ArticlePDF AvailableLiterature Review

Abiotic stresses and induced BVOCs

Authors:

Abstract and Figures

Plants produce a wide spectrum of biogenic volatile organic compounds (BVOCs) in various tissues above and below ground to communicate with other plants and organisms. However, BVOCs also have various functions in biotic and abiotic stresses. For example abiotic stresses enhance BVOCs emission rates and patterns, altering the communication with other organisms and the photochemical cycles. Recent new insights on biosynthesis and eco-physiological control of constitutive or induced BVOCs have led to formulation of hypotheses on their functions which are presented in this review. Specifically, oxidative and thermal stresses are relieved in the presence of volatile terpenes. Terpenes, C6 compounds, and methyl salicylate are thought to promote direct and indirect defence by modulating the signalling that biochemically activate defence pathways.
Content may be subject to copyright.
This article appeared in a journal published by Elsevier. The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
Special Issue: Induced biogenic volatile organic compounds from plants
Abiotic stresses and induced BVOCs
Francesco Loreto
1
and Jo¨ rg-Peter Schnitzler
2
1
Consiglio Nazionale delle Ricerche (CNR), Istituto per la Protezione delle Piante (IPP), Via Madonna del Piano 10, 50019 Sesto
Fiorentino, Firenze, Italy
2
Karlsruhe Institute of Technology (KIT), Institute of Meteorology and Climate Research (IMK-IFU), Kreuzeckbahnstraße 19,
82467 Garmisch-Partenkirchen, Germany
Plants produce a wide spectrum of biogenic volatile
organic compounds (BVOCs) in various tissues above
and below ground to communicate with other plants
and organisms. However, BVOCs also have various func-
tions in biotic and abiotic stresses. For example abiotic
stresses enhance BVOCs emission rates and patterns,
altering the communication with other organisms and
the photochemical cycles. Recent new insights on bio-
synthesis and eco-physiological control of constitutive
or induced BVOCs have led to formulation of hypotheses
on their functions which are presented in this review.
Specifically, oxidative and thermal stresses are relieved
in the presence of volatile terpenes. Terpenes, C6 com-
pounds, and methyl salicylate are thought to promote
direct and indirect defence by modulating the signalling
that biochemically activate defence pathways.
The emission of BVOCs: few biochemical pathways but
many compounds emitted
Biosynthesis of the main BVOCs
Plants produce a wide spectrum of BVOCs in various
tissues above and below ground. Most BVOCs are largely
lipophilic and have enough vapour pressure to be released
into the atmosphere in significant amounts. The availabil-
ity of new methods of head-space sampling (such as solid
phase micro-extraction) in combination with gas chroma-
tographymass spectroscopy and new techniques for on-
line analysis (proton transfer reaction-mass spectrometry)
[1] has led, in the last 15 years, to a significant expansion of
our knowledge on the occurrence and temporal and spatial
distribution of BVOCs emissions. At present, about 1700
substances have been found to be emitted from plants [2].
Nearly all organs from vegetative parts, as well as flowers
[3] and roots [4] emit these compounds. Many BVOCs are
emitted constitutively and the emissions can be observed
throughout the life cycle of the plant or, more often, at
specific developmental stages (e.g. leaf and needle matu-
ration, senescence, flowering, and fruit ripening). The
emission is biosynthetically controlled by abiotic factors
such as light and/or temperature, atmospheric CO
2
con-
centration, or nutrition. Other BVOCs are induced after
wounding and herbivore feeding or after environmental
stresses. Stresses may induce change of constitutive
BVOCs, either stimulating or quenching the emissions
(e.g. [5]) or may induce de novo synthesis and emission
of BVOCs. Induced emissions may occur in a systemic way,
i.e. away from the site of damage [6].
The biosynthesis of most BVOCs can be assigned to the
following three major pathways: terpenes (= isoprenoids),
oxylipins, and shikimate and benzoic acid [7,8]. Low mol-
ecular weight, C1 and C2, compounds, such as methanol,
ethanol, formaldehyde, and acetaldehyde can be synthes-
ized via other biosynthetic routes [9]. Two other BVOCs are
methane and ethylene. Emission of non-microbial methane
by vegetation has been discovered recently [10] and
research on this important topic is still in its infancy.
Recent isotope labelling studies provided evidence that
methane can be generated from methoxyl groups deriving
from breakdown of plant pectins [11].
In this review of the impact of abiotic factors on the
induction of BVOCs emissions, we will mostly concentrate
on volatile terpenes which are the most important com-
pounds for plant biology [12] and atmospheric chemistry
[13] because of their role in plant protection (e.g. in the
protection of photosynthesis against thermal and oxidative
stresses, and in direct and indirect defence against herbi-
vores), as well as in the chemical properties of the atmos-
phere (e.g. entering the cycle of photochemical production/
destruction of ozone, aerosols, and particles). The emission
of volatile terpenes is estimated to account for more than
half of the total emission of BVOCs [14] and is constitu-
tively ten times higher than other emissions, as heavy
emitters can release isoprene at rates of 50100 nmol
m
2
s
1
, representing up to 25% of the photosynthetic
net carbon uptake in tree species. Volatile terpene emis-
sions are far more sustained than the emission of other
induced volatiles, for which emissions are transient by
nature and limited to specific periods after stress, depend-
ing on the damage experienced and on the activation of the
biosynthetic pathways producing the volatiles.
Biosynthesis of volatile terpenes
Terpenes are constitutively formed in some plant families
that store them in massive amounts in internal or external
structures (e.g. the resin ducts of conifers or the glandular
cells of Lamiaceae leaves). They may also be induced in
response to wounding or herbivory attack. The emission of
terpenes from storage structures is generally uncoupled
from photosynthesis as it may occur, for example, at night
[15].
Direct emission of isoprene or monoterpenes from the
mesophyll is common in some tree species (Figure 1), in
Review
Corresponding author: Loreto, F. (francesco.loreto@ipp.cnr.it).
154 1360-1385/$ see front matter ß2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.tplants.2009.12.006 Available online 4 February 2010
Author's personal copy
particular from the Fagaceae and Salicaceae (oaks and
poplars) [16,17]. These emissions are light-dependent
[18,19] and are closely linked to the availability of photo-
synthetic intermediates [19,20]. The investigation of
volatile terpene biosynthesis is a very active area of plant
research, especially since the discovery and complete elu-
cidation of the methylerythritol (MEP) pathway, respon-
sible for the formation of the basic C5 units isopentenyl
diphosphate (IDP) and dimethylallyl diphosphate
(DMADP) in many bacteria and in the plastids of all
organisms from phototrophic phyla [21].
In plants, IDP and DMADP are formed via two alterna-
tive pathways: (i) in the cytosolic mevalonic acid (MVA)
pathway from acetyl-CoA, and (ii) in the plastidic MEP
pathway from pyruvate and glyceraldehyde-3-phosphate
[22]. Generally, the MEP pathway provides IDP and
DMADP for hemiterpene and monoterpene biosynthesis,
while the MVA pathway provides the C5 units for sesqui-
terpene formation. However, a very recent report has
revealed that the MEP pathway also contributes to sesqui-
terpene formation [23]. In addition, some metabolic
crosstalk between both biosynthetic routes is possible
[24] particularly (via IDP) in the direction from chloro-
plasts to the cytosol [25].
Prenyl transferases catalyze the condensation of IDP
and DMADP to form geranyl diphosphate (GDP) and
farnesyl diphosphate (FDP) [26]. Finally, the conversion
of DMADP, GDP, and FDP, into volatile hemi-, mono-, and
sesquiterpenes, respectively, is catalyzed by terpene
synthases (TPS), a large family of enzymes, encoded by
closely related genes [27]. Isolation and characterization of
prenyl transferase [28] and terpene synthase genes [29] is
now giving new insights into the evolutionary origin
[30,31] and genetic and biochemical regulation of terpene
biosynthesis.
Isoprene and 2-methyl-3-buten-2-ol (MBO), a hemiter-
pene common only in American western pines, [32] are
biochemically synthesized in chloroplasts by isoprene [33]
and MBO synthase [34] from DMADP. Due to the import-
ance of isoprene for atmospheric processes and plant func-
tions, isoprene synthase (ISPS) became one of the best
studied TPS [35]. A positive correlation between ISPS
activity and basal standard emission capacity was found
in different isoprene-emitting species [36,37].Uptonow,
five ISPS genes from different poplar (Populus) species or
poplar hybrids have been described [35]. The only ISPS gene
cloned so far from another genus, Pueraria montana (kudzu)
[38], shows only 52% identity with the poplar protein
sequences, although the structure of poplar and kudzu genes
are similar (six introns and seven exons) [38].AllISPS genes
belong to the subgroup b of the class 1 plant TPS-family
which includes monomeric mono-, sesqui-, and diterpene
synthases, grouped in six subgroups (TspaTspfTpsaTpsf?)
[27]. All known ISPS enzymes have a 10100-fold higher
Michaelis constant (k
M
) for its substrate DMADP (in the
millimolar range) than monoterpene synthases for GDP [39]
or prenyltransferases for DMADP [40]. The low k
M
of pre-
nyltransferases may control the metabolic flux within the
MEP pathway because downstream reactions leading to
monoterpene and non-volatile terpene biosynthesis are
favoured over isoprene biosynthesis. Based on this finding,
it was suggested that isoprene emission occurs only when
plants’ need for ‘essential’, higher terpenes (hormones, e.g.
ABA and gibberellins; tocopherol; phytosterols; and photo-
synthetic pigments) are satisfied [41].
Interest in genetic regulation and biochemical proper-
ties of TPS other than ISPS mostly focused on those TPS
involved in the formation in storage structures (e.g. in
gymnosperms and Lamiaceae) of those terpenes with
defensive or attractive functions [42] (see Dicke and Bald-
win, this issue).
In light-dependent monoterpene emitters, the infor-
mation about TPS regulation is scant. Differences in mono-
terpene emission pattern of chemotypes and oak hybrids
[43], the seasonal development of monoterpene emission
[44,45], and the dependence of monoterpene emission on
atmospheric CO
2
[46] all appear to be controlled by TPS
activities. However, our knowledge about the underlying
genes is scarce; in oak, only two mono-TPS genes, a b-
myrcene synthase [47] and a multiproduct a-pinene
synthase [48] have been isolated and functionally charac-
terized up to now.
Figure 1. Origin of volatile terpene emissions from different leaf types. In
deciduous leaves of many tree species (e.g. oaks and poplars) with no specific
storage structures for terpenes, isoprene and monoterpene, BVOCs emissions
originate from mesophyll cells in a light- and temperature-dependent manner. In
conifer needles (e.g. Picea abies - Norway spruce) light-dependent terpene
emission stems from photosynthetic tissue and is superposed by a temperature-
dependent volatilization of terpenes from resin ducts. Lamiaceae (e.g. Ocimum
basilicum basil) leaves release temperature-dependent volatile terpenes from
external glandular cells. Confocal laser scanning microscopic images were taken
from cross-sections of Quercus robur (top), Picea abies (middle) and leaf surface of
Ocimum basilicum (bottom).
Review Trends in Plant Science Vol.15 No.3
155
Author's personal copy
The commonly used model plant Arabidopsis thaliana
(Arabidopsis) is a good example of how modern technol-
ogies can improve knowledge of BVOCs biosynthesis and
functions. In the past classified as ‘non-emitting’ species,
analysis of the Arabidopsis genome revealed the existence
of over 30 putative genes belonging to the multigene family
of TPS [49,50]. Most of these genes are almost exclusively
expressed in flowers [50,51], but low constitutive terpene
emissions from leaves and siliques [50,52] and even emis-
sions from roots (namely 1,8-cineole, [4]) could be detected.
Isoprene synthase overexpression in Arabidopsis allowed
verification of a ‘thermo-protective’ activity of isoprene
[53,54], see below, and indicated an ecological function
of this hemiterpene in plantinsect interactions as a repel-
ling cue [55]. The expression of b-caryophyllene was also
successfully engineered in this plant (J. Gershenzon,
personal communication) and may help assess functions
and fate of sesquiterpenes. Sesquiterpenes are not emitted
in large amounts constitutively, but their biosynthesis can
be induced by biotic stresses, and may be important as an
indirect defence mechanism (see Dicke and Baldwin, this
issue). Even at low concentrations sesquiterpenes are also
important as nucleation factors, eventually leading to
particle formation in the atmosphere [56].
Poplar is another seminal model system to study regu-
lation and function of plant volatiles. Besides ISPS [35],
only one TPS gene [()-germacrene-D synthase] [57] is
characterized. However, full length cDNAs [58], cDNA
microarrays [59,60], and a largely sequenced genome of
Populus trichocarpa [61] are now available, and these tools
will certainly allow for quick progress in the understanding
of BVOCs formation in a near future.
Biosynthesis of C1 and C2 oxygenated compounds
Volatile terpenes are certainly the family of compounds
that contributes the majority of BVOCs [14]. However,
short-chained oxygenated compounds (especially methanol
and acetaldehyde) are important components of constitu-
tive and induced emissions of many plants [5] with a large
presence globally [62].
Methanol is predominantly emitted because of degra-
dation and formation of cell wall pectins, e.g. (i) during cell
expansion in all types of plant tissue and seeds, and (ii)
during leaf abscission, senescence, and seed maturation
(Figure 2). The formation of methanol is catalyzed by
pectin methylesterases (PME) which, among the other
functions, demethoxylates pectin [63,64]. To a minor
extent, methanol emissions originate from protein meth-
yltransferase and protein repair reactions [65], or from
tetrahydrofolate metabolism [66]. Emission of methanol
can be induced by mechanical wounding [67] or herbivore
feeding, due to an upregulation of PME expression. This
finding has stimulated a discussion about whether metha-
nol might act as a signal in plantplant communication
[68].
The metabolic origin of acetaldehyde emitted by forest
trees is still a matter of debate [69,70]. This compound
seems to be predominantly induced by stresses. It is known
that acetaldehyde emission correlates with root flooding
[71,72] and with xylem sap ethanol concentrations [73,74].
Ethanol formed under anoxic conditions in roots is trans-
ported to leaves by the transpiration stream, where it is
oxidized to acetaldehyde by alcohol dehydrogenase (ADH).
However, only a small portion of acetaldehyde is emitted
while the bulk is further metabolized by aldehyde dehydro-
genase (ALDH) to acetate and acetyl-CoA.
In some tree species, strong transient acetaldehyde
bursts during lightdark transitions have been reported
[70,72,75]. These acetaldehyde bursts are thought to be the
result of a ‘pyruvate overflow mechanism’ [75]. In the
proposed mechanism pyruvate decarboxylase (PDC) acts
as a metabolic regulator converting excess cytosolic pyr-
uvate into acetaldehyde, which is subsequently oxidized to
acetate. Such an excess of cytosolic pyruvate may be the
result of transiently decreased transport rates of pyruvate
equivalents [i.e. phosphoenolpyruvate (PEP)] into orga-
nelles, or reduced pyruvate utilization in leaf cells immedi-
ately after darkening [75]. However, emission of
acetaldehyde may also derive from cleavage of moieties
of C6 aldehydes, which emissions are also transiently
stimulated during lightdark transitions [70]. In this case,
acetaldehyde emission is independent of cytosolic pyruvate
and is part of the leaf response to wounding that also
includes biosynthesis of C6 compounds (see below).
Biosynthesis of C6 aldehydes and alcohols
Wounding induces the release of ‘green leaf volatiles’
(GLV) as can be easily sensed in the odour of fresh hay.
In most wounded plants GLV are C6 aldehydes, C6 alco-
hols, and their derivatives, often collectively called C6- or
LOX-products [76]. Physiologically these compounds have
Figure 2. Simplified scheme of the subcellular origin and biosynthesis of volatile
organic compounds upon abiotic stress. Abbreviations: CH
4
, methane; DMADP,
dimethylallyl diphosphate; DXS, 1-deoxy-D-xylulose 5-phosphate synthase (EC
4.1.3.37); FDP, farnesyl diphosphate; GDP, geranyl diphosphate; HMGR, 3-hydroxy-
3-methylglutaryl-CoA reductase (EC 1.1.1.34); 13-HPOT, 13S-hydroperoxy-
9(Z),11(E),15(Z)-octadecatrienoic acid; IDP, isopentenyl diphosphate; ISPS,
isoprene synthase (EC 4.2.3.27); a-LeA, a-linolenic acid; 13-LOX, 13-lipoxygenase
(EC 1.13.11.12); MeOH, methanol; MTS, monoterpene synthase (e.g. myrcene EC
4.2.3.14); PDC, pyruvate decarboxylase (EC 4.1.1.1); PEP, phosphoenolpyruvate;
PME, pectine methylesterase (EC 3.1.1.11); PYR, pyruvate; STS, sesquiterpene
synthase (e.g. epi-aristolochene EC 4.2.3.9); 13-HPL, 13-hydroxyperoxide lyase (not
listed in enzyme classification); TP, triose phosphate. The broken arrow indicates a
proposed, yet unidentified acetaldehyde emission path from chloroplasts.
Review Trends in Plant Science Vol.15 No.3
156
Author's personal copy
antibiotic properties inhibiting the invasion of damaged
tissue [77], and can serve a signalling function within
plants to induce or prime defence [78].
GLV are derived from oxidized polyenoic fatty acids
(PUFA), collectively called oxylipins. The initial formation
is catalyzed by lipoxygenases (LOX), a large gene family of
non-haeme iron containing fatty acid dioxygenases [8]
(Figure 2). LOXs (13-LOX, classified with respect to their
positional specificity of oxidation at carbon-13 of the fatty
acid hydrocarbon backbone) initiate the octadecanoic
pathway by adding O
2
stereospecifically to unsaturated
fatty acids [e.g. linoleic acid (18:2) and a-linolenic acid
(18:3)] generating 13-(S)-hydroperoxides. Subsequently,
13-(S)-hydroperoxide lyase (HPL) catalyzes the cleavage
between C12 and C13 releasing n-hexanal (from linoleic
acid) and (Z)-3-hexenal (from a-linolenic acid) which are
the parent compounds for all other aldehydes and alcohols,
and enzymatically acetylated compounds such as hexyl
acetate and (Z)-3-hexenyl acetate [76].AnalysisofC6
compounds is complicated by their chemical instability
and the transient nature of their formation after wound-
ing. Thanks to modern on-line techniques like proton
transfer reaction mass spectrometry (PTR-MS) that avoid
pre-concentration of the samples on adsorbent phases, the
rapid and transient emission of C6 upon various stimuli,
such as physical damage [79], lightdark transitions, her-
bivory, or senescence processes, can be easily monitored
[70,79].
Plants and abiotic stresses: how stresses affect BVOCs
biosynthesis and emissions
Abiotic stresses affect primary and secondary metabolism
in different ways. Stresses generally inhibit photosyn-
thesis by reducing CO
2
uptake and diffusion inside leaves
to the site of fixation into carbohydrates or by altering the
photochemical or biochemical reactions of the photosyn-
thetic cycle [80]. The impact of stresses on the secondary
metabolisms that produces BVOCs is more controversial,
as some of the pathways may be elicited by stresses. In the
case of volatile terpenes, the uncoupling of photosynthesis
and BVOCs emission is surprising, because of the well-
known high requirement of photosynthetic carbon for ter-
pene biosynthesis [81]. The stimulation of terpene pro-
duction is made possible by the activation of sources of
carbon which are alternatives to freshly fixed photo-
synthates and not yet fully identified. Some labelling
studies have provided evidence that isoprene may also
be formed from xylem-transported glucose and chloroplas-
tic starch [82]. However, under stress conditions, other
13
C
labelling studies point out that, because of starch
depletion, extra-chloroplastic sources of carbon may be
activated and feed carbon to volatile terpenes [83]. Stresses
may also induce damage that elicits or induces the syn-
thesis of other volatiles with important consequences for
plant protection, directly against the environmental con-
straint [12] or indirectly, against possible, associated
bursts of pathogens and herbivores (see Dicke and Bald-
win, this issue). The impact of the main environmental
factors of stress on the biosynthesis and emission of vola-
tiles will be reviewed in the following sections (see also
Figure 3).
Temperature
BVOCs partition between the gas and liquid phase in the
plants according to their Henry’s law constant (k
H
), which
is generally very high (e.g. k
H
7500 Pa m
3
mol
1
at 25 8C
for isoprene) [84]. The equilibrium between gas and aqu-
eous phases is indeed determined by the temperature and
therefore it is expected that more BVOCs enter the gas
phase and are emitted at rising temperatures. This direct
effect of temperature on BVOCs emission is, however,
modulated by diffusion resistances encountered from the
sites of synthesis inside the leaf to the atmosphere. BVOCs
that are stored in specialized structures (ducts or glands)
reach very high concentrations and are tightly separated
by the surrounding cells by an impermeable cell layer,
generally made by sub-cuticular cells [85]. This is because
high concentrations of BVOCs, which have probably
important ecological functions as deterrents of herbivores
and as anti-bacterial and anti-mycotic substances, could be
auto-toxic for plant cells [86]. Temperature affects the
evaporation and release of a minimal part of the pools of
BVOCs that leaks out the impermeable cell layer. Indeed
plants with large storage pools are moderate emitters
[15,17] unless the pools are opened up, e.g. by herbivory
[87], strong winds, or forest fires [88]. High humidity may
also lead to a swelling and to the consequent explosion of
the structures containing the pools. In this case strong
emissions of terpenes may even be sensed, e.g. walking on a
pine forest at sunrise.
On the other hand, plants that do not store BVOCs into
specialized structures have small temporary pools in the
leaf mesophyll that freely diffuse out of the leaf driven by a
concentration gradient. The only important constraint to
this diffusion is at the stomatal conductance to gas
Figure 3. Schematic overview of long-term and short-term regulation of terpene
emissions upon various abiotic stresses. Long-term adaptation of volatile terpene
emission (predisposition and/or priming) relates to differential gene expression
resulting in upregulation (green circles) and/or downregulation (red circles) of
specific enzymes and changed emission pattern and emission rates. Short-term
regulation reflects rapid changes in metabolic fluxes and enzyme kinetic properties
(e.g. temperature, pH and ion shifts, and redox/energy status) due to abiotic factors
like atmospheric CO
2
concentration, light and temperature. Orientation of arrows
in the green and red circles indicate strength of upregulation and downregulation.
Abbreviations: h, hours; min, minutes; s, seconds.
Review Trends in Plant Science Vol.15 No.3
157
Author's personal copy
exchange. High temperatures often affect stomatal beha-
viour, either per se or because this stress is generally
associated with a drought stress. Stomatal opening under
transient heat stress is an important mechanism to dis-
sipate latent heat through transpiration of water and to
uncouple the leaf temperature from air temperature. On
the other hand, stomatal closure improves instantaneous
water use efficiency (the ratio between net CO
2
assimila-
tion and transpiration) and avoids excessive loss of water
driven by increasing transpiration. Stomatal movements
do not affect the steady-state diffusion rates of gases with
high k
H
, such as isoprene and monoterpenes [84]. In the
case of stomatal closure, these compounds build up tran-
siently significant partial pressures inside leaves that
compensate for the increasing resistance at stomatal level
[89,90]. However, diffusion of gases that are mainly parti-
tioned into liquid phase, such as oxygenated VOCs (metha-
nol, C6 aldehydes, and alcohols) might be strongly
restrained by stomatal closure [87].
Temperature has a strong and immediate influence on
the activity of the enzymes that catalyze the synthesis of
many BVOCs. Emissions of volatile terpenes typically
have a Q10 = 24, at temperatures variable between 20
and 40 8C[91]. Thus the main effect of rising temperature
is a direct increment of the terpenes formed through
enzymatic reactions. However, when a heat stress occurs
with temperature above the optimal enzyme temperature
(around 4045 8C for enzymes in the MEP pathway and
most TPS), then a very rapid inhibition of terpene emission
is observed [5]. This is probably due to downregulation or
impairment of primary metabolism and to the con-
sequently insufficient supply of photosynthetic metabolites
into the MEP pathway [92,93]. However, there are cases in
which the emission is not rapidly re-established upon heat
stress removal, and in these cases a heat-induced dena-
turation of the TPS is also likely to occur [5].
Interestingly, when the heat-induced inhibition of
volatile terpenes occurs, then the emission of other BVOCs
is highly enhanced [5]. This is particularly evident in the
case of methanol and C6 compounds, pinpointing that the
inhibition of volatile terpene emission coincides with the
occurrence of damage to cell walls and membranes,
respectively [5]. The emission of C6 compounds is sus-
tained for the whole period of heat stress and may continue
for a long time after temperatures go back to physiological
levels [5]. Significant fluxes of methyl-butenol, ethanol,
and acetaldehyde were found in North American conifers
exposed to high temperature [94], and methanol and
acetone emissions were also observed from bare agricul-
tural soils following a heat wave episode [95]. Thus soil
microorganisms may also contribute to the temperature-
dependent emission of oxygenated BVOCs.
Finally, temperature seems to have an important role in
determining emissions of methyl salicylate (MeSA), a sig-
nalling molecule whose induction is frequent in response to
biotic stresses [74,96]. One study [97] reported induction of
MeSA only after spraying plants with jasmonic acid,
another important signalling molecule activating the
induction of genes involved in hypersensitive responses,
both after biotic and abiotic stresses. In another study a
significant induction of MeSA, with fluxes comparable to
those of monoterpenes was observed in plants exposed to
night chilling temperatures [98] and the induction of this
compound correlated to the difference in temperature
between day and night that was experienced by walnut
(Juglans californica Juglans regia) plants.
Drought and salt
The impact of drought and salt on BVOCs emission is
surprising. These abiotic stresses directly affect stomatal
conductance and produce diffusive and biochemical limita-
tions of photosynthesis [79]. Both the reduction of photo-
synthesis and the stomatal closure are expected to
negatively impact on BVOCs emission by altering the
carbon supply into the MEP pathway and by increasing
resistance to their emission.
In fact, the emission of volatile terpenes is resistant to
these stresses and is often elicited by stress occurrence.
The original observation that isoprene is not reduced by
increasing drought stress until the stress becomes heavy
and almost completely inhibits photosynthesis [99] has
been repeatedly made, also with experiments in which
the drought stress has been controlled differently, e.g.
[90,100]. The sustained emission of isoprene to salt and
drought [101,102] is made possible by the induction of
carbon sources alternative to photosynthesis, possibly
related to respiration [103] or starch breakdown [82].
Labelling experiments with
13
CO
2
[35,83] demonstrate
clearly the preference of ‘old’ unlabeled (
12
C) carbon
skeletons over recently fixed,
13
C enriched photosynthetic
intermediates when photosynthesis (which is heavily
reduced by drought and salt stress) [83,101] and isoprene
biosynthesis become uncoupled. This may also explain why
in some instances re-establishment of photosynthetic
metabolism, by re-watering plants, results in a burst of
isoprene emission [99]. However, it has been shown [83]
that the alternative carbon sources rapidly cease to feed
carbon into the MEP pathway once the water status of
plants is reintegrated. A recent observation indicates that
in plants exposed to severe drought stress and in those
recovering from drought stress the temperature depen-
dency of isoprene emission is lost for at least several weeks,
if not permanently [100]. The mechanism behind this
observation is still unknown, but it is likely that the ISPS
protein is slowly or incompletely re-synthesized after the
stress [100]. Climate-change impact on isoprene emission
has been mainly attributed to positive long-term (enzy-
matic) and short-term (substrate) feedback of rising
temperature [104,105], implying that future emissions of
isoprene will also increase [106]. However, as drought is
often associated with heat waves and summer climate, the
finding that drought suppresses temperature-dependency
of isoprene emission may have consequences for trees’
thermal and ozone tolerance in regions (e.g. the Mediter-
ranean basin) which will be plagued by climate-change-
induced droughts associated with rising temperatures.
Laboratory measurements of the impact of drought on
monoterpene emissions are missing, but in field exper-
iments monoterpene emissions are inhibited by drought
stress [46,107,108]. Interestingly, the inhibition is particu-
larly evident only when drought stress is severe (e.g. the
water potential is less than 2 MPa) again suggesting that
Review Trends in Plant Science Vol.15 No.3
158
Author's personal copy
biosynthesis of volatile terpenes is resistant to mild
drought stress [109]. It is unclear whether alternative
carbon sources can be used to generate monoterpenes
under these conditions. As for temperature, heavily
drought-stressed oak plants lose the capacity to respond
to other environmental factors that are known to modulate
the emission, such as CO
2
(see below) [46].
Drought seems to have a different effect on pools of
monoterpenes stored in specialized structures or non-
stored in the mesophyll. In conifers, if the drought event
occurs in winter, when the biosynthesis of terpenes is
restrained by temperature, the pools appear to be nega-
tively affected. However, summer droughts can further
enrich the monoterpene pools. In non-storing species,
summer drought depletes the pools of terpenes in the
mesophyll which are under direct control of photosynthetic
carbon via the MEP pathway [110].
As described above, the emission of oxygenated BVOCs
depends on stomatal opening [84]. Stomatal closure in
response to drought and salt stress is therefore expected
to reduce particularly the emission of these compounds.
However, morning peaks of acetone and methanol emis-
sions may be very high in drought-stressed leaves, because
the temporary opening of stomata during times of higher
humidity allows the release of large pools of oxygenated
BVOCs that were built inside the leaf mesophyll [111,112].
Thus, drought stress does not inhibit per se the biosyn-
thesis of oxygenated BVOCs. Moreover, if the stress
reaches levels that are able to damage membranes and
cell walls, further increments of the emission of C6 com-
pounds and methanol must be expected. However, emis-
sion of these compounds has not been reported generally in
response to drought and salt [101]. Indeed, C6 emissions
occur in bursts, immediately after the damage to cellular
structures has occurred [5,113]. A recent experiment has
established a correlation between the emission of C6 vola-
tiles and the damage to membranes, as assessed by the ion
leakage, under a developing drought [114].
Ozone and other oxidants
In the atmosphere, volatile terpenes perform a dual action,
depending on the presence of anthropogenic pollutants.
When these compounds are absent, volatile terpenes
cleanse the atmosphere of ozone. In the presence of NO
x
,
however, these BVOCs initiate reactions leading to
increased ozone formation [13]. The chemical reactivity
of terpenes in the atmosphere led to the idea that volatile
terpenes play a similar dual action inside the leaves, before
they are released into the atmosphere [115,116].
In general, terpenes have been demonstrated to reduce
ozone damage and to quench ozone and reactive oxygen
species (ROS) [115118]. The mechanism(s) by which this
protective effect occurs is still under investigation and the
prevalent theories will be discussed below. Here we con-
centrate on the impact of oxidative stresses on the emission
of BVOCs. If isoprene reacts with ozone and other oxidative
species then it is expected that it disappears, concurrently
with the appearance of its reaction products. In the atmos-
phere, these products are mainly methyl-vinyl-ketone and
methacrolein, two compounds that are indeed found in
chamber studies in which isoprene-emitting trees were
fumigated with ozone [119]. In plants exposed to ozone
either a reduction [113,120] or a stimulation of isoprene
and monoterpene emissions is observed. The stimulation of
the emission is more evident in response to acute and
heavy doses of ozone [121123] (e.g. 150300 ppb) whereas
it is often absent when plants are exposed to low doses of
ozone above background [122,124]. Ozone-enhanced emis-
sion of isoprene is due to a higher expression of the ISPS
mRNA which probably upregulates the protein and the
activity of the enzyme [123]. Interestingly, such an upre-
gulation is more evident in leaves that develop under
enriched ozone and that build up a better resistance to
pollutants, as well as in new leaves that develop above
those that have been ozonated [123]. Evidently, impair-
ment of photosynthetic activities prevents older ozonated
leaves from enhancing the secondary metabolism, leading
to volatile terpene formation.
When plants are exposed to low or moderate and chronic
doses of ozone, such an induction of the volatile terpene
pathway is absent, and the expression of ISPS mRNA and
the level of ISPS protein may even be reduced [120,124].
Clearly, the signals that activate the biochemistry of ter-
pene formation, and which are unknown at present, are not
released under these conditions. It has been hypothesized
[125] that the induction of volatile terpenes in response to
ozone follows a hormetic doseresponse relationship, i.e.
that the volatile terpenes are increasingly induced at low
but growing doses of ozone until an ozone threshold is
reached after which the biosynthesis of terpenes is
repressed. The experiments, however, reveal a more com-
plex picture, with biosynthesis of volatile terpenes being
induced only when ozone dose overcomes a threshold that
marks cellular damage but at which photosynthesis is not
yet so heavily suppressed to be unable to supply enough
substrate for volatile terpenes. On the basis of these find-
ings a long-term induction of isoprene biosynthesis, and
the consequent evolutionary hypothesis that high terpene
emitters will be favoured in a future more oxidative atmos-
phere [126] may not hold true.
Ozone is a very damaging pollutant for plant cells, and
one of the first recognized ozone effects is the denaturation
of the lipids in cellular membranes [127]. It is therefore
expected that volatiles that are associated with lipid per-
oxidation are also more emitted in ozone-stressed leaves.
Accordingly, bursts of C6 compound emissions were
observed in ozone-stressed leaves and the lag time with
which these compounds were emitted was proportional to
the ozone dose absorbed by the leaves [128] and to the
ozone-induced injuries [113]. These experiments clearly
show that C6 compounds are in vivo indicators of mem-
brane denaturation and damage as already indicated also
in response to drought [114]. In addition, these studies
[113,128] also revealed transient pulses of methanol and
MeSA. While methanol emission has been mainly attrib-
uted to demethylation of cell wall pectins, and is therefore
likely to be another indicator of ozone damage, the induc-
tion of MeSA is more intriguing. MeSA emissions showed a
weak association with high levels of ozone recorded in non-
manipulative field experiments [98], but acute exposures
to ozone and UV light are apparently also able to induce
bursts of this compound (Velikova et al., personal com-
Review Trends in Plant Science Vol.15 No.3
159
Author's personal copy
munication). Emissions of MeSA induced by oxidative and
other environmental stresses may further contribute to
biogenic secondary organic aerosols [129,130] and may
alter the network of plant communication with other
organisms which is mediated by chemical messengers [98].
Other climate change factors: UV-B radiation
Studies of the effect of UV-B radiation (wavebands of 290
320 nm) on BVOCs emissions are rare, despite the fact that
UV-B radiation is known (i) in moderate doses (environ-
mentally relevant) as a priming abiotic agent, triggering
the formation of UV screening pigments and modulating
plant growth; and (ii) in higher doses (e.g. as a result of the
stratospheric ‘ozone hole’ and indeed transiently present in
certain regions surrounding the South Pole) as an agent
causing severe damage to the photosynthetic machinery,
nucleic acids, and proteins [127].
The impact of UV-B radiation on terpenes seems to
indicate a stimulation of their biosynthesis and emission.
In one study with European oak [131] the higher emission
of isoprene under UV-B radiation was attributed to a
higher biomass density rather than to a higher instan-
taneous photosynthetic rate. A more recent study [132]
again showed increased isoprene emission rates of sub-
arctic peatlands when irradiated with increasing levels of
UV-B radiation, and explained the rising emission as a
consequence of oxidative damage to membranes and to the
induction of the terpene defensive antioxidant pathway
[133].
UV damage to cellular structure may also induce emis-
sion of other BVOCs. Emission of methane under aerobic
conditions originating from plant material [10] is a con-
troversial topic in recent plant and atmospheric research,
e.g. [134,135]. Plant-mediated transport of methane
originating from methanogenic soil microorganisms
through the aerenchyma and out of the leaves of wetland
plants, e.g. rice [136] has been known from many years.
However, there is growing evidence that UV radiation can
mediate non-microbial methane release from cell wall
material, in particular pectin. In the initial work [10] it
was suggested that the methoxyl groups of pectin can be
one, albeit not the only, source of aerobic methane, e.g.
[137,138]. It was recently demonstrated [137139] that the
release of methane from structural cell wall components of
fresh and dried leaf tissue was UV-B dependent. In line
with this observation, the authors [139] also showed that
the removal of methoxyl groups interrupted methane emis-
sion from UV-irradiated pectin. Very recently it was
demonstrated [140] that the emission of methane from cell
wall material is mediated by UV-generated ROS [hydroxyl
radicals (
OH) and singlet oxygen; but not hydrogen per-
oxide or superoxide radicals].
Environmental stresses as well as cellular signalling
processes involve the formation of ROS. Therefore it might
be speculated that aerobic methane formation is a com-
mon, yet overlooked part of plant stress responses com-
plementing the transient burst of BVOCs, i.e. methanol,
acetaldehyde, and C6 alcohols and aldehydes [67,79], and
the long-lasting stimulation of terpene biosynthesis. Sup-
port for this assumption is given by recent work demon-
strating increased emissions of methane upon bacterial
infection, chemical generation of ROS [139] or physical
injury [141].
Other climate change factors: atmospheric CO
2
Rising CO
2
concentration at global level is dramatically
affecting plant life. The primary effect is the increased
availability of substrate for Rubisco, and the consequent
enhancement of photosynthesis [142]. However, CO
2
is one
more factor that uncouples terpene metabolism and emis-
sions from photosynthesis.
As early as 1964, G. Sanadze [143] demonstrated that at
low CO
2
concentration photosynthesis is reduced, but iso-
prene emission increases. In a more recent laboratory
study the highest emission of isoprene was detected at
an intercellular CO
2
concentration (C
i
) of around 150200
ppmv, not far from the C
i
experienced in nature by leaves of
tree species [18], and then progressively decreased at
higher C
i
. Since then, many studies on different plant
species have shown an inhibition of isoprene biosynthesis
under higher than ambient CO
2
concentrations [144148].
For example elevated CO
2
in a natural CO
2
spring reduces
ISPS activity and isoprene emission from common reed
(Phragmites australis)[37]. Compared to isoprene, less is
known about the influence of CO
2
on light-dependent
monoterpene emission. A study on holm oak (Quercus ilex)
growing under CO
2
concentration double than ambient in
open top chambers [46] showed a parallel downregulation
of mono-TPS activities and monoterpene emission. How-
ever, studies conducted in holm oak plants growing for
their entire life in natural CO
2
springs under very high but
not steady CO
2
levels did not reveal substantial inhibition
of the emitted monoterpenes [149], probably because the
treatment was also associated with recurrent drought
stress episodes that could also affect the emission (see
above).
While a downregulation of ISPS and mono-TPS is gener-
ally associated with lower emission of isoprene and mono-
terpenes grown under elevated CO
2
[145], the enzyme
properties are not necessarily the factors controlling the
emissions. This control may be exerted primarily by
reduced substrate availability (DMADP and GDP) result-
ing from a downregulation of the MEP pathway. Although
this assumption has no direct support from the literature, a
simultaneous decrease in isoprene emission and DMADP
content was observed [145]. This reduction in isoprene
(and also monoterpene) emission under elevated CO
2
might result from enforced higher consumption rates of
cytosolic PEP (phosphoenolpyruvate) through PEP
carboxylase activity, thus lowering the rates of PEP trans-
port into the chloroplast, where PEP, in its dephosphory-
lated form (pyruvate) feeds into the MEP pathway. Based
on this work [145] and earlier suggestions by G. Sanadze
[143,150] a double carboxylation hypothesis was proposed
[148,151] for C3 plants (to which all of isoprene emitters
belong) with cytosolic PEP carboxylase and plastidic RuBP
carboxylase as antithetic precursors controlling the flow of
carbon to plastidic terpene metabolism (as well as to the
shikimic acid pathway and fatty acid biosynthesis) in
response to changes in CO
2
concentration. This idea gets
support from a study [152] showing an inverse relationship
of isoprene biosynthesis and PEP carboxylase activity
Review Trends in Plant Science Vol.15 No.3
160
Author's personal copy
when cottonwoods (Populus deltoides) were grown on high
nitrate concentration, a condition favouring cytosolic
organic acids synthesis. In line with this scheme, mito-
chondrial respiration can constitute a growing sink for
cytosolic PEP under rising CO
2
, thus competing with the
import of this substrate in the chloroplast for terpene
biosynthesis [145]. Recent experiments with Free Air
CO
2
Enhancement (FACE) facilities, however, indicated
that the inverse relationship between isoprene and respir-
ation was not straightforward under elevated CO
2
, and
that competition occurs only when oxaloacetate production
from PEP for anabolic support of respiration is strong, such
as in young, expanding leaves [153]. A positive CO
2
effect
on isoprene [144] and monoterpene [154] emissions may be
occasionally observed. Moreover, no effect on isoprene and
monoterpenes was reported on poplar [125,153] or on Scots
pine (Pinus sylvestris)[155].
Currently we do not know whether the CO
2
impact on
light-dependent isoprene and monoterpene emission is a
general effect and which fundamental biochemical mech-
anisms are responsible. If the biosynthesis of volatile
terpenes is ubiquitously enhanced at CO
2
lower than
ambient, then stress conditions in which a lower intercel-
lular CO
2
concentration is set by increasing resistances to
CO
2
diffusion can generically lead to higher emissions.
Why is emission of volatile terpenes induced by
abiotic stress?
Whereas oxygenated BVOCs are mainly catabolic products
of the denaturation of cellular walls and membranes, iso-
prene and monoterpenes are produced through a dedicated
metabolic pathway that is stimulated by several abiotic
stresses (see above). This tightly regulated terpene biosyn-
thesis and the observation that emission of volatile ter-
penes represents a significant loss of photosynthetic
carbon, led to the proposition that these compounds play
important physiological and ecological roles in the protec-
tion of plants from environmental constraints. However,
the debate is still ongoing whether one or more ecological
actions should be attributed to volatile terpenes, and which
are the physiological mechanisms that allow terpenes to
exert their protective action. Two ‘metabolic’ hypotheses
suggest that isoprene acts as a kind of ‘safety valve’ which
allows quenching energy or metabolites (Figure 4). In
particular, one study [150] considers isoprene biosynthesis
as a pathway for dissipation of excess photosynthetic
energy, whereas a more recent one [145] postulates that
isoprene biosynthesis prevents the overflow of chloroplas-
tic DMADP. We consider as a ‘metabolic’ hypothesis also
the ‘opportunistic hypothesis’ [41] which suggests that
volatile terpenes are somehow alternative to ‘essential’
terpenes (such as carotenoids, also formed through the
MEP pathway, and for which an important antioxidant
role is clearly established). The same pool of carbon may
generate volatile and/or essential terpenes according to the
need to face different constraints. This may be true also
within the class of volatile terpenes, as it was recently
demonstrated that isoprene decreases when monoterpenes
are synthesized in poplar leaves attacked by beetles [156].
A second group of hypotheses establishes a more
distinctive functional role for volatile terpenes in plant
protection against abiotic stressors (Figure 4). The main
hypothesis is that isoprene, as well as monoterpenes, are
thermoprotective molecules, able to stabilize chloroplast
membranes during high temperature events [157] there-
fore protecting the photosynthetic apparatus. There is
ample experimental support for this idea. In brief: (i)
experiments in which plants have been fumigated with
terpenes or in which the biosynthesis of these compounds
have been chemically blocked have shown that the photo-
synthetic apparatus of isoprene-emitting plants is better
protected against heat stress [158160] and in particular
against rapid temperature changes [161]. The interesting
idea behind the latter observation is that isoprene acts as a
rapid mechanism of protection before plants can synthes-
ize more complex molecules (including non-volatile ter-
penes) that improve thermal stabilization. (ii) An in
silico experiment has demonstrated that isoprene may
indeed partition into the phospho-lipid bilayer of mem-
branes and maintain their stability, in particular during
exposure to high temperature [162]. (iii) Transgenic plants
that have been engineered to emit isoprene or monoter-
penes, or in which isoprene biosynthesis has been
repressed are now available. Isoprene-emitting transgenic
plants are more resistant to heat stress than wild types
[53,54,163].
Volatile terpenes also appear to have a relevant anti-
oxidant action. Again several lines of evidence support this
hypothesis: (i) isoprene and monoterpenes have been
shown to reduce the damage caused by ozone [115,122]
and ROS [117,133,164]. (ii) Evidence of the reaction be-
tween isoprene and monoterpenes and ozone or other ROS
has been produced by observing the appearance of the
reaction products and the disappearance of the reagents
[120,165]. In plants exposed to oxidative stresses isoprene
has also been shown to quench nitrogen reactive species
(namely NO) that can also have an important role as
messenger molecules of the hypersensitive response to
stress [166,167]. (iii) Transgenic tobacco (Nicotiana taba-
cum) plants that have been engineered to emit isoprene are
more resistant to ozone toxicity than non-emitting wild
types [168], and the ozone sensitivity of poplar clones is
inversely related to their capacity to emit isoprene [169].
However, both functions of isoprene have been chal-
lenged by studies in which the authors were not able to
reproduce improved resistance to the stresses, especially
when using artificial systems or transgenic plants. For
instance, photosynthesis analysis showed that transgenic
Arabidopsis plants do not benefit from isoprene when a
transient heat stress occurs [53]. In transgenic, isoprene-
emitting tobacco plants, tolerance of photosynthesis
against transient high-temperature episodes could only
be observed in lines emitting high levels of isoprene. More-
over, this effect was very mild and could only be identified
over repetitive stress events [168]. Finally, grey poplar
(Populus canescens) plants in which isoprene emission
was efficiently repressed were not more sensible to oxi-
dative stresses [113]. There may be good reasons that
explain why the protection offered by volatile terpenes is
not observed in specific cases: (i) the protection offered by
these molecules may not be achieved when photosynthesis
is already heavily impaired by prolonged or acute stresses,
Review Trends in Plant Science Vol.15 No.3
161
Author's personal copy
or when the stresses are too mild to affect photosynthesis.
Probably a window of stress exists in which these mol-
ecules substantially protect the photosynthetic apparatus.
(ii) In some transgenic plants, especially Arabidopsis
[53,54], the achieved emission of isoprene may be too
low to induce any significant physiological effect. (iii)
Downregulation or repression of the emission of isoprene
in natural emitters may cause a large induction of other
antioxidant molecules (e.g. ascorbate and tocopherol
[113]). This is actually another strong indication that
compensatory mechanisms are operated to replace the
antioxidant action of volatile terpenes.
The observation that volatile terpene emission is resist-
ant and is often induced by stress conditions, including
stress other than heat and strong oxidants (e.g. drought
[100] and salt [102]), has led to the hypothesis that a
unique mechanism, mostly related to the antioxidant
capacity of volatile terpenes, exists [12]. As for higher
terpenes which are known for their antioxidant function
(i.e. tocopherols, carotenoids, and sesquiterpenes), the
antioxidant action of volatile terpenes may be due to the
presence of conjugated double bounds [12]. Indeed, it might
be that the role of isoprene and monoterpenes in thermal
protection also comes from their antioxidant properties.
Hydrogen peroxide may also be produced by enhanced
photorespiration under moderately high temperatures.
The generation of ROS under heat stress has been reported
often [170,171]; however, there are many heat-tolerance
mechanisms in plants [172]. Probably, ROS scavenging
would not be sufficient to protect leaves from heavy heat
stress which directly affects thylakoid structure and the
functions associated with thylakoid intactness, namely
photochemical reactions. Thylakoid functionality would
be explained better by a mechanism counteracting thyla-
koid leakiness and the consequent increase of cyclic elec-
tron flow around photosystem I. Xanthophylls, a group of
non-volatile terpenes also originating from the MEP path-
way, may indeed render the thylakoid membranes more
resistant to heat [173], and this function may also be
carried out by volatile terpenes. The localization of ISPS
in the stromal side of thylakoidal membranes [35,174] and
the hydrophobic nature of isoprene are expected to assist
with its partition into photosynthetic membranes [162].
Lipophilic isoprene partitioned into membranes can also
prevent the formation of water channels responsible for the
membrane leakiness at high temperature [158,175]. Iso-
prene could also enhance hydrophobic interactions within
thylakoids and thereby stabilize interactions between
lipids and/or membrane proteins during episodes of
heat-shock or high temperature stress conditions [176].
Figure 4. Schematic overview of the proposed physiological functions of volatile terpenes: (a) In the ‘Metabolic overflow’ hypothesis cytosolic PEPC activity plays a central
role dividing the PEP pool into fractions (red line indicates the negative impact of PEPC activity on isoprene emission) available for isoprene biosynthesis and mitochondrial
metabolism or carbon metabolism (e.g. amino acid biosynthesis) (cellular scheme adapted from [152]) explaining the downregulation of isoprene emission under enhanced
CO
2
[145].(b) The thermoprotective function of isoprene is evident when comparing the net CO
2
assimilation in wild type and non-isoprene emitting poplars [163] when
short periods of heat were applied. (c) The property of isoprene to quench NO accumulation and ozone injury is demonstrated with transgenic tobacco leaves modified in
isoprene emission potential [168].(d) Interactions of isoprene with biomembranes are indicated by an in silico model study postulating a stabilizing effect of isoprene under
high temperature [162]. (Microscopic images of NO staining are provided by V. Velikova; photos from ozone-stressed tobacco leaves are provided by V. Velikova and C.
Vickers).
Review Trends in Plant Science Vol.15 No.3
162
Author's personal copy
Preliminary research using circular dichroism spec-
troscopy, a valuable tool for probing the molecular archi-
tecture of the complexes and supercomplexes and their
macro-organization in the membrane system [177]
confirms that a higher thermal stability of thylakoid mem-
branes is induced in transgenic plants that are able to emit
isoprene (Fortunati et al., personal communication).
Conclusions and future directions
We have seen that abiotic stresses may induce the emission
of multiple BVOCs. Many of the emitted compounds are
synthesized from the degradation of cellular structures
and may be used as reliable indicators of cell wall degra-
dation (methanol and methane) or membrane denatura-
tion (C6 volatiles). In the case of volatile terpenes the
induced emissions reflect the elicitation of the MEP path-
way, revealing important function(s) of these compounds in
the protection against stresses. The physiological and
ecological functions of volatile terpenes are well estab-
lished; however, more studies are needed to reveal the
molecular and biochemical mechanisms that oversee the
protective role of volatile terpenes. Induction or alteration
of these BVOCs emissions by abiotic stresses and other
climate change factors may also contribute to modify the
communication of plants with other organisms, namely
herbivores and carnivores that use BVOCs emissions as
olfactory cues to retrieve hosts suitable both as food and
shelter. Technical advances have made it possible to detect
induced emissions of MeSA not only in response to biotic
stresses but also in plants subjected to abiotic stresses.
This volatile is therefore emerging as a central molecule for
the signalling of stresses and for the consequent activation
of systemic acquired resistance and hypersensitive
responses.
Acknowledgments
We thank I. Zimmer for critical reading of the manuscript and C. Vickers
and V. Velikova for providing us with tobacco images. The work was
supported by the ESF Project Volatile Organic Compounds in the
Biosphere-Atmosphere System (VOCBAS). The German Research
Foundation (DFG; SCHN653/4 to J.-P.S.) supported the research within
the German joint research group ‘Poplar a model to address tree-specific
questions’.
References
1 Tholl, D. et al. (2006) Practical approaches to plant volatile analysis.
Plant J. 45, 540560
2 Knudsen, J.T. and Gershenzon, J. (2006) The chemical diversity of
floral scent. In Biology of Floral Scent (Dudareva, N. and Pichersky,
E., eds), pp. 2752, Taylor & Francis
3 Knudsen, J.T. et al. (1993) Floral scents: a check-list of volatile
compounds isolated by head-space techniques. Phytochemistry 33,
253280
4 Steeghs, M. et al. (2004) Proton-transfer-reaction mass spectrometry
as a new tool for real time analysis of root-secreted volatile organic
compounds in Arabidopsis.Plant Physiol. 135, 4758
5 Loreto, F. et al. (2006) On the induction of volatile organic compound
emissions by plants as consequence of wounding or fluctuations of
light and temperature. Plant Cell Environ. 29, 18201828
6 Pare
´, P.W. and Tumlinson, J.H. (1999) Plant volatiles as a defense
against insect herbivores. Plant Physiol. 121, 325332
7 Dudareva, N. et al. (2004) Biochemistry of plant volatiles. Plant
Physiol. 135, 19932011
8 Feussner, I. and Wasternack, C. (2002) The lipoxygenase pathway.
Annu. Rev. Plant Biol. 53, 275297
9 Kreuzwieser, J. et al. (1999) Metabolic origin of acetaldehyde emitted
by poplar (Populus tremula P. alba) trees. J. Exp. Bot. 50, 757
765
10 Keppler, F. et al. (2006) Methane emission from terrestrial plants
under aerobic conditions. Nature 439, 187191
11 Keppler, F. et al. (2008) Methoxyl groups of plant pectin as a precursor
of atmospheric methane: evidence from deuterium labelling studies.
New Phytol. 178, 808814
12 Vickers, C.E. et al. (2009) A unified mechanism of action for volatile
isoprenoids in plant abiotic stress. Nat. Chem. Biol. 5, 283291
13 Fehsenfeld, F. et al. (1992) Emissions of volatile organic compounds
from vegetation and the implications for atmospheric chemistry.
Global Biogeochem. Cyc. 6, 389430
14 Guenther, A.B. et al. (1995) A global model of natural volatile organic
compound emissions. J. Geophys. Res. 100, 88738892
15 Ghirardo, A. et al. (2010) Determination of de novo and pool
emissions of terpenes from four common boreal/alpine trees by
13
CO
2
labeling and PTR-MS analysis. Plant Cell Environ DOI:
10.1111/j.1365-3040.2009.02104.x (http://www3.interscience.wiley.
com/journal/123224612/abstract)
16 Loreto, F. et al. (200 9) One species, many terpenes: matching chemical
and biological diversity. Trends Plant Sci. 14, 416420
17 Kesselmeier, J. and Staudt, M. (1999) Biogenic volatile organic
compounds (VOC): an overview on emission, physiology and
ecology. J. Atmos. Chem. 33, 2388
18 Loreto, F. and Sharkey, T.D. (1990) A gas exchange study of
photosynthesis and isoprene emission in red oak (Quercus rubra
L.). Planta 182, 523531
19 Loreto, F. et al. (1996) Influence of environmental factors and air
composition on the emission of a-pinene from Quercus ilex leaves.
Plant Physiol. 110, 267275
20 Delwiche, C.F. and Sharkey, T.D. (1993) Rapid appearance of
13
Cin
biogenic isoprene when
13
CO
2
is fed to intact leaves. Plant Cell
Environ. 16, 587591
21 Rohmer, M. (2008) From molecular fossils of bacterial hopanoids to
the formation of isoprene units: Discovery and elucidation of the
methylerythritol phosphate pathway. Lipids 43, 10951107
22 Lichtenthaler, H.K. (1999) The 1-deoxy-D-xylulose 5-phosphate
pathway of isoprenoid biosynthesis in plants. Annu. Rev. Plant
Physiol. Plant Mol. Biol. 50, 4765
23 Sallaud, C. et al. (2009) A novel pathway for sesquiterpene
biosynthesis from Z,Z-farnesyl pyrophosphate in the wild tomato
Solanum habrochaites. Plant Cell 21, 301317
24 Schuhr, C.A. et al. (2003) Quantitative assessment of crosstalk
between the two isoprenoid biosynthesis pathways in plants by
NMR spectroscopy. Phytochem. Rev. 2, 316
25 Laule, O. et al. (2003) Crosstalk between cytosolic and plastidial
pathways of isoprenoid biosynthesis in Arabidopsis thaliana.Proc.
Natl. Acad. Sci. U. S. A. 100, 68666871
26 Schmidt, A. and Gershenzon, J. (2008) Cloning and characterization
of two different types of geranyl diphosphate synthases from Norway
spruce (Picea abies). Phytochemistry 69, 4957
27 Bohlmann, J. et al. (1998) Plant terpenoid synthases: molecular
biology and phylogenetic analysis. Proc. Natl. Acad. Sci. U. S. A.
95, 41264133
28 Vandermoten, S. et al. (2009) New insights into short-chain
prenyltransferases: structural features, evolutionary history and
potential for selective inhibition. Cell. Mol. Life Sci. 66, 3685
3695
29 Wise, M.L. and Croteau, R. (1999) Monoterpene biosynthesis. In
Comprehensive Natural Products Chemistry: Isoprenoids (Cane,
D.E., ed.), pp. 97153, Elsevier Science
30 Trapp, S.C. and Croteau, R.B. (2001) Genomic organization of plant
terpene synthases and molecular evolutionary implications. Genetics
158, 811832
31 Degenhardt, J. et al. (2009) Monoterpene and sesquiterpene
synthases and the origin of terpene skeletal diversity in plants.
Phytochemistry 70, 16211637
32 Gray, D. et al. (2006) Influences of temperature history, water stress
and needle age on methylbutenol emissions. Ecology 84, 765776
33 Silver, G.M. and Fall, R. (1995) Characterization of aspen isoprene
synthase, an enzyme responsible for leaf isoprene emission to the
atmosphere. J. Biol. Chem. 270, 1301013016
Review Trends in Plant Science Vol.15 No.3
163
Author's personal copy
34 Fisher, A.J. et al. (2000) Enzymatic synthesis of methylbutenol from
dimethylallyl diphosphate in needles of Pinus sabiniana.Arch.
Biochem. Biophys. 383, 128134
35 Schnitzler, J.P. et al. (2010) Poplar volatiles - biosynthesis, regulation
and (eco)physiology of isoprene and stress-induced isoprenoids. Plant
Biol. 12, 302316
36 Lehning, A. et al. (1999) Isoprene synthase activity and its relation to
isoprene emission in Quercus robur L. leaves. Plant Cell Environ. 22,
495504
37 Scholefield, P.A. et al. (2004) Impact of rising CO
2
on VOC emissions:
isoprene emission from Phragmites australis growing at elevated CO
2
in a natural carbon dioxide spring. Plant Cell Environ. 27, 381392
38 Sharkey, T.D. et al. (2005) Evolution of the isoprene biosynthetic
pathway in kudzu. Plant Physiol. 137, 700712
39 Fischbach, R.J. et al. (2000) Monoterpene synthase activities in leaves
of Picea abies (L.) Karst. and Quercus ilex L. Phytochemistry 54,
257265
40 Tholl, D. et al. (2001) Partial purification and characterization of the
short-chain prenyltransferases, geranyl diphosphate synthase and
farnesyl diphosphate synthase, from Abies grandis (Grand Fir).
Arch. Biochem. Biophys. 386, 233242
41 Owen, S. and Pen
˜uelas, J. (2005) Opportunistic emissions of volatile
isoprenoids. Trends Plant Sci. 10, 420426
42 Dudareva, N. et al. (2006) Plant volatiles: Recent advances and future
perspectives. Crit. Rev. Plant Sci. 25, 417440
43 Schnitzler, J.P. et al. (2004) Hybridisation of European oaks (Quercus
ilex xQ. robur) results in a mixed isoprenoid emitter type. Plant Cell
Environ. 27, 585593
44 Fischbach, R.J. et al. (2002) Seasonal pattern of monoterpene
synthase activities in leaves of the evergreen tree Quercus ilex L.
Physiol. Plant. 114, 354360
45 Grote, R. et al. (2009) Modelling the drought impact on monoterpene
fluxes from an evergreen Mediterranean forest canopy. Oecologia 160,
213223
46 Loreto, F. et al. (2001) Monoterpene emission and monoterpene
synthase activities in the Mediterranean evergreen oak Quercus
ilex L. grown at elevated CO
2
concentrations. Global Change Biol.
7, 709717
47 Fischbach, R.J. et al. (2001) Isolation and functional analysis of a
cDNA encoding a myrcene synthase of holm oak (Quercus ilex L.). Eur.
J. Biochem. 268, 56335638
48 Andre
`s-Montaner, D. (2009) Study of gene regulation and function of
monoterpene synthases in Holm oak (Quercus ilex L.) and transgenic
birch (Betula pendula Roth). PhD thesis, FZK-Bericht 7511
49 Aubourg, S. et al. (2002) Genomic analysis of the terpenoid synthase
(AtTPS) gene family of Arabidopsis thaliana.Mol. Genet. Genomics
267, 730745
50 Chen, F. et al. (2003) Biosynthesis and emission of terpenoids volatiles
from Arabidopsis flowers. Plant Cell 15, 481494
51 Tholl, D. et al. (2005) Two sesquiterpene synthases are responsible for
the complex mixture of sesquiterpenes emitted from Arabidopsis
flowers. Plant J. 42, 757771
52 Van Poecke, R.M.P. et al. (2001) Herbivore induced volatile production
by Arabidopsis thaliana leads to attraction of the parasitoid Cotesia
rubecula: Chemical, behavioural and gene-expression analysis. J.
Chem. Ecol. 27, 19111928
53 Loivama
¨ki, M. et al. (2007) Arabidopsis, a model to study biological
functions of isoprene emission? Plant Physiol. 144, 10661078
54 Sasaki, K. et al. (2007) Plants utilize isoprene emission as a
thermotolerance mechanism. Plant Cell Physiol. 48, 12541262
55 Loivama
¨ki, M. et al. (2008) Isoprene interferes with the attraction of
bodyguards by herbaceous plants. Proc. Natl. Acad. Sci. U. S. A. 105,
1743017435
56 Bonn, B. and Moortgart, G.K. (2003) Sesquiterpene ozonolysis: origin
of atmospheric new particle formation from biogenic hydrocarbons.
Geophys. Res. Lett. 30, 15851589
57 Arimura, G. et al. (2004) Forest tent caterpillars (Malacosoma
disstria) induce systemic and diurnal emissions of terpenoid
volatiles in hybrid poplar (Populus trichocarpa xdeltoides): cDNA
cloning, functional characterization, and patterns of gene expression
of ()-germacrene Dsynthase, PtdTPS1. Plant J. 37, 603616
58 Ralph, S.G. et al. (2008) Analysis of 4,664 high-quality sequence-
finished poplar full-length cDNA clones and their utility for the
discovery of genes responding to insect feeding. BMC Genomics 9,
5775
59 Hertzberg, M. et al. (2001) A transcriptional roadmap to wood
formation. Proc. Natl. Acad. Sci. U. S. A. 98, 1473214737
60 Ralph, S.G. et al. (2006) Genomics of hybrid poplar (Populus
trichocarpa x deltoides) interacting with forest tent caterpillars
(Malacosoma disstria): normalized and full-length cDNA libraries,
expressed sequence tags, and a cDNA microarray for the study of
insect-induced defences in poplar. Mol. Ecol. 15, 12751297
61 Tuskan, G. et al. (2006) The genome of black cottonwood, Populus
trichocarpa (Torr & Gray ex. Brayshaw). Science 313, 15961604
62 Guenther, A.B. (2002) The contribution of reactive carbon emissions
from vegetation to the carbon balance of terrestrial ecosystems.
Chemosphere 49, 837844
63 Ricard, J. and Noat, G. (1986) Electrostatic effects and the dynamics of
enzyme reactions at the surface of plant cells. 1. A theory of the ionic
control of a complex multi-enzyme system. Eur. J. Biochem. 155, 183
190
64 Gaffe, J. et al. (1994) Pectin methylesterase isoforms in tomato
(Lycopersicon esculentum) tissues. Effects of expression of a pectin
methylesterase antisense gene. Plant Physiol. 105, 199203
65 Mudgett, M.B. and Clarke, S. (1993) Characterization of plant L-
isoaspartyl methyltransferases that may be involved in seed survival.
Purification, characterization and sequence analysis of the wheat
germ enzyme. Biochemistry 32, 1110011111
66 Cossins, E.A. (1987) Folate biochemistry and the metabolism of one-
carbon units. In The Biochemistry of Plants (Vol 11) (Davies, D.D.,
ed.), In pp. 317353, Academic Press
67 Davison, B. et al. (2008) Cut-induced VOC emissions from agricultural
grasslands. Plant Biol. 10, 7685
68 Baldwin, I.T. et al. (2006) Volatile signaling in plant-plant
interactions: ‘talking trees’ in the genomics era. Science 311, 812
815
69 Fall, R. (2003) Abundant oxygenates in the atmosphere: a biochemical
perspective. Chem. Rev. 103, 49414951
70 Graus, M. et al. (2004) Transient release of oxygenated VOC during
lightdark transitions. Plant Physiol. 135, 19671975
71 Kreuzwieser, J. et al. (2001) Acetaldehyde emission by the leaves of
Trees - correlation with physiological and environmental parameters.
Physiol. Plant. 113, 4149
72 Holzinger, R. et al. (2000) Emissions of volatile organic compounds
from Quercus ilex L. measured by Proton Transfer Reaction Mass
spectrometry (PTR-MS) under different environmental conditions. J.
Geophys. Res. 105, 2057320579
73 Kreuzwieser, J. et al. (2002) Xylem-transported glucose as additional
carbon source for leaf isoprene formation in Quercus robur.New
Phytol. 156, 171178
74 Cojocariu, C. et al. (2004) Correlation of short-chained carbonyls
emitted from Picea abies with physiological and environmental
parameters. New Phytol. 162, 717727
75 Karl, T. et al. (2002) Transient releases of acetaldehyde from tree
leaves products of a pyruvate overflow mechanism? Plant Cell
Environ. 25, 11211131
76 Hatanaka, A. (1993) The biogeneration of green odour by green leaves.
Phytochemistry 34, 12011218
77 Croft, K.P.C. et al. (1993) Volatile products of the lipoxygenase
pathway evolved from Phaseolus vulgaris L. leaves inoculated
with Pseudomonas syringae pv phaseolicola.Plant Physiol. 101,
1324
78 Frost, C.J. et al. (2007) Within-plant signalling via volatiles
overcomes vascular constraints on systemic signalling and primes
responses against herbivores. Ecol. Lett. 10, 490498
79 Fall, R. et al. (1999) Volatile organic compounds emitted after leaf
wounding: on-line analysis by proton-transfer-reaction mass
spectrometry. J. Geophys. Res. 104, 1596315974
80 Flexas, J. et al. (2006) Diffusive and metabolic limitations to
photosynthesis under drought and salinity in C3 plants. Plant Biol.
6, 269279
81 Loreto, F. et al. (1996) Different sources of acetyl CoA contribute to
form three classes of terpenoid emitted by Quercus ilex L. leaves. Proc.
Natl. Acad. Sci. U. S. A. 93, 99669969
82 Schnitzler, J.P. et al. (2004) Contribution of different carbon sources to
isoprene biosynthesis in poplar leaves. Plant Physiol. 135, 152160
Review Trends in Plant Science Vol.15 No.3
164
Author's personal copy
83 Brilli, F. et al. (2007) Response of isoprene emission and carbon
metabolism to drought in white poplar (Populus alba) saplings.
New Phytol. 175, 244254
84 Niinemets, U
¨.et al. (2004) Physiological and physico-chemical
controls on foliar volatile organic compound emissions. Trends
Plant Sci. 9, 180186
85 Gershenzon, J. et al. (2000) Regulation of monoterpene accumulation
in leaves of peppermint. Plant Physiol. 122, 205214
86 Pasqua, G. et al. (2002) The role of isoprenoid accumulation and
oxidation in sealing wounded needles of Mediterranean pines.
Plant Sci. 163, 355359
87 Litvak, M.E. and Monson, R.K. (1998) Induced and constitutive
monoterpene defenses in conifer needles in relation to herbivory
patterns. Oecologia 114, 531540
88 Alessio, G.A. et al. (2004) Direct and indirect impacts of fire on the
isoprenoids emission from Mediterranean vegetation. Funct. Ecol. 18,
357364
89 Fall, R. and Monson, R.K. (1992) Isoprene emission rate and
intercellular isoprene concentration as influenced by stomatal
distribution and conductance. Plant Physiol. 100, 987992
90 Bru¨ ggemann, N. and Schnitzler, J.P. (2002) Comparison of isoprene
emission, intercellular isoprene concentration and photosynthetic
performance in water-limited oak (Quercus pubescens Willd. and
Quercus robur L.) saplings. Plant Biol. 4, 456463
91 Monson, R.K. et al. (199 2) Relationship among isoprene emission rate,
photosynthesis, and isoprene synthase activity as influenced by
temperature. Plant Physiol. 98, 11751180
92 Singsaas, E.L. and Sharkey, T.D. (2000) Regulation of isoprene
synthesis during high temperature stress. Plant Cell Environ. 23,
751757
93 Magel, E. et al. (2007) Determination of the role of products of
photosynthesis in substrate supply of isoprenoid biosynthesis in
poplar leaves. Atmos. Environ. 40, S138S151
94 Schade, G.W. and Goldstein, A.H. (2001) Fluxes of oxygenated volatile
organic compounds from a ponderosa pine plantation. J. Geophys. Res.
106, 31113124
95 Schade, G.W. and Custer, T.G. (2004) OVOC emissions from
agricultural soil in northern Germany during the 2003 European
heat wave. Atmos. Environ. 38, 61056114
96 Dicke, M. et al. (1999) Jasmonic acid and herbivory differentially
induce carnivore-attracting plant volatiles in lima bean plants. J.
Chem. Ecol. 25, 19071922
97 Filella, I. et al. (2006) Dynamics of the enhanced emissions of
monoterpenes and methyl salicylate, and decreased uptake of
formaldehyde, by Quercus ilex leaves after application of jasmonic
acid. New Phytol. 169, 135144
98 Karl, T. et al. (2008) Chemical sensing of plant stress at the ecosystem
scale. Biogeosciences 5, 12871294
99 Sharkey, T.D. and Loreto, F. (1993) Water stress, temperature, and
light effects on the capacity for isoprene emission and photosynthesis
of kudzu leaves. Oecologia 95, 328333
100 Fortunati, A. et al. (2008) Isoprene emission is not temperature-
dependent during and after severe drought-stress: a physiological
and biochemical analysis. Plant J. 55, 687697
101 Teuber, M. et al. (2008) VOC emission of Grey poplar leaves as
affected by salt stress and different N sources. Plant Biol. 10, 8696
102 Loreto, F. and Delfine, S. (2000) Emission of isoprene from
salt-stressed Eucalyptus globulus leaves. Plant Physiol. 123, 1605
1610
103 Loreto, F. et al. (2004)
13
C labelling reveals chloroplastic and extra-
chloroplastic pools of dimethylallyl pyrophosphate and their
contribution to isoprene formation. Plant Physiol. 135, 19031907
104 Lehning, A. et al. (2001) Modeling of annual variations of oak (Quercus
robur L.) isoprene synthase activity to predict isoprene emission
rates. J. Geophys. Res. 106, 31573166
105 Rennenberg, H. et al. (2006) Physiological responses of forest trees to
heat and drought. Plant Biol. 8, 556571
106 Arneth, A. et al. (2008) Effects of species composition, land surface
cover, CO
2
concentration and climate on isoprene emissions from
European forests. Plant Biol. 10, 150162
107 Bertin, N. and Staudt, M. (1996) Effect of water stress on
monoterpene emissions from young potted holm oak (Quercus ilex
L.) trees. Oecologia 107, 456462
108 Llusia, J. and Pen
˜uelas, J. (1998) Changes in terpene content and
emission in potted Mediterranean woody plants under severe
drought. Can. J. Bot. 76, 13661373
109 Lavoir, A.V. et al. (2009) Drought reduced monoterpene emissions
from Quercus ilex trees: Results from a throughfall displacement
experiment within a forest ecosystem. Biogeosciences 6, 863893
110 Llusia, J. et al. (2006) Seasonal contrasting changes of foliar
concentrations of terpenes and other volatile organic compounds in
four dominant species of a Mediterranean shrubland submitted to a
field experimental drought and warming. Physiol. Plant 127, 632649
111 Fares, S. et al. (2009) The ACCENT-VOCBAS field campaign on
biosphere-atmosphere interactions in a Mediterranean ecosystem
of Castelporziano (Rome): site characteristics, climatic and
meteorological conditions, and eco-physiology of vegetation.
Biogeosciences 6, 10431058
112 Filella, I. et al. (2009) Short-chained oxygenated VOC emissions in
Pinus halepensis in response to changes in water availability. Acta
Physiol. Plant. 31, 311318
113 Behnke, K. et al. (2009) RNAi mediated suppression of isoprene
biosynthesis impacts ozone tolerance. Tree Physiol. 29, 725736
114 Capitani, D. et al. (2009) In situ investigation of leaf water status by
portable unilateral NMR. Plant Physiol. 149, 16381647
115 Loreto, F. et al. (2001) Ozone quenching properties of isoprene and its
antioxidant role in leaves. Plant Physiol. 26, 9931000
116 Hewitt, N.C. et al. (1990) Hydroperoxides in plants exposed to ozone
mediate air pollution damage to alkene emitters. Nature 344, 5658
117 Affek, H.P. and Yakir, D. (2002) Protection by isoprene against singlet
oxygen in leaves. Plant Physiol. 129, 269277
118 Loreto, F. and Fares, S. (2007) Is ozone flux inside leaves only a
damage indicator? Clues from volatile isoprenoid studies. Plant
Physiol. 143, 10961100
119 Fares, S. et al. (2008) Stomatal uptake and stomatal deposition of
ozone in isoprene and monoterpene emitting plants. Plant Biol. 10,
4454
120 Ryan, A. et al. (2009) Defining hybrid poplar (Populus
deltoides Populus trichocarpa) tolerance to ozone: identifying key
parameters. Plant Cell Environ. 32, 3145
121 Loreto, F. et al. (2004) Impact of ozone on monoterpene emissions and
evidences for an isoprene-like antioxidant action of monoterpenes
emitted by Quercus ilex (L.) leaves. Tree Physiol. 24, 361367
122 Velikova, V. et al. (2005) Localized ozone fumigation system for
studying ozone effects on photosynthesis, respiration, electron
transport rate and isoprene emission in field-grown Mediterranean
oak species. Tree Physiol. 25, 15231532
123 Fares, S. et al. (2006) Impact of high ozone on isoprene emission,
photosynthesis and histology of developing Populus alba leaves
directly or indirectly exposed to the pollutant. Physiol. Plant. 128,
456465
124 Calfapietra, C. et al. (2007) Isoprene synthase expression and protein
levels are reduced under elevated O
3
but not under elevated CO
2
(FACE) in field-grown aspen trees. Plant Cell Environ. 30, 654
661
125 Calfapietra, C. et al. (2009) Volatile organic compounds from Italian
vegetation and their interaction with ozone. Environ. Poll. 157, 1478
1486
126 Lerdau, M. (2007) A positive feedback with negative consequences.
Science 316, 212213
127 Pell, E.J. et al. (1997) Ozone induced oxidative stress: mechanism of
action and reaction. Physiol. Plant. 100, 264273
128 Beauchamp, J. et al. (2005) Ozone induced emissions of biogenic VOC
from tobacco: relationships between ozone uptake and emission of
LOX products. Plant Cell Environ. 28, 13341343
129 Clays, M. et al. (2004) Formation of secondary organic aerosols
through photooxidation of isoprene. Science 303, 11731175
130 Kanakidou, M. et al. (2005) Organic aerosol and global climate
modeling: a review. Atmos. Chem. Phys. 5, 10531123
131 Harley, P. et al. (1996) Effects of elevated levels of UV-B radiation on
photosynthesis and isoprene emission in Gambel’s oak and velvet
bean. Global Change Biol. 2, 149154
132 Tiiva, P. et al. (2007) Isoprene emission from a subarctic peatland
under enhanced UV-B radiation. New Phytol. 176, 346355
133 Loreto, F. and Velikova, V. (2001) Isoprene produced by leaves
protects the photosynthetic apparatus against ozone damage,
Review Trends in Plant Science Vol.15 No.3
165
Author's personal copy
quenches ozone products, and reduces lipid peroxidation of cellular
membranes. Plant Physiol. 127, 17811787
134 Dueck, T.A. et al. (2007) No evidence for substantial aerobic methane
emission by terrestrial plants: A
13
C-labelling approach. New Phytol.
175, 2935
135 Nisbet, R.E.R. et al. (2009) Emission of methane from plants. Proc. R.
Soc. B-Bio. Sci. 276, 13471354
136 Butterbach-Bahl, K. et al. (1997) Impact of gas transport through rice
cultivars on methane emission from rice paddy fields. Plant Cell
Environ. 20, 11751183
137 Vigano, I. et al. (2008) Effect of UV radiation and temperature on the
emission of methane from plant biomass and structural components.
Biogeosciences 5, 937947
138 Bru¨ ggemann, N. et al. (2009) Nonmicrobial aerobic methane emission
from poplar shoot cultures under low-light conditions. New Phytol. 23,
912918
139 McLeod, A.R. et al. (2008) Ultraviolet radiation drives methane
emissions from terrestrial plant pectins. New Phytol. 180, 124132
140 Messenger, D.J. et al. (2009) The role of ultraviolet radiation,
photosensitizers, reactive oxygen species and ester groups in
mechanisms of methane formation from pectin. Plant Cell Environ.
32, 19
141 Wang, Z.P. et al. (2009) Physical injury stimulates aerobic methane
emissions from terrestrial plants. Biogeosciences 6, 615621
142 Long, S.P. et al. (2004) Rising atmospheric carbon dioxide: plants face
the future. Annu. Rev. Plant Biol. 55, 591628
143 Sanadze, G.A. (1964) Light-dependent excretion of isoprene by plants.
Photosyn. Res. 2, 701707
144 Sharkey, T.D. et al. (1991) High carbon dioxide and sun/shade effects
on isoprene emission from oak and aspen tree leaves. Plant Cell
Environ. 14, 333338
145 Rosenstiel, T.N. et al. (2003) Increased CO
2
uncouples growth from
isoprene emission in an agriforest ecosystem. Nature 421, 256259
146 Centritto, M. et al. (2004) Profiles of isoprene emission and
photosynthetic parameters in hybrid poplars exposed to free-air
CO
2
enrichment. Plant Cell Environ. 27, 403412
147 Pegoraro, E. et al. (2004) Effect of elevated CO
2
concentration and
vapour pressure deficit on isoprene emission from leaves of Populus
deltoides during drought. Funct. Plant Biol. 31, 11371147
148 Wilkinson, M.J. et al. (2009) Leaf isoprene emission rate as a function
of atmospheric CO
2
concentration. Global Change Biol. 15, 11891200
149 Rapparini, F. et al. (2004) Isoprenoid emission in trees of Quercus
pubescens and Quercus ilex with lifetime exposure to naturally high
CO
2
environment. Plant Cell Environ. 27, 381391
150 Sanadze, G.A. (2004) Biogenic isoprene (A review). Russ. J. Plant
Physiol. 51, 729741
151 Monson, R.K. et al. (2009) Biochemical controls of the CO
2
response of
leaf isoprene emission: an alternative view of Sanadze’s double
carboxylation scheme. Ann. Agrar. Sci. 7, 2129
152 Rosenstiel, T.N. et al. (2004) Induction of poplar leaf nitrate
reductase: a test of extrachloroplastic control of isoprene emission
rate. Plant Biol. 6, 1221
153 Loreto, F. et al. (2007) The relationship between isoprene emission
rate and dark respiration rate in white poplar (Populus alba L.)
leaves. Plant Cell Environ. 30, 662669
154 Staudt, M. et al. (2001) Effect of elevated CO
2
on monoterpene
emission of young Quercus ilex L. trees and its relation to
structural and ecophysiological parameters. Tree Physiol. 21, 437
445
155 Ra
¨isa
¨nen, T. et al. (2008) Effects of elevated CO
2
and temperature on
monoterpene emission of Scots pine (Pinus sylvestris L.). Atmos.
Environ. 42, 41604171
156 Brilli, F. et al. (2009) Constitutive and herbivore-induced
monoterpenes emitted by Populus xeuroamericana leaves are key
volatiles that orient Chrysomela populi beetles. Plant Cell Environ.
32, 542552
157 Sharkey, T.D. and Singsaas, E.L. (1995) Why plants emit isoprene.
Nature 374, 7691769
158 Singsaas, E.L. et al. (1997) Isoprene increases thermotolerance of
isoprene-emitting species. Plant Physiol. 115, 14131420
159 Loreto, F. et al. (1998) On the monoterpene emission under heat stress
and on the increased thermotolerance of leaves of Quercus ilex
fumigated with selected monoterpenes. Plant Cell Environ. 21,
101107
160 Velikova, V. and Loreto, F. (2005) On the relationship between
isoprene emission and thermotolerance in Phragmites australis
leaves exposed to high temperatures and during the recovery from
a heat stress. Plant Cell Environ. 28, 318327
161 Sharkey, T.D. et al. (2008) Isoprene emission from plants: Why and
how. Ann. Bot. 101, 518
162 Siwko, M.E. et al. (2007) Does isoprene protect plant membranes from
thermal shock? A molecular dynamics study. Biochim. Biophys. Acta -
Biomembranes 1768, 198206
163 Behnke, K. et al. (2007) Transgenic, non-isoprene emitting poplars
don’t like it hot. Plant J. 51, 485499
164 Velikova, V. et al. (2004) Endogenous isoprene protects Phragmites
australis leaves against singlet oxygen. Physiol. Plant. 122, 219225
165 Pinto, D.M. et al. (2007) The effects of increasing atmospheric ozone
on biogenic monoterpene profiles and the formation of secondary
aerosols. Atmos. Environ. 41, 48774887
166 Velikova, V. et al. (2005) Isoprene decreases the concentration of nitric
oxide in leaves exposed to elevated ozone. New Phytol. 166, 419426
167 Velikova, V. et al. (2008) Isoprene and nitric oxide reduce damages in
leaves exposed to oxidative stress. Plant Cell Environ. 31, 18821894
168 Vickers, C.E. et al. (2009) Isoprene synthesis protects tobacco plants
from oxidative stress. Plant Cell Environ. 32, 520531
169 Calfapietra, C. et al. (2008) Isoprene emission rates under elevated
CO
2
and O
3
in two field-grown aspen clones differing for their
sensitivity to O
3
.New Phytol. 179, 5561
170 Valle
´lian-Bindschedler, L. et al. (1998) Heat-induced resistance in
barley to powdery mildew (Blumeria graminis f.sp. hordei)is
associated with a burst of active oxygen species. Physiol. Mol.
Plant Pathol. 52, 185199
171 Vacca, R.A. et al. (2004) Production of reactive oxygen species,
alteration of cytosolic ascorbate peroxidase, and impairment of
mitochondrial metabolism are early events in heat shock-induced
programmed cell death in tobacco bright-yellow 2 cells. Plant Physiol.
134, 11001112
172 Sharkey, T.D. and Schrader, S.M. (2006) High temperature stress. In
Physiology and Molecular Biology of Stress Tolerance in Plants
(Madhava Rao, K.V. et al., eds), pp. 101129, Springer
173 Havaux, M. and Tardy, F. (1996) Temperature-dependent adjustment
of the thermal stability of photosystem II in vivo: possible
involvement of xanthophyll-cycle pigments. Planta 198, 324333
174 Wildermuth, M.C. and Fall, R. (1998) Biochemical characterization of
stromal and thylakoid-bound isoforms of isoprene synthase in willow
leaves. Plant Physiol. 116, 11111123
175 Sharkey, T.D. et al. (2001) Isoprene increases thermotolerance of
fosmidomycin-fed leaves. Plant Physiol. 125, 20012006
176 Sharkey, T.D. and Yeh, S. (2001) Isoprene emissions from plants.
Annu. Rev. Plant Physiol. Plant Mol. Biol. 52, 407436
177 Garab, G. and van Amerongen, H. (2009) Linear dichroism (LD) and
circular dichroism (CD) in photosynthesis research. Photosyn. Res.
101, 135146
Review Trends in Plant Science Vol.15 No.3
166
... Biogenic volatile organic compounds (BVOCs) are produced by various plant species as part of their physiological metabolism [1]. The high reactivity of BVOCs contributes to the formation of troposphere air pollutants, such as ozone, secondary organic aerosols, and fine particulate matter [2]. ...
Article
Full-text available
Biogenic volatile organic compounds (BVOCs) significantly contribute to atmospheric chemistry at both regional and global scales. The composition and intensity of BVOC emissions vary significantly among different plant species. Previous studies have focused on BVOC emissions from tree species, but the results of research on BVOC emissions from wetland plants are still limited. Therefore, in this study, BVOCs emitted by three aquatic plants (Phragmites australis, Typha angustifolia, and Iris pseudacorus) were sampled and analyzed using a dynamic headspace technique combined with GC-MS at daily scales. The diurnal observation data showed that the total BVOC emission rates of the three plants peaked with the increase in environmental factors (temperature, PAR, and water temperature). P. australis was the only of the three plants that emitted isoprene with a high rate of 48.34 μg·g−1Dw·h−1. Moreover, the peak emission rates of total BVOC (78.45 μg·g−1Dw·h−1) in P. australis were higher than most tree species. The emissions rates of volatile organic compounds, including monoterpenes, oxygenated volatile organic compounds, alkanes, and other volatile organic compounds, were statistically correlated across all species. The emission rates of isoprene from P. australis had significant associations with intercellular CO2 concentration (Ci) (0.58, p < 0.05) and transpiration rate (Tr) (−0.63, p < 0.01). The emission rates of monoterpenes from P. australis were found to have a significantly positive correlation with the net photosynthetic rate (Pn) (0.58, p < 0.05) while T. angustifolia (−0.59, p < 0.05) and I. pseudacorus (−0.47, p < 0.05) showed the opposite trend. Such findings hold significance for the refinement of localized emission inventories and the development of comprehensive emission process models in future research, as BVOC emissions from wetland plants were reported here for the first time.
Chapter
Medicinal and aromatic plants (MAPs) are known for their valuable secondary metabolites, such as terpenoids and steroids, fatty acid-derived substances, and alkaloids, which possess various therapeutic and aromatic properties. However, these plants are often exposed to various abiotic stresses such as drought, salinity, extreme temperatures, and heavy metals, which negatively impact their growth and essential metabolite production. However, in response to these abiotic/environmental stresses, plants evolve various physiological and biochemical adaptation mechanisms for their protection and survival. These mechanisms include adjustment of the plant membrane system, production of reactive oxygen species (ROS) and phytohormone, and accumulation of osmolytes in the cellular compartments of plant cells. These vital compounds act as a defense system for plants in adverse environmental conditions. Secondary metabolites in plants are also affected by stress conditions. However, they play a major role in the adaptation and survival of plants to the changing environment and stress conditions. The level of stress in medicinal and aromatic plants leads to production of the secondary metabolites in the plants. This chapter will describe a detailed review of the effect of abiotic stress on secondary metabolites of different medicinal and aromatic plants.
Article
Full-text available
Detecting volatile organic compounds (VOCs) emitted from different plant species andtheir organs can provide valuable information about plant health and environmental factors thataffect them. For example, limonene emission can be a biomarker to monitor plant health and detectstress. Traditional methods for VOC detection encounter challenges, prompting the proposal ofnovel approaches. In this study, we proposed integrating electrospinning, molecular imprinting,and conductive nanofibers to fabricate limonene sensors. In detail, polyvinylpyrrolidone (PVP)and polyacrylic acid (PAA) served here as fiber and cavity formers, respectively, with multiwalledcarbon nanotubes (MWCNT) enhancing conductivity. We developed one-step monolithic molecularlyimprinted fibers, where S(−)-limonene was the target molecule, using an electrospinning technique.The functional cavities were fixed using the UV curing method, followed by a target molecule washing.This procedure enabled the creation of recognition sites for limonene within the nanofiber matrix,enhancing sensor performance and streamlining manufacturing. Humidity was crucial for sensorworking, with optimal conditions at about 50% RH. The sensors rapidly responded to S(−)-limonene,reaching a plateau within 200 s. Enhancing fiber density improved sensor performance, resulting in alower limit of detection (LOD) of 137 ppb. However, excessive fiber density decreased accessibility toactive sites, thus reducing sensitivity. Remarkably, the thinnest mat on the fibrous sensors createdprovided the highest selectivity to limonene (Selectivity Index: 72%) compared with other VOCs,such as EtOH (used as a solvent in nanofiber development), aromatic compounds (toluene), and twoother monoterpenes (α-pinene and linalool) with similar structures. These findings underscored thepotential of the proposed integrated approach for selective VOC detection in applications such asprecision agriculture and environmental monitoring.
Article
Bryophytes desiccate rapidly when relative humidity decreases. The capacity to withstand dehydration depends on several ecological and physiological factors. Volatile organic compounds (VOCs) may have a role in enhancing tolerance to desiccating bryophytes. However, the functions of VOCs in bryophytes have received little attention so far. We aimed to investigate the impact of a dehydration‐rehydration treatment on primary carbon metabolism and volatile terpenes (VTs) in three bryophytes with contrasting ecological traits: Vessicularia dubyana , Porella platyphylla and Pleurochaete squarrosa . First, we confirmed the desiccation sensitivity gradient of the species. Under fully hydrated conditions, the photosynthetic rate (A) was inversely associated with stress tolerance, with a lower rate in more tolerant species. The partial recovery of A in P. platyphylla and P. squarrosa after rehydration confirmed the desiccation tolerance of these two species. On the other hand, A did not recover after rehydration in V. dubyana . Regarding VT, each species exhibited a distinct VT profile under optimum hydration, with the highest VT pool found in the more desiccation‐sensitive species ( V. dubyana ). However, the observed species‐specific VT pattern could be associated with the ecological habitat of each species. P. squarrosa , a moss of dry habitats, may synthesize mainly non‐volatile secondary metabolites as stress‐defensive compounds. On the other hand, V. dubyana , commonly found submerged, may need to invest photosynthetically assimilated carbon to synthesize a higher amount of VTs to cope with transient water stress occurrence. Further research on the functions of VTs in bryophytes is needed to deepen our understanding of their ecological significance.
Article
Full-text available
Organic aerosol formed from ocimene photooxidation has more oligomers compared to organic aerosol formed from other acyclic terpene precursors.
Article
Full-text available
The effects of water limitations on the emission of biogenic volatile organic compounds are not well understood. Experimental approaches studying drought effects in natural conditions are still missing. To address this question, a throughfall displacement experiment was set up in a natural forest of Quercus ilex, an evergreen Mediterranean oak emitting monoterpenes. Mature trees were exposed in 2005 and 2006 either to an additional drought, to irrigation or to natural drought (untreated control). In both years, absolute monoterpene emission rates as well as the respective standard factors of the trees exposed to normal and additional drought strongly declined during the drought periods. Monoterpene emissions were lower in year 2006 than in year 2005 (factor 2) due to a more pronounced summer drought period in this respective year. We observed a significant difference between the irrigation and additional drought or control treatment: irrigated trees emitted 82% more monoterpenes during the drought period 2006 than the trees of the other treatments. However, no significant effect on monoterpene emission was observed between normal and additional drought treatments, despite a significant effect on leaf water potential and photochemical efficiency. During the development of drought, monoterpene emissions responded exponentially rather than linearly to decreasing leaf water potential. Emissions rapidly declined when the water potential dropped below - 2 MPa and photosynthesis was persistently inhibited. Monoterpene synthase activities measured in vitro showed no clear reduction during the same period. From our results we conclude that drought significantly reduces monoterpene fluxes of Mediterranean Holm oak forest into the atmosphere due to a lack of primary substrates coming from photosynthetic processes.
Chapter
Fragrant flowers emit aroma bouquets containing anywhere from a few to more than one hundred different compounds present in varying amounts. The number and diversity of these constituents make floral scent one of the most variable aspects of plant phenotype. Yet, floral scent compounds have a common function in promoting cross pollination, a vital process in the life cycle of most plants. Accompanied by visual and tactile cues, these substances mediate plant-pollinator interactions, which usually benefit both partners. In return for being pollinated, pollinators are rewarded with pollen, nectar, and oils. However, deceptive pollination systems that involve floral scent as an attractant are also well known. Because of its commercial importance in the perfume and flavor industries, the chemistry of floral scents has been studied for a long time. Among the most striking recent examples is a series of papers published during the last 15 years by Roman Kaiser and collaborators at Givaudan Roure. These authors have contributed significantly to our knowledge of floral scents in the orchid family,1-2 and Kaiser and his collaborators have presented many astonishing new scent molecules from various other plant families.3-14 Apart from many studies of medicinal plants,15-17 e.g., most investigators have focused on species from one or a few plant families, either from the perspective of pollination biology and/or chemotaxonomy.
Chapter
This chapter focuses on folate biochemistry and the metabolism of one-carbon unit. The one-carbon metabolism of prokaryotic and eukaryotic cells involves reactions that are mainly folate-dependent. It is clear that the pathways of c1 metabolism are of prime importance, as the synthesis of several cellular constituents depends directly or indirectly on folate metabolism. The chapter also discusses the techniques for the isolation and assay of folate derivatives. Folate derivatives commonly show different degrees of oxidation, c1 unit substitution, and glutamyl conjugation. Folates are isolated by methods based on the principles of ion exchange, gel filtration, and high-performance chromatography. These methods have provided excellent data, particularly when analyses are carried out on hydrolyzed samples. A variety of chemical cleavage methods, combined with chromatographic separations, have been used successfully in recent studies of naturally occurring polyglutamyl folates. It is still customary to use a standard microbiological assay procedure whenever the investigator requires information on total folate concentrations. In clinical laboratories, where folate assays are carried out for their diagnostic value, the more rapid radio-assay methods are now widely used. These analytical methods are reviewed with special emphasis on their application to studies of plant folates.
Article
We investigated growth, leaf monoterpene emission, gas exchange, leaf structure and leaf chemical composition of 1-year-old Quercus ilex L. seedlings grown in ambient (350 mul l(-1)) and elevated (700 mul l(-1)) CO2 concentrations ([CO2]). Monoterpene emission and gas exchange were determined at constant temperature and irradiance (25 degreesC and 1000 mu mol m(-2) s(-1) of photosynthetically active radiation) at an assay [CO2] of 350 or 700 mul l(-1). Measurements were made on intact shoots after the end of the growing season between mid-October and mid-February. On average, plants grown in elevated [CO2] had significantly increased foliage biomass (about 50%). Leaves in the elevated [CO2] treatment were significantly thicker and had significantly higher concentrations of cellulose and lignin and significantly lower concentrations of nitrogen and minerals than leaves in the ambient [CO2] treatment. Leaf dry matter density and leaf concentrations of starch, soluble sugars, lipids and hemi-cellulose were not significantly affected by growth in elevated [CO2]. Monoterpene emissions of seedlings were significantly increased by elevated [CO2] but were insensitive to short-term changes in assay [CO2]. On average, plants grown in elevated [CO2] had 1.8-fold higher monoterpene emissions irrespective of the assay [CO2]. Conversely, assay [CO2] rapidly affected photosynthetic rate, but there was no apparent long-term acclimation of photosynthesis to growth in elevated [CO2]. Regardless of growth [CO2]. photosynthetic rates of all plants almost doubled when the assay [CO2] was switched from 350 to 700 mul l(-1). At the same assay [CO2], mean photosynthetic rates of seedlings in the two growth CO2 treatments were similar. The percentage of assimilated carbon lost as monoterpenes was not significantly altered by CO2 enrichment. Leaf emission rates were correlated with leaf thickness, leaf concentrations of cellulose, lignin and nitrogen, and total plant leaf area. In all plants, monoterpene emissions strongly declined during the winter independently of CO2 treatment. The results are discussed in the context of the acquisition and allocation of resources by e. ilex seedlings and evaluated in terms of emission predictions.
Article
Atmospheric new aerosol particle formation observed in remote areas (e.g., in Finland, Portugal and in the U.S.), is generally attributed to low-volatile oxidation products of monoterpenes (C10H16), which are emitted by the vegetation. In this article we show that this atmospheric new particle formation is not caused by monoterpene products, but most likely initiated by very low-volatile substances produced during sesquiterpene (C15H24)-ozone reactions. For this purpose, the nucleation times of the most abundant monoterpene reactions have been calculated and discussed exemplarily for the Finnish site Hyytiälä, at which nucleation events have been observed. In addition, the important negative influence of water vapor on the nucleation threshold of the β-caryophyllene-ozone reaction has been studied in detail at different dew points in the laboratory. Therein, the saturation vapor pressure of the nucleating compounds was estimated to be less than 1.2 · 10-10 hPa, which is recommended for atmospheric homogeneous nucleation of non-volatile organics.
Article
In this review we explore several models which might explain ozone (O-3)-induced injury to plant foliage. Ozone enters the cell through the wall and plasma membrane where active oxygen species are generated. If the concentration of O-3 is very high, unregulated cell death will occur. Alternatively, the active oxygen species, or succeeding reaction products, may serve as elicitors of regulated plant responses. These regulated responses include the induction of ethylene which could serve as a primary signal for - or a facilitator of - subsequent responses. The role of regulated suppression of photosynthetic genes and induction of chitinases and beta-1,3-glucanase in programmed cell death is explored. Induction of antioxidants, enzymes of lignification and glutathione-S-transferase are discussed in the context of O-3-induced cell repair or cell protection. A second model is postulated to explain induction of accelerated foliar senescence by low levels of O-3. The notion that O-3-induced elicitation of responses in the nucleus might lead to increased oxidative stress in the chloroplast is considered as a mechanism for accelerating the rate of degradation of ribulose-1, 15-bisphosphate carboxylase/oxygenase (Rubisco). The mechanisms by which O-3 induces loss of Rubisco, and the relationship to accelerated foliar senescence are discussed.
Article
Isoprene is formed in and emitted by plants and the reason for this apparent carbon waste is still unclear. It has been proposed that isoprene stabilizes cell and particularly chloroplast thylakoid membranes. We tested if membrane stabilization or isoprene reactivity with ozone induces protection against acute ozone exposures. The reduction of visible, physiological, anatomical, and ultrastructural (chloroplast) damage shows that clones of plants sensitive to ozone and unable to emit isoprene become resistant to acute and short exposure to ozone if they are fumigated with exogenous isoprene, and that isoprene-emitting plants that are sensitive to ozone do not suffer damage when exposed to ozone. Isoprene-induced ozone resistance is associated with the maintenance of photochemical efficiency and with a low energy dissipation, as indicated by fluorescence quenching. This suggests that isoprene effectively stabilizes thylakoid membranes. However, when isoprene reacts with ozone within the leaves or in a humid atmosphere, it quenches the ozone concentration to levels that are less or non-toxic for plants. Thus, protection from ozone in plants fumigated with isoprene may be due to a direct ozone quenching rather than to an induced resistance at membrane level. Irrespective of the mechanism, isoprene is one of the most effective antioxidants in plants.