Effects of Climate Change

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Agricultural Water Management 159 (2015) 155–164

Contents lists available at ScienceDirect

Agricultural Water Management


journal homepage: www.elsevier.com/locate/agwat

Effects of climate change including elevated CO2 concentration,


temperature and water deficit on growth, water status, and yield
quality of grapevine (Vitis vinifera L.) cultivars
T. Kizildeniz a , I. Mekni a , H. Santesteban a , I. Pascual a , F. Morales b,∗ , J.J. Irigoyen a
a
Grupo de Fisiología del Estrés en Plantas (Dpto. de Biología Ambiental), Unidad Asociada al CSIC, EEAD, Zaragoza e ICVV, Logroño, Facultades de Ciencias y
Farmacia, Universidad de Navarra, Irunlarrea 1, 31008 Pamplona, Spain
b
Estación Experimental de Aula Dei (EEAD), CSIC, Dpto. Nutrición Vegetal, Apdo. 13034, 50080 Zaragoza, Spain

a r t i c l e i n f o a b s t r a c t

Article history: In the Mediterranean area, climate change is associated with atmospheric CO2 concentration increases,
Received 29 October 2014 enhanced temperatures and scarce water availability, limiting seriously crop yield and decreasing quality.
Received in revised form 9 June 2015 The aim of this study was to investigate the effects of elevated CO2 , elevated temperature and water deficit,
Accepted 13 June 2015
acting individually and/or interacting, on vegetative and reproductive growth, substrate and plant water
status, and must quality in fruit-bearing cuttings of two grapevine (Vitis vinifera L.) cultivars (red and
Keywords:
white Tempranillo). In four temperature gradient greenhouses, eight treatments were applied, from fruit
Red and white Tempranillo
set to maturity: CO2 level (400 versus 700 ␮mol mol−1 ), temperature (ambient versus ambient +4 ◦ C),
Anthocyanins
Malic and tartaric acid
and water availability (full irrigation versus cyclic drought). Effects of climate change on grape yield and
Total soluble sugars quality were cultivar dependent. Generally, red Tempranillo had more vegetative growth and grape yield
Total polyphenol index than the white cultivar. Also, grape yield was less affected by the treatments than vegetative growth.
Drought, especially under elevated temperature, drastically reduced vegetative growth, bunch fresh and
dry weights in both cultivars. Interestingly, elevated CO2 attenuated these negative effects of drought.
The effects of climatic factors on yield were not associated with a worse water status of the vegetative
or reproductive organs. In red Tempranillo, the combination of elevated CO2 , elevated temperature and
drought reduced total polyphenol index (TPI), malic acid and increased color density, but did not modify
anthocyanin concentration. In white Tempranillo, the combined action of the three factors associated with
climate change modified only tartaric acid. In this latter cultivar, drought increased TPI under ambient
temperature, regardless of CO2 level, when compared with full-irrigated plants. In conclusion, climate
change-related factors (elevated CO2 , elevated temperature and water deficit) individually (especially
drought) and/or interacting affected to different extent red and white Tempranillo vegetative growth
and yield. Drought combined with elevated temperatures reduced grapevine performance, and elevated
CO2 mitigated such deleterious effect.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction (Collins et al., 2013). Due to its greenhouse effect and if CO2 emis-
sions continue at high level, models predict a temperature rise
CO2 is the most important anthropogenic greenhouse gas. between 1.8 and 4.0 ◦ C in the year 2100 (Collins et al., 2013).
Its atmospheric concentration has increased since pre-industrial Since 1850, the warmest years of global surface temperature have
period until nowadays from 280 to 389–400 ␮mol mol−1 air (ppm). occurred from 1995 and beyond. Furthermore, the assumptions
It is expected to rise to ca. 700 ppm at the end of this century foresee that climate change could reduce plant water availability
and could enhance the agricultural areas under drought, affecting
crop production (Collins et al., 2013). In Europe, the Mediterranean
area is particularly prone to aridness and potentially very vulner-
Abbreviations: ◦ Bx, ◦ Brix; DW, dry weight; FW, fresh weight; IPCC, intergovern- able to a temperature enhancement, along with more frequent
mental panel on climate change; T, ambient temperature; T+4 ◦ C, 4 ◦ C higher than extreme weather (Tubiello et al., 2000). Even if the rainfall lev-
ambient temperature; TGGs, temperature gradient greenhouses. els are unaffected, the risks of intense drought will increase due
∗ Corresponding author. Fax: +34 976 716145.
to the augmentation of the evaporative atmospheric demand as a
E-mail address: fmorales@eead.csic.es (F. Morales).

http://dx.doi.org/10.1016/j.agwat.2015.06.015
0378-3774/© 2015 Elsevier B.V. All rights reserved.
156 T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164

result of warming (Alcamo et al., 2007). For these reasons, crops Effects on berry weight were more severe when the above-
in the future, even some ecosystems, will face climate-related fac- mentioned stress factors were combined (McCarthy, 1997), and
tors such as elevated CO2 , elevated temperature and altered water changes in berry quality were more complex (Schultz, 2000). The
availability. individual effect of elevated CO2 stimulating grapevine production
In grapevine plants, a general response to elevated CO2 is was reduced or completely cancelled when elevated CO2 was com-
increased plant growth and yield (Bowes, 1993). Elevated CO2 may bined with elevated temperature (Bindi et al., 2001). With increases
influence all phases of berry growth and the accumulation of grape in the number of warm days during flowering and veraison and
constituents with potential effects on quality. When grown at ele- a reduced precipitation during maturation, Merlot and Cabernet
vated CO2 , acids and sugars were positively influenced in early Sauvignon would produce grapes with higher sugar to total acid
stages of grape development, although disappeared at maturity ratio and greater berry weights (Jones and Davis, 2000). In the
(Bindi et al., 2001). Furthermore, grape sugars, total anthocyanins case of Tempranillo, phenology was altered, berry malic acid and
and tannins (Gonçalves et al., 2009), and must color tonality and total anthocyanin concentrations were decreased and anthocyanin
intensity (Salazar-Parra et al., 2010), were not affected by elevated extractability was facilitated by climate change conditions (Salazar-
CO2 . Parra et al., 2010). A survey of the literature indicates that effects
Anthocyanin concentration is an important grapevine quality of single stress factors are sometimes contradictory (see above
trait that influences grape color. Temperature affects anthocyanin the case of water stress) and grapevine works focused on multi-
accumulation because several enzymes of this metabolic path- ple stress factors are scarce. In the present work, we were able to
way are temperature sensitive (Cohen et al., 2008; Yamane and investigate the effects of elevated CO2 , elevated temperature and
Shibayama, 2006). In experiments with controlled high tempera- water deficit, either acting individually or interacting among them,
ture, the anthocyanin contents were reduced (Kliewer, 1970; Mori on vegetative and reproductive growth, substrate and plant water
et al., 2007). Furthermore, temperature may decrease grape acid- status, and must quality in fruit-bearing cuttings of two grapevine
ity and especially can modulate grape malic acid concentration (Vitis vinifera L.) cultivars (red and white Tempranillo) in the TGGs.
(Spayd et al., 2002; Sweetman et al., 2014). As a consequence, grape One of the strengths of the present work lies in the assessment
malic acid levels are generally higher in cooler than warmer regions of three-way interactions between CO2 , temperature and water
(Blouin and Guimberteau, 2003). availability under fully controlled conditions using the TGGs.
Soil water availability influences grapevine vegetative growth.
Shoot growth stopping (Matthews and Anderson, 1989; Williams
2. Materials and methods
and Matthews, 1990), reduction in leaf size (Matthews et al.,
1987) and high leaf senescence (Kliewer and Weaver, 1971) are
2.1. Plant material
some of the reported symptoms. Grapevine yield and berry quality
are generally decreased by strong water stress (Myburgh, 2003).
Dormant cuttings of Vitis vinifera L. cv. red (accession T43, Clone
Between flowering and veraison, water stress reduces berry size
RJ-43) and white Tempranillo (accession CI-101 in “La Grajera”
due to restrictions in cell division (Matthews and Anderson, 1989;
germplasm bank, Rioja Government, Spain) were collected from an
McCarthy, 1997; Ojeda et al., 2002). Reduced berry size affects berry
experimental vineyard of the Institute of Sciences of Vine and Wine
quality indirectly via concentration of must phenolics, including
(ICVV) in Logroño (La Rioja, Spain). Tempranillo is a native red grape
anthocyanins (Castellarin et al., 2007; Matthews et al., 1990; Petrie
variety extensively cultivated in northern and central Spain. A new
et al., 2004). Some effects of water stress on berry quality reported
white berry somatic variant of Tempranillo (“white Tempranillo”)
in the literature are rather contradictory, showing (i) increases in
is the result of a natural mutation (chromosomal reorganizations
malic acid concentration (López et al., 2007), sugar content (Antolín
and deletions) from a single stem of red Tempranillo due to envi-
et al., 2006; Koundouras et al., 2006; Matthews et al., 1990; Ojeda
ronmental factors.
et al., 2002) and pH values (Bahar et al., 2011) that affect nega-
tively wine quality, (ii) decreases in malic acid concentration (De
Souza et al., 2005; Intrigliolo and Castel, 2009; Koundouras et al., 2.2. Growth conditions
2006; Salazar-Parra et al., 2010), berry volume, total soluble solids,
potassium and tartaric acid concentration (Bahar et al., 2011) or Cuttings of V. vinifera L. cv. red and white Tempranillo were
(iii) no change on malic acid concentration (Esteban et al., 1999). selected to obtain fruit-bearing cuttings according to Mullins
According to these results, the grape composition could be affected (1966) and modified by Santa María (2004). Indole butyric acid
both directly and indirectly, and positively or negatively depend- was applied to three-node cuttings, which were well covered with
ing on the degree as well as the duration of water stress. For all rock wool in a heat-bed (25–27 ◦ C) and kept moist in a cool room
these reasons, controlling CO2 concentration, temperature, and (5 ◦ C). One month later, the rooted-cuttings were planted in 0.8 L
water availability under greenhouse experiments could be a viable plastic pots containing a mixture of sand, perlite and vermiculite
approach. (1:1:1, v/v) and transferred to the greenhouse. When cuttings were
Climate change studies have been performed on different in inflorescence stage, they were transplanted with a mixture of
growth systems such as open top chambers (OTC), free air CO2 peat and perlite (2:1, v/v) in 2 L plastic pots. At fruit-set, plants were
enrichment (FACE) systems and temperature gradient greenhouses planted in 13 L plastic pots containing the peat-perlite mixture.
(TGG), among others (Morales et al., 2014). In the present work, we Only a single flowering stem was allowed to develop on each
used TGG. TGGs are close to function as growth chambers but with plant during growth, resulting in only one grape bunch per plant.
the advantages of the research-oriented greenhouses in order to The reason is that after bud break and appearance of the first
investigate effects of elevated CO2 , elevated temperature and water inflorescence, the latter grows at the expense of wood reserves
availability separately or in interaction. TGGs are a current way of exclusively and the first leaves have to be removed because at
attributing to one specific factor a physiological response, enabling this stage they compete for reserves. Allowing growing more
for more mechanistic researches on the causes of the physiological than one inflorescence per plant leads to all inflorescences to be
responses to future climate conditions (Morales et al., 2014). unviable (Mullins, 1966). Growth conditions in the greenhouse
Effects of climate change-related factors on grapevine (i.e., ele- were 26/15 ◦ C and 60/80% relative humidity (RH) (day/night) and
vated CO2 , elevated temperature and water deficit) have been a photoperiod of 15 h with natural daylight supplemented with
investigated individually (see above) but rarely their interactions. high-pressure metal halide lamps (OSRAM® , Augsburg, Germany),
T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164 157

Full irrigaon T
Ambient CO2
T+4ºC

Cyclicdrought T

T+4ºC
Greenhouse Temperature Gradient Greenhouse

Full irrigaon T

Elevated CO2 T+4ºC

Cyclicdrought T

T+4ºC

March April May June July August

Roong Bud Flowering Fruit set Veraison Maturity


break

Fig. 1. Experimental design. Red and white Tempranillo grapevine (Vitis vinifera) fruit-bearing cuttings were developed. Soil moisture sensors were placed into each pot.
Plants grew until fruit set under ambient conditions in the greenhouse. Then, they were transferred to four temperature gradient greenhouses (TGGs), two set at elevated
(ca. 700 ppm) and two at current (ca. 400 ppm) CO2 concentrations. The TGGs have modular design, where a gradient of temperature is created from module 1 of ambient
temperature to module 3 of ambient temperature +4 ◦ C (no experimental plants were set in module 2 with intermediate temperature, and therefore treatments used were
ambient temperature and ambient temperature +4 ◦ C). In each TGG, plants were divided into two groups, full irrigation and cyclic drought. Treatments were maintained
until maturity.

providing a minimum photosynthetic photon flux density (PPFD) of varieties (red and white Tempranillo), two CO2 levels (ambient
500 ␮mol m−2 s−1 at inflorescence level. Plants were irrigated with ca. 400 ppm, and elevated ca. 700 ppm), two temperature regimes
the nutrient solution proposed by Ollat et al. (1998) until harvest- (ambient and ambient +4 ◦ C) and two water availability conditions
ing. Plants grew under these conditions until fruit set, from March (full irrigation and cyclic drought) (Fig. 1).
to May 2013.
2.4. Water treatments
2.3. Temperature gradient greenhouses and experimental design
Soil moisture sensors (EC-5 Soil Moisture Sensors, Decagon
Treatments were applied in the TGGs, located in the Campus of Devices Inc., Pullman, WA, USA) were placed into the pots dur-
the University of Navarra (42◦ 48 9.486 N, 1◦ 40 1.5882 W) in Pam- ing the last transplant (at fruit set). Substrate water content was
plona (Navarra, Spain), from June to August 2013 (i.e., from fruit controlled until maturity. Plants under full irrigation were main-
set to maturity). TGGs have a design based on temperature gradi- tained at ca. 80% of the substrate field capacity (sensor value
ent tunnels (Rawson et al., 1995) aimed to investigate the effects between 30 and 40%, (m3 H2 O m−3 soil) × 100). Cyclic drought con-
of global environment change on plants. They were constructed sisted in withholding irrigation until plants showed visual signs of
with a modular design with 3 temperature modules (3.04 m long water deficit such as tendrils and leaves downwards (sensor value
each). CO2 can be injected inside increasing the air CO2 concen- between 10 and 0%, (m3 H2 O m−3 soil) × 100). When cyclic drought
tration as desired, a gradient of temperature is created along them plants had symptoms, they were irrigated with the same amount
(from module 1 of ambient temperature to module 3 of ambient of nutrients that received plants under full irrigation in that cycle.
temperature +4 ◦ C), and plants in pots growing inside can be sub- Irrigations (full irrigation and cyclic drought) were chosen follow-
jected to different levels of irrigation as a research greenhouse with ing our previous experience with grapevine fruit-bearing cuttings
high level of environmental control (Morales et al., 2014). Fans and and TGG conditions for the red cultivar (Salazar-Parra, 2011).
heaters were used to circulate air through the TGGs at required
speed and for maintaining a difference of 4 ◦ C between the two 2.5. Measurements
extreme modules. No experimental plants were set in module 2
with intermediate temperature, and therefore, treatments used Midrib length was measured in fully expanded leaves at matu-
were ambient temperature and ambient temperature +4 ◦ C. CO2 rity to calculate leaf area, according to the formula described by
concentration, temperature, relative humidity and radiation were Santa María (2004), as follows: Leaf area = [−4.65 × (2.23 × midrib
measured and/or controlled by an informatics system. length) + (0.96 × (midrib length)2 )]. Complete bunch of berries
Fig. 1 shows the experimental design. Pruning was used to con- (berries and rachis), samples of 1 cm2 leaf discs (taken from
trol vegetative growth until fruit set, allowing 4 leaves per plant to fully expanded leaves with a calibrated cork borer), leaves, peti-
grow. Plants transferred to the TGGs at fruit set were maintained oles, stems, and roots were harvested at maturity (defined at
with free vegetative growth. Plants were selected in the basis of 21–23◦ Brix), weighted (fresh weight, FW) and then oven-dried (dry
similar grape bunch size, avoiding changes in berry quality due weight, DW) at 80 ◦ C for 48 h. Leaf (from leaf discs), bunch, berry,
to sink demand and were divided into 16 treatments (10 plants and rachis water content was calculated as follows: (FW – DW)/DW
per treatment). Treatments were a combination of two grapevine (g H2 O g−1 DW). In the particular case of berries, they were frozen
158 T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164

in liquid nitrogen and stored at −80 ◦ C until analysis and before 45 Ambient CO – T A White – Full irrigation
40 White – Cyclic drought
being oven-dried. Red – Full irrigation
35
Samples of 30 frozen berries per biological replicate (each bio- 30
Red – Cyclic drought

logical replicate consisted of a pool of berries from three different 25


plants) were extracted and centrifuged for 10 min at 4302 × g 20
15
(Sorvall SS-34 rotor, 6000 rpm) and the supernatants used for
10
technological maturity determinations. Must total soluble solids 5
(◦ Brix) was estimated with a refractometer (Zuzi model n◦ 315, 0
Elevated CO – T
40
Digital ABBE). Must pH was measured with pH meter (Microph 35 B
Crison model 2000). The modified Rebelein’s method was used 30
for quantifying tartaric acid, determined spectrophotometrically 25

(m3 H2O m-3 Substrate) X 100


20

Substrate Water Content


(Model U-2001, Hitachi Instruments Inc., USA) at 530 nm from the 15
red coloration given by the vanadic acid (Rebelein, 1973; Vidal 10
5
and Blouin, 1978). l-Malic acid was measured using an enzy- 0
Ambient CO – T + 4ºC
matic method, the l-malic acid test kit (EnzytecTM l-Malic Acid, 40 C
35
Boehringer Manneheim/R-Biopharm).
30
A second sample of 30 berries per biological replicate (as already 25
mentioned, each biological replicate consisted of a pool of berries 20
15
from three different plants) was weighed and ground in blender.
10
Part of the homogenate was centrifuged for 10 min at 4302 × g (Sor- 5
vall SS-34 rotor, 6000 rpm) and the resultant supernatant used for 0
Elevated CO – T + 4ºC
40
the determination of color density. Color density was quantified 35 D
as the sum of the optical densities at 420, 520, and 620 nm, mea- 30
25
sured with 1 cm optical path (Glories, 1978). Phenolic maturity 20
was measured after homogenate maceration during 4 h follow- 15
10
ing the Glories’ method (Glories and Augustin, 1993). In brief, 5
0 4 8 12 16 20 24 28 32 36 40 44 48 52 56
homogenate and buffer were mixed 1:1 by volume. The buffer
was set at pH 1, extracting total anthocyanins. Subsequently, the Days of treatment
macerated samples were centrifuged for 10 min at 4302 × g (Sor-
vall SS-34 rotor, 6000 rpm), and the supernatants used for the Fig. 2. Pot substrate water content from red (black symbols) and white (grey
symbols) Tempranillo grapevine (Vitis vinifera) grown under different CO2 concen-
following determinations. Total anthocyanin concentration was
trations (ambient, A and C or 700 ppm CO2, B and D) at ambient temperature (A
determined spectrophotometrically according to Ribéreau-Gayon and B) or ambient temperature +4 ◦ C (C and D) and under full irrigation (squares) or
and Stronestreet (1965). Total polyphenol index (TPI) was mea- cyclic drought (circles) measured every day until maturity. Each point is the mean
sured in the supernatant obtained after maceration at pH 3.2 of 3 and 7 measurements in full irrigation and cyclic drought pots, respectively.
(diluted 100 times with distilled water) at 280 nm according to
Ribéreau-Gayon et al. (1972).
strate water content than pots of the white cultivar (Fig. 2C and
D). Cyclic drought plants were grown withholding irrigation until
2.6. Data analysis plants showed visual signs of water deficit, and the sensors read-
ings were between 10 and 0% ((m3 H2 O m−3 soil) × 100) (Fig. 2).
Data were analyzed using two-way ANOVA. Means were com- When cyclic drought plants had symptoms, they were irrigated
pared by Fisher’s least significant difference (LSD) post-hoc test. with nutrient solution. Thus, 8 cycles of drought were used from
Results were considered statistically significant at P < 0.05. Data are fruit set to maturity (Fig. 2). In the first drought cycle, plants had
presented as means, or means ± standard error. Statistical analysis visual signs of water deficit when they had ca. 20–10% of substrate
was carried out using XLStat 7.5.2 Pro® statistical software (XLStat, water content, and spent more time than the other cycles (Fig. 2).
Paris, France), a complete analysis and statistical application adding The substrate water content required for showing drought symp-
soft for Microsoft Excel® . toms continuously decreased along the time (Fig. 2). During the first
half of the experiment (ca. 1 month), when both cyclic drought-
treated red and white Tempranillo showed signs of water stress,
3. Results and discussion
pots of red Tempranillo had less substrate water content than those
of the white ones, when ambient temperature (irrespective of CO2
3.1. Climate change and substrate water status at maturity in V.
conditions) (Fig. 2A and B) or when elevated temperature com-
vinifera cv. red and white Tempranillo
bined with elevated CO2 (Fig. 2D) were used for growing the plants.
Therefore, under most experimental conditions, red Tempranillo
Upon irrigation, a thin water film surrounding individual soil
consumed more water than the white cultivar.
particles remains in soil pores by capillarity, while some amount
of water drains via gravitation. Substrate water capacity is defined
as the amount of water retained in the substrate after irrigation 3.2. Climate change and plant growth at maturity in V. vinifera
and drainage. In presence of water stress and along time, the water cv. red and white Tempranillo
in the soil pores is lost. Plants grown under full irrigation were
watered at ca. 80% of the substrate capacity, showing the sensors Under most experimental conditions, red Tempranillo had more
values ranging from 30 to 40% ((m3 H2 O m−3 soil) × 100) (Fig. 2). leaf area than the white one at maturity (Fig. 3A). Red Tempranillo
Substrate water content of full irrigation treatments combined plants grown in current conditions (ambient CO2 , ambient temper-
with ambient temperature under ambient (Fig. 2A) or elevated ature and full irrigation) had higher leaf area (Fig. 3A) and leaf DW
(Fig. 2B) CO2 showed similar values for both cultivars. In full irriga- (Table 1) than the white ones. This increased leaf growth of the
tion treatments combined with elevated temperature under both red cultivar could be consequence of higher water consumption
CO2 conditions, however, red Tempranillo pots had lower sub- and soil water depletion (Fig. 2). Elevated CO2 increased leaf area
T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164 159

25000
b
a Ambient CO2
Elevated CO2
A
20000

cm2 plant-1
Leaf Area c
15000 de
de d
ef
fg
10000 gh hi hi
ij jk jk
jk k
5000

0
a
ab B
300 abc

250 bcd
cde bcd cde
Bunch Fresh Weight

de de de
de def def
200 ef def
g plant-1

f
150

100

50

0
T T+4ºC T T+4ºC T T+4ºC T T+4ºC

Full irrigation Cyclic drought Full irrigation Cyclic drought

Tempranillo Red Tempranillo White

Fig. 3. Total leaf area (A) and bunch fresh weight (B) in red and white Tempranillo grapevine (Vitis vinifera) grown under different CO2 concentrations (ambient or 700 ppm
CO2 ; black and white bars, respectively), temperature regimes (ambient or ambient +4 ◦ C) and water availability (full irrigation or cyclic drought). Data were plotted as
mean ± standard error (n = 10). Different letters indicate significant differences among treatments (P < 0.05) based on LSD test.

Table 1
Leaf, petioles, stem, root, bunch, berry, and rachis dry weight (DW) in red and white Tempranillo grapevine (Vitis vinifera) grown under different CO2 concentrations (ambient
(A) or 700 ppm CO2 (E)), temperature regimes (ambient (T) or ambient +4 ◦ C (T + 4 ◦ C)) and water availability (full irrigation (FI) or cyclic drought (CD)). Data were plotted as
mean (n = 10). Different letters indicate significant differences among treatments (P < 0.05) based on LSD test.

Treatments Leaf Petioles Stem Root Bunch Berry Rachis


(g DW plant−1 ) (g DW plant−1 ) (g DW plant−1 ) (g DW plant−1 ) (g DW plant−1 ) (g DW plant−1 ) (g DW plant−1 )

Red FI T A CO2 132.9 bc 17.8 bc 123.5 bc 40.1 de 35.5 a 33 ab 2.6 cdefg


E CO2 151 b 20.8 a 146.9 a 78.7 bc 27.8 abcd 25.3 abcdef 2.9 bcdef
T + 4 ◦C
A CO2 130 bc 15.5 cd 121.5 c 90.1 a 33.2 ab 31 abc 3.1 bcd
E CO2 128.2 c 14.5 d 128.6 abc 88.2 ab 33.9 ab 30.7 abc 4a
CD T A CO2 37.8 hi 4.7 fg 32.8 fgh 22.9 gh 32 ab 29.4 abcd 3.3 b
E CO2 71.9 fg 8.2 e 58.3 e 41.7 de 31.1 abc 28.5 abcde 3.2 bc
T + 4 ◦C
A CO2 29.6 i 3.6 g 25.2 gh 16 h 23.1 cd 20.8 def 2.5 defg
E CO2 40.4 hi 4.6 fg 40.7 efg 45.8 de 36.7 a 33.7 a 2.7 bcdefg
White FI T A CO2 98.2 de 13.5 d 85.1 d 42.8 de 25.9 bcd 23.3 cdef 2.4 efg
E CO2 115.1 cd 13.2 d 80.7 d 46.9 d 27.3 bcd 24.6 bcdef 2.5 defg
T + 4 ◦C
A CO2 84.5 ef 12.8 d 83 d 40.4 de 21.8 de 18.7 fg 2.2 g
E CO2 197.6 a 19.7 ab 141.8 ab 68.8 c 27.8 bcd 23.8 cdef 3.2 bcd
CD T A CO2 39.5 hi 5.5 efg 35.6 fgh 21.9 gh 23.7 d 20.3 ef 2.6 cdefg
E CO2 58.6 gh 7 ef 47.1 ef 34.8 ef 33.8 ab 30.5 abc 2.5 defg
T + 4 ◦C
A CO2 29.1 i 3.4 g 20.6 h 15 h 13.6 e 11 g 2.3 fg
E CO2 42.3 hi 5.5 efg 35.8 fgh 27.6 fg 33.2 ab 30.3 abc 3 bcde

A CO2 , ambient CO2 ; E CO2 , elevated CO2 ; T, ambient temperature; T + 4 ◦ C, 4 ◦ C higher than ambient temperature; FI, full irrigation; CD, cyclic drought.

(Fig. 2A) and leaf DW (Table 1) when combined with ambient tem- area when combined with ambient temperature and full irrigation
perature under both full irrigation and drought conditions in the red conditions in white Tempranillo (Fig. 2A). In both cultivars, drought
Tempranillo (differences were not significant in the case of leaf DW decreased leaf area (Fig. 2A) and leaf DW (Table 1). In petioles
with full irrigation). Nevertheless, this CO2 effect was only observed (Table 1), changes in DW due to elevated CO2 , elevated temper-
in the full irrigation and elevated temperature-grown plants in ature and water availability were similar to those described for
white Tempranillo, both in leaf area (Fig. 2A) and leaf DW (Table 1). leaf DW. In full irrigation plants, stem DW of red Tempranillo was
Contrary to these results, elevated CO2 had negative effect on leaf higher than that of the white one, and only when elevated CO2 was
160 T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164

combined with increased temperature white Tempranillo reached white Tempranillo plants grown under elevated temperature and
stem DW values similar to those of red Tempranillo (Table 1). Stem full irrigation (Table 1).
DW was the highest when elevated CO2 was combined with ambi- At the beginning of the berry developmental stage, berry growth
ent temperature and full irrigation treatments in red Tempranillo, is very sensitive to water deficit and cannot be reversed by supple-
and with elevated temperature and full irrigation treatments in the mental irrigation during the following stages. Thus, negative effects
white cultivar (Table 1). Drought treatments decreased markedly of water stress on reproductive growth had been widely reported
stem DW, with respect to full irrigation plants (Table 1). Root DW (McCarthy, 1997; Petrie et al., 2004; Santesteban et al., 2011). How-
of red Tempranillo was generally higher than the root DW of the ever, applying cyclic drought after fruit set, grapevine production in
white cultivar, although exceptionally white Tempranillo grown both cultivars was not significantly affected when compared with
under elevated CO2 , increased temperature and full irrigation had full irrigation plants and the corresponding temperature and CO2
high root DW values (Table 1). In all water regimes, temperature conditions, except in red Tempranillo grown under elevated tem-
treatments and cultivars, elevated CO2 significantly increased root perature and CO2 where decreased in droughted plants (Fig. 3B).
DW, with the exception of elevated temperature and full irrigation With the latter exception, this may be caused by the fact that (i)
treatments of red Tempranillo, and ambient temperature and full in these experiments deficit irrigation did not result in changes
irrigation treatments of white Tempranillo (Table 1). Root DW was on berry size (Mekni, 2014) and (ii) elevated CO2 concentration
also impaired by drought, similarly to other plant growth parame- improves grapevine carbon assimilation (Salazar-Parra et al., 2012,
ters (Table 1). 2015) and may compensate partially for the negative effects of
Elevated CO2 increases grapevine photosynthesis (Salazar-Parra water stress (Williams and Matthews, 1990). Experiments of apply-
et al., 2012, 2015). An enhanced photosynthesis could explain the ing water deficit after veraison demonstrated that Shiraz berries
observed increases in plant growth, reflected in increases in leaf are insensitive to water deficit during the month before harvest
area, and in the DW of the different organs analyzed. However, this (McCarthy, 1997).
effect has only been observed in short-term experiments with ele- Grapevine quality depends on bunch and berry characteristics.
vated CO2 exposure in grapevine plants (Bindi et al., 1996; Schultz, Therefore, in order to avoid changes in berry quality due to sink
2000), whereas long-term exposure to elevated CO2 has negative demand, plants were initially selected homogeneously in all treat-
effects on net photosynthesis (Bindi et al., 1996). If plants have ments to have similar bunch size. This was aimed to avoid indirect
enough nutrients and water, elevated CO2 accelerates plant growth but positive effects on berry quality, through concentration of phe-
(Mullins et al., 1992). This effect may explain the increase in plant nolic compounds due to initial small berry size (Korkutal et al.,
growth of both red Tempranillo grown under ambient temperature 2011), such as those occurring when grapevine plants face water
and full irrigation, and white Tempranillo grown under elevated stress from flowering until veraison (Hardie and Considine, 1976;
temperature and full irrigation conditions (Fig. 3A). With water Matthews et al., 1987; McCarthy, 1997). It is well known that reduc-
scarcity, a reduction of plant growth was observed (Fig. 3A). How- tions in berry size can intensify the berry constituents (Antolín et al.,
ever, it has been reported that elevated CO2 compensates partially 2003; Koundouras et al., 2006; Niculcea et al., 2014; Ojeda et al.,
for the negative effects of water stress (Chaves and Pereira, 1992), 2002). As already discussed, our data of red and white Tempranillo
which could explain the increased growth of grapevine plants support the contention that reproductive growth is hardly sensi-
due to elevated CO2 in water-stressed red Tempranillo grown tive to water deficit after fruit set. However, when combined with
under ambient temperature (Fig. 3A). In the two Tempranillo culti- another climate change-related factor, there were positive effects.
vars tested, water deficit inhibited both shoot and root growths Thus, berry and bunch DW increased in red Tempranillo with ele-
(Table 1), in line with previous reports (Schultz and Matthew, vated CO2 , elevated temperature and water deficit and in white
1988). Furthermore, root growth was less sensitive to water deficit Tempranillo with elevated CO2 combined with water deficit in both
than shoot growth (Table 1) as also indicated by Williams and temperature regimes, when compared to ambient CO2 (Table 1).
Matthews (1990).
Grapevines are generally less sensitive to water stress in repro-
ductive than in vegetative growth (Korkutal et al., 2011; Williams 3.3. Climate change and vegetative and reproductive water
and Matthews, 1990). In the climate change conditions, water status at maturity in V. vinifera cv. red and white Tempranillo
stress is not the only factor that influences grapevine reproductive
growth. It also depends on grapevine cultivars and other environ- Leaf water content was similar in both red and white Tem-
mental factors (CO2 , temperature, etc.). Red Tempranillo produced pranillo in almost all treatments (Fig. 4A). Red Tempranillo grown
more bunch FW (Fig. 3B) and DW (Table 1) than the white culti- under elevated CO2 had decreased leaf water content, when com-
var, related somehow to higher water consumption and soil water bined with drought and ambient temperature condition (Fig. 4A).
depletion (Fig. 2). Elevated CO2 increased bunch FW of the red cul- An unlike behavior for leaf water content compared to leaf area
tivar in the full irrigation and elevated temperature treatments, should be noted. There were no changes in leaf water content in
whereas it did not change in the white Tempranillo (Fig. 3B). both cultivars, but leaf area increased for full irrigation (T + 4 ◦ C
On the contrary, elevated CO2 did not increase bunch FW in the scenario) with elevated CO2 when compared to the ambient one
cyclic drought and elevated temperature treatments of red Tem- in white but not in red Tempranillo (Fig. 3A).
pranillo, while it was increased in the white one (Fig. 3B). Elevated White Tempranillo plants grown under ambient CO2 , elevated
CO2 increased bunch DW when grown under drought in white temperature and drought had the highest bunch water content
Tempranillo, whereas in red Tempranillo the increases were only (Fig. 4B). Elevated CO2 had no effect on almost all treatments, except
noticed in the combination of drought and elevated temperature under elevated temperature with cyclic drought conditions in both
treatments (Table 1). Changes observed in bunch DW data were red and white Tempranillo where bunch water content decreased
mainly due to berry DW (Table 1), being the contribution of the (Fig. 4B).
rachis DW (Table 1) to bunch DW minor. Thus, when elevated CO2 Elevated CO2 significantly decreased white Tempranillo berry
favored berry DW, the effect was not observed in the rachis, except water content under drought condition in both temperature
in drought and elevated temperature-treated plants in white Tem- regimes (Fig. 4C). Furthermore, it increased in red Tempranillo
pranillo (Table 1). Contrary to this, when elevated CO2 had no effect plants grown under elevated temperature and full irrigation, and in
on berry DW, there were increases of rachis DW such as in red and white Tempranillo plants grown under ambient temperature and
full irrigation (Fig. 4C).
T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164 161

4.0 Ambient CO2 3.4. Climate change and grape berry quality at maturity in V.
g H2O g-1 DW
3.5 Elevated CO2 A vinifera cv. red and white Tempranillo
abc abcd a
Leaf WC

3.0 abc ab
abcde abcdeabcde
2.5 bcde abcde abcde abcde 3.4.1. Grape technological maturity
cde e de bcde Malic and tartaric acids are the main acid components of grape
2.0
berries. During the grape respiration, both acids are lost. As a
1.5
consequence of more respiration in hot climates, berry acidity
1.0
decreases (Pandell, 1999). Malic acid concentration decreased in
0.5
the elevated temperature treatment under full irrigation conditions
0.0 a and ambient CO2 concentration in red Tempranillo when com-
10 ab ab B pared to ambient temperature, but this decrease was not observed
g H2O g-1 DW

bc bc
Bunch WC

bcd bcd bc bcd in the white one (Table 2). The malic acid concentration in red
8 cde cde
de
de de de Tempranillo was significantly reduced under the climate change
6 e conditions (elevated temperature, elevated CO2 and drought),
4 when compared to the current situation (ambient temperature,
2 ambient CO2 , and full irrigation) (Table 2), in line with previous
results with the same cultivar (Salazar Parra et al., 2010). The exper-
0 ab a
ab ab ab iments of temperature influence on malic acid demonstrate that
abcd
10 bcde
abcde
abc abcde C elevated temperature stimulates the activity of the enzyme respon-
g H2O g-1 DW

e bcde de cde cde cde


Berry WC

8 sible for the malic acid degradation to glucose during ripening


(Blouin and Guimberteau, 2003). Also, in vitro phosphoenolpyru-
6
vate (PEP) carboxylase activity in immature grape berries had an
4 optimum of temperature at 38 ◦ C, whereas malic enzyme activity
2 increased with rising temperature (between 10 and 46 ◦ C) (Lakso
and Kliewer, 1975). Elevated CO2 had no effect on malic acid con-
0
abc a a centrations, with the exception of white Tempranillo plants grown
ab
4
abcde abcd
bcdef bcdef
D under drought and elevated temperature (Table 2). Under drought
ef
f ef cdef
g H2O g-1 DW

def ef f conditions, malic acid concentration was significantly higher in


Rachis WC

3 white than in red Tempranillo in the current temperature and cur-


g rent CO2 conditions (Table 2). Drought usually decreases malate
2
concentration in grapevine due to malate breakdown (Matthews
1 and Anderson, 1989), as it is reported here for red Tempranillo
under ambient temperature.
0 Under climate change conditions (elevated CO2 , elevated tem-
T T+4ºC T T+4ºC T T+4ºC T T+4ºC
perature and drought), tartaric acid concentration was significantly
Full irrigation Cyclic drought Full irrigation Cyclic drought increased in red Tempranillo with respect to the current condi-
tions, whereas it was decreased in the case of white Tempranillo
Tempranillo Red Tempranillo White (Table 2). Elevated CO2 had no effect on the tartaric acid concen-
tration under the current condition (ambient temperature and full
Fig. 4. Leaf (A), bunch (B), berry (C) and rachis (D) water content in red and white irrigation) in both cultivars but it decreased tartaric acid concen-
Tempranillo grapevine (Vitis vinifera) grown under different CO2 concentrations tration under elevated temperature in red Tempranillo, whereas it
(ambient or 700 ppm CO2 ; black and white bars respectively), temperature regimes
(ambient or ambient +4 ◦ C) and water availability (full irrigation or cyclic drought).
did not change in the case of white Tempranillo (Table 2). Cyclic
Data were plotted as mean ± standard error (n = 10, 5 for leaf water content). Dif- drought decreased tartaric acid concentration under ambient tem-
ferent letters indicate significant differences among treatments (P < 0.05) based on perature and elevated CO2 conditions in the red cultivar, whereas
LSD test. it was increased in the white cultivar (Table 2). Elevated temper-
ature, elevated CO2 and drought had the highest effect, increasing
the concentration in red Tempranillo (Table 2).
Water content of red Tempranillo rachis remained fairly con- Drought in combination with elevated temperature and under
stant under drought conditions (Fig. 4D). However, elevated CO2 current CO2 increased significantly must pH in red as well as
increased rachis water content when plants were grown under in white Tempranillo, when compared to full irrigation plants
ambient temperature and full irrigation (Fig. 4D). White Tem- (Table 2). The lowest values of must pH were found in plants grown
pranillo rachis water content did not change with elevated CO2 under cyclic drought combined with ambient temperature under
in the plants grown at ambient temperature and full irrigation, elevated CO2 in the white cultivar (Table 2).
whereas elevated CO2 decreased significantly rachis water content Bunches of all plants were harvested at maturity, defined as
under elevated temperature in both water regimes and increased >21–23 ◦ Brix in some berries samples used to choose harvest time.
it under ambient temperature with drought (Fig. 4D). However, the range of values found at maturity in the whole bunch
In an experiment with Thompson Seedless, Williams and of berries ranged from 19 to 24.7 ◦ Brix (Table 2). Cyclic drought
Matthews (1990) reported a linear relationship between soil water combined with ambient temperature increased significantly ◦ Brix
content and leaf water potential. However, the relationship of under elevated CO2 conditions both in red and white Tempranillo
soil water content with leaf water content is more complex and (Table 2). The highest values of ◦ Brix were found in plants grown
depends on cultivar. Irrespective of soil water content (Fig. 2), only under cyclic drought combined with ambient temperature under
red Tempranillo leaves from the elevated CO2 , ambient tempera- elevated CO2 in the red cultivar (Table 2). There were no more
ture and drought treatment decreased leaf water content (Fig. 4A). remarkable changes.
In the case of bunch, berry and rachis (reproductive growth; Grape quality largely depends on sugar/acid balance at maturity.
Table 1), the relationships with soil water content were even more Thus, sugar, acidity, and pH are important characteristics for identi-
complex. fying maturity. In the present study, all plants were harvested when
162 T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164

Table 2
Malic acid, tartaric acid, pH, total soluble solids (mainly sugars) (TSS), and total polyphenol index (TPI) in berries of red and white Tempranillo grapevine (Vitis vinifera) grown
under different CO2 concentrations (ambient (A) or 700 ppm CO2 (E)), temperature regimes (ambient (T) or ambient +4 ◦ C (T + 4 ◦ C)) and water availability (full irrigation (FI)
or cyclic drought (CD)). Data were plotted as mean (n = 3). Different letters indicate significant differences among treatments (P < 0.05) based on LSD test.

Treatments Malicacid Tartaricacid pH TSS TPI

(g L-1 ) (g L-1 ) (◦ Bx) (A280 nm)

Red FI T A CO2 4.04 ab 1.22 cdefg 4 cde 21.8 bcd 40.13 def
E CO2 4.11 ab 1.26 cde 3.9 de 21.1 cdef 28.8 fg

T+4 C A CO2 3.15 cde 1.56 b 4.1 bcd 21.6 bcd 41.47 cde
E CO2 3.43 bcde 1.03 gh 4 cde 21.3 bcde 42.33 cde
CD T A CO2 3.01 de 0.95 h 4.1 bcd 21.8 bcd 48.07 bcde
E CO2 2.95 e 1.04 fgh 4 cde 24.7 a 40.8 def
T + 4 ◦C A CO2 3.77 abcd 1.38 bc 4.5 a 23.3 ab 44.47 cde
E CO2 3.08 cde 1.8 a 4 cde 20.2 def 19.73 g
White FI T A CO2 3.86 abc 1.11 efgh 4 bcde 20.3 def 53.27 bc
E CO2 3.47 bcde 1.13 efgh 3.9 de 19 f 46.67 cde
T + 4 ◦C A CO2 3.57 bcde 1.14 defgh 4.2 b 20.1 def 44.33 cde
E CO2 3.77 abcd 1.23 cdef 4 bcde 21 cdef 37.27 ef
CD T A CO2 4.01 ab 0.99 h 4.1 bc 19 f 59.2 ab
E CO2 3.35 bcde 1.34 cd 3.8 e 23.2 abc 66.93 a
T + 4 ◦C A CO2 4.4 a 0.62 i 4.6 a 19.3 ef 49.53 bcd
E CO2 3.14 cde 0.73 i 4.1 bcd 21.8 bcd 45.67 cde

TSS, total soluble solids; TPI, total polyphenol index; A CO2 , ambient CO2 ; E CO2 , elevated CO2 ; T, ambient temperature; T + 4 ◦ C, 4 ◦ C higher than ambient temperature; FI,
full irrigation; CD, cyclic drought.

they reached 21–23 ◦ Brix, as a criterium of full maturity. Accord- This result highlights the importance of the combined effect of
ingly, minor differences in the concentration of TSS (◦ Bx) were different environmental factors. Water deficit in the white culti-
observed (Table 2), in agreement with previous reports (Ojeda et al., var increased TPI under ambient temperature and elevated CO2 ,
2002; Williams and Matthews, 1990). The pH is closely related to in line with Petrie et al. (2004). Nevertheless, when water deficit
total acidity and both of them measure acid levels within the berry was combined with elevated temperature and elevated CO2 did
juice (Pandell, 1999). Grape berries from a hot environment are not influence TPI (Table 2). This demonstrates the cancelation of
likely to have lower acidity than berries from a cool environment, the effects of some environmental factors by others, in this case
ascribed principally to temperature (Sadras and Moran, 2012). As temperature. This reduction or cancelation of the effect of one envi-
mentioned above, the berry juice pH was significantly rose when ronmental factor, such as CO2 , by others on grape yield and quality
red and white cultivars were grown under ambient CO2 , elevated was previously reported by Bindi et al. (2001), highlighting the
temperature and water deficit, but the interaction with elevated importance of investigating multiple environmental factors, and
CO2 inhibited the increase of must pH (Table 2). This increase in their interactions, in grapevine studies.
pH, among other features, may have direct influence on wine chem- Anthocyanins are flavonoids compounds responsible for red,
istry but elevated CO2 can stabilize must pH, avoiding changes on purple, and blue pigmentation of the grape berries. Berry coloration
the wine chemistry. is influenced by anthocyanin composition, and this is why it is
an important grapevine quality in red wine (Kuhn et al., 2013).
3.4.2. Grape phenolic maturity While red grapevines synthesize anthocyanins via the flavonoid
Generally, total polyphenol index (TPI) was significantly higher pathway in the berry skin, white grapevines have lacked off them
in white Tempranillo than in the red cultivar (Table 2). Shading and due to a multi-allelic mutation (Walker et al., 2007). Elevated CO2 ,
temperature of the bunch could influence grape quality, and also it irrespective of temperature, increased the concentration of total
is well known that several grape constituents are depended on light anthocyanins in the red cultivar under full irrigation conditions
and temperature (Blouin and Guimberteau, 2003; Kliewer, 1970; (Table 3), which increased color density in the case of elevated tem-
Lakso and Kliewer, 1975; Mori et al., 2007). Individually, effects perature (Table 3). However, water deficit decreased anthocyanins
of elevated CO2 , elevated temperature or drought had no influ- concentrations (irrespective of temperature) and color density (at
ence on TPI in red Tempranillo, whereas the combination of these ambient temperature) of grapes ripened under elevated CO2 condi-
three factors significantly reduced TPI in this cultivar (Table 2). tions, thus cancelling the effect of CO2 in this quality trait. Therefore,

Table 3
Total anthocyanins and color density in berries of red Tempranillo grapevine (Vitis vinifera) grown under different CO2 concentrations (ambient (A) or 700 ppm CO2 (E)),
temperature regimes (ambient (T) or ambient +4 ◦ C (T + 4 ◦ C)) and water availability (full irrigation (FI) or cyclic drought (CD)). Data were plotted as mean (n = 3). Different
letters indicate significant differences among treatments (P < 0.05) based on LSD test.

Treatments Total anthocyanins Color density


−1
(mg L ) (Sum of A420,520,620 nm )

Red FI T A CO2 274.95 b 1.38 cd


E CO2 406.88 a 1.3 bc
T + 4 ◦C A CO2 334.45 ab 0.96 d
E CO2 414.75 a 2.62 a
CD T A CO2 287.29 b 1.95 bc
E CO2 303.43 b 1.37 cd
T + 4 ◦C A CO2 344.75 ab 1.85 bc
E CO2 267.17 b 2.11 ab

A CO2 , ambient CO2 ; E CO2 , elevated CO2 ; T, ambient temperature; T + 4 ◦ C, 4 ◦ C higher than ambient temperature; FI, full irrigation; CD, cyclic drought.
T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164 163

the combination of elevated CO2 , temperature and drought (climate Antolín, M.C., Ayari, M., Sánchez-Díaz, M., 2006. Effects of partial rootzone drying
change conditions) did not modify the concentration of antho- on yield, ripening and berry ABA in potted Tempranillo grapevines with split
roots. Aust. J. Grape Wine Res. 12, 13–20.
cyanins (Table 3), although increased color density (Table 3). Our Antolín, M.C., Baigorri, H., De Luis, I., Aguirrezabal, F., Geny, L., Broquedis, M.,
data agree with previous reports showing lower anthocyanins Sánchez-Díaz, M., 2003. ABA during reproductive development in
concentrations in non-irrigated than in those subjected to conven- non-irrigated grapevines (Vitis vinifera L. cv. Tempranillo). Aust. J. Grape Wine
Res. 9, 169–176.
tional (or regulated deficit) irrigation in red Tempranillo grapevines Bahar, E., Carbonneau, A., Korkutal, I., 2011. The effect of extreme water stress on
(Kuhn et al., 2013; Zarrouk et al., 2012). leaf drying limits and possibilities of recovering in three grapevine (Vitis
vinifera L.) cultivars. Afr. J. Agric. Res. 6 (5), 1151–1160.
Bindi, M., Fibbi, L., Gozzini, B., Orlandini, S., Miglietta, F., 1996. Modelling the
4. Conclusions impact of future climate change scenarios on yield and yield variability of
grapevine. Clim. Res. 7, 213–224.
Bindi, M., Fibbi, L., Miglietta, F., 2001. Free Air CO2 Enrichment (FACE) of grapevine
Grape yield and quality effects of climate change conditions (Vitis vinifera L.): II growth and quality of grape and wine in response to
(elevated CO2 , elevated temperature and water deficit) were culti- elevated CO2 concentrations. Eur. J. Agron. 14, 145–155.
var dependent, affecting different traits. Red Tempranillo produced Blouin, J., Guimberteau, G., 2003. Maduración. In: Maduración y madurez de la uva,
Mundi-Prensa Libros, pp. 32–40.
more leaf area and had more yield and lower levels of total polyphe- Bowes, G., 1993. Facing the inevitable – plants and increasing atmospheric CO2 .
nol index than white Tempranillo. This increased growth and Annu. Rev. Plant Physiol. Plant Mol. Biol. 44, 309–332.
production of the red cultivar could be a consequence of higher Castellarin, S.D., Pfeiffer, A., Sivilotti, P., Degan, M., Peterlunger, E., Di Gaspero, G.,
2007. Transcriptional regulation of anthocyanins biosynthesis in ripening fruits
water consumption and soil water depletion, which was more of grapevine under seasonal water deficit. Plant Cell Environ. 30, 1381–1399.
evident under elevated than under ambient temperature. Under Chaves, M.M., Pereira, J.S., 1992. Water stress, CO2 and climate change. J. Exp. Bot.
drought, the lower water content of the substrate decreased leaf 43 (8), 1131–1139.
Cohen, S.D., Tarara, J.M., Kennedy, J.A., 2008. Assessing the impact of temperature
area in both cultivars of all treatments. Nevertheless, leaf water on grape phenolic metabolism. Anal. Chim. Acta 621, 57–67.
content showed generally no remarkable differences in both culti- Collins, M.R., Knutti, R., Arblaster, J., Dufresne, J.-L., Fichefet, T., Friedlingstein, P.,
vars in any of the treatments. It can be, therefore, concluded that Gao, X., Gutowski, W.J., Johns, T., Krinner, G., Shongwe, M., Tebaldi, C., Weaver,
A.J., Wehner, M., 2013. Long-term climate change: projections, commitments
the water availability-mediated decreased growth was due to an and irreversibility. In: Stocker, T.F., Qin, D., Plattner, K., Tignor, M., Allen, S.K.,
inhibition of leaf area growth but not to a decrease of the water Boschung, J., Nauels, A., Xia, Y., Bex, V., Midgley, P.M. (Eds.), Climate Change
status of the leaves. Elevated temperature also modulated growth, 2013: the Physical Science Basis. Contribution of Working Group I to the Fifth
Assessment Report of the Intergovernmental Panel on Climate Change.
reducing vegetative growth and leaf area in both cultivars of all
Cambridge University Press, Cambridge, United Kingdom and New York, NY,
treatments. In both red and white Tempranillo, climate change USA.
conditions did not modify significantly grape yield. In the white De Souza, C.R., Maroco, J.P., Dos Santos, T.P., Rodrigues, M.L., Lopes, C.M., Pereira,
cultivar, elevated CO2 compensated the decrease in bunch weight J.S., Chaves, M.M., 2005. Grape berry metabolism in field-grown grapevines
exposed to different irrigation strategies. Vitis 44, 103–109.
induced by elevated temperature and water deficit. In red Tem- Esteban, M.A., Villanueva, M.J., Lissarrague, J.R., 1999. Effects of irrigation on
pranillo: (i) elevated temperature in combination with drought and changes in berry composition of Tempranillo during maturation. Sugars,
ambient CO2 increased pH, whereas elevated CO2 increased antho- organic acids, and mineral elements. Am. J. Enol. Viticult. 50, 418–434.
Glories, Y., 1978. Recherches sur la matière colorante des vins rouges. In: Thèse de
cyanins irrespective of temperature and color density at elevated Doctorat en Sciences. Université de Bordeaux II, France.
temperature under well irrigated conditions, (ii) grapes ripened Glories, Y., Augustin, M., 1993. Maturité phénolique du raisin, conséquences
under climate change conditions had higher levels of tartaric acid technologiques: applications aux millésimes 1991 et 1992. Actes du Colloque:
Journée technique du CIVB. 21 Janvier, Bordeaux, 56.
and decreased malic acid. In white Tempranillo, elevated tempera- Gonçalves, B., Falcon, V., Moutinho-Pereira, J., Bacelar, E., Peixoto, F., Correia, C.,
ture combined with water deficit and ambient CO2 increased pH of 2009. Effects of elevated CO2 on grapevine (Vitis vinifera L.): volatile
berry juice, whereas the combination of the three factors studied composition, phenolic content, and in vitro antioxidant activity of red wines. J.
Agric. Food Chem. 57, 265–273.
only decreased tartaric acid concentration in this cultivar. All these Hardie, W.J., Considine, J.A., 1976. Response of grapes to water-deficit stress in
results underline the usefulness of investigating multiple environ- particular stages of development. Am. J. Enol. Viticult. 27, 55–61.
mental factors, including climate change-related factors and their Intrigliolo, D.S., Castel, J.R., 2009. Response of Vitis vinifera cv. ‘Tempranillo’ to
partial rootzone drying in the field: water relations, growth, yield and fruit and
interactions, in grapevine studies. In summary, red Tempranillo
wine quality. Agric. Water Manage. 96, 282–292.
was more productive than the white cultivar as a consequence of Jones, G.V., Davis, R.E., 2000. Climate influences on grapevine phenology, grape
higher water consumption and soil water depletion. Water deficit composition, and wine production and quality for Bordeaux, France. Am. J.
combined with elevated temperatures were the main factors reduc- Enol. Viticult. 51, 249–261.
Kliewer, W., 1970. Effects of day temperature and light intensity on coloration of
ing grapevine performance, and CO2 alleviated such deleterious Vitis vinifera L. grapes. J. Am. Soc. Hortic. Sci. 95, 693–697.
effect. Kliewer, W.M., Weaver, R.J., 1971. Effect of crop level and leaf area on growth,
composition and coloration of Tokay grapes. Am. J. Enol. Viticult. 22, 172.
Korkutal, I., Bahar, E., Carbonneau, A., 2011. Growth and yield responses of cv.
Acknowledgements Merlot (Vitis vinifera L.) to early water stress. Afr. J. Agric. Res. 6 (29),
6281–6288.
Koundouras, S., Marinos, V., Gkoulioti, A., Kotseridis, Y., Van Leeuwen, C., 2006.
Authors thanks Asociación de Amigos de la Universidad de Influence of vineyard location and vine water status on fruit maturation of
Navarra (T. Kizildeniz PhD Grant), IAMZ-CIHEAM (I. Mekni Master non-irrigated cv. Agiorgitiko (Vitis vinifera L.). Effects of wine phenolic and
aroma components. J. Agric. Food Chem. 54, 5077–5086.
Grant), European project (INNOVINE Call FP7-KBBE-2011-6, Pro- Kuhn, N., Le Guan Dai, Z.W., Wu, B.H., Lauvergeat, V., Gomès, E., Li, S.H., Godoy, F.,
posal N◦ 311775), Spanish Ministry of Economy and Innovation Arce-Johnson, P., Delrot, S., 2013. Berry ripening: recently heard through the
(MINECO BFU2011-26989) and Aragón Government (A03 Research grapevine. J. Exp. Bot. 65, 4543–4559.
Lakso, A.N., Kliewer, W.M., 1975. The influence of temperature on malic acid
Group) for financial support. Authors acknowledge the great tech-
metabolism in grape berries: I. Enzyme responses. Plant Physiol. 56,
nical work of M. Oyarzun, A. Urdiain, and M. Bettoni. 370–372.
López, M.I., Sánchez, M.T., Díaz, A., Ramírez, P., Morales, J., 2007. Influence of a
deficit irrigation regime during ripening on berry composition in grapevines
References (Vitis vinifera L.) grown in semi-arid areas. Int. J. Food Sci. Nutr. 58, 491–507.
Matthews, M.A., Anderson, M.M., 1989. Reproductive development in grape (Vitis
Alcamo, J., Moreno, J.M., Nováky, B., Bindi, M., Corobov, R., Devoy, R.J.N., vinifera L.): responses to seasonal water deficits. Am. J. Enol. Viticult. 40, 52–60.
Giannakopoulos, C., Martin, E., Olesen, J.E., Shvidenko, A., 2007. Climate change Matthews, M.A., Anderson, M.M., Schultz, H.R., 1987. Phenologic and growth
impacts, adaptation and vulnerability. In: Parry, M.L., Canziani, O.F., Palutikof, responses to early and late season water deficits in Cabernet franc. Vitis 26,
J.P., van der Linden, P.J., Hanson, C.E. (Eds.), Contribution of Working Group II 147–160.
to the Fourth Assessment Report of the Intergovernmental Panel on Climate Matthews, M.A., Ishii, R., Anderson, M.M., O’Mahomy, M., 1990. Dependence of
Change. Cambridge University Press, Cambridge, UK, pp. 541–580. wine sensory attributes on vine water status. J. Sci. Food Agric. 51, 321–335.
164 T. Kizildeniz et al. / Agricultural Water Management 159 (2015) 155–164

McCarthy, M.G., 1997. The effect of transient water deficit on berry development of de CO2 elevado, estrés hídrico y temperatura elevada. In: PhD Thesis.
cv. Shiraz (Vitis vinifera L.). Aust. J. Grape Wines Res. 3, 102–108. University of Navarra, Spain.
Mekni, I., 2014. Grapevine and climate change: response of two varieties of Salazar-Parra, C., Aguirreolea, J., Sánchez-Díaz, M., Irigoyen, J.J., Morales, F., 2010.
Tempranillo (read and white). In: Master Thesis. IAMZ-University of Lleida, Effects of climate change scenarios on Tempranillo grapevine (Vitis vinifera L.)
Spain. ripening: response to a combination of elevated CO2 and temperature, and
Morales, F., Pascual, I., Sánchez-Díaz, M., Aguirreolea, J., Irigoyen, J.J., Goicoechea, moderate drought. Plant Soil 337, 179–191.
N., Antolín, M.C., Oyarzun, M., Urdiain, A., 2014. Methodological advances: Salazar-Parra, C., Aguirreolea, J., Sánchez-Díaz, M., Irigoyen, J.J., Morales, F., 2012.
using greenhouses to simulate climate change scenarios. Plant Sci. 226, Photosynthetic response of Tempranillo grapevine to climate change
30–40. scenarios. Ann. Appl. Biol. 161, 277–292.
Mori, K., Goto-Yamamoto, N., Kitayama, M., Hashizume, H., 2007. Loss of Salazar-Parra, C., Aranjuelo, I., Pascual, I., Erice, G., Sanz-Sáez, A., Aguirreolea, J.,
anthocyanins in red-wine grape under high temperature. J. Exp. Bot. 58, Sánchez-Díaz, M., Irigoyen, J.J., Araus, J.L., Morales, F., 2015. Carbon balance,
1935–1945. partitioning and photosynthetic acclimation in fruit-bearing grapevine (Vitis
Mullins, M.G., 1966. Test-plant for investigations of the physiology of fruiting in vinifera L. cv. Tempranillo) grown under simulated climate change (elevated
Vitis vinifera L. Nature 209, 419–420. CO2 , elevated temperature and moderate drought) scenarios in temperature
Mullins, M.G., Bouquet, A., Williams, L.E., 1992. Biology of the Grapevine. gradient greenhouses. J. Plant Physiol. 174, 97–109.
Cambridge University Press, New York. Santa María, E., 2004. Incidencia de Botrytis cinerea en relación con diferentes
Myburgh, P.A., 2003. Responses of Vitis vinifera L. cv. Sultanina to water deficits aspectos fisiológicos de la vid. In: PhD Thesis. University of Navarra, Spain.
during various pre- and post-harvest phases under semi-arid conditions. S. Afr. Santesteban, L.G., Miranda, C., Royo, J.B., 2011. Regulated deficit irrigation effects
J. Enol. Vitic. 24, 25–33. on growth yield, grape quality and individual anthocyanin composition in Vitis
Niculcea, M., López, J., Sánchez-Díaz, M., Antolín, M.C., 2014. Involvement of berry vinifera L. cv. Tempranillo. Agric. Water Manage. 98, 1171–1179.
hormonal content in the response to pre- and post-veraison water deficit in Schultz, H.R., 2000. Climate change and viticulture: a European perspective on
different grapevine (Vitis vinifera L.) cultivars. Aust. J. Grape Wine Res. 20, climatology, carbon dioxide and UV-B effects. Aust. J. Grape Wine Res. 6, 2–12.
281–291. Schultz, H., Matthew, M., 1988. Vegetative growth distribution during water deficit
Ollat, N., Geny, L., Soyer, J., 1998. Les boutures fructifères de vigne: validation dı́un in Vitis vinifera L. Aust. J. Plant Physiol. 15, 641–656.
modele dı́étude du développement de la physiologie de la vigne. I. Spayd, S.E., Tarara, J.M., Mee, D.L., Ferguson, J.C., 2002. Separation of sunlight and
Caractéristiques de lı́appareil végétatif. J. Int. Sci. Vigne Vin. 32, 1–9. temperature effects on the composition of Vitis vinifera cv. Merlot berries. Am.
Ojeda, H., Andary, C., Kraeva, E., Carbonneau, A., Deloire, A., 2002. Influence of pre- J. Enol. Vitic. 53, 171–182.
and postveraison water deficit on synthesis and concentration of skin phenolic Sweetman, C., Sadras, V.O., Hancock, R.D., Soole, K.L., Ford, C.M., 2014. Metabolic
compounds during berry growth of Vitis vinifera cv. Shiraz. Am. J. Enol. Vitic. effects of elevated temperature on organic acid degradation in ripening Vitis
53, 261–267. vinifera fruit. J. Exp. Bot. 65, 5975–5988.
Pandell, A.J., 1999. The Acidity of Wine (visited 05.08.14) http://www. Tubiello, F.N., Donatelli, M., Rosenzweig, C., Stockle, C.O., 2000. Effects of climate
wineperspective.com/the acidity of wine.htm change and elevated CO2 on cropping systems: models predictions at two
Petrie, P.R., Cooley, N.M., Clingeleffer, P.R., 2004. The effect of post-veraison water Italian locations. Eur. J. Agron. 13, 179–189.
deficit on yield components and maturation of irrigated Shiraz (Vitis vinifera L.) Vidal, M., Blouin, J., 1978. Dosage colorimetrique rapide de l’acide tartarique dans
in the current and following season. Aust. J. Grape Wine Res. 10, 203–215. les mouts et les vins. Rev. Fr. Oenol. 16, 39–46.
Rawson, H.M., Gifford, R.M., Condon, B.N., 1995. Temperature-gradient chambers Walker, A.R., Lee, E., Bogs, J., McDavid, D.A.J., Thomasand, M.R., Robinson, S.P.,
for research on global environment change. 1. Portable chambers for research 2007. White grapes arose through the mutation of two similar and adjacent
on short-stature vegetation. Plant Cell Environ. 18, 1048–1054. regulatory genes. Plant J. 49, 772–785.
Rebelein, H., 1973. Rapid quantitative determination of tartaric acid. Chem. Williams, L.E., Matthews, M.A., 1990. In: Stewart, B.A., Nielsen, D.R. (Eds.),
Mikrobiol. Technol. Lebensm. 2, 33–38. Grapevine, 30. Irrigation of Agricultural Crops. Agronomy, Madison, Wisconsin,
Ribéreau-Gayon, J., Peynaud, E., Sudraud, P., Ribéreau-Gayon, P., 1972. Traité USA, pp. 1019–1055.
d’Enologie sciences et techniques dutin. In: Tome I –Analyse et Control des Yamane, T., Shibayama, K., 2006. Effects of changes in the sensitivity to
Vins. Dunod, Paris, France. temperature on skin coloration in ‘Aki Queen’ grape berries. J. Jpn. Soc. Hortic.
Ribéreau-Gayon, J., Stronestreet, E., 1965. Le dosage des anthocyanes dans le vin Sci. 75, 458–462.
rouge. Bull. Soc. Chim. 9, 2649–2652. Zarrouk, O., Francisco, R., Pinto-Marijuan, M., Brossa, R., Santos, R.R., Pinheiro, C.,
Sadras, V.O., Moran, M.A., 2012. Elevated temperature decouples anthocyanins and Costa, J.M., Lopes, C., Chaves, M.M., 2012. Impact of irrigation regime on berry
sugars in berries of Shiraz and Cabernet Franc. Aust. J. Grape Wine Res. 18, development and flavonoids composition in Aragonez (Syn. Tempranillo)
115–122. grapevine. Agric. Water Manage. 114, 18–29.
Salazar-Parra, C., 2011. Vid y cambio climático. Estudio del proceso de maduración
de la baya en esquejes fructíferos de Tempranillo en respuesta a la interacción

You might also like