13.07.2015 Views

Herbivore-induced, indirect plant defences - UPCH

Herbivore-induced, indirect plant defences - UPCH

Herbivore-induced, indirect plant defences - UPCH

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

92G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111with the first contact between <strong>plant</strong> and herbivore, andincluding herbivore recognition, signal elicitation, and earlysignalling steps, and followed by a description of thesignalling cascades and biosynthesis pathways involved,we will end with an overview of the ecology and evolution of<strong>induced</strong>, <strong>indirect</strong> defensive <strong>plant</strong> traits.2. Cell- and long-distance signalling in <strong>plant</strong>s in responseto herbivoryPlants have evolved a large array of interconnected cellsignallingcascades, resulting in local resistance and longdistancesignalling for systemic acquired resistance (SAR).Such responses were initiated with the recognition ofphysical and chemical signals of the feeding herbivores,activate subsequent signal transduction cascades, and finallylead to an activation of genes involved in defence responsesthat consequently enhance feedback signalling and metabolicpathways (Fig. 1). This section describes the aspects ofcell signalling involved in the induction of <strong>indirect</strong>herbivore responses.2.1. ElicitorsIn the 1990s and earlier, many chemical ecologistsbelieved that herbivorous oral secretions and regurgitantselicited insect-<strong>induced</strong> <strong>plant</strong> responses, since simplemechanical wound stimuli, in many cases, could not mimic<strong>plant</strong> responses following insect attack [2–5]. h-Glucosidasederived from regurgitate of Pieris brassicae larvae wasidentified as a potential elicitor of herbivore-<strong>induced</strong> <strong>plant</strong>Fig. 1. Schematic representation of the signalling pathways required for herbivore-<strong>induced</strong> responses in <strong>plant</strong>s. This scheme merges the evidence obtained fromseveral <strong>plant</strong> taxa. The overall scenario may differ in certain <strong>plant</strong>s; in particular the existence and the extent of synergistic and antagonist interaction betweenpathways may vary significantly. Elements in blue represent enzymes. Broken arrows indicate possible steps not yet described. Abbreviations: ACC, 1-aminocyclopropane-1-carboxylic acid; ACS, ACC synthase; DAG; diacylglycerol; FAC, fatty acid-amino acid conjugate; FAD, N-3 fatty acid desaturase;HIPV, herbivore-<strong>induced</strong> <strong>plant</strong> volatiles; JA, jasmonic acid; JMT, JA carboxyl methyl transferase; LOX, lipoxygenase; MAPK, mitogen-activated proteinkinase; MeJA, methyl JA; MeSA, methyl SA; OPDA, 12-oxophytodienoic acid; PL, phospholipase; PA, phosphatidic acid; SA, salicylic acid; SAM, S-adenosyl-methionine; SAMT, SA carboxyl methyl transferase; TF, transcription factor.


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 93volatiles (HIPVs) [6]. The application of this h-glucosidaseto mechanically wounded parts of cabbage leaves resulted inthe emission of a volatile blend that attracted parasitic wasps(Cotesia glomerata). On the other hand, the application ofglucose oxidase, a salivary gland enzyme, inhibited thesynthesis of nicotine in tobacco leaves and may thus beresponsible for the observed resistance of feeding herbivores[7]. The suppressive effect, which maintains the <strong>plant</strong>palatable, may be due to the products of glucose oxidase,namely hydrogen peroxide (H 2 O 2 ) and gluconic acid.N-(17-hydroxylinolenoyl)-l-glutamine (volicitin), a fattyacid-amino acid conjugate (FAC), was detected in the oralsecretion of beet armyworm larvae (Spodoptera exigua) [8](Fig. 2). This FAC was identified as the main elicitorinducing the emission of a volatile blend in maize <strong>plant</strong>s(Zea mays) similar to that emitted when caterpillars fed onthem. While volicitin was isolated from S. exigua, otherFACs with biological activity, such as N-acyl Gln/Glu, havebeen isolated from the regurgitate of several lepidopteranspecies [9–11]. For example, the FAC N-linolenoyl-Glu inthe regurgitate of the tobacco hornworm (Manduca sexta)was characterised as a potential elicitor of volatile emissionin tobacco <strong>plant</strong>s. FACs can induce the accumulation of 7-epi-jasmonic acid (JA) an octadecanoid-derived phytohormonethat elicited transcripts of herbivore-responsive genesin the tobacco <strong>plant</strong>s [9]. 7-Epi-jasmonic acid is theoriginally produced and biologically active form of JA,which is isomerised in <strong>plant</strong>a to the less active trans-isomer(see Fig. 1). Recently, a plasma membrane protein frommaize was suggested as a volicitin-binding protein [12].Plants that have been either pre-treated with methyljasmonate (MeJA) or S. exigua feeding showed increasedbinding activity of labelled volicitin to the plasma membrane.These findings suggest that volicitin-<strong>induced</strong> JAenhanced the interactions of the plasma membrane proteinwith volicitin as a ligand or the <strong>induced</strong> formation of theproteins after the recognition of volicitin. However, volicitinis not generally active. Leaves of Lima bean (Phaseoluslunatus), for example, do not respond to volicitin or to otherFACs with the induction of volatile emission [13].2.2. Early and secondary signal transduction pathwaysAfter mechanical damage, reactive oxygen species weregenerated near cell walls of tomato vascular bundle cells andthe resulting H 2 O 2 was shown to act as a second messengerfor the activation of defensive genes in mesophyll cellswhich are expressed later [14]. In Lima bean leaves infestedwith Spodoptera littoralis larvae, a depolarization of themembrane potential (V m ) and intracellular calcium influxwere observed only at the site of damage [15]. Simplemechanical wounding of the leaves also caused V mdepolarization, yet without a concomitant influx of calcium.Hence, already at this early step of signal recognition andtransduction, the pathways offer different responses towounding and herbivore attack. Thus, both wounding andthe introduction of herbivore-specific elicitors appear to beessential for the full induction of defence responses.On the other hand, recent studies applying a continuousrather than a single instance of mechanical damage (patternwheel) to Lima bean leaves clearly resulted in the emissionof volatile blends resembling those that occur afterherbivore damage [16]. As mentioned above, Lima beanand cotton (Gossypium hirsutum) do not respond to FACswith increased volatile emissions [13]. This observationmay be consistent with a high internal threshold forFig. 2. N-acyl glutamines from oral secretions of lepidopteran larvae.


94G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111mechanical damage which, in some <strong>plant</strong> species, is reachedonly after continuous wounding and which triggers thetypical repertoire of defence responses caused by elicitors.cDNA was isolated from a species of tobacco, encoding amitogen-activated protein kinase (MAPK), whose transcriptbegins to accumulate in the leaves 1 min after a singlemechanical wounding event [17]. This protein kinase(WIPK) is considered essential for JA formation and JA<strong>induced</strong>responses in tobacco, since transgenic <strong>plant</strong>s inwhich WIPK has been silenced show impaired accumulationof JA and MeJA after wounding [17]. In contrast, the loss ofWIPK function in this transgenic line resulted in anincreased accumulation of salicylic acid (SA) and its sugarconjugate salicylic acid h-glucoside after wounding; however,such accumulation was not observed in wild-type<strong>plant</strong>s. The signalling pathways associated with JA and SAare generally thought to cross-communicate antagonistically(see below). Moreover, WIPK has also been reported toelicit the transcription of a gene for a N-3 fatty aciddesaturase (FAD7), which catalyses the conversion oflinoleic acid to linolenic acid, a precursor of JA [18]. Thus,WIPK may be an early activator of the octadecanoidpathway. WIPK and the SA-<strong>induced</strong> protein kinase (SIPK)were found to share an upstream MAPKK, NtMEK2 [19],and to interact with the calmodulin (CaM)-binding MAPKphosphatase (NtMKP1) [20]. Both kinases are commonlyinvolved in wound-mediated defence responses in tobacco[19,21,22]. If the MAPK cascades in tobacco were onlyinterlinked as described above, an active mutant of theupstream kinase of SIPK (NtMEK2 DD ) should haveincreased the accumulation of JA and MeJA, but it didnot [23]. Instead, ethylene was assumed to be involved in<strong>plant</strong> defence responses mediated by the NtMEK2-SIPK/WIPK pathway. This discrepancy may be not only due tothe complicated interactions of protein kinase/phosphatasecascades, but also to cross-talk between the JA-, SA-, andethylene-signalling pathways (see below). Since calciumdependentprotein kinases (CDPKs) are regularly involvedin signal transduction of a variety of biotic and abioticstresses [24], their involvement as active protein cascades inherbivore/wound responses cannot be excluded. CDPKscompose a large family of serine/threonine kinases in <strong>plant</strong>s(for example, 34 members in Arabidopsis) [25,26]. JA hasbeen reported to affect CDPK transcript and activity inpotato <strong>plant</strong>s [27].A transgenic Arabidopsis line in which the expressionlevel of phospholipase Da (PLDa) is suppressed has beenreported to show a decreased wound-<strong>induced</strong> accumulationof phosphatidic acid and JA, as well as JA-inducibletranscript activation [28]. Curiously, this loss-of-functionmutant also featured decreased expression of the genecoding lipoxygenase2 (LOX2) but not of any other geneinvolved in the octadecanoid pathway. LOX2 is one of thekey enzymes for the wound-<strong>induced</strong> synthesis of JA inchloroplasts [29]. Therefore, PLD may specifically regulateLOX2 upstream of the octadecanoid pathway that leads toJA. In addition, the involvement of phospholipase A (PLA),which mediates the release of linolenic acid from cellmembranes and is inducible by wound stimuli, has been alsoproposed for tomato (Lycopersicon esculentum) and otherspecies [30].Altogether, early and secondary cell signalling forherbivore-<strong>induced</strong> <strong>plant</strong> responses comprise: (1) the receptionof an extracellular signal(s) such as high- or lowmolecularweight factors from the herbivore (e.g., FACs),(2) V m depolarization and an intracellular calcium influx, (3)the activation of protein kinase/phosphatase cascades, and(4) the release of linolenic acid from the cell membrane andsubsequent activation of the octadecanoid pathway whichleads finally to the synthesis of JA and other oxylipins.However, multiple signals, signal circulation by, forexample, protein phosphorylation/dephosphorylation, andfeedback regulation of signalling and metabolic pathwaysare likely to complicate the understanding of the accuratemechanism.2.3. Oxylipin, SA, and their antagonismJA and SA function as signalling molecules, whichmediate <strong>induced</strong> <strong>plant</strong> responses toward herbivory andpathogen infection, resulting in the activation of distinctsets of defence genes [31–33]. Besides free JA, also aconjugate with the amino acid isoleucine may be involvedin signalling [34,35], but the significance of suchconjugates still remains to be established. Inhibiting theperformance of JA or SA obviously renders <strong>plant</strong>s moresusceptible to herbivore damage and pathogen infection[36–39]. JA and SA act antagonistically, and are bothrequired for the <strong>induced</strong> response following herbivorefeeding or pathogen attack [40–44]. For instance, thegraduated increase of SA accumulation in Lima beanleaves treated with alamethicin (ALA), a potent fungalelicitor of <strong>plant</strong> volatile emission, interferes with steps inthe biosynthetic pathway downstream of 12-oxophytodienoicacid (OPDA), thereby reducing the rapid accumulationof JA (~40 min) [41] and, finally, the biosynthesisand emission of all JA-linked volatiles. On the other hand,enhanced levels of OPDA selectively induce the emissionof the C 16 -tetranorditerpene 4,8,12-trimethyltrideca-1,3,7,11-tetraene (TMTT) in Lima bean [45]. Consequently,the treatment of Lima bean leaves with variousamounts of JA and SA caused the emission of characteristicblends of volatiles including TMTT, which stronglyattracted the carnivorous natural enemies of spider mites[40,46,47]. Hence, JA and SA pathways may vie with eachother to control and coordinate <strong>induced</strong> <strong>defences</strong>, since theemission of characteristic blends of herbivore-<strong>induced</strong>volatiles from attacked <strong>plant</strong>s strongly depends on thedelicate balance between JA and SA [41]. Alternatively,both pathways may play a role in discriminating insectbiting from mechanical wounding and thus may be a kindof filter that prevents unnecessary defence activation.


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 95Whether or not the recently discovered phytoprostanes,which result from non-enzymatic oxidative transformationof linolenic acid into prostaglandine-type octadecanoids[48], are involved in defence responses against herbivoresremains to be established. At least, gene expression analysisof Arabidopsis cell cultures treated with the phytoprostanesPPB 1 -I or -II revealed that both compounds triggered amassive detoxification and defence response covering theexpression of several glutathione S-transferases, glycosyltransferases and putative ATP-binding cassette transporters[49].2.4. Synergy between JA and ethyleneUndoubtedly, JA is among the most important signallingmolecules of herbivore-<strong>induced</strong> responses. However,the activation of JA formation alone cannot explain allchanges following wounding or herbivory. For example,transgenic potato <strong>plant</strong>s overexpressing an allene oxidesynthase gene, which is involved in the JA formation,failed to express a PI gene (Pin2) constitutively, despitethe fact that transgenic <strong>plant</strong>s exhibited six- to twelve-foldhigher levels of endogenous JA than did non-transgenic<strong>plant</strong>s [50]. This observation indicates that JA is notalways involved in the signalling pathways of wound/herbivore responses and that other mediators may play arole in the respective signalling cascades. As describedabove, SA is an example of a signalling molecule that actsantagonistically towards JA. In addition, other signallingpathways may positively affect JA action. Ethylene is sucha candidate, as it induces the expression of defence genes,and its synthesis is <strong>induced</strong> by wounding, herbivorefeeding, and JA treatment [51,52]. Hints of a synergybetween JA and ethylene in the herbivory- or wound<strong>induced</strong>responses are: (1) <strong>plant</strong> defence genes aresynergistically <strong>induced</strong> by ethylene and JA/MeJA intobacco and in wounded tomato <strong>plant</strong>s [51,53], (2)treatment of maize <strong>plant</strong>s with 1-methylcyclopropene (1-MCP) reduced the production of ethylene and volatileemission following M. sexta feeding [54], and (3) treatmentof maize <strong>plant</strong>s and Lima bean leaves with ethyleneor its precursor 1-aminocyclopropane-1-carboxylic acid(ACC) significantly promoted volatile emission <strong>induced</strong>by JA [55,56]. In addition, simultaneous application of JAand ACC to Lima bean leaves increased the attractivenessof the pre-treated Lima bean leaf for predatory mites (e.g.,Phytoseiulus persimilis) as compared to leaves treated withJA alone [56]. All these results suggest that the cross-talkbetween JA and ethylene is required for both direct and<strong>indirect</strong> defence responses to herbivory. The application ofethylene and ethephon (an ethylene-releasing compound)together with the application of MeJA to tobacco <strong>plant</strong>ssuppressed the induction of nicotine accumulation yethardly affected the emission of bergamotene, a volatilesesquiterpene [57]. Moreover, the application of 1-MCP totobacco <strong>plant</strong>s treated with oral secretions of M. sextalarvae increased the <strong>induced</strong> accumulation of nicotine butreduced fitness-relevant <strong>plant</strong> parameters such as the rateof stalk elongation or lifetime capsule production [58]. Inthis case, the interplay between ethylene and JA mayreduce the level of nicotine present in a <strong>plant</strong>, thusreducing its autotoxicity as well as the physiological costsof this compound. Moreover, ethylene has been reported topositively regulate the induction of allene oxide synthase,an enzyme which catalyzes the production of a precursorof OPDA in the octadecanoid pathway, in Arabidopsis andtomato [51,59,60]. Also, the antagonistic interactionbetween ethylene and JA on defence gene expression,resulting in resistance of Arabidopsis <strong>plant</strong>s to herbivory,has been proposed [61]. These mechanisms are differentamong different <strong>plant</strong> species (e.g., tomato and Arabidopsis)and the other conditions [61,62].Several other signalling molecules may function synergisticallywith JA and ethylene in <strong>plant</strong> defence responses. Forexample, in potato abscisic acid has been shown to increaseJA levels following mechanical damage to leaf tissues [63],whereas abscisic acid did not regulate any defence-relatedgenes by itself [64]. Among other <strong>plant</strong> hormones, auxin is alikely candidate, as it also has a negative effect on the JApathway [65]. Furthermore, spermine is considered as anintegral constituent of wound/herbivory responses, because(1) spermine activates WIPK and SIPK in tobacco leaves andmitochondrial dysfunction via a signalling pathway in whichreactive oxygen species and a calcium influx are involved[66], and (2) the expression of a S-adenosylmethionine(SAM) decarboxylase gene involved in the polyaminesynthesis is <strong>induced</strong> in Lima bean leaves infested withspider mites [52]. Considering that spider mite-infested Limabean leaves do not accumulate endogenous spermine, thispolyamine may, however, play a minor role in wound/herbivory responses. Alternatively, polyamines such asspermine may accumulate locally in intercellular spaces, ashas been shown for tobacco leaves infected with tobaccomosaic virus [67]. More evidence is still required to fullyunderstand the role of polyamines in stress responses.2.5. TranscriptionA subset of cis- and trans-transcription elements, whichare involved in the JA- ethylene- and the SA-signallingpathways, has been characterised. In several <strong>plant</strong> species,the GCC box (AGCCGCC), an ethylene-inducible element,has been found in the promoter region of ethylene-inducibledefence genes [68]. Four and six ethylene-responsivetranscription factors (ERF) that specifically interact withthe GCC box were characterised from tobacco andArabidopsis, respectively [69–71]. ERF genes can respondto various extracellular stimuli, such as wounding, pathogeninfection, and drought stresses, by changing of theirtranscriptional levels and have so far been described forhigher <strong>plant</strong>s but not for yeast and other fungi [68,71]. InArabidopsis, ERF1 and AtERF1, AtERF2, and AtERF5 are


96G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111known as active elements of GCC-associated transcriptions,while AtERF3 and AtERF4 act as suppressive elements ofthese transcriptions [71]. As they bind to the same targetsequence, their transcriptional regulation is thought todepend on either the DNA binding affinity or on the relativeabundance of active ERFs in the nucleus, or on both factors.Moreover, such GCC-associated transcripts have beenshown to be coordinated not only by ethylene, but also byJA [72]. Conversely, JA-inducible transcription factorsORCAs (the octadecanoid-derivative responsive CatharanthusAP2-domain) have been characterised as a JAresponsiveAPETALA2 (AP2)/ERF- domain transcriptionfactor in Catharanthus roseus [73,74]. The overexpressionof Orca3 constitutively increased the expression of genesinvolved in several metabolic pathways (tryptophan decarboxylase,1-deoxy-d-xylulose-5-phosphate synthase anddesacetoxyvindoline 4-hydroxylase) involved in terpenoidand indole alkaloid formations [74]. Recently, a WRKYtranscription factor was shown to bind to a W-box, with thelatter consisting of two reversed TAGC repeats and locatedin the promoter region of the cotton (+)-y-cadinene(sesquiterpene) synthase gene CAD1 [75]. The expressionof this transcription factor was <strong>induced</strong> by JA and an elicitorderived from Verticillium dahliae. Direct/<strong>indirect</strong> synergismsor antagonisms between multiple trans-factors arelikely to regulate genes via signalling mediators such as JA,ethylene, and SA.The recent development of high-throughput microarraytechnology allows the simultaneous and systematic monitoringof the expression pattern of an immense number of<strong>plant</strong> genes. Such analyses have shown comprehensivetranscript profiles in <strong>plant</strong>s responding to insect feeding,mechanical wounding, JA, H 2 O 2 , and <strong>plant</strong> volatiles [76–80]. For instance, the transcription of several genes was<strong>induced</strong> in Arabidopsis following mechanical wounding andPieris rapae feeding [77]. Curiously, one gene encoding ahevein-like protein was <strong>induced</strong> by P. rapae, but not bymechanical wounding. As described above, <strong>induced</strong> <strong>plant</strong>responses to mechanical wounding and feeding herbivoresmay differ due to varying degrees and types of damage andthe presence or absence of salivary compounds.2.6. Systemic responseIt is well known that <strong>plant</strong>s respond to herbivory andmechanical wounding not only at the site of damage, but alsoin remote, undamaged leaves. The best systems studied inthis context are solanaceous species, such as tomato, forwhich an 18-amino-acid peptide called systemin wasdiscovered and claimed to represent an intercellular signallingmolecule [81]. Systemin is derived from a 200-aminoacidprecursor prosystemin that has been originally discoveredin tomato. Transgenic tomato <strong>plant</strong>s that constitutivelyexpress antisense prosystemin RNA reduce wound-<strong>induced</strong>transcript accumulation of PIs in systemic leaves and aremore susceptible to M. sexta attack than are wild-type <strong>plant</strong>s[82]. Furthermore, a 160-kDa plasma membrane-bound,leucine-rich repeat receptor kinase was recently characterisedas a systemin receptor (SR160) from suspension cultures oftomato cells [83]. Surprisingly, this receptor is identical totBRI1, a brassinosteroid receptor in tomato, and homologousto BRI1 in Arabidopsis [84–86]. The dual function of SR160/tBRI1, which remains to be unequivocally established, isvery unique, as this receptor has two ligands (systemin andbrassinosteroid) and apparently functions in developmentand defence responses [86]. Additional analyses demonstratedthat the overexpression of SR160 in suspension-culturedtobacco cells yielded systemin-binding activity and asystemin-<strong>induced</strong> alkalinisation response in the cells. Theseresults indicate that post-receptional steps of the signaltransduction are also present in tobacco, another member ofthe family Solanaceae. In tobacco, two 18-amino-acidhydroxyproline-rich glycoproteins (TobHypSys I and II),which are potential PI-inducers in a similar manner tosystemin, have been discovered [87,88]. These two peptidesare derived from a 165-amino-acid precursor that does notexhibit any homology to prosystemin from tomato, but isanalogous to peptide hormones from animals and yeast.Recently, this type of peptide was also discovered in tomato,raising the question of whether these peptides represent ageneral type of defence signal in the <strong>plant</strong> kingdom [89].Today, systemin is no longer considered as a longdistancesignal. A systemin-insensitive tomato mutant (spr1,suppressor of prosystemin-mediated responses1), deficientin a signalling step that couples systemin perception to thesubsequent activation of the octadecanoid pathway, showedthat the peptide acts at or near the local site of wounding,increasing JA-synthesis above the threshold that is requiredfor the systemic response in tomato [90]. According to theseresults, systemin is now thought to up-regulate the JAbiosynthetic pathway to generate a long-distance signal atthe local site (for example, JA or OPDA as the transmissiblesignals), whose recognition in distal leaves is linked tooctadecanoid signalling. Moreover, it was also found thatthe rapid and transient activation of early genes in responseto leaf excision or wounding was not affected in spr1 <strong>plant</strong>s,indicating the existence of a Spr1- and systemin-independentpathway for wound signalling [90].<strong>Herbivore</strong>-<strong>induced</strong> systemic responses can be observednot only in solanaceous species but also in many other<strong>plant</strong> families. Poplar trees (Populus trichocarpa x deltoides)are able to trigger the activation of gene expressionssystemically in undamaged leaves far from those that hadbeen damaged by the forest tent caterpillar (Malacosomadisstria) [91]. This systemic response revealed acropetal(upper) but not basipetal (lower) direction, but the transportmechanism remains obscure. A putative apoplastic lipidtransfer protein DIR1 has been suggested as a mobilesignal, mediating an interaction between a leaf of Arabidopsisinfected with Pseudomonas syringae and distant,uninfected leaves [92]. It was speculated that DIR1 couldbe a co-signal with other lipids, such as oxylipins (JA),


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 97phosphatidic acid, and N-acylethanolamines, which travelthrough the vascular system of the <strong>plant</strong>. However, actualevidence for a possible involvement of DIR1 in anherbivore-<strong>induced</strong> defence is still lacking. In addition,electrical signals have been noted and claimed to mediatelong-distance interactions in wounded tomato seedlings[93]. Plants may have multiple systems that enable accuratelong-distance signalling. Thus, JA itself can act as asystemic signal in tobacco, which is formed in the woundedleaves and travels to the undamaged distal leaves and rootswhere the expression of PI and the nicotine biosynthesis are<strong>induced</strong>, respectively [94,95].3. Biosynthesis of herbivore-<strong>induced</strong> <strong>plant</strong> volatiles(HIPVs)3.1. Volatile terpenoidsVolatile terpenoids which can be <strong>induced</strong> by herbivorefeedingcomprise monoterpenes (C 10 ), sesquiterpenes (C 15 )and homoterpenes (C 11 or C 16 ). All terpenoids are synthesisedthrough the condensation of isopentenyl diphosphateand its allylic isomer dimethylallyl diphosphate by catalysisof farnesyl diphosphate (FPP) synthase via the mevalonatepathway (cytosol/endoplasmic reticulum) or geranyl diphosphate(GPP) and geranylgeranyl diphosphate via the methyld-erythritol-1-phosphatepathway in plastids [96,97] (Fig.3). A large, structurally diverse number of terpenoids areyielded by a large family of terpene synthases (TPS) usingGPP and FPP as substrates. In Arabidopsis, 32 genesincluding two gibberellin biosynthetic genes are putativemembers of the TPS family [98], some function as mono-TPSs and sesqui-TPSs [99–102]. Terpenoid formation isgenerally assumed to be regulated on the transcript level ofthe TPS genes [91,103–105].However, the regulation mechanism seems to be rathercomplex, because herbivore-<strong>induced</strong> TPS transcripts andterpene emissions are affected by several factors (forexample, by diurnal rhythmicity and distance to herbivoredamagedtissue) [91,106]. Fig. 4 shows temporal patterns ofvolatile emissions in Lima bean leaves following herbivoreattack by S. littoralis over 4 days. The release of terpenoidsand the C 6 -volatile (3Z)-hex-3-enyl acetate follows diurnalcycles with increased emissions during the light period andreduced emissions during darkness. This result is in linewith findings in Lima beans treated with ALA and poplarleaves infested with forest tent caterpillars, where thevolatile emissions or the TPS expressions follows diurnalcycles [91,107]. In this context it would be interesting tostudy to what extent volatile emissions are linked to theendogenous biological clock.On the other hand, a single event of mechanical damageor the application of ALA to Lotus japonicus <strong>plant</strong>s was notFig. 3. Biosynthetic pathways required for herbivore-<strong>induced</strong> <strong>plant</strong> volatiles. Elements in bold are enzymes. Abbreviations: ALDH, aldehyde dehydrogenase;DMAPP, dimethylallyl diphosphate; DMNT, (E)-4,8-dimethyl-1,3,7-nonatriene; DXP, 1-deoxy-d-xylulose-5-phosphate; DXR, DXP reductoisomerase; DXS,DXP synthase; FPP, farnesyl diphosphate; FPS, FPP synthase; GGPP, geranylgeranyl diphosphate; GGPS, GGPP synthase; GPP; geranyl diphosphate; GPS,GPP synthase; HMG-CoA, 3-hydroxy- 3-methyl-glutaryl CoA; HMGR, HMG-CoA reductase; HPL, fatty acid hydroperoxide lyase; IDI, IPP isomerase; IGL,indole-3-glycerol phosphate lyase; IPP, isopentenyl diphosphate; JA, jasmonic acid; JMT, JA carboxyl methyl transferase; LOX, lipoxygenase; MeJA, methyljasmonate; MEP, 2-C-methyl-d-erythritol-4-phosphate; TMTT, 4,8,12-trimethyltrideca-1,3,7,11-tetraene; TPS, terpene synthase.


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 99defence agents or signalling molecules which elicit defencereactions.However, C 6 -volatiles may not always benefit <strong>plant</strong>s.(2E)-Hex-2-enal inhibits the germination and subsequentgrowth of soybeans as well as the germination of applepollen [124,125]. Furthermore, (2E)-hex-2-enal reduces thegermination frequency of the MeJA-insensitive Arabidopsismutant jar1-1, suggesting that (2E)-hex-2-enal and MeJAare recognized in <strong>plant</strong>s by different mechanisms [119]. Arapid formation of C 6 -volatiles after wounding not onlyserves as protection against herbivores or pathogens, butmay also be toxic for the <strong>plant</strong> itself. This effect could beinterpreted as some kind of strategic retreat, as has beenobserved for hypersensitive cell death in response topathogen infection. It remains to be investigated whetherthe transient burst of wound-<strong>induced</strong> C 6 -volatiles isdetrimental for the growth and development of a <strong>plant</strong> inthe long term.In contrast to C 6 -aldehydes and -alcohols, the emissionof (3Z)-hex-3-enyl acetate can frequently be observed a fewhours after the first impact of herbivore feeding ormechanical damage [113,126,127], as seen in the emissionof herbivore-<strong>induced</strong> terpenoids (Fig. 2). Since the emissionof both (3Z)-hex-3-enyl acetate and h-ocimene can be<strong>induced</strong> by JA and ethylene [56], the signalling pathwaymay be identical. However, nothing is known about whetheror not the biological function of hexenyl acetate differs fromthe two early C 6 -volatiles.3.3. Volatile phytohormones and indoleLike JA, MeJA is also considered one of the mostimportant signalling molecules of herbivore-<strong>induced</strong> <strong>plant</strong>defence. An early study showed that synthesis of PIproteins in the leaves of solanaceous species is <strong>induced</strong> byexogenous application of MeJA or natural emission ofMeJA from sagebrush <strong>plant</strong>s (Artemisia tridentata) [128].Transgenic Arabidopsis overexpressing JA carboxylmethyl transferase (JMT), an enzyme that methylates JAto form MeJA, showed a three-fold increase of theendogenous MeJA level without altering JA levels [129].Furthermore, transgenic <strong>plant</strong>s exhibited constitutive transcriptaccumulation of JA-responsible genes and enhancedresistance against the virulent fungus Botrytis cinere [129].These results indicate the importance of MeJA as a cellularsignal regulator and potential inter-cellular, and intra- andinter-<strong>plant</strong> signal transducers. The other JA derivative, (Z)-jasmone, has been found to repel the lettuce aphid(Nasonovia ribisnigri) and to induce (E)-h-ocimene,resulting in an increased attractivity of the <strong>plant</strong>s to theaphid parasitoid Aphidius ervi [130]. However, it has beenalso reported that MeJA and cis-jasmone are onlymoderately <strong>induced</strong> in Nicotiana attenuata <strong>plant</strong>s inresponse to feeding herbivores (9% of the JA accumulation)[131]. Functionally, these JA derivatives may be ofminor importance for defence and signalling. Otherwise,the importance may be limited in a few <strong>plant</strong> species suchas A. tridentata.The Arabidopsis genome contains 24 genes, whichbelong to a structurally related group of methyl transferases[132]. These have been classified as one gene family termedAtSABATH, of which JMT is a member. AtBSMT1,another member of this family, has recently been characterisedas SA carboxyl methyl transferase, an enzyme thatmethylates SA to form MeSA [133]. The transcriptaccumulation of AtBSMT1 can be <strong>induced</strong> by any of thefollowing: the fungal elicitor ALA; feeding Plutellaxylostella; uprooting; mechanical wounding; and exogenousapplication of MeJA to Arabidopsis leaves. In some of thesecases, the emission of MeSA is also <strong>induced</strong>. Besides itsrole as HIPV in attracting carnivores, such as the mites P.persimilis and Amblyseius potentillae [47], MeSA can alsoinduce defence responses in <strong>plant</strong>s [134].Occasionally indole has been observed as a minorconstituent of herbivore-<strong>induced</strong> volatile blends. In maize<strong>plant</strong>s, a gene has been characterised as indole-3-glycerolphosphate (IGP) lyase, which catalyses the formation of freeindole from IGP [135]. Both volicitin and MeJA wereshown as selective inducers of the gene expression and theindole emission from maize <strong>plant</strong>s [135,136].4. Ecology of <strong>induced</strong> <strong>indirect</strong> <strong>defences</strong>4.1. Variability of herbivore-<strong>induced</strong> <strong>plant</strong> volatilesThe emission of volatile organic compounds seems to bevery common across the <strong>plant</strong> kingdom and is not limited tocertain life forms (see [137] for a compilation of species).Odour blends emitted by herbivore-infested <strong>plant</strong>s arecomplex mixtures, often composed of more than 100different compounds, many of which occur as minorconstituents [138]. Different <strong>plant</strong> species vary in theheadspace composition [139,140] yet also share compounds,for example, (E)-h-ocimene, (3Z)-hex-3-enyl acetate, andDMNT in Lima beans and cucumber [141]. Even within aspecies, the volatile blends emitted upon herbivore damagediffer both quantitatively and qualitatively depending on thegenotype or cultivar [142–144], the leaf developmental stage[145], or the <strong>plant</strong> tissue attacked by an herbivore [139], aswell as on abiotic conditions (light intensity, time of year,water stress) [146]. Beyond that, the time of the day alsoinfluences the composition of the emitted volatile blend (Fig.4): for example, Nicotiana tabacum releases several HIPVsexclusively at night. These nocturnally emitted compoundswere repellent to female moths (Heliothis virescens), whichsearch for oviposition sites during the night [147]. Moreover,the blend composition strongly depends on the type ofdamage (e.g., herbivore feeding versus oviposition [148])inflicted upon a given <strong>plant</strong> individual as well as on the timeelapsed after leaf damage (Fig. 4). To all the variability thatalready exists on the <strong>plant</strong> level, several studies have added


100G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111that the emitted blends also vary with the herbivore species[149], and even different ontogenetic stages of the sameherbivore [150,151] may influence the headspace composition.However, differences between volatile blends seem tobe largest among different <strong>plant</strong> species and smallest among<strong>plant</strong>s of one species infested by different herbivores[141,149].The high variability that characterises the chemicalcomposition of emitted blends raises the question of theecological relevance of these differences and whetherarthropods can discriminate among these complex mixtures.Electroantennogram (EAG) analyses allow individual blendcomponents to be differentiated in order to determine whichcompounds or combination of different compounds can beperceived by an arthropod [152,153]. Further informationon the physiological perception of different <strong>plant</strong> odours byarthropods have been obtained from in vivo calciumimagingmeasurements that provide additional informationon how the olfactory information is processed in thearthropod brain [154,155]. However, only the additionalinclusion of behavioural assays, in which the arthropods aresimultaneously exposed to two or more odours amongwhich they have to choose, provides information on whetheran odour has an attracting, repelling, or neutral effect on thebehaviour of the focal arthropod.4.2. Specificity and behavioural responses to HIPVsMuch has been learned from such behavioural assays onthe ability of arthropods to discriminate different odourblends. As the production of HIPVs has been originallydescribed in the context of host-location cues for parasiticwasps [5], most researchers have focussed on the role of<strong>plant</strong>-derived volatiles in determining the ability of differentparasitoid or carnivore species to detect their herbivoroushosts or preys. Several studies of an increasing number oftritrophic <strong>plant</strong>–herbivore–carnivore systems have indicatedthat the ability of the carnivores and parasitoids todiscriminate different odour blends depends very much ontheir degree of dietary specialization [156], and on theirlevel of deprivation, as well as on their previous experience[149]. The particular behavioural response – that is whetheran herbivore or carnivore is attracted or repelled, or reactsneutrally to a specific blend of HIPVs – seems to dependstrongly on the level of <strong>plant</strong> induction [157–159].However, <strong>plant</strong>s do not only interact with otherorganisms above ground. Recently, evidence has shownfor the first time that mechanisms similar to the onesdescribed above- may also work below-ground: Entomopathogenicnematodes (Heterohabditis megidis) are attractedas yet unidentified chemicals that are released from roots ofa coniferous <strong>plant</strong> (Thuja occidentalis) when weevil larvae(Otiorhynchus sulcatus) attack [160]. Moreover, systemicallyreleased HIPVs have been shown to attract thespecialist parasitoids of root-feeding larvae [161]. Althoughcurrently multitrophic below-ground interactions and thelinks between multitrophic above- and below-groundinteractions are poorly understood, future research willundoubtedly address these questions [162].4.3. Ecological function of HIPVs in natureMost studies regarding HIPV emission have beenconducted under artificial greenhouse or laboratory conditions.Facing the above-mentioned variability, the questionarises whether <strong>plant</strong>s benefit from the emission of HIPVsunder natural growing conditions. To date, few studies existthat show predators are attracted to herbivore-infested <strong>plant</strong>sunder field conditions [153,163–165]. Moreover, specialistparasitoid wasps were able to distinguish among maize,cotton, and tobacco <strong>plant</strong>s that were infested by theirherbivorous hosts from those which were under non-hostattack [166]. Caterpillars (S. exigua) placed on tomato<strong>plant</strong>s (L. esculentum var. Ace) that had been grown in anagricultural system and <strong>induced</strong> previously with exogenouslyapplied JA, suffered from higher parasitism rates byan endoparasitic wasp (Hyposoter exiguae) than control<strong>plant</strong>s [167]. A further field study of native tobacco (N.attenuata) demonstrated that, indeed, the release of volatilesincreased the predation rates of tobacco hornworm eggs by ageneralist predator (Geocoris pallens) [168]. Recently, Limabean <strong>plant</strong>s grown in their natural environment, followed byrepeated treatments with JA solutions for 5 weeks wereshown to respond with an increased seed set [169]. Sinceboth the emission of HIPV as well as the secretion ofextrafloral nectar had been upregulated by this treatment, itremained unclear which <strong>indirect</strong> defence was responsible forthe observed beneficial effect. More field trials are clearlynecessary to determine the ecological relevance of individualfactors and to verify results that have been gathered inlaboratory or greenhouse experiments.4.4. Extrafloral nectarIn addition to releasing <strong>induced</strong> volatiles, some <strong>plant</strong>species have developed other ways to attract predatoryarthropods that will defend them. These include both theprovision of shelter in so-called leaf domatia [170], and theoffer of attractive food sources such as food bodies orextrafloral nectar (EFN). Extrafloral (EF) nectaries arespecialized nectar-secreting organs that may occur onvirtually all vegetative and reproductive <strong>plant</strong> structures,yet do not involve pollination [171]. They have beendescribed for approximately a thousand <strong>plant</strong> speciesranging over 93 <strong>plant</strong> families [172] of flowering <strong>plant</strong>sand ferns, but they are absent in gymnosperms [171]. EFNsare aqueous solutions that comprise mainly sucrose,glucose, and fructose, but other sugars, amino acids andother organic compounds may be present in some species[173]. The secreted sugars are mainly derived from thephloem [174,175] or synthesised in the region of the nectary[176]. EF nectaries secrete small amounts of nectar


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 101throughout the day. The secretion of nectar can followcircadian rhythms [177–180] or be relatively constantthroughout day and night [181,182].The mechanisms of nectar secretion are poorly understood:Some researchers have described secretion as apassive process, whereas others discuss it as an eliminationof surplus sugars [173]. However, there is evidence thatnectar secretion is an active secretory process: It requires theexpenditure of metabolic energy [183], and EF nectariesstrongly resemble secretory cells which contain largenumbers of mitochondria and show dense protoplasts andlarge nuclei [173]. Furthermore, the secretion of EFN can be<strong>induced</strong> by JA treatment. Originally discovered in Macarangatanarius [184], nectar secretion is suggested to beinducible by leaf damage. This seems to be a widespreadphenomenon [169,185–188], which may be linked to thesame or similar signalling pathways as those discussed forthe induction of volatile biosynthesis. Inhibitors thatsuppress the release of linolenic acid or interfere with theproduction of linolenic acid hydroperoxides completelysuppress the damage-<strong>induced</strong> flow of EFN and, hence,clearly demonstrate the involvement of oxylipin-basedsignalling in EFN-induction. Interestingly, also the attackof a below-ground herbivore (Agriotes lineatus) on cotton<strong>plant</strong>s (Gossypium herbaceum) <strong>induced</strong> above-groundproduction of EFN [185]. InVicia faba, even the numberof EF nectaries increased following leaf damage [189].Generally, the secretion of EFN is believed to function asan <strong>indirect</strong> anti-herbivore defence by attracting and henceincreasing the presence of putative <strong>plant</strong> defenders onnectar-secreting <strong>plant</strong> parts. Most studies focussed on theprotective effect of ants on entire <strong>plant</strong>s or individual <strong>plant</strong>parts, demonstrating a beneficial function of these insects[172,173]. However, contradictory observations thatdetected no measurable preventive effect of EFN-attractedants also exist [190–194]. In these cases, the lack ofprotection could be explained by, for example, (1) differencesin the aggressiveness of the attracted ant species[194–196], (2) differences in foraging behaviour of antspecies from different habitats [197,198], and (3) a varyingsusceptibility of the herbivores to ant predation [199,200].Besides ants, the EF nectaries attract a diverse spectrum ofother arthropods including Araneae, Diptera, Coleopteraand Hymenoptera (see [172] for review). Due to theirpredatory or parasitoid ways of life, many of these non-ants,such as ichneumonid, braconid and chalcidoid wasps [201–203], jumping spiders (Salticidae) [204], phytoseiid mites[205] or tachinid flies [206], can as well reduce the numberof herbivores. Both ants and wasps exerted beneficial effectson the EF nectary-bearing <strong>plant</strong> Turnera ulmifolia whenselectively excluded, yet no additive increase was observedwhen both insect groups had access [207].The emission of floral odours by <strong>plant</strong>s is very importantfor attracting pollinators to their floral nectars. Some ofthese floral nectars also scent by themselves [208], whereasnothing is known about the headspace of EFNs. Beyondthese odours, which communicate the location, abundance,and quality of these nectars to higher trophic levels, othermechanisms may help to guide EFN-feeding arthropods toremote nectar sources. First, some EF nectaries are colouredproviding visual cues for foraging arthropods [172]. Additionally,increased amounts of both HIPVs and EFN(Table 1) would allow foraging arthropods to use theemitted volatile organic compounds as a cue to detect nectarsources from longer distances. Future research should try todisentangle and quantify the costs and benefits of one ofthese <strong>indirect</strong> <strong>defences</strong> for a <strong>plant</strong> and figure out to whatextent each of them contributes to <strong>plant</strong> defence in nature.4.5. Evolution of HIPVs and EFNThe evolutionary origin of HIPV emission and EFNsecretion remains unknown. As described above, C 6 -volatiles or terpenoids may repel herbivores [123,209], betoxic for phytophages [210,211], or act anti-microbially[212]. HIPVs are therefore discussed as having originallyTable 1Compilation of species that are known to both emit HIPVs and feature extrafloral nectariesFamily Species Nectar inducible HIPVs inducible ReferencesEuphorbiaceae Manihot esculenta n.i. n.i. + H 1 [269,270]Leguminosae Phaseolus lunatus + JA + JA [169]Vicia faba n.i. n.i. + H 2 [271,272]Vignia unguiculata n.i. n.i. + H 3 [273,274]Malvaceae Gossypium herbaceum + M, H 4, 5 + H 5, 6 [185,186,275]Gossypium hirsutum n.i. n.i. + H 6, 7, 8 [202,276–278]Rosaceae Prunus serotina n.i. n.i. + H 9 [256,279]Salicaceae Populus deltoides n.i. n.i. + H 10 [91,280]Populus trichocarpa n.i. n.i. + H 10 [91,281]Solanaceae Solanum nigrum n.i. n.i. + H 11, 12 [282,283]Additional information on the inducibility of the respective trait is presented (for abbreviations see below).H 1, spider mite (Mononychellus tanajoa Bondar); H 2, pentatomid bug (Nezara viridula Linnaeus); H 3, spider mite (Tetranychus urticae Koch); H 4,Mediterranean brocade (Spodoptera littoralis Boisduval); H 5, wireworms (Agriotes lineatus Linnaeus); H 6, beet armyworm (Spodoptera exigua Hqbner); H7, tobacco budworm (Heliothis virescens Fabricius); H 8, corn earworm (Helicoverpa zea Boddie); H 9, forest cockchafe (Melolontha hippocastani Fabricius);H 10, forest tent caterpillar (Malacosoma disstria Hqbner); H 11, Colorado potato beetle (Leptinotarsa decemlineata Say); H 12, death’s head hawkmoth(Acherontia atropos Linnaeus); JA, jasmonic acid; M, mechanical damage; n.i., not investigated.


102G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111functioned as a direct defence against pathogens orherbivores [213]. However, in this case, non-volatiledefensive compounds rather than volatile substances shouldhave evolved to fulfil this purpose, as they would not diffuseso easily in the environment [214]. HIPV emission is anactive process that does not necessarily depend on celldamage at the site of emission [215]. Even if the firstvolatile release by an herbivore-damaged <strong>plant</strong> was incidental,natural enemies might have been immediately able touse this cue to locate the responsible herbivores; hence,natural selection should operate on the <strong>plant</strong> to optimizethese signals [214].The evolution of EF nectaries is considered a very simpleprocess [173] that may have occurred several timesindependently [171,173]. This development has resulted inan extreme variety of <strong>plant</strong> taxa that feature a wide structuraldiversity of EF nectaries [216]; yet these structures may beabsent in one of two very closely related species [173].Recently, a 2-year study has shown that traits of EFnectaries were heritable and that casual ant associates actedas agents of selection on these traits [217]. Even ifcorresponding studies on other EF nectary-bearing <strong>plant</strong>sor <strong>plant</strong>s that feature HIPVs are missing, these two <strong>indirect</strong><strong>defences</strong> seem to be evolutionarily plastic and to remainstable due to strong selection pressure. However, the geneticbasis of these two traits has been investigated only for thedevelopment of EF nectaries in cultivated cotton [218,219]and Acacia koa [220]: For cotton crossing, studies showedthat the inheritance of bnectarylessQ was transferred by twopairs of recessive genes (i.e., ne-1 and ne-2) [218,219].In addition to the benefits discussed above, the conceptof fitness costs provides a powerful explanation for theevolution of <strong>induced</strong> <strong>plant</strong>s <strong>defences</strong> [221,222]: Costscover any trade-off between resistance and other fitnessrelevantprocesses [221]. Inducibility itself (i.e., the abilityto change per se) can raise fitness costs by, for example,increasing the lag time until the defence is up-regulated,during which the <strong>plant</strong>s remains susceptible to herbivores[223,224]. Further evolutionarily relevant costs compriseallocation costs [225], genetic costs [226,227], autotoxicitycosts [228], opportunity costs [229], and ecological costs(Fig. 5) [230,231]. Inducible defence strategies are generallyFig. 5. Potential interactions of <strong>plant</strong>s with higher trophic levels via herbivore-<strong>induced</strong> <strong>plant</strong> volatiles (HIPVs) and extrafloral nectar (EFN). The stipules at thepetioles of the leaves represent extrafloral nectaries. Effects mediated by HIPVs or EFN can be attraction (+) or deterrence ( ) that consequently affect trophicinteractions such as herbivory or predation (Arrows). Additionally, the ecological fitness costs (C) and benefits (B) of each interaction from the <strong>plant</strong>’sperspective are indicated.


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 103assumed to reduce the fitness of a <strong>plant</strong> when expressedunder enemy-free conditions [232] as compared to costs of aconstitutive (i.e., permanent) expression. The physiologicalexpenditures of EFN production that could cause allocationtrade-offs within the <strong>plant</strong> are assumed to be very low, yetthis has rarely been tested: For the Balsa tree Ochromapyramidale, the physiological expenses of EFN secretionhave been calculated to account for 1% of the total energyinvestment per leaf [233]. In contrast, a phylogeneticalanalysis of American Acacia species revealed that constitutiveEFN secretion was found to be the derived state whichdeveloped as an adaptation to obligate myrmecophytism.This may indicate that a constitutive flow of EFN is costlyand only profitable when rewarded by permanent antprotection[187]. The actual costs probably depend on thelevel of EFN excretion and physiological limitations such asthe availability of water or nutrients. EFF secretion evenceases in the absence of nectar feeders, yet this may be apassive process and not necessarily a <strong>plant</strong>-regulated strategyto save costs [179]. The physiological expenditures of HIPVare expected to be relatively high [234–236], yet predictionsremain controversial [237–239] and experimental evidenceis rare. In maize <strong>plant</strong>s, the first hints of allocation costs forthe production of HIPVs have been seen [211]. However, towhat extent simultaneously <strong>induced</strong> direct <strong>defences</strong>accounted for the measured costs remained unknown.Interestingly, these allocation costs were detectable only inyoung <strong>plant</strong>s, with compensation for their metabolic investmentoccurring during maturation. Finite resource availabilitiesfor <strong>plant</strong>s may also lead to trade-offs between direct and<strong>indirect</strong> <strong>defences</strong> [240,241]; these seem to be more commonin obligate ant–<strong>plant</strong> mutualisms [242–244] than in facultativeassociations [241].Both EFN and HIPVs are openly presented cues that maynon-specifically attract members of higher trophic levels.Such loose forms of facultative mutualisms are especiallyprone to exploitation by undesired arthropods. For example,the attraction of non-beneficial or even herbivorous speciesfeeding on EFN results in costs rather than benefits for therespective <strong>plant</strong> (Fig. 5) [245–247]. HIPVs emitted fromstrongly <strong>induced</strong> <strong>plant</strong>s denote not only parasitoids orpredators, the presence of potential prey organisms but,additionally, a hint at increased numbers of competitors[248–250]. Furthermore, by calling on strongly <strong>induced</strong><strong>plant</strong>s, carnivores or parasitoids may well run the risk offalling prey themselves to a hyperparasite or a predator[251]. The resulting deterrent effect of HIPVs on beneficialarthropods represents an ecologically relevant fitness costfor the herbivore-attacked <strong>plant</strong> (Fig. 5).Signals emitted from infested <strong>plant</strong>s can also provideinformation to other trophic levels (Fig. 5): Conspecificherbivores, for example, can be attracted to infested <strong>plant</strong>s[157,252,253], because this cue denotes the location ofpalatable host <strong>plant</strong>s [254] and suitable oviposition sites[255] as well as the presence of potential mating partners[256]. Such a dmiscueT from the <strong>plant</strong>’s perspective resultsas well in ecological fitness costs for the <strong>plant</strong>. On the otherhand, conspecific or heterospecific herbivores may bedeterred by the emitted volatiles [147,252], as strongly<strong>induced</strong> <strong>plant</strong>s may point to competition with conspecific orheterospecific herbivores or to an enemy-dense patch [257].In this case, the release of HIPVs would act as a directdefence [147,215], thereby benefiting the <strong>plant</strong>. Moreover,hyperparasitoides or carnivores of the fourth trophic levelcould use <strong>plant</strong>-derived volatiles for the search on largerspatial scales to find their prey or host by locating potentialhost <strong>plant</strong>s [258]. Such attraction would again impose therespective <strong>plant</strong> with ecological costs (Fig. 5).Several <strong>plant</strong> species have the potential to defendthemselves <strong>indirect</strong>ly using HIPV emission and EFNsecretion; both <strong>indirect</strong> <strong>defences</strong> are inducible via theoctadecanoid pathway (Table 1). Comparing the widespreaddistribution of EF nectaries across the <strong>plant</strong> kingdom [172]with the small number of nearly exclusively crop species forwhich HIPVs have been studied [213,259], an increasingnumber of <strong>plant</strong> species that features both <strong>indirect</strong> <strong>defences</strong>can be expected. Beyond that, a wide taxonomicaldistribution of these <strong>plant</strong> traits may indicate the ecologicalsuccess of these two <strong>induced</strong>, <strong>indirect</strong> <strong>defences</strong>.In the ecological arms race between <strong>plant</strong>s and herbivores,the latter are known for rapidly adapting to novelhost <strong>plant</strong>s or toxins [260–262] by evolving specificmechanisms to circumvent direct <strong>defences</strong> with physiologicaltolerance [263,264] or behavioural responses [265,266].In view of the diverse spectrum of carnivorous specieswhich is attracted to <strong>plant</strong>s expressing <strong>indirect</strong> <strong>defences</strong>, itis hard to imagine any mechanism or strategy thatherbivores might evolve to cope with these highly efficient<strong>plant</strong> defenders.5. ConclusionsWe described what is currently known about molecularmechanisms of cell signalling and signal travelling, as wellas the ecology and evolution of herbivore-<strong>induced</strong> <strong>plant</strong>responses. Despite the effort that has been brought to bearon this topic, the exact mechanisms as well as the ecologicaland evolutionary relevance still remain elusive. One reasonis the diversity that exists on the genetic, biochemical,physiological, and phenotypical levels and that differs withthe evolutionary background of the studied <strong>plant</strong> species. Inecological studies, even the geographical location with itsparticular biotic and abiotic conditions influences theoutcome of a field experiment. We therefore suggestadopting model systems beyond the most frequently usedones, namely tobacco, Arabidopsis, maize, and Lima beanto distinguish in a more comparative way particularphenomena from general mechanisms. Such a biodiversity-orientedapproach should include more non-cropspecies since in this case a stronger co-evolution withnatural herbivores is to be expected.


104G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111Connectivity exists between multiple <strong>defences</strong>, as theyare often regulated via the same signalling cascade[267,268], leading to considerable methodological problemswhen individual defensive <strong>plant</strong> traits are studied. Forexample, the clear distinction between pure pathogen andherbivore defence pathways was undermined by the findingthat the type of damage inflicted by an herbivore (chewingcaterpillars versus sucking aphids) determines whichdefence pathway (SA versus JA) is activated. Furtheridentification of specific elicitors and the incorporation ofresponse mutants and transgenic <strong>plant</strong>s will contribute toseparate the different defence responses. These newapproaches have to be complemented by creative fieldexperiments and comparative approaches to gain betterinsights into emerging areas like, e.g., the macroecology orcommunity ecology of <strong>induced</strong>, <strong>indirect</strong> <strong>defences</strong>. Finally,the interactions of <strong>plant</strong>s with microorganisms inside theherbivores’ gut or with micro- or macrobiota below-groundare promising fields of future research.References[1] P.W. Price, C.E. Bouton, P. Gross, B.A. McPheron, J.N. Thompson,A.E. Weis, Interactions among 3 trophic levels — influence of <strong>plant</strong>son interactions between insect herbivores and natural enemies, Ann.Rev. Ecolog. Syst. 11 (1980) 41–65.[2] I.T. Baldwin, The alkaloidal responses of wild tobacco to real andsimulated herbivory, Oecologia 77 (1988) 378–381.[3] P.W. Paré, J.H. Tumlinson, De novo biosynthesis of volatiles <strong>induced</strong>by insect herbivory in cotton <strong>plant</strong>s, Plant Physiol. 114 (1997)1161–1167.[4] W. Boland, Z. Feng, J. Donath, A. G7bler, Are acyclic C 11 and C 16homoterpenes <strong>plant</strong> volatiles indicating herbivory? Naturwissenschaften79 (1992) 368–371.[5] T.C.J. Turlings, J.H. Tumlinson, W.J. Lewis, Exploitation ofherbivore-<strong>induced</strong> <strong>plant</strong> odors by host-seeking parasitic wasps,Science 250 (1990) 1251–1253.[6] L. Mattiacci, M. Dicke, M.A. Posthumus, h-Glucosidase: an elicitorof herbivore-<strong>induced</strong> <strong>plant</strong> odor that attracts host-searching parasiticwasps, Proc. Natl. Acad. Sci. U. S. A. 92 (1995) 2036–2040.[7] R.O. Musser, S.M. Hum-Musser, H. Eichenseer, M. Peiffer, G.Ervin, J.B. Murphy, G.W. Felton, Herbivory: caterpillar saliva beats<strong>plant</strong> <strong>defences</strong>, Nature 416 (2002) 599–600.[8] H.T. Alborn, T.C.J. Turlings, T.H. Jones, G. Stenhagen, J.H.Loughrin, J.H. Tumlinson, An elicitor of <strong>plant</strong> volatiles from beetarmyworm oral secretion, Science 276 (1997) 945–949.[9] R. Halitschke, U. Schittko, G. Pohnert, W. Boland, I.T. Baldwin,Molecular interactions between the specialist herbivore Manducasexta (Lepidoptera, Sphingidae) and its natural host Nicotianaattenuata: III. Fatty acid-amino acid conjugates in herbivore oralsecretions are necessary and sufficient for herbivore-specific <strong>plant</strong>responses, Plant Physiol. 125 (2001) 711–717.[10] D. Spiteller, W. Boland, N-(17-Acyloxy-acyl)-glutamines: novelsurfactants from oral secretions of lepidopteran larvae, J. Org. Chem.68 (2003) 8743–8749.[11] G. Pohnert, V. Jung, E. Haukioja, K. Lempa, W. Boland, New fattyacid amides from regurgitant of lepidopteran (Noctuidae, Geometridae)caterpillars, Tetrahedron 55 (1999) 11275–11280.[12] C.L. Truitt, H.X. Wei, P.W. Paré, A plasma membrane protein fromZea mays binds with the herbivore elicitor volicitin, Plant Cell 16(2004) 523–532.[13] D. Spiteller, G. Pohnert, W. Boland, Absolute configuration ofvolicitin, an elicitor of <strong>plant</strong> volatile biosynthesis from lepidopteranlarvae, Tetrahedron Lett. 42 (2001) 1483–1485.[14] M.L. Orozco-Cárdenas, J. Narváez-Vásquez, C.A. Ryan, Hydrogenperoxide acts as a second messenger for the induction of defensegenes in tomato <strong>plant</strong>s in response to wounding, systemin, andmethyl jasmonate, Plant Cell 13 (2001) 179–191.[15] M. Maffei, S. Bossi, D. Spiteller, A. Mithöfer, W. Boland, Effects offeeding Spodoptera littoralis on lima bean leaves: I. Membranepotentials, intracellular calcium variations, oral secretions, andregurgitate components, Plant Physiol. 134 (2004) 1752–1762.[16] A. Mithöfer, G. Wanner, W. Boland, Effects of feeding Spodopteralittoralis on Lima Bean leaves: II. Continuous mechanical woundingresembling insect feeding is sufficient to elicit herbivory-relatedvolatile emission, Plant Physiol. 137 (2005) 1160–1168.[17] S. Seo, M. Okamoto, H. Seto, K. Ishizuka, H. Sano, Y. Ohashi,Tobacco MAP kinase: a possible mediator in wound signaltransduction pathways, Science 270 (1995) 1988–1992.[18] H. Kodama, T. Nishiuchi, S. Seo, Y. Ohashi, K. Iba, Possibleinvolvement of protein phosphorylation in the wound-responsiveexpression of Arabidopsis plastid N-3 fatty acid desaturase gene,Plant Sci. 155 (2000) 153–160.[19] Y. Liu, H. Jin, K.Y. Yang, C.Y. Kim, B. Baker, S. Zhang, Interactionbetween two mitogen-activated protein kinases during tobaccodefense signaling, Plant J. 34 (2003) 149–160.[20] H. Yamakawa, S. Katou, S. Seo, I. Mitsuhara, H. Kamada, Y.Ohashi, Plant MAPK phosphatase interacts with calmodulins, J.Biol. Chem. 279 (2004) 928–936.[21] S. Zhang, D.F. Klessig, The tobacco wounding-activated mitogenactivatedprotein kinase is encoded by SIPK, Proc. Natl. Acad. Sci.U. S. A. 95 (1998) 7225–7230.[22] S. Seo, H. Sano, Y. Ohashi, Jasmonate-based wound signal transductionrequires activation of WIPK, a tobacco mitogen-activatedprotein kinase, Plant Cell 11 (1999) 289–298.[23] C.Y. Kim, Y. Liu, E.T. Thorne, H. Yang, H. Fukushige, W.Gassmann, D. Hildebrand, R.E. Sharp, S. Zhang, Activation ofa stress-responsive mitogen-activated protein kinase cascade inducesthe biosynthesis of ethylene in <strong>plant</strong>s, Plant Cell 15 (2003)2707–2718.[24] A.A. Ludwig, T. Romeis, J.D. Jones, CDPK-mediated signallingpathways: specificity and cross-talk, J. Exp. Bot. 55 (2004) 181–188.[25] A.C. Harmon, M. Gribskov, J.F. Harper, CDPKs — a kinase forevery Ca 2+ signal? Trends Plant Sci. 5 (2000) 154–159.[26] S.H. Cheng, M.R. Willmann, H.C. Chen, J. Sheen, Calciumsignaling through protein kinases. The Arabidopsis calciumdependentprotein kinase gene family, Plant Physiol. 129 (2002)469–485.[27] R.M. Ulloa, M. Raíces, G.C. MacIntosh, S. Maldonado, M.T. Téllez-Iñón, Jasmonic acid affects <strong>plant</strong> morphology and calcium-dependentprotein kinase expression and activity in Solanum tuberosum,Physiol. Plant. 115 (2002) 417–427.[28] C. Wang, C.A. Zien, M. Afitlhile, R. Welti, D.F. Hildebrand, X.Wang, Involvement of phospholipase D in wound-<strong>induced</strong> accumulationof jasmonic acid in Arabidopsis, Plant Cell 12 (2000)2237–2246.[29] E. Bell, R.A. Creelman, J.E. Mullet, A chloroplast lipoxygenase isrequired for wound-<strong>induced</strong> jasmonic acid accumulation in Arabidopsis,Proc. Natl. Acad. Sci. U. S. A. 92 (1995) 8675–8679.[30] J. Narváez-Vásquez, J. Florin-Christensen, C.A. Ryan, Positionalspecificity of a phospholipase A activity <strong>induced</strong> by wounding,systemin, and oligosaccharide elicitors in tomato leaves, Plant Cell11 (1999) 2249–2260.[31] P. Reymond, E.E. Farmer, Jasmonate and salicylate as globalsignals for defense gene expression, Curr. Opin. Plant Biol. 1 (1998)404–411.[32] J.G. Turner, C. Ellis, A. Devoto, The jasmonate signal pathway, PlantCell 14 (2002) S153–S164.


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 105[33] K. Maleck, R.A. Dietrich, Defense on multiple fronts: how do <strong>plant</strong>scope with diverse enemies? Trends Plant Sci. 4 (1999) 215–219.[34] T. Krumm, K. Bandemer, W. Boland, Induction of volatile biosynthesisin the Lima bean (Phaseolus lunatus) by leucine- andisoleucine conjugates of 1-oxo- and 1-hydroxyindan-4-carboxylicacid: evidence for amino acid conjugates of jasmonic acidintermediates in the octadecanoid signalling pathway, FEBS Lett.377 (1995) 523–529.[35] P.E. Staswick, I. Tiryaki, The oxylipin signal jasmonic acid isactivated by an enzyme that conjugates it to isoleucine inArabidopsis, Plant Cell 16 (2004) 2117–2127.[36] J. Royo, J. León, G. Vancanneyt, J.P. Albar, S. Rosahl, F. Ortego, P.Castañera, J.J. Sánchez-Serrano, Antisense-mediated depletion of apotato lipoxygenase reduces wound induction of proteinase inhibitorsand increases weight gain of insect pests, Proc. Natl. Acad. Sci.U. S. A. 96 (1999) 1146–1151.[37] M. McConn, R.A. Creelman, E. Bell, J.E. Mullet, J. Browse,Jasmonate is essential for insect defense in Arabidopsis, Proc. Natl.Acad. Sci. U. S. A. 94 (1997) 5473–5477.[38] P. Vijayan, J. Shockey, C.A. Lévesque, R.J. Cook, J. Browse, A rolefor jasmonate in pathogen defense of Arabidopsis, Proc. Natl. Acad.Sci. U. S. A. 95 (1998) 7209–7214.[39] J.A. Ryals, U.H. Neuenschwander, M.G. Willits, A. Molina, H.Y.Steiner, M.D. Hunt, Systemic acquired resistance, Plant Cell 8(1996) 1809–1819.[40] R. Ozawa, G. Arimura, J. Takabayashi, T. Shimoda, T. Nishioka,Involvement of jasmonate- and salicylate-related signaling pathwaysfor the production of specific herbivore-<strong>induced</strong> volatiles in <strong>plant</strong>s,Plant Cell Physiol. 41 (2000) 391–398.[41] J. Engelberth, T. Koch, G. Schuler, N. Bachmann, J. Rechtenbach,W. Boland, Ion channel-forming alamethicin is a potent elicitor ofvolatile biosynthesis and tendril coiling, Cross talk betweenjasmonate and salicylate signaling in lima bean, Plant Physiol. 125(2001) 369–377.[42] P.J. Moran, G.A. Thompson, Molecular responses to aphid feeding inArabidopsis in relation to <strong>plant</strong> defense pathways, Plant Physiol. 125(2001) 1074–1085.[43] S.H. Spoel, A. Koornneef, S.M. Claessens, J.P. Korzelius, J.A. VanPelt, M.J. Mueller, A.J. Buchala, J.P. Métraux, R. Brown, K.Kazan, L.C. Van Loon, X. Dong, C.M.J. Pieterse, NPR1 modulatescross-talk between salicylate- and jasmonate-dependent defensepathways through a novel function in the cytosol, Plant Cell 15(2003) 760–770.[44] D. Cipollini, S. Enright, M.B. Traw, J. Bergelson, Salicylic acidinhibits jasmonic acid-<strong>induced</strong> resistance of Arabidopsis thaliana toSpodoptera exigua, Mol. Ecol. 13 (2004) 1643– 1653.[45] T. Koch, T. Krumm, V. Jung, J. Engelberth, W. Boland, Differentialinduction of <strong>plant</strong> volatile biosynthesis in the lima bean by early andlate intermediates of the octadecanoid-signaling pathway, PlantPhysiol. 121 (1999) 153–162.[46] T. Shimoda, R. Ozawa, G. Arimura, J. Takabayashi, T. Nishioka,Olfactory responses of two specialist insect predators of spidermites toward <strong>plant</strong> volatiles from lima bean leaves <strong>induced</strong> byjasmonic acid and/or methyl salicylate, Appl. Entomol. Zool. 37(2002) 535–541.[47] M. Dicke, T.A. van Beek, M.A. Posthumus, N. Ben Dom, H. vanBokhoven, A.E. de Groot, Isolation and identification of volatilekairomone that affects acarine predator–prey interactions, J. Chem.Ecol. 16 (1990) 381–396.[48] I. Thoma, C. Loeffler, A.K. Sinha, M. Gupta, M. Krischke, B.Steffan, T. Roitsch, M.J. Mueller, Cyclopentenone isoprostanes<strong>induced</strong> by reactive oxygen species trigger defense gene activationand phytoalexin accumulation in <strong>plant</strong>s, Plant J. 34 (2003) 363–375.[49] C. Loeffler, S. Berger, A. Guy, T. Durand, G. Bringmann, M. Dreyer,U. von Rad, J. Durner, M.J. Mueller, B1-phytoprostanes trigger <strong>plant</strong>defense and detoxification responses, Plant Physiol. 137 (2005)328–340.[50] K. Harms, R. Atzorn, A. Brash, H. Kqhn, C. Wasternack, L.Willmitzer, H. Peña-Cortés, Expression of a flax allene oxidesynthase cDNA leads to increased endogenous jasmonic acid (JA)levels in transgenic potato <strong>plant</strong>s but not to a correspondingactivation of JA-responding genes, Plant Cell 7 (1995) 1645–1654.[51] P.J. O’Donnell, C. Calvert, R. Atzorn, C. Wasternack, H.M.O.Leyser, D.J. Bowles, Ethylene as a signal mediating the woundresponse of tomato <strong>plant</strong>s, Science 274 (1996) 1914–1917.[52] G. Arimura, R. Ozawa, T. Nishioka, W. Boland, T. Koch, F.Kqhnemann, J. Takabayashi, <strong>Herbivore</strong>-<strong>induced</strong> volatiles induce theemission of ethylene in neighboring lima bean <strong>plant</strong>s, Plant J. 29(2002) 87–98.[53] Y. Xu, P. Chang, D. Liu, M.L. Narasimhan, K.G. Raghothama, P.M.Hasegawa, R.A. Bressan, Plant defense genes are synergistically<strong>induced</strong> by ethylene and methyl jasmonate, Plant Cell 6 (1994)1077– 1085.[54] E.A. Schmelz, H.T. Alborn, E. Banchio, J.H. Tumlinson, Quantitativerelationships between <strong>induced</strong> jasmonic acid levels and volatileemission in Zea mays during Spodoptera exigua herbivory, Planta216 (2003) 665–673.[55] E.A. Schmelz, H.T. Alborn, J.H. Tumlinson, Synergistic interactionsbetween volicitin, jasmonic acid and ethylene mediate insect-<strong>induced</strong>volatile emission in Zea mays, Physiol. Plant. 117 (2003) 403–412.[56] J. Horiuchi, G. Arimura, R. Ozawa, T. Shimoda, J. Takabayashi, T.Nishioka, Exogenous ACC enhances volatiles production mediated byjasmonic acid in lima bean leaves, FEBS Lett. 509 (2001) 332–336.[57] J. Kahl, D.H. Siemens, R.J. Aerts, R. G7bler, F. Kqhnemann, C.A.Preston, I.T. Baldwin, <strong>Herbivore</strong>-<strong>induced</strong> ethylene suppresses adirect defense but not a putative <strong>indirect</strong> defense against an adaptedherbivore, Planta 210 (2000) 336–342.[58] C. Voelckel, U. Schittko, I.T. Baldwin, <strong>Herbivore</strong>-<strong>induced</strong> ethyleneburst reduces fitness costs of jasmonate- and oral secretion-<strong>induced</strong>defenses in Nicotiana attenuata, Oecologia 127 (2001) 274–280.[59] D. Laudert, E.W. Weiler, Allene oxide synthase: a major controlpoint in Arabidopsis thaliana octadecanoid signalling, Plant J. 15(1998) 675–684.[60] S. Sivasankar, B. Sheldrick, S.J. Rothstein, Expression of alleneoxide synthase determines defense gene activation in tomato, PlantPhysiol. 122 (2000) 1335–1342.[61] E. Rojo, R. Solano, J.J. Sánchez-Serrano, Interactions betweensignaling compounds involved in <strong>plant</strong> defense, J. Plant GrowthRegul. 22 (2003) 82–98.[62] K.L. Wang, H. Li, J.R. Ecker, Ethylene biosynthesis and signalingnetworks, Plant Cell 14 (2002) S131–S151.[63] C. Dammann, E. Rojo, J.J. Sánchez-Serrano, Abscisic acid andjasmonic acid activate wound-inducible genes in potato throughseparate, organ-specific signal transduction pathways, Plant J. 11(1997) 773–782.[64] G.F. Birkenmeier, C.A. Ryan, Wound signaling in tomato <strong>plant</strong>s.Evidence that aba is not a primary signal for defense gene activation,Plant Physiol. 117 (1998) 687–693.[65] J. León, E. Rojo, J.J. Sánchez-Serrano, Wound signaling in <strong>plant</strong>s, J.Exp. Bot. 52 (2001) 1–9.[66] Y. Takahashi, T. Berberich, A. Miyazaki, S. Seo, Y. Ohashi, T.Kusano, Spermine signalling in tobacco: activation of mitogenactivatedprotein kinases by spermine is mediated through mitochondrialdysfunction, Plant J. 36 (2003) 820–829.[67] H. Yamakawa, H. Kamada, M. Satoh, Y. Ohashi, Spermine is asalicylate-independent endogenous inducer for both tobacco acidicpathogenesis-related proteins and resistance against tobacco mosaicvirus infection, Plant Physiol. 118 (1998) 1213–1222.[68] M. Ohme-Takagi, K. Suzuki, H. Shinshi, Regulation of ethylene<strong>induced</strong>transcription of defense genes, Plant Cell Physiol. 41 (2000)1187–1192.[69] M. Ohme-Takagi, H. Shinshi, Ethylene-inducible DNA bindingproteins that interact with an ethylene-responsive element, Plant Cell7 (1995) 173–182.


106G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111[70] R. Solano, A. Stepanova, Q. Chao, J.R. Ecker, Nuclear events inethylene signaling: a transcriptional cascade mediated by ethyleneinsensitiveand ethylene-response-factor 1, Genes Dev. 12 (1998)3703– 3714.[71] S.Y. Fujimoto, M. Ohta, A. Usui, H. Shinshi, M. Ohme-Takagi,Arabidopsis ethylene-responsive element binding factors act astranscriptional activators or repressors of GCC box-mediated geneexpression, Plant Cell 12 (2000) 393–404.[72] R.L. Brown, K. Kazan, K.C. McGrath, D.J. Maclean, J.M.Manners, A role for the GCC-box in jasmonate-mediated activationof the PDF1.2 gene of Arabidopsis, Plant Physiol. 132 (2003)1020–1032.[73] F.L.H. Menke, A. Champion, J.W. Kijne, J. Memelink, A noveljasmonate- and elicitor-responsive element in the periwinklesecondary metabolite biosynthetic gene Str interacts with ajasmonate- and elicitor-inducible AP2-domain transcription factor,ORCA2, EMBO J. 18 (1999) 4455– 4463.[74] L. van der Fits, J. Memelink, ORCA3, a jasmonate-responsivetranscriptional regulator of <strong>plant</strong> primary and secondary metabolism,Science 289 (2000) 295–297.[75] Y.H. Xu, J.W. Wang, S. Wang, J.Y. Wang, X.Y. Chen, Characterizationof GaWRKY1, a cotton transcription factor that regulates thesesquiterpene synthase gene (+)-y-cadinene synthase-A, PlantPhysiol. 135 (2004) 507–515.[76] G. Arimura, K. Tashiro, S. Kuhara, T. Nishioka, R. Ozawa, J.Takabayashi, Gene responses in bean leaves <strong>induced</strong> by herbivoryand by herbivore-<strong>induced</strong> volatiles, Biochem. Biophys. Res.Commun. 277 (2000) 305–310.[77] P. Reymond, H. Weber, M. Damond, E.E. Farmer, Differential geneexpression in response to mechanical wounding and insect feeding inArabidopsis, Plant Cell 12 (2000) 707–720.[78] R. Halitschke, K. Gase, D. Hui, D.D. Schmidt, I.T. Baldwin,Molecular interactions between the specialist herbivore Manducasexta (lepidoptera, sphingidae) and its natural host Nicotianaattenuata: VI. Microarray analysis reveals that most herbivorespecifictranscriptional changes are mediated by fatty acid-aminoacid conjugates, Plant Physiol. 131 (2003) 1894–1902.[79] N. Qu, U. Schittko, I.T. Baldwin, Consistency of Nicotianaattenuata’s herbivore- and jasmonate-<strong>induced</strong> transcriptionalresponses in the allotetraploid species Nicotiana quadrivalvis andNicotiana clevelandii, Plant Physiol. 135 (2004) 539–548.[80] P. Reymond, N. Bodenhausen, R.M. Van Poecke, V. Krishnamurthy,M. Dicke, E.E. Farmer, A conserved transcript pattern inresponse to a specialist and a generalist herbivore, Plant Cell 16(2004) 3132–3147.[81] G. Pearce, D. Strydom, S. Johnson, C.A. Ryan, A polypeptide fromtomato leaves induces wound-inducible proteinase-inhibitor proteins,Science 253 (1991) 895–898.[82] M. Orozco-Cardenas, B. McGurl, C.A. Ryan, Expression of anantisense prosystemin gene in tomato <strong>plant</strong>s reduces resistancetoward Manduca sexta larvae, Proc. Natl. Acad. Sci. U. S. A. 90(1993) 8273–8276.[83] J.M. Scheer, C.A. Ryan Jr., The systemin receptor SR160 fromLycopersicon peruvianum is a member of the LRR receptor kinasefamily, Proc. Natl. Acad. Sci. U. S. A. 99 (2002) 9585–9590.[84] T. Montoya, T. Nomura, K. Farrar, T. Kaneta, T. Yokota, G.J.Bishop, Cloning the tomato curl3 gene highlights the putative dualrole of the leucine-rich repeat receptor kinase tBRI1/SR160 in <strong>plant</strong>steroid hormone and peptide hormone signaling, Plant Cell 14 (2002)3163– 3176.[85] M. Szekeres, Brassinosteroid and systemin: two hormones perceivedby the same receptor, Trends Plant Sci. 8 (2003) 102–104.[86] Z.Y. Wang, J.X. He, Brassinosteroid signal transduction — choicesof signals and receptors, Trends Plant Sci. 9 (2004) 91–96.[87] G. Pearce, D.S. Moura, J. Stratmann, C.A. Ryan, Production ofmultiple <strong>plant</strong> hormones from a single polyprotein precursor, Nature411 (2001) 817–820.[88] G. Pearce, D.S. Moura, J. Stratmann, C.A. Ryan Jr., RALF, a 5-kDaubiquitous polypeptide in <strong>plant</strong>s, arrests root growth and development,Proc. Natl. Acad. Sci. U. S. A. 98 (2001) 12843–12847.[89] G. Pearce, C.A. Ryan, Systemic signaling in tomato <strong>plant</strong>s fordefense against herbivores. Isolation and characterization of threenovel defense-signaling glycopeptide hormones coded in a singleprecursor gene, J. Biol. Chem. 278 (2003) 30044–30050.[90] G.I. Lee, G.A. Howe, The tomato mutant spr1 is defective insystemin perception and the production of a systemic wound signalfor defense gene expression, Plant J. 33 (2003) 567–576.[91] G. Arimura, D.P. Huber, J. Bohlmann, Forest tent caterpillars(Malacosoma disstria) induce local and systemic diurnal emissionsof terpenoid volatiles in hybrid poplar (Populus trichocarpa xdeltoides): cDNA cloning, functional characterization, and patternsof gene expression of ( )-germacrene D synthase, PtdTPS1, Plant J.37 (2004) 603–616.[92] A.M. Maldonado, P. Doerner, R.A. Dixon, C.J. Lamb, R.K.Cameron, A putative lipid transfer protein involved in systemicresistance signalling in Arabidopsis, Nature 419 (2002) 399–403.[93] D.C. Wildon, J.F. Thain, P.E.H. Minchin, I.R. Gubb, A.J. Reilly,Y.D. Skipper, H.M. Doherty, P.J. O’Donnell, D.J. Bowles, Electricalsignaling and systemic proteinase-inhibitor induction in the wounded<strong>plant</strong>, Nature 360 (1992) 62–65.[94] L. Li, C. Li, G.I. Lee, G.A. Howe, Distinct roles for jasmonatesynthesis and action in the systemic wound response of tomato, Proc.Natl. Acad. Sci. U. S. A. 99 (2002) 6416–6421.[95] I.T. Baldwin, E.A. Schmelz, T.E. Ohnmeiss, Wound-<strong>induced</strong>changes in root and shoot jasmonic acid pools correlate with<strong>induced</strong> nicotine synthesis in Nicotiana sylvestris Spegazzini andComes, J. Chem. Ecol. 20 (1994) 2139–2157.[96] H.K. Lichtenthaler, The 1-deoxy-d-xylulose-5-phosphate pathway ofisoprenoid biosynthesis in <strong>plant</strong>s, Annu. Rev. Plant Physiol. PlantMol. Biol. 50 (1999) 47–65.[97] M. Rodríguez-Concepción, A. Boronat, Elucidation of the methylerythritolphosphate pathway for isoprenoid biosynthesis in bacteriaand plastids. A metabolic milestone achieved through genomics,Plant Physiol. 130 (2002) 1079–1089.[98] S. Aubourg, A. Lecharny, J. Bohlmann, Genomic analysis of theterpenoid synthase (AtTPS) gene family of Arabidopsis thaliana,Mol. Gen. Genomics 267 (2002) 730–745.[99] J. Bohlmann, D. Martin, N.J. Oldham, J. Gershenzon, Terpenoidsecondary metabolism in Arabidopsis thaliana: cDNA cloning,characterization, and functional expression of a myrcene/(E)-hocimenesynthase, Arch. Biochem. Biophys. 375 (2000) 261–269.[100] J. F7ldt, G. Arimura, J. Gershenzon, J. Takabayashi, J. Bohlmann,Functional identification of AtTPS03 as (E)-h-ocimene synthase: amonoterpene synthase catalyzing jasmonate- and wound-<strong>induced</strong>volatile formation in Arabidopsis thaliana, Planta 216 (2003)745–751.[101] F. Chen, D. Tholl, J.C. D’Auria, A. Farooq, E. Pichersky, J.Gershenzon, Biosynthesis and emission of terpenoid volatiles fromArabidopsis flowers, Plant Cell 15 (2003) 481–494.[102] F. Chen, D.K. Ro, J. Petri, J. Gershenzon, J. Bohlmann, E. Pichersky,D. Tholl, Characterization of a root-specific Arabidopsis terpenesynthase responsible for the formation of the volatile monoterpene1,8-Cineole, Plant Physiol. 135 (2004) 1956–1966.[103] S.A.B. McKay, W.L. Hunter, K.A. Godard, S.X. Wang, D.M. Martin,J. Bohlmann, A.L. Plant, Insect attack and wounding inducetraumatic resin duct development and gene expression of ( )-pinenesynthase in Sitka spruce, Plant Physiol. 133 (2003) 368–378.[104] N. Dudareva, D. Martin, C.M. Kish, N. Kolosova, N. Gorenstein, J.F7ldt, B. Miller, J. Bohlmann, (E)-h-Ocimene and myrcene synthasegenes of floral scent biosynthesis in snapdragon: function andexpression of three terpene synthase genes of a new terpene synthasesubfamily, Plant Cell 15 (2003) 1227– 1241.[105] L. Sharon-Asa, M. Shalit, A. Frydman, E. Bar, D. Holland, E. Or, U.Lavi, E. Lewinsohn, Y. Eyal, Citrus fruit flavor and aroma


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 107biosynthesis: isolation, functional characterization, and developmentalregulation of Cstps1, a key gene in the production ofthe sesquiterpene aroma compound valencene, Plant J. 36 (2003)664–674.[106] B. Miller, L.L. Madilao, S. Ralph, J. Bohlmann, Insect-<strong>induced</strong>conifer defense. White pine weevil and methyl jasmonate inducetraumatic resinosis, de novo formed volatile emissions, andaccumulation of terpenoid synthase and putative octadecanoidpathway transcripts in sitka spruce, Plant Physiol. 137 (2005)369–382.[107] M. Kunert, A. Biedermann, T. Koch, W. Boland, Ultrafast samplingand analysis of <strong>plant</strong> volatiles by a hand-held miniaturised GC withpre-concentration unit: kinetic and quantitative aspects of <strong>plant</strong>volatile production, J. Sep. Sci. 25 (2002) 677–684.[108] G. Arimura, R. Ozawa, S. Kugimiya, J. Takabayashi, J. Bohlmann,<strong>Herbivore</strong>-<strong>induced</strong> defense response in a model legume: two-spottedspider mites, Tetranychus urticae, induce emission of (E)-h-ocimeneand transcript accumulation of (E)-h-ocimene synthase in Lotusjaponicus, Plant Physiol. 135 (2004) 1976– 1983.[109] J. Lqcker, H.J. Bouwmeester, W. Schwab, J. Blaas, L.H.W. van derPlas, H.A. Verhoeven, Expression of Clarkia S-linalool synthase intransgenic petunia <strong>plant</strong>s results in the accumulation of S-linalyl-hd-glucopyranoside,Plant J. 27 (2001) 315–324.[110] A. Aharoni, A.P. Giri, S. Deuerlein, F. Griepink, W.J. de Kogel,F.W.A. Verstappen, H.A. Verhoeven, M.A. Jongsma, W. Schwab,H.J. Bouwmeester, Terpenoid metabolism in wild-type and transgenicArabidopsis <strong>plant</strong>s, Plant Cell 15 (2003) 2866–2884.[111] J. Donath, W. Boland, Biosynthesis of acyclic homoterpenes inhigher <strong>plant</strong>s parallels steroid hormone metabolism, J. Plant Physiol.143 (1994) 473–478.[112] J. Degenhardt, J. Gershenzon, Demonstration and characterization of(E)-nerolidol synthase from maize: a herbivore-inducible terpenesynthase participating in (3E)-4,8-dimethyl-1,3,7-nonatriene biosynthesis,Planta 210 (2000) 815–822.[113] G. Arimura, R. Ozawa, T. Shimoda, T. Nishioka, W. Boland, J.Takabayashi, Herbivory-<strong>induced</strong> volatiles elicit defence genes inlima bean leaves, Nature 406 (2000) 512–515.[114] S. Seo, H. Seto, H. Koshino, S. Yoshida, Y. Ohashi, A diterpene asan endogenous signal for the activation of defense responses toinfection with tobacco mosaic virus and wounding in tobacco, PlantCell 15 (2003) 863–873.[115] A. Hatanaka, The biogeneration of green odour by green leaves,Phytochemistry 34 (1993) 1201–1218.[116] K. Matsui, S. Kurishita, A. Hisamitsu, T. Kajiwara, A lipidhydrolysingactivity involved in hexenal formation, Biochem. Soc.Trans. 28 (2000) 857–860.[117] E. Alméras, S. Stolz, S. Vollenweider, P. Reymond, L. Méne-Saffrané, E.E. Farmer, Reactive electrophile species activate defensegene expression in Arabidopsis, Plant J. 34 (2003) 205–216.[118] M.A. Farag, P.W. Paré, C 6 -Green leaf volatiles trigger localand systemic VOC emissions in tomato, Phytochemistry 61 (2002)545–554.[119] N.J. Bate, S.J. Rothstein, C 6 -volatiles derived from the lipoxygenasepathway induce a subset of defense-related genes, Plant J. 16 (1998)561–569.[120] G. Arimura, R. Ozawa, J. Horiuchi, T. Nishioka, J. Takabayashi,Plant–<strong>plant</strong> interactions mediated by volatiles emitted from<strong>plant</strong>s infested by spider mites, Biochem. Syst. Ecol. 29 (2001)1049–1061.[121] J. Engelberth, H.T. Alborn, E.A. Schmelz, J.H. Tumlinson, Airbornesignals prime <strong>plant</strong>s against insect herbivore attack, Proc. Natl. Acad.Sci. U. S. A. 101 (2004) 1781–1785.[122] D.F. Hildebrand, G.C. Brown, D.M. Jackson, T.R. Hamilton-Kemp,Effects of some leaf-emitted volatile compounds on aphid populationincrease, J. Chem. Ecol. 19 (1993) 1875–1887.[123] G. Vancanneyt, C. Sanz, T. Farmaki, M. Paneque, F. Ortego, P.Castanera, J.J. Sánchez-Serrano, Hydroperoxide lyase depletion intransgenic potato <strong>plant</strong>s leads to an increase in aphid performance,Proc. Natl. Acad. Sci. U. S. A. 98 (2001) 8139–8144.[124] H.W. Gardner, D.L.J. Dornbos, A.E. Desjardins, Hexanal, trans-2-hexenal, and trans-2-nonenal inhibit soybean, Glycine max, seedgermination, J. Agric. Food Chem. 38 (1990) 1316–1320.[125] T.R. Hamilton-Kemp, J.H. Loughrin, D.D. Archbold, R.A. Andersen,D.F. Hildebrand, Inhibition of pollen germination by volatilecompounds including 2-hexenal and 3-hexenal, J. Agric. FoodChem. 39 (1991) 952–956.[126] T.C. Turlings, J.H. Loughrin, P.J. McCall, U.S. Rose, W.J. Lewis,J.H. Tumlinson, How caterpillar-damaged <strong>plant</strong>s protect themselvesby attracting parasitic wasps, Proc. Natl. Acad. Sci. U. S. A. 92(1995) 4169–4174.[127] U.S.R. Röse, J.H. Tumlinson, Volatiles released from cotton <strong>plant</strong>s inresponse to Helicoverpa zea feeding damage on cotton flower buds,Planta 218 (2004) 824–832.[128] E.E. Farmer, C.A. Ryan, Inter<strong>plant</strong> communication: airborne methyljasmonate induces synthesis of proteinase inhibitors in <strong>plant</strong> leaves,Proc. Natl. Acad. Sci. U. S. A. 87 (1990) 7713–7716.[129] H.S. Seo, J.T. Song, J.J. Cheong, Y.H. Lee, Y.W. Lee, I. Hwang, J.S.Lee, Y.D. Choi, Jasmonic acid carboxyl methyltransferase: a keyenzyme for jasmonate-regulated <strong>plant</strong> responses, Proc. Natl. Acad.Sci. U. S. A. 98 (2001) 4788–4793.[130] M.A. Birkett, C.A.M. Campbell, K. Chamberlain, E. Guerrieri, A.J.Hick, J.L. Martin, M. Matthes, J.A. Napier, J. Pettersson, J.A.Pickett, G.M. Poppy, E.M. Pow, B.J. Pye, L.E. Smart, G.H.Wadhams, L.J. Wadhams, C.M. Woodcock, New roles for cisjasmoneas an insect semiochemical and in <strong>plant</strong> defense, Proc. Natl.Acad. Sci. U. S. A. 97 (2000) 9329–9334.[131] C.C. von Dahl, I.T. Baldwin, Methyl jasmonate and cis-jasmone donot dispose of the herbivore-<strong>induced</strong> jasmonate burst in Nicotianaattenuata, Physiol. Plant. 120 (2004) 474–481.[132] J.C. D’Auria, F. Chen, E. Pichersky, The SABATH family of MTs inArabidopsis thaliana and other <strong>plant</strong> species, in: J.T. Romeo (Ed.),Recent Advances in Phytochemistry, Elsevier Science Ltd., Oxford,2003, pp. 253–283.[133] F. Chen, J.C. D’Auria, D. Tholl, J.R. Ross, J. Gershenzon, J.P. Noel,E. Pichersky, An Arabidopsis thaliana gene for methylsalicylatebiosynthesis, identified by a biochemical genomics approach, has arole in defense, Plant J. 36 (2003) 577–588.[134] V. Shulaev, P. Silverman, I. Raskin, Airborne signalling bymethyl salicylate in <strong>plant</strong> pathogen resistance, Nature 385 (1997)718–721.[135] M. Frey, C. Stettner, P.W. Paré, E.A. Schmelz, J.H. Tumlinson, A.Gierl, An herbivore elicitor activates the gene for indole emission inmaize, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 14801–14806.[136] M. Frey, D. Spiteller, W. Boland, A. Gierl, Transcriptional activationof Igl, the gene for indole formation in Zea mays: a structure–activitystudy with elicitor-active N-acyl glutamines from insects, Phytochemistry65 (2004) 1047–1055.[137] M. Dicke, L.E.M. Vet, Plant–carnivore interactions: evolutionary andecological consequences for <strong>plant</strong>, herbivore and carnivore, in: R.H.Drent (Ed.), <strong>Herbivore</strong>s: Between Plants and Predators, BlackwellScience, Oxford, 1999, pp. 483–520.[138] M. Dicke, Specifity of herbivore-<strong>induced</strong> <strong>plant</strong> <strong>defences</strong>, in: J.A.Goode (Ed.), Insect–Plant Interactions and Induced Plant Defence,John Wiley & Sons, Ltd., Chichester, 1997.[139] T.C.J. Turlings, F.L. W7ckers, L.E.M. Vet, J.M. Lewis, J.H.Tumlinson, Learning of host-finding cues by hymenopterousparasitoids, in: A.C. Lewis (Ed.), Insect Learning — Ecologyand Evolutionary Perspectives, Chapman & Hall, New York, 1993,pp. 51–78.[140] E.N. Barata, J.A. Pickett, L.J. Wadhams, C.M. Woodcock, H.Mustaparta, Identification of host and nonhost semiochemicalsof eucalyptus woodborer Phoracantha semipunctata by gas chromatography-electroantennography,J. Chem. Ecol. 26 (2000)1877–1895.


108G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111[141] J. Takabayashi, M. Dicke, M.A. Posthumus, Variation in compositionof predator-attracting allelochemicals emitted by herbivoreinfested<strong>plant</strong>s: relative influence of <strong>plant</strong> and herbivore, Chemoecology2 (1991) 1 –6.[142] S. Gouinguene, T. Degen, T.C.J. Turlings, Variability in herbivore<strong>induced</strong>odour emissions among maize cultivars and their wildancestors (teosinte), Chemoecology 11 (2001) 9 –16.[143] P. Scutareanu, J. Bruin, M.A. Posthumus, B. Drukker, Constitutiveand herbivore-<strong>induced</strong> volatiles in pear, alder and hawthorn trees,Chemoecology 13 (2003) 63–74.[144] O.E. Krips, P.E.L. Willems, R. Gols, M.A. Posthumus, G. Gort, M.Dicke, Comparison of cultivars of ornamental crop Gerbera jamesoniion production of spider mite-<strong>induced</strong> volatiles, and theirattractiveness to the predator Phytoseiulus persimilis, J. Chem. Ecol.27 (2001) 1355–1372.[145] J. Takabayashi, M. Dicke, S. Takahashi, M.A. Posthumus, T.A.Vanbeek, Leaf age affects composition of herbivore-<strong>induced</strong>synomones and attraction of predatory mites, J. Chem. Ecol. 20(1994) 373–386.[146] J. Takabayashi, M. Dicke, M.A. Posthumus, Volatile herbivore<strong>induced</strong>terpenoids in <strong>plant</strong> mite interactions — variation caused bybiotic and abiotic factors, J. Chem. Ecol. 20 (1994) 1329–1354.[147] C.M. De Moraes, M.C. Mescher, J.H. Tumlinson, Caterpillar<strong>induced</strong>nocturnal <strong>plant</strong> volatiles repel nonspecific females, Nature410 (2001) 577–580.[148] M. Hilker, T. Meiners, Induction of <strong>plant</strong> responses to ovipositionand feeding by herbivorous arthropods: a comparison, Entomol. Exp.Appl. 104 (2002) 181–192.[149] M. Dicke, Are herbivore-<strong>induced</strong> <strong>plant</strong> volatiles reliable indicators ofherbivore identity to foraging carnivorous arthropods? Entomol.Exp. Appl. 91 (1999) 131–142.[150] S. Gouinguene, H. Alborn, T.C.J. Turlings, Induction of volatileemissions in maize by different larval instars of Spodopteralittoralis, J. Chem. Ecol. 29 (2003) 145–162.[151] J. Takabayashi, S. Takahashi, M. Dicke, M.A. Posthumus, Developmentalstage of herbivore Pseudaletia separata affects productionof herbivore-<strong>induced</strong> synomone by corn <strong>plant</strong>s, J. Chem. Ecol. 21(1995) 273–287.[152] B. Weissbecker, J.J.A. Van Loon, M.A. Posthumus, H.J. Bouwmeester,M. Dicke, Identification of volatile potato sesquiterpenoidsand their olfactory detection by the two-spotted stinkbug Perillusbioculatus, J. Chem. Ecol. 26 (2000) 1433–1445.[153] Y.J. Du, G.M. Poppy, W. Powell, J.A. Pickett, L.J. Wadhams, C.M.Woodcock, Identification of semiochemicals released during aphidfeeding that attract parasitoid Aphidius ervi, J. Chem. Ecol. 24(1998) 1355–1368.[154] C.G. Galizia, S. Sachse, H. Mustaparta, Calcium responses topheromones and <strong>plant</strong> odours in the antennal lobe of the male andfemale moth Heliothis virescens, J. Comp. Physiol., A Neuroethol.Sens. Neural Behav. Physiol. 186 (2000) 1049–1063.[155] H.T. Skiri, C.G. Galizia, H. Mustaparta, Representation of primary<strong>plant</strong> odorants in the antennal lobe of the moth Heliothis virescensusing calcium imaging, Chem. Senses 29 (2004) 253–267.[156] L.E.M. Vet, M. Dicke, Ecology of infochemical use by naturalenemies in a tritrophic context, Annu. Rev. Entomol. 37 (1992)141–172.[157] J.I. Horiuchi, G.I. Arimura, R. Ozawa, T. Shimoda, M. Dicke, J.Takabayashi, T. Nishioka, Lima bean leaves exposed to herbivore<strong>induced</strong>conspecific <strong>plant</strong> volatiles attract herbivores in addition tocarnivores, Appl. Entomol. Zool. 38 (2003) 365–368.[158] M. Heil, Direct defense or ecological costs: responses of herbivorousbeetles to volatiles released by wild Lima bean (Phaseolus lunatus),J. Chem. Ecol. 30 (2004) 1289–1295.[159] R. Gols, M. Roosjen, H. Dijkman, M. Dicke, Induction of direct and<strong>indirect</strong> <strong>plant</strong> responses by jasmonic acid, low spider mite densities,or a combination of jasmonic acid treatment and spider miteinfestation, J. Chem. Ecol. 29 (2003) 2651–2666.[160] R. van Tol, A.T.C. van der Sommen, M.I.C. Boff, J. van Bezooijen,M.W. Sabelis, P.H. Smits, Plants protect their roots by alerting theenemies of grubs, Ecol. Lett. 4 (2001) 292–294.[161] N. Neveu, J. Grandgirard, J.P. Nenon, A.M. Cortesero, Systemicrelease of herbivore-<strong>induced</strong> <strong>plant</strong> volatiles by turnips infested byconcealed root-feeding larvae Delia radicum L, J. Chem. Ecol. 28(2002) 1717–1732.[162] W.H. van der Putten, L.E.M. Vet, J.A. Harvey, F.L. Wackers, Linkingabove- and belowground multitrophic interactions of <strong>plant</strong>s, herbivores,pathogens, and their antagonists, Trends Ecol. Evol. 16 (2001)547–554.[163] T. Shimoda, J. Takabayashi, W. Ashihara, A. Takafuji, Response ofpredatory insect Scolothrips takahashii toward herbivore-<strong>induced</strong><strong>plant</strong> volatiles under laboratory and field conditions, J. Chem. Ecol.23 (1997) 2033–2048.[164] P. Scutareanu, B. Drukker, J. Bruin, M.A. Posthumus, M.W. Sabelis,Volatiles from Psylla-infested pear trees and their possible involvementin attraction of anthocorid predators, J. Chem. Ecol. 23 (1997)2241–2260.[165] B. Drukker, P. Scutareanu, M.W. Sabelis, Do anthocorid predatorsrespond to synomones from Psylla-infested pear trees under fieldconditions, Entomol. Exp. Appl. 77 (1995) 193–203.[166] C.M. De Moraes, J.M. Lewis, P.W. Paré, H.T. Alborn, J.H.Tumlinson, Herbivory-infested <strong>plant</strong>s selectively attract parasitoids,Nature 393 (1998) 570–573.[167] J.S. Thaler, Jasmonate-inducible <strong>plant</strong> <strong>defences</strong> cause increasedparasitism of herbivores, Nature 399 (1999) 686–688.[168] A. Kessler, I.T. Baldwin, Defensive function of herbivore-<strong>induced</strong><strong>plant</strong> volatile emissions in nature, Science 291 (2001) 2141–2144.[169] M. Heil, Induction of two <strong>indirect</strong> <strong>defences</strong> benefits Limabean (Phaseolus lunatus, Fabaceae) in nature, J. Ecol. 92 (2004)527–536.[170] A.A. Agrawal, R. Karban, Domatia mediate <strong>plant</strong>–arthropodmutualism, Nature 387 (1997) 562–563.[171] T.S. Elias, Extrafloral nectaries: their structure and distribution, in:T.S. Elias (Ed.), The Biology of Nectaries, Columbia UniversityPress, New York, 1983, pp. 174–203.[172] S. Koptur, Extrafloral nectary-mediated interactions between insectsand <strong>plant</strong>s, in: E. Bernays (Ed.), Insect–Plant Interactions, vol. IV,CRC Press, Boca Raton, 1992, pp. 81–129.[173] B.L. Bentley, Extrafloral nectaries and protection by pugnaciousbodyguards, Ann. Rev. Ecolog. Syst. 8 (1977) 407–427.[174] A. Frey-Wyssling, The phloem supply to nectaries, Acta Bot. Neerl.4 (1955) 358–369.[175] A. Fahn, Structure and function of secretory cells, in: J.C. Gray(Ed.), Advances in Botanical Research: Incorporating Advances inPlant Pathology, Plant Trichomes, vol. 31, Academic Press, London,2000, pp. 37–75.[176] N. Bargoni, Synthesis of sucrose in the nectary of Convolvulussepium, Boll. Soc. Ital. Biol. Sper. 48 (1972) 1159–1160.[177] L. Pascal, M. Belindepoux, On the biological rhythm correlation of<strong>plant</strong>–ant association — the case of extrafloral nectaries of AmericanMalpighiaceae, C. R. Acad. Sci., III-Vie 312 (1991) 49–54.[178] T.S. Elias, Morphology and anatomy of foliar nectaries ofPithecellobium macradenium (Leguminosae), Bot. Gaz. 133(1972) 38–42.[179] M. Heil, B. Fiala, B. Baumann, K.E. Linsenmair, Temporal, spatialand biotic variations in extrafloral nectar secretion by Macarangatanarius, Funct. Ecol. 14 (2000) 749–757.[180] N.E. Raine, P. Willmer, G.N. Stone, Spatial structuring and floralavoidance behavior prevent ant–pollinator conflict in a Mexican antacacia,Ecology 83 (2002) 3086–3096.[181] B.L. Bentley, Plants bearing extrafloral nectaries and the associatedant community: interhabitat differences in the reduction of herbivoredamage, Ecology 57 (1976) 815–820.[182] B.L. Bentley, The protective function of ants visiting the extrafloralnectaries of Bixa orellana (Bixaceae), J. Ecol. 65 (1977) 27–38.


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 109[183] U. Lqttge, Structure and function of <strong>plant</strong> glands, Annu. Rev. PlantPhysiol. 22 (1971) 23–44.[184] M. Heil, T. Koch, A. Hilpert, B. Fiala, W. Boland, K.E.Linsenmair, Extrafloral nectar production of the ant-associated<strong>plant</strong>, Macaranga tanarius, is an <strong>induced</strong>, <strong>indirect</strong>, defensiveresponse elicited by jasmonic acid, Proc. Natl. Acad. Sci. U. S. A.98 (2001) 1083–1088.[185] F.L. W7ckers, T.M. Bezemer, Root herbivory induces an aboveground<strong>indirect</strong> defence, Ecol. Lett. 6 (2003) 9–12.[186] F.L. W7ckers, D. Zuber, R. Wunderlin, F. Keller, The effect ofherbivory on temporal and spatial dynamics of foliar nectarproduction in cotton and castor, Ann. Bot. 87 (2001) 365–370.[187] M. Heil, S. Greiner, H. Meimberg, R. Kruger, J.L. Noyer, G. Heubl,K.E. Linsenmair, W. Boland, Evolutionary change from <strong>induced</strong> toconstitutive expression of an <strong>indirect</strong> <strong>plant</strong> resistance, Nature 430(2004) 205–208.[188] J.H. Ness, Catalpa bignonioides alters extrafloral nectar productionafter herbivory and attracts ant bodyguards, Oecologia 134 (2003)210–218.[189] E.B. Mondor, J.F. Addicott, Conspicuous extra-floral nectaries areinducible in Vicia faba, Ecol. Lett. 6 (2003) 495–497.[190] D.J. O’Dowd, E.A. Catchpole, Ants and extrafloral nectaries: noevidence for <strong>plant</strong> protection in Helichrysum spp.– ant interactions,Oecologia 59 (1983) 191–200.[191] W.J. Boecklen, The role of extrafloral nectaries in the herbivoredefence of Cassia fasciculata, Ecol. Entomol. 9 (1984) 243–249.[192] J.X.I. Becerra, D.L. Venable, Extrafloral nectaries: a defense againstant–Homoptera mutualisms? Oikos 55 (1989) 276–280.[193] V.K. Rashbrook, S.G. Compton, J.H. Lawton, Ant–herbivoreinteractions: reasons for the absence of benefits to a fern with foliarnectaries, Ecology 73 (1992) 2167–2174.[194] K. Mody, K.E. Linsenmair, Plant-attracted ants affect arthropodcommunity structure but not necessarily herbivory, Ecol. Entomol.29 (2004) 217–225.[195] A. Dejean, D. McKey, M. Gibernau, M. Belin, The arboreal antmosaic in a Cameroonian rainforest (Hymenoptera: Formicidae),Sociobiology 35 (2000) 403–423.[196] C.C. Horvitz, D.W. Schemske, Effects of ants and an ant-tendedherbivore on seed production of a neotropical herb, Ecology 65(1984) 1369–1378.[197] D.W. Inouye, O.R.J. Taylor, A temperate region <strong>plant</strong> ant seedpredator system consequences of extra floral nectar secretion byHelianthella quinquenervis, Ecology 60 (1979) 1–7.[198] A.M. Barton, Spatial variation in the effect of ants on an extrafloralnectary <strong>plant</strong>, Ecology 67 (1986) 495–504.[199] P.A. Heads, J.H. Lawton, Bracken ants and extrafloral nectaries: III.How insect herbivores avoid ant predation, Ecol. Entomol. 10 (1985)29–42.[200] S.V. Fowler, M. Macgarvin, The impact of hairy wood ants, Formicalugubris, on the guild structure of herbivorous insects on birch,Betula pubescens, J. Anim. Ecol. 54 (1985) 847–856.[201] R.L. Bugg, R.T. Ellis, R.W. Carlson, Ichneumonidae (Hymenoptera)using extrafloral nectar of Faba bean (Vicia faba L, Fabaceae) inMassachusetts, Biol. Agric. Hortic. 6 (1989) 107–114.[202] J.O. Stapel, A.M. Cortesero, C.M. DeMoraes, J.H. Tumlinson,W.J. Lewis, Extrafloral nectar, honeydew, and sucrose effects onsearching behavior and efficiency of Microplitis croceipes(Hymenoptera: Braconidae) in cotton, Environ. Entomol. 26 (1997)617–623.[203] C. Kost, M. Heil, Increased availability of extrafloral nectar reducesherbivory in Lima bean <strong>plant</strong>s (Phaseolus lunatus, Fabaceae), BasicAppl. Ecol. (in press).[204] S. Ruhren, S.N. Handel, Jumping spiders (Salticidae) enhance theseed production of a <strong>plant</strong> with extrafloral nectaries, Oecologia 119(1999) 227–230.[205] P.C.J. van Rijn, L.K. Tanigoshi, The contribution of extrafloralnectar to survival and reproduction of the predatory mite Iphiseiusdegenerans on Ricinus communis, Exp. Appl. Acarol. 23 (1999)281–296.[206] R.W. Pemberton, J.H. Lee, The influence of extrafloral nectaries onparasitism of an insect herbivore, Am. J. Bot. 83 (1996) 1187–1194.[207] M. Cuautle, V. Rico-Gray, The effect of wasps and ants on thereproductive success of the extrafloral nectaried <strong>plant</strong> Turneraulmifolia (Turneraceae), Funct. Ecol. 17 (2003) 417–423.[208] R.A. Raguso, Why are some floral nectars scented? Ecology 85(2004) 1486–1494.[209] D.A. Avé, P. Gregory, W.M. Tingey, Aphid repellent sesquiterpenesin glandular trichomes of Solanum berthaultii and S. tuberosum,Entomol. Exp. Appl. 44 (1987) 131–138.[210] D.R. Crankshaw, J.H. Langenheim, Variation in terpenes andphenolics through leaf development in hymenaea and its possiblesignificance to herbivory, Biochem. Syst. Ecol. 9 (1981) 115–124.[211] M.E. Hoballah, T.G. Köllner, J. Degenhardt, T.C.J. Turlings, Costs of<strong>induced</strong> volatile production in maize, Oikos 105 (2004) 168–180.[212] K.P.C. Croft, F. Juttner, A.J. Slusarenko, Volatile products of thelipoxygenase pathway evolved from Phaseolus vulgaris (L) leavesinoculated with Pseudomonas syringae pv Phaseolicola, PlantPhysiol. 101 (1993) 13–24.[213] M. Dicke, Evolution of <strong>induced</strong> <strong>indirect</strong> <strong>defences</strong> of <strong>plant</strong>s, in: C.D.Harvell (Ed.), The Ecology and Evolution of Inducible Defenses,Princeton University Press, Princeton, 1999.[214] A. Janssen, M.W. Sabelis, J. Bruin, Evolution of herbivore-<strong>induced</strong><strong>plant</strong> volatiles, Oikos 97 (2002) 134–138.[215] M. Dicke, J.J.A. van Loon, Multitrophic effects of herbivore-<strong>induced</strong><strong>plant</strong> volatiles in an evolutionary context, Entomol. Exp. Appl. 97(2000) 237–249.[216] J.G. Zimmermann, Über die extrafloralen Nektarien der Angiospermen,Bot. Zbl. Beih. 49 (1932) 99–196.[217] J.A. Rudgers, Enemies of herbivores can shape <strong>plant</strong> traits: selectionin a facultative ant–<strong>plant</strong> mutualism, Ecology 85 (2004) 192–205.[218] C.L. Rhyne, Inheritance of extrafloral nectaries in cotton, Adv. Front.Plant Sci. 13 (1965) 121–135.[219] C.R. Meyer, V.G. Meyer, Origin and inheritance in nectariless cotton,Crop Sci. 1 (1961) 167–170.[220] C.C. Daehler, M. Yorkston, W.G. Sun, N. Dudley, Genetic variationin morphology and growth characters of Acacia koa in the Hawaiianislands, Int. J. Plant Sci. 160 (1999) 767–773.[221] M. Heil, Ecological costs of <strong>induced</strong> resistance, Curr. Opin. PlantBiol. 5 (2002) 345–350.[222] D. Cipollini, C.B. Purrington, J. Bergelson, Costs of <strong>induced</strong>responses in <strong>plant</strong>s, Basic Appl. Ecol. 4 (2003) 79–85.[223] A.R. Zangerl, Evolution of <strong>induced</strong> <strong>plant</strong> responses to herbivores,Basic Appl. Ecol. 4 (2003) 91–103.[224] T.J. DeWitt, A. Sih, D.S. Wilson, Costs and limits of phenotypicplasticity, Trends Ecol. Evol. 13 (1998) 77–81.[225] D.A. Herms, W.J. Mattson, The dilemma of <strong>plant</strong>s: to grow or todefend, Q. Rev. Biol. 67 (1992) 283–335.[226] T. Mitchell-Olds, D. Siemens, D. Pedersen, Physiology and costs ofresistance to herbivory and disease in Brassica, Entomol. Exp. Appl.80 (1996) 231–237.[227] A.A. Agrawal, J.K. Conner, M.T.J. Johnson, R. Wallsgrove,Ecological genetics of an <strong>induced</strong> <strong>plant</strong> defense against herbivores:additive genetic variance and costs of phenotypic plasticity,Evolution 56 (2002) 2206–2213.[228] I.T. Baldwin, P. Callahan, Autotoxicity and chemical defense —nicotine accumulation and carbon gain in solanaceous <strong>plant</strong>s,Oecologia 94 (1993) 534–541.[229] I.T. Baldwin, W. Hamilton, Jasmonate-<strong>induced</strong> responses of Nicotianasylvestris results in fitness costs due to impaired competitiveability for nitrogen, J. Chem. Ecol. 26 (2000) 915–952.[230] A.A. Agrawal, A. Janssen, J. Bruin, M.A. Posthumus, M.W.Sabelis, An ecological cost of <strong>plant</strong> defence: attractiveness of bittercucumber <strong>plant</strong>s to natural enemies of herbivores, Ecol. Lett. 5(2002) 377–385.


110G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111[231] P.W. Price, C.E. Bouton, P. Gross, B.A. McPheron, J.N. Thompson,A.E. Weis, Interactions among 3 trophic levels — influence of <strong>plant</strong>son interactions between insect herbivores and natural enemies, Ann.Rev. Ecolog. Syst. 11 (1980) 41–65.[232] E. Haukioja, T. Hakala, <strong>Herbivore</strong> cycles and periodic outbreaks:formulation of a general hypothesis, Rep. Kevo Subarct. 12 (1975)1–9.[233] D.J. O’Dowd, Foliar nectar production and ant activity ona neotropical tree, Ochroma pyramidale, Oecologia 43 (1979)233–248.[234] S.L. Gulmon, H.A. Mooney, Costs of defense and their effects on<strong>plant</strong> productivity, in: T.J. Givnish (Ed.), On the Economy of PlantForm and Function, Cambridge University Press, Cambridge, 1986.[235] J. Gershenzon, Metabolic costs of terpenoid accumulation in higher<strong>plant</strong>s,J. Chem. Ecol. 20 (1994) 1281–1328.[236] J. Gershenzon, The cost of <strong>plant</strong> chemical defense againstherbivores: a biochemical perspective, in: E.A. Bernays (Ed.),Insect–Plant Interactions, vol. 5, CRC Press, Boca Raton, 1994,pp. 105–173.[237] M. Dicke, M.W. Sabelis, Does it pay <strong>plant</strong>s to advertize forbodyguards? Towards a cost–benefit analysis of <strong>induced</strong> synomoneproduction, in: T.L. Pons (Ed.), Variation in Growth Rate andProductivity of Higher Plants, SPB Academic Publishing, TheHague, 1989, pp. 341–358.[238] M.W. Sabelis, M.C.M. de Jong, Should all <strong>plant</strong>s recruit bodyguards?Conditions for a polymorphic ESS of synomone productionin <strong>plant</strong>s, Oikos 53 (1988) 247–252.[239] H.C.J. Godfray, Communication between the first and third trophiclevels: an analysis using biological signalling theory, Oikos 72(1995) 367–374.[240] M. Heil, T. Delsinne, A. Hilpert, S. Schurkens, C. Andary, K.E.Linsenmair, M.S. Sousa, D. McKey, Reduced chemical defence inant–<strong>plant</strong>s? A critical re-evaluation of a widely accepted hypothesis,Oikos 99 (2002) 457–468.[241] J.A. Rudgers, S.Y. Strauss, J.E. Wendel, Trade-offs among antiherbivoreresistance traits: insights from Gossypieae (Malvaceae),Am. J. Bot. 91 (2004) 871–880.[242] M. Heil, C. Staehelin, D. McKey, Low chitinase activity in Acaciamyrmecophytes: a potential trade-off between biotic and chemical<strong>defences</strong>? Naturwissenschaften 87 (2000) 555–558.[243] M. Heil, B. Fiala, T. Boller, K.E. Linsenmair, Reduced chitinaseactivities in ant <strong>plant</strong>s of the genus Macaranga, Naturwissenschaften86 (1999) 146–149.[244] S. Koptur, Alternative defenses against herbivores in Inga(Fabaceae, Mimosoideae) over an elevational gradient, Ecology69 (1985) 278–283.[245] M. Heil, A. Hilpert, R. Krqger, K.E. Linsenmair, Competition amongvisitors to extrafloral nectaries as a source of ecological costs of an<strong>indirect</strong> defence, J. Trop. Ecol. 20 (2004) 201–208.[246] S. Koptur, J.H. Lawton, Interactions among vetches bearingextrafloral nectaries, their biotic protective agents, and herbivores,Ecology 69 (1988) 278–283.[247] P.J. Devries, I. Baker, Butterfly exploitation of an ant-<strong>plant</strong>mutualism adding insult to herbivory, J. N. Y. Entomol. Soc. 97(1989) 332–340.[248] A. Janssen, J.J.M. Vanalphen, M.W. Sabelis, K. Bakker, Odormediatedavoidance of competition in Drosophila parasitoids: theghost of competition, Oikos 73 (1995) 356–366.[249] A. Janssen, J.J.M. Vanalphen, M.W. Sabelis, K. Bakker, Specificityof odor mediated avoidance of competition in Drosophila Parasitoids,Behav. Ecol. Sociobiol. 36 (1995) 229–235.[250] A. Janssen, J. Bruin, G. Jacobs, R. Schraag, M.W. Sabelis, Predatorsuse volatiles to avoid prey patches with conspecifics, J. Anim. Ecol.66 (1997) 223–232.[251] C. Höller, S.G. Micha, S. Schulz, W. Francke, J.A. Pickett,Enemy-<strong>induced</strong> dispersal in a parasitic wasp, Experientia 50 (1994)182–185.[252] M.L. Bernasconi, T.C.J. Turlings, L. Ambrosetti, P. Bassetti, S. Dorn,<strong>Herbivore</strong>-<strong>induced</strong> emissions of maize volatiles repel the cornleaf aphid, Rhopalosiphum maidis, Entomol. Exp. Appl. 87 (1998)133–142.[253] C.J. Bolter, M. Dicke, J.J.A. vanLoon, J.H. Visser, M.A. Posthumus,Attraction of Colorado potato beetle to herbivore-damaged <strong>plant</strong>sduring herbivory and after its termination, J. Chem. Ecol. 23 (1997)1003–1023.[254] Y. Inui, Y. Miyamoto, T. Ohgushi, Comparison of volatile leafcompounds and herbivorous insect communities on three willowspecies, Popul. Ecol. 45 (2003) 41–46.[255] V. Stanjek, C. Herhaus, U. Ritgen, W. Boland, E. St7dler, Changes inthe leaf surface chemistry of Apium graveolens (Apiaceae) stimulatedby jasmonic acid and perceived by a specialist insect, Helv.Chim. Acta 80 (1997) 1408–1420.[256] J. Ruther, A. Reinecke, M. Hilker, Plant volatiles in the sexualcommunication of Melolontha hippocastani: response towards timedependentbouquets and novel function of (Z)-3-hexen-1-ol as asexual kairomone, Ecol. Entomol. 27 (2002) 76–83.[257] J. Horiuchi, G. Arimura, R. Ozawa, T. Shimoda, J. Takabayashi,T. Nishioka, A comparison of the responses of Tetranychusurticae (Acari: Tetranychidae) and Phytoseiulus persimilis (Acari:Phytoseiidae) to volatiles emitted from lima bean leaves withdifferent levels of damage made by T. urticae or Spodopteraexigua (Lepidoptera: Noctuidae), Appl. Entomol. Zool. 38 (2003)109–116.[258] W. Völkl, D.J. Sullivan, Foraging behaviour, host <strong>plant</strong> and hostlocation in the aphid hyperparasitoid Euneura augarus, Entomol.Exp. Appl. 97 (2000) 47–56.[259] R.M.P. van Poecke, M. Dicke, Indirect defence of <strong>plant</strong>s againstherbivores: using Arabidopsis thaliana as a model <strong>plant</strong>, Plant Biol.6 (2004) 387–401.[260] R. Karban, Fine-scale adaptation of herbivorous thrips to individualhost <strong>plant</strong>s, Nature 340 (1989) 60–61.[261] J.D. Fry, Evolutionary adaptation to host <strong>plant</strong>s in a laboratorypopulation of the phytophagous mite Tetranychus urticae Koch,Oecologia 81 (1989) 559–565.[262] A.A. Agrawal, Host-range evolution: adaptation and trade-offs infitness of mites on alternative hosts, Ecology 81 (2000) 500–508.[263] S.S. Duffey, Sequestration of <strong>plant</strong> natural-products by insects,Annu. Rev. Entomol. 25 (1980) 447–477.[264] R. Nishida, Sequestration of defensive substances from <strong>plant</strong>s byLepidoptera, Annu. Rev. Entomol. 47 (2002) 57–92.[265] D.E. Dussourd, T. Eisner, Vein-cutting behavior insect counterployto the latex defense of <strong>plant</strong>s, Science 237 (1987) 898–901.[266] D.E. Dussourd, R.F. Denno, Deactivation of <strong>plant</strong> defense —correspondence between insect behavior and secretory canalarchitecture, Ecology 72 (1991) 1383–1396.[267] I.T. Baldwin, An ecologically motivated analysis of <strong>plant</strong>–herbivore interactions in native tobacco, Plant Physiol. 127(2001) 1449–1458.[268] I.T. Baldwin, C.A. Preston, The eco-physiological complexityof <strong>plant</strong> responses to insect herbivores, Planta 208 (1999)137–145.[269] R.E. Gary, W.A. Foster, Anopheles gambiae feeding and survival onhoneydew and extra-floral nectar of peridomestic <strong>plant</strong>s, Med. Vet.Entomol. 18 (2004) 102–107.[270] D. Gnanvossou, R. Hanna, M. Dicke, S.J. Yaninek, Attraction of thepredatory mites Typhlodromalus manihoti and Typhlodromalus aripoto cassava <strong>plant</strong>s infested by cassava green mite, Entomol. Exp.Appl. 101 (2001) 291–298.[271] S. Colazza, J.S. McElfresh, J.G. Millar, Identification of volatilesynomones, <strong>induced</strong> by Nezara viridula feeding and oviposition onbean spp., that attract the egg parasitoid Trissolcus basalis, J. Chem.Ecol. 30 (2004) 945–964.[272] J. Figier, Fine structure in extrafloral nectary of Vicia faba L., Planta98 (1971) 31–49.


G. Arimura et al. / Biochimica et Biophysica Acta 1734 (2005) 91–111 111[273] J. Kuo, J.S. Pate, The extrafloral nectaries of cowpea (Vignaunguiculata (L.) Walp.): I. Morphology anatomy and fine structure,Planta 166 (1985) 15–27.[274] C.E.M. van den Boom, T.A. van Beek, M.A. Posthumus, A. deGroot, M. Dicke, Qualitative and quantitative variation amongvolatile profiles <strong>induced</strong> by Tetranychus urticae feeding on <strong>plant</strong>sfrom various families, J. Chem. Ecol. 30 (2004) 69–89.[275] T.M. Bezemer, R. Wagenaar, N.M. van Dam, W.H. van der Putten,F.L. W7ckers, Above- and below-ground terpenoid aldehydeinduction in cotton, Gossypium herbaceum, following root and leafinjury, J. Chem. Ecol. 30 (2004) 53–67.[276] J.H. Loughrin, A. Manukian, R.R. Heath, J.H. Tumlinson, Volatilesemitted by different cotton varieties damaged by feeding beetarmyworm larvae, J. Chem. Ecol. 21 (1995) 1217–1227.[277] P.J. Landolt, Effects of host <strong>plant</strong> leaf damage on cabbagelooper moth attraction and oviposition, Entomol. Exp. Appl. 67(1993) 79–85.[278] P.J. McCall, T.C.J. Turlings, J. Loughrin, A.T. Proveaux, J.H.Tumlinson, <strong>Herbivore</strong>-<strong>induced</strong> volatile emissions from cotton(Gossypium hirsutum L.) seedlings, J. Chem. Ecol. 20 (1994)3039–3050.[279] D. Tilman, Cherries, ants and tent caterpillars — timing of nectarproduction in relation to susceptibility of caterpillars to ant predation,Ecology 59 (1978) 686–692.[280] J.D. Curtis, N.R. Lersten, Morphology seasonal variation andfunction of resin glands on buds and leaves of Populus deltoidesSalicaceae, Am. J. Bot. 61 (1974) 835–845.[281] W. Trelease, The foliar nectar gland of Populus, Bot. Gaz. 6 (1881)284–290.[282] D.D. Schmidt, A. Kessler, D. Kessler, S. Schmidt, M. Lim, K. Gase,I.T. Baldwin, Solanum nigrum: a model ecological expressionsystem and its tools, Mol. Ecol. 13 (2004) 981–995.[283] K.H. Keeler, Species with extrafloral nectaries in a temperate flora(Nebraska), Prairie Nat. 11 (1979) 33–38.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!