Next Article in Journal
Stable Reference Gene Selection for qRT-PCR Normalization in Strawberry (Fragaria × ananassa) Leaves under Different Stress and Light-Quality Conditions
Next Article in Special Issue
The Pour-Through Procedure for Monitoring Container Substrate Chemical Properties: A Review
Previous Article in Journal
First Report of Three Tylenchidae Taxa from Southern Alberta, Canada
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Update on the Function, Biosynthesis and Regulation of Floral Volatile Terpenoids

Yunnan Province Engineering Research Center for Functional Flower Resources and Industrialization, College of Landscape Architecture and Horticulture Sciences, Southwest Forestry University, Kunming 650224, China
*
Author to whom correspondence should be addressed.
These two authors contributed equally to this work.
Horticulturae 2021, 7(11), 451; https://doi.org/10.3390/horticulturae7110451
Submission received: 15 September 2021 / Revised: 24 October 2021 / Accepted: 26 October 2021 / Published: 2 November 2021
(This article belongs to the Special Issue Feature Reviews in Floriculture, Nursery and Landscape, and Turf)

Abstract

:
Floral volatile terpenoids (FVTs) belong to a group of volatile organic compounds (VOC) that play important roles in attracting pollinators, defending against pathogens and parasites and serving as signals associated with biotic and abiotic stress responses. Although research on FVTs has been increasing, a systematic generalization is lacking. Among flowering plants used mainly for ornamental purposes, a systematic study on the production of FVTs in flowers with characteristic aromas is still limited. This paper reviews the biological functions and biosynthesis of FVTs, which may contribute a foundational aspect for future research. We highlight regulatory mechanisms that control the production of FVTs in ornamental flowers and the intersection of biosynthetic pathways that produce flower fragrance and color. Additionally, we summarize the opportunities and challenges facing FVT research in the whole genome and -omics eras and the possible research directions that will provide a foundation for further innovation and utilization of flowering ornamental plants and their germplasm resources.

1. Introduction

Plant volatile compounds (VOCs) biosynthesis occurs in almost all plant organs, including the roots, stems, leaves, flowers, fruits and seeds. They are widely used in perfumes, cosmetics and medicines, and seem promising for use in therapeutic gardens because some possess anxiolytic properties [1,2]. VOCs are lipophilic liquids with low molecular weights and high vapor pressures at ambient temperatures. They include terpenoids, phenylpropanoids/benzenoids, fatty acid derivatives and amino acid derivatives, in addition to a few species- and genus-specific compounds not represented in these major classes. Floral volatile terpenoids (FVTs) are the most dominant VOCs, followed by particular phenylpropanoids/benzenoids [3,4].
The main FVTs—released into the air because of their high vapor pressures—are the hemiterpenes (C5), monoterpenes (C10), sesquiterpenes (C15) and a few diterpenes (C20) [5,6,7]. In addition, irregular volatile terpenoids with carbon skeletons ranging from C8 to C18 are derived from carotenoids. The homoterpenes that are often emitted from night-scented flowers and aerial tissues upon herbivore attack form a small part of the FVTs [3]. Among these FVTs, monoterpenes, such as limonene, ocimene, myrcene and linalool, and sesquiterpenes, such as farnesene, nerolidol and caryophyllene, are the most ubiquitous volatiles (Table 1) [4,8]. The FVTs identified so far in flowering plants are detailed in Table 1.
The mechanism of FVT production is complex, influenced by many environmental factors and associated with vital biological functions. Here, we focus on the biological functions of FVTs and the environmental factors, biosynthesis, transcriptional regulation and modification of terpene skeletons that affects FVTs. We also discuss some possible applications of FVT research.

2. Biological Functions of FVTs

An important property of flowering plants is their scent, which plays a vital role in ecology, the economy and aesthetics. To date, numerous FVTs have been studied in flowering woody plants and herbs (Table 1). Most of these are ornamental plants that are very close to people’s lives. Biologically, FVTs are involved in the attraction of pollinators [9], defense against herbivores [10], protection against pathogens and response to abiotic stress [11]. When it comes to human products, FVTs influence the taste and quality of foods and beverages [12,13]. They also contribute to the aromas of flowers and, consequently, to their ornamental value [14]. FVTs serve as major components of essential oils and are, therefore, key to the development of industries that utilize these natural products [15].
Table 1. Major FVTs and genes associated with their biosynthesis in flowering plants.
Table 1. Major FVTs and genes associated with their biosynthesis in flowering plants.
Latin NameFamilyMain FVT compoundsGenesRef.
Actinidia deliciosa ‘Hayward’Actinidiaceae(E,E)-α-farnesene, (E)-β-ocimene, (+)-germacrene D [16]
Albizia julibrissinLeguminosaeα-farnesene, (Z, E)-β-farneseneAjTPS2, AjTPS5, AjTPS7, AjTPS9, AjTPS10[17]
Camellia spp.Theaceaelinalool and its oxides, geraniol, α-farnesene, hedycaryolCbTPS1, ChTPS1, CbTPS18, CbTPS25, CbTPS28, CbTPS33, CbTPS35 CsTPS29, CbTPS47, CbTPS48, CbTPS51, CbTPS52[18,19,20]
Cananga odorata var. fruticosaAnnonaceaelinaloolCoTPS1, CoTPS2, CoTPS3, CoTPS4[21]
Chimonanthus praecox L.Calycanthaceaelinalool, trans-β-ocimene, β-caryophylleneCpTPS1, CpTPS9, CpTPS10, CpTPS14, CpTPS16, CpTPS4, CpTPS9, CpTPS42[22,23,24,25]
Datura wrightiiSolanaceaelinalool and its enantiomers [26]
Eurya japonica ThunbTheaceaeα-pinene, linalool [27]
Gardenia jasminoidesRubiaceaefarnesene, Z-3-hexenyl tiglate, indole [28]
Gelsemium sempervirens (L.) J. St.-Hil.Gelsemiaceae(Z)-α-ocimene, α-farnesene [29]
Gossypium hirsutumMalvaceae(3S)-linaloolGhTPS12[30,31]
Jasminum spp.Oleaceae𝛼-farnesene, linalool, β-ocimene, germacrene-D [32,33,34,35,36]
Laurus nobilisLauraceaesesquiterpenes, γ-cadinene, δ-cadinene [37]
Lonicera japonicaCaprifoliaceaelinalool [38]
Magnolia champacaMagnoliaceae(R)-linalool, linalool and its oxides [39]
Malus domesticaRosaceae(E)-linalool oxide [40]
Murraya paniculataRutaceaeE-β-ocimene, linalool, α-cubebene [41,42]
Myrtus communis L.Myrtaceaeα-pinene, linalool, 1,8-cineole [43]
Osmanthus fragransOleaceaelinalool and its derivatives, α-ionone, β-iononeOfTPS1, OfTPS2, OfTPS3[44,45,46,47]
Paeonia spp.Paeoniaceaeβ-caryophyllene, linalool [48,49]
Psidium guajavaMyrtaceaeα-cadinol, β-caryophyllene, nerolidol [50]
Rosa spp.Rosaceaegeraniol, linalool, nerolidol, myrcene, ocimene, citronellolNEROLIDOL SYNTHASE (NES), RcLIN-NERS1, RcLIN-NERS2[51,52,53,54,55,56]
Styrax japonicas spp.Styracaceaelinalool, α-pincnc, gcrmacrcnc D [57]
Syringa oblata Lindl.OleaceaeD-limonene [58,59]
Penstemon digitalisPlantaginaceaelinalool and its enantiomers, cis-and trans-β-ocimene [60,61,62]
Alstroemeria spp.Alstroemeriaceae(E)-caryophyllene, α-caryophyllene [63]
Anthurium ‘Mystral’Araceaeeucalyptol, β/α-pinene, β-phellandrene, β-Myrcene [64]
Antirhinum majusPlantaginaceaenerolidol, linalool, (E)-β-ocimene, myrcene [65,66]
Arabidopsis thalianaBrassicaceaeα-copaene, α-caryophyllene, β-elemeneAtTPS21, AtTPS11, and other 40 terpenoid synthase genes[11,67,68,69,70]
Aristolochia giganteaAristolochiaceaelinalool, (Z,E)-α-farnesene, geraniol [71]
Caladenia plicataOrchidaceaeβ-citronellol [72]
Cannabis sativaCannabaceae(+)-α-pinene, (−)-limonene, β-caryophyllene [73]
Chrysanthemum indicumAsteraceae1,8-cineole, germacrene D, camphor [74,75]
Citrus L.Rutaceaelinalool, β-myrcene, α-myrcene, limonene [76]
Clarkia breweriOnagraceaeS-linalool, Linalool, linalool oxidelinalool synthase (LIS) gene[77,78]
Clematis florida cv. ‘Kaiser’Ranunculaceaelinalool, linalool oxide, nerolidolCfTPS1, CfTPS2, CfTPS3[79]
Cymbidium spp.Orchidaceae(E)-β-farrnesene, nerolidol, linaloolCgTPS7[80,81]
Dendrobium officinaleOrchidaceaeα-thujene, linalool, α-terpineolDoTPS10[82,83,84]
Dianthus caryophyllus L.Caryophyllaceaecaryophyllene, caryophyllene oxide, linalool [85,86,87]
Freesia hybrida. “Shiny Gold”Iridaceaelinalool, β-ocimene, D-limoneneFhTPS1, FhTPS2, FhTPS3, FhTPS4, FhTPS5, FhTPS6, FhTPS7, FhTPS8[88,89,90]
Gymnadenia conopsea (L.) R. Br.Orchidaceaeβ-myrcene, α-terpineol, (+)-cyclosativene, α-santalene, trans-α-bergamotene, (Z,E)-α-farnesene, (E,E)-α-farnesene [91]
Hedychium coronariumZingiberaceaeβ-ocimene, 1,8-cineole, linaloolHcTPS1, HcTPS3, HcTPS5, HcTPS6, HcTPS7, HcTPS8, HcTPS10, HcTPS11, HcTPS21[92,93,94,95,96]
Hippeastrum spp.Amaryllidaceaeeucalyptol, (Z)-β-ocimene [97]
Lathyrus odoratusLeguminosaeα-bergamotene, linalool, (−)-α-cubebene [98]
Lavandula spp.Lamiaceaelinalool acetate, linalool, lavandulyl acetate, α/β-PineneLaLIMS, LaLINS[98,99,100,101,102]
Lilium spp.Liliaceaelinalool, myrcene, (E)-β-ocimene, α-pinene, limoneneLoTPS1, LoTPS2, LoTPS3, LoTPS4[103,104,105]
Maxillaria tenuifoliaOrchidaceaeβ-caryophyllene, α-copaene, delta-decalacton [106]
Mentha citrataLamiaceaelinalool and its enantiomers [107]
Mimulus spp.Phrymaceae(E)-β-ocimene, d-limonene, β-myrceneOCIMENE SYNTHASE (OS) gene[108,109,110]
Narcissus spp.Amaryllidaceaemyrcene, eucalyptol, linalool [111,112]
Nicotiana spp.Solanaceae(E)-α-bergamotene, (E)-β-ocimene, 1,8-cineoleNaTPS25, NaTPS38[113,114,115,116,117]
Nymphaea subg. HydrocallisNymphaeaceaelinalool, farnesene, nerolidol [118]
Ocimum basilicum L.Lamiaceaelinalool [119]
Petunia hybridaSolanaceaegermacrene D, β-cadinenePhTPS1, PhTPS2, PhTPS3, PhTPS4[120]
Passiflora edulis SimsPassifloraceaelinaloolPeTPS2, PeTPS3, PeTPS4, PeTPS24[14]
Phalaenopsis spp.Orchidaceaeα-pinene, trans-β-ocimene, linalool, geraniol and their derivativesPbTPS5, PbTPS7, PbTPS9, PbTPS10, PbTPS3, PbTPS4[121,122,123]
Plectranthus amboinicus (Lour.) SprengLamiaceaelinalool, nerolidol [124]
Polianthes tuberosa L.Amaryllidaceaegermacrene D, 1, 8- cineole, α-terpineol [125,126,127,128]
Rheum nobilePolygonaceaeα-pinene [129]
Salvia officinalisLabiataemyrcene, (+)-neomenthol, 1,8-cineole [130]
Tanacetum vulgareAsteraceaeα-pinene, 3-hexen-1-ol-acetate [131]

2.1. Attraction of Pollinators

The most important biological function for FVTs is the attraction of pollinators to enhance propagation. Flower morphology, including flower structure, color, size and the arrangement of floral organs, is important for attracting pollinators. Flowers that produce FVTs are commonly pollinated by bees and moths [4]. Linalool and its enantiomers are reported as being attractive to pollinators including the Megalopta bee [132], bumblebees [60], honeybees [39] and hawkmoths, as well as moths and/or hornworms [25] such as the C. praecox L., D. wrightii, M. champaca, P. digitalis and M. citrata flowers (Table 1). Flowers that produce E-β-ocimene have been reported to be especially attractive to honeybees and bumblebees [107,133].
On the other hand, many flowers produce FVTs that repel herbivores or attract the enemies of herbivores [4], because plants will expose their most vulnerable tissues to insects and pathogens during pollination. Indeed, pollinators often become florivores during particular stages of their life cycles [134]. For example, (E)-β-caryophyllene was reported to increase floral resistance [11]; meanwhile, linalool, (E)-β-farnesene, (E)-α-bergamotene and particular homoterpenes or terpene derivatives help to protect plants from microbial pathogens by recruiting the enemies of pollinators and pests [6,61,112,135,136,137].

2.2. Enhancement of Plant Resistance

Scientists infer that plants would not need to produce terpenes to protect their tissues if fungi and pathogens did not exist. Indeed, FVT emission dramatically decreased after flowers were fumigated with a combination of antibiotics. One week after the fumigation, some flowers still did not emit particular FVTs [138]. The amount of the FVT emission reduction depended on the concentration of antibiotics [139]. Meanwhile, microflora affects FVTs [140]. VOCs (including FVTs) can be used as carbon sources by bacteria in some plants and thus, bacteria can reduce the emission rate of FVTs, which may lead to differences in floral aroma [141].
In addition, particular FVTs play key roles in the synergistic and antagonistic interactions among organisms and the environment [142]. For example, (E, E)-α-farnesene, enantiomers of β-ocimene and linalool are reported to influence multiple interactions between plants and other organisms during biological and abiotic trauma [103,104,133,143]. Germacrene D, bicyclogermacrene and germacrene D-4-ol have been reported to possibly mediate communications between floral organs [137].

3. Complexity of FVT Biosynthesis and Emission

The biosynthesis and emission of FVTs in flowering plants and cut flowers is complex and is not only regulated by the spatio–temporal expression of particular genes but is also affected by various environmental factors, such as light intensity, radiation, the composition of the atmosphere, ambient temperature and relative humidity [85,110]. The mechanisms that influence the biosynthesis and emission of FVTs in response to specific environmental factors are still to be studied [86].

3.1. Spatio–Temporal Regulation

The release of FVTs follows a spatio–temporal pattern. Generally, each flowering plant has a unique composition of FVTs and coordinates the rhythm of FVT emission with the activity of its pollinators [70,90,144,145]. When flowers are ready to be pollinated, they emit elevated levels of volatile compounds. Successful pollination leads to fertilization and decreases in the emission of floral scents, resulting in decreases in unproductive visits from pollinators [145,146]. For example, the emission of linalool from P. lemonei ‘High noon’ flowers appears highest—accounting for 40% of total volatiles—at the fully opened stage and decreases as the flower wilts [47]. The diel emission of FVTs and the composition of FVTs produced by fragrant orchid G. conopsea is consistent with the spatial variation of nocturnal and diurnal pollinators in southern Sweden [90]. Meanwhile, the emission of particular FVTs varies during the different stages of flower development. The monoterpenes and relatively few sesquiterpenes mainly form in buds. The proportion of terpenes is greatly reduced in open flowers. This phenomenon occurs in the flowers of most fragrant plants, including Plumeria rubra flowers [147], lemon basil (O. citriodorum Vis) [148], roses [55], J. auriculatum [32], J. grandiflorum flowers [34], styrax flowers [56], M. tenuifolia [105], C. sativa [72] and C. goeringii [79]. Moreover, the circadian rhythm strongly influences the release of FVTs including (Z)-β-ocimene and (+/−)-linalool from lilium ‘Siberia’ [104], myrcene and (E)-β-ocimene from snapdragon flowers, 1,8-cineole from N. suaveolens [116] and linalool and its enantiomers from Jasminum spp. (J. auriculatum, J. grandiflorum, J. multiflorum and J. malabaricum) [149]. The molecular mechanism linking the emission of particular FVTs to the circadian rhythm remains unknown.
Different plants release FVTs from various tissues and organs to serve specific biological functions. The maximum amounts of FVTs are synthesized at the top of the petunia flower tube directly below the unexpanded corolla, close to the developing stigma. This arrangement allows the stigma to absorb the most FVTs for resistance [119]. In Lilium ‘Siberia’ flowers, the expression of the TPS genes was prominent in flowering parts, especially in sepals and petals [104]. Specifically, LoTPS1 and LoTPS3 were localized to plastids and mitochondria, respectively [104]. The bisexual dimorphism of floral aromas reflects the evolution of flowering plants from hermaphroditic to dioecious plants. Researchers have found that the constituents and release of FVTs differ in pistils and stamens. In E. japonica flowers, α-pinene and linalool were identified as the major components of floral scents in females, hermaphrodites and males. The males emit particularly high levels of α-pinene relative to females and hermaphrodites. The emissions from males generally decrease as flowers senescence. In contrast, the emissions from females and hermaphrodites do not change significantly during senescence [26].

3.2. Luminous Intensity

Light directly affects floral scent emission, changing the qualities and quantities of light-induced fluctuations in the FVTs emitted from Lilium ‘siberia’ flowers [150], P. bellina, P. violacea and Phalaenopsis hybrid flowers [123], Narcissus sp. cut flowers [110] and C. sinensis leaves [151]. One study indicated that light intensity and the circadian clock influenced a Ca2+ signal that contributed to the biosynthesis and emission of monoterpenes in Lilium ‘siberia’ tepals [152].

3.3. Radiation

γ radiation greatly influences the biosynthesis and emission of FVTs. The concentration of linalool in the floral scent bouquet from J. auriculatum was increased twofold in 10 Gy gamma-irradiated variants relative to the control [31]. In addition, the researchers observed a significant increase in the expression of FVT biosynthetic pathway genes and enzymes in particular plants that were irradiated with ultraviolet (UV) light [153]. These data provide evidence that UV-B light affects FVT biosynthesis.

3.4. Composition of the Atmosphere

VOCs form the floral scent trails that are essential for plant–insect interactions. Tropospheric ozone (O3) chemically degrades the floral scent trails, thus reducing the distance, specificity and efficiency of the VOC signal [154,155]. Moreover, elevated levels of O3, carbon dioxide (CO2), diesel exhaust and nitrogen inputs (e.g., atmospheric NOx, N deposition and soil N enrichment) contribute to the production of O3—and have recently been reported to rapidly degrade floral volatiles [155,156,157]. Thus, these environmental factors decrease the distance of scent trails and negatively affect the orientation of pollinators toward floral sources [155].

3.5. Ambient Temperature and Relative Humidity

With global climate change, temperature and humidity are increasingly threatening floral maturation and the size, nectar volume, floral scent and pollinator visitation rates associated with it, and thus, the composition of the pollinator community for some flowering plants [32,157,158,159,160]. The enhancement of temperature and humidity have a significant effect on the amounts of floral scent components, especially FVTs from the O. fragrans cultivars [44], P. axillaris [161], J. auriculatum [32], Lilium ‘Siberia’ [162] and seven common Mediterranean species [163]. As a result, the attractive characteristics of the floral fragrances are diminished due to the changed compositions of FVTs and the release rates of particular FVTs, such as (E)-β-ocimene, (E,E)-α-farnesene and α- and β-pinene [160,164,165,166].

4. Biosynthesis of FVTs

FVTs are the dominant and most diverse group of floral VOCs and include 556 scent compounds [167]. They are produced from the C5 carbon precursors isopentenyl diphosphate (IPP) and dimethylallyl diphosphate (DMAPP) through two alternative pathways, the cytosol-localized mevalonic-acid (MVA) and the plastid-localized 2-c-methylerythritol 4-phosphate (MEP) pathway in flowers and other plant organs [3,113,168]. Generally, the MEP pathway requires seven enzymatic steps to produce IPP and DMAPP. It is mainly used for the biosynthesis of hemiterpenes (C5), monoterpenes (C10) and diterpenes (C20) [3]. The MVA pathway uses six enzymatic reactions to produce IPP and is mainly responsible for the biosynthesis of sesquiterpenes (C15) (Figure 1). One exception is that sesquiterpenes are produced from the IPP and DMAP synthesized by the MEP pathway only in snapdragon flowers [169]. Although the MVA pathway produces only IPP, the MEP pathway produces both IPP and DMAPP at a 6: 1 ratio. Isopentenyl diphosphate isomerase (IDI) plays a key role in reversibly converting IPP to its allylic isomer DMAPP and controlling the equilibrium between these two isomers [3]. FPP is the prenyl diphosphate precursor of the sesquiterpenes that are produced in the cytoplasm. FPPS catalyzes a condensation reaction that utilizes IPP and DMAPP as substrates in a 2:1 ratio. GPP is the prenyl diphosphate precursor of monoterpenes and is produced in the plastid by GPPS, which utilizes IPP and DMAPP as substrates in a 1:1 ratio [3]. Notably, IPP synthesized by the MVA pathway can be imported into mitochondria for the production of FPP, GGPP and GFPP, which serve as precursors for the biosynthesis of other terpenoids [170].
Numerous studies have focused on the upstream genes that encode the enzymes of the MEP and MVA pathways. In the MEP pathway, the expression of the DXS and DXR genes is positively correlated with the emission of volatile terpenoids throughout the flowering period in S. oblata and rose [15,57]. The Nudix hydrolase (RhNUDX1) in the cytosol hydrolyzes GPP to GP, which is converted to geraniol and is part of the biosynthetic pathway that produces the free monoterpene alcohols that contribute to fragrance in roses [51]. In Litsea cubeba (Lour.), the products of the MEP pathway are primarily utilized by GGPPS, and thus, more GGPP-derived products are produced relative to GPP-derived products. Recent studies have shown that both homodimeric and heterodimeric GPPSs contribute to the production of monoterpenes in flowering plants, such as A. majus and C. breweri [171], C. praecox L. [23], L. × intermedia [172] and A. thaliana [173]. The heterodimeric GPPS is composed of one large subunit (LSU) and one small subunit (SSU). The activity, function and regulatory mechanisms associated with the subunits of FVT biosynthetic enzymes are being investigated. The metabolic flux flows mainly from the plastic MEP pathway to the cytoplastic MVA pathway [14,174]. MEcPP is a possible rheostat that influences the abundance of upstream enzymes and is instrumental in fine-tuning the flux of the MEP pathway [175,176]. HMGR is the main rate-determining enzyme of the MVA pathway [177]. During flower development in L. angustifolia, the expression of most MEP pathway-associated genes was upregulated, especially genes that encode the synthases that contribute to the biosynthesis of monoterpenes and sesquiterpenes. In contrast, most genes in the MVA pathway were downregulated [178]. In A. thaliana, in both seedlings and mature plants, the expression of MVA pathway genes is more strongly expressed in the radicle, hypocotyl, roots, flowers and seeds relative to MEP pathway genes [179]. HMGS plays an important role in the yield of monoterpenes and sesquiterpenes and the communication between the MEP and MVA pathways. Overexpressing the LcHMGS gene is shown to significantly increase the types and quantities of FVTs and induce early flowering. In contrast, silencing this gene significantly delays flowering in Litsea cubeba (Lour.) [14]. Some of the transporters required for the subcellular transport of IPP and GPP remain unknown [180]. The MEP and MVA pathways ultimately control the availability of substrates for terpene synthases (TPS) and the interactions between the different substrate pools [14].
By catalyzing complex carbocation-driven cyclization, rearrangement and elimination reactions, TPSs have played a major role in converting a few acyclic prenyl diphosphate substrates into the vast chemical library of terpenoids. Plant TPS genes are widely distributed in flower tissues including buds, pistils, petals, stamens, tubes, corollas, sepals, stigmas and pollen [36,181]. Approximately 1000 monoterpenes and more than 7000 sesquiterpenes have been reported [182]. To date, more than 2000 TPS genes are responsible for the biosynthesis of monoterpenes, while sesquiterpenes have been identified in more than 40 species [182,183]. Table 1 lists some of the TPS genes that have been identified in flowering plants recently.
Three major mechanisms are generally thought to be responsible for the high structural diversity of terpenoids: (1) many TPSs utilize multiple substrates or convert one substrate into multiple products [3,87,184]; (2) single amino acid changes in a TPSs can lead to elementary changes in biosynthetic properties and thus, to the production of completely different mixtures of terpenoids [142]; (3) many modifying enzymes (e.g., oxidoreductases, isomerases, acylases, etc.) convert basic terpene skeletons into diverse products [142].
TPS is a superfamily in higher plants, making TPS gene function diverse. The TPS family is generally divided into three classes and seven sub-families. Class I consists of TPS-c, TPS-e/f and TPS-h. Class II consists of TPS-d. Class III consists of TPS-a, TPS-b and TPS-g [54,97]. The specific functions of particular classes of TPSs can vary from plant to plant. In angiosperms, subfamily TPS-b is typically comprised of mono-TPSs. The TPS-a subclass contributes to sesquiterpene synthases. TPS-e/f is responsible for the biosynthesis of particular mono- and sesquiterpenes, such as linalool synthase in C. breweri and C. florida cv. ‘Kaiser’ [76,78], geranyllinalool synthase in A. thaliana and farnesene synthase in kiwifruit [15]. The angiosperm-specific TPS-g is responsible for the biosynthesis of acyclic monoterpenes and sesquiterpenes [3,36,97,183]. The TPS-f subfamily may be peculiar to dicotyledonous plants [122]. An updated phylogeny of the CsTPS gene family from C. sativa has shown three cannabis-specific clades, including a clade of sesquiterpene synthases within the TPS-b subfamily that typically contains mostly monoterpene synthases [72]. In L. odoratus (sweet pea), LoTPS4 and LoTPS7 belong to the TPS-b clade but catalyze the formation of monoterpenes and sesquiterpenes. LoTPS3 and LoTPS8 from the TPS-a clade also produce monoterpenes and sesquiterpenes [97]. Both TPS-b and TPS-e/f contribute to floral monoterpene biosynthesis in P. bellina [122]. However, in N. attenuata, TPS38 belongs to the TPS-b clade and synthesizes sesquiterpenes, despite being expected to synthesize monoterpenes [112]. The substantial evolution and expansion of mono- and sesqui-TPS families [183] has enabled flowers to produce a surprisingly large variety of FVTs.

5. Transcriptional Regulation and Modification of Terpene Skeletons

5.1. Transcriptional Regulation

Transcription factors regulate gene expression by binding cis-acting elements (i.e., they inhibit or enhance the transcription initiation). Although terpenoids are widely used, their content is low. Transcription factors can regulate gene transcriptions that encode enzymes that contribute to the biosynthesis of secondary metabolites. FVT biosynthesis is regulated by transcription factors (TFs). To date, six TF families have been reported to be involved in the biosynthesis of FVTs in various plants, including MYB- [89,185,186], basic helix-loop-helix (bHLH)- [89,121,187,188], WRKY- [189], ethylene response factor (ERF/AP2)- [190], basic leucine zipper (bZIP)- and NAC-type [185] transcription factors. In the biosynthesis of FVTs, certain TFs work independently, while others work cooperatively. Transcription factors have been thoroughly studied in aromatic plants with small, scentless or inconspicuous flowers, such as Cinnamomum Camphora [191], and in less volatile terpenes, such as artemisinin [192].

5.1.1. MYB

MYB transcription factors are ubiquitous in plants and animals. In plants, the MYB transcription factor family consists of four subclasses: 1R-MYB, R2R3-MYB, R1-R2R3-MYB and 4R-MYB. R2R3-MYB is the largest subclass [186]. Evidence suggests that members of the R2R3-MYB subclass may control the biosynthesis of volatile terpenoids during the full flowering stages in H. coronarium, the bud stages in S. oblata [57,93] and flowers from O. fragrans [193]. The production of anthocyanin pigment-1 (PAP1) is an R2R3-MYB transcription factor that contributes to the biosynthesis of phenylpropanoids, benzenoids and anthocyanins in Arabidopsis, tobacco and petunia flowers [194], and regulates the biosynthesis of the phenylpropanoid and terpenoid scent compounds in rose flowers [195]. PAP1 enhances the emission levels of the sesquiterpene germacrene D in rose flowers [195]. HcMYB directly binds the promoters of genes that contribute to volatile terpenoid biosynthesis (HcTPS1, HcTPS3 and HcTPS10) in H. coronarium [94]. In the same plant, a flower-specific HcMYB2 activates the expression of the linalool synthase gene HcTPS5. The expression of HcMYB2 is induced by auxin [95]. However, in M. spicata, a MYB transcription factor is a novel negative regulator of monoterpene biosynthesis that perturbs the production of sesquiterpene- and diterpene-derived metabolites by binding cis-elements in the gene encoding the large subunit of MsGPPS (MsGPPS.LSU) and suppressing its expression [109].

5.1.2. bHLH

The basic helix-loop-helix type (bHLH) transcription factor family is the second-largest transcription factor family found in plants, and it is divided into 26 subfamilies [196]. Recently, MYC2, a member of the bHLH transcription factor family, was found to activate the expression of two sesquiterpene synthase genes, TPS21 and TPS11, in Lilium ‘Siberia’ [197], after they were found to influence both gibberellic acid (GA) and jasmonic acid (JA) signaling in the inflorescence of Arabidopsis [69]. Meanwhile, both MYB21 and MYC2 were confirmed to upregulate the expression of a monoterpene (linalool) synthase gene in F. hybrida [89]. In C. praecox L., the MYC2 and bHLH13 transcription factors possibly contribute to the biosynthesis of monoterpene (linalool) and sesquiterpene (β-caryophyllene) biosynthesis [21,24].

5.1.3. WRKY

The WRKY family consists of three major groups (I, II and III) that are distinguished based on the number of WRKY domains and the type of zinc-finger-like motifs [198]. Members of group I contain two WRKY domains. Members of groups II and III have a single WRKY domain [199]. In O. fragrans flowers, WRKY transcription factors play important roles in regulating the biosynthesis of secondary metabolites. The expression of OfWRKY19/36/84/139 positively correlates with the biosynthesis of volatile monoterpenes; in contrast, the expression of OfWRKY7 negatively correlates with the biosynthesis of volatile monoterpenes [46]. In addition, the OfWRKYs regulate volatile monoterpene biosynthesis in O. fragrans by binding plant zinc cluster domains [45].

5.1.4. Others

The specificity of transcription factors is versatile in different plants. Additionally, different types of transcription factors may work together for a particular purpose in a particular plant. To date, the following types of transcription factors have been found to work together to regulate FVT biosynthesis in flowering plants: bZIP, bHLH, ERF, MYB and NAC. The NAC genes are specific to plants [200] and contain five subdomains: A, B, C, D and E. Among them, A, C and D are highly conserved, while B and E are relatively less conserved [201]. The bZIP transcription factors are widely distributed and highly conserved in eukaryotes. In plants, the bZIP transcription factors usually contribute to stress responses and the regulation of secondary metabolism [202]. In P. bellina, five TFs (PbbHLH4, PbbHLH6, PbbZIP4, PbERF1 and PbNAC1) have been found to transactivate several structural genes involved in monoterpene biosynthesis [121]. The genes encoding geraniol and linalool synthetase, PbTPS5 and PbTPS10, are regulated by several transcription factors (TFs), including PbbHLH4, PbbHLH6, PbbZIP4, PbERF1, PbERF9 and PbNAC1 [123]. PbNAC1 might be regulated by HY5, a bZIP-type transcription factor. In addition, cis-acting elements in the promoters of PbNAC1, structural genes and TFs associated with monoterpene biosynthesis have been identified that are responsive to light signaling and the circadian clock [123].
Some TFs can collaboratively activate a number of genes in particular secondary metabolic pathways and thus contribute to a comprehensive regulatory network. For instance, in L. × intermedia, LiMYB, LibZIP, LiGeBP, LiSBP-2, LiERF-1 and LiERF-2 possibly activate the transcription of genes that encode monoterpene synthases [185]. LiMYB, LibZIP and LiGeBP activate transcription from both the linalool and linalool synthase promoters, while LiNAC activates transcription only from the 1,8-cineole synthase promoter [185]. In P. bellina, PbbHLH4 may interact with PbbZIP4 and/or PbNAC1 to regulate monoterpene biosynthesis to varying degrees, and PbbHLH4 can regulate the expression of PbTPS7 [121].

5.2. Modification of Terpene Skeletons

In addition to a wide range of volatile terpenoids formed directly through the catalytic activities of TPS enzymes, there are a large number of modifying enzymes, such as cytochrome P450 monooxygenases (P450s) and glycosyl- and acyltransferases, that modify the volatile terpenes produced by TPSs. These modifying enzymes often add functional groups to the volatile terpenes using methylation, acylation or oxidation [203,204]. Modifying enzymes helps to produce more than 80,000 distinct natural products and, thus, alter the olfactory properties of flowers [7,183].
The role of cytochrome P450 enzymes in the modification of FVTs has been widely studied. In flowers, cytochrome P450 enzymes are able to perform highly regio- and stereoselective irreversible oxidations of terpenes, and hence contribute to the structural diversity of terpenes [205,206]. Cytochrome P450 enzymes oxidize more than 20 small and highly hydrophobic natural compounds [207]. For instance, they turn volatile terpenes into soluble compounds stored in floral organs, which prevent the invasion of pests and pathogens [119]. The irregular acyclic C11 and C16 homoterpenes, such as 4,8-dimethylnona-1,3,7-triene (DMNT) and 4,8,12-trimethyltrideca-1,3,7,11-tetraene (TMTT), are usually synthesized in leaves and flowers by a pathway that includes an oxidative degradation step catalyzed by a cytochrome P450 monooxygenase and that utilizes the C15 and C20 tertiary alcohols, (E)-nerolidol and (E, E)-geranyl linalool as the substrates [208,209]. DMNT and TMTT are common constituents of floral volatile blends, as well as of herbivore- and pathogen-induced volatile blends in angiosperms. They contribute to the deterrence or attraction of insect pests and the parasites or predators [208,209] in orchids [210] and Arabidopsis [209]. Based on existing research, members of the CYP70, CYP71 and CYP76 clans, including CYP706A3, CYP76C1, CYP71D13, CYP71A, CYP76AH1 and CYP102A1, oxidize limonene to perillyl alcohol, the anti-tumor compound, and have been co-expressed with genes related to volatile terpenoid biosynthesis [137,211,212,213,214].
Studies on cytochrome P450 enzymes in floral organs have mainly focused on their associations with TPS genes. There is sufficient evidence that cytochrome P450 enzymes are co-expressed with monoterpene synthases and involved in monoterpenoid metabolism. In Arabidopsis flowers, the TPS10 and TPS14 are the terpene synthases that synthesize the two different enantiomers of linalool, (−)-(R)-linalool and (+)-(S)-linalool, respectively. TPS10 and TPS14 are co-expressed with two genes that encode cytochrome P450s (CYP71B31 and CYP76C3) at anthesis, mainly in the upper anther filaments and in petals. CYP71B31 and CYP76C3 produce different but overlapping sets of hydroxylated or epoxidized products [211]. TPS enzymes colocalize in vesicular structures associated with the plastid surface. In contrast, cytochrome P450 enzymes are typically found in the endoplasmic reticulum [211]. These oxygenated products are not emitted into the floral headspace but accumulate in floral tissues as conjugated metabolites [206,211]. CYP76C1 is the most abundant and active linalool-metabolizing enzyme in the flowers of Arabidopsis [213]. In the same flowers, the expression of CYP76C1 is tightly coregulated with TPS10 and TPS14. These genes are expressed in the petals and anthers at anthesis. They facilitate and modulate the emission of linalool and the production of soluble carboxylinalool in opening flowers. They also contribute to reductions in floral attractiveness and flavor, thus protecting flowers against florivorous insects [213]. Moreover, in Arabidopsis, TPS11 produces a blend of sesquiterpenes. TPS11 and CYP706A3 are tightly co-expressed in floral tissues during anthesis and floral bud development. Sesquiterpene and most monoterpene emissions from opening flowers are largely suppressed by CYP706A3-mediated metabolism, which produces terpene oxides retained in floral tissues and soluble sesquiterpene oxides in flower buds. These terpenes effectively protect against insect larvae feeding [119,207]. Mutations in CYP706A3 suppress sesquiterpene and monoterpene emissions from Arabidopsis flowers and change the floral microbial operational taxonomic units (OTUs) in the genus Pseudomonas [119]. In addition, it has also been reported that several members of the CYP76 family in Catharanthus roseus and Grapevine catalyzed different oxidation reactions with monoterpenes, including sequential reactions and conversion of different related compounds [12,211,215].

6. Intersection of Synthetic Pathways That Influence Flower Fragrance and Color

Usually, flowering plants use both fragrance and color to attract pollinators by stimulating the sensory systems of flower-visiting animals [216]. A study on Phrygana, a natural Mediterranean scrubland flower, displayed integrated patterns of scent composition (including monoterpenes and sesquiterpenes) and color consistent with the sensory abilities and perceptual biases of bees [216]. Bee visitation rates are associated with color hue and are positively related to terpenoid emissions from flowers, especially the emission of sesquiterpenes [217]. This is consistent with the integration of phenotypes potentially facilitating pollination. Some scientists hypothesized that if particular plant species emit elevated levels of sesquiterpenes, they would appear more vividly colored to bees. However, neither shared metabolic pathways nor a direct pleiotropy between anthocyanins and sesquiterpenes are known [218]. In addition, in the flowers of Ionopsis utricularioides (Sw.) Lindl., no association between color and fragrance has been observed [218].
Studies of phenotypic characteristics propose an association between FVTs and flower color. The pale-colored R. damascena flowers with their low anthocyanin content appear to produce higher levels of carotenoids and monoterpenes during flower development [219]. Although the gene encoding GGPPS contributes to monoterpene biosynthesis and is probably responsible for both the production of volatile terpenes and floral coloration, further research is needed to establish this connection [219]. In wild roses, the early blooming flowers accumulate low levels of geraniol, developing white petals. The petals of geraniol-rich roses are bigger and have longer shelf lives compared to small, monoterpene-poor roses [55]. As for Cymbidium ‘Sael Bit’, small green and light-yellow colored flowers are highly fragrant relative to brightly colored flowers [220].
From a biosynthesis and transcriptional regulation perspective, there are interactive and competitive relationships between flower fragrance and flower color. TFs that influence different biosynthetic pathways in N. tazetta constitute a complex flower scent–color competition-regulatory network. Potential competition exists between biosynthetic pathways that contributes to the production of floral pigments and fragrance during the growth, development and reproduction of N. tazetta. Coronas MYB12, MYB1, AP2-ERF, bZIP, NAC and MYB are associated with a metabolite flux through the: phenylpropanoid pathway that produces flavonols/anthocyanins; the phenylpropanoid pathway that produces volatile benzenoid/phenylpropanoids in the tepal; and the biosynthetic pathways that produce FVTs and carotenoids/carotene derivatives [111]. During flower development, different enzymes are responsible for directing the carbon flux through different biosynthetic pathways. The levels of transcripts that encode ENT-COPALYL DIPHOSPHATE SYNTHASE/TPS, phenylalanine ammonia-lyase (PAL) and PHYTOENE SYNTHASE/GERANYL-GERANYL REDUCTASE impact the metabolite flux through the terpene biosynthetic pathway and, therefore, influence the benzenoid/phenylpropanoid and carotenoid biosynthetic pathways. Branches of the carotenoid biosynthetic pathway produce the yellow color and increases in indole, (E)-β-ocimene and the benzyl acetate scent in the corona of narcissus [111].
In common angiosperm flowers, biochemical connections between the production of compounds that contribute to color and scent are known from early studies on the phenylpropanoid and terpenoid pathways. One such connection is the shared transcription factor PAP1 [194,195]. Our understanding of the relationship between FVTs and the biosynthesis of pigments is still limited due to the complexity of the two biosynthetic pathways. Nonetheless, the whole genome sequences of model plants and common aromatic flowering plants, such as Arabidopsis, petunia, rose and snapdragon, and the reconstruction of the regulation that controls secondary metabolism, will give researchers a better understanding of the interactions between flower fragrance and color [54].

7. Conclusions and Perspectives

The availability of whole genome sequences for many plants and the recent progress in -omics technology has led to a new genomics and -omics era in plant biology research that has contributed to a novel understanding of regulatory mechanisms involved in the biosynthesis of FVTs. The progress achieved in our understanding of VOCs and FVTs highlights the importance of floral volatile terpenes in natural ecosystems, plant reproduction, plant defense, pollination and signal transduction. Recent breakthroughs in the identification of TPS genes, associated TFs, the supplementation of terpene biosynthetic pathways and its derivatives demonstrate that we have reliable genetic techniques and methods that can be used to improve floral fragrance, such as modifying the emissions of FVTs, to greatly facilitate the recruitment of pollinators and control pests and improve the production of targeted FVTs and essential oils. Moreover, growing metabolically and genetically engineered plants in different natural conditions will allow us to determine the species-specific functions of different FVTs.
Generally speaking, the study of flower fragrance has seen considerable progress in recent years, but there are still many gaps in our knowledge. There is a connection between the biosynthetic pathways that produce FVTs and pigments, but the details of the connection remain unclear. Although our knowledge of FVT biosynthesis is substantial, the transcriptional and post-translational regulation of these pathways requires further study. Meanwhile, our understanding of the influence of hormones, such as auxin, ABA and GA, on FVTs is still limited. In addition, the transport of FVTs remains to be explored. To date, the transmembrane transporters of FVTs and their biosynthetic precursors remain unknown. However, we can speculate that ATP-binding cassette (ABC) transporters transport FVTs across membranes because previous research indicates that an ABC transporter transports phenylpropanoid/benzenoid volatiles across the plasma membrane [221,222].
Moreover, extraction and isolation methods of FVTs, especially specific volatile terpenoids, remain ambiguous. There are significant prospects in investigating the extraction technology of specific terpenes due to the chemical and physical properties of FVTs and their promising applications.

Author Contributions

Conceptualization, Z.Q. and H.H.; writing—review and editing, these two authors contributed to this paper equally; visualization, S.S. and X.Y.; co-supervision and funding acquisition, B.Y.; and proofreading and finalization of the manuscript, L.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Program of Yunnan Province, grant number 202001AS070039, Education and Scientific Research Foundation of Yunnan Province, grant number 2021Y252 and Research Start-Up Fund of Southwest Forestry University, grant number 01102-112109.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We would like to thank Robert M. Larkin from Huazhong Agricultural University for revising the final version.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lu, X.; Tang, K.; Li, P. Plant metabolic engineering strategies for the production of pharmaceutical terpenoids. Front. Plant Sci. 2016, 7, 1647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sabitov, A.; Gaweł-Bęben, K.; Sakipova, Z.; Strzępek-Gomółka, M.; Hoian, U.; Satbayeva, E.; Głowniak, K.; Ludwiczuk, A. Rosa platyacantha schrenk from Kazakhstan-Natural source of bioactive compounds with cosmetic significance. Molecules 2021, 26, 2578. [Google Scholar] [CrossRef]
  3. Dudareva, N.; Klempien, A.; Muhlemann, J.K.; Kaplan, I. Biosynthesis, function and metabolic engineering of plant volatile organic compounds. New Phytol. 2013, 198, 16–32. [Google Scholar] [CrossRef]
  4. Farré-Armengol, G.; Fernández-Martínez, M.; Filella, I.; Junker, R.R.; Peñuelas, J. Deciphering the biotic and climatic factors that influence floral scents: A systematic review of floral volatile emissions. Front. Plant Sci. 2020, 11, 1154. [Google Scholar] [CrossRef] [PubMed]
  5. Knudsen, J.T.; Eriksson, R.; Gershenzon, J.; Stahl, B. Diversity and distribution of floral scent. Bot. Rev. 2006, 72, 1–120. [Google Scholar] [CrossRef]
  6. Raguso, R.A. More lessons from linalool: Insights gained from a ubiquitous floral volatile. Curr. Opin. Plant Biol. 2016, 32, 31–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Pichersky, E.; Raguso, R.A. Why do plants produce so many terpenoid compounds? New Phytol. 2018, 220, 692–702. [Google Scholar] [CrossRef]
  8. Brokl, M.; Fauconnier, M.L.; Benini, C.; Lognay, G.; Jardin, P.d.; Focant, J.F. Improvement of ylang-ylang essential oil characterization by GC × GC-TOFMS. Molecules 2013, 18, 1783–1797. [Google Scholar] [CrossRef] [Green Version]
  9. Raguso, R.A. Wake up and smell the roses: The ecology and evolution of floral scent. Annu. Rev. Ecol. Evol. Syst. 2008, 39, 549–569. [Google Scholar] [CrossRef]
  10. Unsicker, S.B.; Kunert, G.; Gershenzon, J. Protective perfumes: The role of vegetative volatiles in plant defense against herbivores. Curr. Opin. Plant Biol. 2009, 12, 479–485. [Google Scholar] [CrossRef] [PubMed]
  11. Huang, M.; Sanchez-Moreiras, A.M.; Abel, C.; Sohrabi, R.; Lee, S.; Gershenzon, J.; Tholl, D. The major volatile organic compound emitted from Arabidopsis thaliana flowers, the sesquiterpene (E)-β-caryophyllene, is a defense against a bacterial pathogen. New Phytol. 2012, 193, 997–1008. [Google Scholar] [CrossRef] [PubMed]
  12. Ilc, T.; Halter, D.; Miesch, L.; Lauvoisard, F.; Kriegshauser, L.; Ilg, A.; Baltenweck, R.; Hugueney, P.; Werck-Reichhart, D.; Duchêne, E.; et al. A grapevine cytochrome P450 generates the precursor of wine lactone, a key odorant in wine. New Phytol. 2017, 213, 264–274. [Google Scholar] [CrossRef]
  13. Denby, C.M.; Li, R.A.; Vu, V.T.; Costello, Z.; Lin, W.; Chan, L.J.G.; Williams, J.; Donaldson, B.; Bamforth, C.W.; Petzold, C.J.; et al. Industrial brewing yeast engineered for the production of primary flavor determinants in hopped beer. Nat. Commun. 2018, 9, 965. [Google Scholar] [CrossRef]
  14. Xia, Z.; Huang, D.; Zhang, S.; Wang, W.; Ma, F.; Wu, B.; Xu, Y.; Xu, B.; Chen, D.; Zou, M.; et al. Chromosome-scale genome assembly provides insights into the evolution and flavor synthesis of passion fruit (Passiflora edulis Sims). Hortic. Res. 2021, 8, 14. [Google Scholar] [CrossRef]
  15. Wu, L.; Zhao, Y.; Zhang, Q.; Chen, Y.; Gao, M.; Wang, Y. Overexpression of the 3-hydroxy-3-methylglutaryl-CoA synthase gene LcHMGS effectively increases the yield of monoterpenes and sesquiterpenes. Tree Physiol. 2020, 40, 1095–1107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. 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. Two terpene synthases are responsible for the major sesquiterpenes emitted from the flowers of kiwifruit (Actinidia deliciosa). J. Exp. Bot. 2009, 60, 3203–3219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Liu, G.; Yang, M.; Yang, X.; Ma, X.; Fu, J. Five TPSs are responsible for volatile terpenoid biosynthesis in Albizia julibrissin. J. Plant Physiol. 2021, 258–259, 153358. [Google Scholar] [CrossRef] [PubMed]
  18. Hattan, J.; Shindo, K.; Ito, T.; Shibuya, Y.; Watanabe, A.; Tagaki, C.; Ohno, F.; Sasaki, T.; Ishii, J.; Kondo, A.; et al. Identification of a novel hedycaryol synthase gene isolated from Camellia brevistyla flowers and floral scent of Camellia cultivars. Planta 2016, 243, 959–972. [Google Scholar] [CrossRef] [PubMed]
  19. Hattan, J.I.; Shindo, K.; Sasaki, T.; Ohno, F.; Tokuda, H.; Ishikawa, K.; Misawa, N. Identification of novel sesquiterpene synthase genes that mediate the biosynthesis of valerianol, which was an unknown ingredient of tea. Sci. Rep. 2018, 8, 12474. [Google Scholar] [CrossRef]
  20. Zhou, H.C.; Shamala, L.F.; Yi, X.K.; Yan, Z.; Wei, S. Analysis of terpene synthase family genes in Camellia sinensis with an emphasis on abiotic stress conditions. Sci. Rep. 2020, 10, 933. [Google Scholar] [CrossRef] [PubMed]
  21. Jin, J.; Kim, M.J.; Dhandapani, S.; Tjhang, J.G.; Yin, J.L.; Wong, L.; Sarojam, R.; Chua, N.H.; Jang, I.C. The floral transcriptome of ylang ylang (Cananga odorata var fruticosa) uncovers biosynthetic pathways for volatile organic compounds and a multifunctional and novel sesquiterpene synthase. J. Exp. Bot. 2015, 66, 3959–3975. [Google Scholar] [CrossRef] [Green Version]
  22. Tian, J.P.; Ma, Z.Y.; Zhao, K.G.; Zhang, J.; Xiang, L.; Chen, L.Q. Transcriptomic and proteomic approaches to explore the differences in monoterpene and benzenoid biosynthesis between scented and unscented genotypes of wintersweet. Physiol. Plant 2019, 166, 478–493. [Google Scholar] [CrossRef]
  23. Shang, J.; Tian, J.; Cheng, H.; Yan, Q.; Li, L.; Jamal, A.; Xu, Z.; Xiang, L.; Saski, C.A.; Jin, S.; et al. The chromosome-level wintersweet (Chimonanthus praecox) genome provides insights into floral scent biosynthesis and flowering in winter. Genome. Biol. 2020, 21, 200. [Google Scholar] [CrossRef]
  24. Kamran, H.M.; Hussain, S.B.; Junzhong, S.; Xiang, L.; Chen, L.Q. Identification and molecular characterization of geranyl diphosphate synthase (GPPS) genes in wintersweet flower. Plants 2020, 9, 666. [Google Scholar] [CrossRef] [PubMed]
  25. Aslam, M.Z.; Lin, X.; Li, X.; Yang, N.; Chen, L. Molecular cloning and functional characterization of CpMYC2 and CpBHLH13 transcription factors from wintersweet (Chimonanthus praecox L.). Plants 2020, 9, 785. [Google Scholar] [CrossRef]
  26. Reisenman, C.E.; Riffell, J.A.; Bernays, E.A.; Hildebrand, J.G. Antagonistic effects of floral scent in an insect-plant interaction. Proc. Biol. Sci. 2010, 277, 2371–2379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Wang, H.; Zheng, P.; Aoki, D.; Miyake, T.; Yagami, S.; Matsushita, Y.; Fukushima, K.; Nakagawa, M. Sexual and temporal variations in floral scent in the subdioecious shrub Eurya japonica Thunb. Ecol. Evol. 2018, 8, 8266–8272. [Google Scholar] [CrossRef]
  28. Kanlayavattanakul, M.; Lourith, N. Volatile profile and sensory property of Gardenia jasminoides aroma extracts. J. Cosmet. Sci. 2015, 66, 371–377. [Google Scholar] [PubMed]
  29. Obi, J.B.; Golonka, A.M.; Blackwell, A.; Vazquez, I.; Wolfram, N. Floral scent variation in the heterostylous species Gelsemium sempervirens. Molecules 2019, 24, 2818. [Google Scholar]
  30. Huang, X.Z.; Xiao, Y.T.; Köllner, T.G.; Jing, W.X.; Kou, J.F.; Chen, J.Y.; Liu, D.F.; Gu, S.H.; Wu, J.X.; Zhang, Y.J.; et al. The terpene synthase gene family in Gossypium hirsutum harbors a linalool synthase GhTPS12 implicated in direct defence responses against herbivores. Plant Cell Environ. 2018, 41, 261–274. [Google Scholar] [CrossRef] [PubMed]
  31. Huang, M.; Fan, R.; Ye, X.; Lin, R.; Luo, Y.; Fang, N.; Zhong, H.; Chen, S. The transcriptome of flower development provides insight into floral scent formation in Freesia hybrida. Plant Growth Regul. 2018, 86, 93–104. [Google Scholar] [CrossRef]
  32. Barman, M.; Kotamreddy, J.N.R.; Agarwal, A.; Mitra, A. Enhanced emission of linalool from floral scent volatile bouquet in Jasminum auriculatum variants developed via gamma irradiation. Ind. Crops. Prod. 2020, 152, 112545. [Google Scholar] [CrossRef]
  33. Barman, M.; Mitra, A. Floral maturation and changing air temperatures influence scent volatiles biosynthesis and emission in Jasminum auriculatum Vahl. Environ. Exp. Bot. 2021, 181, 104296. [Google Scholar] [CrossRef]
  34. Bera, P.; Mukherjee, C.; Mitra, A. Enzymatic production and emission of floral scent volatiles in Jasminum sambac. Plant Sci. 2017, 256, 25–38. [Google Scholar] [CrossRef] [PubMed]
  35. Pragadheesh, V.S.; Chanotiya, C.S.; Rastogi, S.; Shasany, A.K. Scent from Jasminum grandiflorum flowers: Investigation of the change in linalool enantiomers at various developmental stages using chemical and molecular methods. Phytochem. 2017, 140, 83–94. [Google Scholar] [CrossRef]
  36. Joulain, D. Jasminum grandiflorum flowers-phytochemical complexity and its capture in extracts: A review. Flavour Fragr. J. 2021, 36, 526–553. [Google Scholar] [CrossRef]
  37. Yahyaa, M.; Matsuba, Y.; Brandt, W.; Doron-Faigenboim, A.; Bar, E.; McClain, A.; Davidovich-Rikanati, R.; Lewinsohn, E.; Pichersky, E.; Ibdah, M. Identification, functional characterization, and evolution of terpene synthases from a basal dicot. Plant Physiol. 2015, 169, 1683–1697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Miyake, T.; Yamaoka, R.; Yahara, T. Floral scents of hawkmoth-pollinated flowers in Japan. J. Plant Res. 1998, 111, 199–205. [Google Scholar] [CrossRef]
  39. Dhandapani, S.; Jin, J.; Sridhar, V.; Sarojam, R.; Chua, N.H.; Jang, I.C. Integrated metabolome and transcriptome analysis of Magnolia champaca identifies biosynthetic pathways for floral volatile organic compounds. BMC Genom. 2017, 18, 463. [Google Scholar] [CrossRef]
  40. Rachersberger, M.; Cordeiro, G.D.; Schäffler, I.; Dötterl, S. Honeybee pollinators use visual and floral scent cues to find apple (Malus domestica) flowers. J. Agric. Food Chem. 2019, 67, 13221–13227. [Google Scholar] [CrossRef]
  41. Paul, I.; Chatterjee, A.; Maiti, S.; Bhadoria, P.B.S.; Mitra, A. Dynamic trajectories of volatile and non-volatile specialised metabolites in ‘overnight’ fragrant flowers of Murraya paniculata. Plant Biol. 2019, 21, 899–910. [Google Scholar] [CrossRef] [PubMed]
  42. Paul, I.; Mitra, A.; Bhadoria, P. Seasonal and diel variations in scent composition of ephemeral Murraya paniculata (linn.) Jack flowers are contributed by separate volatile components. Biochem. Syst. Ecol. 2020, 89, 104004. [Google Scholar] [CrossRef]
  43. Bouzabata, A.; Cabral, C.; Gonçalves, M.J.; Cruz, M.T.; Bighelli, A.; Cavaleiro, C.; Casanova, J.; Tomi, F.; Salgueiro, L. Myrtus communis L. as source of a bioactive and safe essential oil. Food Chem. Toxicol. 2015, 75, 166–172. [Google Scholar] [CrossRef] [PubMed]
  44. Zeng, X.; Liu, C.; Zheng, R.; Cai, X.; Luo, J.; Zou, J.; Wang, C. Emission and accumulation of monoterpene and the key terpene synthase (TPS) associated with monoterpene biosynthesis in Osmanthus fragrans lour. Front. Plant Sci. 2016, 6, 1–16. [Google Scholar] [CrossRef] [Green Version]
  45. Fu, J.; Hou, D.; Zhang, C.; Bao, Z.; Zhao, H.; Hu, S. The emission of the floral scent of four Osmanthus fragrans cultivars in response to different temperatures. Molecules 2017, 22, 430. [Google Scholar] [CrossRef] [Green Version]
  46. Ding, W.; Ouyang, Q.; Li, Y.; Shi, T.; Li, L.; Yang, X.; Ji, K.; Wang, L.; Yue, Y. Genome-wide investigation of WRKY transcription factors in sweet osmanthus and their potential regulation of aroma synthesis. Tree Physiol. 2020, 40, 557–572. [Google Scholar] [CrossRef] [PubMed]
  47. Zheng, R.; Zhu, Z.; Wang, Y.; Hu, S.; Xi, W.; Xiao, W.; Qu, X.; Zhong, L.; Fu, Q.; Wang, C. UGT85A84 catalyzes the glycosylation of aromatic monoterpenes in Osmanthus fragrans Lour. flowers. Front. Plant Sci. 2019, 10, 1376. [Google Scholar] [CrossRef] [PubMed]
  48. Zhao, J.; Hu, Z.H.; Leng, P.S.; Cheng, F.Y. Developmental and diurnal change of fragrance emission from ‘High noon’ flowers (Paeonia×lemonei ‘High noon’). In Proceedings of the Third Conference on Horticulture Science and Technology (CHST) 2012, Thika, Kenya, 31 July–3 August 2012; pp. 54–61. [Google Scholar]
  49. Song, C.; Wang, Q.; da Silva, J.A.T.; Yu, X. Identification of floral fragrances and analysis of fragrance patterns in Herbaceous peony cultivars. J. Am. Soc. Hort. Sci. 2018, 143, 248–258. [Google Scholar] [CrossRef] [Green Version]
  50. Fernandes, C.C.; Rezende, J.L.; Silva, E.A.J.; Silva, F.G.; Stenico., L.; Crotti, A.E.M.; Esperandim, V.R.; Santiago, M.B.; Martins, C.H.G.; Miranda, M.L.D. Chemical composition and biological activities of essential oil from flowers of Psidium guajava (Myrtaceae). Braz. J. Biol. 2021, 81, 728–736. [Google Scholar] [CrossRef] [PubMed]
  51. Feng, L.; Chen, C.; Li, T.; Wang, M.; Tao, J.; Zhao, D.; Sheng, L. Flowery odor formation revealed by differential expression of monoterpene biosynthetic genes and monoterpene accumulation in rose (Rosa rugosa Thunb.). Plant Physiol. Biochem. 2014, 75, 80–88. [Google Scholar] [CrossRef]
  52. Magnard, J.L.; Roccia, A.; Caissard, J.C.; Vergne, P.; Sun, P.; Hecquet, R.; Dubois, A.; Hibrand-Saint Oyant, L.; Jullien, F.; Baudino, S.; et al. Plant volatiles. Biosynthesis of monoterpene scent compounds in roses. Science 2015, 349, 81–83. [Google Scholar] [CrossRef] [Green Version]
  53. Magnard, J.L.; Bony, A.R.; Bettini, F.; Campanaro, A.; Blerot, B.; Baudino, S.; Jullien, F. Linalool and linalool nerolidol synthases in roses, several genes for little scent. Plant Physiol. Biochem. 2018, 127, 74–87. [Google Scholar] [CrossRef]
  54. Xu, Q.; Liu, Z.J. The genomic floral language of rose. Nat. Genet. 2018, 50, 770–771. [Google Scholar] [CrossRef] [PubMed]
  55. Raymond, O.; Gouzy, J.; Just, J.; Badouin, H.; Verdenaud, M.; Lemainque, A.; Vergne, P.; Moja, S.; Choisne, N.; Pont, C.; et al. The rosa genome provides new insights into the domestication of modern roses. Nat. Genet. 2018, 50, 772–777. [Google Scholar] [CrossRef] [PubMed]
  56. Dani, K.G.S.; Fineschi, S.; Michelozzi, M.; Trivellini, A.; Pollastri, S.; Loreto, F. Diversification of petal monoterpene profiles during floral development and senescence in wild roses: Relationships among geraniol content, petal colour, and floral lifespan. Oecologia 2020, 1–3. [Google Scholar] [CrossRef]
  57. Chen, C.; Cao, Y.; Chen, H.; Ni, M.; Yu, F. Floral scent compounds and emission patterns of three styrax species. Dendrobiology 2021, 85, 30–38. [Google Scholar] [CrossRef]
  58. Zheng, J.; Hu, Z.; Guan, X.; Dou, D.; Bai, G.; Wang, Y.; Guo, Y.; Li, W.; Leng, P. Transcriptome analysis of Syringa oblata Lindl. inflorescence identifies genes associated with pigment biosynthesis and scent metabolism. PLoS ONE 2015, 10, e0142542. [Google Scholar]
  59. Yan, Z.; Ying, Q.; Zheng, J.; Leng, P.; Hu, Z. Gene cloning and expression analysis of limonene synthase in Syringa oblata and S. oblata var. alba. J. For. Res. 2019, 30, 1301–1309. [Google Scholar] [CrossRef]
  60. Parachnowitsch, A.L.; Raguso, R.A.; Kessler, A. Phenotypic selection to increase floral scent emission, but not flower size or color in bee-pollinated Penstemon digitalis. New Phytol. 2012, 195, 667–675. [Google Scholar] [CrossRef]
  61. Burdon, R.C.F.; Raguso, R.A.; Gegear, R.J.; Pierce, E.C.; Kessler, A.; Parachnowitsch, A.L. Scented nectar and the challenge of measuring honest signals in pollination. J. Ecol. 2020, 108, 2132–2144. [Google Scholar] [CrossRef]
  62. Burdon, R.C.F.; Junker, R.R.; Scofield, D.G.; Parachnowitsch, A.L. Bacteria colonising Penstemon digitalis show volatile and tissue-specific responses to a natural concentration range of the floral volatile linalool. Chemoecology 2018, 28, 11–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Aros, D.; Gonzalez, V.; Allemann, R.K.; Müller, C.T.; Rosati, C.; Rogers, H.J. Volatile emissions of scented Alstroemeria genotypes are dominated by terpenes, and a myrcene synthase gene is highly expressed in scented Alstroemeria flowers. J. Exp. Bot. 2012, 63, 2739–2752. [Google Scholar] [CrossRef] [Green Version]
  64. Wei, Q.; Xia, Q.; Wang, Y.; Chen, W.; Liu, C.; Zeng, R.; Xie, L.; Yi, M.; Guo, H. Profiling of volatile compounds and associated gene expression in two Anthurium cultivars and their F1 hybrid progenies. Molecules 2021, 26, 2902. [Google Scholar] [CrossRef]
  65. Dudareva, N.; Martin, D.; Kish, C.M.; Kolosova, N.; Gorenstein, N.; Fäldt, J.; Miller, B.; Bohlmann, J. (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 2003, 15, 1227–1241. [Google Scholar] [CrossRef] [Green Version]
  66. Nagegowda, D.A.; Gutensohn, M.; Wilkerson, C.G.; Dudareva, N. Two nearly identical terpene synthases catalyze the formation of nerolidol and linalool in snapdragon flowers. Plant J. 2008, 55, 224–239. [Google Scholar] [CrossRef]
  67. Aubourg, S.; Lecharny, A.; Bohlmann, J. Genomic analysis of the terpenoid synthase (AtTPS) gene family of Arabidopsis thaliana. Mol. Genet. Genomics 2002, 267, 730–745. [Google Scholar] [CrossRef] [PubMed]
  68. Chen, F.; Tholl, D.; D’Auria, J.C.; Farooq, A.; Pichersky, E.; Gershenzon, J. Biosynthesis and emission of terpenoid volatiles from Arabidopsis flowers. Plant Cell. 2003, 15, 481–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Tholl, D.; Chen, F.; Petri, J.; Gershenzon, J.; Pichersky, E. Two sesquiterpene synthases are responsible for the complex mixture of sesquiterpenes emitted from Arabidopsis flowers. Plant J. 2005, 42, 757–771. [Google Scholar] [CrossRef] [PubMed]
  70. Hong, G.J.; Xue, X.Y.; Mao, Y.B.; Wang, L.J.; Chen, X.Y. Arabidopsis MYC2 interacts with DELLA proteins in regulating sesquiterpene synthase gene expression. Plant Cell 2012, 24, 2635–2648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Martin, K.R.; More, M.; Hipolito, J.; Charlemagne, S.; Schlumpberger, B.O.; Raguso, R.A. Spatial and temporal variation in volatile composition suggests olfactory division of labor within the trap flowers of Aristolochia gigantea. Flora 2016, 232, 153–168. [Google Scholar] [CrossRef] [Green Version]
  72. Wong, D.C.J.; Pichersky, E.; Peakall, R. The biosynthesis of unusual floral volatiles and blends involved in orchid pollination by deception: Current progress and future prospects. Front. Plant Sci. 2017, 8, 1955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Booth, J.K.; Yuen, M.M.S.; Jancsik, S.; Madilao, L.L.; Page, J.E.; Bohlmann, J. Terpene synthases and terpene variation in Cannabis sativa. Plant Physiol. 2020, 184, 130–147. [Google Scholar] [CrossRef] [PubMed]
  74. Hwang, E.S.; Kim, G.H. Safety evaluation of Chrysanthemum indicum L. flower oil by assessing acute oral toxicity, micronucleus abnormalities, and mutagenicity. Prev. Nutr. Food Sci. 2013, 18, 111–116. [Google Scholar] [CrossRef] [Green Version]
  75. Zhou, Z.; Xian, J.; Wei, W.; Xu, C.; Yang, J.; Zhan, R.; Ma, D. Volatile metabolic profiling and functional characterization of four terpene synthases reveal terpenoid diversity in different tissues of Chrysanthemum indicum L. Phytochemistry 2021, 185, 112687. [Google Scholar] [CrossRef] [PubMed]
  76. Jabalpurwala, F.A.; Smoot, J.M.; Rouseff, R.L. A comparison of citrus blossom volatiles. Phytochemistry 2009, 70, 1428–1434. [Google Scholar] [CrossRef]
  77. Dudareva, N.; Cseke, L.; Blanc, V.M.; Pichersky, E. Evolution of floral scent in Clarkia: Novel patterns of S-linalool synthase gene expression in the Clarkia breweri flower. Plant Cell 1996, 8, 1137–1148. [Google Scholar]
  78. Lavy, M.; Zuker, A.; Lewinsohn, E.; Larkov, O.; Ravid, U.; Vainstein, A.; Weiss, D. Linalool and linalool oxide production in transgenic carnation flowers expressing the Clarkia breweri linalool synthase gene. Mol. Breed. 2002, 9, 103–111. [Google Scholar] [CrossRef]
  79. Jiang, Y.; Qian, R.; Zhang, W.; Wei, G.; Ma, X.; Zheng, J.; Köllner, T.G.; Chen, F. Composition and biosynthesis of scent compounds from sterile flowers of an ornamental plant Clematis florida cv. ‘Kaiser’. Molecules 2020, 25, 1711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Ramya, M.; Jang, S.; An, H.R.; Lee, S.Y.; Park, P.M.; Park, P.H. Volatile organic compounds from orchids: From synthesis and function to gene regulation. Int. J. Mol. Sci. 2020, 21, 1160. [Google Scholar] [CrossRef] [Green Version]
  81. Kim, S.M.; Jang, E.J.; Hong, J.W.; Song, S.H.; Pak, C.H. A comparison of functional fragrant components of Cymbidium (Oriental Orchid) Species. Hortic. Sci. 2016, 34, 331–341. [Google Scholar]
  82. Li, N.; Dong, Y.; Lv, M.; Qian, L.; Sun, X.; Liu, L.; Cai, Y.; Fan, H. Combined analysis of volatile terpenoid metabolism and transcriptome reveals transcription factors related to terpene synthase in two cultivars of Dendrobium officinale flowers. Front. Genet. 2021, 12, 661296. [Google Scholar] [CrossRef]
  83. Zhao, C.; Yu, Z.; Silva, J.A.T.D.; He, C.; Wang, H.; Si, C.; Zhang, M.; Zeng, D.; Duan, J. Functional characterization of a Dendrobium officinale geraniol synthase DoGES1 involved in floral scent formation. Int. J. Mol. Sci. 2020, 21, 7005. [Google Scholar] [CrossRef] [PubMed]
  84. Yu, Z.; Zhao, C.; Zhang, G.; da Silva, J.A.T.; Duan, J. Genome-wide identification and expression profile of TPS gene family in Dendrobium officinale and the role of DoTPS10 in linalool biosynthesis. Int. J. Mol. Sci. 2020, 21, 5419. [Google Scholar] [CrossRef] [PubMed]
  85. Kishimoto, K.; Nakayama, M.; Yagi, M.; Onozaki, T.; Oyama-Okubo, N. Evaluation of wild Dianthus species as genetic resources for fragrant carnation breeding based on their floral scent composition. J. Jpn. Soc. Hortic. Sci. 2011, 80, 175–181. [Google Scholar] [CrossRef] [Green Version]
  86. Kishimoto, K.; Inamoto, K.; Ymaguchi, H. Component analysis and sensory evaluation of scent emitted from cut carnation flowers. Bull NARO Veg. & Flor. Sci. 2019, 3, 29–40. [Google Scholar]
  87. Kishimoto, K.; Shibuya, K. Scent emissions and expression of scent emission-related genes: A comparison between cut and intact carnation flowers. Sci. Hortic. 2021, 281, 109920. [Google Scholar] [CrossRef]
  88. Gao, F.; Liu, B.; Li, M.; Gao, X.; Fang, Q.; Liu, C.; Ding, H.; Wang, L.; Gao, X. Identification and characterization of terpene synthase genes accounting for volatile terpene emissions in flowers of Freesia × hybrida. J. Exp. Bot. 2018, 69, 4249–4265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Srinivasan, A.; Ahn, M.S.; Jo, G.S.; Suh, J.N.; Seo, K.H.; Kim, W.H.; Kang, Y.I.; Lee, Y.R.; Choi, Y.J. Analysis of relative scent intensity, volatile compounds and gene expression in Freesia “Shiny Gold”. Plants 2020, 9, 1597. [Google Scholar] [CrossRef]
  90. Yang, Z.; Li, Y.; Gao, F.; Jin, W.; Li, S.; Kimani, S.; Yang, S.; Bao, T.; Gao, X.; Wang, L. MYB21 interacts with MYC2 to control the expression of terpene synthase genes in flowers of Freesia hybrida and Arabidopsis thaliana. J. Exp. Bot. 2020, 71, 4140–4158. [Google Scholar] [CrossRef]
  91. Chapurlat, E.; Anderson, J.; Ågren, J.; Friberg, M.; Sletvold, N. Diel pattern of floral scent emission matches the relative importance of diurnal and nocturnal pollinators in populations of Gymnadenia conopsea. Ann. Bot. 2018, 121, 711–721. [Google Scholar] [CrossRef] [PubMed]
  92. Yue, Y.; Yu, R.C.; Fan, Y.P. Characterization of two monoterpene synthases involved in floral scent formation in Hedychium coronarium. Planta 2014, 240, 745–762. [Google Scholar] [CrossRef]
  93. Li, X.Y.; Zheng, S.Y.; Yu, R.C.; Fan, Y.P. Promoters of HcTPS1 and HcTPS2 genes from Hedychium coronarium direct floral-specific, developmental-regulated and stress-inducible gene expression in transgenic tobacco. Plant Mol. Biol. Rep. 2014, 32, 864–880. [Google Scholar] [CrossRef]
  94. Yue, Y.; Yu, R.C.; Fan, Y.P. Transcriptome profiling provides new insights into the formation of floral scent in Hedychium coronarium. BMC Genom. 2015, 16, 470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Abbas, F.; Ke, Y.; Zhou, Y.; Yu, Y.; Waseem, M.; Ashraf, U.; Wang, C.; Wang, X.; Li, X.; Yue, Y.; et al. Genome-wide analysis reveals the potential role of MYB transcription factors in floral scent formation in Hedychium coronarium. Front. Plant Sci. 2021, 12, 623742. [Google Scholar] [CrossRef]
  96. Ke, Y.; Abbas, F.; Zhou, Y.; Yu, R.; Fan, Y. Auxin-responsive R2R3-MYB transcription factors HcMYB1 and HcMYB2 activate volatile biosynthesis in Hedychium coronarium flowers. Front. Plant Sci. 2021, 12, 710826. [Google Scholar] [CrossRef] [PubMed]
  97. Meerow, A.W.; Reed, S.T.; Dunn, C.; Schnell, E. Fragrance analysis of two scented Hippeastrum species. HortScience 2017, 52, 1853–1860. [Google Scholar] [CrossRef] [Green Version]
  98. Bao, T.; Shadrack, K.; Yang, S.; Xue, X.; Li, S.; Wang, N.; Wang, Q.; Wang, L.; Gao, X.; Cronk, Q. Functional characterization of terpene synthases accounting for the volatilized-terpene heterogeneity in Lathyrus odoratus cultivar flowers. Plant Cell Physiol. 2020, 61, 1733–1749. [Google Scholar] [PubMed]
  99. Wilson, T.M.; Poulson, A.; Packer, C.; Carlson, R.E.; Buch, R.M. Essential oil profile and yield of corolla, calyx, leaf, and whole flowering top of cultivated Lavandula angustifolia Mill. (Lamiaceae) from Utah. Molecules 2021, 26, 2343. [Google Scholar] [CrossRef]
  100. Aprotosoaie, A.C.; Gille, E.; Trifan, A.; Luca, V.S.; Miron, A. Essential oils of Lavandula genus: A systematic review of their chemistry. Phytochem. Rev. 2017, 16, 761–799. [Google Scholar] [CrossRef]
  101. Adal, A.M.; Sarker, L.S.; Malli, R.P.N.; Liang, P.; Mahmoud, S.S. RNA-Seq in the discovery of a sparsely expressed scent-determining monoterpene synthase in lavender (Lavandula). Planta 2019, 249, 271–290. [Google Scholar] [CrossRef]
  102. Guitton, Y.; Nicole, F.; Jullien, F.; Jean-Claude, C.; Saint-Marcoux, D.; Legendre, L.; Pasquier, B.; Moja, S. A comparative study of terpene composition in different clades of the genus Lavandula. Bot. Lett. 2018, 165, 494–505. [Google Scholar] [CrossRef]
  103. Du, F.; Wang, T.; Fan, J.M.; Liu, Z.Z.; Zong, J.X.; Fan, W.X.; Han, Y.H.; Grierson, D. Volatile composition and classification of Lilium flower aroma types and identification, polymorphisms, and alternative splicing of their monoterpene synthase genes. Hortic. Res. 2019, 6, 110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Abbas, F.; Ke, Y.; Zhou, Y.; Ashraf, U.; Li, X.; Yu, Y.; Yue, Y.; Ahmad, K.W.; Yu, R.; Fan, Y. Molecular cloning, characterization and expression analysis of LoTPS2 and LoTPS4 involved in floral scent formation in oriental hybrid Lilium variety ‘Siberia’. Phytochemistry 2020, 173, 112294. [Google Scholar] [CrossRef] [PubMed]
  105. Abbas, F.; Ke, Y.; Yu, R.; Fan, Y. Functional characterization and expression analysis of two terpene synthases involved in floral scent formation in Lilium ‘Siberia’. Planta 2019, 249, 71–93. [Google Scholar] [CrossRef] [PubMed]
  106. Kim, S.Y.; Ramya, M.; An, H.R.; Park, P.M.; Lee, S.Y.; Park, S.Y.; Park, P.H. Floral volatile compound accumulation and gene expression analysis of Maxillaria tenuifolia. Korean J. Hortic. Sci. Technol. 2019, 37, 756–766. [Google Scholar]
  107. Crowell, A.L.; Williams, D.C.; Davis, E.M.; Wildung, M.R.; Croteau, R. Molecular cloning and characterization of a new linalool synthase. Arch. Biochem. Biophys. 2002, 405, 112–121. [Google Scholar] [CrossRef]
  108. Peng, F.; Byers, K.J.R.P.; Bradshaw, H.D., Jr. Less is more: Independent loss-of-function OCIMENE SYNTHASE alleles parallel pollination syndrome diversification in monkeyflowers (Mimulus). Am. J. Bot. 2017, 104, 1055–1059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Byers, K.J.; Vela, J.P.; Peng, F.; Riffell, J.A.; Bradshaw, H.D., Jr. Floral volatile alleles can contribute to pollinator-mediated reproductive isolation in monkeyflowers (Mimulus). Plant J. 2014, 80, 1031–1042. [Google Scholar] [CrossRef] [Green Version]
  110. Reddy, V.A.; Wang, Q.; Dhar, N.; Kumar, N.; Venkatesh, P.N.; Rajan, C.; Panicker, D.; Sridhar, V.; Mao, H.Z.; Sarojam, R. Spearmint R2R3-MYB transcription factor MsMYB negatively regulates monoterpene production and suppresses the expression of geranyl diphosphate synthase large subunit (MsGPPS.LSU). Plant Biotechnol. J. 2017, 15, 1105–1119. [Google Scholar] [CrossRef] [Green Version]
  111. Terry, M.I.; Ruiz-Hernández, V.; Águila, D.J.; Weiss, J.; Egea-Cortines, M. The effect of post-harvest conditions in Narcissus sp. cut flowers scent profile. Front. Plant Sci. 2021, 11, 540821. [Google Scholar] [CrossRef] [PubMed]
  112. Yang, J.; Ren, Y.; Zhang, D.; Chen, X.; Huang, J.; Xu, Y.; Aucapiña, C.B.; Zhang, Y.; Miao, Y. Transcriptome-based WGCNA analysis reveals regulated metabolite fluxes between floral color and scent in Narcissus tazetta flower. Int. J. Mol. Sci. 2021, 22, 8249. [Google Scholar] [CrossRef] [PubMed]
  113. Zhou, W.; Kügler, A.; McGale, E.; Haverkamp, A.; Knaden, M.; Guo, H.; Beran, F.; Yon, F.; Li, R.; Lackus, N.; et al. Tissue-specific emission of (E)-α-Bergamotene helps resolve the dilemma when pollinators are also herbivores. Curr. Biol. 2017, 27, 1336–1341. [Google Scholar] [CrossRef] [Green Version]
  114. Henry, L.K.; Thomas, S.T.; Widhalm, J.R.; Lynch, J.H.; Davis, T.C.; Kessler, S.A.; Bohlmann, J.; Noel, J.P.; Dudareva, N. Contribution of isopentenyl phosphate to plant terpenoid metabolism. Nat. Plants 2018, 4, 721–729. [Google Scholar] [CrossRef]
  115. Xu, S.; Kreitzer, C.; McGale, E.; Lackus, N.D.; Guo, H.; Köllner, T.G.; Schuman, M.C.; Baldwin, I.T.; Zhou, W. Allelic differences of clustered terpene synthases contribute to correlated intraspecific variation of floral and herbivory-induced volatiles in a wild tobacco. New Phytol. 2020, 228, 1083–1096. [Google Scholar] [CrossRef]
  116. Heiling, S.; Llorca, L.C.; Li, J.; Gase, K.; Schmidt, A.; Schäfer, M.; Schneider, B.; Halitschke, R.; Gaquerel, E.; Baldwin, I.T. Specific decorations of 17-hydroxygeranyllinalool diterpene glycosides solve the autotoxicity problem of chemical defense in Nicotiana attenuata. Plant Cell. 2021, 33, 1748–1770. [Google Scholar] [CrossRef]
  117. Roeder, S.; Hartmann, A.M.; Effmert, U.; Piechulla, B. Regulation of simultaneous synthesis of floral scent terpenoids by the 1, 8-cineole synthase of Nicotiana suaveolens. Plant Mol. Biol. 2007, 65, 107–124. [Google Scholar] [CrossRef] [PubMed]
  118. Maia, A.C.D.; de Lima, C.T.; Navarro, D.M.D.A.F.; Chartier, M.; Giulietti, A.M.; Machado, I.C. The floral scents of Nymphaea subg. Hydrocallis (Nymphaeaceae), the New World night-blooming water lilies, and their relation with putative pollinators. Phytochemistry 2014, 103, 67–75. [Google Scholar] [CrossRef] [PubMed]
  119. Filip, S.; Vidovi, S.; Vladi, J.; Pavli, B.; Zekovi, Z. Chemical composition and antioxidant properties of Ocimum basilicum L. extracts obtained by supercritical carbon dioxide extraction: Drug exhausting method. J. Supercrit. Fluids 2015, 109, 20–25. [Google Scholar] [CrossRef]
  120. Boachon, B.; Lynch, J.H.; Ray, S.; Yuan, J.; Caldo, K.M.P.; Junker, R.R.; Kessler, S.A.; Morgan, J.A.; Dudareva, N. Natural fumigation as a mechanism for volatile transport between flower organs. Nat. Chem. Biol. 2019, 15, 583–588. [Google Scholar] [CrossRef]
  121. Chuang, Y.C.; Hung, Y.C.; Tsai, W.C.; Chen, W.H.; Chen, H.H. Pbbhlh4 regulates floral monoterpene biosynthesis in Phalaenopsis orchids. J. Exp. Bot. 2018, 69, 4363–4377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Huang, H.; Kuo, Y.W.; Chuang, Y.C.; Yang, Y.P.; Huang, L.M.; Jeng, M.F.; Chen, W.H.; Chen, H.H. Terpene synthase-b and terpene synthase-e/f genes produce monoterpenes for Phalaenopsis bellina floral scent. Front. Plant Sci. 2021, 12, 700958. [Google Scholar] [CrossRef] [PubMed]
  123. Chuang, Y.C.; Lee, M.C.; Chang, Y.L.; Chen, W.H.; Chen, H.H. Diurnal regulation of the floral scent emission by light and circadian rhythm in the Phalaenopsis orchids. Bot. Stud. 2017, 58, 50. [Google Scholar] [CrossRef] [PubMed]
  124. Ashaari, N.S.; Ab Rahim, M.H.; Sabri, S.; Lai, K.S.; Song, A.A.; Abdul Rahim, R.; Wan Abdullah, W.M.A.N.; Ong Abdullah, J. Functional characterization of a new terpene synthase from Plectranthus amboinicus. PLoS ONE 2020, 15, e0235416. [Google Scholar] [CrossRef] [PubMed]
  125. Fan, R.; Chen, Y.; Ye, X.; Wu, J.; Lin, B.; Zhong, H. Transcriptome analysis of Polianthes tuberosa during floral scent formation. PLoS ONE 2018, 13, e0199261. [Google Scholar] [CrossRef]
  126. Kutty, N.N.; Mitra, A. Profiling of volatile and non-volatile metabolites in Polianthes tuberosa L. flowers reveals intraspecific variation among cultivars. Phytochemistry 2019, 162, 10–20. [Google Scholar] [CrossRef] [PubMed]
  127. Kutty, N.N.; Ghissing, U.; Mitra, A. Revealing floral metabolite network in tuberose that underpins scent volatiles synthesis, storage and emission. Plant Mol. Biol. 2021, 106, 533–554. [Google Scholar] [CrossRef]
  128. Maiti, S.; Mitra, A. Elucidation of headspace volatilome in Polianthes tuberosa flower for identifying non-invasive biomarkers. Hortic. Environ. Biotech. 2019, 60, 269–280. [Google Scholar]
  129. Song., B.; Chen, G.; Stöcklin, J.; Peng, D.L.; Niu, Y.; Li, Z.M.; Sun, H. A new pollinating seed-consuming mutualism between Rheum nobile and a fly fungus gnat, Bradysia sp., involving pollinator attraction by a specific floral compound. New Phytol. 2014, 203, 1109–1118. [Google Scholar] [CrossRef]
  130. Ali, M.; Li, P.; She, G.; Chen, D.; Wan, X.; Zhao, J. Transcriptome and metabolite analyses reveal the complex metabolic genes involved in volatile terpenoid biosynthesis in garden sage (Salvia officinalis). Sci. Rep. 2017, 7, 16074. [Google Scholar] [CrossRef] [Green Version]
  131. Eilers, E.J.; Kleine, S.; Eckert, S.; Waldherr, S.; Müller, C. Flower production, headspace volatiles, pollen nutrients, and florivory in Tanacetum vulgare chemotypes. Front. Plant Sci. 2021, 11, 611877. [Google Scholar] [CrossRef] [PubMed]
  132. Krug, C.; Cordeiro, G.D.; Schäffler, I.; Silva, C.I.; Oliveira, R.; Schlindwein, C.; Dötterl, S.; Alves-Dos-Santos, I. Nocturnal bee pollinators are attracted to guarana flowers by their scents. Front. Plant Sci. 2018, 9, 1072. [Google Scholar] [CrossRef] [Green Version]
  133. Farré-Armengol, G.; Filella, I.; Llusià, J.; Peñuelas, J. β-Ocimene, a key floral and foliar volatile involved in multiple interactions between plants and other organisms. Molecules 2017, 22, 1148. [Google Scholar] [CrossRef] [Green Version]
  134. Abdalsamee, M.K.; Müller, C. Uncovering different parameters influencing florivory in a specialist herbivore. Ecol. Entomol. 2015, 40, 258–268. [Google Scholar] [CrossRef]
  135. Yu, X.D.; Pickett, J.; Ma, Y.Z.; Bruce, T.; Napier, J.; Jones, H.D.; Xia, L.Q. Metabolic engineering of plant-derived (E)-β-farnesene synthase genes for a novel type of aphid-resistant genetically modified crop plants. J. Integr. Plant Biol. 2012, 54, 282–299. [Google Scholar] [CrossRef]
  136. Li, J.; Hu, H.; Mao, J.; Yu, L.; Stoopen, G.; Wang, M.; Mumm, R.; De Ruijter, N.C.A.; Dicke, M.; Jongsma, M.A.; et al. Defense of pyrethrum flowers: Repelling herbivores and recruiting carnivores by producing aphid alarm pheromone. New Phytol. 2019, 223, 1607–1620. [Google Scholar] [CrossRef] [PubMed]
  137. Boachon, B.; Burdloff, Y.; Ruan, J.X.; Rojo, R.; Junker, R.R.; Vincent, B.; Nicolè, F.; Bringel, F.; Lesot, A.; Henry, L.; et al. A promiscuous CYP706A3 reduces terpene volatile emission from arabidopsis flowers, affecting florivores and the floral microbiome. Plant Cell. 2019, 31, 2947–2972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Peñuelas, J.; Farré-Armengol, G.; Llusia, J.; Gargallo-Garriga, A.; Rico, L.; Sardans, J.; Terradas, J.; Filella, I. Removal of floral microbiota reduces floral terpene emissions. Sci. Rep. 2014, 4, 6727. [Google Scholar] [CrossRef] [Green Version]
  139. Saunier, A.; Mpamah, P.; Biasi, C.; Blande, J.D. Microorganisms in the phylloplane modulate the BVOC emissions of Brassica nigra leaves. Plant Signal Behav. 2020, 15, e1728468. [Google Scholar] [CrossRef] [PubMed]
  140. Pang, Z.; Chen, J.; Wang, T.; Gao, C.; Li, Z.; Guo, L.; Xu, J.; Cheng, Y. Linking plant secondary metabolites and plant microbiomes: A review. Front. Plant Sci. 2021, 12, 621276. [Google Scholar] [CrossRef]
  141. Helletsgruber, C.; Dötterl, S.; Ruprecht, U.; Junker, R.R. Epiphytic bacteria alter floral scent emissions. J. Chem. Ecol. 2017, 43, 1073–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Kessler, A. Introduction to a special feature issue-new insights into plant volatiles. New Phytol. 2018, 220, 655–658. [Google Scholar] [CrossRef] [Green Version]
  143. Yang, T.; Stoopen, G.; Thoen, M.; Wiegers, G.; Jongsma, M.A. Chrysanthemum expressing a linalool synthase gene ‘smells good’, but ‘tastes bad’ to western flower thrips. Plant Biotechnol J. 2013, 11, 875–882. [Google Scholar] [CrossRef] [PubMed]
  144. Dötterl, S.; Jahreiß, K.; Jhumur, U.S.; Jürgens, A. Temporal variation of flower scent in Silene otites (Caryophyllaceae): A species with a mixed pollination system. Bot. J. Linn. Soc. 2012, 169, 447–460. [Google Scholar] [CrossRef]
  145. Rodriguez-Saona, C.; Parra, L.; Quiroz, A.; Isaacs, R. 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. Ann. Bot. 2011, 107, 1377–1390. [Google Scholar] [CrossRef] [PubMed]
  146. Muhlemann, J.K.; Waelti, M.O.; Widmer, A.; Schiestl, F.P. Postpollination changes in floral odor in Silene latifolia: Adaptive mechanisms for seed-predator avoidance? J. Chem. Ecol. 2006, 32, 1855–1860. [Google Scholar] [CrossRef]
  147. Dhandapani, S.; Jin, J.; Sridhar, V.; Chua, N.H.; Jang, I.C. CYP79D73 Participates in biosynthesis of floral scent compound 2-phenylethanol in Plumeria rubra. Plant Physiol. 2019, 180, 171–184. [Google Scholar] [CrossRef] [PubMed]
  148. Al-Kateb, H.; Mottram, D.S. The relationship between growth stages and aroma composition of lemon basil Ocimum citriodorum Vis. Food Chem. 2014, 152, 440–446. [Google Scholar] [CrossRef] [PubMed]
  149. Barman, M.; Mitra, A. Temporal relationship between emitted and endogenous floral scent volatiles in summer-and winter-blooming Jasminum species. Physiol. Plant. 2019, 166, 946–959. [Google Scholar] [CrossRef]
  150. Zhang, H.X.; Leng, P.S.; Hu, Z.H.; Zhao, J.; Wang, W.H.; Xu, F. The floral scent emitted from Lilium ‘Siberia’ at different flowering stages and diurnal variation. Acta. Hortic. Sin. 2013, 40, 693–702. [Google Scholar]
  151. Fu, X.; Chen, Y.; Mei, X.; Katsuno, T.; Kobayashi, E.; Dong, F.; Watanabe, N.; Yang, Z. Regulation of formation of volatile compounds of tea (Camellia sinensis) leaves by single light wavelength. Sci. Rep. 2015, 5, 16858. [Google Scholar] [CrossRef]
  152. Hu, Z.; Li, T.; Zheng, J.; Yang, K.; He, X.; Leng, P. Ca2+ signal contributing to the synthesis and emission of monoterpenes regulated by light intensity in Lilium ‘siberia’. Plant Physiol. Biochem. 2015, 91, 1–9. [Google Scholar] [CrossRef]
  153. Shamala, L.F.; Zhou, H.C.; Han, Z.X.; Wei, S. UV-B induces distinct transcriptional re-programing in UVR8-signal transduction, flavonoid, and terpenoids pathways in Camellia sinensis. Front. Plant Sci. 2020, 11, 234. [Google Scholar] [CrossRef] [Green Version]
  154. Blande, J.D.; Holopainen, J.K.; Niinemets, Ü. Plant volatiles in polluted atmospheres: Stress responses and signal degradation. Plant Cell Environ. 2014, 37, 1892–1904. [Google Scholar] [CrossRef] [Green Version]
  155. Farré-Armengol, G.; Peñuelas, J.; Li, T.; Yli-Pirilä, P.; Filella, I.; Llusia, J.; Blande, J.D. Ozone degrades floral scent and reduces pollinator attraction to flowers. New Phytol. 2016, 209, 152–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Girling, R.D.; Lusebrink, I.; Farthing, E.; Newman, T.A.; Poppy, G.M. Diesel exhaust rapidly degrades floral odours used by honeybees. Sci. Rep. 2013, 3, 2779. [Google Scholar] [CrossRef] [PubMed]
  157. Jamieson, M.A.; Burkle, L.A.; Manson, J.S.; Runyon, J.B.; Trowbridge, A.M.; Zientek, J. Global change effects on plant-insect interactions: The role of phytochemistry. Curr. Opin. Insect Sci. 2017, 23, 70–80. [Google Scholar] [CrossRef] [PubMed]
  158. Burkle, L.A.; Runyon, J.B. Drought and leaf herbivory influence floral volatiles and pollinator attraction. Glob. Chang. Biol. 2016, 22, 1644–1654. [Google Scholar] [CrossRef] [PubMed]
  159. Gallagher, M.K.; Campbell, D.R. Shifts in water availability mediate plant-pollinator interactions. New Phytol. 2017, 215, 792–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Glenny, W.R.; Runyon, J.B.; Burkle, L.A. Drought and increased CO2 alter floral visual and olfactory traits with context-dependent effects on pollinator visitation. New Phytol. 2018, 220, 785–798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Sagae, M.; Oyama-Okubo, N.; Ando, T.; Marchesi, E.; Nakayama, M. Effect of temperature on the floral scent emission and endogenous volatile profile of Petunia axillaris. Biosci. Biotechnol. Biochem. 2008, 72, 110–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Hu, Z.; Zhang, H.; Leng, P.; Zhao, J.; Wang, W.; Wang, S. The emission of floral scent from Lilium “siberia” in response to light intensity and temperature. Acta Physiologiae Plantarum 2013, 35, 1691–1700. [Google Scholar] [CrossRef]
  163. Farré-Armengol, G.; Filella, I.; Llusià, J.; Niinemets, Ü.; Peñuelas, J. Changes in floral bouquets from compound-specific responses to increasing temperatures. Glob. Chang. Biol. 2014, 20, 3660–3669. [Google Scholar] [CrossRef] [Green Version]
  164. Campbell, D.R.; Sosenski, P.; Raguso, R.A. Phenotypic plasticity of floral volatiles in response to increasing drought stress. Ann. Bot. 2019, 123, 601–610. [Google Scholar] [CrossRef] [PubMed]
  165. Rering, C.C.; Franco, J.G.; Yeater, K.M.; Mallinger, R.E. Drought stress alters floral volatiles and reduces floral rewards, pollinator activity, and seed set in a global plant. Ecosphere 2020, 11, e03254. [Google Scholar] [CrossRef]
  166. Kuppler, J.; Kotowska, M.M. A meta-analysis of responses in floral traits and flower-visitor interactions to water deficit. Glob. Chang. Biol. 2021, 27, 3095–3108. [Google Scholar] [CrossRef] [PubMed]
  167. Muhlemann, J.K.; Klempien, A.; Dudareva, N. Floral volatiles: From biosynthesis to function. Plant Cell Environ. 2014, 37, 1936–1949. [Google Scholar] [CrossRef] [PubMed]
  168. Zhao, L.; Chang, W.C.; Xiao, Y.; Liu, H.W.; Liu, P. Methylerythritol phosphate pathway of isoprenoid biosynthesis. Annu. Rev. Biochem. 2013, 82, 497–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Dudareva, N.; Andersson, S.; Orlova, I.; Gatto, N.; Reichelt, M.; Rhodes, D.; Boland, W.; Gershenzon, J. The nonmevalonate pathway supports both monoterpene and sesquiterpene formation in snapdragon flowers. Proc. Natl. Acad. Sci. USA 2005, 102, 933–938. [Google Scholar] [CrossRef] [Green Version]
  170. Kopcsayova, D.; Vranova, E. Functional gene network of prenyltransferases in Arabidopsis thaliana. Molecules 2019, 24, 4556. [Google Scholar] [CrossRef] [Green Version]
  171. Tholl, D.; Kish, C.M.; Orlova, I.; Sherman, D.; Gershenzon, J.; Pichersky, E.; Dudareva, N. Formation of monoterpenes in Antirrhinum majus and Clarkia breweri flowers involves heterodimeric geranyl diphosphate synthases. Plant Cell. 2004, 16, 977–992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Adal, A.M.; Mahmoud, S.S. Short-chain isoprenyl diphosphate synthases of lavender (Lavandula). Plant Mol. Biol. 2020, 102, 517–535. [Google Scholar] [CrossRef]
  173. Chen, Q.; Fan, D.; Wang, G. Heteromeric geranyl (geranyl) diphosphate synthase is involved in monoterpene biosynthesis in Arabidopsis flowers. Mol. Plant 2015, 8, 1434–1437. [Google Scholar] [CrossRef] [Green Version]
  174. Mendoza-Poudereux, I.; Kutzner, E.; Huber, C.; Segura, J.; Eisenreich, W.; Arrillaga, I. Metabolic cross-talk between pathways of terpenoid backbone biosynthesis in spike lavender. Plant Physiol. Biochem. 2015, 95, 113–120. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, J.Z.; Lei, Y.; Xiao, Y.; He, X.; Liang, J.; Jiang, J.; Dong, S.; Ke, H.; Leon, P.; Zerbe, P.; et al. Uncovering the functional residues of Arabidopsis isoprenoid biosynthesis enzyme HDS. Proc. Natl. Acad. Sci. USA 2020, 117, 355–361. [Google Scholar] [CrossRef]
  176. Pu, X.; Dong, X.; Li, Q.; Chen, Z.; Liu, L. An update on the function and regulation of methylerythritol phosphate and mevalonate pathways and their evolutionary dynamics. J. Integr. Plant Biol. 2021, 63, 1211–1226. [Google Scholar] [CrossRef] [PubMed]
  177. Rodriguez-Concepcion, M.; Boronat, A. Breaking new ground in the regulation of the early steps of plant isoprenoid biosynthesis. Curr. Opin. Plant Biol. 2015, 25, 17–22. [Google Scholar] [CrossRef] [PubMed]
  178. Guo, D.; Kang, K.; Wang, P.; Li, M.; Huang, X. Transcriptome profiling of spike provides expression features of genes related to terpene biosynthesis in lavender. Sci. Rep. 2020, 10, 6933. [Google Scholar] [CrossRef]
  179. Vranová, E.; Coman, D.; Gruissem, W. Network analysis of the MVA and MEP pathways for isoprenoid synthesis. Annu. Rev. Plant Biol. 2013, 64, 665–700. [Google Scholar] [CrossRef]
  180. Borghi, M.; Fernie, A.R.; Schiestl, F.P.; Bouwmeester, H.J. The sexual advantage of looking, smelling, and tasting good: The metabolic network that produces signals for pollinators. Trends Plant Sci. 2017, 22, 338–350. [Google Scholar] [CrossRef]
  181. Falara, V.; Akhtar, T.A.; Nguyen, T.T.; Spyropoulou, E.A.; Bleeker, P.M.; Schauvinhold, I.; Matsuba, Y.; Bonini, M.E.; Schilmiller, A.L.; Last, R.L.; et al. The tomato terpene synthase gene family. Plant Physiol. 2011, 157, 770–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Huang, L.M.; Huang, H.; Chuang, Y.C.; Chen., W.H.; Wang, C.N.; Chen, H.H. Evolution of terpene synthases in Orchidaceae. Int. J. Mol. Sci. 2021, 22, 6947. [Google Scholar] [CrossRef] [PubMed]
  183. Karunanithi, P.S.; Zerbe, P. Terpene synthases as metabolic gatekeepers in the evolution of plant terpenoid chemical diversity. Front. Plant Sci. 2019, 10, 1166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Christianson, D.W. Structural and chemical biology of terpenoid cyclases. Chem. Rev. 2017, 117, 11570–11648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Sarker, L.S.; Adal, A.M.; Mahmoud, S.S. Diverse transcription factors control monoterpene synthase expression in lavender (Lavandula). Planta 2020, 251, 5. [Google Scholar] [CrossRef]
  186. Ramya, M.; Kwon, O.K.; An, H.R.; Park, P.M.; Baek, Y.S.; Park, P.H. Floral scent: Regulation and role of MYB transcription factors. Phytochem. Lett. 2017, 19, 114–120. [Google Scholar] [CrossRef]
  187. Van Moerkercke, A.; Steensma, P.; Schweizer, F.; Pollier, J.; Gariboldi, I.; Payne, R.; Vanden Bossche, R.; Miettinen, K.; Espoz, J.; Purnama, P.C.; et al. The bHLH transcription factor BIS 1 controls the iridoid branch of the monoterpenoid indole alkaloid pathway in Catharanthus roseus. Proc. Natl. Acad. Sci. USA 2015, 112, 8130–8135. [Google Scholar]
  188. Van Moerkercke, A.; Steensma, P.; Gariboldi, I.; Espoz, J.; Purnama, P.C.; Schweizer, F.; Miettinen, K.; Vanden, B.R.; De Clercq, R.; Memelink, J.; et al. The basic helix- loop- helix transcription factor BIS2 is essential for monoterpenoid indole alkaloid production in the medicinal plant Catharanthus roseus. Plant J. 2016, 88, 3–12. [Google Scholar]
  189. Alfieri, M.; Vaccaro, M.C.; Cappetta, E.; Ambrosone, A.; Tommasi, D.; Leone, A. Coactivation of MEP-biosynthetic genes and accumulation of abietane diterpenes in Salvia sclarea by heterologous expression of WRKY and MYC2 transcription factors. Sci. Rep. 2018, 8, 11009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Paul, P.; Singh, S.K.; Patra, B.; Sui, X.; Pattanaik, S.; Yuan, L. A differentially regulated AP2/ERF transcription factor gene cluster acts downstream of a MAP kinase cascade to modulate terpenoid indole alkaloid biosynthesis in Catharanthus roseus. New Phytol. 2017, 213, 1107–1123. [Google Scholar] [CrossRef]
  191. Yang, Z.; Xie, C.; Huang, Y.; An, W.; Liu, S.; Huang, S.; Zheng, X. Metabolism and transcriptome profiling provides insight into the genes and transcription factors involved in monoterpene biosynthesis of borneol chemotype of Cinnamomum camphora induced by mechanical damage. PeerJ 2021, 9, e11465. [Google Scholar] [CrossRef]
  192. Zhang, F.Y.; Xiang, L.E.; Yu, Q.; Zhang, H.X.; Zhang, T.X.; Zeng, J.L.; Geng, C.; Li, L.; Fu, X.Q.; Shen, Q.; et al. Artemisinin Biosynthesis Promoting Kinase 1 positively regulates artemisinin biosynthesis through phosphorylating Aab-ZIP1. J. Exp. Bot. 2018, 69, 1109–1123. [Google Scholar] [CrossRef] [Green Version]
  193. Li, H.; Yue, Y.Z.; Ding, W.J.; Chen, G.W.; Li, L.; Li, Y.L.; Shi, T.T.; Yang, X.L.; Wang, L.G. Genome-wide identification, classification, and expression profiling reveals R2R3-MYB transcription factors related to monoterpenoid biosynthesis in Osmanthus fragrans. Genes 2020, 11, 353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Ben Zvi, M.M.; Negre-Zakharov, F.; Masci, T.; Ovadis, M.; Shklarman, E.; Ben-Meir, H.; Tzfira, T.; Dudareva, N.; Vainstein, A. Interlinking showy traits: Co-engineering of scent and colour biosynthesis in flowers. Plant Biotechnol. J. 2008, 6, 403–415. [Google Scholar]
  195. Zvi, M.M.B.; Shklarman, E.; Masci, T.; Kalev, H.; Debener, T.; Shafir, S.; Ovadis, M.; Vainstein, A. PAP1 transcription factor enhances production of phenylpropanoid and terpenoid scent compounds in rose flowers. New Phytol. 2012, 195, 335–345. [Google Scholar] [CrossRef] [PubMed]
  196. Feller, A.; Machemer, K.; Braun, E.L.; Grotewold, E. Evolutionary and comparative analysis of MYB and bHLH plant transcription factors. Plant J. 2011, 66, 94–116. [Google Scholar] [CrossRef] [PubMed]
  197. Shi, S.; Duan, G.; Li, D.; Wu, J.; Liu, X.; Hong, B.; Yi, M.; Zhang, Z. Two-dimensional analysis provides molecular insight into flower scent of Lilium ‘Siberia’. Sci. Rep. 2018, 8, 5352. [Google Scholar] [CrossRef] [Green Version]
  198. Eulgem, T.; Rushton, P.J.; Robatzek, S.; Somssich, I.E. The WRKY superfamily of plant transcription factors. Trends Plant Sci. 2000, 5, 199–206. [Google Scholar] [CrossRef]
  199. Zhang, Y.; Wang, L. The WRKY transcription factor superfamily: Its origin in eukaryotes and expansion in plants. BMC Evol. Biol. 2005, 5, 3. [Google Scholar] [CrossRef] [Green Version]
  200. Olsen, A.N.; Ernst, H.A.; Leggio, L.L.; Skriver, K. NAC transcription factors: Structurally distinct, functionally diverse. Trends Plant Sci. 2005, 10, 79–87. [Google Scholar] [CrossRef]
  201. Ooka, H.; Satoh, K.; Doi, K.; Nagata, T.; Otomo, Y.; Murakami, K.; Matsubara, K.; Osato, N.; Kawai, J.; Carninci, P.; et al. Comprehensive analysis of NAC family genes in Oryza sativa and Arabidopsis thaliana. DNA Res. 2003, 10, 239–247. [Google Scholar] [CrossRef]
  202. Hurst, H.C. Transcription factors 1: bZIP proteins. Protein Profile 1994, 1, 123–168. [Google Scholar] [PubMed]
  203. Hamberger, B.; Bak, S. Plant P450s as versatile drivers for evolution of species-specific chemical diversity. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 2013, 368, 20120426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Reed, J.; Osbourn, A. Engineering terpenoid production through transient expression in Nicotiana benthamiana. Plant Cell Rep. 2018, 37, 1431–1441. [Google Scholar] [CrossRef]
  205. Reed, J.; Stephenson, M.J.; Miettinen, K.; Brouwer, B.; Leveau, A.; Brett, P.; Goss, R.J.M.; Goossens, A.; O’Connell, M.A.; Osbourn, A. A translational synthetic biology platform for rapid access to gram-scale quantities of novel drug-like molecules. Metab. Eng. 2017, 42, 185–193. [Google Scholar] [CrossRef] [PubMed]
  206. Yazaki, K.; Arimura, G.I.; Ohnishi, T. ‘Hidden’ terpenoids in plants: Their biosynthesis, localization and ecological roles. Plant Cell Physiol. 2017, 58, 1615–1621. [Google Scholar] [CrossRef] [Green Version]
  207. Abdollahi, F.; Alebrahim, M.T.; Ngov, C.; Lallemand, E.; Zheng, Y.; Villette, C.; Zumsteg, J.; André, F.; Navrot, N.; Werck-Reichhart, D.; et al. Innate promiscuity of the CYP706 family of P450 enzymes provides a suitable context for the evolution of dinitroaniline resistance in weed. New Phytol. 2021, 229, 3253–3268. [Google Scholar] [CrossRef]
  208. Tholl, D.; Lee, S. Terpene specialized metabolism in Arabidopsis thaliana. Arab. Book 2011, 9, e0143. [Google Scholar] [CrossRef] [Green Version]
  209. Sohrabi, R.; Huh, J.H.; Badieyan, S.; Rakotondraibe, L.H.; Kliebenstein, D.J.; Sobrado, P.; Tholl, D. In planta variation of volatile biosynthesis: An alternative biosynthetic route to the formation of the pathogen-induced volatile homoterpene DMNT via triterpene degradation in Arabidopsis roots. Plant Cell. 2015, 27, 874–890. [Google Scholar] [CrossRef] [PubMed]
  210. Donath, J.; Boland, W. Biosynthesis of acyclic homoterpenes: Enzyme selectivity and absolute configuration of the nerolidol precursor. Phytochemistry 1995, 39, 785–790. [Google Scholar] [CrossRef]
  211. Ginglinger, J.F.; Boachon, B.; Höfer, R.; Paetz, C.; Köllner, T.G.; Miesch, L.; Lugan, R.; Baltenweck, R.; Mutterer, J.; Ullmann, P.; et al. Gene coexpression analysis reveals complex metabolism of the monoterpene alcohol linalool in Arabidopsis flowers. Plant Cell 2013, 25, 4640–4657. [Google Scholar] [CrossRef] [Green Version]
  212. Li, H.; Li, J.; Dong, Y.; Hao, H.; Ling, Z.; Bai, H.; Wang, H.; Cui, H.; Shi, L. Time-series transcriptome provides insights into the gene regulation network involved in the volatile terpenoid metabolism during the flower development of lavender. BMC Plant Biol. 2019, 19, 313. [Google Scholar] [CrossRef] [Green Version]
  213. Boachon, B.; Junker, R.R.; Miesch, L.; Bassard, J.E.; Höfer, R.; Caillieaudeaux, R.; Seidel, D.E.; Lesot, A.; Heinrich, C.; Ginglinger, J.F.; et al. CYP76C1 (Cytochrome P450)-mediated linalool metabolism and the formation of volatile and soluble linalool oxides in Arabidopsis flowers: A strategy for defense against floral antagonists. Plant Cell 2015, 27, 2972–2990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Seifert, A.; Antonovici, M.; Hauer, B.; Pleiss, J. An efficient route to selective bio-oxidation catalysts: An iterative approach comprising modeling, diversification, and screening, based on CYP102A1. Chembiochem. 2011, 12, 1346–1351. [Google Scholar] [CrossRef]
  215. Miettinen, K.; Dong, L.; Navrot, N.; Schneider, T.; Burlat, V.; Pollier, J.; Woittiez, L.; van der Krol, S.; Lugan, R.; Ilc, T.; et al. The seco-iridoid pathway from Catharanthus roseus. Nat. Commun. 2014, 5, 3606. [Google Scholar] [CrossRef] [Green Version]
  216. Kantsa, A.; Raguso, R.A.; Dyer, A.G.; Sgardelis, S.P.; Olesen, J.M.; Petanidou, T. Community-wide integration of floral colour and scent in a Mediterranean scrubland. Nat. Ecol. Evol. 2017, 1, 1502–1510. [Google Scholar] [CrossRef] [PubMed]
  217. Kantsa, A.; Raguso, R.A.; Dyer, A.G.; Olesen, J.M.; Tscheulin, T.; Petanidou, T. Disentangling the role of floral sensory stimuli in pollination networks. Nat. Commun. 2018, 9, 1041. [Google Scholar] [CrossRef] [PubMed]
  218. Aguiar, J.M.R.B.V.; Ferreira, G.S.; Sanches, P.A.; Bento, J.M.S.; Sazima, M. What pollinators see does not match what they smell: Absence of color-fragrance association in the deceptive orchid Ionopsis utricularioides. Phytochemistry 2021, 182, 112591. [Google Scholar] [CrossRef]
  219. Rasouli, O.; Ahmadi, N.; Monfared, S.R.; Sefidkon, F. Physiological, phytochemicals and molecular analysis of color and scent of different landraces of Rosa damascena during flower development stages. Sci. Hortic. 2018, 231, 144–150. [Google Scholar] [CrossRef]
  220. Park, P.H.; Ramya, M.; An, H.R.; Park, P.M.; Lee, S.Y. Breeding of Cymbidium ‘Sale Bit’ with bright yellow flowers and floral scent. Korean J. Breed. Sci. 2019, 51, 258–262. [Google Scholar] [CrossRef] [Green Version]
  221. Adebesin, F.; Widhalm, J.R.; Boachon, B.; Lefèvre, F.; Pierman, B.; Lynch, J.H.; Alam, I.; Junqueira, B.; Benke, R.; Ray, S.; et al. Emission of volatile organic compounds from petunia flowers is facilitated by an ABC transporter. Science 2017, 356, 1386–1388. [Google Scholar] [CrossRef] [Green Version]
  222. Liao., P.; Ray, S.; Boachon, B.; Lynch, J.H.; Deshpande, A.; McAdam, S.; Morgan, J.A.; Dudareva, N. Cuticle thickness affects dynamics of volatile emission from petunia flowers. Nat. Chem. Biol. 2021, 17, 138–145. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biosynthesis and regulation of FVT production (adapted from [113,176]). Enzymes are placed in the green frame. Specialized terpenoids are in the blue frame. Dashed lines indicate multiple enzymatic steps and the transport of metabolites between cell compartments. Metabolites acting as signaling molecules are indicated in red. Unidentified transporters are indicated with question marks. Some transcription factors that interact with enzymes in the biosynthetic pathways are indicated with green boxes and arrows. The location of transcription factors in this illustration does not represent their subcellular location. Abbreviations: AACT, acetyl-CoA acetyltransferase; CDP-ME, 4-diphosphocytidyl-2-C-methyl-d-erythritol; CDP-MEP, 2-phospho-4-(cytidine 5′-diphospho)-2-C-methyl-d-erythritol; CMK, 4-(cytidine 5′-diphospho)-2-C-methyl-d-erythritol kinase; DMAPP, dimethylallyl pyrophosphate; DXP, 1-deoxy-d-xylulose 5-phosphate; DXR, 1-deoxy-D-xylulose 5-phosphate reductoisomerase; DXS, DXP synthase; FPP, farnesyl diphosphate; FPPS, farnesyl diphosphate synthase; GAP, D-glyceraldehyde 3-phosphate; GGPP, geranylgeranyl diphosphate (C20); GGPPS, GGPP synthase; GPP, geranyl diphosphate; GPPS, geranyl diphosphate synthase; HDS, 4-hydroxy-3-methylbut-2-en-1-yl diphosphate synthase; HMBPP, 4-hydroxy-3-methylbut-2-enyldiphosphate; HMBPP, 4-hydroxy-3-methylbut-2-enyldiphosphate; HMG-CoA, 3-hydroxy-3-methylglutaryl-CoA; HMGR, HMG-CoA reductase; HMGS, HMG-CoA synthase; IDI, IPP isomerase; IDS, isoprenyl diphosphate synthase; IP, isopentenyl phosphate; IPK, IP kinase; IPP, isopentenyl diphosphate; MCT, 2-C-methyl-D-erythritol 4-phosphate cytidylyltransferase; MDS, 2-C-methyl-d-erythritol 2,4-cyclodiphosphate synthase; MEcPP, 2-C-methyl-d-erythritol-2,4-cyclodiphosphate; MECPS, ME-CDP synthase; MEP, methylerythritol phosphate; MPD, mevalonate phosphate decarboxylase; MPDC, mevalonate-5-diphosphate decarboxylase; MVA, mevalonic acid; MVK, mevalonate kinase; NUDX, Nudix hydrolase; PMK, phosphomevalonate kinase; TPS, terpene synthase.
Figure 1. Biosynthesis and regulation of FVT production (adapted from [113,176]). Enzymes are placed in the green frame. Specialized terpenoids are in the blue frame. Dashed lines indicate multiple enzymatic steps and the transport of metabolites between cell compartments. Metabolites acting as signaling molecules are indicated in red. Unidentified transporters are indicated with question marks. Some transcription factors that interact with enzymes in the biosynthetic pathways are indicated with green boxes and arrows. The location of transcription factors in this illustration does not represent their subcellular location. Abbreviations: AACT, acetyl-CoA acetyltransferase; CDP-ME, 4-diphosphocytidyl-2-C-methyl-d-erythritol; CDP-MEP, 2-phospho-4-(cytidine 5′-diphospho)-2-C-methyl-d-erythritol; CMK, 4-(cytidine 5′-diphospho)-2-C-methyl-d-erythritol kinase; DMAPP, dimethylallyl pyrophosphate; DXP, 1-deoxy-d-xylulose 5-phosphate; DXR, 1-deoxy-D-xylulose 5-phosphate reductoisomerase; DXS, DXP synthase; FPP, farnesyl diphosphate; FPPS, farnesyl diphosphate synthase; GAP, D-glyceraldehyde 3-phosphate; GGPP, geranylgeranyl diphosphate (C20); GGPPS, GGPP synthase; GPP, geranyl diphosphate; GPPS, geranyl diphosphate synthase; HDS, 4-hydroxy-3-methylbut-2-en-1-yl diphosphate synthase; HMBPP, 4-hydroxy-3-methylbut-2-enyldiphosphate; HMBPP, 4-hydroxy-3-methylbut-2-enyldiphosphate; HMG-CoA, 3-hydroxy-3-methylglutaryl-CoA; HMGR, HMG-CoA reductase; HMGS, HMG-CoA synthase; IDI, IPP isomerase; IDS, isoprenyl diphosphate synthase; IP, isopentenyl phosphate; IPK, IP kinase; IPP, isopentenyl diphosphate; MCT, 2-C-methyl-D-erythritol 4-phosphate cytidylyltransferase; MDS, 2-C-methyl-d-erythritol 2,4-cyclodiphosphate synthase; MEcPP, 2-C-methyl-d-erythritol-2,4-cyclodiphosphate; MECPS, ME-CDP synthase; MEP, methylerythritol phosphate; MPD, mevalonate phosphate decarboxylase; MPDC, mevalonate-5-diphosphate decarboxylase; MVA, mevalonic acid; MVK, mevalonate kinase; NUDX, Nudix hydrolase; PMK, phosphomevalonate kinase; TPS, terpene synthase.
Horticulturae 07 00451 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qiao, Z.; Hu, H.; Shi, S.; Yuan, X.; Yan, B.; Chen, L. An Update on the Function, Biosynthesis and Regulation of Floral Volatile Terpenoids. Horticulturae 2021, 7, 451. https://doi.org/10.3390/horticulturae7110451

AMA Style

Qiao Z, Hu H, Shi S, Yuan X, Yan B, Chen L. An Update on the Function, Biosynthesis and Regulation of Floral Volatile Terpenoids. Horticulturae. 2021; 7(11):451. https://doi.org/10.3390/horticulturae7110451

Chicago/Turabian Style

Qiao, Zhenglin, Huizhen Hu, Senbao Shi, Xuemei Yuan, Bo Yan, and Longqing Chen. 2021. "An Update on the Function, Biosynthesis and Regulation of Floral Volatile Terpenoids" Horticulturae 7, no. 11: 451. https://doi.org/10.3390/horticulturae7110451

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop