Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Environmental and Experimental Botany 138 (2017) 184–192

Contents lists available at ScienceDirect

Environmental and Experimental Botany


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

Research Paper

Disproportionate photosynthetic decline and inverse relationship


between constitutive and induced volatile emissions upon feeding of
Quercus robur leaves by large larvae of gypsy moth (Lymantria dispar)
Lucian Copolovicia,* , Andreea Pagb , Astrid Kännastec, Adina Bodescua , Daniel Tomescub ,
Dana Copolovicia , Maria-Loredana Sorand, Ülo Niinemetsc,e
a
Faculty of Food Engineering, Tourism and Environmental Protection, Research Center in Technical and Natural Sciences, “Aurel Vlaicu” University, 2 Elena
Dragoi, Arad 310330, Romania
b
Institute of Technical and Natural Sciences Research, Development of “Aurel Vlaicu” University, 2 Elena Dragoi, Arad 310330, Romania
c
Institute of Agricultural and Environmental Sciences, Estonian University of Life Sciences, Kreutzwaldi 1, Tartu 51014, Estonia
d
National Institute of Research and Development for Isotopic and Molecular Technologies, Cluj-Napoca 400293, Romania
e
Estonian Academy of Sciences, Kohtu 6, 10130 Tallinn, Estonia

A R T I C L E I N F O A B S T R A C T

Article history:
Received 7 January 2017 Gypsy moth (Lymantria dispar L., Lymantriinae) is a major pest of pedunculate oak (Quercus robur) forests
Received in revised form 21 March 2017 in Europe, but how its infections scale with foliage physiological characteristics, in particular with
Accepted 24 March 2017 photosynthesis rates and emissions of volatile organic compounds has not been studied. Differently from
Available online 27 March 2017 the majority of insect herbivores, large larvae of L. dispar rapidly consume leaf area, and can also bite
through tough tissues, including secondary and primary leaf veins. Given the rapid and devastating
Keywords: feeding responses, we hypothesized that infection of Q. robur leaves by L. dispar leads to disproportionate
Green leaf volatiles scaling of leaf photosynthesis and constitutive isoprene emissions with damaged leaf area, and to less
Induced emissions
prominent enhancements of induced volatile release. Leaves with 0% (control) to 50% of leaf area
Isoprene emission
removed by larvae were studied. Across this range of infection severity, all physiological characteristics
Large insect herbivores
Monoterpene emission were quantitatively correlated with the degree of damage, but all these traits changed disproportionately
Photosynthesis with the degree of damage. The net assimilation rate was reduced by almost 10-fold and constitutive
Quantitative responses isoprene emissions by more than 7-fold, whereas the emissions of green leaf volatiles, monoterpenes,
Volatile organic compounds methyl salicylate and the homoterpene (3E)-4,8-dimethy-1,3,7-nonatriene scaled negatively and almost
linearly with net assimilation rate through damage treatments. This study demonstrates that feeding by
large insect herbivores disproportionately alters photosynthetic rate and constitutive isoprene
emissions. Furthermore, the leaves have a surprisingly large capacity for enhancement of induced
emissions even when foliage photosynthetic function is severely impaired.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction Barbosa, 2016) and protection of plants against herbivore attacks


(Heil, 2014; Pichersky and Gershenzon, 2002;), against excess
In field environments, plants are frequently exposed to a temperatures (Becker et al., 2015; Possell and Loreto, 2013), and
multitude of abiotic and biotic stressors. To cope with environ- oxidative stress, e.g. that generated by ozone exposure (Loreto and
mental and biological stressors, more than 100,000 secondary Schnitzler, 2010; Possell and Loreto, 2013; Vickers et al., 2009).
chemical products are synthesized by plants and at least 1700 of While constitutive volatile emissions occur in only a limited
these are known to be volatile (Bauer et al., 1998; Copolovici and number of species (Fineschi et al., 2013), stress-driven volatile
Niinemets, 2016; Kesselmeier and Staudt, 1999; Pichersky and emissions can be induced by abiotic and biotic stresses in all plant
Gershenzon, 2002). These biogenic volatile organic compounds species (Copolovici and Niinemets, 2016; Harrison et al., 2013;
(BVOC) serve many functions such as pollinator attraction (Lucas- Niinemets, 2010). Key biotic stresses eliciting major volatile
emission responses are infestations by fungi, bacteria, herbivores
and insects (Copolovici and Niinemets, 2016; Niinemets et al.,
* Corresponding author.
2013). Different biotic stressors elicit the same major classes of
E-mail address: lucian.copolovici@uav.ro (L. Copolovici). volatile compounds, green leaf volatiles (such as C5 and C6

http://dx.doi.org/10.1016/j.envexpbot.2017.03.014
0098-8472/© 2017 Elsevier B.V. All rights reserved.
L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192 185

alcohols and aldehydes), ubiquitous (e.g., a-pinene, b-pinene, For example, in the Southeastern United States BVOC emissions
D-3-carene) and specific monoterpenes (e.g., b-ocimene, linalool), can have a significant influence on the total aerosol burden due to
sesquiterpenes (e.g. b-caryophyllene) and homoterpenes ((3E)- condensation on acidic sulfate seed aerosols and concomitant
4,8-dimethy-1,3,7-nonatriene (DMNT) and (3E,7E)-4,8,12-trime- growth of particles (Lee et al., 2012; Link et al., 2015). The current
thyltrideca-1,3,7,11-tetraene (TMTT)) (Blande et al., 2014; Copolo- estimates of BVOC emissions and their role in atmospheric
vici and Niinemets, 2016; Holopainen, 2011; Kleist et al., 2012). processes only consider constitutive emissions (Arneth et al.,
However, as comparative experiments among different stresses, 2011; Guenther, 2013; Guenther et al., 2012), but stress-dependent
e.g., among infested, mechanically damaged and jasmonic acid- elicitation of volatiles can significantly modify the overall volatile
treated plants (Giorgi et al., 2015), and among leaf wounding and blend and amount of volatiles released into the atmosphere (Hare,
darkening (Brilli et al., 2011), demonstrate, different biotic stresses 2011; Holopainen and Blande, 2013; Holopainen and Gershenzon,
result in different volatile emission fingerprints. 2010; Niinemets et al., 2010a). In particular, under major outbreaks
Stress-induced volatiles can directly protect plants against of feeding herbivores that regularly occur in nature (Abrams and
biotic stress by serving as repellents of herbivores (Lucas-Barbosa Orwig, 1996; Dwyer et al., 2004; Mattson and Haack, 1987) induced
et al., 2011) or inhibitors of fungal and bacterial growth (Schmidt volatiles can dominate the release of BVOC from vegetation over
et al., 2016). They can also confer indirect defense by attracting large areas, underscoring the importance of gaining a better
predators to their herbivore prey or oviposition host (Arimura knowledge of quantitative scaling of induced emissions with the
et al., 2000a; Dicke and Baldwin, 2010; Johnson and Gilbert, 2015; degree of herbivory damage.
Koski et al., 2015). For both functions, the composition of the Pedunculate oak (Quercus robur L.) is a widely distributed
emission blend as well as the total emission rates play important constitutively isoprene-emitting tree species that is one of the
roles, the first determining the specificity of the signal for repelling most economically important broad-leaved forest trees in Europe.
or attraction, and the second, determining the dispersal of the Oak forests currently exhibit declining productivity and crown
signal as well as its chemical and physiological activity. Herbivore- dieback in several sites throughout Europe. Various abiotic and
induced changes in emissions of volatiles have been shown in biotic factors, including herbivory by the larvae of phyllophagous
numerous papers (Dicke, 2016; Heil, 2014 for reviews), but only a insects have been shown to importantly contribute to the oak
few studies have focused on quantitative relationships between decline (Batos et al., 2014; Thomas et al., 2002; Tonioli et al., 2001).
the degree of damage and amount of compounds emitted (Bruce, The most important defoliating insects feeding on Q. robur are
2015; Copolovici et al., 2014a; Holopainen and Gershenzon, 2010; winter moth (Operophtera brumata L.), tortrix moth (Tortrix
Niinemets et al., 2013). Presence of correlations among the degree viridana L.) and European gypsy moth (Lymantria dispar dispar
of herbivory damage and emission of volatiles seems trivial, L.) (Thomas et al., 2002). Among these, L. dispar is a species with
however, it crucially depends on local and systemic emission large larvae that can grow to the size of 50–90 mm (Milanovic et al.,
responses. While local response in immediately impacted leaf 2014) and that are capable of consuming 10 cm2 leaf area per day
areas is induced rapidly, e.g., emissions of green leaves volatles per larva, including the second order veins and the terminal part of
(GLV) can occur within seconds due to constitutive activity of the mid-rib (Milanovic et al., 2014). In the current study, we used L.
lipoxygenases (Portillo-Estrada et al., 2015), induction of emissions dispar larvae as a model to characterize the influence of large insect
of mono- and sesquiterpenes that need de novo expression of herbivore on constitutive and induced volatile release in Q. robur.
responsible synthases is more time-consuming, from hours to days We tested the hypothesis that the physiological activity of Q. robur
(Copolovici et al., 2014a, 2011; Pazouki et al., 2016). Also, the leaves infected with the large herbivore L. dispar is quantitatively
induction response is quenched relatively rapidly, within a few associated with the degree of damage, in particular, that the
days upon relief of herbivory stress (Copolovici et al., 2014a, 2011). infection leads to a major decline in leaf photosynthetic activity
Thus, the whole leaf response is a complex amalgamate of local and and isoprene emissions and a modest increase in induced
systemic induction and quenching responses. This is biologically emissions.
highly relevant considering the huge diversity of herbivore impacts
plants encounter in the field. While the rate of leaf area 2. Materials and methods
consumption is low for small solitary herbivores, leaving plenty
of time for systemic responses in non-impacted leaf areas, 2.1. Plant material
simultaneous presence of many small herbivores and large solitary
herbivores can rapidly consume a major proportion of leaf area, The field measurements were performed in Lipova forest at
implying that the systemic emission response might not even Arad county, Romania (46 50 3000 N, 21 4103000 E) in May 2014. The
occur in the major part of the leaf. In addition, while small site supports a broad-leaved deciduous forest that is mainly
herbivores mainly consume the intercostal leaf parts, large dominated by Quercus robur (canopy height 4–5 m), whereas Q.
herbivores can also consume major veins and lead to catastrophic petraea, Alnus glutinosa, and Q. rubra are minor canopy compo-
dysfunction of leaf water-, nutrient- and carbohydrate-conducting nents. The forest expands more than 6300 ha and more than half of
networks (Sack et al., 2003, 2004). Infections by small herbivores the trees (51%) were infected at the time of the study. The infection
can lead to compensatory increases of leaf photosynthetic activity was spatially highly heterogeneous, and forest patches with
in remaining leaf parts, but the damage of major veins could mean infested (more than 50 different patches with most trees infected
that feeding by large herbivores can lead to disproportionate observed) and non-infested patches with almost all trees lacking
reductions in foliage physiological activity, including photosyn- herbivores were interspersed. The weather conditions at the time
thetic activity and constitutive isoprenoid emissions as well as of measurements were: average air temperature of 26  2  C,
hindered elicitation of induced emissions (Copolovici et al., 2014a, relative humidity of 62% and atmospheric pressure of 102.4 kPa.
2011). The plants of Q. robur included for measurements were 10–12 years
Quantitative understanding of stress-dependent elicitation of old and 4–5 m tall. At the time of the measurements, the length of
volatile organic compound emission is further relevant for large- L. dispar larvae was about 30–40 mm. We took all measurements
scale processes in biosphere-atmosphere system. This is because with control leaves from the clean plots with healthy trees. These
BVOCs affect atmospheric OH radical and O3 concentrations and control leaves were further checked for presence of eggs and small
participate in the formation of secondary organic aerosols (SOA) larvae, and discarded if there was evidence of biotic interactions.
(Shen et al., 2013; Zhang et al., 2015; Ziemann and Atkinson, 2012). As all experiments were performed in natural conditions, the past
186 L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192

and current larval damage as well as the actual number of 2.4. Leaf pigment analysis
herbivores that had been feeding in the leaf could not be
controlled. Therefore, we only report the relationships of leaf Pigment extraction was performed according to the method of
physiological traits with the degree of leaf damage. Opris et al. (2013) with minor modifications. Briefly, leaf samples of
4 cm2 were taken after leaf gas-exchange measurements and
2.2. Photosynthetic measurements volatile sampling and immediately frozen in liquid nitrogen. The
pigments were extracted in ice-cold 100% acetone with calcium
Foliage photosynthetic characteristics during larval feeding carbonate (Sigma-Aldrich, Steinheim, Germany), and centrifuged
were determined in the field with a portable gas exchange system with a Hettich centrifuge (Hettich 320 R Universal, Hettich GmbH,
GFS-3000 (Waltz, Effeltrich, Germany) as in Niinemets et al. Tuttlingen, Germany) at 0  C and 9500g for 3 min. Then the
(2010b), except for minor modifications in environmental con- supernatant was decanted, and the extraction was repeated till the
ditions as stated below. This system has an environmental- supernatant remained colorless, but at least three times. The
controlled cuvette with 8 cm2 window area and a full-window leaf extracts were pooled and brought to a final volume of 10 ml and
chamber fluorimeter for sample illumination and fluorescence then filtered through a 0.45 mm PTFE membrane filter (VWR
measurements. Each time, a leaf fully filling the cuvette area was International, Radnor, PA, USA). The obtained extracts were
enclosed and standard measurement conditions (light intensity of analyzed for carotenoid and chlorophyll pigments by a high
1000 nmol m2 s1, leaf temperature of 25  C, chamber air humidi- pressure liquid chromatograph (HPLC–MS Shimadzu 2010, Kyoto,
ty of 70%, and CO2 concentration of 385 mmol mol1) were Japan) equipped with a diode array detector (DAD) according to
established. The leaf was stabilized under the standard conditions Niinemets et al. (1998) using a Lichrosorb1 RP-18 column (length
until stomata opened and steady-state CO2 and water vapor 125 mm, inner diameter 4 mm, film thickness 5 mm; Hichrom, UK).
exchange rates were reached. Steady-state values of net assimila- The column temperature was maintained at 10  C and the flow rate
tion (A) and transpiration (E) rates, and stomatal conductance to at 1.5 ml min1. The solvents used for the chromatographic elution
water vapor (gs) were calculated according to von Caemmerer and consisted of ultra-pure water (A) and HPLC grade acetone (B)
Farquhar (1981). (Sigma-Aldrich, Steinheim, Germany). The chromatographic elu-
tion was started by isocratically running a mixture of 25% A and
2.3. Volatile sampling and GC–MS analyses 75% B for the first 7.5 min, followed by a 9.5 min linear gradient to
100% B, which was run isocratically for 3 min. Further, the eluent
Volatile organic compounds (VOC) were sampled via the outlet was changed to the initial composition of 25% A and 75% B by a
of the gas-exchange cuvette at a flow rate of 200 ml min1 for 2 min linear gradient. The HPLC was calibrated using commercially
20 min with a constant flow air sample pump 210-1003MTX (SKC available chlorophyll a, chlorophyll b, and b-carotene standards
Inc., Houston, TX, USA). The leaf chamber air was drawn through a (Sigma-Aldrich, Steinheim, Germany), and the calibration curves
multibed stainless steel cartridge (10.5 cm length, 4 mm inner were developed at corresponding spectral maxima (430 nm for
diameter, Supelco, Bellefonte, USA) filled with Carbotrap C 20/ chlorophyll a, and 455 nm for chlorophyll b and b-carotene).
40 mesh (0.2 g), Carbopack B 40/60 mesh (0.1 g) and Carbotrap X
20/40 mesh (0.1 g) adsorbents (Supelco, Bellefonte, USA) to 2.5. Estimation of the degree of leaf damage
quantitatively sample all volatiles in C5–C15 range (Kännaste
et al., 2014). Volatiles were also collected using L. dispar larvae The leaves were scanned at 200 dpi, and the total leaf area and
without plants in the laboratory conditions using a 3 l glass the leaf area damaged by L. dispar larvae were estimated with the
chambers with a flow rate of 2 l min1 similarly as in Copolovici “Leaf Area Measurement” software (www.plant-image-analysis.
et al. (2011). To estimate the background VOC concentrations org). Leaf dry mass was estimated after oven-drying at 70  C, and
(blank samples), air samples were taken from empty chambers leaf dry mass to leaf area was calculated. Foliage photosynthetic
before and after enclosure of leaves or larvae. rates, volatile emissions and pigment contents per unit leaf area
The adsorbent cartridges were analyzed for volatile lipoxyge- were estimated after correction for the consumed leaf area.
nase (LOX) pathway products (also called green leaf volatiles, GLV),
terpenes and methyl salicylate using a Shimadzu TD20 automated 2.6. Statistical analysis and data handling
cartridge desorber integrated with a Shimadzu 2010 Plus GC–MS
instrument (Shimadzu Corporation, Kyoto, Japan) according to the All measurements have been done in triplicates and data points
method of Toome et al. (2010) and Kännaste et al. (2014). Briefly, correspond to averages of three replicate leaves in each individual
for tube desorption, He purge flow was set at 40 ml min1, primary plant (SE). The data were analyzed by linear and non-linear
desorption temperature at 250  C, and primary desorption time regression analyses, and the herbivory treatment effect at a certain
was 6 min. The mass spectrometer was operated in electron- level of damage severity relative to non-infected leaves was also
impact (EI) mode at 70 eV, with the transfer line temperature set at tested by ANOVA. All statistical analyses were conducted with
240  C and ion-source temperature at 150  C. The identification of ORIGIN 10.0 (OriginLab Corporation, MA, USA) and the statistical
terpenes, green leaf volatiles and isoprene was done using NIST tests were considered significant at P < 0.05.
spectral library ver. 14 and authentic standards (Sigma-Aldrich,
Taufkirchen, Germany). The absolute concentrations of com- 3. Results
pounds were calculated based on an external authentic standard
consisting of known amount of VOCs as described in full detail in 3.1. Effects of herbivory on foliage photosynthetic characteristics
Kännaste et al. (2014). Briefly, 1 ml of calibration sample has been
injected into the multi-bed tube followed by passing a nitrogen gas Herbivore feeding by L. dispar reduced leaf net CO2 assimilation
at 300 ml min1 through the tube for 5 min in order to evaporate rate (A), and even a moderate feeding, ca. 10% of leaf area removed,
the solvent and trap the volatiles on the adsorbent. Finally, the tube resulted in ca. 12% reduction of net assimilation rate, i.e. resulting
has been analyzed in GC–MS using the same program as for the in values of about 10 mmol m2 s1 (Fig. 1a). Net assimilation rate
samples. The background (blank) VOC concentrations were decreased with further increases in insect feeding, reaching values
subtracted from the concentrations with leaf samples and volatile of 1–2 mmol m2 s1 in leaves with ca. 50% of leaf area eaten
emission rates were calculated according to Niinemets et al. (2011). (Fig. 1a).
L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192 187

30
a

Isoprene emission rate


25

(nmol m s )
-2 -1
20

15

10
r = 0.933

5
b
4

GLV emission rate


(nmol m s )
-2 -1
3

2 r = 0.795

6
c

Monoterpenes emission rate


(nmol m s )
4
-2 -1

0
Fig. 1. Foliage net assimilation rate (a), stomatal conductance to water vapor (b) and 0 10 20 30 40 50 60
intracellular CO2 concentration per unit projected leaf area in Quercus robur plants
in relation to the degree of damage by the larvae of the lymantriid moth Lymantria Leaf area damage (%)
dispar (percentage of leaf area consumed). Data points correspond to averages of
three replicate leaves in each individual plant (SE). Data were fitted by linear (a) Fig. 2. Emissions rates of isoprene (a), green leaf volatiles (b; GLV, volatiles of
and non-linear regression in the form y = abx (b). lipoxygenase pathway, LOX volatiles) and monoterpenes (c) from Q. robur leaves
with different degrees of feeding by L. dispar larvae (replicates and data
presentation as in Fig. 1). Total LOX product emission was calculated as the sum
Differently from A, the stomatal conductance to water vapor (gs) of emissions of 1-hexanol, (Z)-3-hexenol, (Z)-2-hexenal, and (Z)-3-hexenyl acetate
and the total monoterpene emission as the sum of emissions of a-pinene, b-pinene,
was only moderately affected by larval feeding (Fig. 1b). In fact, the
camphene, limonene, D-3-carene, p-cymene, and b-phellandrene. Data were fitted
average (SE) gs of 109  20 mmol m2 s1 for strongly damaged by non-linear regressions in the form of y = abx.
leaves with 30–35% leaf area eaten was not different from the
average gs in non-damaged leaves (141  6 mmol m2 s1; P = 0.18 induced emissions of ubiquitous monoterpenes (a-pinene, cam-
for the comparison among the means; Fig. 1b). Nevertheless, the phene, D-3-carene, limonene and b-phellandrene) and typical
regression analysis indicated that through the entire damage range stress-marker monoterpenes such as (E)-b-ocimene and linalool
of 0–50%, gs was reduced by ca. 30%. Given the smaller gs than A (Table 1). The total emission rate of monoterpenes increased with
response to herbivory, the intercellular CO2 concentration (Ci) the degree of leaf damage from close to zero level in control leaves
increased with increasing the degree of herbivory damage (Fig. 1c). to values as high as ca. 5.3 nmol m2 s1 in strongly infected leaves
(Fig. 2c). Herbivory feeding also led to low-level emissions of the
3.2. Elicitation of volatile emissions in herbivory-infected leaves benzenoid methyl salicylate (MeSA) and the homoterpene (3E)-
4,8-dimethyl-1,3,7-nonatriene (DMNT), whereas the emission
Constitutive isoprene emission in herbivore-fed leaves was rates increased curvilinearly with the percentage of leaf damage
reduced from 30.3  0.7 nmol m2 s1 (average  SE) in healthy to maximum values of 0.17 nmol m2 s1 for MeSA and 0.08 nmol
leaves to 4–5 nmol m2 s1 in heavily infected leaves (40–50% of m2 s1 for DMNT (Fig. 3).
leaf area eaten by insects), and a strong negative correlation
between isoprene emission rate and percentage of leaf area eaten 3.3. Responses of foliage chlorophyll and b-carotene contents to
was observed (Fig. 2a). Larval feeding induced emissions of green herbivory
leaf volatiles [GLV; primarily, (Z)-3-hexenol, (E)-2-hexenal, (Z)-3-
hexenyl acetate and 1-hexanol), Fig. 2b], that increased with the Both chlorophyll a and b contents decreased strongly with the
degree of damage, reaching very high values of 3.0–4.4 nmol m2 area eaten by the larvae, e.g., the average chlorophyll a content
s1 (sum of all GLV) in heavily eaten leaves (Fig. 2b). Herbivory also decreased from 231  2 mg m2 in non-infected leaves to 95  3
188 L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192

Table 1
Average  SD emission rates (nmol m2 s1) of different volatiles released from leaves of Q. robur in response to insect damage.

Compound Control Average (SD) degree of damage (%)

12.2  0.6 18.6  1.0 38.4  9.9


(Z)-3-hexenol nd 0.107  0.005 0.213  0.007 0.90  0.13
(E)-2-hexenal nd 0.100  0.003 0.320  0.010 0.82  0.17
(Z)-3-hexenyl acetate nd 0.054  0.007 0.080  0.010 0.223  0.023
1-hexanol nd 0.223  0.022 0.454  0.017 1.84  0.33
isoprene 30.3  3.2 19.1  3.4 14.0  1.7 3.9  1.4
apinene 0.0092  0.0012 0.622  0.011 0.848  0.027 1.58  0.11
camphene nd 0.033  0.002 0.043  0.001 0.111  0.010
D-3-carene 0.011  0.006 0.504  0.004 0.663  0.009 1.07  0.07
limonene 0.0094  0.003 0.305  0.032 0.412  0.008 1.02  0.05
b-phellandrene 0.0106  0.007 1.12  0.14 1.09  0.13 0.98  0.15
(E)-b-ocimene nd 0.052  0.001 0.098  0.001 0.297  0.044
linalool nd 0.029  0.001 0.053  0.002 0.128  0.024
DMNT nd 0.007  0.002 0.021  0.006 0.068  0.015
methyl salicylate nd 0.034  0.006 0.045  0.011 0.145  0.021

nd = not detected.

0.20 30
Emission rate (nmol m s )
-1

DMNT a

Isoprene emission rate


MeSA
-2

r = 0.927 25
0.15

(nmol m s )
-2 -1
20
0.10
15
r = 0.919
0.05 r = 0.931 10 P < 0.05

5
0.00
0 10 20 30 40 50 60 0
400
Leaf area damage (%) b
Chlorophyll a+b (mg m )
-2

Fig. 3. Emissions of the benzenoid methyl salicylate (MeSA) and the homoterpene
(3E)-4,8-dimethyl-1,3,7-nonatriene (DMNT) from leaves of Q. robur in relation to 300
the degree of herbivory by L. dispar larvae. Data presentation and replication as in
Fig. 1. The relationships were fitted by linear regressions.
200
mg m2 in leaves with 30–50% damage (Fig. 4). The chlorophyll a/b r = 0.516
ratio was on average 2.48  0.26 and did not depend on the degree P = 0.06
100
of leaf damage (r = 0.5, P > 0.05). b-Carotene content did not
significantly correlate with the degree of larval damage (Fig. 4).
0
3.4. Negative scaling among constitutive and induced isoprenoids 6
c
Monoterpenes emission rate

Through different damage severities, foliage net assimilation


rate scaled negatively with the emissions of induced volatiles
(nmol m s )

4
-2 -1

(Fig. 5a for total monoterpenes, 5b for isoprene and 5c for total


r = 0.953
P < 0.05
2

0
0 10 20
-2 -1
Net assimilation rate (nmol m s )
Fig. 5. Correlations of net assimilation rate (a), isoprene emission rate (b) and
chlorophyll (a + b) content (c) with monoterpene emission rate in Q. robur leaves
with different degrees of damage by L. dispar larvae (Figs. 1 a, 2 a,c and 4 for the
correlations of given traits with the degree of damage). Data were fitted by non-
linear regressions.

Fig. 4. Effects of feeding by the larvae of L. dispar on the contents of chlorophyll a


and b, and carotene in Q. robur leaves. Statistical replicates and data presentation as
in Fig. 1. Data were fitted by linear regressions.
L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192 189

chlorophylls, r = 0.83, P < 0.01 for GLV; r = 0.89, P < 0.01 for MeSA; activation of alternative carbon sources can have resulted in the
r = 0.87, P < 0.01 for DMNT, for linear regression used). Constitutive reduction of the pool size of the immediate isoprene precursor,
isoprene emissions were negatively correlated with induced dimethylallyl diphosphate (DMADP) in chloroplasts (Rasulov et al.,
emissions (r = 0.97, P < 0.001 for total monoterpenes; r = 0.88, 2011, 2009), thereby reducing the emission rate. In addition,
P < 0.01 for GLV; r = 0.93, P < 0.01 for MeSA; r = 0.91, P < 0.01 for simultaneous reduction of isoprene synthase activity and isoprene
DMNT). emission can indicate overall decreases in foliage primary
metabolism and constitutive isoprene synthesis in non-consumed
4. Discussion leaf areas. Such a reduction is supported by significant declines in
foliage chlorophyll contents in infected leaves (Fig. 4).
4.1. Changes in photosynthetic function in leaves infected by large On the other hand, chloroplastic isoprene and monoterpene
larvae syntheses rely on the same plastidic DMADP pool, but the in vivo
effective Michaelis-Menten constant for DMADP (Km) is much
Plant photosynthetic responses to biotic stresses such as smaller for monoterpenes than for isoprene (Rasulov et al., 2014).
herbivory can range from stimulating effects, no effect or even Lower Km for monoterpene synthesis implies that the competition
significant impairment (Attaran et al., 2014; Gog et al., 2005; for DMADP is one-sided, and thus, elicitation of monoterpene
Nabity et al., 2009, 2013; Roslin et al., 2006; Zhou et al., 2015). synthesis upon herbivory feeding, can also partly explain the
Especially, colonization by solitary small herbivores can lead to decline in isoprene emission rates (Fig. 5b).
compensatory enhancement of photosynthesis in remaining leaf
parts (Copolovici et al., 2014a, 2011; Delaney et al., 2008). In 4.3. Induction of green leaf volatile emissions in larval-eaten leaves
Quercus robur in our study, we observed a drastic, disproportionate
reduction in net assimilation rate (Fig. 1a), consistent with the As a result of an attack by insects, specific elicitor molecules are
hypothesis that herbivory by large larvae capable of biting through generated by chemical or physical damage to plant membranes
and consumption of also major veins can seriously impair (Heil, 2014). Our study demonstrated elicitation of all key stress-
photosynthesis. Yet, stomatal conductance was surprisingly little induced volatile compound classes, green leaf volatiles (GLV),
affected, seemingly inconsistent with this hypothesis (Fig. 1b). mono- and sesquiterpenes, homoterpenes and methyl salicylate in
However, the low change in calculated stomatal conductance herbivory-infected leaves (Table 1). GLV are synthesized in a
might be an artifact due to evaporation of water from free water process where free octadecanoid fatty acids (linoleic acid = 18:2
surfaces generated upon herbivory, as well as due to transient and linolenic acid = 18:3) are released from plant membranes by
increases in stomatal conductance upon relaxation of epidermal phospholipases. Upon release of these free fatty acids, lip-
tension due to damage and concomitant opening of stomata (so- oxygenases (LOX) then produce 9- or 13-hydroperoxylinoleic or
called Ivanov’s effect, Moldau et al., 1993). –linolenic acid or a mixture of both (Matsui, 2006). A hydroperox-
On the other hand, longer-term reductions in photosynthesis in ide lyase further catalyzes the breakdown of 13-hydroperoxylinole
herbivore-fed leaves have also been associated with impaired (n)ic acid to a C6-compound, (Z)-3-hexenal, and a C12-product (12-
electron transport rate (Nabity et al., 2013). In our study, the oxo-(Z)-9-dodecenoic acid). (Z)-3-Hexenal can further give rise to
chlorophyll content decreased with leaf damage severity (Fig. 4). (Z)-3-hexenol, (E)-2-hexenol, (E)-3-hexenol or (E)-2-hexenal in
Such a reduction has been observed in some studies looking at consequent reactions (Feussner and Wasternack, 2002; Matsui,
diffuse massive leaf infection, e.g. mustard (Brassica juncea) leaf 2006). The emission of green leaf volatiles is a reliable marker of
infection by phloem-sucking mustard aphid (Lipaphis erysimi) oxidative stress and membrane-level damage (Porta and Rocha-
(Rehman et al. (2014) and tomato (Solanum lycopersicum) infection Sosa, 2002). GLV are rapidly released upon herbivory feeding due
by cotton mealybug (Phenacoccus solenopsis) (Huang et al. (2013)), to constitutive activity of LOX, and their almost immediate release
but not necessarily upon infection by tissue-removing herbivores. has been typically associated with mechanical damage upon
However, damage of major veins and concomitant reduction in wounding (Matsui et al., 2012; Portillo-Estrada et al., 2015; Scala
water availability of isolated mesophyll areas can well lead to the et al., 2013). In our study, we observed a strong correlation
start of senescence processes (Munné-Bosch, 2007, 2008). Given between the emissions of green leaf volatiles and the percentage of
the reduction of leaf pigment content (Fig. 4), changes character- leaf area damaged (Fig. 2b). Similarly to our study, increases in the
istic to programmed cell death such as inhibition of photosynthetic emission of GLV with the degree of herbivore feeding were also
electron transport and reductions in the amount or activity of observed in experiments with Caberia pusaria feeding on grey alder
photosynthetic rate-limiting enzymes were also likely responsible (Alnus incana) (Copolovici et al., 2011) and in experiments with
for reduction in net-assimilation rate in larval-eaten leaves. Epirrita autumnata feeding on hybrid aspen (Populus tremula x P.
In contrast, b-carotene content was weakly affected by larval tremuloides) (Schaub et al., 2010). Given that in these studies and in
feeding (Fig. 4). Differently from chlorophylls, b-carotene primari- our study, the degree of damage quantified as the percentage of
ly functions as antioxidant and its sustained level might serve leaf area removed includes both fresh and somewhat older
protective function. Although carotenoids can be destroyed under damage, the question is how such a correlation can occur if GLV
severe abiotic stress conditions (Ashraf and Harris, 2013), release is exclusively associated with immediate damage. Howev-
carotenoids are maintained longer than leaf chlorophylls in leaf er, conversion of (Z)-3-hexenal to more reduced and less toxic
tissues though senescence Niinemets et al., 2012). volatiles can also occur in non-impacted leaf areas (Matsui et al.,
2012), and GLV release can continue for a certain period of time
4.2. Modification of constitutive isoprene emissions by herbivory after the immediate biotic impact (Copolovici et al., 2011; Jiang
et al., 2016). Thus, scaling of GLV emissions with the degree of
Concomitant reductions of net assimilation rates and constitu- damage (Fig. 2b) might be explained by the sustained GLV release
tive isoprene emissions as observed in herbivore-fed leaves in our from non-impacted leaf areas.
study (Fig. 1b) have been reported in several other herbivory There is evidence that GLV release is higher upon damage of
studies (Laothawornkitkul et al., 2008; Loivamaeki et al., 2008; veins than upon damage of intercostal tissues (Portillo-Estrada
Loreto et al., 2014) as well as in fungal-infected leaves (Copolovici et al., 2015). Interestingly, the increase of GLV release with the
et al., 2014b; Jiang et al., 2016). Such a simultaneous reduction degree of damage in leaves with minor to moderate damage of 5–
might indicate that limited plastidic carbon availability or delayed 20% was much less than in leaves with moderate to extensive
190 L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192

damage of 20–50% (Fig. 2b). Such a difference might reflect the complex interplay between jasmonate- and salicylate-dependent
circumstance that in leaves with 5–20% damage, there was limited signalling pathways upon herbivore infestations.
big vein severance by herbivores, while herbivores also bit through
major veins in leaves with extensive damage. This result is in a 5. Conclusions
marked contrast with the linear relationship of GLV release vs. leaf
area damage in the study with small herbivores Caberia pusaria The results of the current study highlight a major negative
(Copolovici et al., 2011) and Monsoma pulveratum (Copolovici et al., scaling of foliage photosynthetic rates and constitutive isoprene
2014a) that are incapable of biting through major veins. emissions, and concomitant increase in induced volatile emissions
upon feeding by large herbivore larvae (Fig. 1,2,5). While the
damage-dependent increase of green leaf volatiles was dispropor-
4.4. Elicitation of emissions of terpenoids and MeSA tionately greater in leaves with extensive degree of damage than in
moderately damaged leaves (Fig. 2b), emission rates of terpenoids
As a widespread consensus, isoprene and monoterpenes are leveled off at higher degrees of leaf damage (Fig. 2c). This suggests
thought to be synthesized via 2-C-methyl-D-erythritol 4-phos- that faster and more severe damage, especially major vein
phate (MEP) pathway in plastids (Copolovici et al., 2014a; Fineschi severance by large herbivores can much more strongly influence
et al., 2013; Vranova et al., 2012) and sesquiterpenes via foliage physiological activity than feeding by smaller herbivores.
mevalonate (MVA) pathway in cytosol (Lombard and Moreira, These contrasting herbivore responses need consideration in
2011), although there is recent evidence of possible monoterpene modeling herbivore elicited emissions in large-scale biosphere-
synthesis in cytosol depending on substrate availability (Pazouki atmosphere models.
and Niinemets, 2016). DMNT is synthesized from the sesquiter-
pene (E)-nerolidol likely in the cytosol (Baldwin et al., 2006; Tholl Acknowledgements
et al., 2011). Although present in different cellular compartments,
both isoprenoid synthesis pathways are upregulated upon Funding for this study has been provided by the Estonian
herbivory stress and several terpene synthase genes involved Ministry of Science and Education (institutional grant IUT-8-3), the
have already been identified (Arimura et al., 2000a, 2000b; European Commission through the European Regional Fund (the
Baldwin et al., 2006; Loreto et al., 2014). Thus, elicitation of Center of Excellence EcolChange), the European Research Council
emissions of mono-, sesqui-, and homoterpenes has been found in (advanced grant 322603, SIP-VOL+) and the European Commission
different deciduous trees under herbivore stress (Stam et al., 2014; and the Romanian Government (project POSCCE 621/2014). This
Zhu et al., 2014 for reviews). The interplay between different work was also supported by a grant of the Romanian National
terpene synthase pathways is still somewhat unclear, especially Authority for Scientific Research, CNCS – UEFISCDI (project
given that different compounds serve different ecological func- number PN-II-RU-TE-2011-3-0022). We thank the Forestry Direc-
tions. In Q. robur infected by the moth Tortrix viridana, Giraldo et al. torate Arad (Lipova branch) for all their support during the time of
(2012) demonstrated that the larvae were attracted to the plants samples collection.
releasing higher amounts of homoterpene DMNT and monoter-
pene (E)-b-ocimene, while sesquiterpenes a-farnesene and References
germacrene D acted as a repellent.
We observed that the monoterpene emission rate increased Abrams, M.D., Orwig, D.A., 1996. A 300-year history of disturbance and canopy
recruitment for co-occurring white pine and hemlock on the Allegheny Plateau,
with increasing leaf damage (Fig. 2c), indicating that the activity of
USA. J. Ecol. 84, 353–363.
monoterpene synthases increased upon herbivore feeding. In Arimura, G., Ozawa, R., Shimoda, T., Nishioka, T., Boland, W., Takabayashi, J., 2000a.
addition, as discussed above, monoterpene synthesis could have Herbivory-induced volatiles elicit defence genes in lima bean leaves. Nature
been further favored by greater competitive capacity for chloro- 406, 512–515.
Arimura, G.I., Tashiro, K., Kuhara, S., Nishioka, T.O.R., Takabayashi, J., 2000b. Gene
plastic DMADP compared with isoprene synthesis (Fig. 5b) and responses in bean leaves induced by herbivory and by herbivory-induced
inhibition of pigment synthesis as evident in the reduction in leaf volatiles. Biochem. Biophys. Res. Commun. 277, 305–310.
pigment content (Fig. 5c). Scaling of mono- and sesquiterpene Arneth, A., Schurgers, G., Lathière, J., Duhl, T., Beerling, D.J., Hewitt, C.N., Martin, M.,
Guenther, A., 2011. Global terrestrial isoprene emission models: sensitivity to
emissions with the degree of damage has been observed in several variability in climate and vegetation. Atmos. Chem. Phys. 11, 8037–8052.
experiments looking at lepidopteran larval feeding effects on Ashraf, M., Harris, P.J.C., 2013. Photosynthesis under stressful environments: an
volatile release (for example Copolovici et al., 2014a, 2011). As we overview. Photosynthetica 51, 163–190.
Attaran, E., Major, I.T., Cruz, J.A., Rosa, B.A., Koo, A.J.K., Chen, J., Kramer, D.M., He, S.Y.,
hypothesized, there was evidence of leveling off of monoterpene Howe, G.A., 2014. Temporal dynamics of growth and photosynthesis
and DMNT vs. damage severity relationships at higher severity of suppression in response to jasmonate signaling. Plant Physiol. 165, 1302–1314.
damage (Figs. 2 c, 3). In fact, the ratios of monoterpene to GLV Baldwin, I.T., Halitschke, R., Paschold, A., von Dahl, C.C., Preston, C.A., 2006. Volatile
signaling in plant–plant interactions: talking trees in the genomics era. Science
emissions and DMNT to GLV emissions decreased with increasing
311, 812–815.
the degree of damage, indicating that the induction response was Batos, B., Seslija Jovanovic, D., Miljkovic, D., 2014. Spatial and temporal variability of
relatively less prominent in more severely damaged leaves flowering in the pedunculate oak (Quercus robur L.). Sumarski List 138, 371–379.
Bauer, K., Garbe, D., Surburh, H., 1998. Flavors and fragrances, Ullmann's
compared with the immediate stress response (or with the rapidly
Encyclopedia of Industrial Chemistry. The CD-ROM Edition. 5th ed. Wiley-VCH,
induced stress response). This is different from infection by small Verlag, Berlin.
larvae (Copolovici et al., 2014a, 2011) or from infection by slowly Becker, C., Desneux, N., Monticelli, L., Fernandez, X., Michel, T., Lavoir, A.-V., 2015.
developing biotic stresses such as fungal infections (Copolovici Effects of abiotic factors on HIPV-mediated interactions between plants and
parasitoids. BioMed Res. Int. 2015 doi:http://dx.doi.org/10.1155/2015/342982
et al., 2014b; Jiang et al., 2016), where GLV and monoterpene and Article ID 342982, 18 pages.
GLV emissions are almost proportional. Blande, J.D., Holopainen, J.K., Niinemets, U., 2014. Plant volatiles in polluted
The emission of shikimate pathway derived compound, methyl atmospheres: stress responses and signal degradation. Plant Cell Environ. 37,
1892–1904.
salicylate (MeSA), is typically observed for sap-sucking herbivores Brilli, F., Ruuskanen, T.M., Schnitzhofer, R., Mueller, M., Breitenlechner, M., Bittner,
such as aphids or whiteflies (Li et al., 2006; Zarate et al., 2007) and V., Wohlfahrt, G., Loreto, F., Hansel, A., 2011. Detection of plant volatiles after leaf
is not usually considered as part of chewing herbivore response. wounding and darkening by proton transfer reaction time-of-flight mass
spectrometry (PTR-TOF). PLoS One 6.
However, as in our study, several previous studies have demon- Bruce, T.J.A., 2015. Interplay between insects and plants: dynamic and complex
strated the release of methyl salicylate upon chewing herbivore interactions that have coevolved over millions of years but act in milliseconds. J.
attacks (Cardoza et al., 2002; Dicke et al., 1999), indicating a Exp. Bot. 66, 455–465.
L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192 191

Cardoza, Y.J., Alborn, H.T., Tumlinson, J.H., 2002. In vivo volatile emissions from birds use volatile organic compounds from plants as olfactory foraging cues?
peanut plants induced by simultaneous fungal infection and insect damage. J. Three experimental tests. Ethology 121, 1131–1144.
Chem. Ecol. 28, 161–174. Laothawornkitkul, J., Paul, N.D., Vickers, C.E., Possell, M., Taylor, J.E., Mullineaux, P.
Copolovici, L., Niinemets, Ü., 2016. Environmental impacts on plant volatile M., Hewitt, C.N., 2008. Isoprene emissions influence herbivore feeding
emission. In: Blande, J., Glinwood, R. (Eds.), Deciphering Chemical Language of decisions. Plant Cell Environ. 31, 1410–1415.
Plant Communication. Springer International Publishing, Berlin, pp. 35–59. Lee, A.K.Y., Hayden, K.L., Herckes, P., Leaitch, W.R., Liggio, J., Macdonald, A.M., Abbatt,
Copolovici, L., Kannaste, A., Remmel, T., Vislap, V., Niinemets, U., 2011. Volatile J.P.D., 2012. Characterization of aerosol and cloud water at a mountain site
emissions from Alnus glutionosa induced by herbivory are quantitatively during WACS 2010: secondary organic aerosol formation through oxidative
related to the extent of damage. J. Chem. Ecol. 37, 18–28. cloud processing. Atmos. Chem. Phys. 12, 7103–7116.
Copolovici, L., Kannaste, A., Remmel, T., Niinemets, U., 2014a. Volatile organic Li, Q., Xie, Q.G., Smith-Becker, J., Navarre, D.A., Kaloshian, I., 2006. Mi-1-mediated
compound emissions from Alnus glutinosa under interacting drought and aphid resistance involves salicylic acid and mitogen-activated protein kinase
herbivory stresses. Environ. Exp. Bot. 100, 55–63. signaling cascades. Mol. Plant Microbe Interact. 19, 655–664.
Copolovici, L., Vaartnou, F., Portillo Estrada, M., Niinemets, U., 2014b. Oak powdery Link, M., Zhou, Y., Taubman, B., Sherman, J., Morrow, H., Krintz, I., Robertson, L., Cook,
mildew (Erysiphe alphitoides)-induced volatile emissions scale with the degree R., Stocks, J., West, M., Sive, B.C., 2015. A characterization of volatile organic
of infection in Quercus robur. Tree Physiol. 34, 1399–1410. compounds and secondary organic aerosol at a mountain site in the
Delaney, K.J., Haile, F.J., Peterson, R.K.D., Higley, L.G., 2008. Impairment of leaf Southeastern United States. J. Atmos. Chem. 72, 81–104.
photosynthesis after insect herbivory or mechanical injury on common Loivamaeki, M., Mumm, R., Dicke, M., Schnitzler, J.-P., 2008. Isoprene interferes with
milkweed, asclepias syriaca. Environ. Entomol. 37, 1332–1343. the attraction of bodyguards by herbaceous plants. Proc. Natl. Acad. Sci. U. S. A.
Dicke, M., Baldwin, I.T., 2010. The evolutionary context for herbivore-induced plant 105, 17430–17435.
volatiles: beyond the ‘cry for help’. Trends Plant Sci. 15, 167–175. Lombard, J., Moreira, D., 2011. Origins and early evolution of the mevalonate
Dicke, M., Gols, R., Ludeking, D., Posthumus, M.A., 1999. Jasmonic acid and herbivory pathway of isoprenoid Biosynthesis in the three domains of life. Mol. Biol. Evol.
differentially induce carnivore-attracting plant volatiles in lima bean plants. J. 28, 87–99.
Chem. Ecol. 25, 1907–1922. Loreto, F., Schnitzler, J.-P., 2010. Abiotic stresses and induced BVOCs. Trends Plant
Dicke, M., 2016. Plant phenotypic plasticity in the phytobiome: a volatile issue. Curr. Sci. 15, 154–166.
Opin. Plant Biol. 32, 17–23. Loreto, F., Dicke, M., Schnitzler, J.-P., Turlings, T.C.J., 2014. Plant volatiles and the
Dwyer, G., Dushoff, J., Yee, S.H., 2004. The combined effects of pathogens and environment. Plant Cell Environ. 37, 1905–1908.
predators on insect outbreaks. Nature 430, 341–345. Lucas-Barbosa, D., van Loon, J.J.A., Dicke, M., 2011. The effects of herbivore-induced
Feussner, I., Wasternack, C., 2002. The lipoxygenase pathway. Annu. Rev. Plant Biol. plant volatiles on interactions between plants and flower-visiting insects.
53, 275–297. Phytochemistry 72, 1647–1654.
Fineschi, S., Loreto, F., Staudt, M., Peñuelas, J., 2013. Diversification of volatile Lucas-Barbosa, D., 2016. Integrating studies on plant-pollinator and plant-herbivore
isoprenoid emissions from trees: evolutionary and ecological perspectives. In: interactions. Trends Plant Sci. 21, 125–133.
Niinemets, Ü., Monson, R.K. (Eds.), Biology, Controls and Models of Tree Volatile Matsui, K., Sugimoto, K., Mano, J.i., Ozawa, R., Takabayashi, J., 2012. Differential
Organic Compound Emissions. Springer, Berlin. metabolisms of green leaf volatiles in injured and intact parts of a wounded leaf
Giorgi, A., Manzo, A., Nanayakkara, N.N.M.C., Giupponi, L., Cocucci, M., Panseri, S., meet distinct ecophysiological requirements. PLoS One 7, e36433.
2015. Effect of biotic and abiotic stresses on volatile emission of Achillea collina Matsui, K., 2006. Green leaf volatiles: hydroperoxide lyase pathway of oxylipin
Becker ex Rchb. Nat. Prod. Res. 29, 1695–1702. metabolism. Curr. Opin. Plant Biol. 9, 274–280.
Giraldo, A., Heller, W., Fladung, M., Schnitzler, J.-P., Schroeder, H., 2012. Function of Mattson, W.J., Haack, R.A., 1987. The role of drought in outbreaks of plant-eating
defensive volatiles in pedunculate oak (Quercus robur) is tricked by the moth insects. Bioscience 37, 110–118.
Tortrix viridana. Plant Cell Environ. 35, 2192–2207. Milanovic, S., Lazarevic, J., Popovic, Z., Miletic, Z., Kostic, M., Radulovic, Z., Karadzic,
Gog, L., Berenbaum, M.R., DeLucia, E.H., Zangerl, A.R., 2005. Autotoxic effects of D., Vuleta, A., 2014. Preference and performance of the larvae of Lymantria
essential oils on photosynthesis in parsley, parsnip, and rough lemon. dispar (Lepidoptera: Lymantriidae) on three species of European oaks. Eur. J.
Chemoecology 15, 115–119. Entomol. 111, 371–378.
Guenther, A.B., Jiang, X., Heald, C.L., Sakulyanontvittaya, T., Duhl, T., Emmons, L.K., Moldau, H., Wong, S.-C., Osmond, C.B., 1993. Transient depression of photosynthesis
Wang, X., 2012. The Model of Emissions of Gases and Aerosols from Nature in bean leaves during rapid water loss. Aust. J. Plant Physiol. 20, 45–54.
version 2.1 (MEGAN2.1): an extended and updated framework for modeling Munné-Bosch, S., 2007. Ageing in perennials. Crit. Rev. Plant Sci. 26, 123–138.
biogenic emissions. Geosci. Model. Dev. 5, 1471–1492. Munné-Bosch, S., 2008. Do perennials really senesce? Trends Plant Sci. 13, 216–220.
Guenther, A., 2013. Upscaling biogenic volatile compound emissions from leaves to Nabity, P.D., Zavala, J.A., DeLucia, E.H., 2009. Indirect suppression of photosynthesis
landscapes. In: Niinemets, Ü., Monson, R.K. (Eds.), Biology, Controls and Models on individual leaves by arthropod herbivory. Ann. Bot. 103, 655–663.
of Tree Volatile Organic Compound Emissions. Springer, Berlin, pp. 391–414. Nabity, P.D., Zavala, J.A., DeLucia, E.H., 2013. Herbivore induction of jasmonic acid
Hare, J.D., 2011. Ecological role of volatiles produced by plants in response to and chemical defences reduce photosynthesis in Nicotiana attenuata. J. Exp. Bot.
damage by herbivorous insects. Annu. Rev. Entomol. 56, 161–180. 64, 685–694.
Harrison, S.P., Morfopoulos, C., Dani, K.G.S., Prentice, I.C., Arneth, A., Atwell, B.J., Niinemets, U., Bilger, W., Kull, O., Tenhunen, J.D., 1998. Acclimation to high
Barkley, M.P., Leishman, M.R., Loreto, F., Medlyn, B.E., Niinemets, Ü., Possell, M., irradiance in temperate deciduous trees in the field: changes in xanthophyll
Peñuelas, J., Wright, I.J., 2013. Volatile isoprenoid emissions from plastid to cycle pool size and in photosynthetic capacity along a canopy light gradient.
planet. New Phytol. 197, 49–57. Plant Cell Environ. 21, 1205–1218.
Heil, M., 2014. Herbivore- induced plant volatiles: targets, perception and Niinemets, Ü., Arneth, A., Kuhn, U., Monson, R.K., Peñuelas, J., Staudt, M., 2010a. The
unanswered questions. New Phytol. 204, 297–306. emission factor of volatile isoprenoids: stress, acclimation, and developmental
Holopainen, J.K., Blande, J.D., 2013. Where do herbivore-induced plant volatiles go? responses. Biogeosciences 7, 2203–2223.
Front. Plant Sci. 4. Niinemets, Ü., Copolovici, L., Hüve, K., 2010b. High within-canopy variation in
Holopainen, J.K., Gershenzon, J., 2010. Multiple stress factors and the emission of isoprene emission potentials in temperate trees: implications for predicting
plant VOCs. Trends Plant Sci. 15, 176–184. canopy-scale isoprene fluxes. J. Geophys. Res.  Biogeosci. 115, G04029.
Holopainen, J.K., 2011. Can forest trees compensate for stress-generated growth Niinemets, Ü., Kuhn, U., Harley, P.C., Staudt, M., Arneth, A., Cescatti, A., Ciccioli, P.,
losses by induced production of volatile compounds? Tree Physiol. 31, 1356– Copolovici, L., Geron, C., Guenther, A.B., Kesselmeier, J., Lerdau, M.T., Monson, R.
1377. K., Peñuelas, J., 2011. Estimations of isoprenoid emission capacity from
Huang, J., Zhang, P.-J., Zhang, J., Lu, Y.-B., Huang, F., Li, M.-J., 2013. Chlorophyll enclosure studies: measurements, data processing, quality and standardized
content and chlorophyll fluorescence in tomato leaves infested with an invasive measurement protocols. Biogeosciences 8, 2209–2246.
mealybug, Phenacoccus solenopsis (Hemiptera: Pseudococcidae). Environ. Niinemets, Ü., García-Plazaola, J.I., Tosens, T., 2012. Photosynthesis during leaf
Entomol. 42, 973–979. development and ageing. In: Flexas, J., Loreto, F., Medrano, H. (Eds.), Terrestrial
Jiang, Y., Ye, J., Veromann, L.-L., Niinemets, Ü., 2016. Scaling of photosynthesis and Photosynthesis in a Changing Environment. A Molecular, Physiological and
constitutive and induced volatile emissions with severity of leaf infection by Ecological Approach. Cambridge University Press, Cambridge, pp. 353–372.
rust fungus (Melampsora larici-populina) in Populus balsamifera var. suaveolens. Niinemets, Ü., Kännaste, A., Copolovici, L., 2013. Quantitative patterns between
Tree Physiol. 38, 856–872. plant volatile emissions induced by biotic stresses and the degree of damage.
Johnson, D., Gilbert, L., 2015. Interplant signalling through hyphal networks. New Frontiers in Plant Science. Front. Plant-Microb. Interact. 4, 262.
Phytol. 205, 1448–1453. Niinemets, Ü., 2010. Mild versus severe stress and BVOCs: thresholds, priming and
Kännaste, A., Copolovici, L., Niinemets, Ü., 2014. Gas chromatography mass- consequences. Trends Plant Sci. 15, 145–153.
spectrometry method for determination of biogenic volatile organic Opris, O., Copaciu, F., Soran, M.L., Ristoiu, D., Niinemets, U., Copolovici, L., 2013.
compounds emitted by plants. In: Rodríguez-Concepción, M. (Ed.), Plant Influence of nine antibiotics on key secondary metabolites and physiological
Isoprenoids: Methods and Protocols. Humana Press, New York, pp. 161–169. characteristics in Triticum aestivum: Leaf volatiles as a promising new tool to
Kesselmeier, J., Staudt, M., 1999. Biogenic volatile organic compounds (VOC): an assess toxicity. Ecotoxicol. Environ. Saf. 87, 70–79.
overview on emission, physiology and ecology. J. Atmos. Chem. 33, 23–88. Pazouki, L., Niinemets, Ü., 2016. Multi-substrate terpenoid synthases: their
Kleist, E., Mentel, T.F., Andres, S., Bohne, A., Folkers, A., Kiendler-Scharr, A., Rudich, Y., occurrence and physiological significance. Front. Plant Sci. 7, 1019.
Springer, M., Tillmann, R., Wildt, J., 2012. Irreversible impacts of heat on the Pazouki, L., Kanagendran, A., Li, S., Kännaste, A., Rajabi Memari, H., Bichele, R.,
emissions of monoterpenes, sesquiterpenes, phenolic BVOC and green leaf Niinemets, Ü., 2016. Mono- and sesquiterpene release from tomato (Solanum
volatiles from several tree species. Biogeosciences 9, 5111–5123. lycopersicum) leaves upon mild and severe heat stress and through recovery:
Koski, T.-M., Laaksonen, T., Mantyla, E., Ruuskanen, S., Li, T., Giron-Calva, P.S., from gene expression to emission responses. Environ. Exp. Bot. 132, 1–15.
Huttunen, L., Blande, J.D., Holopainen, J.K., Klemola, T., 2015. Do insectivorous
192 L. Copolovici et al. / Environmental and Experimental Botany 138 (2017) 184–192

Pichersky, E., Gershenzon, J., 2002. The formation and function of plant volatiles: Schmidt, R., Etalo, D.W., de Jager, V., Gerards, S., Zweers, H., de Boer, W., Garbeva, P.,
perfumes for pollinator attraction and defense. Curr. Opin. Plant Biol. 5, 237– 2016. Microbial small talk: volatiles in fungal-bacterial interactions. Front.
243. Microbiol. 6.
Porta, H., Rocha-Sosa, M., 2002. Plant lipoxygenases: physiological and molecular Shen, X., Zhao, Y., Chen, Z., Huang, D., 2013. Heterogeneous reactions of volatile
features. Plant Physiol. 130, 15–21. organic compounds in the atmosphere. Atmos. Environ. 68, 297–314.
Portillo-Estrada, M., Kazantsev, T., Talts, E., Tosens, T., Niinemets, Ü., 2015. Emission Stam, J.M., Kroes, A., Li, Y., Gols, R., van Loon, J.J.A., Poelman, E.H., Dicke, M., 2014.
timetable and quantitative patterns of wound-induced volatiles across different Plant interactions with multiple insect herbivores: from community to genes.
damage treatments in aspen (Populus tremula). J. Chem. Ecol. 41, 1105–1117. Annu. Rev. Plant Biol. 65, 689–713.
Possell, M., Loreto, F., 2013. The role of volatile organic compounds in plant Tholl, D., Sohrabi, R., Huh, J.-H., Lee, S., 2011. The biochemistry of homoterpenes
resistance to abiotic stresses: responses and mechanisms. In: Niinemets, Ü., common constituents of floral and herbivore-induced plant volatile bouquets.
Monson, R.K. (Eds.), Biology, Controls and Models of Tree Volatile Organic Phytochemistry 72, 1635–1646.
Compound Emissions. Springer, Berlin, pp. 209–235. Thomas, F.M., Blank, R., Hartmann, G., 2002. Abiotic and biotic factors and their
Rasulov, B., Hüve, K., Välbe, M., Laisk, A., Niinemets, Ü., 2009. Evidence that light, interactions as causes of oak decline in Central Europe. For. Pathol. 32, 277–307.
carbon dioxide and oxygen dependencies of leaf isoprene emission are driven Tonioli, M., Escarre, J., Lepart, J., Speranza, M., 2001. Facilitation and competition
by energy status in hybrid aspen. Plant Physiol. 151, 448–460. affecting the regeneration of Quercus pubescens Willd. Ecoscience 8, 381–391.
Rasulov, B., Hüve, K., Laisk, A., Niinemets, Ü., 2011. Induction of a longer-term Toome, M., Randjärv, P., Copolovici, L., Niinemets, Ü., Heinsoo, K., Luik, A., Noe, S.M.,
component of isoprene release in darkened aspen leaves: origin and regulation 2010. Leaf rust induced volatile organic compounds signalling in willow during
under different environmental conditions. Plant Physiol. 156, 816–831. the infection. Planta 232, 235–243.
Rasulov, B., Bichele, I., Laisk, A., Niinemets, Ü., 2014. Competition between isoprene Vickers, C.E., Gershenzon, J., Lerdau, M.T., Loreto, F., 2009. A unified mechanism of
emission and pigment synthesis during leaf development in aspen. Plant Cell action for volatile isoprenoids in plant abiotic stress. Nat. Chem. Biol. 5, 283–
Environ. 37, 724–741. 291.
Rehman, F., Khan, F.A., Anis, S.B., 2014. Assessment of aphid infestation levels in von Caemmerer, S., Farquhar, G.D., 1981. Some relationships between the
some cultivars of mustard with varying defensive traits. Arch. Phytopathol. biochemistry of photosynthesis and the gas exchange of leaves. Planta 153, 376–
Plant Protect. 47, 1866–1874. 387.
Roslin, T., Gripenberg, S., Salminen, J.P., Karonen, M., O'Hara, R.B., Pihlaja, K., Vranova, E., Coman, D., Gruissem, W., 2012. Structure and dynamics of the
Pulkkinen, P., 2006. Seeing the trees for the leaves – oaks as mosaics for a host- isoprenoid pathway network. Mol. Plant 5, 318–333.
specific moth. Oikos 113, 106–120. Zarate, S.I., Kempema, L.A., Walling, L.L., 2007. Silverleaf whitefly induces salicylic
Sack, L., Cowan, P.D., Holbrook, N.M., 2003. The major veins of mesomorphic leaves acid defenses and suppresses effectual jasmonic acid defenses. Plant Physiol. 43,
revisited: testing for conductive overload in Acer saccharum (Aceraceae) and 866–875.
Quercus rubra (Fagaceae). Am. J. Bot. 90, 32–39. Zhang, X., McVay, R.C., Huang, D.D., Dalleska, N.F., Aumont, B., Flagan, R.C., Seinfeld,
Sack, L., Streeter, C.M., Holbrook, N.M., 2004. Hydraulic analysis of water flow J.H., 2015. Formation and evolution of molecular products in alpha-pinene
through leaves of sugar maple and red oak. Plant Physiol. 134, 1824–1833. secondary organic aerosol. Proc. Natl. Acad. Sci. U. S. A. 112, 14168–14173.
Scala, A., Allmann, S., Mirabella, R., Haring, M.A., Schuurink, R.C., 2013. Green leaf Zhou, S., Lou, Y.-R., Tzin, V., Jander, G., 2015. Alteration of plant primary metabolism
volatiles: a plant’s multifunctional weapon against herbivores and pathogens. in response to insect herbivory. Plant Physiol. 169, 1488–1498.
Int. J. Mol. Sci. 14, 17781–17811. Zhu, F., Poelman, E.H., Dicke, M., 2014. Insect herbivore- associated organisms affect
Schaub, A., Blande, J.D., Graus, M., Oksanen, E., Holopainen, J.K., Hansel, A., 2010. plant responses to herbivory. New Phytol. 204, 315–321.
Real-time monitoring of herbivore induced volatile emissions in the field. Ziemann, P.J., Atkinson, R., 2012. Kinetics, products, and mechanisms of secondary
Physiol. Plant. 138, 123–133. organic aerosol formation. Chem. Soc. Rev. 41, 6582–6605.

You might also like