13
UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl) UvA-DARE (Digital Academic Repository) Structural diversity in the strigolactones Wang, Y.; Bouwmeester, H.J. DOI 10.1093/jxb/ery091 Publication date 2018 Document Version Final published version Published in Journal of Experimental Botany Link to publication Citation for published version (APA): Wang, Y., & Bouwmeester, H. J. (2018). Structural diversity in the strigolactones. Journal of Experimental Botany, 69(9), 2219-2230. https://doi.org/10.1093/jxb/ery091 General rights It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open content license (like Creative Commons). Disclaimer/Complaints regulations If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You will be contacted as soon as possible. Download date:23 Jul 2021

UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

UvA-DARE (Digital Academic Repository)

Structural diversity in the strigolactones

Wang, Y.; Bouwmeester, H.J.DOI10.1093/jxb/ery091Publication date2018Document VersionFinal published versionPublished inJournal of Experimental Botany

Link to publication

Citation for published version (APA):Wang, Y., & Bouwmeester, H. J. (2018). Structural diversity in the strigolactones. Journal ofExperimental Botany, 69(9), 2219-2230. https://doi.org/10.1093/jxb/ery091

General rightsIt is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s)and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an opencontent license (like Creative Commons).

Disclaimer/Complaints regulationsIf you believe that digital publication of certain material infringes any of your rights or (privacy) interests, pleaselet the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the materialinaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letterto: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. Youwill be contacted as soon as possible.

Download date:23 Jul 2021

Page 2: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

REVIEW PAPER

Structural diversity in the strigolactones

Yanting Wang and Harro J. Bouwmeester*

Plant hormone biology, Swammerdam Institute for Life Sciences, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands

* Correspondence: [email protected]

Received 17 September 2017; Editorial decision 27 February 2018; Accepted 5 March 2018

Editor: Cristina Prandi, Universitá di Torino, Italy

Abstract

Strigolactones (SLs) are a class of signalling molecules secreted by the roots of plants into the rhizosphere. On the one hand, they serve as the signal for recruiting arbuscular mycorrhizal fungi which have a symbiotic relationship with plants. On the other hand, they are also host detection signals for the non-symbiotic, pathogenic, root parasitic plants, which use the SLs as germination stimulants. Finally, recently the SLs were discovered to be a new class of plant hormones that regulate processes such as branching/tillering and root architecture. Intriguingly, >25 different SLs have already been discovered that all have the so-called D-ring but otherwise display many differences in struc-ture and functional groups. In this review, we will critically discuss the structural diversity in the SLs. How are they synthesized in plants; how has this structural diversity possibly evolved; what is the biological relevance of this diver-sity; and what does this imply for the perception of the SLs by receptors in the plant itself and in other organisms? Finally, we conclude that only little is known about the biological significance of this structural diversity, and we will give an outlook on how to elucidate their importance further.

Keywords: Biological relevance, biosynthesis, evolution, perception, receptor, strigolactones, structural diversity.

Introduction

Already in the first half of the previous century scientists had observed that there is a compound or compounds present in the root exudates of clover, maize, sorghum, and linseed that induce(s) the germination of seeds of parasitic plants (Saunders, 1933; Brown et al., 1949, 1951). In the 1960s, the first germination stimulant for witchweed (Striga lutea Lour.) named strigol was isolated from the root exudates of cot-ton (Gossypium hirsutum L.) (Cook et al., 1966, 1972). Since then, >20 different molecules with similar properties to str-igol, collectively called the strigolactones (SLs) (see below), have been identified in the plant kingdom. In this review, we will discuss the structural diversity of the SLs in the differ-ent plant species and what is known about their biosynthesis. We discuss how this structural diversity has possibly evolved and what the biological relevance may be. Finally, we discuss what this structural diversity implies for the perception of these molecules in plants and the other organisms that plants

communicate with through the SLs. Furthermore, we will give an opinion on how to elucidate further the biological sig-nificance of this structural diversity.

Structural diversity in the strigolactones

As described above, the first germination stimulant, strigol, was discovered in the root exudate of cotton. Cotton is not a host of this Striga species (even though it does induce germi-nation of its seeds), but strigol was later also identified in true Striga hosts such as sorghum [Sorghum bicolor (L.) Moench], maize (Zea mays L.), and proso millet (Panicum miliaceum L.) (Siame et al., 1993). Apparently, not only host plant spe-cies but also non-host species produce these germination stimulants. At around the same time, two germination stimu-lants that are chemically closely related to strigol were discov-ered in the root exudates of sorghum [Sorghum bicolor (L.)

© The Author(s) 2018. Published by Oxford University Press on behalf of the Society for Experimental Biology. All rights reserved. For permissions, please email: [email protected]

Journal of Experimental Botany, Vol. 69, No. 9 pp. 2219–2230, 2018doi:10.1093/jxb/ery091 Advance Access publication 7 March 2018

Dow

nloaded from https://academ

ic.oup.com/jxb/article-abstract/69/9/2219/4924111 by U

niversiteit van Amsterdam

user on 11 February 2020

Page 3: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

cv. Haygrazer]—sorgolactone (Hauck et  al., 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol (which was later identified as orobanchyl acetate; Xie et  al., 2008b; Ueno et  al., 2015). In 1995, Butler suggested to call all these strigol-related compounds ‘strigolactones’ (Butler, 1995). Three years later, the first germination stimu-lant of the broomrape, Orobanche minor Smith, was isolated from the root exudate of red clover (Trifolium pratense L.) and named orobanchol (Yokota et al., 1998). This discovery demonstrated that both Striga and Orobanche spp. utilize SLs produced by their host as germination cues to ensure they germinate in the presence of a host root.

In 2008, a novel SL, sorgomol (Fig. 1), a germination stim-ulant for parasitic plants Striga hermonthica and O.  minor, was isolated from Sorghum bicolor (Table  1) (Xie et  al., 2008a). Fabacol and fabacyl acetate (Fig. 1), which stimulate germination of O.  minor, were originally purified from the root exudates of pea (Pisum sativum L.) (Table 1; Xie et al., 2009). Strigone (Fig. 1) that exhibited a highly potent activ-ity on S. hermonthica was isolated from root exudates of the herb, and non-host, Houttuynia cordata (Kisugi et al., 2013). Highly intriguing and surprising was the discovery of new, so-called non-canonical, SLs that lack the so far consistent ABCD-ring structure (Table 1). In Arabidopsis, SL-like 1 (or methyl carlactonoate, MeCLA) was reported (Fig. 1; Table 1) (Abe et  al., 2014; Seto and Yamaguchi, 2014). Another

SL-like molecule, called heliolactone (Fig.  1), was isolated from the root exudate of sunflower (Table  1) (Ueno et  al., 2014). A non-canonical SL, zealactone (called methyl zealac-tonoate by Xie et al., 2017) (Fig. 1), was isolated from maize root exudate (Table 1) and was shown to stimulate S. hermon-thica germination (Charnikhova et al., 2017). Another non-canonical SL, avenaol (Fig.  1), was isolated from the root exudate of black oats (Avena strigosa Schreb.) (Table 1) (Kim et  al., 2014). Avenaol is a potent stimulant for Phelipanche ramosa seeds, but has only a weak effect on the germination of S. hermonthica and O. minor (Kim et al., 2014).

So far, all plant species examined have been shown to exude mixtures of several SLs. Different species have different SL profiles, and sometimes the profile differs between different cultivars within the same species. Furthermore, the amounts and ratios of SLs may vary with different growth conditions and developmental stages (Yoneyama et al., 2009).

Strigolactones are not only germination stimulants

An intriguing question that emerged upon the discovery of strigol was why plants would secrete SLs if these compounds would only have a negative consequence for the host. It took 50  years, until 2005, to answer that question, when Kohki

Fig. 1. The structures of strigolactones. The top row shows examples of strigol-type SLs; the middle two rows show examples of orobanchol-type SLs; and the bottom row shows examples of non-cannonical SLs. Both strigol- and orobanchol-type SLs are canonical SLs.

2220 | Wang and BouwmeesterD

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 4: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

Tab

le 1

. Th

e di

strib

utio

n of

str

igol

acto

nes

in d

iffer

ent p

lant

spe

cies

Str

igo

lact

one

sP

lant

sp

ecie

s

Ara

bid

op

sisa

Tom

ato

Ric

eTo

bac

coS

org

hum

Mai

zeP

etun

iab

Pea

Oat

sS

unfl

ow

erP

opu

lus

Co

tto

n

Can

oni

cal

SLs

Oro

ban

cho

l-ty

pe

SLs

4-D

eoxy

oro

ban

cho

l√

√√

√O

rob

anch

ol

√√

√√

√√

√O

rob

anch

yl a

ceta

te√

√√

√√

√7-

Oxo

oro

ban

cho

l√

7-O

xoo

rob

anch

yl a

ceta

te√

√7-

Hyd

roxy

oro

ban

cho

l√

7-H

ydro

xyo

rob

anch

yl

acet

ate

Fab

aco

l√

Fab

acyl

ace

tate

√S

ola

naco

l√

√S

ola

nacy

l ace

tate

√D

ideh

ydro

-oro

ban

cho

l is

om

ers

(ten

tati

ve)c

√√

Str

igo

l-ty

pe

SLs

5-D

eoxy

stri

go

l√

√√

√4-

Hyd

roxy

-5-d

eoxy

stri

go

l√

4-A

ceto

xy-5

-deo

xyst

rig

ol

√S

org

om

ol

√√

Str

igo

l√

√√

So

rgo

lact

one

√S

trig

yl a

ceta

te√

Did

ehyd

ro-o

rob

anch

ol

iso

mer

s (t

enta

tive

)a

√√

No

n-ca

noni

cal S

LsC

LA√

√√

MeC

LA√

√A

vena

ol

√H

elio

lact

one

√Z

eala

cto

nes

√M

etho

xy-5

-deo

xyst

rig

ol

iso

mer

s (t

enta

tive

)a

a Oro

banc

hol,

orob

anch

yl a

ceta

te, a

nd 5

-deo

xyst

rigol

hav

e be

en re

port

ed in

the

root

exu

date

of A

rabi

dops

is (K

ohle

n et

 al.,

201

0), b

ut th

ey h

ave

not b

een

confi

rmed

by

othe

rs.

b Oro

banc

hyl a

ceta

te, 7

-oxo

orob

anch

yl a

ceta

te, a

nd 7

-hyd

roxy

orob

anch

yl a

ceta

te h

ave

been

repo

rted

in p

etun

ia (Y

oney

ama

et a

l., 2

011)

, but

it h

as n

ot b

een

confi

rmed

by

othe

rs.

c Str

uctu

res

are

unid

entifi

ed; c

lass

ifica

tions

are

tent

ativ

e.

Structural diversity in the strigolactones | 2221D

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 5: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

Akiyama reported that SLs induce hyphal branching in arbus-cular mycorrhizal (AM) fungi and hence promote the sym-biosis of plants with these fungi (Akiyama et al., 2005). AM fungi can engage in a symbiotic interaction with the majority of land plants, and supply water and nutrients, particularly phosphate, which they efficiently obtain from the soil, to the plants, in return for photoassimilates. This symbiotic relation-ship is postulated to be >460 million years old, dating back to when plants moved from water to land. The important role that these fungi play in the acquisition of phosphate (and to a lesser extent nitrogen) from the soil explains why phosphate starvation (and to a lesser extent nitrogen) elevate(s) the SL level so markedly in the root exudates of many plant species (Yoneyama et  al., 2007a, b, 2008; López-Ráez et  al., 2008; Umehara et al., 2008; Jamil et al., 2011).

Three years after the discovery of the role of SLs in sym-biosis, another role for these molecules was discovered. The production of SLs was significantly reduced in a series of mutants with a highly branched/tillering phenotype, which demonstrated that SLs act as a new plant hormone that reg-ulates the above-ground plant architecture (Gomez-Roldan et al., 2008; Umehara et al., 2008). The up-regulation of SL production under conditions of limited phosphate avail-ability mentioned above was later shown also to be key in the reduction of branching/tillering of plants under low phosphate conditions (Umehara et al., 2010; Kohlen et al., 2011). Further studies discovered that SLs also regulate other aspects of plant development including root architecture,

secondary stem growth, and leaf senescence (Ruyter-Spira et al., 2011; Kapulnik and Koltai 2014; Yamada et al., 2014; Sun et al., 2015).

Strigolactone biosynthesis

The naturally occurring SLs can be divided into strigol- and orobanchol-type SLs, which only differ in the stereochemistry of the C-ring [(3aR,8bS) in strigol and (3aS,8bR) in oroban-chol] while the D-ring is always R configured (Figs  1, 2). Together these two types of SLs have been termed ‘canoni-cal strigolactones’ because they exhibit the A-, B-, C-, and D-rings that originally were used to define the SLs (Butler, 1995). It logically appears that the strigol-type SLs are derived from 5-deoxystrigol (5DS); so far there is no direct biosynthetic evidence for this even though it has been reported that 5DS is bioconverted to sorgomol by sorghum (Motonami et al., 2013) (Fig. 2), while orobanchol-type SLs are derived from 4-deoxyorobanchol (Zhang et  al., 2014) (Fig. 2). In contrast, there are also SL-like compounds that do not have the canonical A-, B-, and/or C-part and are therefore termed ‘non-canonical SLs’ (Fig.  2). However, in all cases, the butenolide (D-ring) is connected to the rest of the molecule through an enol ether bridge (Alder et al., 2012) (Figs 1, 2). In our view, it would be better to revise the definition of an SL to ‘a carotenoid-derived molecule with a butenolide D-ring’ such that it includes both the canoni-cal and non-canonical SLs. In canonical SLs, the AB-rings

Fig. 2. Strigolactone biosynthesis. The elucidated core pathway producing the ubiquitous precusror, carlactone, is shown in black. The putative biosynthesis of strigol-type canonical strigolactones is shown in blue, the biosynthesis of non-canonical strigolactones is shown in red, and the biosynthesis of orobanchol-type strigolactones is shown in green.

2222 | Wang and BouwmeesterD

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 6: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

can be modified by demethylation, hydroxylation, epoxida-tion, acetoxylation, and ketolation (Bhattacharya et al., 2009; Al-Babili and Bouwmeester, 2015). Together all these reac-tions give rise to the structural diversification we see in the canonical SLs (Figs 1, 2).

Plants often produce a blend of different SLs (Table  1). Both canonical and non-canonical SLs are produced in some plants such as Arabidopsis (Kohlen et  al., 2011; Abe et al., 2014), maize (Awad et al., 2006; Yoneyama et al., 2015; Charnikhova et al., 2017), and populus (Xie, 2016) (Table 1). Some species, such as tomato, petunia, pea, and populus, only have the orobanchol-type SLs (Table  1), and some species such as tobacco (Xie et al., 2007) and sorghum (Hauck et al., 1992; Siame et al., 1993; Mohemed et al., 2016; Gobena et al., 2017) produce both types of canonical SLs (Table 1). The SL blend can be quite complex. Tomato root exudate, for exam-ple, contains orobanchol, solanacol, hydroxyorobanchol, oxoorobanchol, orobanchyl acetate, and didehydro-oroban-chol isomers (Table 1; Kohlen et al., 2013). Also tobacco has a complex SL composition, not only consisting of various SLs but also with both types: orobanchol-type SLs such as 4-deoxyorobanchol (4DO), orobanchol, orobanchyl acetate, solanacol, and solanacyl acetate, and strigol-type SLs such as 5DS, 4α-hydroxy-5DS, and 4α-acetoxy-5DS (Table 1) (Xie et al., 2007, 2013).

The elucidation of SL biosynthesis started with the discov-ery that root exudates of carotenoid biosynthesis mutants in maize, and of wild-type maize seedlings treated with the carot-enoid biosynthesis inhibitor fluridone, induced lower germi-nation of Striga (Matusova et al., 2005), which suggested that the SLs are derived from a carotenoid. The large structural diversity in the SLs suggests the involvement of many genes in their biosynthesis. Nevertheless, just a handful of SL bio-synthetic genes have been identified, mostly using genetics, particularly from the core SL pathway (Table  2). Mutants with a highly branched phenotype in Arabidopsis showed

that MORE AXILLARY GROWTH 1 (MAX1), MORE AXILLARY GROWTH 3 (MAX3), and MORE AXILLARY GROWTH 4 (MAX4) are essential for the biosynthesis of a shoot branching inhibition signal (Booker et al., 2004, 2005; Gomez-Roldan et  al., 2008; Umehara et  al., 2008). The orthologues of MAX3 and MAX4 are also characterized in other species, such as rice, DWARF17 (D17) (Zou et al., 2006) and DWARF10 (D10) (Arite et al., 2007); pea, RAMOSOUS 5 (RMS5) (Johnson et al., 2006) and RAMOSOUS 1 (RMS1) (Sorefan et al., 2003); and petunia, DECREASED APICAL DOMINANCE 3 (DAD3) (Drummond et  al., 2009) and DECREASED APICAL DOMINANCE 1 (DAD1) (Snowden et  al., 2005). MAX3 and MAX4 encode CAROTENOID CLEAVAGE DIOXYGENASEs 7 and 8 (CCD7 and CCD8), respectively. Consistent with the postulated carotenoid ori-gin of the SLs, it was shown in pea and rice that these two CCDs are required to produce SLs and the shoot branching-inhibiting signal, which may be an SL, an SL precursor, or a compound derived from any of these (Gomez-Roldan et al., 2008; Umehara et  al., 2008). As a third enzyme required for SL biosynthesis, DWARF 27 (D27) was identified, an iron-containing protein with at that time unknown catalytic function (Lin et al., 2009). Later it was shown that D27, an all-trans/9-cis-β-carotene isomerase, catalyses the first dedi-cated step in SL biosynthesis by forming the substrate for the next enzyme in the pathway, CCD7 (Alder et al., 2012). In land plants, there is evidence for the presence of other D27-like proteins (Waters et al., 2012). For example, in the Arabidopsis genome, there are three genes encoding proteins that show superficial similarity to the rice D27. To date, all CCD7 genes identified are single copy (Vallabhaneni et  al., 2010). In contrast, CCD8 has four, two, and six copies in rice, maize, and sorghum, respectively. These copies of CCD8 are divided into group CCD8a and group CCD8b based on their separate clustering in a phylogenetic tree. Intriguingly, CCD8b is not present in Arabidopsis. CCD8a was shown to

Table 2. Genes related to strigolactone biosynthesis and signalling in model plants

Arabidopsis thaliana Oryza sativa Petunia hybrida Pisum sativum

Function of the proteins

SL biosynthesisAtD27 (At1g03350) D27 (Os11g0587000) ? ? Carotenoid isomeraseMAX3 (At2g44990) OsCCD7/D17/HTD1 (Os04g0550600) DAD3 RMS5 Carotenoid cleavage dioxygenaseMAX4 (At4g32810) OsCCD8/D10 (Os01g0746400) DAD1 RMS1 Carotenoid cleavage dioxygenaseMAX1 (At2g26170) OsMAX1 (Os01g0700900, ? ? Cytochrome P450 enzyme; catalytic activity varies

and includes carlactone oxidation to carlactonoic acid (Arabidopsis) and 5-deoxystrigol (rice), and 5-deoxystrigol oxidation to orobanchol

Os01g0701400, Os01g0701500,Os2g0221900, Os06g0565100)

LBO (At3g21420) Os01g0935400 ? ? 2-Oxoglutarate- and Fe(II)-dependent dioxygenase and acting downstream of MAX1 on MeCLA

SL and KAR signallingAtD14 (At3g03990) D14/D88/HTD2 (Os03g0203200) DAD2 RMS4 α/β-Fold hydrolase and receptor for SLMAX2 (At2g42620) D3 (Os06g0154200) PhMAX2A, RMS3 F-box protein component of SCF complex

PhMAX2B

KAI2 (At4g37470) D14L (Os03g0437600) ? ? α/β-Fold hydrolase and receptor for KAR signalingSMXL6 (At1g07200), SMXL7 (At2g29970), SMXL8 (At2g40130)

D53 (Os11g0104300), D53-LIKE (Os12g0104300)

? ? Targets of SL signalling and involved in regulating transcription

Structural diversity in the strigolactones | 2223D

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 7: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

convert 10′-apo-β-carotenal into 13′-apo-β-carotenal. The enzymatic role of the CCD8b group remains elusive.

Together, D27, CCD7, and CCD8 form the core pathway of SL biosynthesis which results in the formation of carlac-tone (CL) (Alder et al., 2012). By now it seems that CL is the common intermediate in the biosynthesis of all SLs discov-ered so far (Al-Babili and Bouwmeester, 2015; Charnikhova et al., 2017). Alder et al. (2012) also speculated that MAX1 would catalyse the next step in the pathway, namely the oxi-dation of CL, to a next at that time unknown pathway inter-mediate. Indeed, it was demonstrated that in rice, one MAX1 homologue, CL oxidase (Os900), catalyses B–C ring closure and stereoselective conversion of CL into 4DO (Zhang et al., 2014) (Fig.  2). Subsequently, another MAX1 homologue, 4DO hydroxylase (Os1400), catalyses the hydroxylation of 4DO to orobanchol (Fig. 2). Orobanchol seems to be a cen-tral intermediate in SL biosynthesis, and it has been postu-lated that further modification of orobanchol can result in the formation of fabacol (Fig. 2) (Xie et al., 2009), orobanchyl acetate (Fig. 2; Ueno et al., 2011b), solanacol (Fig. 2) (Chen et al., 2010), and so on (Fig. 2). In the parallel biosynthetic pathway of the strigol-type SLs, 5DS is postulated to be the precursor for strigol and other strigol-type SLs (Al-Babili and Bouwmeester, 2015) such as sorgomol (Xie et al., 2008a) and strigol (Fig. 2). These can be further modified into other strigol-type SLs, such as strigone (Kisugi et  al., 2013) and strigyl acetate (Sato et al., 2005) from strigol, and sorgolac-tone (Xie et al., 2008a) from sorgomol (Fig. 2).

In Arabidopsis, MAX1 catalyses quite a different reaction, the oxidation of CL to carlactonoic acid (CLA) (Abe et al., 2014). It was shown that CLA can be converted by an as yet unknown methyl transferase into the so-called SL-like1 or methyl carlactonoate (MeCLA) (Abe et al., 2014; Seto and Yamaguchi, 2014). Later, studies showed that an oxoglutarate-dependent dioxygenase, called LATERAL BRANCHING OXIDOREDUCTASE (LBO), facilitates additional process-ing, converting MeCLA into an unknown SL-like product (MeCLA+16 Da) (Brewer et al., 2016) (Fig. 2). Recently, in maize, it was proposed that MeCLA can be converted into zealactone (named methyl zealactonoate by Xie et al., 2017) through hydroxylation of MeCLA at C3, further oxidation (to form heliolactone), double bond epoxidation, and ring cleavage (Charnikhova et  al., 2017) (Fig.  2). Heliolactone, an SL initially isolated from sunflower and closely related to zealactone (Ueno et al., 2014; Charnikhova et al., 2017), was previously also postulated to be derived from MeCLA (Ueno et al., 2014; Fig. 2). In conclusion, although a number of enzymatic steps after CL have now been elucidated, our knowledge of how the large structural diversification in the canonical and non-canonical SLs arises is still very limited.

What evolutionary pressure has caused the structural diversity in the strigolactones?

The evolutionary mechanisms underlying the production of so many different SLs are unknown, but it is likely that the multiple roles of SLs as a signalling molecule in rhizosphere

communication and as endogenous plant hormones have contributed to the diversification. For many plant species, the root parasitic plants—which take water, assimilates, and nutrients from their host—are disastrous enemies that should be avoided. AM fungi, on the contrary, have a symbiotic inter-action with the majority of land plant species, which is ben-eficial. When plants are exposed to phosphate shortage, they produce and release more SLs into the rhizosphere to stimu-late hyphal branching and thus root colonization by the AM fungi that will subsequently help the plant to obtain water and nutrients from the soil. On the one hand, there is a selec-tive pressure for exudation of high amounts of biologically active SLs to induce this hyphal branching effectively in the AM fungi. On the other hand, however, to reduce infection by root parasitic plants, there is a selection pressure to produce SLs that are not or are less active with regard to the para-sitic plant seed germination-inducing activity. Involvement of SLs in additional benign or noxious signalling relationships in the rhizosphere—for which indications are available in the literature (Schlemper et al., 2017)—may further contribute to this selection pressure on the creation of biological specificity through structural changes.

Changes in structure also require concomitant changes in the receptors of these signals. For the role as a plant hor-mone this may be less of an issue as plants could selectively transport SLs to the rhizosphere, independently from the transport to the shoot. This would separate the internal need for SLs to regulate development from the high risk of releas-ing them into the rhizosphere, which can potentially attract enemies, and/or from unnecessary secretion when the plant is already colonized by AM fungi.

From an evolutionary viewpoint, it is beneficial for a host plant to produce an SL with low parasitic plant germina-tion stimulatory activity, but with high AM fungi hyphal branching stimulatory activity. Recently, an example of this was reported in sorghum. Mutant alleles at the LGS1 (LOW GERMINATION STIMULANT 1) locus result in reduced 5DS and enhanced orobanchol levels in sorghum root exu-date which drastically reduces Striga germination stimulant activity but does not negatively affect AM fungi coloniza-tion (Gobena et al., 2017). On the other hand, in the para-sitic plants there is a strong selection pressure to adapt their perception to the changes in the structure of the signalling molecule. The enormous amounts of seeds produced by these parasitic plants form a perfect reservoir for genetic variation and a rapid expansion of a certain genotype. This antago-nistic co-evolution of host and parasites may function as a driver for structural diversification in the SLs of the host plant species.

Biological relevance of structural diversity

The above speculation about the driving force of structural diversification in the SLs only makes sense if there are indi-cations for a biological relevance of this process. There are sufficient indications that there is biological specificity for these structural variants in the processes they control. Some

2224 | Wang and BouwmeesterD

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 8: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

SLs have a high activity in one biological process but much lower in another, and vice versa. Indeed, there are large dif-ferences in the efficiency of different SLs to induce hyphal branching in the AM fungus Gigaspora margarita (Fig.  3). For example, orobanchol and 4α-hydroxy-5DS will induce hyphal branching at only 1 pg per disc, making them effec-tive inducers of hyphal branching in this AM fungus, while strigol and sorgomol are relatively weak inducers with a mini-mum required amount of 100 pg per disc (Akiyama et  al., 2010). Parasitic plant species also vary considerably in their germination responses to different host root exudates and SLs (Fernández-Aparicio et  al., 2009, 2010). In O.  minor, for example, 10 pM 4α-hydroxy-5DS and orobanchol induce 80% and 86% germination, respectively (Kim et  al., 2010) (Fig. 3), while in the same concentration strigol only induces 24% germination. In S.  hermonthica, 0.1  μM 5DS induced 50.9% germination, while 4α-hydroxy-5DS and 4α-acetoxy-5DS in the same concentration only induced 26.5% and 10.9% germination, respectively (Ueno et al., 2011a) (Fig. 3). Intriguingly, this shows that the same SL displays quite dif-ferent germination-inducing activity in different parasitic plants. 4α-Hydroxy-5DS is a low inducer of germination in

Striga but a high inducer of germination in Orobanche seeds. Orobanchol and 4α-hydroxy-5DS are also strong inducers of hyphal branching in the AM fungus, G.  margarita (Fig.  3). That this is relevant in the field is clear from the study on sorghum described above, where a difference in the SL profile between genotypes results in Striga resistance without nega-tively affecting AM colonization (Gobena et al., 2017). Also in the inhibition of shoot branching, different SLs exhibit dif-ferent activities. According to a study by Boyer et al. (2012), orobanchyl acetate and 5DS are relatively active in inhibiting shoot branching, while strigol and orobanchol showed less activity. From the above, we can conclude that different bio-logical processes often have a different SL specificity.

The perception of strigolactones

The different specificities for different SLs in different biolog-ical processes, such as plant development, hyphal branching of AM fungi, and germination of parasitic plants, suggests that the receptor involved in the recognition of the SLs is highly specific. It has indeed been demonstrated that the

Fig. 3. The biological activity of different strigolactones. SLs are 5-deoxystrigol (5DS) (a), 2'-epi-5DS (b), orobanchol (c), 4α-hydroxy-5DS (d), strigol (e), sorgomol (f), orobanchyl acetate (g), 4α-acetoxy-5DS (h), 7-oxoorobanchyl acetate (i), and strigyl acetate (j), respectively. The black columns represent the minimum effective concentration (MEC) of SLs for hyphal branching-inducing activity in Gigaspora margarita (Akiyama et al., 2010); the dark grey columns represent germination-stimulating activity of the strigolactones on Orobanche minor seeds (Kim et al., 2010); the light grey columns represent the germination-stimulating activity of the strigolactones on Striga hermonthica seeds (Ueno et al., 2011a); the white columns represent the node 4 bud outgrowth inhibition activity of the strigolactones in garden pea (Pisum sativum) (Boyer et al., 2012); the concentration of the stigolactones used in Orobanche, Striga, and bud outgrowth assays are 10 pM, 0.1 µM, and 100 nM, respectively. The names of the SLs given above represent the revised stereochemistry (Xie et al., 2013; Zwanenburg and Pospíšil 2013; Ćavar et al., 2015). In Akiyama et al. (2010) a, b, c, d, e, f, g, and i are named (+)-5DS, (+)-2′-epi-5DS, (+)-orobanchol, (+)-2′-epi-orobanchol, (+)-strigol, (+)-sorgomol, (+)-orobanchyl acetate, and (+)-7-oxoorobanchyl acetate, respectively. In Ueno et al. (2011a), d and h are called orobanchol and orobanchyl acetate, respectively. In Kim et al. (2010), c and d are called 2-epi-orobanchol and orobanchol, respectively.

Structural diversity in the strigolactones | 2225D

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 9: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

effect of SLs on plant development and germination of para-sitic plants proceeds via a receptor-mediated mechanism; and it is—considering the extremely low concentrations of SLs that the AM fungi respond to—likely that this also holds for the AM fungi, although there is no direct proof for that yet (Akiyama et  al., 2010; Hamiaux et  al., 2012; Seto and Yamaguchi 2014; Toh et al., 2015; Yao et al., 2017).

Similar to other plant hormones, SL signal transduction is based on hormone-activated proteolysis (Morffy et al., 2016). In this mechanism, the F-box component, first discovered in Arabidopsis and termed MAX2 (see Table 2 for orthologues in other species), of the SCF (Skp1–Cullin–F-box) complex targets a specific protein for the polyubiquitination and deg-radation by the 26S proteasome (Yan et al., 2013; Seto and Yamaguchi, 2014; Yao et al., 2016; Li et al., 2017). The percep-tion of the SLs occurs through the α/β-hydrolase fold super-family protein D14, first identified in rice (for orthologues in other species, see Table 2) (Hamiaux et al., 2012; Nakamura et  al., 2013; Zhao et  al., 2013; Al-Babili and Bouwmeester 2015; de Saint Germain et  al., 2016; Flematti et  al., 2016; Waters et al., 2017). D14 functions as both an enzyme and a receptor (Snowden and Janssen 2016; Yao et al., 2016). It has a conserved catalytic triad (Ser-His-Asp) which is required for the hydrolysis of the SL ligand and its signalling func-tion (Al-Babili and Bouwmeester, 2015; Waters et al., 2017). A conformational change in D14 upon binding and hydroly-sis of the SL ligand allows recruitment of MAX2, which can then target proteins for proteasomal degradation (Al-Babili and Bouwmeester, 2015; Flematti et al., 2016; Waters et al., 2017). Target proteins are D53, that was first identified in rice (Jiang et al., 2013; Zhou et al., 2013), and SMXL6/7/8 [iden-tified as proteolytic targets of SL signaling in Arabidopsis (Wang et  al., 2015)]; for orthologues in other species, see Table 2 (Hamiaux et al., 2012; Soundappan et al., 2015). As a result, D53/SMXLs are ubiquitinated and degraded; at the same time the receptor D14 is degraded and the SL ligand hydrolysed (Jiang et al., 2013; Wang et al., 2015; Yao et al., 2016, 2018). Upon the degradation of D53 and the SMXLs that are repressors of SL signalling (Wang et al., 2015; Waters et al., 2017), expression of the initially repressed genes is acti-vated, which results in the physiological changes incurred by the SLs.

Parasitic plant species such as S.  hermonthica and P.  aegyptiaca also express a seemingly functional D14, the SL receptor (Das et  al., 2015; Y.  Zhang, C.  Ruyter-Spira, H.  Bouwmeester, et  al., unpublished results), and MAX2 (Liu et  al., 2014). Intriguingly, however, the perception of SLs in the seeds of these parasitic plants proceeds dif-ferently from that in all other higher plants. Just like other higher plants, parasitic plants, such as S.  hermonthica and P.  aegyptiaca, have a KAI2 (KARIKIN INSENSITIVE 2) or HYPOSENSITIVE TO LIGHT (HTL) which is a par-alogue of D14 and was discovered in a mutant screen for the induction of germination of Arabidopsis by karrikin (KAR) (Nelson et  al., 2011). KARs are a class of molecules pro-duced by burning vegetation that stimulate the germination of seeds of species pioneering after a fire has destroyed the vegetation (Flematti et  al., 2009; Nelson et  al., 2012), and

structurally display some resemblance to the SLs (Flematti et  al., 2004). Like D14, KAI2 requires the presence of MAX2, in order to be functional. Surprisingly, S. hermon-thica and P. aegyptiaca contain a whole family of KAI2-like proteins, HTLs/KAI2s, of which a subclass (class KAI2d) have evolved to bind SLs as their ligand rather than KARs (Tsuchiya et al., 2015). With a series of elegant experiments they—and not the normal SL receptor D14—were proven to be responsible for the SL-induced seed germination in para-sitic plants (Conn et al., 2015; Tsuchiya et al., 2015; Flematti et al., 2016). There are 11 HTL/KAI2 homologues in S. her-monthica. Using chemical and structural biology, these 11 receptors were clustered into groups with different chemical responsiveness (Conn et  al., 2015). The class KAI2d that includes ShHTL4 to ShHTL9 is most sensitive to a number of natural SLs. ShHTL7 responds to picomolar concen-trations of 5DS, 4DO, and sorgolactone, but responded to strigol in the nanomolar range (Toh et al., 2015). The crys-tal structure of this highly sensitive SL receptor revealed a larger binding pocket than that of the original KAI2 recep-tor (Toh et al., 2015; Yao et al., 2017). ShHTLs seem to have evolved differential SL binding affinities through modula-tion of their active site architecture (Yao et al., 2017). Striga receptors with this larger and modified active site architec-ture had higher sensitivities to SLs compared with that of Arabidopsis HTL, and these evolutionary changes seem to explain their increased SL sensitivity (Toh et al., 2015). With the exception of ShHTL7, the class KAI2d ShHTLs are dif-ferentially sensitive to different SLs (Toh et al., 2015). That is to say, the different receptors of the Striga parasitic plants have different specificity for different SLs.

For unknown reasons, different host plant species pro-duce different blends of SLs, which are characteristic for that plant species. From an evolutionary point of view, it could be that these plant species evolved new SLs that induce less germination in the seeds of the parasitic plants. However, the SL receptors in parasitic plants, the HTLs, appear to have evolved rapidly, possibly allowing the parasites again to rec-ognize the new SLs to continue parasitism of that host. Thus, it seems that the diversity in the receptors in the parasites is a result of the antagonistic co-evolution of the host plants pro-ducing new SLs and of the parasites to adapt their receptor to new structural SLs.

As there seems to be no easily recognizable D14/KAI2 receptor in AM fungi, the discovery of the SL receptor in the AM fungi remains a challenge. Genome and transcriptome sequencing approaches in symbiotic AM fungi will be crucial to assist in the elucidation of their SL perception mechanism. From the biological activity of SLs on AM fungal hyphal branching, we know that different SLs exhibit differences in their stimulating effect. That implies that AM fungi also exhibit SL specificity and thus that there must be a specific receptor. This could also imply that there is host specificity in AM fungi, based on differences in the SL profile, although so far there is no evidence for this. If host specificity does exist, this could help the elucidation of the perception mechanisms of the SLs in AM fungi just as it did in the parasitic plants. The presence of multiple copies of the elusive AM fungi SL

2226 | Wang and BouwmeesterD

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 10: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

receptor would provide additional evidence for co-evolution of structural diversity in the SLs and specificity in the corre-sponding receptors in friends and enemies.

Perspectives

In vitro assays have begun to shed light on the relevance of the structural diversity in the SLs, and in a number of cases their importance was also demonstrated in planta (Gobena et al., 2017). However, for a rigorous evaluation of the biological importance of the different SLs, we need to be able to per-form in planta studies with genotypes differing only in the SL blend that they produce. Mutants and transgenic lines with deletions of single or multiple enzymes will be the material of choice to use in such studies. However, until now, only the core SL biosynthetic pathway is known, and most of the enzymes responsible for the structural diversification of the SLs are still unknown. Elucidating those unknown diversification enzymes by forward genetics approaches will be challenging, as mutations in those enzymes probably only have weak phe-notypes owing to functional redundancy of the different SLs. Transcriptomics and reverse genetics could be an alternative option to elucidate fully the biosynthetic pathway of the SLs. The discovery of LBO in Arabidopsis is an example of this approach (Brewer et al., 2016). With new genes to hand, using TILLING (Targeting Induced Local Lesions IN Genome) or genetic modification, we can change the composition of the SLs in planta and study the biological consequences, such as changes in the recruitment of AM fungi or the infection with parasitic plants. It will also be exciting to use those modified plants to study the rhizosphere microbiome, which could lead to the discovery of new relationships between SLs and other microorganisms.

Elucidation of the role of the HTL receptors in the genera-tion of host-specific races of root parasitic species, and if and how the expression of these receptors are regulated, would increase our understanding of how parasites rapidly evolve new host specificities and become virulent agricultural weeds. A  better understanding of how host specificity is mediated by SLs and how their receptors are regulated would also help to develop selective herbicides for parasitic weeds such as SL analogues and antagonists that highly specifically trigger or inhibit, respectively, the germination of parasitic plant seeds (Johnson et  al., 1976; Babiker and Hamdoun, 1982; Kgosi et  al., 2012; Holbrook-Smith et  al., 2016; Lumbroso et  al., 2016; Samejima et al., 2016; Zwanenburg et al., 2016). The use of such analogues will help to eliminate the seeds from the field and reduce parasitic plant infestation. Engineering the SL profile of crop plants can also be an attractive target for breeders who want to breed parasitic plant-resistant vari-eties of crops. This could possibly be achieved using exist-ing natural variation (such as was done in sorghum; Gobena et al., 2017), through the selection of mutants in target genes using TILLING, or through the use of genetic modification. Since there is a potential possibility that the parasites will adapt to recognize the new structural SLs of resistant varie-ties, farmers and breeders need to take measures to prevent

that by stacking several resistance mechanisms and/or cul-tural measures.

Finally, in the future, more and more new SLs will be iden-tified, further complicating the elucidation of the complete picture of the biological relevance of the structural diversity in the SLs. The elucidation of the biosynthesis of this diver-sity, the possibility to modify it, and a rigorous understand-ing of the perception in the host as well as in other organisms should help us understand how plants and their partners try to balance the positive and negative effects of their chemical communication.

AcknowledgementsThis work was supported by the ERC [Advanced grant CHEMCOMRHIZO (670211) to HJB] and a CSC scholarship (to YW). We thank Kristyna Flokova for her help in drawing structures of SLs, and Yanxia Zhang for helpful discussions.

ReferencesAbe S, Sado A, Tanaka K, et al. 2014. Carlactone is converted to carlactonoic acid by MAX1 in Arabidopsis and its methyl ester can directly interact with AtD14 in vitro. Proceedings of the National Academy of Sciences, USA 111, 18084–18089.

Akiyama K, Matsuzaki K, Hayashi H. 2005. Plant sesquiterpenes induce hyphal branching in arbuscular mycorrhizal fungi. Nature 435, 824–827.

Akiyama K, Ogasawara S, Ito S, Hayashi H. 2010. Structural requirements of strigolactones for hyphal branching in AM fungi. Plant and Cell Physiology 51, 1104–1117.

Al-Babili S, Bouwmeester HJ. 2015. Strigolactones, a novel carotenoid-derived plant hormone. Annual Review of Plant Biology 66, 161–186.

Alder A, Jamil M, Marzorati M, et al. 2012. The path from β-carotene to carlactone, a strigolactone-like plant hormone. Science 335, 1348–1351.

Arite T, Iwata H, Ohshima K, Maekawa M, Nakajima M, Kojima M, Sakakibara H, Kyozuka J. 2007. DWARF10, an RMS1/MAX4/DAD1 ortholog, controls lateral bud outgrowth in rice. The Plant Journal 51, 1019–1029.

Awad AA, Sato D, Kusumoto D, Kamioka H, Takeuchi Y, Yoneyama K. 2006. Characterization of strigolactones, germination stimulants for the root parasitic plants Striga and Orobanche, produced by maize, millet and sorghum. Plant Growth Regulation 48, 221–227.

Babiker AGT, Hamdoun AM. 1982. Factors affecting the activity of GR7 in stimulating germination of Striga hermonthica (Del.) Benth. Weed Research 22, 111–115.

Bhattacharya C, Bonfante P, Deagostino A, Kapulnik Y, Larini P, Occhiato EG, Prandi C, Venturello P. 2009. A new class of conjugated strigolactone analogues with fluorescent properties: synthesis and biological activity. Organic and Biomolecular Chemistry 7, 3413–3420.

Booker J, Auldridge M, Wills S, McCarty D, Klee H, Leyser O. 2004. MAX3/CCD7 is a carotenoid cleavage dioxygenase required for the synthesis of a novel plant signaling molecule. Current Biology 14, 1232–1238.

Booker J, Sieberer T, Wright W, et al. 2005. MAX1 encodes a cytochrome P450 family member that acts downstream of MAX3/4 to produce a carotenoid-derived branch-inhibiting hormone. Developmental Cell 8, 443–449.

Boyer FD, de Saint Germain A, Pillot JP, et al. 2012. Structure–activity relationship studies of strigolactone-related molecules for branching inhibition in garden pea: molecule design for shoot branching. Plant Physiology 159, 1524–1544.

Brewer PB, Yoneyama K, Filardo F, et al. 2016. LATERAL BRANCHING OXIDOREDUCTASE acts in the final stages of strigolactone biosynthesis in Arabidopsis. Proceedings of the National Academy of Sciences, USA 113, 6301–6306.

Structural diversity in the strigolactones | 2227D

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 11: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

Brown R, Greenwood AD, Johnson AW, Long AG. 1951. The stimulant involved in the germination of Orobanche minor Sm. I. Assay technique and bulk preparation of the stimulant. The Biochemical Journal 48, 559–564.

Brown R, Johnson AW, Robinson E, Todd AR. 1949. The stimulant involved in the germination of Striga hermonthica. Proceedings of the Royal Society B: Biological Sciences 136, 1–12.

Butler LG. 1995. Chemical communication between the parasitic weed Striga and its crop host. American Chemical Society 582, 158–168.

Ćavar S, Zwanenburg B, Tarkowski P. 2015. Strigolactones: occurrence, structure, and biological activity in the rhizosphere. Phytochemistry Reviews 14, 691–711.

Charnikhova TV, Gaus K, Lumbroso A, Sanders M, Vincken JP, De Mesmaeker A, Ruyter-Spira CP, Screpanti C, Bouwmeester HJ. 2017. Zealactones. Novel natural strigolactones from maize. Phytochemistry 137, 123–131.

Chen VX, Boyer FD, Rameau C, Retailleau P, Vors JP, Beau JM. 2010. Stereochemistry, total synthesis, and biological evaluation of the new plant hormone solanacol. Chemistry 16, 13941–13945.

Conn CE, Bythell-Douglas R, Neumann D, et al. 2015. Plant evolution. Convergent evolution of strigolactone perception enabled host detection in parasitic plants. Science 349, 540–543.

Cook CE, Whichard LP, Turner B, Wall ME, Egley GH. 1966. Germination of witchweed (Striga lutea Lour.): isolation and properties of a potent stimulant. Science 154, 1189–1190.

Cook CE, Whichard LP, Wall M, Egley GH, Coggon P, Luhan PA, McPhail AT. 1972. Germination stimulants. II. Structure of strigol, a potent seed germination stimulant for witchweed (Striga lutea Lour.). Journal of the American Chemical Society 94, 6198–6199.

Das M, Fernnal of the Ame M, Yang Z, et al. 2015. Parasitic plants Striga and Phelipanche dependent upon exogenous strigolactones for germination have retained genes for strigolactone biosynthesis. American Journal of Plant Science 6, 1151–1166.

de Saint Germain A, Clavé G, Badet-Denisot MA, et al. 2016. An histidine covalent receptor and butenolide complex mediates strigolactone perception. Nature Chemical Biology 12, 787–794.

Drummond RS, Martínez-Sánchez NM, Janssen BJ, Templeton KR, Simons JL, Quinn BD, Karunairetnam S, Snowden KC. 2009. Petunia hybrida CAROTENOID CLEAVAGE DIOXYGENASE7 is involved in the production of negative and positive branching signals in petunia. Plant Physiology 151, 1867–1877.

Fernández-Aparicio M, Flores F, Rubiales D. 2009. Recognition of root exudates by seeds of broomrape (Orobanche and Phelipanche) species. Annals of Botany 103, 423–431.

Fernández-Aparicio M, Yoneyama K, Rubiales D. 2010. The role of strigolactones in host specificity of Orobanche and Phelipanche seed germination. Seed Science Research 21, 55–61.

Flematti GR, Ghisalberti EL, Dixon KW, Trengove RD. 2004. A compound from smoke that promotes seed germination. Science 305, 977.

Flematti GR, Ghisalberti EL, Dixon KW, Trengove RD. 2009. Identification of alkyl substituted 2H-furo[2,3-c]pyran-2-ones as germination stimulants present in smoke. Journal of Agricultural and Food Chemistry 57, 9475–9480.

Flematti GR, Scaffidi A, Waters MT, Smith SM. 2016. Stereospecificity in strigolactone biosynthesis and perception. Planta 243, 1361–1373.

Gobena D, Shimels M, Rich PJ, Ruyter-Spira C, Bouwmeester H, Kanuganti S, Mengiste T, Ejeta G. 2017. Mutation in sorghum LOW GERMINATION STIMULANT 1 alters strigolactones and causes Striga resistance. Proceedings of the National Academy of Sciences, USA 114, 4471–4476.

Gomez-Roldan V, Fermas S, Brewer PB, et al. 2008. Strigolactone inhibition of shoot branching. Nature 455, 189–194.

Hamiaux C, Drummond RS, Janssen BJ, Ledger SE, Cooney JM, Newcomb RD, Snowden KC. 2012. DAD2 is an α/β hydrolase likely to be involved in the perception of the plant branching hormone, strigolactone. Current Biology 22, 2032–2036.

Hauck C, Müller S, Schildknecht H. 1992. A germination stimulant for parasitic flowering plants from Sorghum bicolor, a genuine host plant. Journal of Plant Physiology 139, 474–478.

Holbrook-Smith D, Toh S, Tsuchiya Y, McCourt P. 2016. Small-molecule antagonists of germination of the parasitic plant Striga hermonthica. Nature Chemical Biology 12, 724–729.

Jamil M, Rodenburg J, Charnikhova T, Bouwmeester HJ. 2011. Pre-attachment Striga hermonthica resistance of New Rice for Africa (NERICA) cultivars based on low strigolactone production. New Phytologist 192, 964–975.

Jiang L, Liu X, Xiong G, et al. 2013. DWARF 53 acts as a repressor of strigolactone signalling in rice. Nature 504, 401–405.

Johnson AW, Rosebery G, Parker C. 1976. A novel approach to Striga and Orobanche control using synthetic germination stimulants. Weed Research 16, 223–227.

Johnson X, Brcich T, Dun EA, Goussot M, Haurogné K, Beveridge CA, Rameau C. 2006. Branching genes are conserved across species. Genes controlling a novel signal in pea are coregulated by other long-distance signals. Plant Physiology 142, 1014–1026.

Kapulnik Y, Koltai H. 2014. Strigolactone involvement in root development, response to abiotic stress, and interactions with the biotic soil environment. Plant physiology 166, 560–569.

Kgosi RL, Zwanenburg B, Mwakaboko AS, Murdoch AJ. 2012. Strigolactone analogues induce suicidal seed germination of Striga spp. in soil. Weed Research 52, 197–203.

Kim HI, Kisugi T, Khetkam P, et al. 2014. Avenaol, a germination stimulant for root parasitic plants from Avena strigosa. Phytochemistry 103, 85–88.

Kim HI, Xie X, Kim HS, Chun JC, Yoneyama K, Nomura T, Takeuchi Y, Yoneyama K. 2010. Structure–activity relationship of naturally occurring strigolactones in Orobanche minor seed germination stimulation. Journal of Pesticide Science 35, 344–347.

Kisugi T, Xie X, Kim HI, et al. 2013. Strigone, isolation and identification as a natural strigolactone from Houttuynia cordata. Phytochemistry 87, 60–64.

Kohlen W, Charnikhova T, Bours R, López-Ráez JA, Bouwmeester H. 2013. Tomato strigolactones: a more detailed look. Plant Signaling and Behavior 8, e22785.

Kohlen W, Charnikhova T, Liu Q, et al. 2011. Strigolactones are transported through the xylem and play a key role in shoot architectural response to phosphate deficiency in nonarbuscular mycorrhizal host Arabidopsis. Plant Physiology 155, 974–987.

López-Ráez JA, Charnikhova T, Gómez-Roldán V, et al. 2008. Tomato strigolactones are derived from carotenoids and their biosynthesis is promoted by phosphate starvation. New Phytologist 178, 863–874.

Li H, Yao R, Ma S, Hu S, Li S, Wang Y, Yan C, Xie D, Yan J. 2017. Efficient ASK-assisted system for expression and purification of plant F-box proteins. The Plant Journal 92, 736–743.

Lin H, Wang R, Qian Q, et al. 2009. DWARF27, an iron-containing protein required for the biosynthesis of strigolactones, regulates rice tiller bud outgrowth. The Plant Cell 21, 1512–1525.

Liu Q, Zhang Y, Matusova R, et al. 2014. Striga hermonthica MAX2 restores branching but not the very low fluence response in the Arabidopsis thaliana max2 mutant. New Phytologist 202, 531–541.

Lumbroso A, Villedieu-Percheron E, Zurwerra D, et al. 2016. Simplified strigolactams as potent analogues of strigolactones for the seed germination induction of Orobanche cumana Wallr. Pest Management Science 72, 2054–2068.

Matusova R, Rani K, Verstappen FW, Franssen MC, Beale MH, Bouwmeester HJ. 2005. The strigolactone germination stimulants of the plant-parasitic Striga and Orobanche spp. are derived from the carotenoid pathway. Plant Physiology 139, 920–934.

Mohemed N, Charnikhova T, Bakker EJ, van Ast A, Babiker AG, Bouwmeester HJ. 2016. Evaluation of field resistance to Striga hermonthica (Del.) Benth. in Sorghum bicolor (L.) Moench. The relationship with strigolactones. Pest Management Science 72, 2082–2090.

Morffy N, Faure L, Nelson DC. 2016. Smoke and hormone mirrors: action and evolution of karrikin and strigolactone signaling. Trends in Genetics 32, 176–188.

Motonami N, Ueno K, Nakashima H, Nomura S, Mizutani M, Takikawa H, Sugimoto Y. 2013. The bioconversion of 5-deoxystrigol to sorgomol by the sorghum, Sorghum bicolor (L.) Moench. Phytochemistry 93, 41–48.

2228 | Wang and BouwmeesterD

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 12: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

Nakamura H, Xue YL, Miyakawa T, et al. 2013. Molecular mechanism of strigolactone perception by DWARF14. Nature Communications 4, 2613.

Nelson DC, Flematti GR, Ghisalberti EL, Dixon KW, Smith SM. 2012. Regulation of seed germination and seedling growth by chemical signals from burning vegetation. Annual Review of Plant Biology 63, 107–130.

Nelson DC, Scaffidi A, Dun EA, Waters MT, Flematti GR, Dixon KW, Beveridge CA, Ghisalberti EL, Smith SM. 2011. F-box protein MAX2 has dual roles in karrikin and strigolactone signaling in Arabidopsis thaliana. Proceedings of the National Academy of Sciences, USA 108, 8897–8902.

Ruyter-Spira C, Kohlen W, Charnikhova T, et al. 2011. Physiological effects of the synthetic strigolactone analog GR24 on root system architecture in Arabidopsis: another belowground role for strigolactones? Plant Physiology 155, 721–734.

Samejima H, Babiker AG, Takikawa H, Sasaki M, Sugimoto Y. 2016. Practicality of the suicidal germination approach for controlling Striga hermonthica. Pest Management Science 72, 2035–2042.

Sato D, Awad AA, Takeuchi Y, Yoneyama K. 2005. Confirmation and quantification of strigolactones, germination stimulants for root parasitic plants Striga and Orobanche, produced by cotton. Bioscience, Biotechnology, and Biochemistry 69, 98–102.

Saunders AR. 1933. Studies in phanerogamic parasitism, with particular reference to Striga lutea Lour. South African Department of Agriculture Science Bulletin No. 128.

Schlemper TR, Leite MFA, Lucheta AR, Shimels M, Bouwmeester HJ, van Veen JA, Kuramae EE. 2017. Rhizobacterial community structure differences among sorghum cultivars in different growth stages and soils. FEMS Microbiology Ecology 93, doi: 10.1093/femsec/fix096.

Seto Y, Yamaguchi S. 2014. Strigolactone biosynthesis and perception. Current Opinion in Plant Biology 21, 1–6.

Siame BA, Weerasuriya Y, Wood K, Ejeta G, Butler LG. 1993. Isolation of strigol, a germination stimulant for Striga asiatica, from host plants. Journal of Agricultural and Food Chemistry 41, 1486–1491.

Snowden KC, Janssen BJ. 2016. Structural biology: signal locked in. Nature 536, 402–404.

Snowden KC, Simkin AJ, Janssen BJ, et al. 2005. The Decreased apical dominance1/Petunia hybrida CAROTENOID CLEAVAGE DIOXYGENASE8 gene affects branch production and plays a role in leaf senescence, root growth, and flower development. The Plant Cell 17, 746–759.

Sorefan K, Booker J, Haurogné K, et al. 2003. MAX4 and RMS1 are orthologous dioxygenase-like genes that regulate shoot branching in Arabidopsis and pea. Genes and Development 17, 1469–1474.

Soundappan I, Bennett T, Morffy N, Liang Y, Stanga JP, Abbas A, Leyser O, Nelson DC. 2015. SMAX1-LIKE/D53 family members enable distinct MAX2-dependent responses to strigolactones and karrikins in Arabidopsis. The Plant Cell 27, 3143–3159.

Sun H, Tao J, Hou M, et al. 2015. A strigolactone signal is required for adventitious root formation in rice. Annals of Botany 115, 1155–1162.

Toh S, Holbrook-Smith D, Stogios PJ, Onopriyenko O, Lumba S, Tsuchiya Y, Savchenko A, McCourt P. 2015. Structure–function analysis identifies highly sensitive strigolactone receptors in Striga. Science 350, 203–207.

Tsuchiya Y, Yoshimura M, Sato Y, et al. 2015. Parasitic plants. Probing strigolactone receptors in Striga hermonthica with fluorescence. Science 349, 864–868.

Ueno K, Fujiwara M, Nomura S, Mizutani M, Sasaki M, Takikawa H, Sugimoto Y. 2011a. Structural requirements of strigolactones for germination induction of Striga gesnerioides seeds. Journal of Agricultural and Food Chemistry 59, 9226–9231.

Ueno K, Furumoto T, Umeda S, Mizutani M, Takikawa H, Batchvarova R, Sugimoto Y. 2014. Heliolactone, a non-sesquiterpene lactone germination stimulant for root parasitic weeds from sunflower. Phytochemistry 108, 122–128.

Ueno K, Nomura S, Muranaka S, Mizutani M, Takikawa H, Sugimoto Y. 2011b. Ent-2'-epi-Orobanchol and its acetate, as germination stimulants for Striga gesnerioides seeds isolated from cowpea and red clover. Journal of Agricultural and Food Chemistry 59, 10485–10490.

Ueno K, Sugimoto Y, Zwanenburg B. 2015. The genuine structure of alectrol: end of a long controversy. Phytochemistry Reviews 14, 835–847.

Umehara M, Hanada A, Magome H, Takeda-Kamiya N, Yamaguchi S. 2010. Contribution of strigolactones to the inhibition of tiller bud outgrowth under phosphate deficiency in rice. Plant and Cell Physiology 51, 1118–1126.

Umehara M, Hanada A, Yoshida S, et al. 2008. Inhibition of shoot branching by new terpenoid plant hormones. Nature 455, 195–200.

Vallabhaneni R, Bradbury LM, Wurtzel ET. 2010. The carotenoid dioxygenase gene family in maize, sorghum, and rice. Archives of Biochemistry and Biophysics 504, 104–111.

Wang L, Wang B, Jiang L, et al. 2015. Strigolactone signaling in Arabidopsis regulates shoot development by targeting D53-Like SMXL repressor proteins for ubiquitination and degradation. The Plant Cell 27, 3128–3142.

Waters MT, Brewer PB, Bussell JD, Smith SM, Beveridge CA. 2012. The Arabidopsis ortholog of rice DWARF27 acts upstream of MAX1 in the control of plant development by strigolactones. Plant Physiology 159, 1073–1085.

Waters MT, Gutjahr C, Bennett T, Nelson DC. 2017. Strigolactone signaling and evolution. Annual Review of Plant Biology 68, 291–322.

Xie X. 2016. Structural diversity of strigolactones and their distribution in the plant kingdom. Journal of Pesticide Science 41, 175–180.

Xie X, Kisugi T, Yoneyama K, Nomura T, Akiyama K, Uchida K, Yokota T, McErlean CSP, Yoneyama K. 2017. Methyl zealactonoate, a novel germination stimulant for root parasitic weeds produced by maize. Journal of Pesticide Science 42, 58–61.

Xie X, Kusumoto D, Takeuchi Y, Yoneyama K, Yamada Y, Yoneyama K. 2007. 2'-Epi-orobanchol and solanacol, two unique strigolactones, germination stimulants for root parasitic weeds, produced by tobacco. Journal of Agricultural and Food Chemistry 55, 8067–8072.

Xie X, Yoneyama K, Harada Y, Fusegi N, Yamada Y, Ito S, Yokota T, Takeuchi Y, Yoneyama K. 2009. Fabacyl acetate, a germination stimulant for root parasitic plants from Pisum sativum. Phytochemistry 70, 211–215.

Xie X, Yoneyama K, Kisugi T, et al. 2013. Confirming stereochemical structures of strigolactones produced by rice and tobacco. Molecular Plant 6, 153–163.

Xie X, Yoneyama K, Kusumoto D, Yamada Y, Takeuchi Y, Sugimoto Y, Yoneyama K. 2008a. Sorgomol, germination stimulant for root parasitic plants, produced by Sorghum bicolor. Tetrahedron Letters 49, 2066–2068.

Xie X, Yoneyama K, Kusumoto D, Yamada Y, Yokota T, Takeuchi Y, Yoneyama K. 2008b. Isolation and identification of alectrol as (+)-orobanchyl acetate, a germination stimulant for root parasitic plants. Phytochemistry 69, 427–431.

Yamada Y, Furusawa S, Nagasaka S, Shimomura K, Yamaguchi S, Umehara M. 2014. Strigolactone signaling regulates rice leaf senescence in response to a phosphate deficiency. Planta 240, 399–408.

Yan J, Li H, Li S, Yao R, Deng H, Xie Q, Xie D. 2013. The Arabidopsis F-box protein CORONATINE INSENSITIVE1 is stabilized by SCFCOI1 and degraded via the 26S proteasome pathway. The Plant Cell 25, 486–498.

Yao R, Ming Z, Yan L, et al. 2016. DWARF14 is a non-canonical hormone receptor for strigolactone. Nature 536, 469–473.

Yao R, Wang F, Ming Z, Du X, Chen L, Wang Y, Zhang W, Deng H, Xie D. 2017. ShHTL7 is a non-canonical receptor for strigolactones in root parasitic weeds. Cell Research 27, 838–841.

Yao R, Wang L, Li Y, et al. 2018. Rice DWARF14 acts as an unconventional hormone receptor to restore strigolactone signaling in the Arabidopsis d14 mutant. Journal of Experimental Botany 69, 2355–2365.

Yokota T, Sakai H, Okuno K, Yoneyama K, Takeuchi Y. 1998. Alectrol and orobanchol, germination stimulants for Orobanche minor, from its host red clover. Phytochemistry 49, 1967–1973.

Yoneyama K, Arakawa R, Ishimoto K, et al. 2015. Difference in Striga-susceptibility is reflected in strigolactone secretion profile, but not in compatibility and host preference in arbuscular mycorrhizal symbiosis in two maize cultivars. New Phytologist 206, 983–989.

Yoneyama K, Xie X, Kisugi T, Nomura T, Sekimoto H, Yokota T, Yoneyama K. 2011. Characterization of strigolactones exuded by Asteraceae plants. Plant Growth Regulation 65, 495–504.

Yoneyama K, Xie X, Kusumoto D, Sekimoto H, Sugimoto Y, Takeuchi Y, Yoneyama K. 2007a. Nitrogen deficiency as well as phosphorus deficiency in sorghum promotes the production and exudation

Structural diversity in the strigolactones | 2229D

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020

Page 13: UvA-DARE (Digital Academic Repository) Structural diversity ...cv. Haygrazer]—sorgolactone (et al.Hauck , 1992)—and in the root exudate of cowpea [Vigna unguiculata cv (Walp)]—alectrol

of 5-deoxystrigol, the host recognition signal for arbuscular mycorrhizal fungi and root parasites. Planta 227, 125–132.

Yoneyama K, Xie X, Sekimoto H, Takeuchi Y, Ogasawara S, Akiyama K, Hayashi H, Yoneyama K. 2008. Strigolactones, host recognition signals for root parasitic plants and arbuscular mycorrhizal fungi, from Fabaceae plants. New Phytologist 179, 484–494.

Yoneyama K, Xie X, Yoneyama K, Takeuchi Y. 2009. Strigolactones: structures and biological activities. Pest Management Science 65, 467–470.

Yoneyama K, Yoneyama K, Takeuchi Y, Sekimoto H. 2007b. Phosphorus deficiency in red clover promotes exudation of orobanchol, the signal for mycorrhizal symbionts and germination stimulant for root parasites. Planta 225, 1031–1038.

Zhang Y, van Dijk AD, Scaffidi A, et al. 2014. Rice cytochrome P450 MAX1 homologs catalyze distinct steps in strigolactone biosynthesis. Nature Chemical Biology 10, 1028–1033.

Zhao LH, Zhou XE, Wu ZS, et al. 2013. Crystal structures of two phytohormone signal-transducing α/β hydrolases: karrikin-signaling KAI2 and strigolactone-signaling DWARF14. Cell Research 23, 436–439.

Zhou F, Lin Q, Zhu L, et al. 2013. D14-SCF(D3)-dependent degradation of D53 regulates strigolactone signalling. Nature 504, 406–410.

Zou J, Zhang S, Zhang W, et al. 2006. The rice HIGH-TILLERING DWARF1 encoding an ortholog of Arabidopsis MAX3 is required for negative regulation of the outgrowth of axillary buds. The Plant Journal 48, 687–698.

Zwanenburg B, Mwakaboko AS, Kannan C. 2016. Suicidal germination for parasitic weed control. Pest Management Science 72, 2016–2025.

Zwanenburg B, Pospísil T. 2013. Structure and activity of strigolactones: new plant hormones with a rich future. Molecular Plant 6, 38–62.

2230 | Wang and BouwmeesterD

ownloaded from

https://academic.oup.com

/jxb/article-abstract/69/9/2219/4924111 by Universiteit van Am

sterdam user on 11 February 2020