ArticlePDF AvailableLiterature Review

The Effects of Domestication on Secondary Metabolite Composition in Legumes

Frontiers
Frontiers in Genetics
Authors:

Abstract and Figures

Legumes are rich in secondary metabolites, such as polyphenols, alkaloids, and saponins, which are important defense compounds to protect the plant against herbivores and pathogens, and act as signaling molecules between the plant and its biotic environment. Legume-sourced secondary metabolites are well known for their potential benefits to human health as pharmaceuticals and nutraceuticals. During domestication, the color, smell, and taste of crop plants have been the focus of artificial selection by breeders. Since these agronomic traits are regulated by secondary metabolites, the basis behind the genomic evolution was the selection of the secondary metabolite composition. In this review, we will discuss the classification, occurrence, and health benefits of secondary metabolites in legumes. The differences in their profiles between wild legumes and their cultivated counterparts will be investigated to trace the possible effects of domestication on secondary metabolite compositions, and the advantages and drawbacks of such modifications. The changes in secondary metabolite contents will also be discussed at the genetic level to examine the genes responsible for determining the secondary metabolite composition that might have been lost due to domestication. Understanding these genes would enable breeding programs and metabolic engineering to produce legume varieties with favorable secondary metabolite profiles for facilitating adaptations to a changing climate, promoting beneficial interactions with biotic factors, and enhancing health-beneficial secondary metabolite contents for human consumption.
This content is subject to copyright.
Frontiers in Genetics | www.frontiersin.org 1 September 2020 | Volume 11 | Article 581357
REVIEW
published: 18 September 2020
doi: 10.3389/fgene.2020.581357
Edited by:
Jiayin Pang,
University of Western Australia,
Australia
Reviewed by:
Nicolas Rispail,
Spanish National Research Council,
Spain
Eric Von Wettberg,
University of Vermont, UnitedStates
*Correspondence:
Gyuhwa Chung
chung@chonnam.ac.kr
Hon-Ming Lam
honming@cuhk.edu.hk
Specialty section:
This article was submitted to
Evolutionary and Population
Genetics,
a section of the journal
Frontiers in Genetics
Received: 08 July 2020
Accepted: 31 August 2020
Published: 18 September 2020
Citation:
Ku Y-S, Contador CA, Ng M-S, Yu J,
Chung G and Lam H-M (2020)
The Effects of Domestication on
Secondary Metabolite
Composition in Legumes.
Front. Genet. 11:581357.
doi: 10.3389/fgene.2020.581357
The Effects of Domestication on
Secondary Metabolite Composition
in Legumes
Yee-ShanKu1, CarolinaA.Contador1, Ming-SinNg1, JeongjunYu2, GyuhwaChung2* and
Hon-MingLam1
*
1 Centre for Soybean Research of the State Key Laboratory of Agrobiotechnology and School of Life Sciences, The Chinese
University of Hong Kong, Shatin, China, 2 Department of Biotechnology, Chonnam National University, Yeosu, South Korea
Legumes are rich in secondary metabolites, such as polyphenols, alkaloids, and saponins,
which are important defense compounds to protect the plant against herbivores and
pathogens, and act as signaling molecules between the plant and its biotic environment.
Legume-sourced secondary metabolites are well known for their potential benets to
human health as pharmaceuticals and nutraceuticals. During domestication, the color,
smell, and taste of crop plants have been the focus of articial selection by breeders.
Since these agronomic traits are regulated by secondary metabolites, the basis behind
the genomic evolution was the selection of the secondary metabolite composition. In this
review, wewill discuss the classication, occurrence, and health benets of secondary
metabolites in legumes. The differences in their proles between wild legumes and their
cultivated counterparts will beinvestigated to trace the possible effects of domestication
on secondary metabolite compositions, and the advantages and drawbacks of such
modications. The changes in secondary metabolite contents will also bediscussed at
the genetic level to examine the genes responsible for determining the secondary
metabolite composition that might have been lost due to domestication. Understanding
these genes would enable breeding programs and metabolic engineering to produce
legume varieties with favorable secondary metabolite proles for facilitating adaptations
to a changing climate, promoting benecial interactions with biotic factors, and enhancing
health-benecial secondary metabolite contents for human consumption.
Keywords: legume, domestication, secondary metabolite, defense, health benet
INTRODUCTION
Climate change, farmland deterioration, and the resulting food insecurity are major challenges
facing the world. An increase in food supply is required to feed the expanding human population.
e cultivation of high-yield crops has been used as a strategy to improve food supply. Grain
legumes have been suggested as the potential solution to maintaining food and protein security
(Considine etal., 2017). Legumes are also benecial for sustainable agriculture due to the reduced
release of greenhouse gases compared to other crops (Stagnari et al., 2017). Besides the benecial
eects on the improvement of soil fertility, legumes could enhance the resistance of soil to
ecosystem disturbance, possibly due to the enhanced soil food web complexity (Gao etal., 2020).
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 2 September 2020 | Volume 11 | Article 581357
In agriculture, legumes are common candidates for crop rotation
for promoting the growth of other crops such as cereals (Bagayoko
et al., 2000; Uzoh et al., 2019). In addition, legumes produce
unique secondary metabolites such as isoavones, which are
benecial to human health (Gepts et al., 2005; Ku et al., 2020).
Legumes are known to protect humans from chronic diseases,
including cardiovascular diseases, diabetes, obesity, osteoporosis,
or even cancer (Kushi etal., 1999; Al-Anazi et al., 2011). Based
on the mode of consumption, legumes can be classied into
four groups: oil seeds, pulses, vegetable crops, and feed crops
(McCrory et al., 2010). Examples of oil seeds are soybean and
peanut (McCrory et al., 2010). Pulses are legumes, which are
exclusively harvested as dry seeds, such as chickpea, lentils, and
peas. Green bean and garden pea are examples of vegetable
crops while clover and alfalfa are examples of feed crops (McCrory
et al., 2010). Human selection of legumes during domestication
has resulted in the alteration, and even loss of diversity, of
secondary metabolite contents in these crops, directly and indirectly
through the selection pressure on the genes that control the
production of secondary metabolites. Understanding the dierences
in secondary metabolites, and the underlying genetic dierences,
between the domesticated legume cultivars and their wild
progenitors would promote the preservation of legume accessions,
which possess the genes for the biosynthesis of benecial secondary
metabolites. is knowledge will facilitate breeding programs
and metabolite engineering to produce legume crops with favorable
traits for adapting to the changing climate and for human
pharmaceutical/nutraceutical use.
SEVERAL DOMESTICATION-RELATED
TRAITS ALTERED THE SECONDARY
METABOLITE CONTENTS
Domestication traits refer to morphological, biochemical,
developmental, or physiological traits that are dierent between
domesticated plants and their immediate wild progenitors
(Abbo et al., 2014). A key part of domestication is the
improvement of crop yield and harvestability compared to
the wild progenitors (Dehaan etal., 2016). Several crop traits,
including pod shattering, peduncle length, oral color, days
to owering, 100-seed weight, pod length, leaf length, leaf
width, and seed number per pod, have been regarded as
domestication-related traits (Lo et al., 2018).
Besides yield and harvestability related traits, other
agronomic traits, such as seed size, appearance, and taste,
are also subject to selection by breeders. ese traits could
be regarded as improvements due to post-domestication
selection (Abbo et al., 2014). It has been suggested that the
selection for larger seeds is related to facilitating single-seed
planting (Kaplan, 1981). Breeders have also selected seeds of
light colors. e ease of sowing and religious reasons have
been proposed to bebehind such conscious selections (Heiser,
1988). erefore, seeds of modern legumes tend to have larger
sizes and lighter colors compared to their wild counterparts.
Moreover, the bitter taste of seeds has been intentionally
eliminated through breeding (Muzquiz et al., 1994). Behind the
loss of bitter taste is the loss of the corresponding bitter-tasting
secondary metabolites such as alkaloids (Muzquiz etal., 1994).
During domestication, secondary metabolite compositions
which facilitate cultivation and improve the appearance and
taste of food grains were intentionally selected for by breeders.
In some cases, the secondary metabolite composition may
be unintentionally selected due to the close proximity of the
genes or quantitative trait loci (QTLs) for secondary metabolite
biosynthesis to those regulating other traits such as major
nutrients and yield. e selection of favorable cultivation areas
and the protection by breeders during crop growth limit natural
selection pressures due to abiotic and biotic stresses. Domestication
brings forth better yield, better taste, and better appearance but
also reduces the availability of secondary metabolites in legumes.
As a result, domesticated legumes are usually less resistant to
biotic stresses compared to their wild counterparts (Muzquiz
et al., 1994; Pavan etal., 2016; Bazghaleh et al., 2018; Abraham
etal., 2019). e reduced availability of health-benecial secondary
metabolites (Muzquiz etal., 1994; Wang etal., 2010; Fernández-
marín et al., 2014; Kaur et al., 2019) also limits the potential
of legumes as sources of bioactive compounds for pharmaceutical
use. For the growth of the legume plants, the loss of the secondary
metabolites in modern cultivars possibly renders the plants more
susceptible to abiotic stress and biotic stress. e importance
of the secondary metabolites to combating these stresses will
be introduced in section “e Roles of Secondary Metabolites
in Combating Abiotic and Biotic Interactions.
INTRODUCTION TO SECONDARY
METABOLITES IN LEGUMES
Denition of Plant Secondary Metabolites
Secondary metabolites are organic compounds derived from
primary metabolism that serves key roles in defense and signaling
in plants. ey contribute to adaptive traits and ecological
tness, including defense mechanisms, tolerance to abiotic/biotic
stresses, and interactions with insect pollinators, root-associated
microbes, and herbivores. In contrast, primary metabolites are
essential for cellular functions, such as growth, development,
and reproduction. For example, secondary metabolites can attract
insects for pollination or symbiotic rhizobia for nitrogen-xing
nodule formation. ey can also bepart of the defense mechanisms
against herbivores, disease-causing bacteria, fungi, viruses, and
parasites. ere are also a wide range of secondary metabolites
with pharmaceutical, nutraceutical, and toxicological values for
humans (Wink, 2013). e contents of secondary metabolites
vary among dierent plant species (Böttger etal., 2018). Legumes
are rich in secondary metabolites, such as polyphenols, alkaloids,
and saponins (Gupta, 1987).
The Health Benets of Secondary
Metabolites From Legumes
In the recent past, many secondary metabolites in legumes
were considered non-nutritive. For example, tannins,
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 3 September 2020 | Volume 11 | Article 581357
glycosides, alkaloids, and saponins affect the digestibility of
beans (Gupta, 1987). However, more and more evidence
suggests there are health benefits from the secondary
metabolites of legumes (Dixon and Sumner, 2003). The
health benefits of carotenoids, polyphenols, alkaloids, and
saponins, all abundant in legumes, are discussed below and
summarized in Tab l e 1 .
Carotenoids
Carotenoids are a type of tetraterpenoids, ranging from bright
yellow and orange to red, found in algae, photosynthetic bacteria,
and plants, including carrot, pumpkin, tomato, sweet potato,
and papaya. ey can be classied into two groups, carotenes
and xanthophylls (Roberts et al., 2009). Xanthophylls dier
from carotenes by having oxygenated substituents in their
molecules (Roberts etal., 2009). Carotenoids with unsubstituted
β-rings, including α-carotene, β-carotene, and β-cryptoxanthin,
act as provitamin A (Roberts et al., 2009). e carotenoid
compositions in various legume seeds have been previously
summarized (Tee etal., 1995). Lutein and zeaxanthin constitute
the macular pigments in the retina of the mammalian eye.
e oxygenated nature of the lutein and zeaxanthin molecules
provides antioxidative protection for the eye from damage by
free radicals (Roberts etal., 2009). e prevention of age-related
macular degeneration by the consumption of carotenoid-rich
foods has been recommended (Bernstein et al., 2016).
Polyphenols
Polyphenols are the major determinants of tissue colors, and
generally possess antioxidative activities (Abbas et al., 2017).
Polyphenols in plants can be classied into two groups: phenolic
acids and avonoids (Abbas et al., 2017). e occurrence and
TABLE1 | Classication of secondary metabolites in legumes and their benets to human health.
Groups Sub-groups Examples in legumes Occurrence in
legumes
Benet(s) to human health References
Polyphenols Flavonoids Quercetin, kaempferol Widely distributed Reduction in ischemic heart
disease, reduction in body
weight
(Knekt etal., 2002)
Isoavones Genistein, daidzin Soybean seeds Phytoestrogen, antioxidant,
antimicrobial and anti-
inammatory properties,
reduction of risk in
cardiovascular diseases,
diabetes, obesity, and
osteoporosis
(Křížová etal., 2019)
Catechin Catechin, epicatechin,
gallo-catechin
Broad bean, chickpea,
cowpea, kidney-bean,
lentil, peanut
Reduction in heart disease,
improvement of sperm motility
and viability
(Arts etal., 2000;
Hollman and Arts, 2000;
Ojwang etal., 2013; Dias
etal., 2016; López-cortez
etal., 2016; Quintero-
soto etal., 2018)
Anthocyanins Pelargonidin, cyanidin,
malvidin, petunidin
Widely distributed Antioxidant and anti-
inammatory properties, lipid
peroxidation, DNA cleavage
protection
(Acquaviva etal., 2003;
Pietta etal., 2003; Rossi
etal., 2012)
Terpenoids and steroid Triterpenoid saponins Saponins Chickpea, soybean,
lentils, peanut, common
bean, and alfalfa sprouts
Reduction of cholesterol
content, antimicrobial and
anti-cancer properties
(Shi etal., 2004, 2014;
Hassan etal., 2010; Man
etal., 2010; Marrelli
etal., 2016)
Tetraterpenes Carotenoids Widely distributed Antioxidant, better visual
function, reduction of
cardiovascular diseases
(Voutilainen etal., 2006;
Roberts etal., 2009)
Alkaloids Quinolizidine alkaloids (QA) Sparteine Lupinus spp. Antimicrobial properties (Romeo etal., 2018)
Pyrroloindole alkaloids Physostigmine Ordeal bean Treatment of Alzheimer’s
disease and Parkinson’s
disease
(Zhu etal., 2014; Kumar
etal., 2015)
Peptides Polypeptide Lunasin Soybean anti-inammatory properties,
reduction of cholesterol
content, antioxidant,
anticancer and anti-
atherosclerotic activities
(Jeong etal., 2002,
2003, 2007, 2009; Hsieh
etal., 2017; Fernández-
tomé and Hernández-
ledesma, 2019)
Protease inhibitors Angiotensin-I converting
enzyme inhibitors
Pea, chickpea, mung
bean, soybean, lentil
Lowering blood pressure and
risk of heart failure
(Zhang etal., 2018)
Amines Polyamine spermine, spermidine Common bean, white
clover, mung bean
Antioxidant activities,
reduction of cardiovascular
diseases
(Soda, 2010; Menéndez
etal., 2019; Muñoz-Esparza
etal., 2019)
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 4 September 2020 | Volume 11 | Article 581357
health benets of phenolic acids in grain legumes have been
previously summarized (Singh etal., 2017). Flavonoids are classied
into several sub-classes: avones, avonols, avanones, avanonols,
anthocyanins, avanols, and isoavones (Ku etal., 2020). Among
avonoids, isoavones are only found in legumes. Flavonoids
have multiple functions in plants, for example, mediating the
responses to biotic and abiotic stresses, controlling the transport
of auxins, acting as UV radiation-absorbing pigments to protect
the plant against UV damage, attracting pollinating insects,
interacting with rhizobia to initiate nodulation for symbiotic
nitrogen xation, and regulating defense against pathogens and
herbivores through phytoalexin activities (Kumar and Pandey,
2013). For human health, it has been reported that avonoids
can act as protectants against cellular oxidation, inammation,
viral infections, and cancer (Kleemann etal., 2011). e molecular
mechanisms of the health benets of avonoids have been recently
reviewed (Ku et al., 2020).
Alkaloids
Alkaloids are nitrogen-containing organic heterocyclic
compounds that are biologically active. Many alkaloids have
pharmaceutical properties. For example, some alkaloids were
found to have anti-malarial activities (Onguéné et al., 2013),
anticancer activities (Gupta etal., 2015), and abilities to facilitate
blood circulation in the brain and to prevent stroke (Kumar
and Khanum, 2012). Moreover, several studies reported that
alkaloids have potential therapeutic eects on neurodegenerative
diseases, such as Alzheimer’s disease, Parkinsons disease, and
Huntington disease (Amirkia and Heinrich, 2014).
Saponins
Saponins are a group of terpenoids found in plants, including
onion, ginger, garlic, ginseng, fenugreek, and legumes (Oakenfull,
1981; Sauvaire etal., 1996). ese crops are important sources
of saponins in the human diet (Oakenfull, 1981; Sauvaire etal.,
1996). Chickpea, soybean, lentils, peanut, garden pea, broad
bean, and alfalfa are rich in saponins (Oakenfull, 1981). e
antibacterial and foaming properties of saponins led to the
use of saponins as vaccine adjuvants (Marciani, 2018). In the
human body, saponins can bind to bile salts to reduce cholesterol
absorption (Marrelli et al., 2016). Moreover, in rats, it was
shown that a saponin-rich diet resulted in the reduction of
body weight, total cholesterol, triglycerides, very-low-density
lipoproteins (VLDL), and low-density lipoproteins (LDL) in
serum (Latha et al., 2011; Reddy et al., 2012). Alfalfa saponin
extract (ASE) was found to have cholesterol-lowering eects
(Wang et al., 2011; Marrelli et al., 2016). e treatment of
rats with ASE led to the enhanced expression of cholesterol
7-alpha-hydroxylase (Cyp7a1), an enzyme involved in the bile
acid biosynthetic pathway in the livers of hyperlipidemic rats
(Marrelli et al., 2016). Besides, ASE treatment also enhanced
the expression of low-density lipoprotein receptor (Ldlr), which
promotes the uptake and clearance of LDL cholesterol in plasma
(Marrelli et al., 2016). Moreover, saponins also have anti-
microbial and antioxidant properties, and exhibit cancer-related
immunomodulatory eects (Avato et al., 2006).
The Roles of Secondary Metabolites in
Combating Abiotic and Biotic Interactions
Polyphenols
Plant roots communicate actively with the soil microbes for
mutualistic cycles. Flavonoids are important signaling molecules
for the legume-microbe interactions. e ability to form nitrogen
xing nodules with rhizobia is a unique characteristic of legumes
(Hirsch et al., 2001). Such mutualism between legume and
rhizobium is initiated by avonoids. Flavonoids released from
roots attract rhizobia to migrate toward the roots and stimulate
the nod genes, which are essential genes to synthesize Nod
factors for infecting the plants (Spaink, 1995). Flavonoids in
the root exudates of various legumes for attracting rhizobia
have been summarized in a previous review (Haldar and
Sengupta, 2015). Moreover, avonoids stimulate the germination
of mycorrhizal fungus spores and enhance hyphal growth
(Abdel-lateif etal., 2012). Mycorrhizal fungi form hyphae which
penetrate plant roots for the transport of nutrients in rhizosphere
to the host plant (Harrison, 2005).
e importance of polyphenols to combating abiotic stress
has been discussed in recent reviews (Di Ferdinando etal., 2014;
Isah, 2019; Sharma etal., 2019). e antioxidating characteristics
of polyphenols help alleviate the oxidative stress brought forth
by abiotic stress (Di Ferdinando et al., 2014; Isah, 2019; Sharma
et al., 2019). A recent method for screening legume crops for
abiotic stress tolerance suggested the accumulation of anthocyanin,
which is also an osmolyte, as one of the indicators of abiotic
stress tolerance of legume crops (Sinha et al., 2020).
Strigolactones
Based on the molecular structure, strigolactones belong to a group
of lactone, which is derived from carotenoid (Jia et al., 2018).
Functionally, strigolactones are plant hormones that are released
by roots to attract symbiotic arbuscular mycorrhizal fungi and
induce the germination of parasitic weed seeds (Jia et al., 2018).
Strigolactones have been identied from a broad range of legumes,
including Arachis hypogaea, Astragalus sinicus, Cicer arietinum,
Glycine max, Lupinus albus, Medicago sativa, Phaseolus vulgaris,
Pisum sativum, Psophocarpus tetragonolobus, Trifolium incarnatum,
Vicia faba, and Vigna angularis (Yoneyama et al., 2008).
A study showed that the expression of several secretory
proteins of Rhizophagus irregularis, an arbuscular mycorrhizal
fungus, was induced by strigolactone treatment (Tsuzuki
et al., 2016). Among these proteins, Strigolactone-Induced
Putative Secreted Protein 1 (SIS1) showed the highest induction
fold by both strigolactone treatment and Medicago truncatula
root symbiosis. SIS1 is important for colonization and the
formation of stunted arbuscules (Tsuzuki etal., 2016). erefore,
the strigolactone-induced is an essential protein for the
symbiosis (Tsuzuki et al., 2016).
Broomrapes, especially Orobanche crenata, are believed
to be the major parasitic weeds of legumes. The effects of
the parasitic weeds on legumes include local damage of the
plants and yield loss (Rubiales and Fernández-Aparicio, 2012).
The germination of Orobanche seeds is induced by
strigolatones (Yoneyama et al., 2008).
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 5 September 2020 | Volume 11 | Article 581357
Alkaloids and Saponins
Alkaloids and saponins are known for contributing to the bitter
taste of plants (Drewnowski and Gomez-Carneros, 2000). e
toxicity of alkaloids has been reported (Wink, 2013). Several
studies report that alkaloids and saponins are related to the
resistance to herbivores. For example, yellow lupin cultivar
with higher level of alkaloids in the leaves is more resistant
to aphid than the cultivar with lower level of alkaloids (Adhikari
et al., 2012). e removal of the bitter taste from modern
lupin cultivars has enabled them to be a protein source in
animal feed to reduce the dependence on soybean (Abraham
et al., 2019). However, “sweet” lupins are more susceptible to
predators (Muzquiz et al., 1994). Saponins have been thought
to be responsible for the resistance to insect attacks, as the
saponin preparation garden pea (Pisum sativum L.) resistant
to Azuki bean beetle (Callosobruchus chinensis L.) inhibited
the development of the beetle (Applebaum et al., 1969).
ALTERATIONS OF SECONDARY
METABOLITE PROFILES IN LEGUMES
DURING DOMESTICATION
Polyphenols and Carotenoids Determine
the Colors of Seeds and Flowers
e seed coat color is mainly determined by polyphenols such
as tannins (Heiser, 1988; Espinosa-Alonso et al., 2006). It is
common for the pigmentation patterns of domesticated crops
to be altered compared to their wild relatives. e loss of
pigment in the seed coat of cultivated P. v u l g a r i s is an obvious
example of the eects of domestication (McClean etal., 2018).
In a survey of 18 Lablab purpureus (L. purpureus) germplasms,
including wild, semi-domesticated, and cultivated accessions,
it was found that all the wild accessions have gray-brown and
mottled seed coat (Maass, 2006). However, cultivated accessions
display a spectrum of seed coat colors, including cream-white,
cream, tan, and black (Maass, 2006). Unlike the wild accessions,
some cultivated accessions do not have mottled seed coats
(Maass, 2006). Among 11 landraces and two cultivated accessions
of peanut (Arachis hypogaea L.), it was found that all the
cultivated accessions have a single seed coat color: tan (Husain
and Mallikarjuna, 2012), while the landraces are either red or
tan (Husain and Mallikarjuna, 2012). Some landraces even
have variegated seed coats (Husain and Mallikarjuna, 2012).
In another study, it was shown that cultivated peanut
(A. hypogaea) could have purple, brown, red, or white seed
coats and some have variegated seed coats (Bertioli et al.,
2011). In a survey of a soybean population consisting of 1,957
domesticated and 1,079 wild accessions, it was found that
almost all wild accessions have purple owers and black seed
coats (Jeong et al., 2019), whereas the domesticated soybean
accessions have more diverse seed coat colors, including colorless
(yellow or green seeds), brown, or black, and more diverse
oral colors, including white or purple (Jeong et al., 2019).
In another study on 110 cultivated, 130 landrace, and 62 wild
soybean accessions, it was reported that all cultivated accessions
have yellow seeds, and landrace accessions have yellow, green,
brown, or black seeds, while all the wild accessions have black
seeds (Wang et al., 2018). Similarly, the modern cultivated
pea cv. Cameor (P. s at i v u m ) has transparent seed coat while
the wild accession (P. s a t iv um subsp. elatius JI64) has pigmented
seed coat (Smýkal et al., 2014). In another study on cultivated
(Lens culinaris ssp. Culinaris) and wild lentils (Lens culinaris
ssp. orientalis, L. culinaris ssp. odemensis, L. culinaris ssp.
tomentosus, Lens nigricans, and Lens ervoides, Lens lamottei),
although wild accessions do not necessarily have darker seed
coats, wild accessions have more complexed patterns on the
seed coats (Singh etal., 2014). e seed coats of the cultivated
accessions have either no or dotted patterns (Singh etal., 2014).
However, many of the wild accessions have marbled pattern
on seed coats (Singh etal., 2014). For chickpea (C. arietinum),
the light color of the cultivated seeds is thought to
be non-existing in wild accessions (Penmetsa et al., 2016).
e seed coat color is related to the defense against herbivore.
It has been suggested that a black seed coat protects the seed
from night-time foragers (Porter, 2013).
Polyphenols also give rise to the colors of owers (Wiesner
et al., 2017). In cowpea, cultivated accessions have a wide
range of oral colors while most of the wild accessions have
only purple owers (Lo etal., 2018). Similarly, cultivated soybean
accessions have purple, white, or other colors of owers
(Sundaramoorthy etal., 2015; Jeong etal., 2019), whereas most
of the wild soybean accessions have only purple owers
(Sundaramoorthy et al., 2015). On the contrary, in common
bean, most of the cultivated accessions have only white owers
while the wild accessions have white, pink, or purple owers
(García etal., 1997). Cultivated lentils (L. culinaris ssp. Culinaris)
have white or purple owers but some wild lentils, L. culinaris
ssp. odemensis, L. culinaris ssp. tomentosus and Lentil ervoides,
have only purple owers (Singh etal., 2014). e white owers
of cultivated chickpea (C. arietinum) is thought to benon-existing
in wild accessions (Penmetsa et al., 2016). For pea, cultivated
peas (P. s a ti vum ) usually have white owers while purple owers
are found in wild peas (Hellens et al., 2010). e contrasting
ower colors contributed to the establishment of the Mendel’s Laws.
The Co-Evolution of Seed and Floral Colors With
Foragers and Pollinators
As discussed above, cultivated legumes usually have lighter seed
coat colors compared to the wild counterparts. During
domestication, light seed coat colors have been preferred by
farmers. e loss of color is associated with the loss of secondary
metabolites, such as tannins (Heiser, 1988). As mentioned before,
a dark seed coat may protect the seeds from night-time foragers
in the wild (Porter, 2013). However, the potential increase in
loss of sown seeds to wild animals may not be signicant as
farmers usually have measures to keep foragers away from crops.
Another example is the loss of bitter compounds, such as alkaloids
and saponins in domesticated legumes. e loss of such compounds
would have enhanced the loss of seeds due to foraging by
animals and is usually not advantageous for the survival of the
crops without the protection provided by farmers. erefore,
the loss of bitter compounds in domesticated legumes is also
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 6 September 2020 | Volume 11 | Article 581357
known as a conscious selection by breeders during domestication.
In soybean, most of the elite cultivated soybean seeds are yellow.
It was found that the stay-green G gene is associated with green
seeds and it controls seed dormancy, but is lost in elite cultivated
soybean seeds (Wang etal., 2018). In the survey of 110 cultivated,
130 landrace, and 62 wild soybean accessions, it was found
that the G genotype is present in only 4% of the cultivated
accessions, 21% of the landraces while it is found in 100% of
the wild accessions (Wang et al., 2018).
It has been suggested that oral color has co-evolved with
pollinators such as birds and bees. Bees tend to be attracted
to yellow owers while birds tend to prefer red owers due
to their dierent visual sensitivities (Toon et al., 2014).
Bee-pollinated plants usually have yellow, white, or blue owers
while bird-pollinated plants usually have red owers (Toon
et al., 2014). e transition from bee-pollination to bird-
pollination of Australian egg-and-bacon pea is related to the
number of bird species in the geographical region where the
plants grow (Toon et al., 2014). e yellow color of the Lotus
ower, together with the orientation, size, petal morphology,
sucrose-dominant nectar composition, and scent of the ower,
was reported as a factor contributing to the transition to
pollination by birds (Cronk and Ojeda, 2008).
Carotenoid Level Is Related to Seed Dispersal
by Animals
Besides polyphenols, carotenoids also play a role in determining
tissue colors. During domestication, the protability of seeds
is a major concern for farmers. erefore, genotypes with
reduced seed dispersal, including through pod shattering and
seed dispersal by animals, were actively selected for by breeders
and farmers. In a study on the seeds of 10 legume genera:
Arachis (peanut), Cicer (chickpea), Glycine (soybean), Lathyrus
(vetch), Lens (lentil), Lupinus (lupin), Phaseolus (bean), Pisum
(pea), Vicia (fava bean), and Vig na (cowpea), drastic changes
in the levels and compositions of carotenoids in seeds were
found in domesticated cultivars compared to their wild
counterparts (Fernández-marín et al., 2014). An average of
48% reduction in carotenoids was found in the seeds of these
10 legumes. Besides, the compositions of carotenoids were
more complex in the wild species of Cicer, Glycine, Lathyrus,
Lens, Lupinus, and Vigna. In the study, neoxanthin, violaxanthin,
lutein epoxide, and antheraxanthin were only found in the
wild species but not the domesticated varieties. It was suggested
that seeds with lower carotenoid levels are less attractive to
seed dispersers (Fernández-marín et al., 2014). In contrast,
attracting seed dispersers has been suggested to bean adaptation
of wild legumes (Brǿnnvik and von Wettberg, 2019). It was
suggested that seed dispersal by birds is an important factor
contributing to the widespread of P. v u l g a r i s from Mexico to
South America (Brǿnnvik and von Wettberg, 2019).
Isoavones Are Unique to Legumes
Isoavones are a sub-class of avonoid uniquely found in
legumes. Soybean is a rich and common source of isoavones
for human consumption (Ku et al., 2020). In a study of seed
isoavone contents using 209 wild, 580 landrace, and 106
cultivated soybean accessions, it was found that landraces had
the highest average level of total seed isoavone, followed by
wild accessions and then cultivated accessions (Wang et al.,
2010). e higher average total seed isoavone content in
landraces compared to cultivated accessions was also reported
in another study using 927 landraces and 241 cultivars (Azam
etal., 2020). For individual isoavone contents, it was suggested
that high genistin and glycitin contents, with low daidzin levels,
were articially selected for. e signicantly lower daidzin
contents lead to the lower average total seed isoavone levels
in cultivated accessions compared to wild accessions (Wang
et al., 2010). ere are debates over the reasons behind the
articial selection of such seed isoavone traits in domesticated
legumes. Regarding seed nutrient content, a negative correlation
between the total isoavone level and the protein level has
been reported in seeds (Primomo et al., 2005; Morrison et al.,
2008; Liang etal., 2010; Smallwood etal., 2014), and a positive
correlation between total seed isoavone level and seed oil
level has also been reported (Morrison et al., 2008; Liang
et al., 2010). However, there has also been a report on the
negative correlation between total seed isoavone level and
seed oil level (Smallwood etal., 2014). e total seed isoavone
has also been correlated to yield (Primomo et al., 2005;
Smallwood et al., 2014; Zhang et al., 2014), as well as the
resistance against pathogens (Carter et al., 2018). When two
soybean cultivars, RCAT1004 and DH4202, which are resistant
and sensitive to cyst nematodes respectively, were grown in
a cyst nematode-infested environment, the resistant cultivar
had a higher seed isoavone level (Carter et al., 2018). A
putative QTL related to cyst nematode susceptibility was found
close to that related to total seed isoavone content (Carter
et al., 2018). During domestication, besides the deliberate
selection for reduced seed isoavone level to reduce the bitterness
of the seed, the isoavone level may also be unintentionally
selected together with other desirable traits, such as nutrient
composition, yield, and resistance to biotic stress.
Alkaloid and Saponin Contents Are
Related to Taste-Focused Breeding
e bitter taste of legume seeds tends to be eliminated during
domestication. For example, domesticated lupin cultivars are less
bitter than the wild relatives, which have signicantly higher levels
of alkaloids in their seeds. Modern lupin cultivars are referred
to as “sweet” lupins. In a survey of 20 sweet lupins and 29 bitter
lupins, the bitter taste of lupins was found to bepositively correlated
to the seed alkaloid content, with lupanine being the main alkaloid
(Muzquiz et al., 1994). Although seed saponin level has been
correlated to the bitterness of seeds in general (Mohan et al.,
2016), it may not be related to the bitterness of lupin seeds. In
a survey of the seed saponin contents in sweet vs. bitter lupins,
the level of saponin was undetectable in the seeds of both sweet
and bitter varieties of L. albus (Shim et al., 2003).
Saponin is also a contributing factor to the bitterness of
seeds (Okubo et al., 1992). It was found that many of the
wild ancestors of Vigna spp. are more resistant than
their cultivated counterparts to Callosobruchus chinensis or
Callosobruchus maculatus (Tomooka etal., 2000). It is possible
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 7 September 2020 | Volume 11 | Article 581357
that the drop in saponin contents when legumes became
domesticated is related to the loss of insect resistance capability
in cultivated species. Several wild chickpea accessions have
higher seed saponin levels than cultivated chickpea accessions
(Kaur etal., 2019). Several wild pigeonpea (Cajanus scarabaeoides)
accessions have higher seed saponin contents than the cultivated
pigeonpea accessions (Cajanus cajan; Sekhon et al., 2017).
However, cultivated pigeonpea accessions do not necessarily
have lower seed saponin contents than the wild accessions
(Sekhon et al., 2017). Seed saponin content is not the sole
factor leading to the insect resistance of legumes.
Besides total seed saponin content, individual saponin
components in legumes are also studied. Saponins can beclassied
into four groups: group A saponins, group B saponins, group E
saponins, and 2,3-dihydro-2,5-dihydroxy-6-methyl-4H-pyran-4-one
(DDMP) saponins (Sawai and Saito, 2011; Krishnamurthy
etal., 2013). e aglycone form of group A saponins is named
soyasapogenol A, while that of DDMP saponins is named
soyasapogenol B. e basic structure of soyasapogenol A and
B is β-amyrin. Soyasapogenol A is a β-amyrin with a hydroxyl
group at C-21, C-22, and C-24, while soyasapogenol B has a
hydroxyl group at C-22 and C-24 only (Sawai and Saito, 2011).
DDMP saponins are relatively unstable and are oen degraded
into group B and group E saponins during food processing
(Sundaramoorthy et al., 2019). Among the various groups
of saponins, group A saponins, which have an acetylated
oligosaccharide chain attached to C-22 of soyasapogenol A, are
thought to be mostly responsible for the undesirable taste of
soybean seeds (Shiraiwa et al., 1991).
In a survey of saponin compositions among 800 cultivated
soybean accessions and 329 wild soybean accessions, it was
found that the saponin type Aa was predominant in cultivated
soybean accessions, while the saponin type AaBc was predominant
in wild soybean accessions (Tsukamoto etal., 1993). In another
survey of the total seed saponin levels in 17 wild and one
cultivated legumes, it was found that the total saponin level
was highest in Glycine soja (G. soja; wild soybean; Shim et al.,
2003). In a study of seed saponin composition of 3,025 G. soja
accessions, diverse compositions of seed saponins were found
among the accessions (Krishnamurthy et al., 2013). Moreover,
naturally occurring wild soybean mutants that lack group A
saponins were found (Krishnamurthy et al., 2013; Takahashi
etal., 2016; Rehman etal., 2018). Wild legumes do not necessarily
have higher seed saponin contents. Instead, the diverse genetic
backgrounds among wild legumes allow the discovery of novel
allelic forms for desirable seed saponin compositions.
Polyphenols and Strigolactones Are
Related to Biotic Interactions
Flavonoids are signaling molecules for legume-microbe interaction
(Abdel-lateif et al., 2012). In a test of nodulating capability of
Rhizobium japonicum (R. japonicum), it was found that all the
strains of R. japonicum in the test could nodulate cowpea,
sirato, and wild soybean (Heront and Pueppket, 1984). However,
nine out of the 11 strains could not form infection threads
with two of the three commercial soybean cultivars in the test
(Heront and Pueppket, 1984). In another study, aer inoculating
36 G. soja (wild soybean) accessions with R. japonicum, 20
formed normal nodules while 16 could not form nodules or
formed abnormal nodule-like structures (Pueppke et al., 1998).
It was hypothesized that the dierent nodulating phenotypes
were due to the dierent avonoid proles in the root exudates
(Pueppke et al., 1998). However, the avonoid proles of root
exudates are similar between the nodulating group and the
non-nodulating group (Pueppke et al., 1998). e avonoid
proles of root exudates were also compared between wild
soybean accessions and the cultivated soybean Peking (Pueppke
et al., 1998). Although many of the wild soybeans showed a
more complexed root exudate prole, a strong correlation
between the dierent root exudates and the nodulating phenotypes
was not found (Pueppke etal., 1998). e eects of domestication
on the avonoid proles in legume root exudates remain unclear.
On the other hand, the root polyphenol compositions of wild
lentil (Lens ervoides) and cultivated lentil (Lens culinaris) were
compared aer the infection of Aphanomyces euteiches, which
is a legume pathogen (Bazghaleh et al., 2018). e wild lentil
was more tolerant to A. euteiches than the cultivated lentil
pathogen (Bazghaleh et al., 2018). e wild lentil generally
had higher levels of polyphenols compared to the cultivated
lentil (Bazghaleh et al., 2018). Although the amount of legume
species and accessions is not enough to conclude the eect of
domestication on the root polyphenol compositions aer pathogen
infection, genotypic dierence exists between wild and cultivated
legumes and is associated with polyphenol accumulation in
roots under biotic stress.
Strigolactones are stimulants of seed germination (Brun
et al., 2018). e yield of faba bean (V. fab a ) is also limited
by parasitic weeds. Faba bean (V. f a b a ) germplasms resistant
to parasitic weeds, broomrape (Orobanche and Phelipanche
spp.) were found (Fernández-Aparicio etal., 2014). e resistant
germplasms have low or undetectable levels of strigolactones
in the root exudates at all plant ages (Fernández-Aparicio
etal., 2014). It was suggested that the screening of germplasms
with low strigalactone levels in root exudates is a strategy to
breed for weed resistant germplasms. Like faba bean (V. faba ),
most of the commercial pea (Pisum sativum L.) cultivars are
susceptible to the attack by crenate broompape (O. crenata
Forsk.), which is a parasitic weed of legumes (Pavan et al.,
2016). In a screen of O. crenata resistant pea germplasms, a
landrace pea germplasm was selected. Repeated self-pollination
of the landrace germplasm resulted in the O. crenata resistant
line ROR12 (Pavan etal., 2016), which exhibited several unique
characters: (1) compared to a O. crenata susceptible cultivar,
the root exudates of ROR12, which had a lower strigolactone
level, had a lower capability to stimulate the germination of
O. crenata seeds; (2) in the eld, the number of O. crenata
shoots per host plant of ROR12 was lower; and (3) the
emergence of O. crenata on ROR12 was delayed. It was proposed
that the resistance to O. crenata was related to the reduced
strigolactone level in the root exudates (Pavan et al., 2016).
Strigolactones are also involved in legume-microbe
interaction. e treatment of synthetic strigolactone (GR24)
to pea (Pisum sativum L.) roots enhanced the nodule number
on the roots due to Rhizobium leguminosarum bv. viciae
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 8 September 2020 | Volume 11 | Article 581357
(RLV248) inoculation (Foo and Davies, 2011). Mutant rms1
of pea (Pisum sativum L.) had undetectable levels of orobanchol
and orobanchyl acetate and a low level of fabacyl acetate in
the root exudates (Foo and Davies, 2011). Compared to the
wild type, rms1 mutant had less nodules on the roots aer
being inoculated with R. leguminosarum b v. viciae (RLV248;
Foo and Davies, 2011). Commercial legume germplasms usually
have lower levels of strigolactones in the root exudate
(Fernández-Aparicio et al., 2014; Pavan et al., 2016).
e low levels of strigolactones may result in the reduced
number of nodules on the roots. However, the nodulating
phenotype may not be of consideration during domestication
as the application of nitrogen fertilizer is a common practice
during domestication.
GENES THAT REGULATE THE DIFFERENT
SECONDARY METABOLITE-RELATED
TRAITS
Several methods are currently being employed to identify the
genes and mutations underlying legume domestication
phenotypes (Olsen and Wendel, 2013). In general, secondary
metabolites are present at higher levels in wild progenitors
than in the domesticated counterparts as a result of articial
selection (Nagl etal., 1997; Lindig-Cisneros etal., 2002; Gepts,
2014). e selection of cultivars based on ease of farming and
other commercial attributes may have occurred at the expense
of potentially benecial secondary metabolites. e reduction
in genetic diversity is one of the main impacts of domestication.
However, the genetic richness of wild populations can beused
to improve cultivated legumes. Traditional plant breeding is a
millenary process for the improvement and development of
new crop varieties. According to breeding objectives, new
legume varieties are produced by crossing parents with desired
traits and selecting among segregating progenies those individuals
with both high yield and the target trait. In this way, pest-
resistant varieties have been developed with genetic resistance
to pathogens (Lavaud etal., 2015). Traits related to pigmentation
and defense against pathogens or herbivores are characteristically
domestication-related traits governed by secondary metabolites.
Besides biosynthesis-related genes, transport-related genes are
also important. e roles of transporters, including ATP-binding
cassette (ABC) transporters and multidrug and toxic compound
extrusion (MATE) transporters in secondary metabolite secretion
and accumulation have been summarized in previous reviews
(Yazaki, 2005; Ku et al., 2020). In this section, examples of
genes and loci controlling secondary metabolite biosynthesis
and transport in legumes will be discussed.
Biosynthesis-Related Genes
Pigmentation-Related Traits
Polyphenols are the major determinants of tissue colors,
including the colors of seed coats of legumes, both by their
presence and their quantities (Espinosa-Alonso et al., 2006).
e major polyphenols responsible for seed coat color in
legumes are avonoids, such as anthocyanins, avonol glycosides,
and proanthocyanidins (condensed tannins). Flavonoid quantities
vary according to the seed developmental stages, genotypes,
and species. e biosynthetic pathway leading to the biosynthesis
of avonoids has been elucidated and is conserved among
seed-producing plants. Flavonoids and isoavonoids are derived
from the phenylpropanoid pathway (Dastmalchi and Dhaubhadel,
2014). Many genes in this pathway, including enzymes,
transporters, and regulatory factors, have been characterized.
e rst committed step is the formation of a bicyclic
tetrahydroxy chalcone (naringenin chalcone) catalyzed by a
chalcone synthase (CHS). Legumes produce an additional
trihydroxy chalcone (THC), isoliquiritigenin chalcone
(Dastmalchi and Dhaubhadel, 2014). is THC is the end
product of the coupled activities of CHS and the legume-
specic chalcone reductase (CHR). Compounds such as daidzein,
medicarpin, and glyceollin are derived from isoliquiritigenin.
Flavonoid production follows the conversion of naringenin
chalcone to (2S)-naringenin by chalcone isomerase (CHI).
Flavone 3-hydroxylase (F3H) catalyzes the hydroxylation of
(2S)-naringenin, eryodictyol, and pentahydroxyl avanones to
yield (2R,3R)-dihydrokaempferol, dihydroquercetin, and
dihydromyricetin, respectively (Tanaka et al., 2008). Flavonoid
3'-hydroxylase (F3'H) and avonoid 3',5'-hydroxylase (F3'5'H)
catalyze the hydroxylation of flavanones, flavanols, and
flavones, and determine the structures of flavonoids and
anthocyanins (Tanaka, 2006). Other enzymes in the pathway
include dihydroflavonol 4-reductase (DFR) and anthocyanidin
synthase (ANS). The biosynthesis pathway of flavonoids is
illustrated in Figure 1.
Pigmentation mechanisms have been studied in dierent
legumes. A transcriptomic analysis was performed to identify
the genes associated with seed coat color in peanut (A. hypogaea;
Wan et al., 2016). Lower proanthocyanidin and anthocyanin
contents were detected in a peanut mutant with a brown cracking
seed coat (pscb). Transcriptomic analyses revealed that the structural
genes of the phenylpropanoid biosynthetic pathway were
downregulated in the pscb mutant, while the genes related to
melanin production were upregulated at the late developmental
stages. is expression pattern was consistent with the higher
melanin content in the pscb mutant compared to the wild type.
Dierential expression analyses of RNA-seq data between the
wild type and pscb mutant revealed three candidate genes (c36498_
g1, c40902_g2, and c33560_g1) as being responsible for the seed
coat color trait. C33560_g1 encodes a R2R3-MYB transcription
factor. Its homologs in Arabidopsis and apple are associated with
the regulation of the phenylpropanoid biosynthesis pathway (Rowan
etal., 2009; Vimolmangkang etal., 2013). C36498_g1 and c40902_g2
encode a caeoyl-CoA O-methyltransferase and a kinesin-4-like
protein, respectively. Putative functions of the encoded proteins
were associated with cell wall organization.
Soybeans cultivated for the commercial market are either
completely yellow or have pigmentation restricted to the hilum
(Palmer etal., 2004). Wild soybeans accumulate avonoids and
anthocyanins within the entire epidermal layer of the seed coat,
giving them a black or brown color (Todd and Vodkin, 1993;
Song et al., 2016). Quantitative trait loci (QTL) governing
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 9 September 2020 | Volume 11 | Article 581357
seed coat color in soybean have been identied using genetic
and genomic analyses to elucidate the genetic changes that
resulted in this domestication trait (Todd and Vodkin, 1996;
Tuteja et al., 2004, 2009; Song et al., 2016). e I, R, and T
loci were found to be involved in the avonoid biosynthesis
pathway (Palmer et al., 2004; Yang et al., 2010). e I locus
on chromosome eight inhibits pigmentation of the seed coat.
ere are four alleles (I, ii, ik, and i) at the I locus where I
and ii are the two dominant forms (Song et al., 2016). e
presence of the I allele results in the absence of pigmentation
and a yellow seed coat at maturity. is allele contains an
inverted repeat of the CHS gene cluster. is structure triggers
posttranscriptional gene silencing (PTGS), which inhibits the
expression of CHS gene family members and their functions
in the avonoid biosynthesis pathway (Tuteja et al., 2004).
e ii allele inhibits pigmentation, resulting in a yellow seed
coat with a pigmented hilum (Palmer et al., 2004). Meanwhile,
the recessive ik and i alleles allow pigment production, with
the ik allele restringing pigments to the saddle and hilum
regions of the seed coat (Palmer et al., 2004). e R and T
loci determine the type and accumulation of pigments in the
seed coat (Buzzetl etal., 1987; Todd and Vodkin, 1993). Higher
avonoid and anthocyanin contents of seeds are currently of
great interest due to the antioxidant properties and avors of
these compounds. Recently, the wild soybean reference genome
of G. soja W05 was used to identify additional alleles of the
causal structural gene variation that controls soybean seed coat
pigmentation (Xie et al., 2019). e analysis of a seed coat
color QTL that overlaps with the known I locus showed that
the W05 reference genome possesses the same inverted repeat
of the CHS gene cluster as the domesticated soybean reference
genome, G. max (Williams 82). is indicates that additional
factors also played a role in causing the seed color changes
during domestication. A comparative genomic analysis of W05
against two domesticated soybeans (Wm82 and ZH13) revealed
the generation of a small interfering RNA (siRNA) from a
large structural rearrangement next to the CHS gene cluster
in Wm82 and ZH13. rough experimental validation, a subtilisin
promoter was shown to drive the expression of a chimeric
transcript that reads through a subtilisin gene fragment and
an anti-CHS1 gene region, resulting in PTGS and inhibits the
expression of CHS genes.
Flavonoids also contribute to oral pigmentation (Tanaka,
2006; Tanaka et al., 2008). Domesticated cowpea (Vigna
unguiculata L. Walp) shows phenotypic variation compared to
its wild relatives. Among the domestication traits, a wide range
of oral and seed coat colors can be found in the cultivated
cowpea. e wild variety shows purple owers and dark seed
pigmentation. Purple owers are the results of diacylated
delphinidin-based anthocyanins (Tanaka et al., 2008). A QTL
analysis of the determinants of oral color in cowpea was performed
in a biparental mapping population (wild × cultivated crosses;
FIGURE1 | Schematic representation of the avonoid biosynthetic pathway. Enzymes involved in the pathway are indicated in bold: chalcone synthase (CHS),
chalcone reductase (CHR), avone 3-hydroxylase (F3H), avonoid 3'-hydroxylase (F3'H), avonoid 3',5'-hydroxylase (F3'5'H), dihydroavonol 4-reductase (DFR), and
anthocyanidin synthase (ANS).
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 10 September 2020 | Volume 11 | Article 581357
FIGURE2 | Schematic representation of the pathway leading to the
synthesis of quinolizidine alkaloid compounds. Enzymes involved in the
pathway are indicated in bold: lysine decarboxylase (LDC), copper amine
oxidase (CuAO), (+)-epilupinine O-coumaroyltransferase (ECT), (+)-epilupinine
feruloyltransferase (EFT), ()-lupinine O-coumaroyltransferase (LCT),
()-lupinine feruloyltransferase (LFT), ()-13α-hydroxymultiorine transferase
(HMT), and (+)-13α-hydroxylupanine O-tigloyltransferas (HLT).
Lo et al., 2018). A single major QTL for oral color, CFcol7,
was mapped in a 64-cM region on chromosome Vu07 containing
254 annotated genes, among which a transcription factor,
Vigun07g110700, was identied as a homolog of Arabidopsis
AT4G09820.1 and Medicago truncatula (Mt) TT8, involved in
the regulation of avonoid biosynthesis (Nesi et al., 2000; Li
et al., 2016). In soybean, one QTL for oral pigmentation was
identied on linkage group G (Josie et al., 2007).
Defense-Related Traits
Toxic secondary metabolites in legumes confer resistance against
pathogens and herbivores. However, the accumulation of these
compounds in legume crops is not desirable for human
consumption or as animal feed since they impart a bitter taste
and could present acute toxicity if ingested in sucient quantities
(Daverio etal., 2014). Alkaloids are some of the main secondary
metabolites produced and stored by legumes with characteristic
toxicity (Wink, 2013). Examples are quinolizidine alkaloids (QAs)
produced by the genera Lupinus, Baptisia, ermopsis, Genista,
Cytisus, Echinosophora, and Sophora (Ohmiya et al., 1995). e
breeding of low-alkaloid (sweet) varieties changed the agronomic
roles of lupins from green manure and forage crops to grain
legumes with high protein and ber contents and health benets
(Sweetingham and Kingwell, 2008; Arnoldi et al., 2015). Four
species within the genus Lupinus have been domesticated and
are important legume crops: Lupinus angustifolius (narrow-leafed
lupin or blue lupin), L. albus (white lupin), Lupinus luteus (yellow
lupin), and Lupinus mutabilis (Andean lupin) (Gustafsson and
Gadd, 1965; Reinhard et al., 2006). QAs produced by lupins
include lupanine, angustifoline, lupinine, sparteine, multiorine,
and aphylline (Frick et al., 2017). e use of lupins for food
purposes depends on their QA levels and each species has a
characteristic alkaloid composition. Domestication has reduced
the amount of alkaloids in lupins, but it has also increased their
susceptibility to several aphid species (Philippi etal., 2015). QAs
are derived from the decarboxylation of L-lysine (Lys) by a lysine
decarboxylase (LDC, EC 4.1.1.18) to form cadaverine (Bunsupa
et al., 2012a,b), which is then converted to 5-aminopentanal via
oxidative deamination by a copper amine oxidase (CuAO, EC
1.4.3.22). 5-aminopentanal is spontaneously cyclized to a Δ1
piperideine Schi base formation (Leistner and Spenser, 1973;
Gupta et al., 1979; Golebiewski and Spenser, 1988; Bunsupa
et al., 2012b), which then undergoes further modications (e.g.,
oxygenation, dehydrogenation, hydroxylation, or esterication)
to produce a range of Lys-derived alkaloids, including lupinine,
sparteine, lupanine, and multiorane (Ohmiya etal., 1995; Bunsupa
et al., 2012b; Frick et al., 2017). QAs are stored in the form of
QA esters. In lupins, QA esters are converted from lupinine/
epilupinine and 13α-tigloyloxymultiorine/13α-tigloyloxylupanine
by two types of acetyltransferases (ATs): (+)-epilupinine/
()-lupinine O-coumaroyl/feruloyltransferase (ECT/EFT-LCT/
LFT) and ()-13α-hydroxymultiorine/(+)-13α-hydroxylupanine
O-tigloyltransferase (HMT/HLT; EC 2.3.1.93), respectively (Saito
et al., 1992; Okada et al., 2005; Bunsupa et al., 2012a). e
biosynthesis pathway of QAs is illustrated in Figure2. An HMT/
HLT-type acetyltransferase was isolated and characterized at the
molecular level in L. albus (Okada et al., 2005) while an
acyltransferase-like gene, Lupinus angustifolius acyltransferase
(LaAT ), has been proposed to be involved in the QA biosynthetic
pathway (Bunsupa et al., 2011).
Domestication-related genetic modications resulting in
low-alkaloid phenotypes are generally results of naturally
occurring (spontaneous) mutations (Gustafsson and Gadd,
1965). e domestication of lupins led to the active selection
by farmers/breeders for sweet varieties which were low in
alkaloids. In the late 1920s, the rst low-alkaloid lines were
obtained from wild germplasms of L. luteus and L. angustifolius
(von Sengbusch, 1942). Subsequently, sweet types were also
obtained for L. albus and L. mutabilis in the 1930s (Taylor
etal., 2020). Several recessive low-alkaloid mutations have been
discovered in L. angustifolius: iucundus (iuc), esculentus (es),
depressus (depr), and tantalus (Swiecicki and Swiecicki, 1995;
Kurlovich, 2002; Taylor etal., 2020), among which, the iucundus
locus is the most prevalent allele in cultivars (Taylor et al.,
2020). Molecular mapping eorts have allowed researchers to
map the iucundus locus to a 746-kb region on chromosome
NLL-07 (Nelson et al., 2006, 2010; Hane et al., 2017). e
reference L. angustifolius genome also facilitated the identication
of markers linked to iucundus that are suitable for marker-
assisted selection (MAS). Specically, an allele marker, IucLi,
has been identied for the iucundus locus, and could be used
for MAS in wild × domesticated crosses in lupin breeding
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 11 September 2020 | Volume 11 | Article 581357
programs (Li et al., 2011). Recently, 12 candidate genes for
the alkaloid locus iucundus and the major QTLs associated
with total QA contents were identied using a transcriptomic
approach (Kroc etal., 2019b). e most promising candidate,
RAP2-7, encodes an ethylene-responsive transcription factor
(ERF) that co-segregated with the iucundus locus and is likely
to be involved in the regulation of QA biosynthesis in
L. angustifolius (Kroc et al., 2019a). Other candidate genes
include a 4-hydroxy-tetrahydrodipicolinate synthase (DHDPS)
involved in Lys biosynthesis as well as genes involved in plant
secondary metabolism (Kroc et al., 2019b).
Recessive low-alkaloid mutations in L. albus have also been
identied: pauper, mitis, reductus, exiguus, and nutricius
(Hackbarth, 1957; Troll, 1958; Porsche, 1964). As in the case
of L. angustifolius, one locus, pauper, is the most studied in
L. albus (Rychel and Książkiewicz, 2019). e Kiev
mutant × P27174 recombinant inbred lines (RILs) population
was used for the rst genetic map of L. albus where the pauper
locus was located on linkage group 11 (Phan et al., 2007).
Recently, a high-resolution map was developed to provide a
high-resolution QTL assay of the agronomic traits of L. albus
(Michał et al., 2017). e pauper locus was localized in the
linkage group ALB18 (Michał et al., 2017). e Lup021586
gene was identied in the region and showed 100% nucleotide
identity to La AT , the acyltransferase gene previously identied
in L. angustifolius (Bunsupa et al., 2011). LAGI01_35805, an
L. albus homolog of La AT that is highly similar to L. angustifolius
Lup021586 gene, has been proposed as a molecular marker
for the pauper locus (Rychel and Książkiewicz, 2019). Meanwhile,
four low-alkaloid alleles have been identied in L. luteus,
including dulcis, amoenus, liber, and v (von Sengbusch, 1942;
Gustafsson and Gadd, 1965). However, there is limited
information on the genetic basis for the low-alkaloid trait in
this species. Eorts to improve the genomic resources of L.
luteus are underway. e rst genetic map for L. luteus has
been recently released (Iqbal et al., 2019). A high-quality
reference genome will help to implement MAS and identify
loci responsible for the low-alkaloid content in L. luteus.
On the other hand, phytoalexins are a class of secondary
metabolites with antimicrobial activities that are synthesized de
novo aer biotic and abiotic stresses (Walton, 1997). Phytoalexin
biosynthesis can be induced by pathogens or a type of stress-
mimicking compounds called elicitors (Angelova et al., 2006),
and are produced by a range of crops including those in the
Fabaceae family (Ahuja etal., 2012). Phytoalexins produced by
the family Leguminosae comprise a variety of chemical compounds,
including avonoid phytoalexins derived from the shikimic acid
pathway. In species such as soybean, prenylated pterocarpans,
i.e., glyceollins, are synthesized in response to fungal pathogens
such as Phytophthora sojae and Macrophomina phaseolina (Lygin
et al., 2013). Soybean produces six forms of the isoavonoid
phytoalexin, glyceollin, where glyceollin I, glyceollin II, and
glyceollin III are the predominant isomers (Banks and Dewick,
1983), derived from the addition of a dimethylallyl chain to
(6aS,11aS)-3,9,6a-trihydroxypterocarpan (glycinol) at either C-4
or C-2 by prenyltransferases (PTs). Two isoavonoid PTs have
been identied in soybean: 4-dimethylallyltransferase (G4DT)
and glycinol 2-dimethylallyltransferase (G2DT; Akashi et al.,
2009; Yoneyama et al., 2016). Molecular characterization of PT
genes revealed that G4DT and G2DT are paralogs resulting
from a whole-genome duplication (Yoneyama et al., 2016). A
genome-wide analysis of PT genes in G. max Wm82 identied
77 PT-encoding genes with 11 putative isoavonoid-specic PTs
(Sukumaran etal., 2018). One of the candidate genes, GmPT01
(G2DT-2) was induced by P. s o j a e infection and AgNO3, which
mimics pathogen attack and lies in the QTL linked to P. s o j a e
resistance. It was suggested that GmPT01 is one of the genes
involved in the partial resistance and could be used in breeding
for increased fungal resistance. Other genes related to P. s o j a e
resistance include a CHS gene, GmCHR2A, located near a QTL
linked to P. s o j a e resistance (Sepiol et al., 2017). Additionally,
studies have shown that resistant and susceptible genotypes
dier in their timing of activating glyceollin biosynthesis
(Yoshikawa et al., 1978; Hahn et al., 1985). A rapid activation
of the biosynthetic pathway allows a high level of accumulation
of these low-molecular weight compounds and confers resistance
to pathogens. Soybean genotypes encoding the P. s oj a e resistance
gene, Rps1k, have shown a rapid activation of glyceollin
biosynthesis and higher resistance to the pathogen (Yoshikawa
et al., 1978; Hahn et al., 1985). Recently, a member of the
NAC (NAM/ATAF1/2/CUC2)-family of transcription factor (TF)
genes, GmNAC42-1, was identied using comparative
transcriptomics (Jahan et al., 2020). GmNAC42-1 binds the
promoter of G4DT and plays a role in the accumulation of
glyceollin I. However, additional TFs are expected to participate
in the regulation of glyceollin biosynthesis.
e phytoalexins in pea (P. s at i v um ) are pisatin and maackiain
(Ahuja et al., 2012). Pisatin, a 6α-hydroxyl-pterocarpan
phytoalexin, is the major phytoalexin in pea produced in response
to fungal infections by Nectria haematococca, Botrytis cinerea,
and Mycosphaerella pinodes (Van den Heuvel and Glazener,
1975; Shiraishi et al., 1978; Etebu and Osborn, 2010). Its
biosynthetic pathway has been partially characterized (Paiva
etal., 1994; DiCenzo and Vanetten, 2006; Kaimoyo and VanEtten,
2008; Celoy and VanEtten, 2014). Pisatin and maackiain are
synthetized via two chiral intermediates, ()-7,2'-dihydroxy-
4',5'-methylenedioxyisoavanone [()-sophorol] and ()-7,2'-
dihydroxy-4',5'-methylenedioxyisoavanol [()-DMDI; Preisig
et al., 1989; Akashi et al., 2006; DiCenzo and Vanetten, 2006].
Sophorol reductase (SOR) is responsible for the production of
sophorol, and it can be inactivated by RNA-mediated genetic
interference (RNAi) which inhibits the production of pisatin
in transgenic pea hairy roots (Kaimoyo and VanEtten, 2008).
e pathway diverges for pisatin production aer ()-DMDI
formation. In the last step of the pathway, the methylation of
(+)-6α-hydroxymaackiain (6α-HMK) at the C-3 position by
6α-hydroxymaackiain methyltransferase (HMM2) results in the
formation of (+)-pisatin. ()-DMDI is converted to ()-maackiain
by hydroxisoavanol dehydratase (HILD). e biosynthesis
pathway of (+)-pisatin and ()-maackiain is illustrated in
Figure3. M. pinodes causes ascochyta blight, the most important
foliar disease of eld pea, which responds by accumulating
pisatin (Shiraishi et al., 1978). Eorts have been made to
elucidate the QTLs associated with the disease resistance and
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 12 September 2020 | Volume 11 | Article 581357
to facilitate the introgression of resistance into pea cultivars
(Wroth, 1999; Timmerman-Vaughan etal., 2004; Prioul-Gervais
et al., 2007; Fondevilla et al., 2011). However, only moderate
resistance has been reported with such eorts in pea cultivars
(Kra et al., 1998). Wild relatives present a high phytoalexin
diversity (Lindig-Cisneros et al., 2002). e genetic controls
for the resistance to M. pinodes were studied in a wild P. s a t i v u m
subsp. syriacum accession P665 (Fondevilla et al., 2008). Six
QTLs associated with M. pinodes resistance were detected in
linkage groups II, III, IV, and V (Timmerman-Vaughan et al.,
2004; Prioul-Gervais etal., 2007). Quantitative trait loci MpV.1
and MpII.1 were specic for seedlings under growth chamber
conditions and MpIII.3 and MpIV.1 for adult plant resistance
in eld conditions (Timmerman-Vaughan et al., 2004; Prioul-
Gervais et al., 2007). In contrast, QTLs MpIII.1 and MpIII.2
were detected both for seedling and eld resistance (Timmerman-
Vaughan et al., 2004; Prioul-Gervais et al., 2007). MpIII.2
overlaps with a QTL previously reported to be related to the
resistance against ascochyta blight complex Asc3.1 (Timmerman-
Vaughan etal., 2004; Prioul-Gervais et al., 2007). A resistance-
gene analog (RGA1.1) was identied in the vicinity of this
QTL using P. s a t iv u m populations (Timmerman-Vaughan etal.,
2004; Prioul-Gervais et al., 2007). QTLs associated with partial
resistance to the root rot-causing A. euteiches have also been
identied in pea, and would be useful for improving and
facilitating the existing recurrent selection-based breeding
program (Kra, 1988; Lewis and Gritton, 1992). Pilet-Nayel
et al. (2002) crossed susceptible lines with partially resistant
ones in order to map the QTLs associated with resistance
against A. euteiches. e genetic map revealed seven such
genomic regions and Aph1 located on the linkage group IVb
was the most promising. Other minor QTLs were also identied
in the 13 linkage groups obtained in the genetic mapping
(Pilet-Nayel et al., 2002). Meta-analyses of four RILs of pea
revealed 27 meta-Aphanomyces resistance QTLs, including 11
with high consistency across populations, locations, years, and
isolates (Hamon etal., 2013). Seven highly consistent genomic
regions were identied with the potential for use in MAS for
pea improvement. Resistance QTLs located in these seven
regions were further validated (Lavaud etal., 2015). Backcross-
assisted selection programs were used to generate near-isogenic
lines (NILs) carrying the resistance alleles of individual or
combined resistance QTLs. e eects of two major QTLs,
Ae-Ps4.5, and Ae-Ps7.6, were validated. e NILs carrying the
resistance alleles of these two QTLs showed the highest resistance
to A. euteiches strains. Several minor-eect QTLs were also
validated, including Ae-Ps2.2 and Ae-Ps5.1. Genome-wide analyses
further validated most of these resistance QTLs and detected
additional novel resistance loci (Desgroux etal., 2016). Putative
candidate genes in these loci were related to biotic stress responses.
Transporters
ATP-binding cassette transporters and multidrug and toxic
compound extrusion transporters play important roles in
the secretion and accumulation of secondary metabolites
(Yazaki, 2005; Ku et al., 2020). These transporters are
associated with microbe interaction and nutrient accumulation
of legumes (Sugiyama et al., 2007; Zhang et al., 2010;
Fondevilla et al., 2011; Li et al., 2016).
ABC Transporter
In soybean (G. max), an ABC transporter was reported to
be involved in the root secretion of genistein, which is an
important signaling molecule for mediating the symbiosis with
rhizobia (Sugiyama etal., 2007). In M. truncatula, two half-ABC
transporters, STR and STR2, are essential for arbuscule development
in arbuscular mycorrhizal symbiosis (Zhang et al., 2010). e
expression of the STR and STR2 genes was induced in cortical
FIGURE3 | Schematic representation of the pathway leading to the
synthesis of (+)-pisatin and ()-maackiain. Enzymes involved in the pathway
are indicated in bold: isoavone reductase (IFR), sophorol reductase (SOR),
(+)-6α-hydroxymaackiain 3-O-methyltransferase (HMM2), and
hydroxisoavanol dehydratase (HILD). The steps to convert
()-7,2'-dihydroxy-4',5'-methylenedioxyisoavanol (DMDI) to
(+)-6α-hydroxymaackiain are unknown (dotted arrow).
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 13 September 2020 | Volume 11 | Article 581357
cells containing arbuscules (Zhang etal., 2010). STR and STR2
dimerize to form a transporter, which is located in the peri-
arbuscular membrane and is important for the arbuscule
development and therefore the symbiosis (Zhang etal., 2010).
e str mutant and STR2-silenced transgenic roots exhibited
stunted arbuscules aer inoculating with Glomus versiforme
(Zhang et al., 2010). In pea, using microarray technology, an
ABC transporter was found to have a higher expression in P.
sativum ssp. syriacum accession P665, which is resistant to
Mycosphaerella pinodes, than the sensitive accession Messire
(Fondevilla et al., 2011).
MATE Transporter
Seeds of wild soybeans (G. soja) generally have higher antioxidant
contents than cultivated soybeans (G. max; Li et al., 2016).
Statistical analysis showed the high correlation among the levels
of seed total antioxidants, phenolics, and avonoids (Li et al.,
2016). Using RILs resulted from the cross between the wild
soybean W05 (G. soja) and the cultivated soybean C08 (G. max),
QTLs regulating the contents of antioxidants, phenolics, and
avonoids in soybean seeds were identied, which share a
common genomic region (Li etal., 2016). In the target genomic
region, three genes, GmMATE1, GmMATE2, and GmMATE4,
were predicted to encode MATE transporters (Li et al., 2016).
ese MATE genes are possible candidates for investigating the
basis behind the dierent seed antioxidant contents between
wild soybeans (G. soja) and cultivated soybeans (G. max).
MOLECULAR BREEDING AND
SECONDARY METABOLITE CONTENT
As covered in this review, legumes produce a diverse array
of secondary metabolites including a large subset of compounds
with biopharmaceutical/nutraceutical properties. e production
of these phytochemicals can be increase through crop
improvement using classical breeding to genetic approaches
(Jacob et al., 2016). Legumes with increased health-benecial
secondary metabolites are potential raw materials for producing
pharmaceutical products.
e genetic variability of legume species is fundamental to
identify parental lines to be used in breeding programs and
exploit legume secondary metabolites. Modern targeted breeding
programs use tools, such as quantitative trait loci, marked-assisted
selection, and genomics applications (Collard and Mackill, 2008;
Jacob et al., 2016). DNA-based molecular markers are used to
characterize genomic regions (insertions, deletions, mutations)
controlling a particular trait or gene to dierentiate individuals
for germplasm identication and characterization (Nadeem etal.,
2018). Molecular markers provide breeders with a valuable resource
to accelerate selection programs and mark complex traits, which
are inuenced by environmental factors or not observable at
early stages of plant development. Flavonoids have pharmacological
eects, such as antioxidants for human nutrition or anti-
inammatory eects among others. Also, nutritional value of
legumes can beenhanced by increasing avonoid content though
breeding selection (D’Amelia et al., 2018). In this case, molecular
markers have been used to study genetic variability in legumes
to obtain varieties with high total avonoid content. Genetic
heritability of avonoids is high and germplasms with dierent
avonoid content can lead to the identication of potential markers
to use in breeding (Caseys et al., 2015). Flavonoid content was
determined in 57 peanut accessions to evaluate the association
between molecular markers and avonoid content (Hou et al.,
2017). Four expressed sequence tag-simple sequence repeat
(EST-SSRs) markers were identied related to high avonoid
content in Chinese peanut germplasm. Functions of these markers
were analyzed and related to outer membrane protein porin,
heat-shock transcription factor, and lectins (Hou et al., 2017).
Further studies are required to conrm the functions of these
ESTs in avonoid synthesis in peanuts. In soybean, three novel
alleles were identied associated to avonoid hydroxylase genes,
F3'H and F3'5'H, related to pigmentation traits (Guo and Qiu,
2013). ese molecular markers were identied using a set of
gene-tagged markers based on the sequence variation of GmF3'H
and GmF3'5'H in dierent soybean accessions, including cultivars,
landraces, and wild soybeans (Guo and Qiu, 2013). Domestication
process does not appear to erode diversity since four GmF3'H
alleles were identied among cultivated soybeans, while G. soja
contained only the GmF3'H allele. In the case of GmF3'5'H, 92.2%
of wild soybean contained the GmF3'5'H-a allele, while three
GmF3'5'H alleles occurred among cultivated soybeans (Guo and
Qiu, 2013). In white clover (Trifolium repens), diversity array
technology (DArT) and microsatellite markers were used to discover
marker-trait associations for avonoid accumulation and biomass
(Ballizany et al., 2016). Signicant associations to concentrations
of avonols quercetin, kaempferol, and Quercetin:Kaempferol ratio
were found to markers on linkage group 1–2. Additionally, the
study revealed deleterious alleles in an elite cultivar indicating
that genetic variability from wild germplasm could be used for
white clover improvement (Ballizany et al., 2016).
ENGINEERING SECONDARY
METABOLITE CONTENTS IN LEGUMES
In addition to breeding programs to improve domesticated
varieties and broaden the gene pool of cultivars, secondary
metabolite contents can also bemodied through plant metabolic
engineering (DellaPenna, 2001). e identication of genes
involved in the biosynthesis pathways of diverse secondary
metabolites has helped to drive strategies to optimize the
production of target compounds. Increased production of target
metabolites can beachieved by altering the primary or secondary
metabolism of an organism, for example, through the
overexpression of genes in biosynthetic pathways or by knocking
out gene expressions and hence the enzymatic activities of
competing pathways. In soybean, the manipulation of the (iso)
avonoid pathway and its eect on the resistance to P. s o j a e
has been studied (Cheng et al., 2015; Chen et al., 2018; Zhou
et al., 2018). GmIFR, encoding an isoavone reductase (IFR),
was identied and overexpressed in soybean (Cheng et al.,
2015). IFR catalyzes an intermediate step in the biosynthesis
of glyceollins (Graham etal., 1990) and its constitutive expression
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 14 September 2020 | Volume 11 | Article 581357
in transgenic soybean plants enhances the resistance to P. s o j a e ,
along with higher glyceollin contents. Similar eects on pathogen
resistance were obtained by the overexpression of coenzyme
A ligase (GmPI4L) in transgenic soybean plants (Chen et al.,
2018). Further attempts to elucidate enzymes/genes responsible
for the resistance to pathogens include the overexpression of
a chalcone isomerase, GmCHI1A, in soybean hairy roots, which
enhanced daidzein accumulation and resistance to P. s o j a e
strain P6497R compared to the control (Zhou et al., 2018).
In alfalfa, nutritional value was increased by engineering genistein
glucoside production (Deavours and Dixon, 2005). A transgenic
alfalfa was developed by constitutively expressing an isoavone
synthase, MtIFS1, from M. truncatula. However, in the MtIFS1-
expressing transgenic alfalfa, isoavonoid production and
accumulation was tissue-specic and aected by environmental
factors such as UV-B and the disease-causing pathogen, Phoma
medicaginis (Deavours and Dixon, 2005). RNAi-mediated gene
silencing of isoavone reductase, SOR, and hydroxymaackiain-
3-O-methyltransferase in pea (P. s a t i v u m ) allowed the
identication of DMDI, an intermediary in the production of
pisatin and maackiain (Kaimoyo and VanEtten, 2008). Other
viable strategies of engineering secondary metabolite pathways
include biosynthesis in microorganisms and modulation of
gene expressions through manipulating the expressions of
transcription factors (Du et al., 2010). Recently, genome-scale
models have been used to represent the metabolic capabilities
of legumes, including alfalfa and soybean (Pfau et al., 2018;
Moreira et al., 2019). is approach allows the integration of
dierent kinds of omics data to get new insights into plant-
microbe interactions (diCenzo et al., 2016; Pfau et al., 2018;
Contador et al., 2020). Models of plant metabolic pathways
could also be used in the design of optimal-use biosynthesis
pathways of secondary metabolites.
CONCLUSION
Domestication generally results in the reduction in secondary
metabolites, which are oen related to the bitter taste of seeds
and the resistance of plants to biotic stresses. e phenomenon
is consistent with the reported decrease in crop biodiversity
due to domestication (Food and Agricultural Organization of
the United Nations, 2010). Having numerous health-benecial
secondary metabolites, legumes have the great potential to
be employed as the sources of bioactive compounds for
pharmaceutical use. On the other hand, besides abiotic stresses,
the changing climate may also bring forth unpredictable biotic
stresses such as insect infestations. From these perspectives,
it is important to retain the biodiversity of legumes in order
to maintain a healthy gene pool to produce new cultivars that
can respond to future changes in their environments.
Understanding the genes that govern the benecial secondary
metabolite compositions in legumes will facilitate the use of
wild legumes in breeding programs or metabolic engineering
to promote crop diversity, as well as to produce legumes with
favorable secondary metabolite proles.
AUTHOR CONTRIBUTIONS
H-ML planned and coordinated the writing. Y-SK put together
the rst complete dra and prepared the table. CC drew
Figures 1, 2. M-SN prepared the table and drew Figure 3.
All authors contributed to the search of literature and writings.
All authors contributed to the article and approved the
submitted version.
FUNDING
is work was supported by grants from the Hong Kong
Research Grants Council General Research Fund (14143916)
and Area of Excellence Scheme (AoE/M-403/16).
ACKNOWLEDGMENTS
J. Chu copy-edited this manuscript.
REFERENCES
Abbas, M., Saeed, F., Anjum, F. M., Afzaal, M., Tufail, T., Bashir, M. S., et al.
(2017). Natural polyphenols: an overview. Int. J. Food Prop. 20, 1689–1699.
doi: 10.1080/10942912.2016.1220393
Abbo, S., van-Oss, P. R., Gopher, A., Saranga, Y., Ofner, I., and Peleg, Z.
(2014). Plant domestication versus crop evolution: a conceptual framework
for cereals and grain legumes. Trends Plant Sci. 19, 351–360. doi: 10.1016/j.
tplants.2013.12.002
Abdel-lateif, K., Bogusz, D., and Hocher, V. (2012). e role of avonoids in
the establishment of plant roots endosymbioses with arbuscular mycorrhiza
fungi, rhizobia and Frankia bacteria. Plant Signal. Behav. 7, 636–641. doi:
10.4161/psb.20039
Abraham, E. M., Ganopoulos, I., Madesis, P., Mavromatis, A., Mylona, P.,
Nianiou-Obeidat, I., et al. (2019). e use of lupin as a source of protein
in animal feeding: genomic tools and breeding approaches. Int. J. Mol. Sci.
20:851. doi: 10.3390/ijms20040851
Acquaviva, R., Russo, A., Galvano, F., Galvano, G., Barcellona, M. L., Volti, G. L.,
et al. (2003). Cyanidin and cyanidin 3-O-β-D-glucoside as DNA cleavage
protectors and antioxidants. Cell Biol. Toxicol. 19, 243–252. doi: 10.1023/b:
cbto.0000003974.27349.4e
Adhikari, K. N., Edwards, O. R., Wang, S., Ridsdill-Smith, T. J., and Buirchell, B.
(2012). e role of alkaloids in conferring aphid resistance in yellow
lupin (Lupinus luteus L.). Crop Pasture Sci. 63, 444–451. doi: 10.1071/
CP12189
Ahuja, I., Kissen, R., and Bones, A. M. (2012). Phytoalexins in defense against
pathogens. Trends Plant Sci. 17, 73–90. doi: 10.1016/j.tplants.2011.11.002
Akashi, T., Sasaki, R., Aoki, T., Ayabe, S., and Yazaki, K. (2009). Molecular
cloning and characterization of a cDNA for pterocarpan
4-dimethylallyltransferase catalyzing the key prenylation step in the biosynthesis
of glyceollin, a soybean phytoalexin. Plant Physiol. 149, 683–693. doi: 10.1104/
pp.108.123679
Akashi, T., VanEtten, H. D., Sawada, Y., Wasmann, C. C., Uchiyama, H., and
Ayabe, S. (2006). Catalytic specicity of pea O-methyltransferases suggests
gene duplication for (+)-pisatin biosynthesis. Phytochemistry 67, 2525–2530.
doi: 10.1016/j.phytochem.2006.09.010
Al-Anazi, A. F., Qureshi, V. F., Javaid, K., and Qureshi, S. (2011). Preventive
eects of phytoestrogens against postmenopausal osteoporosis as compared
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 15 September 2020 | Volume 11 | Article 581357
to the available therapeutic choices: an overview. J. Nat. Sci. Bi ol. Me d. 2,
154–163. doi: 10.4103/0976-9668.92322
Amirkia, V., and Heinrich, M. (2014). Alkaloids as drug leads – a predictive
structural and biodiversity-based analysis. Phytochem. Lett. 10, xlviii–liii.
doi: 10.1016/j.phytol.2014.06.015
Angelova, Z., Georgiev, S., and Roos, W. (2006). Elicitation of plants. Biotechnol.
Biotechnol. Equip. 20, 72–83. doi: 10.1080/13102818.2006.10817345
Applebaum, S. W., Marco, S., and Birk, Y. (1969). Saponins as possible factors
of resistance of legume seeds to the attack of insects. J. Agric. Food Chem.
17, 618–622. doi: 10.1021/jf60163a020
Arnoldi, A., Boschin, G., Zanoni, C., and Lammi, C. (2015). e health benets
of sweet lupin seed ours and isolated proteins. J. Funct. Foods 18, 550–563.
doi: 10.1016/j.j.2015.08.012
Arts, I. C. W., van de Putte, B., and Hollman, P. C. H. (2000). Catechin
contents of foods commonly consumed in the Netherlands. 1. Fruits, vegetables,
staple foods, and processed foods. J. Agric. Food Chem. 48, 1746–1751. doi:
10.1021/jf000025h
Avato, P., Bucci, R., Tava, A., Vitali, C., Rosato, A., Bialy, Z., et al. (2006).
Antimicrobial activity of saponins from Medicago sp.: structure-activity
relationship. Phytother. Res. 20, 454–457. doi: 10.1002/ptr.1876
Azam, M., Zhang, S., Abdelghany, A. M., Shaibu, A. S., Feng, Y., Li, Y., et al.
(2020). Seed isoavone proling of 1168 soybean accessions from major
growing ecoregions in China. Food Res. Int. 130:108957. doi: 10.1016/j.
foodres.2019.108957
Bagayoko, M., Buerkert, A., Lung, G., Bationo, A., and Römheld, V. (2000).
Cereal/legume rotation eects on cereal growth in Sudano-Sahelian West
Africa: soil mineral nitrogen, mycorrhizae and nematodes. Plant Soil 218,
103–116. doi: 10.1023/A:1014957605852
Ballizany, W. L., Griths, A. G., Franzmayr, B. K., Jahufer, M. Z. Z., Hofmann, R. W.,
and Barrett, B. A. (2016). “Marker-trait associations for avonoids and
biomass in white clover (Trifolium repens L.)” in Breeding in a world of
scarcity. eds. I. Roldán-Ruiz, J. Baert and D. Reheul (Cham: Springer
International Publishing), 225–229.
Banks, S. W., and Dewick, P. M. (1983). Biosysnthesis of glyceollins I, II and
III in soybean. Phytochemistry 22, 2729–2733. doi: 10.1016/S0031-9422(00)97682-9
Bazghaleh, N., Prashar, P., Purves, R. W., and Vandenberg, A. (2018). Polyphenolic
composition of lentil roots in response to infection by Aphanomyces euteiches.
Front. Plant Sci. 9:1131. doi: 10.3389/fpls.2018.01131
Bernstein, P. S., Li, B., Vachali, P. P., Gorusupudi, A., Shyma, R., Henriksen, B. S.,
et al. (2016). Lutein, zeaxanthin, and meso-zeaxanthin: the basic and clinical
science underlying carotenoid-based nutritional interventions against ocular
disease. Prog. Retin. Eye Res. 50, 34–66. doi: 10.1016/j.preteyeres.2015.10.003
Bertioli, D. J., Seijo, G., Freitas, F. O., Valls, J. F. M., Leal-Bertioli, S. C. M.,
and Moretzsohn, M. C. (2011). An overview of peanut and its wild relatives.
Plant Genet. Resour. 9, 134–149. doi: 10.1017/S1479262110000444
Böttger, A., Vothknecht, U., Bolle, C., and Wolf, A. (eds.) (2018). “Plant secondary
metabolites and their general function in plants” in Lessons on caffeine,
cannabis & co: Plant derived drugs and their interaction with human receptors
(Switzerland: Springer Nature), 3–17.
Brǿnnvik, H., and von Wettberg, E. J. (2019). Bird dispersal as a pre-adaptation
for domestication in legumes: insights for neo-domestication. Front. Plant
Sci. 10:1293. doi: 10.3389/fpls.2019.01293
Brun, G., Braem, L., oiron, S., Gevaert, K., Goormachtig, S., and Delavault, P.
(2018). Seed germination in parasitic plants what insights can we expect from
strigolactone research? J. Exp. Bo t. 69, 2265–2280. doi: 10.1093/jxb/erx472
Bunsupa, S., Katayama, K., Ikeura, E., Oikawa, A., Toyooka, K., Saito, K., et al.
(2012a). Lysine decarboxylase catalyzes the rst step of quinolizidine alkaloid
biosynthesis and coevolved with alkaloid production in leguminosae. Plant
Cell 24, 1202–1216. doi: 10.1105/tpc.112.095885
Bunsupa, S., Okada, T., Saito, K., and Yamazaki, M. (2011). An acyltransferase-
like gene obtained by dierential gene expression proles of quinolizidine
alkaloid-producing and nonproducing cultivars of Lupinus angustifolius. Plant
Biotechnol. J. 28, 89–94. doi: 10.5511/plantbiotechnology.10.1109b
Bunsupa, S., Yamazaki, M., and Saito, K. (2012b). Quinolizidine alkaloid
biosynthesis: recent advances and future prospects. Front. Plant Sci. 3:239.
doi: 10.3389/fpls.2012.00239
Buzzetl, R. I., Buttery, B. R., and MacTavish, D. C. (1987). Biochemical genetics
of black pigmentation of soybean seed. J. Hered. 78, 53–54. doi: 10.1093/
oxfordjournals.jhered.a110309
Carter, A., Tenuta, A., Rajcan, I., Welacky, T., Woodrow, L., and Eskandari, M.
(2018). Identication of quantitative trait loci for seed isoavone concentration
in soybean (Glycine max) against soybean cyst nematode stress. Plant Breed.
137, 721–729. doi: 10.1111/pbr.12627
Caseys, C., Stritt, C., Glauser, G., Blanchard, T., and Lexer, C. (2015). Eects
of hybridization and evolutionary constraints on secondary metabolites: the
genetic architecture of phenylpropanoids in European Populus species. PLoS
One 10:e0128200. doi: 10.1371/journal.pone.0128200
Celoy, R. M., and VanEtten, H. D. (2014). (+)-Pisatin biosynthesis: from () enantiomeric
intermediates via an achiral 7,2'dihydroxy-4',5'-methyllenedioxyisoav-3-ene.
Phytochemistry 98, 120–127. doi: 10.1016/j.phytochem.2013.10.017
Chen, X., Fang, X., Zhang, Y., Wang, X., Zhang, C., Yan, X., et al. (2018).
Overexpression of a soybean 4-coumaric acid: coenzyme a ligase (GmPI4L)
enhances resistance to Phytophthora sojae in soybean. Funct. Plant Biol. 46,
304–313. doi: 10.1071/FP18111
Cheng, Q., Li, N., Dong, L., Zhang, D., Fan, S., Jiang, L., et al. (2015). Overexpression
of soybean isoavone reductase (GmIFR) enhances resistance to Phytophthora
sojae in soybean. Front. Plant Sci. 6:1024. doi: 10.3389/fpls.2015.01024
Collard, B. C. Y., and Mackill, D. J. (2008). Marker-assisted selection: an approach
for precision plant breeding in the twenty-rst century. Philos . Trans. R .
Soc. Lond . Ser. B Biol. Sci. 363, 557–572. doi: 10.1098/rstb.2007.2170
Considine, M. J., Siddique, K. H. M., and Foyer, C. H. (2017). Nature’ s pulse
power: legumes, food security and climate change. J. Exp. Bot. 68, 1815–1818.
doi: 10.1093/jxb/erx099
Contador, C. A., Lo, S. -K., Chan, S. H. J., and Lam, H. -M. (2020). Metabolic
analyses of nitrogen xation in the soybean microsymbiont Sinorhizobium
fredii using constraint-based modeling. mSystems 5, e00516–e00519. doi:
10.1128/mSystems.00516-19
Cronk, Q., and Ojeda, I. (2008). Bird-pollinated owers in an evolutionary
and molecular context. J. Exp. Bot. 59, 715–727. doi: 10.1093/jxb/ern009
D’Amelia, V., Aversano, R., Chiaiese, P., and Carputo, D. (2018). e antioxidant
properties of plant avonoids: their exploitation by molecular plant breeding.
Phytochem. Rev. 17, 611–625. doi: 10.1007/s11101-018-9568-y
Dastmalchi, M., and Dhaubhadel, S. (2014). “Soybean seed isoavonoids:
biosynthesis and regulation” in Phytochemicals – Biosynthesis, function and
application. ed. R. Jetter (Cham, Switzerland: Springer), 1–21.
Daverio, M., Cavicchiolo, M. E., Grotto, P., Lonati, D., Cananzi, M., and Da
Dalt, L. (2014). Bitter lupine beans ingestion in a child: a disregarded cause
of acute anticholinergic toxicity. Eur. J. Pediatr. 173, 1549–1551. doi: 10.1007/
s00431-013-2088-2
Deavours, B. E., and Dixon, R. A. (2005). Metabolic engineering of isoavonoid
biosynthesis in alfalfa. Plant Physiol. 138, 2245–2259. doi: 10.1104/pp.105.062539
Dehaan, L. R., Van Tassel, D. L., Anderson, J. A., Asselin, S. R., Barnes, R.,
Baute, G. J., et al. (2016). A pipeline strategy for grain crop domestication.
Crop Sci. 56, 917–930. doi: 10.2135/cropsci2015.06.0356
DellaPenna, D. (2001). Plant metabolic engineering. Plant Physiol. 125, 160–163.
doi: 10.1104/pp.125.1.160
Desgroux, A., Anthoëne, V. L., Roux-Duparque, M., Rivière, J. -P., Aubert, G.,
Tayeh, N., et al. (2016). Genome-wide association mapping of partial resistance
to Aphanomyces euteiches in pea. BMC Genomics 17:124. doi: 10.1186/
s12864-016-2429-4
Di Ferdinando, M., Brunetti, C., Fini, A., and Tattini, M. (2014). “Flavonoids
as antioxidants in plants under abiotic stresses” in Abiotic stress responses
in plants: Metabolism, productivity and sustainability. eds. P. Ahmad and
M. N. V. Prasad (New York, NY: Springer), 159–179.
Dias, T. R., Alves, M. G., Casal, S., Silva, B. M., and Oliveira, P. F. (2016).
e single and synergistic eects of the major tea components caeine,
epigallocatechin-3-gallate and L-theanine on rat sperm viability. Food Funct.
7, 1301–1305. doi: 10.1039/c5fo01611h
diCenzo, G. C., Checcucci, A., Bazzicalupo, M., Mengoni, A., Viti, C., Dziewit, L.,
et al. (2016). Metabolic modelling reveals the specialization of secondary
replicons for niche adaptation in Sinorhizobium meliloti. Nat. Commun.
7:12219. doi: 10.1038/ncomms12219
DiCenzo, G. L., and Vanetten, H. D. (2006). Studies on the late steps of (+)
pisatin biosynthesis: evidence for () enantiomeric intermediates.
Phytochemistry 67, 675–683. doi: 10.1016/j.phytochem.2005.12.027
Dixon, R. A., and Sumner, L. W. (2003). Legume natural products: understanding
and manipulating complex pathways for human and animal health. Plant
Physiol. 131, 878–885. doi: 10.1104/pp.102.017319
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 16 September 2020 | Volume 11 | Article 581357
Drewnowski, A., and Gomez-Carneros, C. (2000). Bitter taste, phytonutrients,
and the consumer: a review. Am. J. Clin. Nutr. 72, 1424–1435. doi: 10.1093/
ajcn/72.6.1424
Du, H., Huang, Y., and Tang, Y. (2010). Genetic and metabolic engineering
of isoavonoid biosynthesis. Appl. Microbiol. Biotechnol. 86, 1293–1312.
doi: 10.1007/s00253-010-2512-8
Espinosa-Alonso, L. G., Lygin, A., Widholm, J. M., Valverde, M. E., and
Paredes-Lopez, O. (2006). Polyphenols in wild and weedy Mexican common
beans (Phaseolus vulgaris L.). J. Agric. Food Chem. 54, 4436–4444.
doi: 10.1021/jf060185e
Etebu, E., and Osborn, A. M. (2010). Molecular quantication of the pea
footrot disease pathogen (Nectria haematococca) in agricultural soils.
Phytoparasitica 38, 447–454. doi: 10.1007/s12600-010-0122-8
Fernández-Aparicio, M., Kisugi, T., Xie, X., Rubiales, D., and Yoneyama, K.
(2014). Low strigolactone root exudation: a novel mechanism of broomrape
(Orobanche and Phelipanche spp.) resistance available for faba bean breeding.
J. Agric. Food Chem. 62, 7063–7071. doi: 10.1021/jf5027235
Fernández-marín, B., Milla, R., Martín-robles, N., Arc, E., Kranner, I., Becerril, J. M.,
et al. (2014). Side-eects of domestication: cultivated legume seeds contain
similar tocopherols and fatty acids but less carotenoids than their wild
counterparts. BMC Plant Biol. 14:1599. doi: 10.1186/s12870-014-0385-1
Fernández-tomé, S., and Hernández-ledesma, B. (2019). Current state of art
aer twenty years of the discovery of bioactive peptide lunasin. Food Res.
Int. 116, 71–78. doi: 10.1016/j.foodres.2018.12.029
Fondevilla, S., Küster, H., Krajinski, F., Cubero, J. I., and Rubiales, D. (2011).
Identication of genes dierentially expressed in a resistant reaction to
Mycosphaerella pinodes in pea using microarray technology. BMC Genomics
12:28. doi: 10.1186/1471-2164-12-28
Fondevilla, S., Satovic, Z., Rubiales, D., Moreno, M. T., and Torres, A. M.
(2008). Mapping of quantitative trait loci for resistance to Mycosphaerella
pinodes in Pisum sativum subsp. syriacum. Mol. Breed. 21, 439–454.
doi: 10.1007/s11032-007-9144-4
Foo, E., and Davies, N. W. (2011). Strigolactones promote nodulation in pea.
Planta 234, 1073–1081. doi: 10.1007/s00425-011-1516-7
Food and Agricultural Organization of the United Nations (2010). Crop
biodiversity: use it or lose it. Available at: http://www.fao.org/news/story/
en/item/46803/icode/ [Accessed July 2, 2020].
Frick, K. M., Kamphuis, L. G., Kadambot, S. H. M., Singh, K. B., and
Foley, R. C. (2017). Quinolizidine alkaloid biosynthesis in lupins and
prospects for grain quality improvement. Front. Plant Sci. 8:87. doi: 10.3389/
fpls.2017.00087
Gao, D., Wang, X., Fu, S., and Zhao, J. (2020). Legume plants enhance the
resistance of soil to ecosystem disturbance site description and experimental.
Front. Plant Sci. 8:1295. doi: 10.3389/fpls.2017.01295
García, E. H., Peña-valdivia, C. B., Aguirre, J. R. R., and Muruaga, J. S. M.
(1997). Morphological and agronomic traits of a wild population and an
improved cultivar of common bean (Phaseolus vulgaris L.). Ann. Bot. 79,
207–213. doi: 10.1006/anbo.1996.0329
Gepts, P. (2014). “Domestication of plants” in Encyclopedia of agriculture and
food systems. ed. N. K. Van Alfen (Cambridge, Massachusetts: Academic
Press), 474–486.
Gepts, P., Beavis, W. D., Brummer, E. C., Shoemaker, R. C., Stalker, H. T.,
Weeden, N. F., et al. (2005). Legumes as a model plant family.
Genomics for food and feed report of the cross-legume advances through
genomics conference. Plant Physiol. 137, 1228–1235. doi: 10.1104/
pp.105.060871.1228
Golebiewski, W. M., and Spenser, I. D. (1988). Biosynthesis of the lupine
alkaloids. II. Sparteine and lupanine. Can. J. Chem. 66, 1734–1748.
doi: 10.1139/v88-280
Graham, T. L., Kim, J. E., and Graham, M. Y. (1990). Role of constitutive
isoavone conjugates in the accumulation of glyceollin in soybean infected
with Phytophthora megasperma. Mol. Plant-Microbe Interact. 3, 157–166.
doi: 10.1094/MPMI-3-157
Guo, Y., and Qiu, L. -J. (2013). Allele-specic marker development and selection
eciencies for both avonoid 3'-hydroxylase and avonoid 3',5'-hydroxylase
genes in soybean subgenus soja. eor. Appl. Gen et. 126, 1445–1455.
doi: 10.1007/s00122-013-2063-3
Gupta, Y. P. (1987). Anti-nutritional and toxic factors in food legumes: a review.
Plant Foods Hum. Nutr. 37, 201–228. doi: 10.1007/BF01091786
Gupta, R. N., Horsewood, P., Koo, S. H., and Spenser, I. D. (1979). e
biosynthesis of the Lythraceae alkaloids. I. e lysine-derived fragment. Can .
J. Chem. 57, 1606–1614. doi: 10.1139/v79-260
Gupta, A. P., Pandotra, P., Kushwaha, M., Khan, S., Sharma, R., and Gupta, S.
(2015). “Alkaloids: a source of anticancer agents from nature” in Studies in
natural products chemistry. ed. Atta-ur-Rahman (Amsterdam, Netherlands:
Elsevier B.V.), 341–445.
Gustafsson, A., and Gadd, I. (1965). Mutations and crop improvement. II. e
genus Lupinu s (Leguminosae). Hereditas 53, 15–39. doi: 10.1111/j.16
01-5223.1965.tb01977.x
Hackbarth, J. (1957). Die Gene der Lupinenarten. III Weiße Lupine (Lupinus
albus). Zeitschri für Pflanzenzüchtung 37, 185–191.
Hahn, M. G., Bonho, A., and Grisebach, H. (1985). Quantitative locaglyceollin
I in relation to fungal hyphae in soybean roots infected with Phytophthora
megasperma f. sp. glycinea. Plant Physiol. 77, 591–601. doi: 10.1104/pp.77.3.591
Haldar, S., and Sengupta, S. (2015). Plant-microbe cross-talk in the rhizosphere:
insight and biotechnological potential. Open Microbiol. J. 9, 1–7. doi:
10.2174/1874285801509010001
Hamon, C., Coyne, C. J., MacGee, R. J., Lesné, A., Esnault, R., Mangin, P.,
et al. (2013). QTL meta-analysis provides a comprehensive view of loci
controlling partial resistance to Aphanomyces euteichesin four sources of
resistance in pea. BMC Plant Biol. 13:45. doi: 10.1186/1471-2229-13-45
Hane, J. K., Ming, Y., Kamphuis, L. G., Nelson, M. N., Garg, G., Atkins, C. A.,
et al. (2017). A comprehensive dra genome sequence for lupin (Lupinus
angustifolius), an emerging health food: insights into plant–microbe interactions
and legume evolution. Plant Biotechnol. J. 15, 318–330. doi: 10.1111/pbi.12615
Harrison, M. J. (2005). Signaling in the arbuscular mycorrhizal symbiosis. Annu.
Rev. Microbiol. 59, 19–42. doi: 10.1146/annurev.micro.58.030603.123749
Hassan, S. M., Byrd, J. A., Cartwright, A. L., and Bailey, C. A. (2010). Hemolytic
and antimicrobial activities dier among saponin-rich extracts from guar,
quillaja, yucca, and soybean. Appl. Biochem. Biotechnol. 162, 1008–1017.
doi: 10.1007/s12010-009-8838-y
Heiser, C. B. (1988). Aspects of unconscious selection and the evolution of
domesticated plants. Euphytica 37, 77–81. doi: 10.1007/BF00037227
Hellens, R. P., Moreau, C., Lin-Wang, K., Schwinn, K. E., omson, S. J.,
Fiers, M. W. E. J., et al. (2010). Identication of Mendel’s white ower
character. PLoS One 5:e13230. doi: 10.1371/journal.pone.0013230
Heront, D. S., and Pueppket, S. G. (1984). Mode of infection, nodulation
specicity, and indigenous plasmids of 11 fast-growing Rhizobium japonicum
strains. J. Bacteri ol. 160, 1061–1066. doi: 10.1128/JB.160.3.1061-1066.1984
Hirsch, A. M., Lum, M. R., and Downie, J. A. (2001). What makes the rhizobia-
legume symbiosis so special? Plant Physiol. 127, 1484–1492. doi: 10.1104/
pp.010866.1484
Hollman, P. C. H., and Arts, I. C. W. (2000). Flavonols, avones and avanols – nature,
occurrence and dietary burden. J. Sci. Food Agric. 80, 1081–1093. doi:
10.1002/(SICI)1097-0010(20000515)80:7<1081::AID-JSFA566>3.0.CO;2-G
Hou, M., Mu, G., Zhang, Y., Cui, S., Yang, X., and Liu, L. (2017). Evaluation
of total avonoid content and analysis of related EST-SSR in Chinese peanut
germplasm. Crop Breed. Appl. Biotechnol. 17, 221–227. doi: 10.1590/1984-703
32017v17n3a34
Hsieh, C., Martínez-Villaluenga, C., De Lumen, B. O., and Hernández-Ledesma, B.
(2017). Updating the research on the chemopreventive and therapeutic role
of the peptide lunasin. J. Sci. Food Agric. 98, 2070–2079. doi: 10.1002/jsfa.8719
Husain, F., and Mallikarjuna, N. (2012). Genetic diversity in Bolivian landrace
lines of groundnut (Arachis hypogaea L.). Indian J. Genet. Plant Breed. 72,
384–389.
Iqbal, M. M., Huynh, M., Udall, J. A., Kilian, A., Adhikari, K. N., Berger, J. D.,
et al. (2019). e rst genetic map for yellow lupin enables genetic dissection
of adaptation traits in an orphan grain legume crop. BMC Genet. 20:68.
doi: 10.1186/s12863-019-0767-3
Isah, T. (2019). Stress and defense responses in plant secondary metabolites
production. Biol. R es. 52:39. doi: 10.1186/s40659-019-0246-3
Jacob, C., Carrasco, B., and Schwember, A. R. (2016). Advances in breeding
and biotechnology of legume crops. Plant Cell Tissue Organ Cult. 127,
561–584. doi: 10.1007/s11240-016-1106-2
Jahan, M. A., Harris, B., Lowery, M., Infante, A. M., Percield, R. J., and
Kovinich, N. (2020). Glyceollin transcription factor GmMYB29A2 regulates
soybean resistance to Phytophthora sojae. Plant Physiol. 183, 530–546. doi:
10.1104/pp.19.01293
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 17 September 2020 | Volume 11 | Article 581357
Jeong, H. J., Jeong, J. B., Kin, D. S., and De Lumen, B. O. (2007). Inhibition
of core histone acetylation by the cancer preventive peptide lunasin. J. Agric.
Food Chem. 55, 632–637. doi: 10.1021/jf062405u
Jeong, H. J., Lam, Y., and De Lumen, B. O. (2002). Barley lunasin suppresses
ras-induced colony formation and inhibits core histone acetylation in
mammalian cells. J. Agric. Food Chem. 50, 5903–5908. doi: 10.1021/jf0256945
Jeong, H. J., Lee, J. R., Jeong, J. B., Park, J. H., and De Lumen, B. O. (2009).
e cancer preventive seed peptide lunasin from rye is bioavailable and
bioactive. Nutr. Cancer 61, 680–686. doi: 10.1080/01635580902850082
Jeong, S. C., Moon, J. -K., Park, S. -K., Kim, M. -S., Lee, K., Lee, S. R., et al.
(2019). Genetic diversity patterns and domestication origin of soybean. e or.
Appl. Gene t. 132, 1179–1193. doi: 10.1007/s00122-018-3271-7
Jeong, H. J., Park, J. H., Lam, Y., and De Lumen, B. O. (2003). Characterization
of lunasin isolated from soybean. J. Agric. Food Chem. 51, 7901–7906. doi:
10.1021/jf034460y
Jia, K., Baz, L., and Al-babili, S. (2018). From carotenoids to strigolactones.
J. Exp. Bot. 69, 2189–2204. doi: 10.1093/jxb/erx476
Josie, J., Alcivar, A., Rainho, J., and Kassem, M. A. (2007). Genomic regions
containing QTL for plant height, internodes length, and fower color in
soybean [Glycine max (L.) Merr.]. BIOS 78, 119–126. doi: 10.1893/00
05-3155(2007)78[119:RAGRCQ]2.0.CO;2
Kaimoyo, E., and VanEtten, H. D. (2008). Inactivation of pea genes by RNAi
supports the involvement of two similar O-methyltransferases in the
biosynthesis of (+)-pisatin and of chiral intermediates with a conguration
opposite that found in (+)-pisatin. Phytochemistry 69, 76–87. doi: 10.1016/j.
phytochem.2007.06.013
Kaplan, L. (1981). What is the origin of the common bean? Econ. Bot. 35,
240–254. doi: 10.1007/BF02858692
Kaur, K., Grewal, S. K., Gill, P. S., and Singh, S. (2019). Comparison of cultivated
and wild chickpea genotypes for nutritional quality and antioxidant potential.
J. Food Sci. Technol. 56, 1864–1876. doi: 10.1007/s13197-019-03646-4
Kleemann, R., Verschuren, L., Morrison, M., Zadelaar, S., van Erk, M. J.,
Wielinga, P. Y., et al. (2011). Anti-inammatory, anti-proliferative and anti-
atherosclerotic eects of quercetin in human in vitro and in vivo models.
Atherosclerosis 218, 44–52. doi: 10.1016/j.atherosclerosis.2011.04.023
Knekt, P., Kumpulainen, J., Järvinen, R., Rissanen, H., Heliövaara, M., Reunanen, A.,
et al. (2002). Flavonoid intake and risk of chronic diseases. Am. J. C lin.
Nut r. 76, 560–568. doi: 10.1093/ajcn/76.3.560
Kra, J. M. (1988). Aphanomyces root rot resistance in peas. Phytopathology
78:1545.
Kra, J. M., Dunne, B., Goulden, D., and Armstrong, S. (1998). A search for
resistance in peas to Mycosphaerella pinodes. Plant Dis. 82, 251–253. doi:
10.1094/PDIS.1998.82.2.251
Krishnamurthy, P., Chung, G., and Singh, R. J. (2013). Saponin polymorphism
in the Korean wild soybean (Glycine soja Sieb. And Zucc.). Plant Breed.
132, 121–126. doi: 10.1111/pbr.12016
Křížová, L., Dadáková, K., Kašparovská, J., and Kašparovský, T. (2019). Isoavones.
Molecules 24:1076. doi: 10.3390/molecules24061076
Kroc, M., Czepiel, K., Wilczura, P., Mokrzycka, M., and Swięcicki, W. (2019a).
Development and validation of a gene-targeted dCAPS marker for marker-
assisted selection of low-alkaloid content in seeds of narrow-leafed lupin
(Lupinus angustifolius L.). Genes 10:428. doi: 10.3390/genes10060428
Kroc, M., Koczyk, G., Kamel, K. A., Czepiel, K., Fedorowicz-Strońska, O.,
Paweł, K., et al. (2019b). Transcriptome-derived investigation of biosynthesis
of quinolizidine alkaloids in narrow-leafed lupin (Lupinus angustifolius L.)
highlights candidate genes linked to iucundus locus. Sci. Rep. 9:2231. doi:
10.1038/s41598-018-37701-5
Ku, Y. -S., Ng, M. -S., Cheng, S. -S., Lo, A. W. -Y., Xiao, Z., Shin, T. -S.,
et al. (2020). Understanding the composition, biosynthesis, accumulation
and transport of avonoids in crops for the promotion of crops as healthy
sources of avonoids for human consumption. Nutrients 12:1717. doi: 10.3390/
nu12061717
Kumar, G. P., and Khanum, F. (2012). Neuroprotective potential of phytochemicals.
Pharmacogn. Rev. 6, 81–90. doi: 10.4103/0973-7847.99898
Kumar, S., and Pandey, A. K. (2013). Chemistry and biological activities of
avonoids: an overview. Sci. World J. 2013:162750. doi: 10.1155/2013/162750
Kumar, A., Singh, A., and Ekavali (2015). A review on Alzheimer’s disease
pathophysiology and its management: an update. Pharmacol. Rep. 67, 195–203.
doi: 10.1016/j.pharep.2014.09.004
Kurlovich, B. S. (2002). “Genetics of lupins” in Lupins: Geography, classification,
genetic resources and breeding. ed. B. S. Kurlovich (St Petersburg, Russia;
Pellosniemi, Finland: OY International North Express), 313–350.
Kushi, L. H., Meyer, K. A., and Jacobs, D. R. Jr. (1999). Cereals, legumes,
and chronic disease risk reduction: evidence from epidemiologic studies.
Am. J. Clin. Nut r. 70(3 suppl), 451S–458S. doi: 10.1093/ajcn/70.3.451s
Latha, B. P., Reddy, I. R. M., Vijaya, T., Rao, S. D., Ismail, S. M., and Girish, B. P.
(2011). Eect of saponin rich extract of Achyranthes aspera on high fat
diet fed male wistar rats. J. Pharm. Res. 4, 3190–3193.
Lavaud, C., Lesné, A., Piriou, C., Le Roy, G., Boutet, G., Moussart, A., et al.
(2015). Validation of QTL for resistance to Aphanomyces euteiches in dierent
pea genetic backgrounds using near-isogenic lines. eor. Appl. Genet. 128,
2273–2288. doi: 10.1007/s00122-015-2583-0
Leistner, E., and Spenser, I. D. (1973). Biosynthesis of the piperidine nucleus.
Incorporation of chirally labeled [1-3H]cadaverine. J. Am. Chem. S oc. 95,
4715–4725. doi: 10.1021/ja00795a041
Lewis, M. E., and Gritton, E. T. (1992). Use of one cycle of recurrent selection
per year for increasing resistance to Aphanomyces root rot in peas. J. Am.
Soc. Hortic. Sci. 117, 638–642. doi: 10.21273/JASHS.117.4.638
Li, M. W., Muñoz, N. B., Wong, C. F., Wong, F. L., Wong, K., Wong, J. W. H.,
et al. (2016). QTLs regulating the contents of antioxidants, phenolics, and
avonoids in soybean seeds share a common genomic region. Front. Plant
Sci. 7:854. doi: 10.3389/fpls.2016.00854
Li, X., Yang, H., Buirchell, B., and Yan, G. (2011). Development of a DNA
marker tightly linked to low-alkaloid gene iucundus in narrow-leafed lupin
(Lupinus angustifolius L.) for marker-assisted selection. Crop Pasture Sci.
62, 218–224. doi: 10.1071/CP10352
Liang, H., Yu, Y., Wang, S., Lian, Y., Wang, T., Wei, Y., et al. (2010). QTL
mapping of isoflavone, oil and protein contents in soybean (Glycine
max L. Merr.). Agric. Sci. China 9, 1108–1116. doi: 10.1016/S1671-
2927(09)60197-8
Lindig-Cisneros, R., Dirzo, R., and Espinosa-Garcia, F. J. (2002). Eects of
domestication and agronomic selection on phytoalexin antifungal
defense in Phaseolus beans. Ecol. R es. 17, 315–321. doi: 10.1046/
j.1440-1703.2002.00491.x
Lo, S., Muñoz-amatriaín, M., Boukar, O., He, I., Cisse, N., Guo, Y., et al.
(2018). Identication of QTL controlling domestication-related traits in
cowpea (Vigna unguiculata L. Walp). Sci. Rep. 8:6261. doi: 10.1038/s415
98-018-24349-4
López-cortez, M. D. S., Rosales-martínez, P., Arellano-cárdenas, S., and
Cornejo-Mazón, M. (2016). “Antioxidants properties and eect of processing
methods on bioactive compounds of legumes” in Grain legumes. ed. A. K. Goyal
(Rijeka: IntechOpen).
Lygin, A. V., Zernova, O. V., Hill, C. B., Kholina, N. A., Widholm, J. M.,
Hartman, G. L., et al. (2013). Glyceollin is an important component of
soybean plant defense against Phytophthora sojae and Macrophomina phaseolina.
Phytopathology 103, 984–994. doi: 10.1094/PHYTO-12-12-0328-R
Maass, B. L. (2006). Changes in seed morphology, dormancy and germination
from wild to cultivated hyacinth bean germplasm (Lablab purpureus:
Papilionoideae). G enet. Res our. Crop. Evol. 53, 1127–1135. doi: 10.1007/
s10722-005-2782-7
Man, S., Gao, W., Zhang, Y., Huang, L., and Liu, C. (2010). Chemical study
and medical application of saponins as anti-cancer agents. Fitoterapia 81,
703–714. doi: 10.1016/j.tote.2010.06.004
Marciani, D. J. (2018). Elucidating the mechanisms of action of saponin-
derived adjuvants. Trends Pharmacol. Sci. 39, 573–585. doi: 10.1016/j.
tips.2018.03.005
Marrelli, M., Conforti, F., Araniti, F., and Statti, G. A. (2016). Eects of saponins
on lipid metabolism: a review of potential health benets in the treatment
of obesity. Molecules 21:1404. doi: 10.3390/molecules21101404
McClean, P. E., Bett, K. E., Stonehouse, R., Lee, R., Pieger, S., Moghaddam, S. M.,
et al. (2018). White seed color in common bean (Phaseolus vulgaris) results
from convergent evolution in the P (pigment) gene. New Phytol. 219,
1112–1123. doi: 10.1111/nph.15259
McCrory, M. A., Hamaker, B. R., Lovejoy, J. C., and Eichelsdoerfer, P. E.
(2010). Pulse consumption, satiety, and weight management. Adv. Nutr. 1,
17–30. doi: 10.3945/an.110.1006
Menéndez, A. B., Calzadila, P. I., Sansberro, P. A., Espasandin, F. D., Gazquez, A.,
Bordenave, C. D., et al. (2019). Polyamines and legumes: joint stories of
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 18 September 2020 | Volume 11 | Article 581357
stress, nitrogen xation and environment. Front. Plant Sci. 10:1415. doi:
10.3389/fpls.2019.01415
Michał, K., Nazzicari, N., Yang, H., Nelson, M. N., Renshaw, D., Rychel, S.,
et al. (2017). A high-density consensus linkage map of white lupin highlights
synteny with narrow-leafed lupin and provides markers tagging key agronomic
traits. Sci. Rep. 7:15335. doi: 10.1038/s41598-017-15625-w
Mohan, V. R., Tresina, P. S., and Daodil, E. D. (2016). Antinutritional factors
in legume seeds: Characteristics and determination. 1st Edn. Cambridge,
Massachusetts: Academic Press.
Moreira, T. B., Shaw, R., Luo, X., Ganguly, O., Kim, H. -S., Ferreira Coelho, L. G.,
et al. (2019). A genome-scale metabolic model of soybean (Glycine max)
highlights metabolic uxes in seedlings. Plant Physiol. 180, 1912–1929.
doi: 10.1104/pp.19.00122
Morrison, M. J., Cober, E. R., Saleem, M. F., McLaughlin, N. B., Frégeau-Reid, J.,
Ma, B. L., et al. (2008). Changes in isoavone concentration with 58 years
of genetic improvement of short-season soybean cultivars in Canada. Crop
Sci. 48, 2201–2208. doi: 10.2135/cropsci2008.01.0023
Muñoz-Esparza, N., Latorre-Moratalla, M. L., Comas-Basté, O., Toro-Funes, N.,
Veciana-Nogués, M. T., and Vidal-Carou, C. V. (2019). Polyamines in food.
Fro nt . Nu tr. 6:108. doi: 10.3389/fnut.2019.00108
Muzquiz, M., Cuadrado, C., Ayet, G., De Cuadra, C., Burbano, C., and Osagie, A.
(1994). Variation of alkaloid components of lupin seeds in 49 genotypes
of Lupinus albus L. from dierent countries and locations. J. Agric. Food
Chem. 42, 1447–1450. doi: 10.1021/jf00043a011
Nadeem, M. A., Nawaz, M. A., Shahid, M. Q., Doğan, Y., Comertpay, G.,
Yıldız, M., et al. (2018). DNA molecular markers in plant breeding: current
status and recent advancements in genomic selection and genome editing.
Biotechnol. Biotechnol. Equip. 32, 261–285. doi: 10.1080/13102818.2017.1400401
Nagl, W., Ignacimuthu, S., and Becker, J. (1997). Genetic engineering and
regeneration of Phaseolus and Vig na. State of the art and new attempts. J.
Plant Physiol. 150, 625–644. doi: 10.1016/S0176-1617(97)80277-5
Nelson, M. N., Moolhuijzen, P. M., Boersma, J. G., Chudy, M., Lesniewska, K.,
Bellgard, M., et al. (2010). Aligning a new reference genetic map of Lupinus
angustifolius with the genome sequence of the model legume Lotus japonicus.
DNA Res. 17, 73–83. doi: 10.1093/dnares/dsq001
Nelson, M. N., Phan, H. T. T., Ellwood, S. R., Moolhuijzen, P. M., Hane, J.,
Williams, A., et al. (2006). e rst gene-based map of Lupinus angustifolius
L.-location of domestication genes and conserved synteny with Medicago
truncatula. eor. Appl . Genet. 113, 225–238. doi: 10.1007/s00122-006-0288-0
Nesi, N., Debeaujon, I., Jond, C., Pelletier, G., Caboche, M., and Lepiniec, L.
(2000). e TT8 gene encodes a basic helix-loop-helix domain protein
required for expression of DFR and BAN genes in Arabidopsis siliques.
Plant Cell 12, 1863–1878. doi: 10.1105/tpc.12.10.1863
Oakenfull, D. (1981). Saponins in food-a review. Food Chem. 7, 19–40. doi:
10.1016/0308-8146(81)90019-4
Ohmiya, S., Saito, K., and Murakoshi, I. (1995). “Lupine alkaloids” in e
alkaloids: Chemistry and pharmacology. ed. G. A. Cordell (San Diego, CA:
Academic Press), 1–114.
Ojwang, L. O., Yang, L., Dykes, L., and Awika, J. (2013). Proanthocyanidin prole
of cowpea (Vigna unguiculata) reveals catechin-O-glucoside as the dominant
compound. Food Chem. 139, 35–43. doi: 10.1016/j.foodchem.2013.01.117
Okada, T., Hirai, M. Y., Suzuki, H., Yamazaki, M., and Saito, K. (2005). Molecular
characterization of a novel quinolizidine akaloid O-tigloyltransferase: cDNA
cloning, catalytic activity of recombinant protein and expression analysis
in Lupinus plants. Plant Cell Physiol. 46, 233–244. doi: 10.1093/pcp/pci021
Okubo, K., Iijima, M., Kobayashi, Y., Yoshikoshi, M., Uchida, T., and Kudou, S.
(1992). Components responsible for the undesirable taste of soybean seeds.
Biosci. Biotechnol. Biochem. 56, 99–103. doi: 10.1271/bbb.56.99
Olsen, K. M., and Wendel, J. F. (2013). A bountiful harvest: genomic insights
into crop domestication phenotypes. Annu. Rev. Plant Biol. 64, 47–70. doi:
10.1146/annurev-arplant-050312-120048
Onguéné, P. A., Ntie-kang, F., Lifongo, L. L., Ndom, J. C., Sippl, W., and
Mbaze, L. M. (2013). e potential of anti-malarial compounds derived
from African medicinal plants. Part I: a pharmacological evaluation of
alkaloids and terpenoids. Mal ar. J. 12:449. doi: 10.1186/1475-2875-13-81
Paiva, N. L., Sun, Y., Dixon, R. A., VanEtten, H. D., and Hrazdina, G. (1994).
Molecular cloning of isoavone reductase from pea (Pisum sativum L):
evidence for a 3R-isoavanone intermediate in (+)-pisatin biosynthesis. Arch.
Biochem. Biophys. 312, 501–510. doi: 10.1006/abbi.1994.1338
Palmer, R. G., Pfeier, T. W., Buss, G. R., and Kilen, T. C. (2004). “Qualitative
genetics” in Soybeans: Improvement, production, and uses. eds. R. M. Shibles,
J. E. Harper, R. F. Wilson and R. C. Shoemaker (Madison, WI: American
Society of Agronomy, Crop Science Society of America, Soil Science Society
of America), 137–233.
Pavan, S., Schiavulli, A., Marcotrigiano, A. R., Bardaro, N., Bracuto, V., Ricciardi, F.,
et al. (2016). Characterization of low-strigolactone germplasm in pea (Pisum
sativum L.) resistant to crenate broomrape (Orobanche crenata Forsk.). Mol.
Plant-Microbe Interact. 29, 743–749. doi: 10.1094/MPMI-07-16-0134-R
Penmetsa, R. V., Carrasquilla-Garcia, N., Bergmann, E. M., Vance, L., Castro, B.,
Kassa, M. T., et al. (2016). Multiple post-domestication origins of kabuli
chickpea through allelic variation in a diversication-associated transcription
factor. New Phytol. 211, 1440–1451. doi: 10.1111/nph.14010
Pfau, T., Christian, N., Masakapalli, S. K., Sweetlove, L. J., Poolman, M. G.,
and Ebenhöh, O. (2018). e intertwined metabolism during symbiotic
nitrogen xation elucidated by metabolic modelling. Sci. Rep. 8:12504. doi:
10.1038/s41598-018-30884-x
Phan, H. T. T., Ellwood, S. R., Adhikari, K. N., Nelson, M. N., and Oliver, R. P.
(2007). e rst genetic and comparative map of white lupin (Lupinus albus L.):
identication of QTLs for anthracnose resistance and owering time, and a
locus for alkaloid content. DNA Res. 14, 59–70. doi: 10.1093/dnares/dsm009
Philippi, J., Schliephake, E., Jürgens, E., Jansen, H. U., and Ordon, F. (2015).
Feeding behavior of aphids on narrow-leafed lupin (Lupinus angustifolius)
genotypes varying in the content of quinolizidine alkaloids. Entomol. Exp.
Appl. 156, 37–51. doi: 10.1111/eea.12313
Pietta, P., Minoggio, M., and Bramati, L. (2003). Plant polyphenols: structure,
occurrence and bioactivity. Stud. Nat. Prod. Chem. 28, 257–312. doi: 10.1016/
S1572-5995(03)80143-6
Pilet-Nayel, M. -L., Muehlbauer, F. J., Mcgee, R. J., Kra, J. M., Baranger, A.,
and Coyne, C. J. (2002). Quantitative trait loci for partial resistance to
Aphanomyces root rot in pea. eor. Appl. Gene t. 106, 28–39. doi: 10.1007/
s00122-002-0985-2
Porsche, W. (1964). Untersuchungen über die Vererbung der Alkaloidarmut
und die Variabilität des Restalkaloidgehaltes Bei lupinus albus L. Der Züchter
34, 251–256. doi: 10.1007/BF00705826
Porter, S. S. (2013). Adaptive divergence in seed color camouage in contrasting
soil environments. New Phytol. 197, 1311–1320. doi: 10.1111/nph.12110
Preisig, C. L., Matthews, D. E., and VanEtten, H. D. (1989). Purication
and characterization of S-Adenosyl-L-methionine:6a-hydroxymaackiain
3-O-methyltransferase from Pisum sativum. Plant Physiol. 91, 559–566.
doi: 10.1104/pp.91.2.559
Primomo, V. S., Poysa, V., Ablett, G. R., Jackson, C. -J., Gijzen, M., and Rajcan, I.
(2005). Mapping QTL for individual and total isoavone content in soybean
seeds. Crop Sci. 45, 2454–2464. doi: 10.2135/cropsci2004.0672
Prioul-Gervais, S., Deniot, G., Receveur, E. -M., Frankewitz, A., Fourmann, M.,
Rameau, C., et al. (2007). Candidate genes for quantitative resistance to
Mycosphaerella pinodes in pea (Pisum sativum L.). eor. Appl. G enet. 114,
971–984. doi: 10.1007/s00122-006-0492-y
Pueppke, S. G., Bolaños-Vásquez, M. C., Werner, D., Bec-Ferté, M. -P., Promé, J. -C.,
and Krishnan, H. B. (1998). Release of avonoids by the soybean cultivars
McCall and Peking and their perception as signals by the nitrogen-xing
symbiont Sinorhizobium fredii. Plant Physiol. 117, 599–608. doi: 10.1104/
pp.117.2.599
Quintero-soto, M. F., Saracho-peña, A. G., Chavez-ontiveros, J., and
Garzon-tiznado, J. A. (2018). Phenolic proles and their contribution to
the antioxidant activity of selected chickpea genotypes from Mexico and
ICRISAT collections. Plant Foods Hum. Nutr. 73, 122–129. doi: 10.1007/
s11130-018-0661-6
Reddy, M. R. I., Latha, P. B., Vijaya, T., and Rao, D. S. (2012). e saponin-
rich fraction of a Gymnema sylvestre R. Br. Aqueous leaf extract reduces
cafeteria and high-fat diet-induced obesity. Z. Naturforsch. C Biosci. 67,
39–46. doi: 10.1515/znc-2012-1-206
Rehman, H. M., Nawas, M. A., Shah, Z. H., Yang, S. H., and Chung, G.
(2018). Characterization of naturally occurring wild soybean mutant (sg-5)
lacking astringent saponins using whole genome sequencing approach. Plant
Sci. 267, 148–156. doi: 10.1016/j.plantsci.2017.11.014
Reinhard, H., Rupp, H., Sager, F., Streule, M., and Zoller, O. (2006). Quinolizidine
alkaloids and phomopsins in lupin seeds and lupin containing food.
J. Chromato gr. A 1112, 353–360. doi: 10.1016/j.chroma.2005.11.079
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 19 September 2020 | Volume 11 | Article 581357
Roberts, R. L., Green, J., and Lewis, B. (2009). Lutein and zeaxanthin in eye and
skin health. Clin. Dermatol. 27, 195–201. doi: 10.1016/j.clindermatol.2008.01.011
Romeo, F. V., Fabroni, S., Ballistreri, G., Muccilli, S., Spina, A., and Rapisarda, P.
(2018). Characterization and antimicrobial activity of alkaloid extracts from
seeds of dierent genotypes of Lupinus spp. Sustainability 10:788. doi: 10.3390/
su10030788
Rossi, A., Serraino, I., Dugo, P., Di Paola, R., Mondello, L., Genovese, T., et al.
(2012). Protective eects of anthocyanins from blackberry in a rat model of
acute lung inammation. Free Radic. Res. 37, 891–900. doi: 10.1080/10715
76031000112690
Rowan, D. D., Cao, M., Lin-Wang, K., Cooney, J. M., Jensen, D. J., Austin, P. T.,
et al. (2009). Environmental regulation of leaf colour in red 35S:PAP1 Arabidopsis
thaliana. New Phytol. 182, 102–115. doi: 10.1111/j.1469-8137.2008.02737.x
Rubiales, D., and Fernández-Aparicio, M. (2012). Innovations in parasitic weeds
management in legume crops. A review. Agron . Sustain. Dev. 32, 433–449.
doi: 10.1007/s13593-011-0045-x
Rychel, S., and Książkiewicz, M. (2019). Development of gene-based molecular
markers tagging low alkaloid pauper locus in white lupin (Lupinus albus L.).
J. Appl. Gen et. 60, 269–281. doi: 10.1007/s13353-019-00508-9
Saito, K., Suzuki, H., Takamatsu, S., and Murakoshi, I. (1992). Acyltransferases
for lupin alkaloids in Lupinus hirsutus. Phytochemistry 32, 87–91. doi:
10.1016/0031-9422(92)80112-R
Sauvaire, Y., Baissac, Y., Leconte, O., Petit, P., and Ribes, G. (1996). “Steroid
saponins from fenugreek and some of their biological properties” in Saponins
used in food and agriculture. Advances in experimental medicine and biology.
eds. G. R. Waller and K. Yamasaki (Boston, MA: Springer).
Sawai, S., and Saito, K. (2011). Triterpenoid biosynthesis and engineering in
plants. Front. Plant Sci. 2:25. doi: 10.3389/fpls.2011.00025
Sekhon, J., Grewal, S. K., Singh, I., and Kaur, J. (2017). Evaluation of nutritional
quality and antioxidant potential of pigeonpea genotypes. J. Food Sci. Technol.
54, 3598–3611. doi: 10.1007/s13197-017-2818-y
Sepiol, C. J., Yu, J., and Dhaubhadel, S. (2017). Genome-wide identication of
Chalcone Reductase gene family in soybean: insight into root-specic GmCHRs
and Phytophthora sojae resistance. Front. Plant Sci. 8:2073. doi: 10.3389/
fpls.2017.02073
Sharma, A., Shahzad, B., Rehman, A., Bhardwaj, R., Landi, M., and Zheng, B.
(2019). Response of phenylpropanoid pathway and the role of polyphenols
in plants under abiotic stress. Molecules 24:2452. doi: 10.3390/molecules
24132452
Shi, J., Arunasalam, K., Yeung, D., Kakuda, Y., Mittal, G., and Jiang, Y. (2004).
Saponins from edible legumes: chemistry, processing, and health benets.
J. Med. Fo od 7, 67–78. doi: 10.1089/109662004322984734
Shi, Y., Guo, R., Wang, X., Yuan, D., Zhang, S., Wang, J., et al. (2014). e
regulation of alfalfa saponin extract on key genes involved in hepatic
cholesterol metabolism in hyperlipidemic rats. PLoS One 9:e88282. doi:
10.1371/journal.pone.0088282
Shim, S. I., Jun, W. J., and Kang, B. H. (2003). Evaluation of nutritional and
antinutritional components in Korean wild legumes. Plant Foods Hum. Nutr.
58, 1–11. doi: 10.1023/B:QUAL.0000041166.10069.6f
Shiraishi, T., Oku, H., Tsuji, Y., and Ouchi, S. (1978). Inhibitory eect of
pisatin on infection process of Mycosphaerella pinodes on pea. Ann. Phy topath.
Soc. Japan 44, 641–645.
Shiraiwa, M., Kudo, S., Shimoyamada, M., Harada, K., and Okubo, K. (1991).
Composition and structure of “group a saponin” in soybean seed. Agr ic.
Biol. Chem. 55, 315–322. doi: 10.1080/00021369.1991.10870574
Singh, M., Bisht, I. S., Kumar, S., Dutta, M., Bansal, K. C., Karale, M., et al.
(2014). Global wild annual Lens collection: a potential resource for lentil
genetic base broadening and yield enhancement. PLoS One 9:e107781. doi:
10.1371/journal.pone.0107781
Singh, B., Jatinder, S. P., Kaur, A., and Singh, N. (2017). Phenolic composition
and antioxidant potential of grain legume seeds: a review. Food Res. Int.
101, 1–16. doi: 10.1016/j.foodres.2017.09.026
Sinha, R., Bala, M., Kumar, M., Sharma, T. R., and Singh, A. K. (2020). “Methods
for screening legume crops for abiotic stress tolerance through physiological
and biochemical approaches” in Legume genomics: Methods and protocols,
methods in molecular biology. eds. M. Jain and R. Garg (New York:
Springer US).
Smallwood, C. J., Nyinyi, C. N., Kopsell, D. A., Sams, C. E., West, D. R.,
Chen, P., et al. (2014). Detection and conrmation of quantitative trait loci
for soybean seed isoavones. Crop Sci. 54, 595–606. doi: 10.2135/
cropsci2013.05.0340
Smýkal, P., Vernoud, V., Blair, M. W., Soukup, A., and ompson, R. D. (2014).
e role of the testa during development and in establishment of dormancy
of the legume seed. Front. Plant Sci. 5:351. doi: 10.3389/fpls.2014.00351
Soda, K. (2010). Polyamine intake, dietary pattern, and cardiovascular disease.
Med. Hypotheses 75, 299–301. doi: 10.1016/j.mehy.2010.03.008
Song, J., Liu, Z., Hong, H., Ma, Y., Tian, L., Li, X., et al. (2016). Identication
and validation of loci governing seed coat color by combining association
mapping and bulk segregation analysis in soybean. PLoS One 11:e0159064.
doi: 10.1371/journal.pone.0159064
Spaink, H. P. (1995). e molecular basis of infection and nodulation by
rhizobia: the ins and outs of sympathogenesis. Annu. Rev. Phytop athol. 33,
345–368. doi: 10.1146/annurev.py.33.090195.002021
Stagnari, F., Maggio, A., Galieni, A., and Pisante, M. (2017). Multiple benets
of legumes for agriculture sustainability: an overview. Chem. Bio l. Techno l.
Agri c. 4:2. doi: 10.1186/s40538-016-0085-1
Sugiyama, A., Shitan, N., and Yazaki, K. (2007). Involvement of a soybean
ATP-binding cassette-type transporter in the secretion of genistein, a signal
avonoid in legume-Rhizobium symbiosis. Plant Physiol. 144, 2000–2008.
doi: 10.1104/pp.107.096727
Sukumaran, A., Mcdowell, T., Chen, L., Renaud, J., and Dhaubhadel, S. (2018).
Isoavonoid-specic prenyltransferase gene family in soybean: GmPT01, a
pterocarpan 2-dimethylallyltransferase involved in glyceollin biosynthesis.
Plant J. 96, 966–981. doi: 10.1111/tpj.14083
Sundaramoorthy, J., Palaniswamy, S., Park, G. T., Son, H. R., Tsukamoto, C.,
Lee, J. -D., et al. (2019). Characterization of a new sg-5 variant with reduced
biosynthesis of group a saponins in soybean (Glycine max (L.) Merr.). Mol.
Breed. 39:144. doi: 10.1007/s11032-019-1066-4
Sundaramoorthy, J., Park, G. T., Lee, J. -D., Kim, J. H., Seo, H. S., and Song, J. T.
(2015). Genetic and molecular regulation of ower pigmentation in soybean.
J. Korean Soc. Appl . Biol . Chem. 58, 555–562. doi: 10.1007/s13765-015-0077-z
Sweetingham, M., and Kingwell, R. (2008). “Lupins– reections and future
possibilities.” in Proceedings 12th International Lupin Conference; September
14–18, 2008; Geraldton: International Lupin Association, 514–524.
Swiecicki, W., and Swiecicki, W. K. (1995). Domestication and breeding
improvement of narrow-leafed lupin (L. angustifolius L.). J. Appl . Genet. 36,
155–167.
Takahashi, Y., Li, X. -H., Tsukamoto, C., and Wang, K. -J. (2016). Identication
of a novel variant lacking group a soyasaponin in a Chinese wild soybean
(Glycine soja Sieb. & Zucc.): implications for breeding signicance. Plant
Breed. 135, 607–613. doi: 10.1111/pbr.12403
Tanaka, Y. (2006). Flower colour and cytochromes P450. Phytochem. Rev. 5,
283–291. doi: 10.1007/s11101-006-9003-7
Tanaka, Y., Sasaki, N., and Ohmiya, A. (2008). Biosynthesis of plant pigments:
anthocyanins, betalains and carotenoids. Plant J. 54, 733–749. doi: 10.1111/j.
1365-313X.2008.03447.x
Taylor, J. L., De Angelis, G., and Nelson, M. N. (2020). “How have narrow-
leafed lupin genomic resources enhanced our understanding of lupin
domestication” in e lupin genome. eds. K. B. Singh, L. G. Kamphuis and
M. N. Nelson (Cham, Switzerland: Springer International Publishing),
95–108.
Tee, E. S., Goh, A. H., and Khor, S. C. (1995). Carotenoid composition and
content of legumes, tubers and starchy roots by HPLC. Malays . J. Nutr. 1,
63–74.
Timmerman-Vaughan, G. M., Frew, T. J., Butler, R., Murray, S., Gilpin, M.,
Fallon, K., et al. (2004). Validation of quantitative trait loci for Ascochyta
blight resistance in pea (Pisum sativum L.), using populations from two
crosses. eor. App l. Genet. 109, 1620–1631. doi: 10.1007/s00122-004-1779-5
Todd, J. J., and Vodkin, L. O. (1993). Pigmented soybean (Glycine max) seed
coats accumulate proanthocyanidins during development. Plant Physiol. 102,
663–670. doi: 10.1104/pp.102.2.663
Todd, J. J., and Vodkin, L. O. (1996). Duplications that suppress and deletions
that restore expression from a chalcone synthase multigene family. Plant
Cell 8, 687–699. doi: 10.1105/tpc.8.4.687
Tomooka, N., Kashiwaba, K., Vaughan, D. A., Ishimoto, M., and Egawa, Y.
(2000). e eectiveness of evaluating wild species: searching for sources
of resistance to bruchid beetles in the genus Vigna subgenus Ceratotropis.
Euphytica 115, 27–41. doi: 10.1023/A:1003906715119
Ku et al. Legume Domestication and Secondary Metabolites
Frontiers in Genetics | www.frontiersin.org 20 September 2020 | Volume 11 | Article 581357
Toon, A., Cook, L. G., and Crisp, M. D. (2014). Evolutionary consequences
of shis to bird-pollination in the Australian pea-owered legumes (Mirbelieae
and Bossiaeeae). BMC Evol. Biol. 14:43. doi: 10.1186/1471-2148-14-43
Troll, H. J. (1958). Erbgänge des Alkaloidgehaltes und Beobachtungen über
Heterosiswirkung bei Lupinus albus. Zeitschri für Pflanzenzüchtung 39,
35–46.
Tsukamoto, C., Kikuchi, A., Harada, K., Kitamura, K., and Okubo, K. (1993).
Genetic and chemical polymorphisms of saponins in soybean seed.
Phytochemistry 34, 1351–1356. doi: 10.1016/0031-9422(91)80028-y
Tsuzuki, S., Handa, Y., Takeda, N., and Kawaguchi, M. (2016). Strigolactone-
induced putative secreted protein 1 is required for the establishment of
symbiosis by the arbuscular mycorrhizal fungus Rhizophagus irregularis. Mol.
Plant-Microbe Interact. 29, 277–286. doi: 10.1094/MPMI-10-15-0234-R
Tuteja, J. H., Clough, S. J., Chan, W., and Vodkin, L. O. (2004). Tissue-specic
gene silencing mediated by a naturally occurring chalcone synthase gene
cluster in Glycine max. Plant Cell 16, 819–835. doi: 10.1105/tpc.021352.1
Tuteja, J. H., Zabala, G., Varala, K., Hudson, M., and Vodkin, L. O. (2009).
Endogenous, tissue-specic short interfering RNAs silence the chalcone
synthase gene family in Glycine max seed coats. Plant Cell 21, 3063–3077.
doi: 10.1105/tpc.109.069856
Uzoh, I. M., Igwe, C. A., Okebalama, C. B., and Babalola, O. O. (2019).
Legume-maize rotation eect on maize productivity and soil fertility parameters
under selected agronomic practices in a sandy loam soil. Sci. Rep. 9:8539.
doi: 10.1038/s41598-019-43679-5
Van den Heuvel, J., and Glazener, J. A. (1975). Comparative abilities of fungi
pathogenic and nonpathogenic to bean (Phaseolus vulgaris) to metabolize
phaseollin. Neth. J. Plant Pathol. 81, 125–137. doi: 10.1007/BF01976803
Vimolmangkang, S., Han, Y., Wei, G., and Korban, S. S. (2013). An apple
MYB transcription factor, MdMYB3, is involved in regulation of anthocyanin
biosynthesis and ower development. BMC Plant Biol. 13:176. doi:
10.1186/1471-2229-13-176
von Sengbusch, R. (1942). Sweet lupins and oil lupins. e history of the
origin of some new crop plants. Landwirtsch. Jahrb. 91, 719–880.
Voutilainen, S., Nurmi, T., Mursu, J., and Rissanen, T. H. (2006). Carotenoids
and cardiovascular health. Am. J. Clin. Nutr. 83, 1265–1271. doi: 10.1093/
ajcn/83.6.1265
Walton, J. D. (1997). “Biochemical plant pathology” in Plant biochemistry. eds.
P. M. Dey and J. B. Harborne (London, UK: Academic Press), 487–502.
Wan, L., Li, B., Pandey, M. K., Wu, Y., Lei, Y., Yan, L., et al. (2016). Transcriptome
analysis of a new peanut seed coat mutant for the physiological regulatory
mechanism involved in seed coat cracking and pigmentation. Front. Plant
Sci. 7:1491. doi: 10.3389/fpls.2016.01491
Wang, M., Li, W., Fang, C., Xu, F., Liu, Y., Wang, Z., et al. (2018). Parallel
selection on a dormancy gene during domestication of crops from multiple
families. Nat. Genet. 50, 1435–1441. doi: 10.1038/s41588-018-0229-2
Wang, C., Wang, Y., Shi, Y., Yan, X., He, Y., and Fan, W. (2011). Eects of
alfalfa saponins on the lipid metabolism, antioxidation and immunity of
weaned piglets. Acta Pratacul. Sin. 20, 210–218.
Wang, C., Zhao, T., and Gai, J. (2010). Genetic variability and evolutionary
peculiarity of isoavone content and its components in soybean germplasm
from China. Sci. Agric. Sin. 43, 3919–3929.
Wiesner, M., Hanschen, F. S., Maul, R., Neugart, S., Schreiner, M., and
Baldermann, S. (2017). “Nutritional quality of plants for food and fodder”
in Encyclopedia of applied plant sciences. eds. B. omas and D. J. Murphy
(Cambridge, Massachusetts: Academic Press), 285–291.
Wink, M. (2013). Evolution of secondary metabolites in legumes (Fabaceae).
S. Afr. J. Bo t. 89, 164–175. doi: 10.1016/j.sajb.2013.06.006
Wroth, J. M. (1999). Evidence suggests that Mycosphaerella pinodes infection
of Pisum sativum is inherited as a quantitative trait. Euphytica 107, 193–204.
doi: 10.1023/A:1003688430893
Xie, M., Chung, C. Y., Li, M., Wong, F., Wang, X., Liu, A., et al. (2019). A
reference-grade wild soybean genome. Nat. Commun. 10:1216. doi: 10.1038/
s41467-019-09142-9
Yang, K., Jeong, N., Moon, J. K., Lee, Y. H., Lee, S. H., Kim, H. M., et al. (2010).
Genetic analysis of genes controlling natural variation of seed coat and ower
colors in soybean. J. Hered. 101, 757–768. doi: 10.1093/jhered/esq078
Yazaki, K. (2005). Transporters of secondary metabolites. Curr. Opin. Plant
Biol. 8, 301–307. doi: 10.1016/j.pbi.2005.03.011
Yoneyama, K., Akashi, T., and Aoki, T. (2016). Molecular characterization of
soybean pterocarpan 2-dimethylallyltransferase in glyceollin biosynthesis:
local gene and whole-genome duplications of prenyltransferase genes led
to the structural diversity of soybean prenylated isoavonoids. Plant Cell
Physiol. 57, 2497–2509. doi: 10.1093/pcp/pcw178
Yoneyama, K., Xie, X., Sekimoto, H., Takeuchi, Y., Ogasawara, S., Akiyama, K.,
et al. (2008). Strigolactones, host recognition signals for root parasitic plants
and arbuscular mycorrhizal fungi, from Fabaceae plants. New Phytol. 179,
484–494. doi: 10.1111/j.1469-8137.2008.02462.x
Yoshikawa, M., Yamauchi, K., and Masago, H. (1978). Glyceollin: its role in
restricting fungal growth in resistant soybean hypocotyls infected with
Phytophthora megasperma var. sojae. Physiol. Plant Pathol. 12, 73–82. doi:
10.1016/0048-4059(78)90020-6
Zhang, Q., Blaylock, L. A., and Harrison, M. J. (2010). Two Medicago truncatula
half-ABC transporters are essential for arbuscule development in arbuscular
mycorrhizal symbiosis. Plant Cell 22, 1483–1497. doi: 10.1105/tpc.110.074955
Zhang, J., Ge, Y., Han, F., Li, B., Yan, S., Sun, J., et al. (2014). Isoavone
content of soybean cultivars from maturity group0 to VI grown in northern
and Southern China. J. Am. Oil Che m. So c. 91, 1019–1028. doi: 10.1007/
s11746-014-2440-3
Zhang, Y., Pechan, T., and Chang, S. K. C. (2018). Antioxidant and angiotensin-I
converting enzyme inhibitory activities of phenolic extracts and fractions
derived from three phenolic-rich legume varieties. J. Funct. Foods 42, 289–297.
doi: 10.1016/j.j.2017.12.060
Zhou, Y., Huang, J., Zhang, X., Zhu, L., Wang, X., Guo, N., et al. (2018).
Overexpression of chalcone isomerase (CHI) increases resistance against
Phytophthora sojae in soybean. J. Plant Biol. 61, 309–319. doi: 10.1007/
s12374-018-0017-7
Zhu, H., Wan, J., Wang, Y., Li, B., Xiang, C., He, J., et al. (2014). Medicinal
compounds with antiepileptic/anticonvulsant activities. Epilepsia 55, 3–16.
doi: 10.1111/epi.12463
Disclaimer: Any opinions, ndings, conclusions or recommendations expressed
in this publication do not reect the views of the Government of the Hong
Kong Special Administrative Region or the Innovation and Technology Commission.
Conflict of Interest: e authors declare that the research was conducted in
the absence of any commercial or nancial relationships that could beconstrued
as a potential conict of interest.
Copyright © 2020 Ku, Contador, Ng, Yu, Chung and Lam. is is an open-access
article distributed under the terms of the Creative Commons Attribution License
(CC BY). e use, distribution or reproduction in other forums is permitted, provided
the original author(s) and the copyright owner(s) are credited and that the original
publication in this journal is cited, in accordance with accepted academic practice. No
use, distribution or reproduction is permitted which does not comply with these terms.
... In meta-analyses of prospective observational studies, the consumption of legumes is associated with anti-carcinogenic properties (mainly for colon, prostate, stomach and pancreatic cancer), cardiovascular protection (both blood pressure and blood lipid values), reduction/delay of the ageing process, improvement of immune response, protection against type 2 diabetes, weight management, protection against osteoporosis, protection against gastrointestinal diseases and psychological health benefits [54][55][56][57][58][59][60][61][62][63]. The mechanisms behind these health benefits are multiple: (1) legumes contain complex carbohydrates with a low glycemic index, (2) they contain antioxidant properties [64], (3) the fibre serves as a food source and during the fermentation process of fibre by the intestinal bacteria secrete short-chain fatty acids [65], (3) they contain pectin, a fibre that reduces LDL cholesterol [65], (4) typical nutraceutical properties of legumes are attributed to bioactive substances such as (a) polyphenols (antioxidant, antimicrobial and anti-inflammatory properties [66,67]), (b) alkaloids (anti-carcinogenic activities and possess the ability to improve blood circulation in the brain [66]), (c) phytates (antioxidant [68]), saponins (beneficial influence on cholesterol levels and antimicrobial, antioxidant and anti-carcinogenic properties [66,69,70]), (d) enzymatic amylase inhibitors (inhibition of the digestive enzyme alpha-amylase, which prevents complex carbohydrates from being converted into simple carbohydrates such as glucose [71]), (e) lectins (anti-carcinogenic properties, stimulating the immune system, binding to tumor cell membranes, reducing cell proliferation and induce apoptosis, antimicrobial and insecticidal mechanisms [72,73]), (f) storage proteins and small peptides (exert hormone-like activities) [74]. Protein-rich soy products such as tofu and tempeh can reduce total serum cholesterol and LDL cholesterol but could also be effective in attenuating the effects of type 2 diabetes, blood pressure and cancer-related issues [75,76]. ...
... In meta-analyses of prospective observational studies, the consumption of legumes is associated with anti-carcinogenic properties (mainly for colon, prostate, stomach and pancreatic cancer), cardiovascular protection (both blood pressure and blood lipid values), reduction/delay of the ageing process, improvement of immune response, protection against type 2 diabetes, weight management, protection against osteoporosis, protection against gastrointestinal diseases and psychological health benefits [54][55][56][57][58][59][60][61][62][63]. The mechanisms behind these health benefits are multiple: (1) legumes contain complex carbohydrates with a low glycemic index, (2) they contain antioxidant properties [64], (3) the fibre serves as a food source and during the fermentation process of fibre by the intestinal bacteria secrete short-chain fatty acids [65], (3) they contain pectin, a fibre that reduces LDL cholesterol [65], (4) typical nutraceutical properties of legumes are attributed to bioactive substances such as (a) polyphenols (antioxidant, antimicrobial and anti-inflammatory properties [66,67]), (b) alkaloids (anti-carcinogenic activities and possess the ability to improve blood circulation in the brain [66]), (c) phytates (antioxidant [68]), saponins (beneficial influence on cholesterol levels and antimicrobial, antioxidant and anti-carcinogenic properties [66,69,70]), (d) enzymatic amylase inhibitors (inhibition of the digestive enzyme alpha-amylase, which prevents complex carbohydrates from being converted into simple carbohydrates such as glucose [71]), (e) lectins (anti-carcinogenic properties, stimulating the immune system, binding to tumor cell membranes, reducing cell proliferation and induce apoptosis, antimicrobial and insecticidal mechanisms [72,73]), (f) storage proteins and small peptides (exert hormone-like activities) [74]. Protein-rich soy products such as tofu and tempeh can reduce total serum cholesterol and LDL cholesterol but could also be effective in attenuating the effects of type 2 diabetes, blood pressure and cancer-related issues [75,76]. ...
... In meta-analyses of prospective observational studies, the consumption of legumes is associated with anti-carcinogenic properties (mainly for colon, prostate, stomach and pancreatic cancer), cardiovascular protection (both blood pressure and blood lipid values), reduction/delay of the ageing process, improvement of immune response, protection against type 2 diabetes, weight management, protection against osteoporosis, protection against gastrointestinal diseases and psychological health benefits [54][55][56][57][58][59][60][61][62][63]. The mechanisms behind these health benefits are multiple: (1) legumes contain complex carbohydrates with a low glycemic index, (2) they contain antioxidant properties [64], (3) the fibre serves as a food source and during the fermentation process of fibre by the intestinal bacteria secrete short-chain fatty acids [65], (3) they contain pectin, a fibre that reduces LDL cholesterol [65], (4) typical nutraceutical properties of legumes are attributed to bioactive substances such as (a) polyphenols (antioxidant, antimicrobial and anti-inflammatory properties [66,67]), (b) alkaloids (anti-carcinogenic activities and possess the ability to improve blood circulation in the brain [66]), (c) phytates (antioxidant [68]), saponins (beneficial influence on cholesterol levels and antimicrobial, antioxidant and anti-carcinogenic properties [66,69,70]), (d) enzymatic amylase inhibitors (inhibition of the digestive enzyme alpha-amylase, which prevents complex carbohydrates from being converted into simple carbohydrates such as glucose [71]), (e) lectins (anti-carcinogenic properties, stimulating the immune system, binding to tumor cell membranes, reducing cell proliferation and induce apoptosis, antimicrobial and insecticidal mechanisms [72,73]), (f) storage proteins and small peptides (exert hormone-like activities) [74]. Protein-rich soy products such as tofu and tempeh can reduce total serum cholesterol and LDL cholesterol but could also be effective in attenuating the effects of type 2 diabetes, blood pressure and cancer-related issues [75,76]. ...
Article
Full-text available
Background: The interest in meat alternatives has increased over the years as people embrace more varied food choices because of different reasons. This study aims to analyse the nutritional composition of ready-to-use meat alternatives and compare them with meat (products). Methods: Nutritional composition values were collected in 2022 of all ready-to-use meat alternatives in Belgian supermarkets, as well as their animal-based counterparts. A one-sample t-test was performed to test the nutritional composition of ready-to-use meat alternatives against norm values, while an independent samples t-test was used to make the comparison with meat. Results: Minced meat and pieces/strips/cubes scored favourably on all norm values. Cheeseburgers/schnitzels, nut/seed burgers and sausages contained more than 10 g/100 g total fat. The saturated fat and salt content was lower than the norm value in each category. Legume burgers/falafel contained less than 10 g/100 g protein. Vegetarian/vegan minced meat and bacon contained fewer calories, total and saturated fat, and more fibre compared to their animal-based counterparts. Conclusions: Minced meat and pieces/strips/cubes came out as the most favourable categories regarding nutritional composition norm values. Vegetarian/vegan steak came out the least favourable compared to steak, while vegetarian/vegan minced meat and vegetarian/vegan bacon came out the most favourable compared to their animal-based counterparts.
... The accumulation is spatiotemporally regulated at transcriptional and post-transcriptional levels (Corso et al., 2020;Lepiniec et al., 2006). Flavonoids and lignins have received particular attention for being altered during plant domestication (Alseekh et al., 2021;Ku et al., 2020;Paauw et al., 2019) as the result of human selection. Tannin-rich seeds have often astringent taste and affect food digestibility (as a result of their defensive role to herbivory) and bioavailability. ...
... Three classes of phenylpropanoid-derived compounds are particularly abundant in seeds: flavonoids, lignins, and lignans (Corso et al., 2020;Francoz et al., 2018) contributing also to seed color variation. These metabolites were found to play a role in seed dormancy in several plants such as pomegranate (Qin et al., 2020), Rubus Comparative analysis of wild progenitors, landraces, and modern varieties has shown a reduction in the content of these (reviewed in Ku et al., 2020). The question is if these have been selected directly, such as visual appearance or palatability (Cosson et al., 2022), or because of relation to other seed compounds, namely, storage proteins and sugars (Sm ykal et al., 2018;Sm ykal & Parker, 2023). ...
Article
Full-text available
The mature seed in legumes consists of an embryo and seed coat. In contrast to knowledge about the embryo, we know relatively little about the seed coat. We analyzed the gene expression during seed development using a panel of cultivated and wild pea genotypes. Gene co‐expression analysis identified gene modules related to seed development, dormancy, and domestication. Oxidoreductase genes were found to be important components of developmental and domestication processes. Proteomic and metabolomic analysis revealed that domestication favored proteins involved in photosynthesis and protein metabolism at the expense of seed defense. Seed coats of wild peas were rich in cell wall‐bound metabolites and the protective compounds predominated in their seed coats. Altogether, we have shown that domestication altered pea seed development and modified (mostly reduced) the transcripts along with the protein and metabolite composition of the seed coat, especially the content of the compounds involved in defense. We investigated dynamic profiles of selected identified phenolic and flavonoid metabolites across seed development. These compounds usually deteriorated the palatability and processing of the seeds. Our findings further provide resources to study secondary metabolism and strategies for improving the quality of legume seeds which comprise an important part of the human protein diet.
... Domestication or cultivation efforts in any medicinal crop will favour the production of secondary metabolites due to various factors like optimal growing conditions and proper management techniques, stable environmental factors, adaptive stress etc. (Li et al. 2020;Isah 2019). Under domestication efforts, stress induced in the plant in response to adaptation will increase the secondary metabolism and yield of metabolites (Ku et al. 2020;Kleinwachter et al. 2015). Another advantage of cultivation over wild collection is that the cultivated genotypes exhibit a stable content unlike wild collections in which batch-to-batch variations occur (Bruni and Sacchetti 2009). ...
Article
Full-text available
Gymnema sylvestre (Retz.) R.Br. ex Sm. (Gudmar) is a medicinal crop that is highly valued because of its antidiabetic and diverse therapeutic properties. In this study, 35 accessions of G. sylvestre collected from different regions of Peninsular India (Karnataka, Tamil Nadu, Kerala, Telangana, and Andhra Pradesh) were characterized in terms of morphological traits and the active biocomponent gymnemagenin. Morphological characterization for five qualitative and 11 quantitative traits indicated the existence of significant variability between the accessions. Pearson’s correlation showed that the leaf size and yield exhibited significant negative correlations with the gymnemagenin content. Thus, small-leaved accessions with a low leaf yield exhibited a higher bioactive content and vice versa. Principal component analysis for 11 quantitative traits revealed that fresh leaf yield per plant, dry leaf yield per plant, gymnemagenin percentage, leaf area, leaf area index and chlorophyll content contributed maximum to the variability. Dendrogram based on Tocher method grouped the 35 accessions of G. sylvestre into six major clusters mostly according to the leaf traits and irrespective of the area of collection. The intracluster distance (34.85) was the highest for Cluster 3, which indicated the presence of highly diverse genotypes within it, and the highest intercluster distance was observed between Cluster 2 and Cluster 6 (688.04). Few promising genotypes viz., IIHR-GS-27 and IIHR-GS-9 were identified for highest fresh leaf yield per plant with 1.76 kg and 1.74 kg leaves/plant, respectively, and IIHR-GS-44 for high gymnemagenin content (1.95%) followed by IIHR-GS-37 (1.72%). Promising morphotypes and chemotypes identified in the diverse genepool can be further deployed in crop improvement programmes to promote the cultivation and commercialization of this species.
... Wild chickpea (Cicer arietinum L.) progenitors, for example, are known to exhibit seed-coat pigmentation that mimics the soil color, thus emphasizing the evolutionary role of seed-coat color [13]. Pigmentation loss has also been linked to the level of bitter-tasting compounds [14]. ...
Article
Full-text available
Legume–rhizobia symbiosis is the most important plant–microbe interaction in sustainable agriculture due to its ability to provide much needed N in cropping systems. This interaction is mediated by the mutual recognition of signaling molecules from the two partners, namely legumes and rhizobia. In legumes, these molecules are in the form of flavonoids and anthocyanins, which are responsible for the pigmentation of plant organs, such as seeds, flowers, fruits, and even leaves. Seed-coat pigmentation in legumes is a dominant factor influencing gene expression relating to N2 fixation and may be responsible for the different N2-fixing abilities observed among legume genotypes under field conditions in African soils. Common bean, cowpea, Kersting’s groundnut, and Bambara groundnut landraces with black seed-coat color are reported to release higher concentrations of nod-gene-inducing flavonoids and anthocyanins compared with the Red and Cream landraces. Black seed-coat pigmentation is considered a biomarker for enhanced nodulation and N2 fixation in legumes. Cowpea, Bambara groundnut, and Kersting’s bean with differing seed-coat colors are known to attract different soil rhizobia based on PCR-RFLP analysis of bacterial DNA. Even when seeds of the same legume with diverse seed-coat colors were planted together in one hole, the nodulating bradyrhizobia clustered differently in the PCR-RFLP dendrogram. Kersting’s groundnut, Bambara groundnut, and cowpea with differing seed-coat colors were selectively nodulated by different bradyrhizobial species. The 16S rRNA amplicon sequencing also found significant selective influences of seed-coat pigmentation on microbial community structure in the rhizosphere of five Kersting’s groundnut landraces. Seed-coat color therefore plays a dominant role in the selection of the bacterial partner in the legume–rhizobia symbiosis.
... Such plants are stronger, less prone to diseases and healthier. Research on the effect of cultivations and environmental factors on plants' secondary metabolites has identified that wild plants contain more and/or stronger medicinal compounds than conventionally grown plants or greenhouse-grown plants (Clemensen et al., 2020;Ku et al., 2020;Strzemski et al., 2020;Pant et al., 2021): the evidence is aligned with agroecology's philosophy (as well as many indigenous and traditional cosmovisions, from Pacha Mama to Taoism) that soil health, plant health and human health are connected (Tornaghi et al., 2023). ...
Chapter
Full-text available
Building on 15 years of research in the field of urban agriculture, this chapter discusses key issues that refrain urban agriculture from achieving its full potential, in terms of human and non-human health, and offers a few pointers for informing policy and practice. First, the author offers a brief overview of how the key challenges of western urbanisation (climate change, people’s health and the destruction of nature) intersect with urban agriculture and what promises lay ahead for this rediscovered practice. The second part highlights a number of shortcomings in urban planning policy, including (i) the failure to consider urban agriculture as a food-producing practice, (ii) the invisibility of soil and lack of understanding of the role of living soils and (iii) the failure to consider the role that soil-cares and food-producing specialists (farmers!) can play in advising on healthy public land design and soil management. Part three discusses how agroecological considerations can deepen and expand the ambition of urban agriculture in public space and bring about more-than-human health. The conclusions offer a summary of the challenges, critical issues, and learning points discussed in the previous three sections and highlight their connection to the concept of ‘agroecological urbanism’.
... Wild chickpea (Cicer arietinum L.) progenitors, for example, are known to exhibit seed coat pigmentation that mimicks soil color, thus emphasizing the evolutionary role of seed coat color [13]. Pigmentation loss has however been linked to the level of bitter-tasting compounds, which suggest a potential relationship between seed coat color and metabolite concentration [14]. ...
Preprint
Full-text available
The legume-rhizobia symbiosis is one of the most important plant-microbe interactions in sustainable agriculture due to its ability to provide much needed N to cropping systems. This interaction is mediated by the mutual recognition of signaling molecules from the two partners, namely legumes and rhizobia. With legumes, these molecules are in the form of flavonoids and anthocyanins, which are responsible for the pigmentation of plants parts, such as seeds, flowers, fruits and even leaves. Seed coat pigmentation in legumes is a dominant factor influencing gene expression relating to N2 fixation, and may be responsible for the different N2-fixing abilities observed among legume genotypes under field conditions in African soils. Cowpea, Kersting’s bean and Bambara groundnut landraces with black seed coat color are reported to release higher concentrations of nod-gene-inducing flavonoids and anthocyanins when compared to the Red and Cream landraces, hence the black seed coat pigmentation is considered a biomarker for enhanced nodulation and N2 fixation. Cowpea, Bambara groundnut and Kersting’s bean with differing seed coat colors are known to attract different native soil rhizobia, confirmable by PCR-RFLP analysis of bacterial DNA from root nodules of these legumes. Even when seeds of the same legume with diverse seed coat colors were planted together in one hole, the nodulating bradyrhizobia clustered differently in the PCR-RFLP dendrograms. In one study, Kersting’s groundnut, Bambara groundnut and cowpea with differing seed coat colors were selectively nodulated by different bradyrhizobial species. Multilocus sequence analysis showed that different Bradyrhizobium species nodulated the Kersting's bean based on seed coat color. Phylogenetic analysis also placed the bradyrhizobial isolates in close proximity to different Bradyrhizobium species such as B. vignae 7-2T, B. subterraneum 58 2-1T, B. kavangense 14-3T, B. liaoningense 2281 (USDA 3622)T, B. yuanmingense LMG 21827T, B. huanghuaihaiense CCBAU 23303T, B. pachyrhizi PAC48T, and a reference type strain of B. elkanii according to seed coat color. Using 16S rDNA amplicon sequencing, we also found significant selective influences of seed coat pigmentation on microbial community structure in the rhizosphere of five Kersting’s groundnut landraces. For example, the rhizosphere of Belane Mottled landrace was dominated by Proteobacteria, while Bacteroidetes dominated the rhizospheres of the other landraces. With legumes, seed coat pigmentation therefore plays a dominant role in the selection of the bacterial symbiotic partner.
... This study sheds light on the intriguing impact of irrigation on the secondary metabolite contents of Calotropis flowers across different regions within the same Agro-Climatic Zone. The results of our study aligns with the reported literature that cultivated crops have lower levels of active constituents (Ku et al, 2020;Li et al, 2023). This also suggests that irrigation practices may influence the biosynthesis or accumulation of sterols and alkaloids in Calotropis flowers, highlighting the potential impact of environmental factors on the secondary metabolite composition. ...
Article
Calotropis procera, a species of family Asclepiadaceae is known in traditional system of medicine from centuries due to its property of curing a number of ailments. It is a xerophytic shrub found in tropical regions of Asia and Africa. It’s flower is useful in anorexia, indigestion and intestinal worm infestation, dysentry, abdominal disorders etc. It is used in many Ayurvedic, Unani and Siddha formulations. Hence investigation on the regional and irrigation-induced variations in secondary metabolites of Calotropis procera flowers across diverse agro-climatic zones in Rajasthan, India, was taken up to find best region for their collection. The study shows that ACZ IIIA has the highest petroleum ether extract content and ACZ V displays elevated methanol extract yields. Within ACZ IB, involving Hanumangarh and Suratgarh, contrasting results in petroleum ether extract and methanol extract yields are observed, with petroleum ether extract higher in the irrigated area and methanol higher in the non-irrigated region. ACZ V demonstrates the maximum total extractive yield, while ACZ IB shows minimum values. Flowers from the irrigated area in Hanumangarh (ACZ IB) exhibit lower values for sterols and alkaloids compared to those from Suratgarh, the non-irrigated region, indicating a complex interplay of environmental factors. The observed variations provide insights into the environmental influences on secondary metabolite content of Calotropis procera flowers, contributing to a deeper understanding from ecological and agricultural perspectives.
... Leguminosae; commonly known as the bean, pea, or legume family; includes trees, shrubs, and perennial or annual herbaceous plants; are flowering plants that play a key role in the diversity and density of forest ecosystems [1] also environmentally sustainable agriculture [2]. Legumes have a distinctive floral structure, podded fruit, protein-rich physiology, and provide many secondary metabolites [3,4]. It has unique ecological role in nitrogen acquisition, with 88% of the species having the ability to form nodules with Rhizobium bacteria to fix atmospheric nitrogen into compounds available to the biotic world [5]. ...
... The resulting effect may be not only positive, but may also compensate for, attenuate, or cancel out possible undesirable effects of other components [9,10,16]. MOCS components may modulate their mutual resorption but may also interact with specific compounds ingested from food or medicines and may also affect typical characteristics of smell and taste [18,19]. ...
Article
Full-text available
Medicinal mushrooms are multicomponent mixtures (MOCSs). They consist of a large number of individual compounds, each with different chemical structures, functions, and possible pharmacological activities. In contrast to the activity of an isolated pure substance, the effects of the individual substances in a mushroom or its extracts can influence each other; they can strengthen, weaken, or complement each other. This results in both advantages and disadvantages for the use of either a pure substance or a multicomponent mixture. The review describes the differences and challenges in the preparation, characterization, and application of complex mixtures compared to pure substances, both obtained from the same species. As an example, we use the medicinal and culinary mushroom Lentinula edodes, shiitake, and some of its isolated compounds, mainly lentinan and eritadenine.
Article
Full-text available
Flavonoids are a class of polyphenolic compounds that naturally occur in plants. Sub-groups of flavonoids include flavone, flavonol, flavanone, flavanonol, anthocyanidin, flavanol and isoflavone. The various modifications on flavonoid molecules further increase the diversity of flavonoids. Certain crops are famous for being enriched in specific flavonoids. For example, anthocyanins, which give rise to a purplish color, are the characteristic compounds in berries; flavanols are enriched in teas; and isoflavones are uniquely found in several legumes. It is widely accepted that the antioxidative properties of flavonoids are beneficial for human health. In this review, we summarize the classification of the different sub-groups of flavonoids based on their molecular structures. The health benefits of flavonoids are addressed from the perspective of their molecular structures. The flavonoid biosynthesis pathways are compared among different crops to highlight the mechanisms that lead to the differential accumulation of different sub-groups of flavonoids. In addition, the mechanisms and genes involved in the transport and accumulation of flavonoids in crops are discussed. We hope the understanding of flavonoid accumulation in crops will guide the proper balance in their consumption to improve human health.
Article
Full-text available
Nitrogen is the most limiting macronutrient for plant growth, and rhizobia are important bacteria for agriculture because they can fix atmospheric nitrogen and make it available to legumes through the establishment of a symbiotic relationship with their host plants. In this work, we studied the nitrogen fixation process in the microsymbiont Sinorhizobium fredii at the genome level. A metabolic model was built using genome annotation and literature to reconstruct the symbiotic form of S. fredii . Genes controlling the nitrogen fixation process were identified by simulating gene knockouts. Additionally, the nitrogen-fixing capacities of S. fredii CCBAU45436 in symbiosis with cultivated and wild soybeans were evaluated. The predictions suggested an outperformance of S. fredii with cultivated soybean, consistent with published experimental evidence. The reconstruction presented here will help to understand and improve nitrogen fixation capabilities of S. fredii and will be beneficial for agriculture by reducing the reliance on fertilizer applications.
Article
Full-text available
Soybean (Glycine max L. Merrill) isoflavones are secondary metabolites of great interest because of their beneficial impact on human health. We profiled the seed isoflavone composition of 1168 soybean accessions collected from diverse ecoregions of China in three locations over two years. We observed significant differences in isoflavone content among the accessions, accession types, years of growth and ecoregions of origin. Total isoflavone (TIF) concentration of the soybean accessions ranged from 745 μg g−1 to 5253.98 μg g−1, which represents a 7-fold difference. The highest mean TIF concentration (2689.27 μg g−1) was observed in the Huang Huai Hai Valley Region (HR) accessions, followed by accessions from the Southern Region (SR) and Northern Region (NR) with TIF concentration of 2518.91 μg g−1 and 1942.78 μg g−1, respectively. Thirty-five accessions were identified as elite soybean resources based on their higher TIF concentration (4024.74 μg g−1 to 5253.98 μg g−1). Pairwise correlation analysis showed significant positive correlations between individual isoflavones and TIF concentrations. Malonyldaidzin and malonylgenistin showed the highest correlations with TIF concentration (r = 0.90 and r = 0.92, respectively), whereas acetyldaidzin showed the lowest correlation. The main isoflavone components had significant negative correlations with latitude and longitude, indicating that the geographical origin of the accessions influenced their seed isoflavone composition. Based on principal component analysis, glycosides and malonylglycosides of isoflavones were the major discriminative components for the soybean accessions. The present study demonstrated the geographical distribution of soybean seed isoflavone concentrations across the main ecoregion of China. The identified soybean accessions with both high and low TIF concentrations, which are desirable materials for industrial uses and could also be used as parents to breed soybean lines with improved isoflavone quantity and composition in the seeds.
Article
Full-text available
Triterpenoid saponins are major secondary metabolites in soybean; they are classified into group A and DDMP saponins. Group A saponins cause bitterness and an astringent aftertaste, whereas DDMP saponins are less bitter and more beneficial to human health. Therefore, a need exists for genetic improvement to develop bitter-free soybeans without losing other beneficial saponins. In this study, an ethyl methanesulfonate (EMS)–induced mutant, PE1327, was isolated and characterized as having a reduced level of group A saponins by chromatography. The PE1327 phenotype, which exhibits no accumulation of group A saponins, resembles that of a previously reported sg-5 wild soybean mutant whose biosynthesis is controlled by the Sg-5 (Glyma.15G243300) gene. The Sg-5 gene sequence in PE1327 revealed a single-nucleotide polymorphism (G1036A) that causes an amino acid change from Glu346 to Lys346 in the sg-5 protein. Multiple alignment analysis of cytochrome P450 enzymes from numerous organisms revealed that Glu346 is highly conserved in eukaryotes, including plants and animals, but not in prokaryotes. The predicted 3D model of the Sg-5 protein showed that the Glu346 residue is located in the J helix and is likely involved in the linkage between the J and K helices and the stabilization of the J-K loop, suggesting that the amino acid substitution from Glu346 to Lys346 in the PE1327 mutant results in hypofunction of Sg-5. Co-segregation analysis revealed that the Sg-5 locus is tightly linked to reduced group A saponin biosynthesis in PE1327 and that the novel sg-5 variant is recessive to Sg-5.
Article
Full-text available
Polyamines (PAs) are natural aliphatic amines involved in many physiological processes in almost all living organisms, including responses to abiotic stresses and microbial interactions. On other hand, the family Leguminosae constitutes an economically and ecologically key botanical group for humans, being also regarded as the most important protein source for livestock. This review presents the profuse evidence that relates changes in PAs levels during responses to biotic and abiotic stresses in model and cultivable species within Leguminosae and examines the unreviewed information regarding their potential roles in the functioning of symbiotic interactions with nitrogen-fixing bacteria and arbuscular mycorrhizae in this family. As linking plant physiological behavior with “big data” available in “omics” is an essential step to improve our understanding of legumes responses to global change, we also examined integrative MultiOmics approaches available to decrypt the interface legumes-PAs-abiotic and biotic stress interactions. These approaches are expected to accelerate the identification of stress tolerant phenotypes and the design of new biotechnological strategies to increase their yield and adaptation to marginal environments, making better use of available plant genetic resources.
Article
Full-text available
The first chapter of Darwin’s On the Origin of Species famously used examples of human selection of domesticated plants and animals to lay the groundwork for the Theory of Evolution (Darwin 1859). Since at least the works of Darwin (1883), domestication of plants and animals have been used as major examples of strong selection radically altering the morphology, architecture, and behavior of organisms on which our contemporary society relies for food, fiber and fuel. Consequently, it is not surprising that crop domestication remains a vibrant area of research. Despite this ongoing interest in the field of domestication, we greatly lack ecological and natural history studies of crop wild relatives in their wild settings.
Article
The composition of “group A saponin” in the seed hypocotyl of soybean was analyzed, and six kinds of “group A saponins,” named Aa, Ab, Ac, Ad, Ae and Af according to their elution order from HPLC, were detected. On the basis of chemical and physicochemical evidence, Aa, Ab, Ae and Af are shown to be identical with the known saponins, acetyl-soyasaponins A4, A1, A5 and A2, respectively, whilst Ac and Ad were novel saponins possessing different sugar chains. In all these saponins, the terminal sugar of the oligosaccharide chain attached to C-22 of an aglycone (soyasapogenol A) was acetylated in the soybean seed.
Article
Glyceollin isomers I, II, and III are the major pathogen-elicited secondary metabolites (i.e. phytoalexins) of soybean (Glycine max) that, collectively with other 5-deoxyisoflavonoids, provide race-specific resistance to Phytophthora sojae. The NAC-family transcription factor (TF) GmNAC42-1 is an essential regulator of some but not all glyceollin biosynthesis genes, indicating other essential TF(s) of the glyceollin gene regulatory network remain to be identified. Here, we conducted comparative transcriptomics on soybean hairy roots of the variety Williams 82 (W82) and imbibing seeds of Harosoy 63 (H63) upon treatment with wall glucan elicitor (WGE) from P. sojae and identified two homologous R2R3-type MYB TF genes, GmMYB29A1 and GmMYB29A2, upregulated during the times of peak glyceollin biosynthesis. Overexpression and RNAi silencing of GmMYB29A2 increased and decreased expression of GmNAC42-1, GmMYB29A1, and glyceollin biosynthesis genes and metabolites, respectively, in response to WGE. By contrast, overexpressing or silencing GmMYB29A1 decreased glyceollin I accumulation with marginal or no effects on the expressions of glyceollin synthesis genes, suggesting a preferential role in promoting glyceollin turnover and/or competing biosynthetic pathways. GmMYB29A2 interacted with the promoters of two glyceollin I biosynthesis genes in vitro and in vivo. Silencing GmMYB29A2 in W82, a soybean variety that encodes the resistance gene Rps1k, rendered it compatible with race 1 P. sojae, whereas overexpressing GmMYB29A2 rendered the susceptible Williams variety incompatible. Compatibility and incompatibility coincided with reduced and enhanced accumulations of glyceollin I but not other 5-deoxyisoflavonoids. Thus, GmMYB29A2 is essential for accumulation of glyceollin I and expression of Phytophthora resistance.
Chapter
Lupins provide an insightful model for plant domestication with five species domesticated over a wide range of time and geography. The most intensively studied species is narrow-leafed lupin, a twentieth-century domesticate where the addition of each successive domestication trait was documented in the scientific literature. Foundational to the advances made in our understanding of lupin domestication was the availability of excellent genetic resources: Well-annotated wild seed collections, published pedigrees of Australian narrow-leafed lupin cultivars and a suite of wild × domesticated cross populations. Rapid developments in genomic technologies culminating in the reference genome for narrow-leafed lupin have greatly increased our understanding of the origins of domesticated lupins, how diversity has been profoundly affected and the molecular control of domestication genes. This chapter provides an overview of our current understanding of lupin domestication and how this knowledge can equip lupin breeders to create more diverse and productive cultivars.
Chapter
Legume crops are subjected to a wide range of abiotic stresses, which stimulate an array of physiological, biochemical, and molecular responses. However, different genotypes may exhibit significant variations between individual responses, which can determine their tolerance or susceptibility to these stresses. The present chapter suggests a broad range of assays that can help in understanding stress perception by plants at cellular and molecular levels. The genotypes may be sorted depending on their tolerance potential, by broadly analysing morphological, physiological, biochemical, and enzyme kinetics parameters. These assays are very beneficial in revealing the mechanism of stress perception and response in varied plant types, and have helped in discriminating contrasting genotypes. Here, we have described detailed protocols of assays which may be carried out to assess tolerance or susceptibility to abiotic stresses. The analysis, as a whole, can help researchers understand the effect of abiotic stresses on plant biochemical pathways, be it photosynthesis, redox homeostasis, metabolite perturbation, signaling, transcription, and translation. These protocols may be beneficial in identification of suitable donors for breeding programs, as well as for identifying promising candidate genes or pathways for developing stress tolerant legume crops through genetic engineering.