ArticlePDF AvailableLiterature Review

The Layers of Plant Responses to Insect Herbivores

Authors:

Abstract and Figures

Plants collectively produce hundreds of thousands of specialized metabolites that are not required for growth or development. Each species has a qualitatively unique profile, with variation among individuals, growth stages, and tissues. By the 1950s, entomologists began to recognize the supreme importance of these metabolites in shaping insect herbivore communities. Plant defense theories arose to address observed patterns of variation, but provided few testable hypotheses because they did not distinguish clearly among proximate and ultimate causes. Molecular plant-insect interaction research has since revealed the sophistication of plant metabolic, developmental, and signaling networks. This understanding at the molecular level, rather than theoretical predictions, has driven the development of new hypotheses and tools and pushed the field forward. We reflect on the utility of the functional perspective provided by the optimal defense theory, and propose a conceptual model of plant defense as a series of layers each at a different level of analysis, illustrated by advances in the molecular ecology of plant-insect interactions. Expected final online publication date for the Annual Review of Entomology Volume 61 is January 07, 2016. Please see http://www.annualreviews.org/catalog/pubdates.aspx for revised estimates.
Content may be subject to copyright.
EN61CH20-Schuman ARI 3 December 2015 21:34
R
E
V
I
E
W
S
I
N
A
D
V
A
N
C
E
The Layers of Plant Responses
to Insect Herbivores
Meredith C. Schuman
and Ian T. Baldwin*
Department of Molecular Ecology, Max Planck Institute for Chemical Ecology, 07745 Jena,
Germany; email: mschuman@ice.mpg.de, baldwin@ice.mpg.de
German Centre for Integrative Biodiversity Research (iDiv), Halle-Jena-Leipzig,
04103 Leipzig, Germany
Annu. Rev. Entomol. 2016. 61:373–94
The Annual Review of Entomology is online at
ento.annualreviews.org
This article’s doi:
10.1146/annurev-ento-010715-023851
Copyright
c
2016 by Annual Reviews.
All rights reserved
Corresponding authors
Keywords
levels of analysis, Tinbergen’s four questions, plant defense theory,
herbivore-associated elicitors, plant-insect interactions
Abstract
Plants collectively produce hundreds of thousands of specialized metabolites
that are not required for growth or development. Each species has a qual-
itatively unique profile, with variation among individuals, growth stages,
and tissues. By the 1950s, entomologists began to recognize the supreme
importance of these metabolites in shaping insect herbivore communities.
Plant defense theories arose to address observed patterns of variation, but
provided few testable hypotheses because they did not distinguish clearly
among proximate and ultimate causes. Molecular plant-insect interaction
research has since revealed the sophistication of plant metabolic, develop-
mental, and signaling networks. This understanding at the molecular level,
rather than theoretical predictions, has driven the development of new hy-
potheses and tools and pushed the field forward. We reflect on the utility
of the functional perspective provided by the optimal defense theory, and
propose a conceptual model of plant defense as a series of layers each at a
different level of analysis, illustrated by advances in the molecular ecology
of plant-insect interactions.
373
Review in Advance first posted online
on December
11, 2015. (Changes may
still occ
ur before final publication
online and in print.)
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
Molecular chemical
ecology: the field
investigating the
evolution of ecological
interactions, and the
functional traits and
molecular and
chemical mechanisms
mediating these
interactions
Levels of analysis:
explanatory levels of
proximate and
ultimate causation
INTRODUCTION
Most plant species are sessile and, during a large portion of their life histories, literally rooted to
the ground. They must either adapt themselves to their environment or adapt their environment
to their needs. It is therefore not surprising that plants are among nature’s most accomplished
chemists, producing a cornucopia of secondary, or specialized, metabolites with which they
manipulate the physiology and behavior of insects and other organisms. In contrast to primary, or
general, metabolites commonly required for cell growth (77), which include most carbohydrates,
amino acids, and fatty acids, specialized metabolites such as alkaloids are by definition unevenly
distributed, both across phylogenetic groups and among individuals within a species (e.g., 9, 32,
40, 44, 48, 63, 75, 87, 113). Of course animals and microbes also produce specialized metabolites,
e.g., defensive glandular exudates or antibiotics; and in the animal kingdom there are many
examples of co-opting, or sequestering specialized metabolites or their precursors, usually from
food and thus often from plants (reviewed in 9, 53). This complex chemical web of life is the
subject of molecular chemical ecology, which originated in studies revealing that phytochemicals
can determine the occurrence and outcome of plant-herbivore interactions (reviewed in 40, 51).
Many early chemical ecologists were zoologists who were interested in the distribution and
evolution of specialized plant metabolites because of the apparent consequences for distribution
and evolution of insect herbivores (32, 40) (Figure 1, 1950s–1960s). The theories that emerged
(Figure 1, 1970s–1990s) posited that plants encounter trade-offs between the costs of producing
these metabolites, in the face of limited resources and coevolving herbivores, and their benefits
via reduced herbivore pressure on plant fitness (9, 123). However, the physiological mechanism
through which a trait occurs does not reveal its function, just as function does not dictate a
mechanism—although each may hint at a reasonable range of possibilities for the other. The
hypotheses derived from these theories were thus frequently neither testable nor mutually exclusive
(8, 9, 123).
In 1988, Sherman (118, p. 616) addressed other, similar instances of “talk[ing] past each other”
in biology, building on earlier work by Tinbergen (130) and Mayr (84) to point out that there are
four levels at which biological phenomena can be understood: (a) Ontogenetic mechanisms and
(b) physiological or cognitive processes determine how a phenomenon happens; while (c) evo-
lutionary trajectories and (d ) ecological pressures determine why it happens that way, or at all
(Figure 1). Biological phenomena can, and ideally should, be understood at all four levels, but
hypotheses may only compete and be tested within a level (118). For example, “increased produc-
tion of specialized compound A entails a reduction in growth” and “the function of compound A
is to defend plants attacked by insect herbivores” are not competing hypotheses. One, both, or
neither may be true, and different types of data must be generated to test each: comparing growth
of plants differing only in their production of compound A in the first case; and comparing fitness
of plants under herbivore attack differing only in their production of compound A in the second
case.
Classical plant defense theories and their insights and shortcomings have been thoroughly
reviewed elsewhere (e.g., 8, 9, 20, 39, 101, 123). Here, we focus on how the utility of these
theories has been limited by their failure to cleanly separate biological levels of analysis; this is
both a universal problem in biology (84, 118) and a topic that has rarely been handled (but see
8). From this point of view, we critically discuss the optimal defense framework and the idea of
growth-defense trade-offs, which are perhaps the most useful of the body of classical plant defense
theories (8, 9, 123). We then provide a synthetic overview of the advances in molecular physiology,
which, in the absence of clear guidance from theory, have driven the field forward (8, 9). Finally,
we offer a perspective on maintaining clear divisions of hypotheses by level of analysis as key
374 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
Coevolution
(Ehrlich & Raven 1964)
Apparency
(Feeney 1976)
Growth:dierentiation balance
(Loomis 1932, 1953; Herms & Mattson 1992)
Spatially varying selection
(Gloss 2013)
Community genetics
(Whitham et al. 2006, Wymore et al. 2011)
Carbon:nutrient balance
(Bryant et al. 1983, Tuomi et al. 1988)
Growth rate/resource availability
(Coley 1985a,b; Coley et al. 1985)
Physiological
mechanisms
Ontogenetic
processes
Evolutionary
origins
Functional
consequences
Why is there natural variation
in plant specialized metabolism?
How are plant specialized metabolites
distributed in nature?
Raison d’être
(Fraenkl 1959)
1950
1960
1970
1980
1990
2000
2010
Optimal defense
(McKey 1974, 1979; Rhoades 1979; Rhoades & Cates 1976)
Figure 1
Several hypotheses from different theoretical frameworks have been put forth over six decades to explain
how and why observed variation in specialized plant metabolites exists. None of these frameworks attempts
to cleanly integrate the levels of analysis at which biological hypotheses can be formulated (8, 118). Because
the frameworks address different levels of how or why (and some blend multiple levels), they generally do
not provide alternative hypotheses but rather different, to an extent complementary, points of view.
discoveries in molecular physiology are extrapolated to the level of ecological interactions, in an
attempt to avoid similar problems of unfalsifiability and talking past each other in the future.
Plant Defense Theory: What Level Are We On?
Figure 1 portrays key concepts and frameworks in classical plant defense theory as well as selected
newer frameworks in terms of how they address, or mix, levels of analysis. In 1959, Fraenkel (40)
argued that plant secondary metabolites evolved primarily under ecological pressure to thwart
herbivores, while herbivores in turn evolved to avoid, detoxify, or sequester these metabolites or
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 375
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
Defense trait: a trait
that increases a plant’s
Darwinian fitness
under stress such as
herbivory
OD: optimal defense
CNB: carbon:nutrient
balance
GR/RA: growth
rate/resource
availability
GDB:
growth:differentiation
balance
Indirect defense:
plant defense trait that
indirectly affects
herbivores by
attracting their
enemies through
provision of resources
or information
even use them to identify suitable host plants. Five years later, Ehrlich & Raven (32) proposed that
coevolution between plant defense and herbivore counteradaptations could explain the diversity
of plants and their herbivores: Plants evolve new defense traits in response to herbivore pressure,
the resulting reduction in pressure permits a radiation in plant diversity, and that radiation creates
new host plants to which herbivores then adapt (adapt-and-radiate coevolution; 123, 129). Fox (39)
later proposed that ephemeral plants interacting with a few, primarily specialized herbivores may
undergo stepwise coevolution in a true evolutionary arms race, while more persistent plant species
may undergo diffuse coevolution, developing a complex suite of defense traits to hedge against
fitness loss to a large, complex, and variable herbivore community. Usually plant defense traits
are taken to refer to specialized plant metabolites, but a robust definition of defense incorporates
other fitness-enhancing responses (discussed below). Coevolutionary hypotheses are difficult to
test because the phylogenetic distribution of traits cannot be predicted from their ecological
functions, or vice versa. However, Fraenkel’s (40) and Erlich & Raven’s theses (32) stimulated
entomologists to ask why, and how, plant defenses are distributed as they are in nature, and how
this distribution has influenced the evolution, behavior, and physiology of insect herbivores.
Perhaps the most successful of these theories has been the optimal defense (OD) theory (85,
86, 102, 103), which has provided testable functional hypotheses. By quantifying plant defense
in the currency of Darwinian fitness, OD provides a link to evolutionary theory (9, 25, 30). This
is not true of carbon:nutrient balance (CNB) theory (15, 133), growth rate/resource availability
(GR/RA) theory (21–23), or growth:differentiation balance (GDB) theory (58, 76, 77), none of
which addresses functional questions (97, 123). The two main hypotheses of OD are (a) defenses
evolve and are allocated to maximize individual inclusive fitness, and (b) defenses have a fitness
cost because they divert resources from other processes, mainly growth and reproduction (102,
123). The first hypothesis is rather the assumption that most traits are adaptive, positing that any
pattern of defense is possible as long as it is adaptive (85, 86, 102, 123). The second hypothesis is
falsifiable (if imprecisely worded; see 34) and produced several subhypotheses. (a) The allocation
hypothesis states that within a plant, tissues are defended in direct proportion to their fitness
value and risk of predation, and in inverse proportion to the cost of allocation. (b) The plasticity
hypothesis states that defenses decrease in the absence of enemies and increase in their presence.
(c) The stress hypothesis states that individuals are less defended when under additional stresses
because they cannot afford to allocate as many resources to defense against herbivores (36, 102,
103, 123).
The hypotheses of OD theory are compatible with apparency theory (36), and the concept of
apparency is implicit in the allocation hypothesis (risk of predation) and the plasticity hypothesis
(presence of enemies) (103, 123). Apparency theory (36) predicts that defenses evolve in direct
proportion to predation risk and inverse proportion to fitness cost. It has largely been criticized
as impractical: Although Feeney (36) defined apparency as the quality of being predictable in time
and space, to quantify the apparency of a host plant or tissue, one must understand the prediction
mechanism used by its herbivores, and thus apparency may be a quality of an interaction and
not of a plant per se (39, 123). However, apparency theory has been remarkably useful to our
understanding of the widespread phenomenon of indirect defenses, which function by increasing
the apparency of herbivores to their enemies (68). The apparency and OD theories have together
provided a handful of testable hypotheses posed at a functional level of analysis. Thus, among the
classical plant defense theories, these likely have the most empirical support (reviewed in 89, 123).
Nevertheless, the trade-off concept invoked by both OD and apparency theories may encourage
researchers to confuse physiological mechanism with functional roles.
As an illustration, evidence for the stress hypothesis of OD is mixed: Plants subjected to non-
herbivory stresses such as drought are predicted to be less well defended, but often are not, and
376 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
SVS: spatially varying
selection
sometimes appear to be better defended (52). Admittedly, this literature is obfuscated by the chal-
lenge of stressing plants without stressing herbivores, as well as imprecise definitions of stress
(73). Nevertheless, the mixed evidence is consistent with observations indicating that plant de-
fense traits may have multiple functions, and that the expression of these traits, as well as their
fitness consequences, depends on multiple ecological factors in addition to herbivore pressure (28,
36, 39). Patterns of plant defense are likely better understood as positions on a changing fitness
landscape, rather than as solutions to an optimization problem (87, 98).
Similarly, observed plant phenotypes may not be best explained mechanistically by a trade-off
between defense traits and growth and development. Plants excel at coordinating their metabolism
interactively with their biotic and abiotic environments, both responding to their environment
and changing it, to avoid situations in which resource availability would constrain their options
(see, e.g., 28, 48, 74, 87, 98, 111, 145). The extent to which any defense trait limits growth and
development depends on the interaction of multiple physiologically and ontogenetically regulated
processes, each of which individually contributes to, but does not solely determine, the observed
growth and defense phenotype. Each of these mechanisms involves some clearly defined trade-off
at the level of, e.g., cell expansion or proliferation versus differentiation, allocation of nutrients,
substrate channeling, or protein stability; but they may interact to buffer potential trade-offs at the
ecological or functional level (74, 76, 77). In the end, the fitness costs and benefits of any trait, and
whether it supports or competes with other traits’ contribution to plant fitness, depend on current
(ecological) and historical (evolutionary) pressures (Figure 2) (6, 88). Despite other failings, GDB
theory recognizes that physiological trade-offs at a cellular level do not equate to trade-offs at an
ecological level (74): It predicts, for example, that high levels of nutrients and light will result in
high rates of both growth and defense (123).
Functional population genetics approaches have now been added to the defense theory
repertoire (Figure 1, 2000 onward). Spatially varying selection (SVS) theory (44) draws a direct
connection between the environment-dependent Darwinian fitness consequences of genetically
controlled defense traits and traits’ predicted frequency in a population, but it does not consider
ecological functions or physiological mechanisms of these traits. Similarly, community genetics
theory (145) proposes equations derived from population genetics that can be used to scale up
from the function of a genetic trait in an individual to its effects in ecological communities. But
although it considers ecological mechanisms cleanly divided into evolutionary and functional
levels, this approach neglects physiological mechanisms of traits. Thus, neither approach provides
a full biological picture, but each offers insights into how ecological contexts might be explicitly
integrated into updated plant defense theory.
This is important because the role of ecological context in determining phenotypes is poorly
understood, most obviously in the case of biological invasions, in which a species out of context
dominates its new community. For over two decades, ecologists have sought with little success
phenotypic markers of invasiveness (136). Feeney (36) suggested that weediness of invasive plants
might be connected to their unapparency in a novel environment, and in 1995, Blossey & Notzold
(10) proposed that invasive species evolve increased competitive ability (EICA) when resources
are not allocated to defense. Data have not supported this hypothesis, which has misled its field
for twenty years (37).
Where Do Trade-Offs Occur, and How Do They Relate to Function?
The simple trade-off between growth and defense invoked in classical plant defense theories
such as OD has been supplanted by an increasingly detailed picture portraying a sophisticated
reconfiguration of metabolism in attacked plants, revealed by molecular chemical studies (see,
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 377
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
a Cellular physiology
b Regulatory feedback loops c Metabolic reconguration
PS II
H
2
O CO
2
(CHOH)
n
O
2
ETC
PS I
NADPH
ATP
ADP
NADP
+
G3P
RuBP 3PG
Sucrose
(export)
Amino acids,
fatty acids
Starch
(storage)
Chloroplast
Abiotic stress
Biotic stress
Mutations
Development
Unfolded proteins
UPR genes
Ub
ATP depletion
Vigorous protein synthesis
Aberration of Ca
2+
or redox homeostasis
Inhibition of protein modication
Inhibition of protein transfer to Golgi apparatus
ERAD
stimulate
promote
restore
prevent
relieve
Nucleus
ER
Interlocked
feedback loops
Diurnal and
other
external cues
Metabolite
accumulation
ZT 0 12 24 36 48 0 12 24 36 48 0 12 24 36 48
1° metabolites Intermediates 2° metabolites
Herbivores
Herbivores
Mycorrhizae
Systemic
HIPVs
HIPVs
Plant pathogens
Competitor
or parasitic plant
Entomopathogenic
nematode
Carnivores
Pollinators
Endophytes
26s
d Ecological phenotypes
378 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
e.g., 8, 9, 12, 54, 94, 110, 122, 125, 147). The individual physiological and ontogenetic trade-offs
that contribute to this complex network differ by how much they influence fitness trade-offs at the
ecological level. A few master switches regulate many components, and in many cases, molecular
interaction studies have already implicated these key players (e.g., 16, 17, 33, 38, 50, 96, 104, 116,
117, 128, 132, 150). However, even when measuring effects of master regulators, the magnitude
and even direction of a growth-defense trade-off are likely to depend strongly on environmental
variables. Furthermore, traits that attract enemies of herbivores, transport and storage mechanisms
that make nutrients inaccessible, and even phenological shifts that confer herbivore avoidance are
part of the network regulated by herbivory, and must be considered as components of defense,
alongside toxic, repellent, and antinutritive secondary metabolites (8, 36, 64, 69, 112, 114, 115).
Defense traits are any traits that defend plant fitness in the face of stress, and the more ecologically
realistic the environment, the more robust the demonstration of defensive function (64, 74). As
has been discussed elsewhere (e.g., 9, 20), it is often difficult to establish ecologically realistic assays
and to measure fitness consequences, but challenges encountered when measuring the right thing
cannot be remedied by measuring the wrong thing. We would argue that suitable compromises
can be found, so long as these are also clearly defined and their shortcomings acknowledged (e.g.,
74, 112).
In fact, we still know astonishingly little about what plants must do in order to grow, defend,
reproduce, and keep fit in their native ecologies (Figure 2). Although plants are autotrophic
organisms, harnessing energy from sunlight and water, fixing carbon from the air, and taking up
mineral nutrients from the soil, they also must operate as heterotrophs, metabolizing their fixed
carbon and nutrients to form a plethora of other metabolites in a system of precisely coordinated
metabolic networks (126). The circadian clock plays a crucial role in regulating these processes,
for example, preparing the photosynthetic machinery just before daylight and allowing plants to
pace the metabolism of starch during the night (31).
The next section provides an updated overview on the molecular mechanisms of plant de-
fense responses that points out where trade-offs take place, and describes how a multitude of
physiological and ontogenetic events are integrated into observed phenotypes, i.e., to incorporate
mechanistic, functional, and evolutionary levels of analysis into an integrated understanding with-
out confounding these levels. We continue to take a phytocentric view (8) because this reflects our
expertise and because we believe that entomologists benefit from an accurate understanding of
plants just as plant scientists benefit from understanding insect herbivores (32, 40). Furthermore,
we assume that plants and their herbivores are strategic players in a set of games that play out in
ecological time, with consequences for evolutionary trajectories (87, 98). The view we present dif-
fers from descriptions of plant defense syndromes (2), which acknowledge the interdependency of
defense- and growth-related traits but have not integrated physiological and ecological plasticity.
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
Figure 2
An overview of the physiological constraints plants face as photosynthetic organisms when responding to attack from, and manipulating
the behavior of heterotrophs: coordination of the light and dark reactions of photosynthesis with (a) subcellular physiology of general
and specialized metabolism via the (b) circadian clock, and the resulting (c) chemical and (d ) ecological phenotypic parameters that we
can measure. Arrows in panel c indicate the timing of an herbivore induction event either at midday (left arrow)ormidnight
(right arrow). Abbreviations: ER, endoplasmic reticulum; ERAD, ER-associated degradation; ETC, electron transport chain; HIPVs,
herbivore-induced plant volatiles; PSI/II, photosystem I/II; Ub, ubiquitin; UPR, unfolded protein response; ZT, Zeitgeber time (e.g.,
time since dawn on the first day of the experiment). Images in panel a are modified from References 92 (left) and 151 (right), image in
panel b is modified from Reference 50, and images in panel d are modified from Reference 28. Conceptual graphs in panel c arose from
discussions with D. Li (75).
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 379
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
HAE:
herbivore-associated
elicitor
FAC: fatty
acid–amino acid
conjugate
THE LAYERS OF PLANT DEFENSE RESPONSES TO HERBIVORES
Plant defense responses to herbivores can be understood at any of several layers, from genetic
structure through ecological influence (Table 1). These layers are separate but not independent:
Each emerges from layers below it, and receives feedback from layers above it, but each layer
represents a different level of analysis (118). Genetic programming, ontogeny, and environmental
factors have all been considered, if not cleanly separated, in plant defense theory. We focus on
the molecular physiology of plant responses to insect herbivores, an area in which advances in
knowledge have led to an updated experimental and theoretical understanding of plant defense
against herbivores. This layer can be divided into (a) perception of herbivores, (b) early signaling
responses, (c) hormone-mediated signaling, (d ) reconfiguration of metabolism, and (e) changes to
phenology. The transcriptional layer of regulation that connects hormonal signaling to phenotypic
changes has been reviewed (54) and so we do not repeat it here.
Perception of Insect Herbivores
Specificity in response to herbivores begins with plants’ perception of damage patterns and
herbivore-associated elicitors (HAEs) in amounts as little as femtomolar concentrations (see side-
bar, PAMP Envy) (reviewed in 12, 33, 60, 93, 147). Plants respond to characteristic spatiotemporal
patterns of damage (60, 95) and furthermore can perceive damage to the leaf surface and trichomes
when herbivores land or walk (reviewed in 33, 59). Plants seem to count the number of herbivore-
associated damage events, and the dynamics of the signaling response to each individual event
determine the characteristics of the response to repeated events (reviewed in 41). However, the
HAEs found in secretions and fluids from herbivores are essential for many components of plant
responses (reviewed in 12, 59, 60, 93, 147).
The first HAEs were discovered in the 1980s and 1990s (reviewed in 12). HAEs may be digested
plant molecules such as inceptins, proteolytic products of the plant chloroplastic ATP synthase γ-
subunit found in Spodoptera frugiperda (fall armyworm) oral secretions (108); or herbivore-derived
molecules thought to be essential for metabolism, such as fatty acid–amino acid conjugates (FACs),
which are probably required for nitrogen metabolism in Lepidoptera (151). The first fully charac-
terized HAE was the hydroxyl FAC volicitin [N-(17-hydroxylinolenoyl)-
L-glutamine], identified
as the plant-volatile-inducing factor in Spodoptera exigua (beet armyworm) oral secretions (3). Data
support the existence of a membrane-bound volicitin receptor that is upregulated by herbivory
(131), but the protein has not been identified; however, a lectin receptor kinase gene (LecRK1)
that is rapidly, transcriptionally upregulated in response to the FAC N-linolenoyl-glutamate from
Manduca sexta (tobacco hornworm) was identified in Nicotiana attenuata (coyote tobacco) (42). Oral
secretions from eight different lepidopteran species furthermore formed ion-permeable channels
in artificial membranes, although FACs alone were unable to form stable channels (82).
Plants may further metabolize HAEs as in the case of N-linolenoyl-glutamate from M. sexta,
which is rapidly oxidized by a lipoxygenase upon application to N. attenuata leaves (139). This may
be one mechanism endowing specificity of plant responses to herbivores. Plant species respond
differently to the same HAEs as measured, e.g., by phytohormone production (109), or they may
not respond at all to specific HAEs (121). Within plant species, the magnitude and quality of
response to the same HAEs vary: In N. attenuata, a genotype native to Arizona had a reduced
signaling and metabolic response to M. sexta oral secretions compared with a genotype from Utah
(149), and neighboring plants from single populations in Utah displayed differences of a similar
magnitude (63, 113).
380 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
Table 1 Layers of plant defense against herbivores, corresponding levels of analysis, and examples of observed phenomena
and testable hypotheses. Levels of analysis are from Sherman (118)
Layer Level of analysis Phenomena Example hypotheses
Ontogenetic changes Ontogeny (proximate) Changes in tissue-specific
hormonal signaling over life
stages
Flowering shuts down herbivore-elicited
ethylene signaling.
Tissue- and life-stage-specific
accumulation of metabolites
Herbivore-induced vegetative volatile
emission is downregulated when flowers
open.
Stage-specific growth and
differentiation patterns
Sensitivity to herbivore induction as
measured by production of hormones,
transcripts, or metabolites differs with
growth stage.
Changes in phenology Senescence processes are induced by
herbivory.
Flower opening time is changed by
herbivory.
Physiological changes Physiology (proximate) Perception of herbivory Transmembrane receptor proteins detect
FACs.
Early signaling events The waveform of [Ca
2+
]
cyt
fluctuations
determines the identity of the protein
messenger activated by Ca
2+
.
Phytohormone signaling JAs other than JA-Ile also bind specific
SCF
COI1
-InsP
5
-JAZ complexes to
release inhibition of JA-upregulated
genes.
Reconfiguration of metabolism Systemically transported specialized
metabolites accumulate in proportion to
tissue sink strength.
Changes in growth rate Different herbivore elicitors result in
different magnitudes of reduced growth
rates within a plant genotype.
Genetic and epigenetic
architecture
Evolution (ultimate) Inter- and intrapopulation
variation in defense gene alleles
The frequency of a defensive gene in a
plant population is proportional to risk
of herbivory by a susceptible herbivore.
Epigenetic signatures of
herbivory
Herbivory in the current generation of
plants causes a predictable epigenetic
signature in the next generation.
Ecological interactions
and realized phenotype
Function (ultimate) Defensive functions of specialized
metabolites
WT accumulation patterns of specialized
metabolites increase plant fitness under
herbivory.
Resistance function of specialized
metabolites
WT accumulation patterns of specialized
metabolites reduce herbivore damage.
Fitness costs of producing
specialized metabolites
WT accumulation patterns of specialized
metabolites reduce plant fitness in the
absence of herbivore attack.
Abbreviations: FAC, fatty acid–amino acid conjugate; JA, jasmonate; WT, wild type.
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 381
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
PAMP ENVY
The perception mechanisms for characteristic microbe- or pathogen-associated molecular patterns (MAMPs/
PAMPs) and effectors from pathogenic microbes is well understood, having produced the gene-for-gene and guard
hypotheses for which there is extensive experimental support (62). This has produced a phenomenon that could
be called PAMP envy in which every review of HAEs must begin with a summary of how similar HAEs are to
MAMPS, PAMPS, and effectors, yet how relatively little is yet known about HAE perception. This sorry state
of affairs is a consequence of where the two fields started: The chemical ecology of plant-insect interactions was
developed by zoologists, phytochemists, and plant physiologists, and plant pathology investigations were begun by
molecular biologists and geneticists. Plant-insect interaction research generally has a better understanding of the
mechanisms, in particular the specific chemistries by which plants and sometimes ecological communities shape
interactions, whereas plant pathologists have a more epidemiological understanding: Some genotypes are more
resistant than others, but the organism-level mechanism and context dependence of resistance are often not known.
An exception is the field of ecological immunology, which attempts to synthesize the cellular, immunological level
with the community-based, ecological level of an organism’s responses to pathogens (83).
MAPK:
mitogen-activated
protein kinase
ROS: reactive oxygen
species
JA: jasmonate
Early Signaling Response
Upon feeding damage or wounding and application of regurgitant, a membrane potential (V
m
)
depolarization at the feeding site spreads isotropically through the leaf in minutes, as first charac-
terizedindetailforPhaseolus lunatus (Lima bean) damaged by Spodoptera littoralis (Egyptian cotton
leafworm) (81). This is rapidly followed by influx of Ca
2+
at the feeding site, which could also be
stimulated in P. lunatus by wounding and application of volicitin or the detergent sodium dodecyl
sulfate (SDS) (81). Within minutes of feeding damage, elevated mitogen-activated protein kinase
(MAPK) activity can be detected proximate to the feeding site and in remote locations within
the damaged leaf, although the distribution of MAPK activity in N. attenuata was constrained by
vasculature and vascular flow (148). MAPK activation occurs both up- and downstream of calcium
signaling (reviewed in 147), and calcium signaling appears to occur both up- and downstream of
reactive oxygen species (ROS) signaling, which is also associated with herbivory but not with me-
chanical damage; as in pathogen attack, induced ROS production is likely controlled by NADPH
oxidases (reviewed in 12, 147).
ROS production, calcium signaling, and MAPK signaling are essential components of responses
to a plethora of biotic and abiotic stresses and cues. Thus, the integration of other environmental
or stress-related signals likely occurs during the early signaling response to herbivores, but the
mechanism of signal integration at this stage is not well understood (reviewed in 12, 147). MAPKs
initiate signaling cascades, resulting in activation of transcription factors and phytohormone-
mediated signaling (reviewed in 147), and signal integration via phytohormone interactions has
been more closely investigated. Interestingly, herbivory-induced MAPK activity patterns corre-
spond to patterns of jasmonate ( JA) stress hormone accumulation in N. attenuata leaves (127, 148).
Hormonal Signaling Response
Most herbivore resistance traits are induced by JAs (reviewed in 60, 147), and JAs were shown to
negatively regulate tolerance of herbivory in N. attenuata (80). JAs are synthesized in plastids and
peroxisomes from linolenic acid cleaved from plastidial membranes by glycerolipase (GLA) and
peroxidated by lipoxygenase (13). cis-(+)-12-Oxo-phytodienoic acid (OPDA), which is synthe-
sized in the plastid, may be the first active hormone in the pathway and is stored, for example, as
382 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
arabidopsides for rapid release upon wounding in Arabidopsis thaliana (mouse-ear cress) (reviewed
in 26). The 18-carbon OPDA is transported to the peroxisome, where it is converted to the
12-carbon (+)-7-iso-jasmonic acid, which may be methylated, hydroxylated, or conjugated to
amino acids or sugars to form active and inactive JAs that may be modified further (reviewed in
60, 143, 144).
The conjugate (+)-7-iso-jasmonoyl-
L-isoleucine ( JA-Ile) is thought to be the primary active
form (38); however, JA signaling, but not JA-Ile, is required in Solanum nigrum (black nightshade)
for resistance against herbivores (138); and JA-Ile is not required for JA-induced volatile emission
in A. thaliana (137). JA-Ile is perceived by the Skp/Cullin/F-box SCF
COI1
E3 ubiquitin ligase
complex (65, 150) interacting with inositol pentakisphosphate (InsP
5
) (117) and one or more
jasmonate ZIM-domain ( JAZ) proteins (17, 128): Different SCF
COI1
-InsP
5
-JAZ complexes may
have altered binding specificities (19, 117). Fifteen JAZ/TIFY proteins in A. thaliana have been
characterized to date (reviewed in 66), and similar numbers were reported in other plant species
(99). The JAZ/TIFY proteins can form complexes with each other and with the JA-regulated
transcription factor MYC2 as well as with proteins involved in other hormone signaling pathways
(reviewed in 67, 104). In a mechanism similar to auxin and gibberellin signaling, when levels of
active JAs are low, JAZ proteins bind to and repress JA-regulated transcription factors; when JA
levels are high, the activated SCF
COI1
-InsP
5
-JAZ complex labels JAZ for degradation, activating
gene transcription (reviewed in 104). Alternative splicing of JAZ may tune sensitivity by producing
variants with reduced affinity for the SCF
COI1
receptor complex (18).
JAs interact extensively with other plant hormones and function in growth and fertility as well
as defense (reviewed in 143, 144, 147). The interaction of JAs with other hormones permits in-
tegration of multiple stress and developmental signals (reviewed in 33, 104). Ethylene modifies
JA-mediated responses, and auxins and the pathogen defense hormone salicylic acid suppress the
action of JAs, whereas interactions between JAs and gibberellins (57), cytokinins (106, 107), ab-
scisic acid, and nitric oxide are less well understood (reviewed in 33, 104). Herbivores, similar to
pathogens, can manipulate these hormone interactions to their own advantage, often by inducing
salicylic acid accumulation to repress JA-mediated responses (reviewed in 104). Some herbivores
and pathogens manipulate cytokinins to change the sink strength and nutritional quality of dam-
aged tissues: Agrobacterium tumefaciens and galling herbivores use cytokinins to create local sinks
in plants, and leafminers form green islands in senescent leaves (reviewed in 33, 43); and more
mobile herbivores such as M. sexta and Tupiocoris notatus (tobacco suckfly) may also manipulate
cytokinin signaling (105, 107). Cytokinins from herbivores are often produced by symbiotic mi-
crobes (reviewed in 43).
Systemic signaling via the vascular system induces defenseproduction in remote plant parts after
attack, similar to the phenomenon of induced systemic resistance after pathogen attack (reviewed
in 35, 60, 147). Classical grafting experiments in Solanum lycopersicum (tomato) indicated that JA
biosynthesis was necessary to produce a long-distance wounding signal, while JA perception was
required for its perception in distal leaves, and a JA-derived phloem-mobile signal was proposed
(reviewed in 60). Within minutes of wounding, electrical signals spread to systemic tissue via
vascular connections and glutamate receptor-like (GLR) ion channels (35, 96, 147). In A. thaliana,
a lipoxygenase expressed in cells adjacent to xylem is responsible for synthesizing JAs in systemic
tissue (16, 35). It is now thought that systemic JA biosynthesis in A. thaliana is triggered similarly
to JA biosynthesis in distal regions of a wounded leaf: by V
m
depolarization and/or Ca
2+
flux
transmitted by GLRs and resulting downstream signaling events; however, systemic (between-
leaf) signaling is suggested to also involve temporary changes in xylem pressure (35, 81).
The take-home message of the advances in systemic signaling research is that when an herbivore
attacks a plant, the entire plant recognizes and responds to this attack at some level, even if
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 383
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
Specialized
metabolism:
metabolic pathways
and products that are
not strictly required
for growth and
production of new
cells; also called
secondary metabolism
Direct defense: plant
defense trait that
directly affects the
ability of an herbivore
to assimilate plant
tissue, e.g., via a
repellent, toxic, or
antidigestive
mechanism
some responses are highly localized to the attacked tissues. In some plants, priming of systemic
defense responses by perception of herbivore-induced volatiles may be a mechanism to avoid
vascular constraints on systemic perception of damage (55). The arena of the induced responses is
potentially extended to neighboring plants sufficiently close and sensitive to emitted volatiles or
that share mycorrhizal connections (7, 55).
Reconfiguration of Metabolism
Herbivory induces extensive changes to both general and specialized metabolism via a transcrip-
tional reconfiguration specific to each herbivore (reviewed in 54, 114). This reconfiguration can
be interpreted as a downregulation of growth and an upregulation of defense in the trade-off
framework. Alternatively, it can be understood as the coordinated reconfiguration of metabolism
to prevent the allocation of energy and nutrients to herbivore fitness, rather than to plant fit-
ness (Figure 2, Table 1). These interpretations are not incompatible, but the second suggests
specific hypotheses that are closely connected to known mechanisms of plant physiology and
plant-herbivore interactions, and is therefore more useful.
One option for plants to avoid converting their fitness to herbivore fitness is of course to avoid
attack, and one way to do this is to “smell bad.” Plant volatiles comprise thousands of structures,
and volatile blends are specific to particular stress events and even configurations of herbivores
and their parasitization status (28, 100, 119). Volatiles may increase the foraging efficiency of
predators and parasitoids of herbivores or prevent herbivores from ovipositing, but herbivores
may use the same compounds as cues to locate host plants (reviewed in 28). However, in different
ratios, these compounds become more attractive to predators and deterrent to herbivores (4, 5). A
volatile blend may comprise the emission from multiple plants, meaning that volatile emission of
close neighbors can determine a plant’s attractiveness to herbivores and predators (111). There is
evidence that plants can detect neighbors’ volatiles (reviewed in 28, 55), and one yet unexplored
function of plants’ sense of smell may be that they adjust their own volatile emission to what the
neighbor is producing. Plant volatiles are a double-edged sword.
Once a plant is attacked, the goal of retaining energy and nutrients for plant fitness can be
achieved in two ways: by channeling metabolism toward synthesis of secondary metabolites, in-
cluding direct and indirect defenses, ideally selected for effectiveness against the attacking her-
bivore; and by enacting tolerance mechanisms to mobilize nutrients away from the attack site
and store them in inaccessible forms for later remobilization. The ever-expanding list of known
specialized plant metabolites has been reviewed by many others (e.g., 8, 36, 40, 78, 86, 94) and, in
addition to more classical examples of plant volatiles, toxins, and digestibility reducers, includes
compounds closer to general metabolism such as diterpene glycosides (56, 120) and antinutritive
amino acids (1). The discovery that JA signaling dramatically regulates leaf sugar levels (79), in
addition to regulating specialized metabolites, underscores how everything in metabolism can be
recruited to enhance defense.
Herbivore assimilation rates depend on the composition of nutritive and antinutritive metabo-
lites rather than on the absolute amounts of either: This key insight explained differences in
mechanism as well as consequences for digestibility-reducing defenses versus toxic direct defenses
proposed by OD and apparency theories (36, 39), and it explains why defense theories such as
CNB, RA/GR, and GDB, which at most predict total amounts of secondary metabolites, are not
useful. Herbivore consumption rates depend on palatability, which may reflect tissue nutritional
quality, but are more likely determined by specific metabolites acting as phagostimulants that may
also be important for host location (14, 40, 141), such as the six-carbon green leaf volatiles (46,
384 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
Specialist herbivore:
herbivore that feeds on
a restricted set of
plants
47, 90). When it comes to understanding the chemical underpinnings of defense, the devil lies in
the chemical details: Composition matters more than total amounts.
The composition of metabolites in tissues can be altered not only by induction of specialized
metabolite production and the reallocation of standing pools of antinutritive metabolites, but also
by the reallocation of nutritive metabolites through senescence or other processes (discussed in
23, 88, 89, 105). Changes induced by the shoot herbivore M. sexta in the expression of the SnRK1
subunit GAL83 increased carbon partitioning to N. attenuata roots, reducing loss of energy to
herbivory and permitting later remobilization of more carbon from roots for seed production
(115). Interestingly, members of the SnRK1 family are involved in regulating plant developmental
transitions based on carbon availability and thus may provide a link between growth and defense
(132). The ability of plants to reallocate nutritive metabolites will depend on the sink strength
of the attacked tissue, which may explain why some (perhaps many) herbivores and pathogens
manipulate cytokinin signaling (43, 107).
Particular feeding patterns may also allow herbivores to avoid induced responses in their host
plants, akin to the laticifer-cutting behavior employed by herbivores feeding on latex-producing
plants (2). By contrast, plant metabolites may manipulate herbivore behavior and susceptibility to
attack. For example, ingestion of trypsin protease inhibitors has a small effect on the growth of the
specialist herbivore M. sexta, but a large effect on its ability to defend against predator attack (112),
thus potentially rendering larvae more susceptible to plant indirect defense, although perhaps also
rendering them less nutritious prey. Toxic metabolites such as the eponymous alkaloid nicotine
in N. attenuata may determine the potential predator community, selecting for predators that can
avoid ingesting, or that can detoxify, nicotine (72). For the generalist S. exigua, consumption of
N. attenuata tissue is limited by its capacity to detoxify nicotine, which is much less than the capacity
of the nicotine-adapted M. sexta;thisrendersS. exigua more susceptible to digestibility-reducing
protease inhibitors because it cannot compensate by consuming more tissue (124).
The plant circadian clock has been proposed to synchronize the expression of defense traits
with herbivore activity (45, 61). However, a key insight from plant defense theory—and one that
has proven robust—is that herbivore populations vary with plant chemistry and life history, and
any individual plant species is likely to face attack from an herbivore community that varies in time,
composition, or both (36, 39). Just as herbivores differ in their ability to cope with plant defense
traits, they also differ in the timing of their activity (32, 61). The internal clock tunes sensitivity to
diurnally predictable stress factors, but herbivory often does not fit into that category, in contrast
to events such as drought stress and pathogen attack, which strongly depend on diurnal changes
in abiotic factors such as humidity (61, 116, 142). Thus, the role of the clock in defense against
herbivory is likely to be more complex than simple rhythmic control of secondary metabolism. For
example, clock regulation of metabolic flux may result in differences in the temporal accumulation
of defense metabolites when plants are attacked at different times of day (Figure 2b) (75). Stress
factors may even disrupt the clock (116, 142).
Phenology and Timing
OD theory predicts that changes in phenology to reduce apparency could be an aspect of plant
defense responses; however, it envisioned these changes over evolutionary and not ecological
time (36, 102, 103). More recent studies have shown that growth rate, allocation of resources,
flowering time, and even flower phenology are altered in ecological time in response to herbivory
(69, 115, 116; and see references in 48, 123). For N. attenuata, the specialist herbivore M. sexta is
also an important pollinator. When attacked by M. sexta larvae, the plant changes flowering time
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 385
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
from nighttime to morning, thereby switching to likely less effective hummingbird pollination
but avoiding further oviposition (69). Leaf removal and elicitation with methyl jasmonate can
alter flowering time in competing plants (134); perhaps plants genetically programmed with an
earlier flowering time, though likely to be smaller and set less seed, are more likely to successfully
distribute seed in seasons with high herbivory pressure, thus stabilizing variation in flowering time
in populations.
In A. thaliana, natural variation in the sequence of pseudo response regulator (PRR) genes
that link external inputs to the internal clock can enhance fitness (91). Variation in the expression
of the light-signaling gene gigantea (GI) under long-day conditions is widespread and has been
proposed to cause conditional variation in the temporal dynamics of diurnally regulated gene
expression, which may permit temporal variation in phenotypes while avoiding pleiotropic effects
from disrupting the clock (27). Such differences in timing could be mimicked using transgenic
plants expressing clock genes under the control of an inducible promoter (105). Sensitivity of
plants to herbivores and production of defense metabolites may furthermore fluctuate throughout
the day as a result of regulation by the plant circadian clock (reviewed in 116, 146) and also depends
strongly on plant developmental stage (11, 29, 49, 71, 135), the progression of which is determined
by the clock interacting with environmental cues (24).
CONCLUSIONS AND FUTURE DIRECTIONS
Plants tailor their defense responses to different herbivores on the basis of HAEs and patterns of
damage from, e.g., feeding or oviposition (reviewed in 12, 33, 60, 93). Different plant species re-
spond differently to the same HAEs (109). Within a species, the magnitude and quality of response
to an herbivore, as measured by activation of signal transduction components and accumulation
of stress hormones, may vary greatly (113, 149), as may the resulting profile of defense metabo-
lites (70; reviewed in 28, 147). What causes this variation? This question can be answered at four
different levels of analysis (Figure 1).
Ontogenetic changes cause large changes in the hormonal and metabolic response to her-
bivory. For example, new buds on an oak (Quercus robur) will not produce tannins, although the
surrounding leaves do (36), and although M. sexta feeding on N. attenuata produces a burst of
ethylene on rosette-stage plants, this burst disappears with the first flower, along with some of the
plant’s inducible defenses, representing a switch from inducible to constitutive metabolite deploy-
ment with the transition to reproductive growth (29)—at least for plants in pots. Mechanistically,
different individuals and species have an array of stages in their signaling response to an herbivore
that may be tuned toward higher or lower sensitivity, for example, by mutations in cis-regulatory
DNA elements that control the transcription of regulatory genes, or by mutations in the alterna-
tive splicing of JAZ genes and other regulators. At the evolutionary level, balancing selection for
two or more alleles of such mutated genes may exist owing to unpredictable or cycling herbivore
populations, for example, or plant populations may undergo purifying selection for one allele or
another (44). Finally, at the functional level, the effect of any one defense trait on the individual’s
inclusive fitness will be determined by a changing and interactive Darwinian fitness landscape,
meaning that plasticity and honed powers of perception are likely very important, but, as assumed
by OD, otherwise anything goes, as long as it is adaptive (98, 102).
Often the evolutionary success of a genotype is determined by its relative advantages compared
with those of its neighbors: Can it grow faster, produce more seed, and defend better (87)? How-
ever, there are important examples, such as the survival of seeds in the seed bank, in which fitness
differences approach categorical distinctions rather than small relative advantages: Is the seed resis-
tant to microbial attack or not? The layered view of plant defense permits separate consideration of
386 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
physiological, ontogenetic, ecological, and evolutionary phenomena while recognizing that each
layer contains emergent properties from the one(s) below it, and that both diurnal timing and on-
togenetic processes are important modifiers of plant responses to the environment. The resulting
integrated concept relies on the definition of defense as a set of ecological context-dependent
responses that increase plant fitness in the face of herbivore attack. We note that this definition of
plant defense is compatible with the concept from ecological immunology of adaptive responses
to pathogens (83), and consistent with the attempt to scale up from genes to ecosystems embodied
in the community genetics approach (145).
SUMMARY POINTS
1. Plant defense theory emerged from efforts to understand the distribution of specialized
plant metabolites in nature, and the role of these metabolites in structuring ecological
communities. However, progress was stymied because the theories frequently posited
hypotheses posed at different levels of analysis, which made them impossible to rigorously
test.
2. Plant defense phenotypes emerge from the coordination of subcellular and whole-
organism physiological processes, and the trade-offs involved are more complex than
balancing an account with photosynthetic carbon fixation and nutrient uptake on the
one side, and growth and secondary metabolite production on the other.
3. Plant defenses are configured in response to particular attackers via recognition of specific
elicitors and patterns of damage, leading to tailored reconfiguration of metabolism that
extends to the entire plant for some responses.
4. Specialized plant metabolites are an essential component of plant defense responses,
which also include physical barriers, tolerance mechanisms, repellent traits, spatiotem-
poral escape, and any other traits that increase fitness under herbivore attack.
5. Plant defense responses and their fitness outcomes are shaped by interactions of plants
with their ecological communities and their abiotic environment, likely in that order.
FUTURE ISSUES
1. What are the receptors of HAEs? How can they be identified? Can we learn from
plant-pathogen interaction research methods, which have identified multiple receptors
of MAMPS and PAMPS and targets of effectors?
2. To what extent is there specificity in the early signaling response to different attackers?
How is it shaped by the perception of other stress factors? How do early signaling
responses produce tailored phytohormone signaling?
3. How is specificity encoded at the level of phytohormone signaling? What is the signifi-
cance of the several JAZ/TIFY proteins and their interactions with each other and other
proteins? What role do JAs other than JA-Ile play? How are multiple hormone signaling
pathways integrated? How are they manipulated by herbivores?
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 387
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
4. What are the roles of general and specialized metabolism in determining nutritional
quality and palatability for more or less specialized insect herbivores? How do direct and
indirect defense strategies interact? How is this metabolic reconfiguration constrained
by ontogeny? What are the cellular constraints on launching induced defenses?
5. How do the phenotypes of neighbors alter a plant’s own realized phenotype and fitness
landscape? What are the consequences for biodiversity and ecosystem function?
6. How can we use this knowledge to address pressing practical concerns such as biological
invasions and agricultural sustainability?
7. Can we avoid hypothesis conflation, unfalsifiability, and “talking past each other” in the
future by applying Sherman’s levels of analysis and conceptual frameworks based on these
levels, such as the concept of plant defense as a series of layers?
DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.
ACKNOWLEDGMENTS
The authors thank C. Diezel for assistance with panels a and b of Figure 2; N. Whiteman for dis-
cussion about functional population genetics and plant defense theory; A. Steppke, M. Berenbaum,
and three anonymous reviewers for comments that improved the concept, focus, and precision of
the manuscript; an ERC Advanced Grant (no. 293926) to I.T.B; the German Centre for Integra-
tive Biodiversity Research (iDiv) Halle-Jena-Leipzig funded by the German Research Foundation
(FZT 118); and the Max Planck Society for funding.
LITERATURE CITED
1. Adio AM, Casteel CL, De Vos M, Kim JH, Joshi V, et al. 2011. Biosynthesis and defensive function of
Nδ-acetylornithine, a jasmonate-induced Arabidopsis metabolite. Plant Cell 23:3303–18
2. Agrawal AA, Fishbein M. 2006. Plant defense syndromes. Ecology 87:132–49
3. Alborn HT. 1997. An elicitor of plant volatiles from beet armyworm oral secretion. Science 276:945–49
4. Allmann S, Baldwin IT. 2010. Insects betray themselves in nature to predators by rapid isomerization of
green leaf volatiles. Science 329:1075–78
5. Allmann S, Sp
¨
athe A, Bisch-Knaden S, Kallenbach M, Reinecke A, et al. 2013. Feeding-induced re-
arrangement of green leaf volatiles reduces moth oviposition. eLife 2:e00421
6. Augner M, Fagerstroem T, Tuomi J. 1991. Competition, defense and games between plants. Behav. Ecol.
Sociobiol. 29:231–34
7. Babikova Z, Gilbert L, Bruce TJA, Birkett M, Caulfield JC, et al. 2013. Underground signals carried
through common mycelial networks warn neighbouring plants of aphid attack. Ecol. Lett. 16:835–43
8. Baldwin IT. 1994. Chemical changes rapidly induced by folivory. In Insect-Plant Interactions,ed.EA
Bernays, pp. V:2–23. Boca Raton, FL: CRC Press
9. Berenbaum MR. 1995. The chemistry of defense: theory and practice. PNAS 92:2–8
10. Blossey B, Notzold R. 1995. Evolution of increased competitive ability in invasive nonindigenous plants:
a hypothesis. J. Ecol. 83:887–89
11. Boege K, Marquis RJ. 2005. Facing herbivory as you grow up: the ontogeny of resistance in plants. Trends
Ecol. Evol. 20:441–48
388 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
12. Bonaventure G. 2014. Plants recognize herbivorous insects by complex signalling networks. See
Ref. 140, pp. 1–35
13. Bonaventure G, Baldwin IT. 2010. Transduction of wound and herbivory signals in plastids. Commun.
Integr. Biol. 3:313–17
14. Bruce TJA, Wadhams LJ, Woodcock CM. 2005. Insect host location: a volatile situation. Trends Plant
Sci. 10:269–74
15. Bryant JP, Chapin FS, Klein DR. 1983. Carbon/nutrient balance of boreal plants in relation to vertebrate
herbivory. Oikos 40:357–68
16. Chauvin A, Caldelari D, Wolfender J-L, Farmer EE. 2012. Four 13-lipoxygenases contribute to rapid
jasmonate synthesis in wounded Arabidopsis thaliana leaves: a role for LIPOXYGENASE 6 in responses
to long-distance wound signals. New Phytol. 197:566–75
17. Chini A, Fonseca S, Fern
´
andez G, Adie B, Chico JM, et al. 2007. The JAZ family of repressors is the
missing link in jasmonate signalling. Nature 448:666–71
18. Chung HS, Cooke TF, Depew CL, Patel LC, Ogawa N, et al. 2010. Alternative splicing expands the
repertoire of dominant JAZ repressors of jasmonate signaling. Plant J. 63:613–22
19. Chung HS, Koo AJK, Gao X, Jayanty S, Thines B, et al. 2008. Regulation and function of Arabidopsis
JASMONATE ZIM-DOMAIN genes in response to wounding and herbivory. Plant Physiol. 146:952–64
20. Cipollini D, Walters D, Voelckel C. 2014. Costs of resistance in plants: from theory to evidence. See
Ref. 140, pp. 263–337
21. Coley PD. 1987. Interspecific variation in plant anti-herbivore properties: the role of habitat quality and
rate of disturbance. New Phytol. 106(Suppl.):251–63
22. Coley PD. 1988. Effects of plant growth rate and leaf lifetime on the amount and type of anti-herbivore
defense. Oecologia 74:531–36
23. Coley PD, Bryant JP, Chapin FSI. 1985. Resource availability and plant antiherbivore defense. Science
230:895–99
24. Covington MF, Maloof JN, Straume M, Kay SA, Harmer SL. 2008. Global transcriptome analysis reveals
circadian regulation of key pathways in plant growth and development. Genome Biol. 9:R130
25. Darwin C. 1876. The Origin of Species by Means of Natural Selection; or, The Preservation of Favoured Races
in the Struggle for Life. London: Murray
26. Dave A, Graham IA. 2012. Oxylipin signaling: a distinct role for the jasmonic acid precursor cis-(+)-12-
oxo-phytodienoic acid (cis-OPDA). Front. Plant Sci. 3:42
27. De Montaigu A, Giakountis A, Rubin M, T
´
oth R, Cremer F, et al. 2014. Natural diversity in daily
rhythms of gene expression contributes to phenotypic variation. PNAS 112:905–10
28. Dicke M, Baldwin IT. 2010. The evolutionary context for herbivore-induced plant volatiles: beyond the
“cry for help.” Trends Plant Sci. 15(3):167–75
29. Diezel C, Allmann S, Baldwin IT. 2011. Mechanisms of optimal defense patterns in Nicotiana attenuata:
Flowering attenuates herbivory-elicited ethylene and jasmonate signaling. J. Integr. Plant Biol. 53:971–83
30. Dobzhansky T. 1973. Nothing in biology makes sense except in the light of evolution. Am. Biol. Teach.
35:125–29
31. Dodd AN, Dalchau N, Gardner MJ, Baek S, Webb AAR. 2014. The circadian clock has transient plasticity
of period and is required for timing of nocturnal processes in Arabidopsis. New Phytol. 201:168–79
32. Ehrlich PR, Raven PH. 1964. Butterflies and plants: a study in coevolution. Evolution 18:586–608
33. Erb M, Meldau S, Howe GA. 2012. Role of phytohormones in insect-specific plant reactions. Trends
Plant Sci. 17:250–59
34. Fagerstr
¨
om
T, Larsson S
, Tenow O. 1987. On optimal defense in plants. Funct. Ecol. 1:73–81
35. Farmer EE, Gasperini D, Acosta IF. 2014. The squeeze cell hypothesis for the activation of jasmonate
synthesis in response to wounding. New Phytol. 204:282–88
36. Feeney P. 1976. Plant apparency and chemical defense. Recent Adv. Phytochem. 10:1–40
37. Felker-Quinn E, Schweitzer JA, Bailey JK. 2013. Meta-analysis reveals evolution in invasive plant species
but little support for evolution of increased competitive ability (EICA). Ecol. Evol. 3:739–51
38. Fonseca S, Chini A, Hamberg M, Adie B, Porzel A, et al. 2009. (+)-7-iso-jasmonoyl-
L-isoleucine is the
endogenous bioactive jasmonate. Nat. Chem. Biol. 5:344–50
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 389
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
39. Fox LR. 1981. Defense and dynamics in plant-herbivore systems. Am. Zool. 21:853–64
40. Fraenkel GS. 1959. The raison d’
ˆ
etre of secondary plant substances. Science 129:1466–70
41. G
´
alis I, Gaquerel E, Pandey SP, Baldwin IT. 2009. Molecular mechanisms underlying plant memory in
JA-mediated defence responses. Plant Cell Environ. 32:617–27
42. Gilardoni PA, Hettenhausen C, Baldwin IT, Bonaventure G. 2011. Nicotiana attenuata LECTIN RE-
CEPTOR KINASE1 suppresses the insect-mediated inhibition of induced defense responses during
Manduca sexta herbivory. Plant Cell. 23:3512–32
43. Giron D, Frago E, Glevarec G, Pieterse CMJ, Dicke M. 2013. Cytokinins as key regulators in plant-
microbe-insect interactions: connecting plant growth and defence. Funct. Ecol. 27:599–609
44. Gloss AD, Nelson Dittrich AC, Goldman-Huertas B, Whiteman NK. 2013. Maintenance of genetic
diversity through plant-herbivore interactions. Curr. Opin. Plant Biol. 16:443–50
45. Goodspeed D, Chehab EW, Min-Venditti A, Braam J, Covington MF. 2012. Arabidopsis synchronizes
jasmonate-mediated defense with insect circadian behavior. PNAS 109:4674–77
46. Halitschke R, Stenberg JA, Kessler D, Kessler A, Baldwin IT. 2008. Shared signals: “Alarm calls” from
plants increase apparency to herbivores and their enemies in nature. Ecol. Lett. 11:24–34
47. Halitschke R, Ziegler J, Kein
¨
anen M, Baldwin IT. 2004. Silencing of HYDROPEROXIDE LYASE and
ALLENE OXIDE SYNTHASE reveals substrate and defense signaling crosstalk in Nicotiana attenuata.
Plant J. 40:35–46
48. Haloin JR, Strauss SY. 2008. Interplay between ecological communities and evolution: review of feed-
backs from microevolutionary to macroevolutionary scales. Ann. N. Y. Acad. Sci. 1133:87–125
49. Hare JD. 2010. Ontogeny and season constrain the production of herbivore-inducible plant volatiles in
the field. J. Chem. Ecol. 36:1363–74
50. Harmer SL. 2009. The circadian system in higher plants. Annu. Rev. Plant Biol. 60:357–77
51. Hartmann T. 2008. The lost origin of chemical ecology in the late 19th century. PNAS 105:4541–46
52. Haugen R, Steffes L, Wolf J, Brown P, Matzner S, Siemens DH. 2008. Evolution of drought tolerance and
defense: dependence of tradeoffs on mechanism, environment and defense switching. Oikos 117:231–44
53. Heckel DG. 2014. Insect detoxification and sequestration strategies. See Ref. 140, pp. 77–114
54. Heidel-Fischer HM, Musser RO, Vogel H. 2014. Plant transcriptomic responses to herbivory. See
Ref. 140, pp. 155–96
55. Heil M, Karban R. 2010. Explaining evolution of plant communication by airborne signals. Trends Ecol.
Evol. 25:137–44
56. Heiling S, Schuman MC, Schoettner M, Mukerjee P, Berger B, et al. 2010. Jasmonate and ppHsystemin
regulate key malonylation steps in the biosynthesis of 17-hydroxygeranyllinalool diterpene glycosides,
an abundant and effective direct defense against herbivores in Nicotiana attenuata. Plant Cell 22(1):273–92
57. Heinrich M, Hettenhausen C, Lange T, W
¨
unsche H, Fang J, et al. 2013. High levels of jasmonic acid
antagonize the biosynthesis of gibberellins and inhibit the growth of Nicotiana attenuata stems. Plant J.
73:591–606
58. Herms DA, Mattson WJ. 1992. The dilemma of plants: to grow or defend? Q. Rev. Biol. 67:283–335
59. Hilker M, Meiners T. 2010. How do plants “notice” attack by herbivorous arthropods? Biol. Rev. Camb.
Philos. Soc. 85:267–80
60. Howe G, Jander G. 2008. Plant immunity to insect herbivores. Annu. Rev. Plant Biol. 59:41–66
61. Jander G. 2012. Timely plant defenses protect against caterpillar herbivory. PNAS 109:4343–44
62. Jones JDG, Dangl JL. 2006. The plant immune system. Nature 444:323–29
63.
Kallenbach M, Bonaventure
G, Gilardoni PA, Wissgott A, Baldwin IT. 2012. Empoasca leafhoppers
attack wild tobacco plants in a jasmonate-dependent manner and identify jasmonate mutants in natural
populations. PNAS 109:E1548–57
64. Karban R, Baldwin IT. 1997. Induced Responses to Herbivory. Chicago: Univ. Chicago Press
65. Katsir L, Schilmiller AL, Staswick PE, He SY, Howe GA. 2008. COI1 is a critical component of a
receptor for jasmonate and the bacterial virulence factor coronatine. PNAS 105:7100–5
66. Kazan K, Manners JM. 2012. JAZ repressors and the orchestration of phytohormone crosstalk. Trends
Plant Sci. 17:22–31
67. Kazan K, Manners JM. 2013. MYC2: the master in action. Mol. Plant 6:686–703
390 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
68. Kessler A, Heil M. 2011. The multiple faces of indirect defences and their agents of natural selection.
Funct. Ecol. 25:348–57
69. Kessler D, Diezel C, Baldwin IT. 2010. Changing pollinators as a means of escaping herbivores. Curr.
Biol. 20:237–42
70. Keurentjes JJB, Fu J, de Vos CHR, Lommen A, Hall RD, et al. 2006. The genetics of plant metabolism.
Nat. Genet. 38:842–49
71. Kim J, Chang C, Tucker ML. 2015. To grow old: regulatory role of ethylene and jasmonic acid in
senescence. Front. Plant Sci. 6:1–7
72. Kumar P, Pandit SS, Steppuhn A, Baldwin IT. 2014. Natural history-driven, plant-mediated RNAi-based
study reveals CYP6B46’s role in a nicotine-mediated antipredator herbivore defense. PNAS 111:1245–52
73. Larsson S. 1989. Stressful times for the plant stress: insect performance hypothesis. Oikos 56:277–83
74. Lekberg Y, Koide RT. 2014. Integrating physiological, community, and evolutionary perspectives on
the arbuscular mycorrhizal symbiosis. Can. J. Bot. 251:241–51
75. Li D, Baldwin IT, Gaquerel E. 2015. Navigating natural variation in herbivory-induced secondary
metabolism in coyote tobacco populations using MS/MS structural analysis. PNAS 112:E4147–55
76. Loomis WE. 1932. Growth-differentiation balance vs carbohydrate-nitrogen ratio. Proc. Am. Soc. Hortic.
Sci. 29:240–45
77. Loomis WE. 1953. Growth and differentiation—an introduction and summary. In Growth and Differen-
tiation in Plants, ed WE Loomis, pp. 1–17. Ames: Iowa State Coll. Press
78. Maag D, Erb M, K
¨
ollner TG, Gershenzon J. 2015. Defensive weapons and defense signals in plants:
Some metabolites serve both roles. BioEssays 37:167–74
79. Machado RAR, Arce CCM, Ferrieri AP, Baldwin IT, Erb M. 2015. Jasmonate-dependent depletion of
soluble sugars compromises plant resistance to Manduca sexta. New Phytol. 207:91–115
80. Machado RAR, Ferrieri AP, Robert CAM, Glauser G, Kallenbach M, et al. 2013. Leaf-herbivore attack
reduces carbon reserves and regrowth from the roots via jasmonate and auxin signaling. New Phytol.
200:1234–46
81. Maffei M, Bossi S, Spiteller D, Mithoefer A, Boland W. 2004. Effects of feeding Spodoptera littoralis on
lima bean leaves. Plant Physiol. 134:1752–62
82. Maischak H, Grigoriev PA, Vogel H, Boland W, Mith
¨
ofer A. 2007. Oral secretions from herbivorous
lepidopteran larvae exhibit ion channel-forming activities. FEBS Lett. 581:898–904
83. Martin LB, Hawley DM, Ardia DR. 2011. An introduction to ecological immunology. Funct. Ecol. 25:1–4
84. Mayr E. 1961. Cause and effect in biology. Science 134:1501–6
85. McKey D. 1974. Adaptive patterns in alkaloid physiology. Am. Nat. 108:305–20
86. McKey D. 1979. The distribution of secondary compounds within plants. In Herbivores: Their Interactions
with Secondary Plant Metabolites, ed GA Rosenthal, DH Janzen, pp. 55–133. New York: Academic
87. McNickle GG, Dybzinski R. 2013. Game theory and plant ecology. Ecol. Lett. 16:545–55
88. Meldau S, Baldwin IT. 2013. Just in time: circadian defense patterns and the optimal defense hypothesis.
Plant Signal. Behav. 8:e24410
89. Meldau S, Erb M, Baldwin IT. 2012. Defence on demand: mechanisms behind optimal defence patterns.
Ann. Bot. 110:1503–14
90. Meldau S, Wu J, Baldwin IT. 2009. Silencing two herbivory-activated MAP kinases, SIPK and WIPK,
does not increase Nicotiana attenuata’s susceptibility to herbivores in the glasshouse and in nature. New
Phytol. 181:161–73
91. Michael TP, Salom
´
e PA, Yu HJ, Spencer TR, Sharp EL, et al. 2003. Enhanced fitness conferred by
naturally occurring variation in the circadian clock. Science 302:1049–53
92. Milo R, Phillips R. 2015. Cell Biology by the Numbers.
New York: Garland
Sci.
93. Mith
¨
ofer A, Boland W. 2008. Recognition of herbivory-associated molecular patterns. Plant Physiol.
146:825–31
94. Mith
¨
ofer A, Boland W. 2012. Plant defense against herbivores: chemical aspects. Annu. Rev. Plant Biol.
63:431–50
95. Mith
¨
ofer A, Wanner G, Boland W. 2005. Effects of feeding Spodoptera littoralis on lima bean leaves.
II. Continuous mechanical wounding resembling insect feeding is sufficient to elicit herbivory-related
volatile emission. Plant Physiol. 137:1160–68
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 391
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
96. Mousavi SAR, Chauvin A, Pascaud F, Kellenberger S, Farmer EE. 2013. GLUTAMATE RECEPTOR-
LIKE genes mediate leaf-to-leaf wound signalling. Nature 500:422–26
97. Nitao JK, Zangerl AR, Berenbaum MR. 2002. CNB: requiescat in pace? Oikos 98:540–46
98. Nowak MA, Sigmund K. 2004. Evolutionary dynamics of biological games. Science 303:793–99
99. Oh Y, Baldwin IT, G
´
alis I. 2012. NaJAZH regulates a subset of defense responses against herbivores
and spontaneous leaf necrosis in Nicotiana attenuata plants. Plant Physiol. 159:769–88
100. Poelman EH, Bruinsma M, Zhu F, Weldegergis BT, Boursault AE, et al. 2012. Hyperparasitoids use
herbivore-induced plant volatiles to locate their parasitoid host. PLOS Biol. 10:e1001435
101. Price PW, Bouton CE, Gross P, McPheron BA, Thompson JN, Weis AE. 1980. Interactions among
three trophic levels: influence of plants on interactions between insect herbivores and natural enemies.
Annu. Rev. Ecol. Syst. 11:41–65
102. Rhoades DF. 1979. Evolution of plant chemical defense against herbivores. In Herbivores: Their Interaction
with Secondary Plant Metabolites, ed GA Rosenthal, DH Janzen, pp. 1–55. New York: Academic
103. Rhoades DF, Cates RG. 1976. Toward a general theory of plant antiherbivore chemistry. Recent Adv.
Phytochem. 10:168–213
104. Robert-Seilaniantz A, Grant M, Jones JDG. 2011. Hormone crosstalk in plant disease and defense: more
than just jasmonate-salicylate antagonism. Annu. Rev. Phytopathol. 49:317–43
105. Sch
¨
afer M, Br
¨
utting C, Gase K, Reichelt M, Baldwin I, Meldau S. 2013. “Real time” genetic manipulation:
a new tool for ecological field studies. Plant J. 76:506–18
106. Sch
¨
afer M, Meza Canales ID, Br
¨
utting C, Baldwin IT, Meldau S. 2015. Cytokinin levels and NaCHK2-
and NaCHK 3-mediated perception modulate herbivory-induced defense-signaling and defenses in
Nicotiana attenuata. New Phytol. 207:645–58
107. Sch
¨
afer M, Meza-Canales ID, Navarro-Quezada A, Br
¨
utting C, Vankov
´
a R, et al. 2015. Cytokinin levels
and signaling respond to wounding and the perception of herbivore elicitors in Nicotiana attenuata. J.
Integr. Plant Biol. 57:198–212
108. Schmelz EA, Carroll MJ, Leclere S, Phipps SM, Meredith J, et al. 2006. Fragments of ATP synthase
mediate plant perception of insect attack. PNAS 103:8894–99
109. Schmelz EA, Engelberth J, Alborn HT, Tumlinson JH, Teal PEA. 2009. Phytohormone-based activity
mapping of insect herbivore-produced elicitors. PNAS 106:653–57
110. Schultz JC, Appel HM, Ferrieri AP, Arnold TM. 2013. Flexible resource allocation during plant defense
responses. Front. Plant Sci. 4:324
111. Schuman MC, Allmann S, Baldwin IT. 2015. Plant defense phenotypes determine the consequences of
volatile emission for individuals and neighbors. eLife 4:e04490
112. Schuman MC, Barthel K, Baldwin IT. 2012. Herbivory-induced volatiles function as defenses increasing
fitness of the native plant Nicotiana attenuata in nature. eLife 1:e00007
113. Schuman MC, Heinzel N, Gaquerel E, Svatos A, Baldwin IT. 2009. Polymorphism in jasmonate signaling
partially accounts for the variety of volatiles produced by Nicotiana attenuata plants in a native population.
New Phytol. 183:1134–48
114. Schwachtje J, Baldwin IT. 2008. Why does herbivore attack reconfigure primary metabolism? Plant
Physiol. 146:845–51
115.
Schwachtje J, Minchin
PEH, Jahnke S, Van Dongen JT, Schittko U, Baldwin IT. 2006. SNF1-related
kinases allow plants to tolerate herbivory by allocating carbon to roots. PNAS 103:12935–40
116. Seo PJ, Mas P. 2015. STRESSing the role of the plant circadian clock. Trends Plant Sci. 20:230–37
117. Sheard LB, Tan X, Mao H, Withers J, Ben-Nissan G, et al. 2010. Jasmonate perception by inositol-
phosphate-potentiated COI1–JAZ co-receptor. Nature 468:400–7
118. Sherman PW. 1988. The levels of analysis. Anim. Behav. 36:616–19
119. Shiojiri K, Ozawa R, Kugimiya S, Uefune M, Van Wijk M, et al. 2010. Herbivore-specific, density-
dependent induction of plant volatiles: honest or “cry wolf” signals? PLOS One. 5:e12161
120. Snook ME, Johnson AW, Severson RF, Teng Q, White RA, et al. 1997. Hydroxygeranyllinalool gly-
cosides from tobacco exhibit antibiosis activity in the tobacco budworm [Heliothis virescens (F.)]. J Agric.
Food. Chem. 45:2299–308
121. Spiteller D, Pohnert G, Boland W. 2001. Absolute configuration of volicitin, an elicitor of plant volatile
biosynthesis from lepidopteran larvae. Tetrahedron Lett. 42:1483–85
392 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
122. Stam JM, Kroes A, Li Y, Gols R, van Loon JJA, et al. 2014. Plant interactions with multiple insect
herbivores: from community to genes. Annu. Rev. Plant Biol. 65:689–713
123. Stamp N. 2003. Out of the quagmire of plant defense hypotheses. Q. Rev. Biol. 78:23–55
124. Steppuhn A, Baldwin IT. 2007. Resistance management in a native plant: Nicotine prevents herbivores
from compensating for plant protease inhibitors. Ecol. Lett. 10:499–511
125. Steppuhn A, Baldwin IT. 2008. Induced defenses and the cost-benefit paradigm. In Induced Plant Resistance
to Herbivory, ed. A Schaller, pp. 61–83. Dordrecht, Neth.: Springer
126. Stitt M, Sulpice R, Keurentjes J. 2010. Metabolic networks: how to identify key components in the
regulation of metabolism and growth. Plant Physiol. 152:428–44
127. Stork W, Diezel C, Halitschke R, G
´
alis I, Baldwin IT, et al. 2009. An ecological analysis of the herbivory-
elicited JA burst and its metabolism: plant memory processes and predictions of the moving target model.
PLOS ONE 4(3):e4697
128. Thines B, Katsir L, Melotto M, Niu Y, Mandaokar A, et al. 2007. JAZ repressor proteins are targets of
the SCF(COI1) complex during jasmonate signalling. Nature 448:661–65
129. Thompson JN. 1989. Concepts of coevolution. Trends Ecol. Evol. 4:179–83
130. Tinbergen N. 1963. On aims and methods of ethology. Z. Tierpsychol. 20:410–33
131. Truitt CL, Wei H-X, Par
´
e PW. 2004. A plasma membrane protein from Zea mays binds with the
herbivore elicitor volicitin. Plant Cell 16:523–32
132. Tsai AY-L, Gazzarrini S. 2014. Trehalose-6-phosphate and SnRK1 kinases in plant development and
signaling: the emerging picture. Front. Plant Sci. 5:1–11
133. Tuomi J, Niemela P, Chapin FSI, Bryant JP, Siren S. 1988. Defensive responses of trees in relation to
their carbon/nutrient balance. In Mechanisms of Woody Plant Defenses Against Insects: Search for Pattern,
ed. WJ Mattson, J Levieux, C Bernard-Dagan, pp. 57–72. New York: Springer
134. Van Dam NM, Baldwin IT. 2001. Competition mediates costs of jasmonate-induced defences, nitrogen
acquisition and transgenerational plasticity in Nicotiana attenuata. Funct. Ecol. 15:406–15
135. Van Dam NM, Horn M, Mares M, Baldwin IT. 2001. Ontogeny constrains systemic protease inhibitor
response in Nicotiana attenuata. J. Chem. Ecol. 27:547–68
136. Van Kleunen M, Weber E, Fischer M. 2010. A meta-analysis of trait differences between invasive and
noninvasive plant species. Ecol. Lett. 13:235–45
137. Van Poecke R, Dicke M. 2003. Signal transduction downstream of salicylic and jasmonic acid in
herbivory-induced parasitoid attraction by Arabidopsis is independent of JAR1 and NPR1. Plant. Cell
Environ. 26:1541–48
138. VanDoorn A, Bonaventure G, Rogachev I, Aharoni A, Baldwin IT. 2011. JA-Ile signaling in Solanum
nigrum is not required for defense responses in nature. Plant Cell Environ. 34:2159–71
139. VanDoorn A, Kallenbach M, Borquez AA, Baldwin IT, Bonaventure G. 2010. Rapid modification of
the insect elicitor N-linolenoyl-glutamate via a lipoxygenase-mediated mechanism on Nicotiana attenuata
leaves. BMC Plant Biol. 10:164
140. Voelckel C, Jander G, eds. 2014. Annual Plant Reviews Volume 47: Insect-Plant Interactions. Chichester,
UK: Wiley
141. Waldbauer GP, Fraenkel G. 1961. Feeding on normally rejected plants by maxillectomized larvae of the
tobacco hornworm, Protoparce sexta (Lepidoptera, Sphingidae). Ann. Entomol. Soc. Am. 54:477–85
142. Wang W, Barnaby JY, Tada Y, Li H, To M, et al. 2011. Timing of plant immune responses by a central
circadian
regulator. Nature 470:110–15
143. Wasternack C.
2007. Jasmonates: an update on biosynthesis, signal transduction and action in plant stress
response, growth and development. Ann. Bot. 100:681–97
144. Wasternack C, Hause B. 2013. Jasmonates: biosynthesis, perception, signal transduction and action in
plant stress response, growth and development. An update to the 2007 review in Annals of Botany. Ann.
Bot. 111:1021–58
145. Whitham TG, Bailey JK, Schweitzer JA, Shuster SM, Bangert RK, et al. 2006. A framework for com-
munity and ecosystem genetics: from genes to ecosystems. Nat. Rev. Genet. 7:510–23
146. Wijnen H, Young MW. 2006. Interplay of circadian clocks and metabolic rhythms. Annu. Rev. Genet.
40:409–48
www.annualreviews.org
Layers of Plant Responses to Insect Herbivores 393
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
EN61CH20-Schuman ARI 3 December 2015 21:34
147. Wu J, Baldwin IT. 2010. New insights into plant responses to the attack from insect herbivores. Annu.
Rev. Genet. 44:1–24
148. Wu J, Hettenhausen C, Meldau S, Baldwin IT. 2007. Herbivory rapidly activates MAPK signaling in
attacked and unattacked leaf regions but not between leaves of Nicotiana attenuata. Plant Cell 19:1096–122
149. Wu J, Hettenhausen C, Schuman MC, Baldwin IT. 2008. A comparison of two Nicotiana attenuata
accessions reveals large differences in signaling induced by oral secretions of the specialist herbivore
Manduca sexta. Plant Physiol. 146:927–39
150. Xie D-X, Feys BF, James S, Nieto-Rostro M, Turner JG. 1998. COI1: an Arabidopsis gene required for
jasmonate-regulated defense and fertility. Science 280:1091–94
151. Yoshinaga N, Aboshi T, Abe H, Nishida R, Alborn HT, et al. 2008. Active role of fatty acid amino acid
conjugates in nitrogen metabolism in Spodoptera litura larvae. PNAS 105:18058–63
152. Zhang L, Wang A. 2012. Virus-induced ER stress and the unfolded protein response. Front Plant Sci.
3:293
RELATED RESOURCES
Farmer EE. 2014. Leaf Defense. Oxford, UK: Oxford Univ. Press
Trewavas A. 2014. Plant Behavior and Intelligence. Oxford, UK: Oxford Univ. Press
394 Schuman
·
Baldwin
Changes may still occur before final publication online and in print
Annu. Rev. Entomol. 2016.61. Downloaded from www.annualreviews.org
Access provided by WIB6417 - Max-Planck-Gesellschaft on 01/07/16. For personal use only.
... Plants employ constitutive defenses to prevent herbivore infestation; when attacked by insect herbivores, plants activate induced defenses by perceiving herbivore-and/or damage-associated molecular patterns (HAMPs/DAMPs) [1,2]. These induced defenses are produced locally and systemically, and may influence the performance of conspecific and non-conspecific herbivores that share the same host plant, either simultaneously or successively, directly and/or indirectly, by attracting the natural enemies of insect herbivores [3][4][5]. In response to plants, insect herbivores have evolved various counter-defense strategies, such as secreting effectors or factors to inhibit plant defenses or enhance plant susceptibility, and detoxifying or accommodating plant defensive compounds [4,6,7]. ...
... These induced defenses are produced locally and systemically, and may influence the performance of conspecific and non-conspecific herbivores that share the same host plant, either simultaneously or successively, directly and/or indirectly, by attracting the natural enemies of insect herbivores [3][4][5]. In response to plants, insect herbivores have evolved various counter-defense strategies, such as secreting effectors or factors to inhibit plant defenses or enhance plant susceptibility, and detoxifying or accommodating plant defensive compounds [4,6,7]. These molecular interactions between plants and insect herbivores largely determine the consequences of plant-herbivore interactions that are observable at the macroscopic level: the plant either resists or is susceptible to the herbivore, which in turn affects the population density and diversity of both plants and insect herbivores. ...
... Plant defense responses are accompanied by the activation and suppression of defense genes, with transcription factors (TFs) playing a crucial role in this process [4,6,7]. Hitherto, TFs associated with rice defense responses have been reported to be mainly in the WRKY, AP2/ERF (APETALA2/ethylene responsive factor), MYB, bZIP (basic leucine zipper factor), and bHLH (basic helixloop-helix) subfamilies (Fig. 1). ...
Article
Full-text available
To adapt to each other, plants and insect herbivores have developed sophisticated molecular interactions. Here, we summarize current knowledge about such molecular interactions between rice, a globally important food crop, and insect herbivores. When infested by insect herbivores, rice perceives herbivore- and/or damage-associated molecular patterns (HAMPs/DAMPs) via receptors that activate early signaling events such as the influx of Ca2+, the burst of reactive oxygen species, and the activation of MPK cascades. These changes result in specific rice defenses via signaling networks that mainly include phytohormones (jasmonic acid, salicylic acid, ethylene, and abscisic acid) and transcription factors. Some compounds, including flavonoids, phenolamides, defensive proteins, and herbivore-induced rice volatiles, have been reported to be used by rice against insects. Insect herbivores can deliver effectors or factors to inhibit rice defenses or enhance rice susceptibility. Although the number of HAMPs and defense-suppressing effectors from rice piercing-sucking insects has increased rapidly, none from rice chewing insects has been identified. Moreover, herbivore effectors or factors that induce rice susceptibility, and rice immune receptors recognizing HAMPs or effectors, are not well characterized. We point out future research directions in this area and highlight the importance of elucidating the mechanisms for rice sensing of insect herbivores and for insect counter-defenses against plants.
... In their habitats, plants are often confronted with the attacks of insect herbivores, which requires proper balancing of growth and defense responses (McMillan, 2023). Consequently, plants have evolved several mechanical and biochemical defense strategies; some of which are constitutive, and others inducible to counter herbivores and preserve their own fitness (Schuman and Baldwin, 2016). Mechanical defense includes various inherent structural traits, such as waxy cuticles, trichomes, hairs, and spines which collectively deter insects from feeding on plants (Mitchell et al., 2016;Kaur and Kariyat, 2023;Balakrishnan et al., 2024). ...
Article
Full-text available
Silicon (Si) uptake is generally beneficial for plants that need protection from insect herbivores. In pursue of mechanisms involved in Si-mediated defense, we comprehensively explored the impact of Si on several defensive and metabolic traits in rice exposed to simulated and real herbivory of Mythimna loreyi Duponchel larvae. Hydroponic experiments showed that Si-deprived rice supplemented with Si 72 h prior to insect infestation were similarly resistant to larvae as plants continuously grown in Si-containing media. Both Si and herbivory altered primary metabolism in rice, including the levels of several sugars, amino acids, and organic acids. While the accumulation of sugars was generally positively correlated with Si presence, multiple amino acids showed a negative correlation trend with Si supplementation. The levels of secondary metabolites, including isopentylamine, p-coumaroylputrescine and feruloylputrescine, were typically higher in the leaves of Si-supplemented plants exposed to herbivory stress compared to Si-deprived plants. In addition, simulated herbivory treatment in Si-supplemented plants induced more volatile emissions relative to Si-deprived plants, which was consistent with the increased transcripts of key genes involved in volatile biosynthesis. In ecological interactions, Si alone did not affect the oviposition choice of M. loreyi but gravid females showed a significant preference for simulated herbivory-treated/Si-deprived compared to Si-supplemented plants. Our data suggest that apart from mechanical defense, Si may affect rice metabolism in multiple ways that might enhance/modulate defense responses of rice under herbivory stress.
... Compared with the CK treatment, the expression levels of DEGs such as CPK25, ATRBOHB, CALM, and CML were significantly upregulated in the plant-pathogen interaction pathway (map 04626) and MAPK plant signaling pathway (map 04016) in "Cabernet Sauvignon" leaves under AI treatment. External stresses, such as insect feeding, have been shown to affect the processes within plant cells as well as the growth and reproduction of plants (Schuman & Baldwin, 2016) and also affect plant intracellular processes and plant growth and reproduction. The plant cell wall F I G U R E 7 Validation map for RT-qPCR includes the following: (a) PCR amplification electrophoresis; (b) melting curves of 20 differential genes and internal reference genes; and (c) linear regression analysis of the RNA-seq and RT-qPCR data framework is formed by the interaction of xylan and microfilaments. ...
Article
Full-text available
To investigate the molecular mechanism of the defense response of “Cabernet Sauvignon” grapes to feeding by Apolygus lucorum, high‐throughput sequencing technology was used to analyze the transcriptome of grape leaves under three different treatments: feeding by A. lucorum, puncture injury, and an untreated control. The research findings indicated that the differentially expressed genes were primarily enriched in three aspects: cellular composition, molecular function, and biological process. These genes were found to be involved in 42 metabolic pathways, particularly in plant hormone signaling metabolism, plant‐pathogen interaction, MAPK signaling pathway, and other metabolic pathways associated with plant‐induced insect resistance. Feeding by A. lucorum stimulated and upregulated a significant number of genes related to jasmonic acid and calcium ion pathways, suggesting their crucial role in the defense molecular mechanism of “Cabernet Sauvignon” grapes. The consistency between the gene expression and transcriptome sequencing results further supports these findings. This study provides a reference for the further exploration of the defense response in “Cabernet Sauvignon” grapes by elucidating the expression of relevant genes during feeding by A. lucorum.
... Herbivore-induced changes in plant phenotype have been shown to affect the plant's susceptibility to later attacks, connecting plant herbivores through a network of indirect interactions (Huang et al., 2017;Kessler & Halitschke, 2007;Ohgushi, 2005). These changes in the phenotype of plants emerge through induced plant responses involving myriad chemical and morphological traits to enhance resistance to current or future attackers (Karban, 2011;Schuman & Baldwin, 2016), or result from antagonists manipulating plant responses or modifying the plant tissues on which they feed (Behmer, 2009;Dussourd, 2017;Lill & Marquis, 2003). In addition to shifts in the identity of attackers, induced plant responses that occur early in the season can affect the trajectory of insect population growth that persists throughout the development of the plant (Karban, 1993;Wold & Marquis, 1997). ...
Article
Full-text available
Insect herbivores can directly affect plant reproduction by feeding on reproductive tissues, or indirectly by feeding on vegetative tissues for which plants are unable to compensate. Additionally, early arriving herbivores may have cascading effects on plant reproduction by altering the later arriving community. However, the dynamic interplay between plant development and the assembly of herbivore communities remains underexplored. Hence, it is unclear whether non‐outbreak levels of ambient herbivory early in the development of plants can impact plant fitness and to what extent these effects are mediated through changes in plant development and subsequent herbivory. By excluding the herbivore community in an exclosure experiment and by manipulating early‐season herbivory in a common garden field experiment replicated across four Brassicaceae species and 2 years, we tested whether early‐season herbivory by caterpillars (Pieris rapae) or aphids (Myzus persicae) affected development, reproduction, and the herbivore communities associated with individual plants. In addition, we tested a causal hypothesis to assess the relative importance and temporal interplay between variation in herbivore communities and variation in plant development in determining plant reproduction. Early‐season herbivory affected plant reproduction in the exclosure experiment, with effects being highly dependent on the plant species, the herbivore species and the year. However, we found no such effects in the field experiment. The exploratory path analysis indicated that variation in plant reproduction is best predicted by variation in plant development, explaining 80% of the total effect on seed production. This suggests early‐season herbivory had limited effects on later plant development, and plants were able to attenuate the impact of early‐season herbivory. However, no clear compensatory mechanism could be identified. While early‐season herbivory has the potential to affect plant reproduction through changes in plant development or the subsequent development of the associated community, these effects were small and varied across closely related species. This suggests that plant species may be exposed to different levels of natural selection by early‐season herbivores through plant‐ or community‐mediated effects on reproduction. Read the free Plain Language Summary for this article on the Journal blog.
... Moreover, insect herbivore faces plant defenses comprising unpalatable substances (e.g., secondary metabolites [14], proteinase inhibitor [15], and hardened cell walls [16]), and natural enemies of herbivores [17]. A substantial part of plant defense is activated by attack primarily through jasmonic acid (JA) signaling [18], which is referred induced defense. ...
Article
Full-text available
Background Preingestive behavioral modulations of herbivorous insects on the host plant are abundant over insect taxa. Those behaviors are suspected to have functions such as deactivation of host plant defenses, nutrient accumulation, or modulating plant-mediated herbivore interactions. To understand the functional consequence of behavioral modulation of insect herbivore, we studied the girdling behavior of Phytoecia rufiventris Gautier (Lamiinae; Cerambycidae) on its host plant Erigeron annuus L. (Asteraceae) that is performed before endophytic oviposition in the stem. Results The girdling behavior significantly increased the larval performance in both field monitoring and lab experiment. The upper part of the girdled stem exhibited lack of jasmonic acid induction upon larval attack, lowered protease inhibitor activity, and accumulated sugars and amino acids in compared to non-girdled stem. The girdling behavior had no effect on the larval performance of a non-girdling longhorn beetle Agapanthia amurensis , which also feeds on the stem of E. annuus during larval phase. However, the girdling behavior decreased the preference of A. amurensis females for oviposition, which enabled P. rufiventris larvae to avoid competition with A. amurensis larvae. Conclusions In conclusion, the girdling behavior modulates plant physiology and morphology to provide a modulated food source for larva and hide it from the competitor. Our study implies that the insect behavior modulations can have multiple functions, providing insights into adaptation of insect behavior in context of plant-herbivore interaction.
... caterpillars, differ from responses to sucking insects, e.g. aphids or whitefly (Wu & Baldwin 2010;Schuman & Baldwin 2016). These different plant responses also lead to an expectation, and finding, that not all rhizobacteria will be effective against all insect herbivores (Blubaugh et al. 2018;Friman et al. 2021). ...
Preprint
Rhizobacteria inoculation of plants has shown promising potential for enhancing resistance against insect pests by reducing herbivore fitness and altering herbivore-natural enemy interactions. Understanding interactions among specific rhizobacterial species and their plants in the rhizosphere is crucial for developing effective strategies to harness these benefits for pest management in agriculture. We present a meta-analysis examining the impact of rhizobacteria inoculation on herbivore interactions with plants. The findings indicate that rhizobacteria inoculation generally reduces herbivore fitness and host choice behaviours. Rhizobacteria inoculation may also enhance the recruitment of natural enemies of herbivores, thus increasing top-down predator control. The effects on herbivores varied significantly depending on the rhizobacterial species, with Bacillus spp. showing stronger effects compared to other commonly studied Pseudomonas spp. Rhizobacteria notably reduced traits such as host choice, leaf consumption, survival, and reproduction of chewing herbivores, while primarily impacting sucking herbivores by reducing reproduction. Single-strain inoculants tended to perform better, especially for sucking herbivores, suggesting potential strain incompatibility issues with multi-strain inoculants. Furthermore, field trials showed less impact on insect fitness reduction compared to experiments under controlled conditions, possibly due to soil diversity and environmental factors affecting inoculant persistence. These results underscore the need for considering broader environmental interactions when developing effective rhizobacteria-based pest management strategies. Understanding specific and generalist rhizosphere interactions can aid in developing synthetic microbial communities with broad protective functions across various plants and environments.
... Changes in amino acid metabolism, especially the increased synthesis of aspartic acid and asparagine, are related to maturation, senescence, ammonia accumulation, and detoxification in plants [30]. Alpha-linolenic acid metabolism produces the precursors for jasmonic acid biosynthesis [31]. Jasmonic acid promotes leaf senescence in Arabidopsis [32]. ...
Article
Full-text available
NaHCO3 accelerates the aging of tobacco leaves; however, the underlying molecular mechanisms have not been elucidated. This study aimed to explore the mechanism of NaHCO3 in the promotion of tobacco leaf maturation using transcriptome analysis. Leaves on plants or detached leaves of the tobacco variety, Honghua Dajinyuan, were sprayed with or without 1% NaHCO3. The leaf yellowing was observed, the pigment content and enzyme activities were determined and RNA sequencing (RNA-seq) was performed. Spraying NaHCO3 onto detached leaves was found to promote leaf yellowing. Pigment content, catalase activity, and superoxide dismutase activity significantly decreased, whereas peroxidase activity and malondialdehyde content significantly increased. RNA-seq demonstrated that spraying with NaHCO3 upregulated genes associated with cysteine and methionine metabolism; alpha-linolenic acid metabolism; and phenylalanine, tyrosine, and tryptophan biosynthesis and downregulated genes related to photosynthesis and carotenoid biosynthesis. Genes correlated with autophagy-other, valine, leucine, and isoleucine degradation, and the MAPK signaling pathway were upregulated while those correlated with DNA replication, phenylalanine, and tyrosine and tryptophan biosynthesis were downregulated in detached leaves sprayed with NaHCO3 compared with the plant leaves sprayed with NaHCO3. Overall, this study is the first to elucidate the molecular and metabolic mechanisms of NaHCO3 in the promotion of tobacco leaf maturation.
Chapter
For plants, insect attack is a complex stimulus which can be divided into three conceptual phases: pest recognition, signal transduction and deployment of defences. Here, we give an overview of the different stages of pest recognition by the plant, focusing on the transcriptomic aspects of these responses. Separating wound‐ and herbivore‐specific components of insect attack is as important as the distinction between attack from different feeding guilds such as chewing and piercing‐sucking herbivores to mount appropriate plant defence responses. We present a meta‐analysis of the existent microarray studies, intending to provide a better overview over plant gene expression data from a wider range of insect‐plant interaction studies. While the existent microarray studies give valid information about gene regulation in attacked plants, they also highlight the importance of synchronizations of experimental designs, as different model systems, controls and experimental time points make it hard to generalize results. Most of the past and current research has focused on single herbivore attack. In nature, however, simultaneous or subsequent attacks by insects are frequently observed, so we provide a brief overview over the current stage of knowledge of simultaneous attacks, ending with an outlook to future challenges in this field of science.
Chapter
The recognition of phytophagous insects by plants induces a set of very specific responses aimed at deterring tissue consumption and reprogramming plant metabolism and development to tolerate herbivory. This recognition requires the plant's ability to perceive chemical cues generated by the insects and to distinguish a particular pattern of tissue disruption. Relatively little is known about the molecular basis of insect perception by plants and the signalling mechanisms directly associated with this perception. Importantly, the insect feeding behaviour (piercing‐sucking versus chewing) is a decisive determinant of the plant's defence response, and the mechanisms used to perceive insects from different feeding guilds may be distinct. During insect feeding, components of the saliva of chewing or piercing‐sucking insects come into contact with plant cells, and elicitors or effectors present in this insect‐derived fluid are perceived by plant cells to initiate the activation of specific signalling cascades.
Chapter
Plants defend themselves from insect herbivores with a vast array of chemical defences, yet insects have evolved several mechanisms for detoxifying these and even sequestering them for their own use. Gene duplication within the large cytochrome P450 gene family plays diverse roles in detoxification, employing fine‐tuning of substrate specificity, transcriptional control and insensitivity to plant‐derived P450 inhibitors. Multiple mechanisms exist for avoiding activation of glucosinolates and cyanogenic glucosides, and for sequestering or even synthesizing the latter. Housekeeping enzymes have been re‐directed to novel detoxicative functions. Often, the primary molecular target of the plant toxin has mutated to a less sensitive form, and multiple parallel evolutionary responses of this type can be seen in communities of herbivores consuming plants posing the same chemical challenges. These and other mechanisms have enabled insects to continue to exploit plants for food throughout the long history of their co‐evolutionary struggle.
Book
This latest volume in Wiley Blackwell's prestigious Annual Plant Reviews brings together articles that describe the biochemical, genetic, and ecological aspects of plant interactions with insect herbivores.. The biochemistry section of this outstanding volume includes reviews highlighting significant findings in the area of plant signalling cascades, recognition of herbivore-associated molecular patterns, sequestration of plant defensive metabolites and perception of plant semiochemicals by insects. Chapters in the genetics section are focused on genetic mapping of herbivore resistance traits and the analysis of transcriptional responses in both plants and insects. The ecology section includes chapters that describe plant-insect interactions at a higher level, including multitrophic interactions, investigations of the cost-benefit paradigm and the altitudinal niche-breadth hypothesis, and a re-evaluation of co-evolution in the light of recent molecular research. Written by many of the world's leading researchers in these subjects, and edited by Claudia Voelckel and Georg Jander, this volume is designed for students and researchers with some background in plant molecular biology or ecology, who would like to learn more about recent advances or obtain a more in-depth understanding of this field. This volume will also be of great use and interest to a wide range of plant scientists and entomologists and is an essential purchase for universities and research establishments where biological sciences are studied and taught.
Chapter
Haukioja and Hakala (1975), Rhoades (1979), and Haukioja (1980) outlined the hypothesis that herbivore cycles are associated with changes in the inducible defense level of individual plants. For example, the herbivore population’s increase begins when host-plant resistance decreases to a low level. On the other hand, the herbivore population’s decrease is precipitated by an increasing plant resistance. The minimum length of the latent phase depends on the rate of relaxation of resistance as the plant recovers from defoliation (May 1975, Haukioja 1980). Although the detailed mechanisms of these changes in plant resistance are still unclear, defoliation has been shown to induce changes in trees that have adverse effects on the growth, reproduction, and survival of lepidopteran defoliators (Haukioja and Niemelä 1977, Wallner and Walton 1979, Werner 1979, Haukioja 1980) and which can have relaxation times of several years (Benz 1974, Haukioja 1982). However, in other cases, defoliation caused no observable increase in plant resistance (Myers 1981), and defoliation may even reduce resistance of the host plant (Niemelä et al. 1984). Furthermore, no evidence for defensive responses to clipping was found in dormant twigs of juvenile woody plants when twig age and diameter were controlled (Chapin et al. 1985).
Article
The degree of herbivory and the effectiveness of defenses varies widely among plant species. Resource availability in the environment is proposed as the major determinant of both the amount and type of plant defense. When resources are limited, plants with inherently slow growth are favored over those with fast growth rates; slow rates in turn favor large investments in antiherbivore defenses. Leaf lifetime, also determined by resource availability, affects the relative advantages of defenses with different turnover rates. Relative limitation of different resources also constrains the types of defenses. The proposals are compared with other theories on the evolution of plant defenses.