)-rotundone, by aerial oxidation of alpha-guaiene.Journal of Agricultural and Food Chemistry, 62,10809–10815.Kliewer W.M. and Dokoozlian N.K., 2005. Leafarea/crop weight ratios of grapevines: influence onfruit composition and wine quality. American Journalof Enology and Viticulture, 56, 170–181.Koundouras S., Tsialtas I.T., Zioziou E. andNikolaou N., 2008. Rootstock effects on the adaptivestrategies of grapevine (Vitis vinifera L. cv.Cabernet–Sauvignon) under contrasting water status:leaf physiological and structural responses.Agriculture, Ecosystems & Environment, 128, 86–96.doi: 10.1016/j.agee.2008.05.006Mayer A.M. and Staples R.C., 2002. Laccase: newfunctions for an old enzyme. Phytochemistry, 60, 551–565.Murisier F.M., 1996. Optimalisation du rapportfeuille-fruit de la vigne pour favoriser la qualité duraisin et l’accumulation des glucides de réserve:relation entre le rendement et la chlorose. Thesis,Eidgenössische Technische Hochschule, Zürich.Organisation Internationale de la Vigne et du Vin,2009. Recueil des Méthodes Internationalesd’Analyse des Vins et des Moûts. OrganisationInternationale de la Vigne et du Vin, Paris, France.Robinson A.L., Boss P.K., Solomon P.S.,Trengove R.D., Heymann H. and Ebeler S.E., 2014.Origins of grape and wine aroma. Part 1. Chemicalcomponents and viticultural impacts. AmericanJournal of Enolology and Viticulture, 65, 1–24.Sala C., Mestres M., Martı́ M.P., Busto O. andGuasch J., 2000. Headspace solid-phase micro-extraction method for determining 3-alkyl-2-methoxypyrazines in musts by means of

polydimethylsiloxane–divinylbenzene fibres. Journalof Chromatography A, 880, 93–99. doi:10.1016/S0021-9673(00)00262-4Scarlett N.J., Bramley R.G.V. and Siebert T.E., 2014.Within-vineyard variation in the ‘pepper’ compoundrotundone is spatially structured and related tovariation in the land underlying the vineyard.Australian Journal of Grape and Wine Research, 20,214–222. doi: 10.1111/ajgw.12075Scheiner J.J., Vanden Heuvel J.E., Pan B. andSacks G.L., 2011. Modeling impacts of viticulturaland environmental factors on 3-isobutyl-2-methoxypyrazine in Cabernet Franc grapes. AmericanJournal of Enology and Viticulture, 63, 94–105. doi:10.5344/ajev.2011.11002Schilling B., Granier T. and Locher E., 2013. 1-Hydroxy-octahydroazulenes as fragrances. US patentapplication US13/703,761. Publication no.20130089904 A1.Smart R.E., 1985. Principles of grapevine canopymicroclimate manipulation with implications foryield and quality. American Journal of Enology andViticulture, 36, 230–239.Takase H., Sasaki K., Shinmori H., Shinohara A.,Mochizuki C., Kobayashi H., Ikoma G., Saito H.,Matsuo H., Suzuki S. and Takata R., 2016.Cytochrome P450 CYP71BE5 in grapevine (Vitisvinifera) catalyzes the formation of the spicy aromacompound (–)-rotundone. Journal of ExperimentalBotany, 67, 787–798. doi: 10.1093/jxb/erv496Tominaga T., Peyrot des Gachons C. andDubourdieu D., 1998. A new type of flavorprecursors in Vitis vinifera L. cv. Sauvignon Blanc:S-cysteine conjugates. Journal of Agricultural andFood Chemistry, 46, 5215–5219. doi:10.1021/jf980481uTonietto J. and Carbonneau A., 2002. Régimethermique en période de maturation du raisin dans legéoclimat viticole: indice de fraîcheur des nuits – IFet amplitude thermique. In Proceedings of theSymposium International sur le Zonage Vitivinicole,17–20 June 2002, Avignon, France (OrganisationInternationale de la Vigne et du Vin, Paris, France),pp. 279–289.van Leeuwen C., Trégoat O., Choné X., Bois B.,Pernet D. and Gaudillère J.P., 2009. Vine water statusis a key factor in grape ripening and vintage qualityfor red Bordeaux wine. How can it be assessed forvineyard management purposes? JournalInternational des Sciences de la Vigne et du Vin, 43,121–134. doi: 10.20870/oeno-one.2009.43.3.798Williams P.J., Strauss C.R. and Wilson B., 1981.Classification of the monoterpenoid composition ofMuscat grapes. American Journal of Enology andViticulture, 32, 230–235.

© 2019 International Viticulture and Enology Society - IVESOENO One 2019, 3, 457-470 469

Page 14: A 2-year multisite study of viticultural and environmental ...

Wold S., Ruhe A., Wold H. and Dunn I.W.J., 1984.The collinearity problem in linear regression. ThePartial Least Squares (PLS) approach to generalizedinverses. SIAM Journal of Scientific and StatisticalComputing, 5, 735–743. doi: 10.1137/0905052Wood C., Siebert T.E., Parker M., Capone D.L.,Elsey G.M., Pollnitz A.P., Eggers M., Meier M.,Vossing T., Widder S., Krammer G., Sefton M.A. andHerderich M.J., 2008. From wine to pepper:rotundone, an obscure sesquiterpene, is a potent spicyaroma compound. Journal of Agricultural and FoodChemistry, 56, 3738–3744. doi: 10.1021/jf800183kZhang P., Barlow S., Krstic M., Herderich M.,Fuentes S. and Howell K., 2015a. Within-vineyard,within-vine, and within-bunch variability of therotundone concentration in berries of Vitis vinifera L.cv. Shiraz. Journal of Agricultural and FoodChemistry, 63, 4276–4283. doi: 10.1021/acs.jafc.5b00590Zhang P., Howell K., Krstic M., Herderich M.,Barlow E.W.R. and Fuentes S., 2015b.

Environmental factors and seasonality affect theconcentration of rotundone in Vitis vinifera L. cv.Shiraz wine. PLoS One 10, e0133137. doi:10.1371/journal.pone.0133137Zhang P., Fuentes S., Wang Y., Deng R., Krstic M.,Herderich M., Barlow E.W.R. and Howell K., 2016a.Distribution of rotundone and possible translocationof related compounds amongst grapevine tissues inVitis vinifera L. cv. Shiraz. Frontiers in PlantScience, 7, 859. doi: 10.3389/fpls.2016.00859Zhang P., Fuentes S., Siebert T., Krstic M.,Herderich M., Barlow E.W.R. and Howell K., 2016b.Terpene evolution during the development of Vitisvinifera L. cv. Shiraz grapes. Food Chemistry, 204,463–474. doi: 10.1016/j.foodchem.2016.02.125Zhang P., Luo F. and Howell K., 2017. Fortificationand elevated alcohol concentration affect theconcentration of rotundone and volatiles in Vitisvinifera cv. Shiraz wine. Fermentation, 3, 29. doi:10.3390/fermentation3030029

Olivier Geffroy et al.

© 2019 International Viticulture and Enology Society - IVES OENO One 2019, 3, 457-470470