Floral volatiles: from biosynthesis to function1

Accepted Article

1 2 3

Joëlle K. Muhlemann, Antje Klempien, Natalia Dudareva

4 5

Department of Biochemistry, Purdue University, 175 South University Street, West

6

Lafayette, IN 47907, USA

7

Author for correspondence:

8

Natalia Dudareva

9

Tel: +1 765 494 1325

10

Email: [email protected]

11 12

Keywords: floral scent; volatile organic compounds; terpenoids; phenylpropanoids;

13

benzenoids; regulation; pollination; florivory

14

This article has been accepted for publication and undergone full peer review but has not been through the copyediting, typesetting, pagination and proofreading process, which may lead to differences between this version and the Version of Record. Please cite this article as doi: 10.1111/pce.12314 1 This article is protected by copyright. All rights reserved.

Accepted Article

15 16

Abstract

17

Floral volatiles have attracted humans’ attention since antiquity and have since

18

then permeated many aspects of our lives. Indeed, they are heavily used in perfumes,

19

cosmetics, flavorings and medicinal applications. However, their primary function is to

20

mediate ecological interactions between flowers and a diverse array of visitors, including

21

pollinators, florivores and pathogens. As such, they ultimately ensure the plants’

22

reproductive and evolutionary success. To date, over 1,700 floral volatile organic

23

compounds have been identified. Interestingly, they are derived from only a few

24

biochemical networks, which include the terpenoid, phenylpropanoid/benzenoid and fatty

25

acid biosynthetic pathways. These pathways are intricately regulated by endogenous

26

and external factors to enable spatially and temporally controlled emission of floral

27

volatiles, thereby fine-tuning the ecological interactions facilitated by floral volatiles. In

28

this review, we will focus on describing the biosynthetic pathways leading to floral volatile

29

organic compounds (VOCs), the regulation of floral volatile emission, as well as

30

biological functions of emitted volatiles.

31 32 33

Introduction Plants are sessile organisms that need to constantly adapt to changing

34

environments for their survival and reproduction. For this environmental adaptation,

35

plants have evolved a wide array of specialized metabolites, also called plant secondary

36

metabolites or plant natural products. To date, over 200,000 specialized metabolites

37

have been described (Dixon & Strack 2003), out of which approximately 1% corresponds

38

to floral VOCs identified in 90 different angio- and gymnosperms families (Knudsen et al.

39

2006). VOCs are lipophilic liquids with low molecular weight and high vapor pressure at

40

ambient temperatures. Physical properties of these compounds allow them to freely

41

cross cellular membranes and be released into the surrounding environment (Pichersky,

42

Noel & Dudareva 2006). Biosynthesis of VOCs occurs in all plant organs: roots, stems,

43

leaves, fruits, seeds, as well as flowers, which were found to release the highest

44

amounts and diversity of VOCs. In contrast to VOCs released from other plant organs,

45

which are exclusively involved in plant defense, floral VOCs assume functions in both

46

attraction of pollinators and defense against florivores and pathogens. Based on their

47

biosynthetic origin, floral VOCs can be divided into three major classes: terpenoids,

48

phenylpropanoids/benzenoids, fatty acid derivatives (Fig. 1). In addition, sulfur- and

2 This article is protected by copyright. All rights reserved.

nitrogen-containing compounds contribute to the attraction of pollinators to flowers by

50

mimicking food or brood sources such as carrion or dung (Wiens, 1978; Faegri & Van

51

der Pijl, 1979; Juergens et al., 2006). However, to date little is known about the

52

biosynthetic pathways leading to the formation of these compounds.

Accepted Article

49

53 54

Biosynthetic Pathways and Genes Involved in the Formation of Floral Volatiles

55 56 57

Biosynthesis of Terpenoid Compounds Terpenoids are the largest class of floral volatiles and encompass 556 scent

58

compounds, which are derived from two common interconvertible five-carbon (C5)

59

precursors, isopentenyl diphosphate (IPP) and its allylic isomer dimethylallyl diphosphate

60

(DMAPP) (McGarvey & Croteau 1995). In plants, these C5-precursors are synthesized

61

from two independent and compartmentally separated pathways, the mevalonic acid

62

(MVA) and the methylerythritol-phosphate (MEP) pathways, which contribute to

63

terpenoid biosynthesis in a species- and/or organ-specific manner (Vranova, Coman &

64

Gruissem 2013). The MEP pathway operates in plastids (Hsieh et al. 2008) and is mainly

65

responsible for the formation of volatile mono- (C10) and diterpenes (C20) (~53% and

66

~1% of total floral terpenoids, respectively) (Knudsen & Gershenzon 2006), whereas the

67

MVA pathway is distributed between the cytosol, endoplasmatic reticulum and

68

peroxisomes (Simkin et al. 2011; Pulido, Perello & Rodriguez-Concepcion 2012) and

69

gives rise to precursors for volatile sesquiterpenes (C15) (~28% of total floral terpenoids).

70

While being compartmentally separated, these isoprenoid biosynthetic pathways are

71

connected via a metabolic “crosstalk” mediated by yet unidentified transporter(s) (Bick &

72

Lange 2003; Flügge & Gao 2005). Such connectivity of the pathways allows the MEP

73

pathway, often with a higher carbon flux than the MVA route, to support biosynthesis of

74

cytosolically formed terpenoids as was demonstrated in vegetative tissue (Laule et al.

75

2003; Ward et al. 2011), fruits (Gutensohn, Nagegowda & Dudareva 2013) and flowers

76

(Laule et al. 2003; Dudareva et al. 2005; Ward et al. 2011). Indeed, the MEP pathway

77

alone supports sesquiterpene biosynthesis in snapdragon flowers (Dudareva et al. 2005).

78

Terpenoid research in flowers has predominantly focused on the isolation and

79

characterization of terpene synthase (TPS) genes responsible for the final steps in

80

terpenoid biosynthesis, while genes and cognate enzymes of the MVA and MEP

81

pathways were mainly characterized from vegetative tissues (Cane 1999; Wise &

82

Croteau 1999; Lange et al. 2000; Rohdich et al. 2003; Guirimand et al. 2012) with 3 This article is protected by copyright. All rights reserved.

several excellent reviews were devoted to this subject (McGarvey & Croteau 1995;

84

Chappell 2002; Vranova et al. 2013). In brief, the MVA pathway starts from a stepwise

85

condensation of three molecules of acetyl-CoA and consists of six enzymatic reactions

86

while the MEP pathway begins with the condensation of D-glyceraldehyde 3-phosphate

87

and pyruvate and involves seven enzymatic reactions.

Accepted Article

83

88

Volatile terpenoids are synthesized from prenyl diphosphate precursors, which

89

are produced from condensation of IPP and DMAPP by prenyltransferases. Sequential

90

head-to-tail condensation of two IPP and one DMAPP molecules by farnesyl

91

diphosphate (FPP) synthase in the cytosol leads to the formation of FPP, the precursor

92

for sesquiterpenes (Fig. 2). Head-to-tail condensation of one DMAPP with one IPP

93

molecule in plastids results in GPP formation, the precursor of monoterpenes, and is

94

catalyzed by the GPP synthase (Fig. 2). This enzyme was found to be heterodimeric

95

in Antirrhinum majus (snapdragon) and Clarkia breweri, both of which have a floral scent

96

bouquet rich in monoterpene compounds (Tholl et al. 2004). Analyses of tissue-specific,

97

developmental, and rhythmic expression of the GPPS small subunit showed positive

98

correlation between expression and monoterpene emission in snapdragon flowers (Tholl

99

et al. 2004), whereas no such correlation was found for the large subunit, suggesting

100

that the small subunit is responsible for the regulation of GPP and subsequently

101

monoterpene formation. Interestingly, a homodimeric GPPS with dual prenyltransferase

102

activity (GPP synthase and FPP synthase activities) was reported in the orchid

103

Phalaenopsis bellina and demonstrated to be linked to the emission of linalool and

104

geraniol (Hsiao et al. 2008).

105

FPP and GPP serve as substrates for terpene synthases and cyclases (Cane

106

1999; Wise & Croteau 1999), which in plants are responsible for the production of a vast

107

variety of volatile terpenoid compounds (Fig. 2). TPSs are highly diversified throughout

108

the plant kingdom and form a mid-size gene family (Bohlmann, Meyer-Gauen & Croteau

109

1998; Chen et al. 2011a), which is comprised of more than 100 genes identified in a

110

variety of plant species, with one-third being isolated from flowers or fruits. Almost half of

111

the known TPSs are capable of synthesizing multiple products from a single prenyl

112

diphosphate precursor (Degenhardt, Köllner & Gershenzon 2009). For instance, the

113

floral volatile blend of Arabidopsis consists of 20 different sesquiterpenes, almost all of

114

which are synthesized by only two sesquiterpene synthases, TPS11 and TPS21 (Tholl et

115

al. 2005). The same is true for the flowers of kiwifruit (Actinidia deliciosa), where almost

116

all the sesquiterpenes released from flowers are the products of either germacrene D 4 This article is protected by copyright. All rights reserved.

synthase1 (AdGDS1) or α-farnesene synthase1 (AdAFS1) (Nieuwenhuizen et al. 2009).

118

In addition, some TPSs exhibit substrate promiscuity resulting in formation of different

119

products. However, in the case of these TSPs their subcellular localization and the

120

availability of a particular substrate determine the type of product formed (Tholl 2006;

121

Nagegowda et al. 2008). In addition, the diversity of formed volatile terpenoids is not

122

only dependent on TPSs, but is also increased by enzymes modifying TPS products by

123

hydroxylation, dehydrogenation, and acylation, which enhance their volatility and

124

olfactory properties (Dudareva, Pichersky & Gershenzon 2004).

Accepted Article

117

125

To date, multiple flower-specific TPSs have been isolated and characterized (see

126

Table 1). They were shown to be responsible for the formation of the monoterpenes

127

linalool (Clarkia breweri, Antirrhinum majus, and Arabidopsis thaliana) (Dudareva et al.

128

1996a; Nagegowda et al. 2008; Ginglinger et al. 2013), E-(β)-ocimene (Antirrhinum

129

majus and Citrus unshiu) (Dudareva et al. 2003; Shimada et al. 2005), myrcene

130

(Antirrhinum majus and Alstromeria peruviana) (Dudareva et al. 2003; Aros et al. 2012)

131

and 1,8-cineole (Nicotiana suaveolens and Citrus unshiu) (Shimada et al. 2005; Roeder

132

et al. 2007), as well as of the sesquiterpenes nerolidol (Antirrhinum majus and Actinidia

133

chinensis) (Nagegowda et al. 2008; Green et al. 2012), α-farnesene (Actinidia deliciosa)

134

(Nieuwenhuizen et al. 2009), germacrene D (Actinidia deliciosa, Rosa hybrid and Vitis

135

vinifera) (Guterman et al. 2002; Lucker, Bowen & Bohlmann 2004; Nieuwenhuizen et al.

136

2009) and valencene (Vitis vinifera) (Lucker et al. 2004).

137

Besides mono- and sesquiterpenes, certain flowers also emit irregular terpenoids

138

(C8 to C18). They constitute a minor class of floral terpenoids (~7% of all floral

139

terpenoids), which are formed via a three-step modification including a dioxygenase

140

cleavage, enzymatic transformation and acid-catalyzed conversion into volatile

141

compounds (Winterhalter & Rouseff 2001). Interestingly, the dioxygenase cleavage step

142

itself can already result in volatile products, such as α- and β-ionone, geranylacetone

143

and pseudoionone, as was found to be the case in petunia flowers (Simkin et al. 2004).

144 145 146

Biosynthesis of phenylpropanoid/benzenoid compounds Phenylpropanoids and benzenoids represent the second largest class of plant

147

VOCs (Knudsen et al. 2006) and are exclusively derived from the aromatic amino acid

148

phenylalanine (Phe) (Fig. 3), which is synthesized via two alternative pathways (Maeda

149

et al. 2010; Maeda, Yoo & Dudareva 2011; Maeda & Dudareva 2012; Yoo et al. 2013).

150

Depending on the structure of their carbon skeleton, this class is divided into three 5 This article is protected by copyright. All rights reserved.

subclasses: phenylpropanoids (with a C6-C3 backbone), benzenoids (C6-C1) and

152

phenylpropanoid-related compounds (C6-C2).

Accepted Article

151 153

Phenylpropanoid-related compounds originate directly from Phe and constitute

154

approximately 24% of all described phenylpropanoid/benzenoid compounds (Knudsen &

155

Gershenzon 2006). So far, only genes and enzymes involved in the biosynthesis of

156

phenylacetaldehyde and 2-phenylethanol have been isolated and characterized

157

(Kaminaga et al. 2006; Sakai et al. 2007; Farhi et al. 2010; Chen et al. 2011b;

158

Gutensohn et al. 2011; Hirata et al. 2012) (Fig. 3). In petunia petals, phenylacetaldehyde

159

is produced via an unusual combined decarboxylation-amine oxidation reaction

160

catalyzed by phenylacetaldehyde synthase (Kaminaga et al. 2006). In roses, however, it

161

is formed via two alternative routes: the first involves a phenylacetaldehyde synthase

162

similar to the one described in petunia, while the second route employs Phe deamination

163

by an aromatic amino acid aminotransferase followed by decarboxylation of the formed

164

phenylpyruvate intermediate (Sakai et al. 2007; Farhi et al. 2010; Hirata et al. 2012).

165

Further conversion of phenylacetaldehyde to 2-phenylethanol is catalyzed by a

166

phenylacetaldehyde reductase as was shown in roses (Sakai et al. 2007; Chen et al.

167

2011b).

168

The first committed step in benzenoid (C6-C1) and phenylpropanoid (C6-C3)

169

biosynthesis is catalyzed by a well-characterized and widely distributed enzyme, L-

170

phenylalanine ammonia-lyase (PAL), which deaminates Phe to trans-cinnamic acid (CA)

171

and competes with phenylacetaldehyde synthase for Phe utilization (Fig. 3). Benzenoid

172

formation from CA involves shortening of the propyl-side chain by a C2 unit and was

173

shown to proceed via a β-oxidative, a non-β-oxidative pathway or a combination of both

174

(Boatright et al. 2004; Orlova et al. 2006) (Fig. 3). The β-oxidative pathway has only

175

recently been fully elucidated in petunia flowers and appears to be analogous to fatty

176

acid catabolism and is localized in peroxisomes. The pathway begins with an activation

177

of CA to cinnamoyl-CoA, followed by hydration, oxidation and cleavage of the β-keto

178

thioester with subsequent formation of benzoyl-CoA (Van Moerkercke et al. 2009;

179

Klempien et al. 2012; Qualley et al. 2012). Benzaldehyde acts as the key intermediate in

180

the alternative non-β-oxidative pathway and is oxidized to benzoic acid by a NAD+-

181

dependent benzaldehyde dehydrogenase, which has been isolated and characterized

182

from snapdragon flowers (Long et al. 2009). However, the enzymatic reactions leading to

183

benzaldehyde formation remain unknown.

6 This article is protected by copyright. All rights reserved.

Accepted Article

184

Formation of floral phenylpropanoids (C6-C3), including (iso)eugenol and

185

methyl(iso)eugenol, shares the initial biosynthetic steps with the lignin biosynthetic

186

pathway up to the coniferyl alcohol stage. This monolignol precursor then undergoes two

187

enzymatic reactions that eliminate the oxygen functionality at the C9 position. The first

188

reaction involves acetylation by an acyltransferase from the BAHD superfamily as was

189

shown for the formation of coniferyl acetate from coniferyl alcohol in petunia petals

190

(Dexter et al. 2007). Coniferyl acetate is then converted to the phenylpropanoids

191

eugenol and isoeugenol by eugenol and isoeugenol synthases, respectively, which both

192

belong to the PIP family of NADPH-dependent reductases (Koeduka et al. 2006;

193

Koeduka et al. 2008) (Fig. 3).

194

In flowers the diversity of phenylpropanoid/benzenoid compounds is further

195

increased by modifications such as methylation, hydroxylation and acetylation of direct

196

scent precursors. These modifications enhance the volatility or olfactory properties of

197

scent compounds. Methylation reactions are catalyzed by either O-methyltransferases or

198

carboxyl methyltransferases. O-methyltransferases were shown to be responsible for the

199

synthesis of a diverse array of benzenoids/phenylpropanoids, including veratrole in

200

Silene flowers (Akhtar & Pichersky ; Gupta et al. 2012), 3,5-dimethoxytoluene and 1,3,5-

201

trimethoxybenzene in roses (Lavid et al. 2002; Scalliet et al. 2002), methyleugenol and

202

isomethyleugenol in Clarkia (Wang & Pichersky 1998). Carboxyl methyltransferases,

203

many of which belong to the SABATH family (D'Auria et al. 2003), are involved in the

204

biosynthesis of volatile esters like methylbenzoate in snapdragon and petunia flowers

205

(Murfitt et al. 2000; Negre et al. 2003) and methylsalicylate in Clarkia and petunia (Ross

206

et al. 1999; Negre et al. 2003). Enzymes from the BAHD superfamily of acyltransferases

207

(D'Auria 2006) were shown to be responsible for the biosynthesis of acetylated scent

208

compounds such as benzylacetate in Clarkia (Dudareva et al. 1998), benzoylbenzoate in

209

Clarkia and petunia (D'Auria, Chen & Pichersky 2002; Boatright et al. 2004; Orlova et al.

210

2006) and phenylethylbenzoate in petunia flowers (Boatright et al. 2004; Orlova et al.

211

2006).

212 213 214

Biosynthesis of volatile fatty acid derivatives Fatty acid derivatives constitute the third class of flower VOCs, which derive from

215

the unsaturated C18 fatty acids linolenic and linoleic acid. Biosynthesis of volatile fatty

216

acid derivatives is initiated by a stereo-specific oxygenation of the octadecanoid

217

precursors, catalyzed by a lipoxygenase (LOX) and leads to formation of 9- and 137 This article is protected by copyright. All rights reserved.

hydroperoxy intermediates (Schaller 2001; Feussner & Wasternack 2002). These

219

intermediates can enter two different branches of the LOX pathway, which in turn leads

220

to the formation of volatile compounds. Allene oxide synthase (AOS) exclusively utilizes

221

the 13-hydroperoxy intermediate as substrate, converting it to an unstable epoxide,

222

which is then subjected to a cyclization followed by a reduction and a series of

223

cyclization reactions to yield jasmonic acid (JA). In contrast to the allene oxide synthase

224

branch, hydroperoxide lyase can convert both types of hydroperoxide fatty acid

225

derivatives into volatile C6 and C9 aldehydes. These saturated or unsaturated C6 and C9

226

aldehydes are often substrates for alcohol dehydrogenases giving rise to volatile

227

alcohols, which can be further converted to their esters. These C6 and C9 aldehydes and

228

alcohols are commonly referred to as green leaf volatiles, as they are usually

229

synthesized in vegetative tissues. However they are also important constituents in the

230

floral volatile bouquet of several plant species such as carnation and wild snapdragon

231

(Schade, Legge & Thompson 2001; Suchet et al. 2011).

Accepted Article

218

232

The orchids of the genus Ophrys produce an array of fatty acid derived volatiles as

233

well. Within their bouquet, alkenes are particularly important mediators in the interaction

234

between orchids and their pollinators. Production of alkenes requires desaturation of

235

fatty acids, a step that is likely mediated by acyl-acyl carrier protein (ACP) desaturases.

236

Two isoforms of a stearoyl-acyl carrier protein (ACP) desaturase (SAD), namely SAD1

237

and SAD2, were identified in Ophrys sphegodes and O. exaltata. However, only

238

expression of SAD2 was positively correlated with the formation of alkenes in flowers

239

(Schlüter et al. 2011). OsSAD2 is a functional desaturase capable of producing 18:1Δ9

240

(ω-9) and 16:1Δ4 (ω-12) fatty acid intermediates from which 9-alkenes and 12-alkenes

241

could be derived.

242 243

Regulation of floral volatile emission

244 245 246

Spatial, rhythmic and developmental regulation of floral scent emission Flowers have evolved many complex olfactory and visual guides for pollinator

247

attraction. In order to maximize pollinator attraction, floral scent emission is often

248

restricted to particular flower tissues and is developmentally and rhythmically regulated.

249

Tissue-specific emission of floral VOCs is a characteristic feature of many

250

species. In general, petals are the primary source of floral volatiles, although other

251

tissues (stamens, pistils, sepals and nectaries) also contribute to the floral bouquet in 8 This article is protected by copyright. All rights reserved.

certain plant species (Dobson, Bergström & Groth 1990; Bergström, Dobson & Groth

253

1995; Dobson, Groth & Bergström 1996; Flamini, Cioni & Morelli 2003; Dötterl & Jürgens

254

2005; Farré-Armengol et al., 2013). Emitted from petals, VOCs often enable long-

255

distance attraction of pollinators, while VOCs produced in nectaries or pollen signal

256

availability of food sources. Tissue-specificity of scent emission is regulated at the level

257

of scent biosynthetic gene expression and enzyme activity. Indeed, many of the scent

258

biosynthetic genes isolated so far show a very specific expression profile, with the

259

highest level found in the scent producing parts of the flower (see e.g. Dudareva et al.

260

1996a; Murfitt et al. 2000; Dudareva et al. 2003; Negre et al. 2003; Nagegowda et al.

261

2008; Rodriguez-Saona et al. 2011). Within scent-emitting tissues, formation of VOCs is

262

often restricted to specific cell types or layers. In snapdragon flowers for example,

263

biosynthesis of the major volatile benzenoid compound methyl benzoate is restricted to

264

the inner epidermal layer of the upper and lower petal lobes (Kolosova et al. 2001b).

265

Similar cell-specific expression of scent biosynthetic genes was also reported for roses

266

and Clarkia breweri (Dudareva et al. 1996b; Bergougnoux et al. 2007).

Accepted Article

252

267

Rhythmicity of floral scent emission has been shown to occur in numerous

268

species and often correlates with the activity of the respective pollinators (see e.g.

269

(Raguso et al. 2003; Dötterl, Wolfe & Jürgens 2005; Effmert et al. 2005; Hoballah et al.

270

2005; Rodriguez-Saona et al. 2011). Rhythmic emission allows plants to conserve

271

valuable carbon and energy during times of the day when their primary pollinators are

272

inactive. Different modes of rhythmic scent release have been described so far. Diurnal

273

rhythmicity in floral VOC emission was observed in plants pollinated during the day,

274

whereas nocturnally emitting plants are visited by pollinators foraging at night (Kolosova

275

et al. 2001a; Waelti et al. 2008). Interestingly, the total amounts of emitted VOCs do not

276

change over the day/night cycle in Dianthus inoxianus, however the levels of compounds

277

contributing to pollinator attraction vary according to visitor activity (Balao et al. 2011).

278

Rhythmicity of scent emission is often transcriptionally regulated, similarly to its tissue-

279

specificity (see e.g. Kolosova et al. 2001a; Hendel-Rahmanim et al. 2007; Nagegowda et

280

al. 2008; Nieuwenhuizen et al. 2009), although substrate availability for scent

281

biosynthetic enzymes was shown to play a regulatory role in the emission of some

282

compounds as well (Kolosova et al. 2001a).

283

In addition to being spatially and rhythmically regulated, floral scent emission

284

often changes over the lifespan of flowers. Usually, emission levels are highest when

285

flowers are ready for pollination, i.e. when anthers are dehisced, and decrease during 9 This article is protected by copyright. All rights reserved.

senescence. Once pollinated, single flowers change or reduce the level of produced

287

volatiles to prevent further visits potentially damaging the flower and to redirect visitors to

288

the remaining unpollinated flowers (Schiestl & Ayasse, 2001; Negre et al., 2003;

289

Muhlemann et al., 2006; Rodriguez-Saona et al., 2011). Developmental regulation of

290

scent emission occurs at several levels, including orchestrated expression of scent

291

biosynthetic genes (Colquhoun et al. 2010), enzyme activities (see e.g. Pichersky,

292

Lewinsohn & Croteau 1995; Dudareva et al. 2000; Shalit et al. 2003; Boatright et al.

293

2004; Nagegowda et al. 2008) and substrate availability (Dudareva et al. 2000).

Accepted Article

286

294 295 296

Transcriptional network controlling volatile emission in flowers Orchestrated formation of volatiles from several independent pathways is not

297

only a function of biochemical properties of biosynthetic enzymes, but also requires the

298

involvement of transcription factors (TFs). Indeed, coordinated transcriptional regulation

299

of entire scent biosynthetic networks has been recently shown (Colquhoun & Clark 2011;

300

Muhlemann et al. 2012), implying that transcription factors control scent emission.

301

Despite their importance, only a few TFs regulating the expression of scent

302

biosynthetic genes have been identified to date. TFs controlling the flux through the

303

phenylpropanoid/ benzenoid network have recently been isolated from petunia flowers.

304

ODORANT1 (ODO1), a R2R3-type MYB TF, is exclusively expressed in petunia petal

305

tissue and regulates the transcription of a major portion of the shikimate pathway as well

306

as entry points into both the Phe (i.e., chorismate mutase) and phenylpropanoid (i.e.,

307

phenylalanine ammonia-lyase) branchways (Verdonk et al. 2005). ODO1 was also found

308

to activate the promoter of an ABC transporter of unknown function localized at the

309

plasma membrane (Van Moerkercke et al. 2012a). In petunia flowers, ODO1 is positively

310

regulated by another R2R3-type MYB TF, EMISSION OF BENZENOIDS II (EOBII),

311

which also activates the promoter of the biosynthetic gene isoeugenol synthase (Spitzer-

312

Rimon et al. 2010; Colquhoun et al. 2011b; Van Moerkercke, Haring & Schuurink 2011).

313

The recently identified petunia EOBI was shown to be a flower-specific R2R3-type TF,

314

which acts downstream of EOBII and upstream of ODO1 (Spitzer-Rimon et al. 2012; Van

315

Moerkercke, Haring & Schuurink 2011). Silencing of EOBI expression lead to down-

316

regulation of numerous genes in the shikimate pathway (5-enolpyruvylshikimate-3-

317

phosphate synthase, 3-deoxy-D-arabinoheptulosonate 7-phosphate synthase,

318

chorismate synthase, chorismate mutase, arogenate dehydratase, and prephenate

319

aminotransferase), as well as downstream scent-related genes (PAL, isoeugenol 10 This article is protected by copyright. All rights reserved.

synthase, and benzoic acid/salicylic acid carboxyl methyltransferase) (Spitzer-Rimon et

321

al. 2012). In contrast to ODO1, EOBI and EOBII, the MYB4 TF was found to be a

322

repressor of only a single enzyme in the phenylpropanoid pathway, cinnamate-4-

323

hydroxylase, thus controlling the flux toward phenylpropanoid volatile compounds in

324

petunia flowers (Colquhoun et al. 2011a).

Accepted Article

320

325

While several transcription factors regulating the phenylpropanoid/benzenoid

326

network have been isolated and characterized, transcriptional regulation of the terpenoid

327

pathways remains elusive. MYC2, a basic helix-loop-helix TF, was recently identified in

328

Arabidopsis inflorescences and shown to activate the expression of two sesquiterpene

329

synthase genes TPS11 and TPS21 via the gibberellic and jasmonic acid signaling

330

pathways (Hong et al. 2012). Despite the identification of several TFs for individual

331

pathways, master regulators, which orchestrate formation of diverse volatile blends and

332

act upstream of multiple metabolic pathways, are yet to be discovered. Recently,

333

upregulation of terpenoid and phenylpropanoid pathways was achieved by

334

overexpression of the Production of Anthocyanin Pigment1 TF in roses (Zvi et al. 2012).

335

However, it remains unknown whether promoters of genes involved in terpenoid

336

formation are the natural targets for this TF.

337 338 339

Changes in floral volatile emission upon diverse biotic interactions In addition to internal regulatory mechanisms, several external factors are known

340

to influence composition, quantity and timing of volatile emission. These factors include

341

pollination and interactions with floral antagonists. Successful pollination leads to a

342

decrease or alterations in floral VOC emission in a variety of plant species (Tollsten &

343

Bergström 1989; Tollsten 1993; Schiestl & Ayasse 2001; Negre et al. 2003; Theis &

344

Raguso 2005; Muhlemann et al. 2006; Rodriguez-Saona et al. 2011). Besides reducing

345

carbon loss caused by emission, post-pollination changes in VOCs redirect pollinators to

346

yet unpollinated flowers (Schiestl & Ayasse 2001) or prevent additional visitation of

347

pollinated flowers by flower antagonists (Muhlemann et al. 2006). Post-pollination

348

decrease in emission was shown to occur within 24-96 hours of pollinator visitation

349

(depending on the plant species) and appears to be triggered once the pollen tubes

350

reach the ovary (Negre et al. 2003). Upon fertilization, different molecular mechanisms

351

were found to trigger the decrease in scent emission. In petunia flowers, reduced

352

methylbenzoate emission after pollination is the result of transcriptional downregulation

353

of the cognate gene in an ethylene-dependent manner (Negre et al. 2003). In 11 This article is protected by copyright. All rights reserved.

snapdragon flowers, however, the post-pollination decrease in methylbenzoate emission

355

largely depends on reduced S-adenosyl-L-methionine:benzoic acid carboxyl

356

methyltransferase (BAMT) activity and substrate availability (Negre et al. 2003).

Accepted Article

354

357

Altered volatile emission was also reported in the context of above- and

358

belowground plant-fungus interactions. Infection of Silene latifolia flowers by the anther

359

smut fungus Microbotryum violaceum results in decreased total scent emission and

360

discrimination against infected flowers by the pollinator (Dötterl et al. 2009). The intensity

361

and chemodiversity of floral scent emission also decrease as a function of colonization

362

by arbuscular mycorrhizal fungi in Polemonium viscosum, suggesting that plant-microbe

363

interactions occurring outside of floral tissues can also modulate flower traits providing

364

an additional layer of external regulatory mechanism (Becklin et al. 2011).

365

Herbivore-induced plant volatiles (HIPVs) have been extensively studied in

366

vegetative tissues. Signal perception and transduction mechanisms, as well as kinetics

367

of HIPV release are well characterized in these tissues (Dicke & Baldwin 2010).

368

However, only few studies have linked florivory with induction of defense VOCs in

369

flowers. One of the few known examples of florivore-induced volatile emission is

370

represented by wild parsnip flowers, which emit higher amounts of octyl esters upon

371

infestation with the parsnip webworm (Zangerl & Berenbaum 2009). Likewise,

372

Helicoverpa zea larvae feeding on cotton flower buds induce emission of terpenes and

373

fatty acid derivatives (Rose & Tumlinson 2004).

374 375 376

Floral volatiles as mediators in biotic interactions Floral VOCs possess multifaceted functions significantly contributing to attraction

377

of pollinators and serving as defense compounds against pathogens and florivores.

378

Although it was proposed that floral VOCs first served in protecting reproductive

379

structures against antagonists and only later acquired pollinator attracting capacities

380

(Pellmyr & Thien 1986), their latter function has been studied the most. Pollinator

381

attraction is mostly mediated by benzenoids, whereas defense functions are

382

predominantly assured by terpenoid and benzenoid VOCs (Schiestl 2010).

383 384 385

Floral volatiles in pollinator attraction For countless cross-pollinating plant species, mating involves the movement of

386

pollen from one individual to another. In many cases, animals such as insects, birds and

387

mammals significantly contribute to pollination by serving as vectors in pollen transfer. 12 This article is protected by copyright. All rights reserved.

Flowers employ a diverse palette of signals to mediate attraction of pollinators to flowers

389

for ensuring successful reproduction. For pollinators, this multisensory (visual, olfactory,

390

thermal, electromagnetic) input is essential to locate food and breeding sites. Over the

391

last decade, mounting evidence was accumulated for the role of floral volatiles in plant-

392

pollinator communication.

Accepted Article

388

393

It has been shown that the information conveyed by floral volatiles depends on

394

amount, composition and context of their emission, and elicits distinct behavioral

395

responses in the respective pollinators. Long-distance emission of volatiles mainly

396

contributes to guiding pollinators to flowers and is especially important for night-emitting

397

plants where production of volatiles has to be of high intensity to overcome decreased

398

conspicuousness of flowers under low illumination. In fact, the moth-pollinated Petunia

399

axillaris and Silene latifolia emit higher amounts of volatiles than day-emitting bee

400

pollinated plants within the same genus, like Petunia integrifolia and Silene dioica (Ando

401

et al. 2001; Waelti et al. 2008). In contrast, volatiles emitted over short distances trigger

402

landing, feeding and reproductive behavior. Exposure even to a single floral volatile of

403

the host plant Silene latifolia elicited landing and feeding behavior in Hadena bicruris

404

moths (Dötterl et al. 2006) while subjection of nocturnal hawkmoth Manduca sexta to an

405

olfactory stimulus induced proboscis extension (Goyret, Markwell & Raguso 2007). Not

406

only volatiles emitted from flower petals but also pollen odor can contribute to pollinator

407

foraging. Indeed, pollen odor artificially added to antherless Rosa rugosa flowers

408

increased frequency of bumblebees’ pollen-collecting behavior relative to odor- and

409

antherless flowers (Dobson, Danielson & Van Wesep 1999).

410

Volatiles emitted from flowers not only advertise food availability, but also mating

411

and oviposition opportunities. Examples include certain orchids which employ flower

412

scent to imitate pheromone blends of female pollinators, thereby triggering copulation

413

attempts of male pollinators with flowers (Schiestl et al. 1999). The dioecious species

414

Silene latifolia represents another example where floral volatiles provide oviposition cues

415

for the females of the nursery moth pollinator Hadena bicruris, which lay eggs while

416

pollinating the flowers (Brantjes 1976; Waelti et al. 2009). To mimic oviposition sites, a

417

number of species within the five plant families (Araceae, Rafflesiaceae, Annonaceae,

418

Apocynaceae and Orchidaceae) emit sulfur-containing volatile compounds to attract

419

necrophagous, saprophagous and caprophagous insects. Interestingly, these volatiles,

420

typically released by decomposing plant and animal organic matter, were found to arise

421

independently via convergent evolution (Jürgens et al. 2013). These emitted volatiles 13 This article is protected by copyright. All rights reserved.

also represent a characteristic feature of plants pollinated by necrophagous flies and

423

beetles, thus showing a clear link between certain pollinator guilds and specific floral

424

scent chemistries.

Accepted Article

422

425

Attempts to correlate floral volatile profiles with pollination syndromes have

426

succeeded for certain plant-pollinator interactions, however, clear-cut predictions of

427

pollinator guilds based on floral scent chemistry still remain unattainable. It was

428

nevertheless established that plant species relying on moths for pollination

429

predominantly release benzenoid, terpenoid and nitrogen-containing compounds

430

(Knudsen & Tollsten 1993; Dobson 2006) and that bat-pollinated species mostly emit

431

sulfur-containing volatiles (von Helversen, Winkler & Bestmann 2000). While

432

involvement of volatiles in the attraction of hummingbirds remains debated, certain

433

ornithophilous plants were found to emit minute amounts of terpenoids and fatty acid-

434

derivatives (Knudsen et al. 2004). Furthermore, hummingbirds were shown to display

435

different behaviors (attraction or aversion) depending on the volatiles contained in the

436

nectar of artificial flowers (Kessler & Baldwin 2007; Kessler, Gase & Baldwin 2008).

437 438 439

Floral volatiles in flower defense Flowers are generally deficient in physical barriers such as a highly lignified cell

440

wall and/or an impermeable cuticle, making them highly susceptible to pathogens and

441

florivores. Moreover, they typically carry higher densities of microorganisms than other

442

aerial plant surfaces due to their high moisture and nutrient content (Johnson &

443

Stockwell 1998). Thus, flowers employ volatiles as an alternative mechanism that

444

prevents damage to their reproductive structures.

445

Many VOCs were shown to exhibit antimicrobial and antifungal activities in vitro

446

(Bakkali et al. 2008) or inferred to have these antimicrobial activities due to tissue-

447

specific expression patterns (e.g. in nectaries and/or stigmas) of their biosynthetic genes

448

(Dudareva et al. 1996a; Chen et al. 2003). However, only few VOCs have been

449

investigated for their role in defense against pathogens. (E)--caryophyllene emitted

450

from stigmas of Arabidopsis flowers was shown to limit bacterial growth. Indeed,

451

Arabidopsis plants lacking (E)--caryophyllene emission displayed denser bacterial

452

populations on their stigmas and reduced seed weight compared to wild type plants,

453

indicating that (E)--caryophyllene acts in the defense against pathogenic bacteria and is

454

important for plant fitness (Huang et al. 2012). VOCs emitted by Saponaria officinalis

14 This article is protected by copyright. All rights reserved.

petals were shown to inhibit bacterial growth and suggested to control diversity of

456

bacterial communities in petals (Junker et al. 2011b).

Accepted Article

455 457

Besides pathogens, florivores also cause substantial damage to reproductive

458

tissues. Florivores are detrimental to plant fitness as they feed on reproductive structures,

459

often displace potential pollinators and alter flower morphology (McCall 2008; Sõber,

460

Mooraa & Tederb 2010). Similarly to green tissue volatiles, floral volatiles are capable of

461

deterring insects that are detrimental to pollinator visitation and/or are feeding on flower

462

tissues. Ants are very inefficient pollinators (due to their morphology and mobility), often

463

exhibit aggressive behavior against pollinators and occasionally feed on floral structures.

464

It was shown that many common European ants are deterred by volatiles emitted from

465

temperate flowers (Willmer et al. 2009) and that inhibition of terpene biosynthesis leads

466

to loss of ant-repellent properties (Junker, Gershenzon & Unsicker 2011a). The

467

facultative florivore Metrioptera bicolor displays a strong aversion to linalool and (E)--

468

caryophyllene, both of which are widespread terpenoid constituents of scent bouquets

469

emitted by flowering species (Junker, Heidinger & Blüthgen 2010). These examples

470

strongly suggest involvement of floral volatiles in deterrence against undesirable floral

471

visitors.

472 473

Balance between defensive and attractive functions of floral volatiles

474

Pollinators and florivores navigate the same visual and olfactory landscape to

475

locate host plants. To avoid visitation by florivores while advertising their flowers to

476

pollinators, plants have to balance attracting and deterring functions of floral volatiles.

477

Unbalanced production of volatiles involved in different functions may result in negative

478

impacts on plant fitness, as was shown in the cucurbit Cucurbita pepo where

479

enhancement of floral fragrance led to higher attraction of florivores significantly affecting

480

reproductive success (Theis & Adler 2012).

481

Several strategies employed by flowers to optimize these functions and therefore

482

maximize reproductive success have been described so far. In Cirsium arvense, some

483

floral volatiles are responsible for attraction of both pollinators and florivores (Theis

484

2006). In this species, timing of emission is fine-tuned to increase likelihood of pollinator

485

rather than florivore visitation. Indeed, developmental and diel timing of Cirsium arvense

486

volatile emission was found to be positively correlated with the flowers' reproductive

487

maturity and peak activity of pollinators, respectively, while negative correlation between

488

diel emission and activity of florivores was observed (Theis, Lerdau & Raguso 2007). 15 This article is protected by copyright. All rights reserved.

Similarly to Cirsium arvense, VOCs of petunia flowers attract both pollinators and

490

florivores. Within the petunia VOC profile some compounds specifically control

491

infestation rate by florivores (i.e. isoeugenol and benzylbenzoate), whereas

492

methylbenzoate, for example, is involved in pollinator attraction. Indeed, utilization of

493

various petunia transgenic lines with downregulation of different floral scent biosynthetic

494

genes allowed elucidation of the distinct roles of individual volatile compounds in

495

attraction of mutualists and deterrence of antagonists (Kessler et al. 2013). A similarly

496

complex picture was found in the cucurbit Cucurbita moschata, where some compounds

497

are attracting both pollinators and florivores, while other compounds are only mediating

498

one type of interaction (Andrews, Theis & Adler 2007). As a consequence of interactions

499

with both mutualists and antagonists, different types of selection may act on the different

500

volatile compounds. Indeed, compounds involved in attraction of both mutualists and

501

antagonists will more likely be under balancing selection, while compounds involved in

502

interactions solely with one type of flower visitor are under directional selection pressure.

503

This diversity of selection mechanisms is predicted to lead to the evolution of very

504

complex floral volatile profiles.

Accepted Article

489

505

Nursery pollination systems represent a special case where insects simultaneously

506

pollinate the flowers and use them as breeding sites. Flowers in these systems have

507

evolved adaptive mechanisms to reduce damage caused by developing larvae that feed

508

on the plant's reproductive tissues (usually seeds) (Dufaÿ & Anstett 2003). In general,

509

damage to seeds is predicted to be proportional to the number of visiting pollinators.

510

Therefore, fast cessation of floral advertisement after successful pollination is essential

511

to avoid further damage to the plant's reproductive success. Indeed, a post-pollination

512

decrease specifically in pollinator-attracting volatiles was observed in Silene, thereby

513

providing a mechanism to prevent further loss of seeds (Muhlemann et al. 2006).

514 515 516

Evolution of floral scent The evolution of angiosperms has resulted in an immense diversity of flower traits

517

such as shape, size, color and scent. To date, more than 1700 floral volatile compounds

518

have been described in over 900 flowering plant species (Knudsen et al. 2006).

519

Interestingly, the quality and quantity of emitted volatiles are species-specific and vary

520

among different populations of a given species (Raguso 2008). While much effort has so

521

far been invested in describing scent composition in various flowering species, the

522

mechanisms driving the evolution and diversification of floral scent remain underexplored. 16 This article is protected by copyright. All rights reserved.

Evolution of floral scent is potentially shaped by two factors that mutually influence each

524

other: 1) genomic changes allowing catalytic expansion and differential regulation of the

525

enzymatic machinery underlying floral scent formation and 2) ecological constraints such

526

as pollinator-mediated selection.

Accepted Article

523

527

To date, only a few studies have examined the genetic basis for odor differences

528

between closely related flowering species. Petunia axillaris and P. exserta represent a

529

good example of a closely related species with distinct pollination syndromes. While P.

530

axillaris flowers are colorless, emit benzenoid compounds and are moth-pollinated, P.

531

exserta flowers are red, devoid of scent and attract hummingbirds. Analysis of the

532

genetic basis for differences in scent profiles between these two species revealed that

533

only two quantitative trait loci are responsible for the distinct scent phenotypes (Klahre et

534

al. 2011). One of these loci maps to the MYB transcription factor ODO1, which controls

535

flux through the shikimate pathway and hence the amount of precursors available for

536

benzenoid biosynthesis (Verdonk et al. 2005), while the genetic identity of the second

537

locus is presently unknown. Clarkia breweri and C. concinna provide another example of

538

two closely related scented and non-scented species, which rely on different pollinators.

539

Although non-scented C. concinna contains genes responsible for formation of VOCs

540

(i.e., linalool, (iso)methyleugenol and benzylacetate), transcriptional and/or post-

541

transcriptional regulatory mechanisms lead to lack of their expression in flowers and

542

elimination of floral scent (Raguso & Pichersky 1995; Dudareva et al. 1996a; Cseke,

543

Dudareva & Pichersky 1998; Nam, Dudareva & Pichersky 1999). Differences in

544

transcriptional levels were also found in the flowers of the orchid genus Ophrys, which

545

emit a blend of fatty acid-derived volatiles (Schiestl et al. 1997; Schiestl et al. 1999;

546

Schiestl & Ayasse 2002). Within this blend, alkenes are the key components for the

547

attraction of pollinators to flowers and each Ophrys species emits a unique alkene profile,

548

resulting in reproductive isolation between species through attraction of distinct and

549

highly specific pollinators (Schiestl & Ayasse 2002). Ophrys sphegodes mostly emits

550

27:1Δ9 and 27:1Δ12 alkenes, while O. exaltata mainly produces a 25:1Δ7 alkene

551

(Schlüter et al. 2011). Formation of double bonds at position 9 and 12 of these alkenes

552

requires the action of a stearoyl-acyl carrier protein desaturase (SAD). Differences in

553

transcript levels, as well as changes in the tertiary structure of SAD2 between O.

554

sphegodes and O. exaltata were proposed as possible mechanisms underlying the

555

distinct alkene profiles and reproductive isolation (Schlüter et al. 2011). While large intra-

556

and interspecific variations in alkene profiles were detected within the Ophrys genus, 17 This article is protected by copyright. All rights reserved.

only limited genetic variation among species and populations was observed with

558

microsatellite markers (Mant, Peakall & Schiestl 2005). These findings suggest that

559

divergent, pollinator-mediated selection rather than genetic drift explains strong

560

differences in volatile profiles. Taken together, the above examples demonstrate that

561

small genetic variations can have large effects on floral scent chemistry and interactions

562

with pollinators.

Accepted Article

557

563

Besides genetic polymorphisms, selection by pollinators is also capable of driving

564

floral diversification and specialization. Pollinator-mediated selection is constrained by

565

the pollinator’s pre-existing preferences and sensory abilities, and can occur only when

566

the traits under selection are heritable and exhibit variation (Schiestl 2010; Schiestl &

567

Dötterl, 2012). Selection on floral scent by pollinators has been described in Penstemon

568

digitalis, which displays marked inter-population variation in its emitted floral VOCs.

569

Quantification of phenotypic selection by pollinators revealed that the monoterpene

570

linalool is a direct target of selection within this scent profile (Parachnowitsch, Raguso &

571

Kessler 2012). Several studies in other plant species have also uncovered volatile

572

compounds which are important for plant-pollinator interaction and thus serving as

573

potential targets for pollinator-mediated selection (see e.g. Dötterl et al. 2006;

574

Shuttleworth & Johnson 2010; Schiestl, Huber & Gomez 2011).

575

Interestingly, changes in floral scent through external manipulation (genetic or

576

chemical) or intrinsic processes (mutations or hybridization) were shown to drive

577

pollinator niche shifts, which are a proximate cause of floral and reproductive isolation.

578

Indeed, Shuttleworth and Johnson (Shuttleworth & Johnson 2010) have demonstrated

579

that the presence/absence of sulfur compounds within the bouquets of four Eucomis

580

species determines whether wasps or flies pollinate individual species. Also, creation of

581

more similar floral volatile profiles between Silene dioica and S. latifolia by artificial

582

manipulation of a single compound resulted in higher interspecific pollen transfer (Waelti

583

et al. 2008). Thus, the aforementioned studies show that floral scent is an important

584

factor in defining and maintaining species boundaries. They also demonstrate that loss

585

or generation of reproductive isolation can occur through relatively simple changes in

586

floral scent profiles. Evolution of reproductive isolation through changes in floral scent

587

chemistry can occur within very short time frames, as was illustrated in Ophrys

588

arachnitiformis x O. lupercalis hybrids (Vereecken, Cozzolino & Schiestl 2010). Indeed,

589

generation of novel floral volatile profiles in these hybrids led to attraction of a pollinator

18 This article is protected by copyright. All rights reserved.

species that does not pollinate any of the parent species (Vereecken et al. 2010),

591

thereby leading to rapid floral isolation.

Accepted Article

590 592 593 594

Conclusions and future perspectives Over the last two decades, the field of floral volatile research has acquired an

595

ever-increasing amount of knowledge on the functions and biosynthesis of floral scent.

596

Numerous floral volatile compounds have been identified to date from nearly 1,000 plant

597

species and their importance in mediating ecological interactions with floral mutualists

598

and antagonists has been highlighted in many plant species. Recent advances in the

599

isolation and characterization of genes and enzymes involved in different scent

600

biosynthetic pathways, as well as in the elucidation of regulatory networks controlling

601

these pathways, have also enhanced our understanding of how floral volatile

602

compounds are synthesized. Despite recent progress in floral volatile research, many

603

aspects of floral volatile function and biosynthesis remain largely unknown. In particular,

604

we still do not know how the majority of floral VOCs are synthesized, how their

605

orchestrated emission is regulated, and the specific roles of floral VOCs in large plant-

606

mutualist/antagonist interaction webs. We therefore anticipate that future research efforts

607

will focus on providing insights into these specific aspects of floral scent biology and

608

allow for development of defensive strategies as well as enhancement of yields in insect-

609

pollinated plants.

610

19 This article is protected by copyright. All rights reserved.

Figure Legends

613

Figure 1. Major volatile classes emitted by flowers. Based on their biosynthetic origin,

614

volatiles emitted by flowers can be grouped into one of the three major volatile classes:

615

terpenoids, phenylpropanoids/benzenoid and fatty acid derivatives. Each volatile class is

616

here represented by a few typical floral scent compounds.

Accepted Article

611 612

617 618

Figure 2. Schematic representation of terpenoid VOCs biosynthesis. Synthesis of

619

terpenoid VOCs occurs via the cytosolic mevalonic acid (MVA) and the plastidial

620

methylerythritol phosphate (MEP) pathways, the former giving rise to sesquiterpenes

621

and the latter to monoterpenes, diterpenes and volatile carotenoid derivatives. Cross-talk

622

between both pathways is facilitated by the export of isopentenyl pyrophosphate (IPP)

623

from the plastid to the cytosol. Stacked arrows represent multiple biosynthetic steps.

624

Volatile compounds are highlighted with a yellow background. Abbreviations: DMAPP,

625

dimethylallyl pyrophosphate; FPP, farnesyl pyrophosphate; FPPS, FPP synthase; G3P,

626

glyceraldehyde-3-phosphate; GGPP, geranylgeranyl pyrophosphate; GGPPS, GGPP

627

synthase; GPP, geranyl pyrophosphate; GPPS, GPP synthase; IPP, isopentenyl

628

pyrophosphate; TPS, terpene synthase.

629 630

Figure 3. Schematic representation of the VOC phenylpropanoid/benzenoid biosynthetic

631

pathway. Phenylpropanoid/benzenoid VOCs are derived from phenylalanine, which itself

632

is synthesized via the shikimate/phenylalanine biosynthetic pathways. Benzoic acid is

633

the central precursor of various benzenoid VOCs and is synthesized via two biosynthetic

634

routes, the beta-oxidative pathway (orange background) and non-oxidative route.

635

Stacked arrows indicate multiple enzymatic reactions. Volatile compounds are

636

highlighted with a yellow background. Abbreviations: AAAT, aromatic amino acid

637

aminotransferase; BA, benzoic acid; BA-CoA, benzoyl-CoA; BAlc, benzylalcohol; BAld,

638

benzaldehyde; BALDH, benzaldehyde dehydrogenase; BB, benzylbenzoate; BEAT,

639

acetyl-CoA:benzylalcohol acetyltransferase; BPBT, benzoyl-CoA:benzylalcohol/2-

640

phenylethanol benzoyltransferase; BSMT, benzoic acid/salicylic acid carboxyl

641

methyltransferase; CA, cinnamic acid; CA-CoA, cinnamoyl-CoA; C4H, cinnamate-4-

642

hydroxylase; CNL, cinnamoyl-CoA ligase; Eug, eugenol; IEug, isoeugenol; KAT, 3-

643

ketoacyl-CoA thiolase; MB, methylbenzoate; 3O3PP-CoA, 3-oxo-3-phenylpropionyl-CoA;

644

PAAS, phenylacetaldehyde synthase; PAL, phenylalanine ammonia lyase; pCA, p-

20 This article is protected by copyright. All rights reserved.

coumaric acid; PEB, phenylethylbenzoate; PhA, phenylacetaldehyde; Phe, L-

646

phenylalanine; PhEth, 2-phenylethanol; PhPyr, phenypyruvic acid.

Accepted Article

645 647

21 This article is protected by copyright. All rights reserved.

Table 1: List of biosynthetic genes involved in final steps of floral volatile formation. Gene Species Reference Volatile Monoterpenoids CitMTSL1 Citrus unshiu (Shimada et 1,8-Cineole al. 2005) NsCIN Nicotiana suaveolens (Roeder et al. 2007) CbLIS Clarkia breweri (Dudareva Linalool et al. 1996a) AmNES/LIS-1 Antirrhinum majus (Nagegowda et al. 2008) TPS10 Arabidopsis thaliana (Ginglinger et al. 2013) TPS14 Arabidopsis thaliana (Ginglinger et al. 2013) Am1e20 Antirrhinum majus (Dudareva Myrcene et al. 2003) AmOc15 Antirrhinum majus (Dudareva et al. 2003) AlstroTPS Alstromeria peruviana (Aros et al. 2012) Am0e23 Antirrhinum majus (Dudareva E-(β)-Ocimene et al. 2003) CitMTSL4 Citrus unshiu (Shimada et al. 2005) Sesquiterpenoids AdAFS1 Actinidia deliciosa (Nieuwenhui α-Farnesene zen et al. 2009) AdGDS1 Actinidia deliciosa (Nieuwenhui Germacrene D zen et al. 2009) VvGerD Vitis vinifera (Lucker et al. 2004) FC0592 Rosa hybrida (Guterman et al. 2002) AmNES/LIS-2 Antirrhinum majus (Nagegowda Nerolidol et al. 2008) AcNES1 Actinidia chinensis (Green et al. 2012) VvVal Vitis vinifera (Lucker et Valencene al. 2004) Benzenoids/ Phenylpropanoids AmBALDH Antirrhinum majus (Long et al. Benzaldehyde 2009) CbBEAT Clarkia breweri (Dudareva Benzylacetate et al. 1998) PhBPBT Petunia hybrida (Boatright et Benzylbenzoate al. 2004) PhEGS Petunia hybrida (Koeduka et Eugenol

Accepted Article

648

22 This article is protected by copyright. All rights reserved.

Accepted Article

Isoeugenol

PhIGS

Petunia hybrida

Isomethyleugenol

CbIEMT

Clarkia breweri

Methylbenzoate

AmBAMT

Antirrhinum majus

PhBSMT1

Petunia hybrida

PhBSMT2

Petunia hybrida

Methyleugenol

CbIEMT

Clarkia breweri

Phenylacetaldehyde

PhPAAS

Petunia hybrida

RhPAAS

Rosa hybrida

2-Phenylethanol

RdPAR

Rosa damascena

Phenylethylbenzoate

PhBPBT

Petunia hybrida

Veratrole

SlGOMT1

Silene latifolia

649 650

23 This article is protected by copyright. All rights reserved.

al. 2006) (Koeduka et al. 2006) (Wang et al. 1997) (Murfitt et al. 2000) (Negre et al. 2003) (Negre et al. 2003) (Wang et al. 1997) (Kaminaga et al. 2006) (Farhi et al. 2010) (Chen et al. 2011b) (Boatright et al. 2004) (Gupta et al. 2012)

References

Accepted Article

651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700

Akhtar T.A. & Pichersky E. Veratrole biosynthesis in white campion. Plant Physiology 162, 52-62. Ando T., Nomura M., Tsukahara J., Watanabe H., Kokubun H., Tsukamoto T., Hashimoto G., Marchesi E. & Kitching I.J. (2001) Reproductive isolation in a native population of Petunia sensu Jussieu (Solanaceae). Annals of Botany 88, 403-413. Andrews E.S., Theis N. & Adler L.S. (2007) Pollinator and herbivore attraction to Cucurbita floral volatiles. Journal of Chemical Ecology 33, 1682-1691. Aros D., Gonzalez V., Allemann R.K., Muller C.T., Rosati C. & Rogers H.J. (2012) Volatile emissions of scented Alstroemeria genotypes are dominated by terpenes, and a myrcene synthase gene is highly expressed in scented Alstroemeria flowers. Journal of Experimental Botany 63, 2739-2752. Bakkali F., Averbeck S., Averbeck D. & Idaomar M. (2008) Biological effects of essential oils-a review. Food and Chemical Toxicology 46, 446-475. Balao F., Herrera J., Talavera S. & Dotterl S. (2011) Spatial and temporal patterns of floral scent emission in Dianthus inoxianus and electroantennographic responses of its hawkmoth pollinator. Phytochemistry 72, 601-609. Becklin K.M., Gamez G., Uelk B., Raguso R.A. & Galen C. (2011) Soil fungal effects on floral signals, rewards, and aboveground interactions in an alpine pollination web. American Journal of Botany 98, 1299-1308. Bergougnoux V., Caissard J.C., Jullien F., Magnard J.L., Scalliet G., Cock J.M., Hugueney P. & Baudino S. (2007) Both the adaxial and abaxial epidermal layers of the rose petal emit volatile scent compounds. Planta 226, 853-866. Bergström G., Dobson H.E.M. & Groth I. (1995) Spatial fragrance patterns within the flowers of Ranunculus acris (Ranunculaceae). Plant Systematics and Evolution 195, 221-242. Bick J.A. & Lange B.M. (2003) Metabolic cross talk between cytosolic and plastidial pathways of isoprenoid biosynthesis: unidirectional transport of intermediates across the chloroplast envelope membrane. Archives of Biochemistry and Biophysics 415, 146-154. Boatright J., Negre F., Chen X.L., Kish C.M., Wood B., Peel G., Orlova I., Gang D., Rhodes D. & Dudareva N. (2004) Understanding in vivo benzenoid metabolism in petunia petal tissue. Plant Physiology 135, 1993-2011. Bohlmann J., Meyer-Gauen G. & Croteau R. (1998) Plant terpenoid synthases: molecular biology and phylogenetic analysis. Proceedings of the National Academy of Sciences, USA 95, 4126-4133. Brantjes N.B.M. (1976) Riddles around pollination of Melandrium album (Mill.) Garcke (Caryophyllaceae) during oviposition by Hadena bicruris Hufn. (Noctuidae, Lepidoptera) .2. Proceedings of the Koninklijke Nederlandse Akademie Van Wetenschappen Series C-Biological and Medical Sciences 79, 127-141. Cane D.E. (1999) Sesquiterpene biosynthesis: Cyclization mechanisms. In: Comprehensive Natural Products Chemistry: Isoprenoids Including Carotenoids and Steroids (ed D.D. Cane), pp. 155-200. Elsevier, Amsterdam. Chappell J. (2002) The genetics and molecular genetics of terpene and sterol origami. Current Opinion in Plant Biology 5, 151-157. Chen F., Tholl D., Bohlmann J. & Pichersky E. (2011a) The family of terpene synthases in plants: a mid-size family of genes for specialized metabolism that is highly diversified throughout the kingdom. Plant Journal 66, 212-229.

24 This article is protected by copyright. All rights reserved.

Chen F., Tholl D., D'Auria J.C., Farooq A., Pichersky E. & Gershenzon J. (2003) Biosynthesis and emission of terpenoid volatiles from Arabidopsis flowers. Plant Cell 15, 481-494. Chen X.M., Kobayashi H., Sakai M., Hirata H., Asai T., Ohnishi T., Baldermann S. & Watanabe N. (2011b) Functional characterization of rose phenylacetaldehyde reductase (PAR), an enzyme involved in the biosynthesis of the scent compound 2-phenylethanol. Journal of Plant Physiology 168, 88-95. Colquhoun T.A. & Clark D.G. (2011) Unraveling the regulation of floral fragrance biosynthesis. Plant Signaling and Behavior 6, 378-381. Colquhoun T.A., Kim J.Y., Wedde A.E., Levin L.A., Schmitt K.C., Schuurink R.C. & Clark D.G. (2011a) PhMYB4 fine-tunes the floral volatile signature of Petunia x hybrida through PhC4H. Journal of Experimental Botany 62, 1133-1143. Colquhoun T.A., Schwieterman M.L., Wedde A.E., Schimmel B.C., Marciniak D.M., Verdonk J.C., Kim J.Y., Oh Y., Galis I., Baldwin I.T. & Clark D.G. (2011b) EOBII controls flower opening by functioning as a general transcriptomic switch. Plant Physiology 156, 974-984. Colquhoun T.A., Verdonk J.C., Schimmel B.C., Tieman D.M., Underwood B.A. & Clark D.G. (2010) Petunia floral volatile benzenoid/phenylpropanoid genes are regulated in a similar manner. Phytochemistry 71, 158-167. Cseke L., Dudareva N. & Pichersky E. (1998) Structure and evolution of linalool synthase. Molecular Biology and Evolution 15, 1491-1498. D'Auria J.C. (2006) Acyltransferases in plants: a good time to be BAHD. Current Opinion in Plant Biology 9, 331-340. D'Auria J.C., Chen F. & Pichersky E. (2002) Characterization of an acyltransferase capable of synthesizing benzylbenzoate and other volatile esters in flowers and damaged leaves of Clarkia breweri. Plant Physiology 130, 466-476. D'Auria J.C., Chen F., Pichersky E. & John T.R. (2003) The SABATH family of MTS in Arabidopsis thaliana and other plant species. In: Recent Advances in Phytochemistry, pp. 253-283. Elsevier. Degenhardt J., Köllner T.G. & Gershenzon J. (2009) Monoterpene and sesquiterpene synthases and the origin of terpene skeletal diversity in plants. Phytochemistry 70, 1621-1637. Dexter R., Qualley A., Kish C.M., Ma C.J., Koeduka T., Nagegowda D.A., Dudareva N., Pichersky E. & Clark D. (2007) Characterization of a petunia acetyltransferase involved in the biosynthesis of the floral volatile isoeugenol. Plant Journal 49, 265-275. Dicke M. & Baldwin I.T. (2010) The evolutionary context for herbivore-induced plant volatiles: beyond the 'cry for help'. Trends in Plant Science 15, 167-175. Dixon R.A. & Strack D. (2003) Phytochemistry meets genome analysis, and beyond. Phytochemistry 62, 815-816. Dobson H.E.M. (2006) Relationship between floral fragrance composition and type of pollinator. In: Biology of Floral Scent (eds N. Dudareva & E. Pichersky), pp. 147198. CRC Press. Dobson H.E.M., Bergström G. & Groth I. (1990) Differences in fragrance chemistry between flower parts of Rosa rugosa Thunb. (Rosaceae). Israel Journal of Botany 39, 143-156. Dobson H.E.M., Danielson E.M. & Van Wesep I.D. (1999) Pollen odor chemicals as modulators of bumble bee foraging on Rosa rugosa Thunb. (Rosaceae). Plant Species Biology 14, 153-166.

Accepted Article

701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749

25 This article is protected by copyright. All rights reserved.

Dobson H.E.M., Groth I. & Bergström G. (1996) Pollen advertisement: Chemical contrasts between whole-flower and pollen odors. American Journal of Botany 83, 877-885. Dötterl S. & Jürgens A. (2005) Spatial fragrance patterns in flowers of Silene latifolia: Lilac compounds as olfactory nectar guides? Plant Systematics and Evolution 255, 99-109. Dötterl S., Jürgens A., Seifert K., Laube T., Weissbecker B. & Schutz S. (2006) Nursery pollination by a moth in Silene latifolia: the role of odours in eliciting antennal and behavioural responses. New Phytologist 169, 707-718. Dötterl S., Jürgens A., Wolfe L. & Biere A. (2009) Disease status and population origin effects on floral scent:: potential consequences for oviposition and fruit predation in a complex interaction between a plant, fungus, and noctuid moth. Journal of Chemical Ecology 35, 307-319. Dötterl S., Wolfe L.M. & Jürgens A. (2005) Qualitative and quantitative analyses of flower scent in Silene latifolia. Phytochemistry 66, 203-213. Dudareva N., Andersson S., Orlova I., Gatto N., Reichelt M., Rhodes D., Boland W. & Gershenzon J. (2005) The nonmevalonate pathway supports both monoterpene and sesquiterpene formation in snapdragon flowers. Proceedings of the National Academy of Sciences, USA 102, 933-938. Dudareva N., Cseke L., Blanc V.M. & Pichersky E. (1996a) Evolution of floral scent in Clarkia: novel patterns of S-linalool synthase gene expression in the C. breweri flower. Plant Cell 8, 1137-1148. Dudareva N., Cseke L., Blanc V.M. & Pichersky E. (1996b) Molecular characterization and cell type-specific expression of linalool synthase gene from Clarkia. Plant Physiology 111, 815-815. Dudareva N., D'Auria J.C., Nam K.H., Raguso R.A. & Pichersky E. (1998) AcetylCoA:benzylalcohol acetyltransferase-an enzyme involved in floral scent production in Clarkia breweri. Plant Journal 14, 297-304. Dudareva N., Martin D., Kish C.M., Kolosova N., Gorenstein N., Faldt J., Miller B. & Bohlmann J. (2003) (E)-beta-ocimene and myrcene synthase genes of floral scent biosynthesis in snapdragon: function and expression of three terpene synthase genes of a new terpene synthase subfamily. Plant Cell 15, 1227-1241. Dudareva N., Murfitt L.M., Mann C.J., Gorenstein N., Kolosova N., Kish C.M., Bonham C. & Wood K. (2000) Developmental regulation of methyl benzoate biosynthesis and emission in snapdragon flowers. Plant Cell 12, 949-961. Dudareva N., Pichersky E. & Gershenzon J. (2004) Biochemistry of plant volatiles. Plant Physiology 135, 1893-1902. Dufaÿ M. & Anstett M.-C. (2003) Conflicts between plants and pollinators that reproduce within inflorescences: evolutionary variations on a theme. Oikos 100, 3-14. Effmert U., Grosse J., Röse U.S.R., Ehrig F., Kägi R. & Piechulla B. (2005) Volatile composition, emission pattern, and localization of floral scent emission in Mirabilis jalapa (Nyctaginaceae). American Journal of Botany 92, 2-12. Faegri K. & van der Pijl L. (1979) The principles of pollination ecology. Oxford, UK: Pergamon Press. Farhi M., Lavie O., Masci T., Hendel-Rahmanim K., Weiss D., Abeliovich H. & Vainstein A. (2010) Identification of rose phenylacetaldehyde synthase by functional complementation in yeast. Plant Molecular Biology 72, 235-245. Farré-Armengol G., Filella I., Llusia J. & Peñuelas J. (2013) Floral volatile organic compounds: Between attraction and deterrence of visitors under global change. Perspectives in Plant Ecology, Evolution and Systematics 15, 56-67

Accepted Article

750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799

26 This article is protected by copyright. All rights reserved.

Feussner I. & Wasternack C. (2002) The lipoxygenase pathway. Annual Review of Plant Biology 53, 275-297. Flamini G., Cioni P.L. & Morelli I. (2003) Differences in the fragrances of pollen, leaves, and floral parts of garland (Chrysanthemum coronarium) and composition of the essential oils from flower heads and leaves. Journal of Agricultural and Food Chemistry 51, 2267-2271. Flügge U.I. & Gao W. (2005) Transport of isoprenoid intermediates across chloroplast envelope membranes. Plant Biology 7, 91-97. Ginglinger J.F. Boachon B. Hofer R.et al. (2013) Gene coexpression analysis reveals complex metabolism of the monoterpene alcohol linalool in Arabidopsis flowers. Plant Cell, Goyret J., Markwell P.M. & Raguso R.A. (2007) The effect of decoupling olfactory and visual stimuli on the foraging behavior of Manduca sexta. Journal of Experimental Biology 210, 1398-1405. Green S.A., Chen X., Nieuwenhuizen N.J., Matich A.J., Wang M.Y., Bunn B.J., Yauk Y.K. & Atkinson R.G. (2012) Identification, functional characterization, and regulation of the enzyme responsible for floral (E)-nerolidol biosynthesis in kiwifruit (Actinidia chinensis). Journal of Experimental Botany 63, 1951-1967. Guirimand G., Guihur A., Phillips M.A., Oudin A., Glevarec G., Melin C., Papon N., Clastre M., St-Pierre B., Rodriguez-Concepcion M., Burlat V. & Courdavault V. (2012) A single gene encodes isopentenyl diphosphate isomerase isoforms targeted to plastids, mitochondria and peroxisomes in Catharanthus roseus. Plant Molecular Biology 79, 443-459. Gupta A.K., Akhtar T.A., Widmer A., Pichersky E. & Schiestl F.P. (2012) Identification of white campion (Silene latifolia) guaiacol O-methyltransferase involved in the biosynthesis of veratrole, a key volatile for pollinator attraction. BMC Plant Biology 12, 158. Gutensohn M., Klempien A., Kaminaga Y., Nagegowda D.A., Negre-Zakharov F., Huh J.-H., Luo H., Weizbauer R., Mengiste T., Tholl D. & Dudareva N. (2011) Role of aromatic aldehyde synthase in wounding/herbivory response and flower scent production in different Arabidopsis ecotypes. Plant Journal 66, 591-602. Gutensohn M., Nagegowda D.A. & Dudareva N. (2013) Involvement of compartmentalization in monoterpene and sesquiterpene biosynthesis in plants. In: Isoprenoid Synthesis in Plants and Microorganisms (eds T.J. Bach & M. Rohmer), pp. 155-169. Springer New York, New York. Guterman I., Shalit M., Menda N., Piestun D., Dafny-Yelin M., Shalev G., Bar E., Davydov O., Ovadis M., Emanuel M., Wang J., Adam Z., Pichersky E., Lewinsohn E., Zamir D., Vainstein A. & Weiss D. (2002) Rose scent: genomics approach to discovering novel floral fragrance-related genes. Plant Cell 14, 23252338. Hendel-Rahmanim K., Masci T., Vainstein A. & Weiss D. (2007) Diurnal regulation of scent emission in rose flowers. Planta 226, 1491-1499. Hirata H., Ohnishib T., Ishidac H., Tomidac K., Sakaic M., Harad M. & Watanabe N. (2012) Functional characterization of aromatic amino acid aminotransferase involved in 2-phenylethanol biosynthesis in isolated rose petal protoplasts. Journal of Plant Physiology 169, 444–451. Hoballah M.E., Stuurman J., Turlings T.C., Guerin P.M., Connetable S. & Kuhlemeier C. (2005) The composition and timing of flower odour emission by wild Petunia axillaris coincide with the antennal perception and nocturnal activity of the pollinator Manduca sexta. Planta 222, 141-150.

Accepted Article

800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849

27 This article is protected by copyright. All rights reserved.

Hong G.J., Xue X.Y., Mao Y.B., Wang L.J. & Chen X.Y. (2012) Arabidopsis MYC2 interacts with DELLA proteins in regulating sesquiterpene synthase gene expression. Plant Cell 24, 2635-2648. Hsiao Y.Y., Jeng M.F., Tsai W.C., Chuang Y.C., Li C.Y., Wu T.S., Kuoh C.S., Chen W.H. & Chen H.H. (2008) A novel homodimeric geranyl diphosphate synthase from the orchid Phalaenopsis bellina lacking a DD(X)2-4D motif. Plant Journal 55, 719733. Hsieh M.H., Chang C.Y., Hsu S.J. & Chen J.J. (2008) Chloroplast localization of methylerythritol 4-phosphate pathway enzymes and regulation of mitochondrial genes in IspD and IspE albino mutants in Arabidopsis. Plant Molecular Biology 66, 663-673. Huang M., Sanchez-Moreiras A.M., Abel C., Sohrabi R., Lee S., Gershenzon J. & Tholl D. (2012) The major volatile organic compound emitted from Arabidopsis thaliana flowers, the sesquiterpene (E)-beta-caryophyllene, is a defense against a bacterial pathogen. New Phytologist 193, 997-1008. Johnson K.B. & Stockwell V.O. (1998) Management of fire blight: A case study in microbial ecology. Annual Review of Phytopathology 36, 227-248. Junker R.R., Gershenzon J. & Unsicker S.B. (2011a) Floral odor bouquet loses its ant repellent properties after inhibition of terpene biosynthesis. Journal of Chemical Ecology 37, 1323-1331. Junker R.R., Heidinger I.M.M. & Blüthgen N. (2010) Floral scent terpenoids deter the facultative florivore Metrioptera bicolor (Ensifera, Tettigoniidae, Decticinae). Journal of Orthoptera Research 19, 69-74. Junker R.R., Loewel C., Gross R., Dötterl S., Keller A. & Blüthgen N. (2011b) Composition of epiphytic bacterial communities differs on petals and leaves. Plant Biology 13, 918-924. Jürgens A., Dötterl S. & Meve U. (2006) The chemical nature of fetid floral odours in stapeliads (Apocynaceae-Asclepiadoideae-Ceropegieae). New Phytologist 172, 452-468 Jürgens A., Wee S.L., Shuttleworth A. & Johnson S.D. (2013) Chemical mimicry of insect oviposition sites: a global analysis of convergence in angiosperms. Ecology Letters 16, 1157-1167. Kaminaga Y., Schnepp J., Peel G., Kish C.M., Ben-Nissan G., Weiss D., Orlova I., Lavie O., Rhodes D., Wood K., Porterfield D.M., Cooper A.J.L., Schloss J.V., Pichersky E., Vainstein A. & Dudareva N. (2006) Plant phenylacetaldehyde synthase is a bifunctional homotetrameric enzyme that catalyzes phenylalanine decarboxylation and oxidation. Journal of Biological Chemistry 281, 23357-23366. Kessler D. & Baldwin I.T. (2007) Making sense of nectar scents: the effects of nectar secondary metabolites on floral visitors of Nicotiana attenuata. Plant Journal 49, 840-854. Kessler D., Diezel C., Clark D.G., Colquhoun T.A. & Baldwin I.T. (2013) Petunia flowers solve the defence/apparency dilemma of pollinator attraction by deploying complex floral blends. Ecology Letters 16, 299-306. Kessler D., Gase K. & Baldwin I.T. (2008) Field experiments with transformed plants reveal the sense of floral scents. Science 321, 1200-1202. Klahre U., Gurba A., Hermann K., Saxenhofer M., Bossolini E., Guerin P.M. & Kuhlemeier C. (2011) Pollinator choice in petunia depends on two major genetic loci for floral scent production. Current Biology 21, 730-739. Klempien A., Kaminaga Y., Qualley A., Nagegowda D.A., Widhalm J.R., Orlova I., Shasany A.K., Taguchi G., Kish C.M., Cooper B.R., D'Auria J.C., Rhodes D.,

Accepted Article

850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899

28 This article is protected by copyright. All rights reserved.

Pichersky E. & Dudareva N. (2012) Contribution of CoA ligases to benzenoid biosynthesis in petunia flowers. Plant Cell 24, 2015-2030. Knudsen J.T., Eriksson R., Gershenzon J. & Stahl B. (2006) Diversity and distribution of floral scent. Botanical Review 72, 1-120. Knudsen J.T. & Gershenzon J. (2006) The chemical diversity of floral scent. In: Biology of Floral Scent (eds N. Dudareva & E. Pichersky), pp. 27-52. CRC Press. Knudsen J.T. & Tollsten L. (1993) Trends in floral scent chemistry in pollination syndromes: floral scent composition in moth-pollinated taxa. Botanical Journal of the Linnean Society 113, 263–284. Knudsen J.T., Tollsten L., Groth I., Bergstrom G. & Raguso R.A. (2004) Trends in floral scent chemistry in pollination syndromes: floral scent composition in hummingbird-pollinated taxa. Botanical Journal of the Linnean Society 146, 191199. Koeduka T., Fridman E., Gang D.R., Vassao D.G., Jackson B.L., Kish C.M., Orlova I., Spassova S.M., Lewis N.G., Noel J.P., Baiga T.J., Dudareva N. & Pichersky E. (2006) Eugenol and isoeugenol, characteristic aromatic constituents of spices, are biosynthesized via reduction of a coniferyl alcohol ester. Proceedings of the National Academy of Sciences, USA 103, 10128-10133. Koeduka T., Louie G.V., Orlova I., Kish C.M., Ibdah M., Wilkerson C.G., Bowman M.E., Baiga T.J., Noel J.P., Dudareva N. & Pichersky E. (2008) The multiple phenylpropene synthases in both Clarkia breweri and Petunia hybrida represent two distinct protein lineages. Plant Journal 54, 362-374. Kolosova N., Gorenstein N., Kish C.M. & Dudareva N. (2001a) Regulation of circadian methyl benzoate emission in diurnally and nocturnally emitting plants. Plant Cell 13, 2333-2347. Kolosova N., Sherman D., Karlson D. & Dudareva N. (2001b) Cellular and subcellular localization of S-adenosyl-L-methionine:benzoic acid carboxyl methyltransferase, the enzyme responsible for biosynthesis of the volatile ester methylbenzoate in snapdragon flowers. Plant Physiology 126, 956-964. Lange B.M., Rujan T., Martin W. & Croteau R. (2000) Isoprenoid biosynthesis: the evolution of two ancient and distinct pathways across genomes. Proceedings of the National Academy of Sciences, USA 97, 13172-13177. Laule O., Furholz A., Chang H.S., Zhu T., Wang X., Heifetz P.B., Gruissem W. & Lange M. (2003) Crosstalk between cytosolic and plastidial pathways of isoprenoid biosynthesis in Arabidopsis thaliana. Proceedings of the National Academy of Sciences, USA 100, 6866-6671. Lavid N., Wang J., Shalit M., Guterman I., Bar E., Beuerle T., Menda N., Shafir S., Zamir D., Adam Z., Vainstein A., Weiss D., Pichersky E. & Lewinsohn E. (2002) Omethyltransferases involved in the biosynthesis of volatile phenolic derivatives in rose petals. Plant Physiology 129, 1899-1907. Long M.C., Nagegowda D.A., Kaminaga Y., Ho K.K., Kish C.M., Schnepp J., Sherman D., Weiner H., Rhodes D. & Dudareva N. (2009) Involvement of snapdragon benzaldehyde dehydrogenase in benzoic acid biosynthesis. Plant Journal 59, 256-265. Lucker J., Bowen P. & Bohlmann J. (2004) Vitis vinifera terpenoid cyclases: functional identification of two sesquiterpene synthase cDNAs encoding (+)-valencene synthase and (-)-germacrene D synthase and expression of mono- and sesquiterpene synthases in grapevine flowers and berries. Phytochemistry 65, 2649-2659. Maeda H. & Dudareva N. (2012) The shikimate pathway and aromatic amino acid biosynthesis in plants. Annual Review of Plant Biology 63, 73-105.

Accepted Article

900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950

29 This article is protected by copyright. All rights reserved.

Maeda H., Shasany A.K., Schnepp J., Orlova I., Taguchi G., Cooper B.R., Rhodes D., Pichersky E. & Dudareva N. (2010) RNAi suppression of Arogenate Dehydratase1 reveals that phenylalanine is synthesized predominantly via the arogenate pathway in petunia petals. Plant Cell 22, 832-849. Maeda H., Yoo H. & Dudareva N. (2011) Prephenate aminotransferase directs plant phenylalanine biosynthesis via arogenate. Nature Chemical Biology 7, 19-21. Mant J., Peakall R. & Schiestl F.P. (2005) Does selection on floral odor promote differentiation among populations and species of the sexually deceptive orchid genus Ophrys? Evolution 59, 1449-1463. McCall A.C. (2008) Florivory affects pollinator visitation and female fitness in Nemophila menziesii. Oecologia 155, 729-737. McGarvey D.J. & Croteau R. (1995) Terpenoid metabolism. Plant Cell 7, 1015-1026. Muhlemann J.K., Maeda H., Chang C.Y., San Miguel P., Baxter I., Cooper B., Perera M.A., Nikolau B.J., Vitek O., Morgan J.A. & Dudareva N. (2012) Developmental changes in the metabolic network of snapdragon flowers. PLoS One 7, e40381. Muhlemann J.K., Waelti M.O., Widmer A. & Schiestl F.P. (2006) Postpollination changes in floral odor in Silene latifolia: adaptive mechanisms for seed-predator avoidance? Journal of Chemical Ecology 32, 1855-1860. Murfitt L.M., Kolosova N., Mann C.J. & Dudareva N. (2000) Purification and characterization of S-adenosyl-L-methionine:benzoic acid carboxyl methyltransferase, the enzyme responsible for biosynthesis of the volatile ester methyl benzoate in flowers of Antirrhinum majus. Archives of Biochemistry and Biophysics 382, 145-151. Nagegowda D.A., Gutensohn M., Wilkerson C.G. & Dudareva N. (2008) Two nearly identical terpene synthases catalyze the formation of nerolidol and linalool in snapdragon flowers. Plant Journal 55, 224-239. Nam K.-H., Dudareva N. & Pichersky E. (1999) Characterization of benzylalcohol acetyltransferases in scented and non-scented Clarkia species. Plant and Cell Physiology 40, 916-923. Negre F., Kish C.M., Boatright J., Underwood B., Shibuya K., Wagner C., Clark D.G. & Dudareva N. (2003) Regulation of methylbenzoate emission after pollination in snapdragon and petunia flowers. Plant Cell 15, 2992-3006. Nieuwenhuizen N.J., Wang M.Y., Matich A.J., Green S.A., Chen X., Yauk Y.K., Beuning L.L., Nagegowda D.A., Dudareva N. & Atkinson R.G. (2009) Two terpene synthases are responsible for the major sesquiterpenes emitted from the flowers of kiwifruit (Actinidia deliciosa). Journal of Experimental Botany 60, 3203-3219. Orlova I., Marshall-Colon A., Schnepp J., Wood B., Varbanova M., Fridman E., Blakeslee J.J., Peer W.A., Murphy A.S., Rhodes D., Pichersky E. & Dudareva N. (2006) Reduction of benzenoid synthesis in petunia flowers reveals multiple pathways to benzoic acid and enhancement in auxin transport. Plant Cell 18, 3458-3475. Parachnowitsch A.L., Raguso R.A. & Kessler A. (2012) Phenotypic selection to increase floral scent emission, but not flower size or colour in bee-pollinated Penstemon digitalis. New Phytologist 195, 667-675. Pellmyr O. & Thien L.B. (1986) Insect reproduction and floral fragrances - Keys to the evolution of the angiosperms. Taxon 35, 76-85. Pichersky E., Lewinsohn E. & Croteau R. (1995) Purification and characterization of Slinalool synthase, an enzyme involved in the production of floral scent in Clarkia breweri. Archives of Biochemistry and Biophysics 316, 803-807. Pichersky E., Noel J.P. & Dudareva N. (2006) Biosynthesis of plant volatiles: Nature's diversity and ingenuity. Science 311, 808-811.

Accepted Article

951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001

30 This article is protected by copyright. All rights reserved.

Pulido P., Perello C. & Rodriguez-Concepcion M. (2012) New insights into plant isoprenoid metabolism. Molecular Plant 5, 964-967. Qualley A.V., Widhalm J.R., Adebesin F., Kish C.M. & Dudareva N. (2012) Completion of the core β-oxidative pathway of benzoic acid biosynthesis in plants. Proceedings of the National Academy of Sciences, USA 109, 16383-16388. Raguso R.A. (2008) Wake up and smell the roses: The ecology and evolution of floral scent. Annual Review of Ecology, Evolution, and Systematics 39, 549–569. Raguso R.A., Levin R.A., Foose S.E., Holmberg M.W. & McDade L.A. (2003) Fragrance chemistry, nocturnal rhythms and pollination "syndromes" in Nicotiana. Phytochemistry 63, 265-284. Raguso R.A. & Pichersky E. (1995) Floral volatiles from Clarkia breweri and C. concinna (Onagraceae): Recent evolution of floral scent and moth pollination. Plant Systematics and Evolution 194, 55-67. Rodriguez-Saona C., Parra L., Quiroz A. & Isaacs R. (2011) Variation in highbush blueberry floral volatile profiles as a function of pollination status, cultivar, time of day and flower part: implications for flower visitation by bees. Annals of Botany 107, 1377-1390. Roeder S., Hartmann A.M., Effmert U. & Piechulla B. (2007) Regulation of simultaneous synthesis of floral scent terpenoids by the 1,8-cineole synthase of Nicotiana suaveolens. Plant Molecular Biology 65, 107-124. Rohdich F., Zepeck F., Adam P., Hecht S., Kaiser J., Laupitz R., Gräwert T., Amslinger S., Eisenreich W., Bacher A. & Arigoni D. (2003) The deoxyxylulose phosphate pathway of isoprenoid biosynthesis: Studies on the mechanisms of the reactions catalyzed by IspG and IspH protein. Proceedings of the National Academy of Sciences, USA 100, 1586-1591. Rose U.S. & Tumlinson J.H. (2004) Volatiles released from cotton plants in response to Helicoverpa zea feeding damage on cotton flower buds. Planta 218, 824-832. Ross J.R., Nam K.H., D'Auria J.C. & Pichersky E. (1999) S-Adenosyl-Lmethionine:salicylic acid carboxyl methyltransferase, an enzyme involved in floral scent production and plant defense, represents a new class of plant methyltransferases. Archives of Biochemistry and Biophysics 367, 9-16. Sakai M., Hirata H., Sayama H., Sekiguchi K., Itano H., Asai T., Dohra H., Hara M. & Watanabe N. (2007) Production of 2-phenylethanol in roses as the dominant floral scent compound from L-phenylalanine by two key enzymes, a PLPdependent decarboxylase and a phenylacetaldehyde reductase. Bioscience, Biotechnology, and Biochemistry 71, 2408-2419. Scalliet G., Journot N., Jullien F., Baudino S., Magnard J.L., Channeliere S., Vergne P., Dumas C., Bendahmane M., Cock J.M. & Hugueney P. (2002) Biosynthesis of the major scent components 3,5-dimethoxytoluene and 1,3,5-trimethoxybenzene by novel rose O-methyltransferases. FEBS Letters 523, 113-118. Schade F., Legge R.L. & Thompson J.E. (2001) Fragrance volatiles of developing and senescing carnation flowers. Phytochemistry 56, 703-710. Schaller F. (2001) Enzymes of the biosynthesis of octadecanoid-derived signalling molecules. Journal of Experimental Botany 52, 11-23. Schiestl F.P. (2010) The evolution of floral scent and insect chemical communication. Ecology Letters 13, 643-656. Schiestl F.P. & Ayasse M. (2001) Post-pollination emission of a repellent compound in a sexually deceptive orchid: a new mechanism for maximising reproductive success? Oecologia 126, 531-534. Schiestl F.P. & Ayasse M. (2002) Do changes in floral odor cause speciation in sexually deceptive orchids? Plant Systematics and Evolution 234, 111–119.

Accepted Article

1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052

31 This article is protected by copyright. All rights reserved.

Schiestl F.P., Ayasse M., Paulus H.F., Erdmann D. & Francke W. (1997) Variation of floral scent emission and postpollination changes in individual flowers of Ophrys sphegodes subsp. sphegodes. Journal of Chemical Ecology 23, 2881-2895. Schiestl F.P., Ayasse M., Paulus H.F., Löfstedt C., Hansson B.S., Ibarra F. & Francke W. (1999) Orchid pollination by sexual swindle. Nature 399, 421-422. Schiestl F.P. & Dötterl S. (2012) The evolution of floral scent and olfactory preferences in pollinators: coevolution or pre-existing bias? Evolution 66, 2042-2055. Schiestl F.P., Huber F.K. & Gomez J.M. (2011) Phenotypic selection on floral scent: trade-off between attraction and deterrence? Evolutionary Ecology 25, 237-248. Schlüter P.M., Xu S., Gagliardini V., Whittle E., Shanklin J., Grossniklaus U. & Schiestl F.P. (2011) Stearoyl-acyl carrier protein desaturases are associated with floral isolation in sexually deceptive orchids. Proceedings of the National Academy of Sciences, USA 108, 5696-56701. Shalit M., Guterman I., Volpin H., Bar E., Tamari T., Menda N., Adam Z., Zamir D., Vainstein A., Weiss D., Pichersky E. & Lewinsohn E. (2003) Volatile ester formation in roses. Identification of an acetyl-coenzyme A. Geraniol/Citronellol acetyltransferase in developing rose petals. Plant Physiology 131, 1868-1876. Shimada T., Endo T., Fujii H., Hara M. & Omura M. (2005) Isolation and characterization of (E)-beta-ocimene and 1,8 cineole synthases in Citrus unshiu Marc. Plant Science 168, 987-995. Shuttleworth A. & Johnson S.D. (2010) The missing stink: sulphur compounds can mediate a shift between fly and wasp pollination systems. Proceedings of the Royal Society B: Biological Sciences 277, 2811-2819. Simkin A.J., Guirimand G., Papon N., Courdavault V., Thabet I., Ginis O., Bouzid S., Giglioli-Guivarc'h N. & Clastre M. (2011) Peroxisomal localisation of the final steps of the mevalonic acid pathway in planta. Planta 234, 903-914. Simkin A.J., Underwood B.A., Auldridge M., Loucas H.M., Shibuya K., Schmelz E., Clark D.G. & Klee H.J. (2004) Circadian regulation of the PhCCD1 carotenoid cleavage dioxygenase controls emission of beta-ionone, a fragrance volatile of petunia flowers. Plant Physiology 136, 3504-3514. Sõber V., Mooraa M. & Tederb T. (2010) Florivores decrease pollinator visitation in a self-incompatible plant. Basic and Applied Ecology 11, 669–675. Spitzer-Rimon B., Farhi M., Albo B., Cna'ani A., Ben Zvi M.M., Masci T., Edelbaum O., Yu Y., Shklarman E., Ovadis M. & Vainstein A. (2012) The R2R3-MYB-like regulatory factor EOBI, acting downstream of EOBII, regulates scent production by activating ODO1 and structural scent-related genes in petunia. Plant Cell 24, 5089-5105. Spitzer-Rimon B., Marhevka E., Barkai O., Marton I., Edelbaum O., Masci T., Prathapani N.-K., Shklarman E., Ovadis M. & Vainstein A. (2010) EOBII, a gene encoding a flower-specific regulator of phenylpropanoid volatiles' biosynthesis in petunia. Plant Cell 22, 1961-1976. Suchet C., Dormont L., Schatz B., Giurfa M., Simon V., Raynaud C. & Chave J. (2011) Floral scent variation in two Antirrhinum majus subspecies influences the choice of naïve bumblebees. Behavioural Ecology and Sociobiology 65, 1015–1027. Theis N. (2006) Fragrance of Canada thistle (Cirsium arvense) attracts both floral herbivores and pollinators. Journal of Chemical Ecology 32, 917-927. Theis N. & Adler L.S. (2012) Advertising to the enemy: enhanced floral fragrance increases beetle attraction and reduces plant reproduction. Ecology 93, 430-435. Theis N., Lerdau M. & Raguso R.A. (2007) The challenge of attracting pollinators while evading floral herbivores: Patterns of fragrance emission in Cirsium arvense and

Accepted Article

1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102

32 This article is protected by copyright. All rights reserved.

Cirsium repandum (Asteraceae). International Journal of Plant Sciences 168, 587-601. Theis N. & Raguso R.A. (2005) The effect of pollination on floral fragrance in thistles. Journal of Chemical Ecology 31, 2581-2600. Tholl D. (2006) Terpene synthases and the regulation, diversity and biological roles of terpene metabolism. Current Opinion in Plant Biology 9, 297-304. Tholl D., Chen F., Petri J., Gershenzon J. & Pichersky E. (2005) Two sesquiterpene synthases are responsible for the complex mixture of sesquiterpenes emitted from Arabidopsis flowers. Plant Journal 42, 757-771. Tholl D., Kish C.M., Orlova I., Sherman D., Gershenzon J., Pichersky E. & Dudareva N. (2004) Formation of monoterpenes in Antirrhinum majus and Clarkia breweri flowers involves heterodimeric geranyl diphosphate synthases. Plant Cell 16, 977-992. Tollsten L. (1993) A multivariate approach to postpollination changes in the floral scent of Platanthera bifolia (Orchidaceae). Nordic Journal Of Botany 13, 495-499. Tollsten L. & Bergström J. (1989) Variation and post-pollination changes in floral odors released by Platanthera bifolia (Orchidaceae). Nordic Journal Of Botany 9, 359362. Van Moerkercke A., Galvan-Ampudia C.S., Verdonk J.C., Haring M.A. & Schuurink R.C. (2012a) Regulators of floral fragrance production and their target genes in petunia are not exclusively active in the epidermal cells of petals. Journal of Experimental Botany 63, 3157-3171. Van Moerkercke A., Haring M.A. & Schuurink R.C. (2011) The transcription factor EMISSION OF BENZENOIDS II activates the MYB ODORANT1 promoter at a MYB binding site specific for fragrant petunias. Plant Journal 67, 917-928. Van Moerkercke A., Haring M.A. & Schuurink R.C. (2012b) A model for combinatorial regulation of the petunia R2R3-MYB transcription factor ODORANT1. Plant Signaling and Behavior 7, 518-520. Van Moerkercke A., Schauvinhold I., Pichersky E., Haring M.A. & Schuurink R.C. (2009) A plant thiolase involved in benzoic acid biosynthesis and volatile benzenoid production. Plant Journal 60, 292-302. Verdonk J.C., Haring M.A., van Tunen A.J. & Schuurink R.C. (2005) ODORANT1 regulates fragrance biosynthesis in petunia flowers. Plant Cell 17, 1612-1624. Vereecken N.J., Cozzolino S. & Schiestl F.P. (2010) Hybrid floral scent novelty drives pollinator shift in sexually deceptive orchids. BMC Evolutionary Biology 10, 103. von Helversen O., Winkler L. & Bestmann H.J. (2000) Sulphur-containing "perfumes" attract flower-visiting bats. Journal of Comparative Physiology A-Sensory Neural and Behavioral Physiology 186, 143-153. Vranova E., Coman D. & Gruissem W. (2013) Network analysis of the MVA and MEP pathways for isoprenoid synthesis. Annual Review of Plant Biology 64, 665-700. Waelti M.O., Muhlemann J.K., Widmer A. & Schiestl F.P. (2008) Floral odour and reproductive isolation in two species of Silene. Journal of Evolutionary Biology 21, 111-121. Waelti M.O., Page P.A., Widmer A. & Schiestl F.P. (2009) How to be an attractive male: floral dimorphism and attractiveness to pollinators in a dioecious plant. BMC Evolutionary Biology 9, 190. Wang J., Dudareva N., Bhakta S., Raguso R.A. & Pichersky E. (1997) Floral scent production in Clarkia breweri (Onagraceae) (II. Localization and developmental modulation of the enzyme S-Adenosyl-L-Methionine:(Iso)Eugenol OMethyltransferase and phenylpropanoid emission). Plant Physiology 114, 213221.

Accepted Article

1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153

33 This article is protected by copyright. All rights reserved.

Wang J. & Pichersky E. (1998) Characterization of S-adenosyl-Lmethionine:(iso)eugenol O-methyltransferase involved in floral scent production in Clarkia breweri. Archives of Biochemistry and Biophysics 349, 153-160. Ward J.L., Baker J.M., Llewellyn A.M., Hawkins N.D. & Beale M.H. (2011) Metabolomic analysis of Arabidopsis reveals hemiterpenoid glycosides as products of a nitrate ion-regulated, carbon flux overflow. Proceedings of the National Academy of Sciences, USA 108, 10762-10767. Wiens D. (1978) Mimicry in plants. Evolutionary Biology 11, 365-403 Willmer P.G., Nuttman C.V., Raine N.E., Stone G.N., Pattrick J.G., Henson K., Stillman P.-., McIlroy L., Potts S.G. & Knudsen J.T. (2009) Floral volatiles controlling ant behaviour. Functional Ecology 23, 888–900. Winterhalter P. & Rouseff R. (2001) Carotenoid-derived aroma compounds: An introduction. In: Carotenoid-Derived Aroma Compounds (eds P. Winterhalter & R. Rouseff), pp. 1-17. American Chemical Society. Wise M.L. & Croteau R. (1999) Monoterpene biosynthesis. In: Comprehensive Natural Products Chemistry: Isoprenoids Including Carotenoids and Steroids (ed D.D. Cane), pp. 97-153. Elsevier, Amsterdam. Yoo H., Widhalm J.R., Qian Y., Maeda H., Cooper B.R., Jannasch A.S., Gonda I., Lewinsohn E., Rhodes D. & Dudareva N. (2013) An alternative pathway contributes to phenylalanine biosynthesis in plants via a cytosolic tyrosine:phenylpyruvate aminotransferase. Nature Communications 4, 2833. Zangerl A.R. & Berenbaum M.R. (2009) Effects of florivory on floral volatile emissions and pollination success in the wild parsnip. Arthropod-Plant Interactions 3, 181191. Zvi M.M., Shklarman E., Masci T., Kalev H., Debener T., Shafir S., Ovadis M. & Vainstein A. (2012) PAP1 transcription factor enhances production of phenylpropanoid and terpenoid scent compounds in rose flowers. New Phytologist 195, 335-345.

Accepted Article

1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183

34 This article is protected by copyright. All rights reserved.

Accepted Article

1184

1185 1186 1187

pce_12314_f1

35 This article is protected by copyright. All rights reserved.

Accepted Article

1188

1189 1190 1191

pce_12314_f2

36 This article is protected by copyright. All rights reserved.

Accepted Article

1192

1193 1194 1195

pce_12314_f3

37 This article is protected by copyright. All rights reserved.

Floral volatiles: from biosynthesis to function.

Floral volatiles have attracted humans' attention since antiquity and have since then permeated many aspects of our lives. Indeed, they are heavily us...
849KB Sizes 0 Downloads 3 Views