ArticlePDF AvailableLiterature Review

AGL15 Promotion of Somatic Embryogenesis: Role and Molecular Mechanism

Frontiers
Frontiers in Plant Science
Authors:

Abstract and Figures

Plants have amazing regenerative properties with single somatic cells, or groups of cells able to give rise to fully formed plants. One means of regeneration is somatic embryogenesis, by which an embryonic structure is formed that “converts” into a plantlet. Somatic embryogenesis has been used as a model for zygotic processes that are buried within layers of maternal tissues. Understanding mechanisms of somatic embryo induction and development are important as a more accessible model for seed development. We rely on seed development not only for most of our caloric intake, but also as a delivery system for engineered crops to meet agricultural challenges. Regeneration of transformed cells is needed for this applied work as well as basic research to understand gene function. Here we focus on a MADS-domain transcription factor, AGAMOUS-Like15 (AGL15) that shows a positive correlation between accumulation levels and capacity for somatic embryogenesis. We relate AGL15 function to other transcription factors, hormones, and epigenetic modifiers involved in somatic embryo development.
This content is subject to copyright.
SYSTEMATIC REVIEW
published: 28 March 2022
doi: 10.3389/fpls.2022.861556
Frontiers in Plant Science | www.frontiersin.org 1March 2022 | Volume 13 | Article 861556
Edited by:
Huihui Guo,
Shandong Agricultural
University, China
Reviewed by:
Wei Gao,
Henan University, China
Xiyan Yang,
Huazhong Agricultural
University, China
*Correspondence:
Sharyn E. Perry
sperr2@uky.edu
Specialty section:
This article was submitted to
Plant Cell Biology,
a section of the journal
Frontiers in Plant Science
Received: 24 January 2022
Accepted: 01 March 2022
Published: 28 March 2022
Citation:
Joshi S, Paul P, Hartman JM and
Perry SE (2022) AGL15 Promotion of
Somatic Embryogenesis: Role and
Molecular Mechanism.
Front. Plant Sci. 13:861556.
doi: 10.3389/fpls.2022.861556
AGL15 Promotion of Somatic
Embryogenesis: Role and Molecular
Mechanism
Sanjay Joshi 1, Priyanka Paul 2, Jeanne M. Hartman 1and Sharyn E. Perry 1
*
1Department of Plant and Soil Sciences, University of Kentucky, Lexington, KY, United States, 2Kentucky Tobacco Research
and Development Center, University of Kentucky, Lexington, KY, United States
Plants have amazing regenerative properties with single somatic cells, or groups of
cells able to give rise to fully formed plants. One means of regeneration is somatic
embryogenesis, by which an embryonic structure is formed that “converts” into a
plantlet. Somatic embryogenesis has been used as a model for zygotic processes that
are buried within layers of maternal tissues. Understanding mechanisms of somatic
embryo induction and development are important as a more accessible model for
seed development. We rely on seed development not only for most of our caloric
intake, but also as a delivery system for engineered crops to meet agricultural
challenges. Regeneration of transformed cells is needed for this applied work as well
as basic research to understand gene function. Here we focus on a MADS-domain
transcription factor, AGAMOUS-Like15 (AGL15) that shows a positive correlation
between accumulation levels and capacity for somatic embryogenesis. We relate AGL15
function to other transcription factors, hormones, and epigenetic modifiers involved in
somatic embryo development.
Keywords: somatic embryo, regeneration, MADS-box, transcription factor, chromatin immunoprecipitation,
transcriptome, Arabidopsis thaliana
INTRODUCTION
Somatic embryogenesis (SE) is a powerful resource and biotechnological tool used for plant
improvement, micropropagation, and clonal regeneration (Guan et al., 2016; Ochoa-Alejo and
Loyola-Vargas, 2016). Because zygotic embryos (ZE) are embedded deep in maternal tissues, SE has
become a widely used technique to understand processes that occur during zygotic embryogenesis.
Somatic cells that undergo reprogramming to induce embryonic development produce structures
called somatic embryos.
Understanding SE is important for several other basic and applied scientific approaches that are
distinct from using SE as a model for ZE. Earth is in jeopardy due to climate change. Not only will
crops suffer from abiotic stresses such as drought and heat, but also “new” biotic stresses that can
dominate new territories due to the altered climate. While one hopes we can deal with this problem
Joshi et al. AGL15 and SE
at the source, it seems increasingly likely we will need alternative
strategies to deal with the evolving climate. One such strategy
to continue to feed our growing population is biotechnology to
engineer crops to cope with different environmental stressors
and to propagate the most valuable genotypes as efficiently as
possible (Ikeda and Kamada, 2005; Tian et al., 2020a). One
challenge to genetic engineering is the introduction of the desired
modifications into the genome. Another is the regeneration
of a plant from the modified cell(s), and SE is one means
to accomplish this latter task, although many plants and even
particular genotypes do not regenerate well by SE. SE is of
particular interest in plant developmental biology because of
its involvement in plant plasticity, especially controlling the
totipotency and pluripotency in somatic cells (Ikeuchi et al., 2016;
Fehér, 2019). Interestingly, there are some parallels with animal
cells acquiring the ability to dedifferentiate in response to stress
(Grafi, 2009).
Several studies have been performed to decipher the
regulatory mechanism of SE, and the model system Arabidopsis
thaliana (Arabidopsis) has been used for much of this work
(Yang and Zhang, 2010). Arabidopsis is a model plant due to
its available genetic resources, with a small genome size and
whole-genome sequence available from an early date (Kaul
et al., 2000). These facts and its ability to be transformed
very easily compared to other species that require intensive
tissue culture (Clough and Bent, 1998) means that there is a
database of mutants with T-DNAs inserted into nearly every
gene-of-interest, generating a loss-of-function mutant that
facilitates understanding gene function (https://conf.arabidopsis.
org/display/COM/Mutant$+$and$+$Mapping$+$Resources).
Arabidopsis zygotic embryogenesis has a predictable pattern
and basic body plan from early division (Palovaara et al., 2016;
Plotnikova et al., 2019), and SE generally follows the same stages.
While there are differences in molecular profiles between ZE and
SE, there are also congruencies worthy of further investigation
(Tian et al., 2020a). Like ZE, SE generates bipolar structures
from somatic cells with no vascular connection to the explant
tissue (Thorpe and Stasolla, 2001; Von Arnold et al., 2002). A
plethora of literature discusses genes encoding transcription
factors (TF) that are involved in seed development and can
promote SE or embryo-programs when ectopically expressed,
including the so-called LAFL factors (LEAFY COTYLEDON1
[LEC1], LEAFY COTYLEDON2 [LEC2], FUSCA3 [FUS3]
and ABSCISIC ACID INDPENDENT3 [ABI3]) (Wójcikowska
et al., 2013), BABY BOOM (BBM) (Boutilier et al., 2002; Jha
and Kumar, 2018), WUSCHEL (WUS) (Jha et al., 2020), and
AGAMOUS-LIKE15 (AGL15) (Harding et al., 2003; Zheng
et al., 2013a,b). Here we focus on AGL15 but include some
discussion of relation to the other TFs as they interact in a
complex network.
AGL15 is a member of the MADS domain TFs (named
for MCM1 from Saccharomyces cerevisiae,AGAMOUS from
Arabidopsis, DEFICIENS from Antirrhinum majus,SRF from
Homo sapiens) (Gramzow and Theissen, 2010; Chen et al.,
2017) that is expressed and primarily accumulates during early
stages of seed development, and mostly within the embryo
(Heck et al., 1995; Perry et al., 1996). The transcript level of
AGL15 in Brassica napus, and Arabidopsis, remains high during
embryo morphogenesis until the seeds start to dry (Perry et al.,
1996). Ectopic constitutive expression of AGL15 promotes SE in
Arabidopsis, Glycine max (soybean), and Gossypium hirsutum
(cotton), while mutants (loss-of-function) of both AGL15 and
AGL18 (closely related to AGL15) significantly reduce SE in
Arabidopsis (Harding et al., 2003; Thakare et al., 2008; Zheng and
Perry, 2014). Like AGL15,AGL18 is able to promote SE when
ectopically expressed in soybean and Arabidopsis (Zheng and
Perry, 2014; Paul et al., 2021).
Here we focus on genes directly regulated by AGL15 to
hypothesize about the stage of SE impacted by overexpression
of AGL15. We highlight how AGL15 takes part in different
processes involved in expression of genes to promote
somatic embryogenesis.
STAGES OF SOMATIC EMBRYOGENESIS
(SE)
In SE, the process initiates when somatic cells are induced to
change fate and form embryogenic cells that can eventually
regenerate into a complete plant. This process can be direct or
indirect (Sharp, 1980) with the latter including a callus phase.
Many factors influence the ability to undergo SE, including
plant type and even particular cultivars within a species,
physiological status of the donor plant, tissue source and age
of the explant, media components, and a variety of stresses
(Yang and Zhang, 2010).
Recent reviews have addressed the terms dedifferentiation vs.
trans-differentiation. While dedifferentiation has been used to
describe the switch from a somatic cell to an “undifferentiated”
callus, it is increasingly recognized that there are no truly
undifferentiated cells in plant regenerative processes. In fact,
callus was recently recognized as having molecular profiles
similar to root meristem tissue (Sugimoto et al., 2010).
Dedifferentiation is still used to refer to a decrease in
specialization, whereas trans-differentiation refers to a change
from one cell type to a different cell type (Fehér, 2019; Bidabadi
and Jain, 2020). While particulars to induce SE vary, a “theme
to induce SE is the treatment of explants with auxin, usually
2,4-D, a synthetic auxin that may act as an auxin, a stressing
reagent, and/or to induce endogenous auxin production. Stress
can also take the form of wounding, temperature, and nutritional
stresses, among others. This process is thought to promote
developmental switching via a series of molecular events and
signal cascades resulting in the transition of somatic cells to
embryo identity, or callus for indirect systems (Yang and Zhang,
2010; Fehér, 2015). The unique developmental process can
be distinguished into two phases: the initiation/induction step
and the developmental phase. Cell proliferation and increased
plasticity of somatic cells occur in the first step, whereas cells
differentiate to form somatic embryos when the culture cells get
the right stimuli during the developmental phase (Magnani et al.,
2017). A brief overview of different steps follows and is shown in
Figure 1.
Frontiers in Plant Science | www.frontiersin.org 2March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
FIGURE 1 | Diagram of stages of SE. Please see the text for a description of
what the numbers mean.
1. Initiation/induction—This is the step where the establishment
of pre-embryogenic and embryogenic cells or groups of cells
from the primary explant takes place. In this stage, plant
growth hormones like auxin, and in some systems, cytokinin,
play a vital role in generating cells responsive for the transition.
Stresses are also commonly applied. Large epigenetic changes
occur in response to these stimuli, and these can be seen at
a cellular level by increased nuclear volume. The competent
cells start activating the genes responsible for producing
embryogenic cells (Su et al., 2021), including genes encoding
TFs that then activate downstream pathways to generate the
SE (Quiroz-Figueroa et al., 2006).
2. Proliferation/maintenance/multiplication—Once the cells
are responsive and have entered embryogenetic fate, the
proliferation of cells initiates, which can be bulked up usually
on semisolid media or solid supports. Maintaining physical
and chemical conditions, are crucial to maintaining the
embryogenic process.
3. Maturation (histodifferentiation)—This involves
morphological stages of somatic embryogenesis analogous
to zygotic embryo development including establishment of
shoot and root apical meristems. During this step, auxin is
often removed from the media. For instance, often in indirect
systems, SEs are only obvious upon removal or a decrease
in the concentration of plant growth regulators (PGR). It is
thought molecular patterns supporting development to the
globular stage are established on the high PGR medium, and
the decrease in exogenously added PGRs is needed to continue
development (Zimmerman, 1993).
4. Post-maturation (drying, desiccation)—This stage involves
mild drying that can switch the embryo from development to
germination or for storage purposes.
5. In vitro germination (radicle elongation).
6. Conversion (establishment of the new plant).
SYSTEMS WHERE AGL15 HAS BEEN
FOUND TO PROMOTE SOMATIC
EMBRYOGENESIS (SE)
Ectopic overexpression of AGL15 has been shown to promote SE
in several plants (Figure 2). First, a system where Arabidopsis
zygotic embryo explants are placed onto Murashige and Skoog
(MS) medium lacking any exogenous plant growth regulators
(PGR) show a positive correlation between AGL15 accumulation
and ability to produce SE and maintain this development for
long time periods (Harding et al., 2003). We refer to this tissue
as Embryo Culture Tissue (ECT). Little to no apparent callus
is observed in this system. While the Arabidopsis ecotypes
Columbia (Col) and Wassilewskija (WS) wild-type (wt) embryos
can transiently produce SE tissue in this system, they switch to
leaves early. Meanwhile 35S:AGL15 increases production of ECT
in WS and leads to long term maintenance for both ecotypes
(25 years to date for WS; 10 years to date for Col) as SE tissue.
Loss-of-function or dominant negative forms of AGL15 reduce
initiation of this tissue. Because AGL15 transcript accumulation
is induced in response to auxin treatment (Zhu and Perry, 2005),
one possibility is the overexpression of transgene removes the
need for exogenous auxin. Kurczynska et al. (2007) demonstrated
that the cells competent for SE are in the protodermis and
subprotodermis of the adaxial side of the cotyledons. Whether
the 35S:AGL15 transgene enhances this competence or expands
it to other tissues remains a question for the future.
Apart from ECT, there has been another SE system where
the mature seeds are allowed to complete germination in
liquid media where synthetic auxin 2,4-D is present. After 21
days of culture, development at the shoot apex region of the
seedlings is counted for those with SE tissue and those lacking
this tissue. Other than SE development at the meristem, the
remainder of the “seedling” is callused. We refer to this system as
Shoot Apical Meristem Somatic Embryogenesis (SAM SE). The
findings show that 35S:AGL15 produced approximately twice
the frequency of seedlings with SAM SE compared to Col, wt
(for example, 56% for 35S:AGL15, compared to 29% for Col,
wt; Zheng et al., 2009). In SAM SE, 2,4-D is required to obtain
SE development; medium lacking 2,4-D simply produces liquid
grown seedlings. However, different sensitivity to 2,4-D was
observed for 35S:AGL15 compared to Col, wt (Zheng et al., 2016).
We expanded this work to Glycine max (soybean). Genes
encoding orthologs of AGL15 from soybean (GmAGL15) and
the closely related MADS-factor that has redundant functions
(GmAGL18) were cloned and introduced into SE tissue by
biolistic transformation. We found that increased expression
of these genes via a 35S promoter could 1. Enhance recovery
of transgenic lines, presumably by promoting regeneration of
transformed cells by SE and 2. Produce transgenic plants that
have enhanced ability to form SE tissue. Meanwhile, Yang
et al. (2014) cloned three AGL15 orthologs from cotton and
demonstrated that overexpression could promote SE in this plant.
While this summarizes the studies where AGL15
accumulation has been manipulated to show effects on SE
development, other studies show a correlation between SE
Frontiers in Plant Science | www.frontiersin.org 3March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
FIGURE 2 | Systems where manipulation of AGL15 levels have impacted on SE. Modified from Tian et al. (2020a).(A,B) two systems in Arabidopsis; (C) in soybean;
(D) in cotton. D40, medium with 40 mg/L 2,4-D, D20, medium with 20 mg/L 2,4-D, CIM, callus induction medium.
and AGL15 accumulation. Potential orthologs of AGL15 are
shown in Figure 3.Perry et al. (1999) demonstrated that
embryos from different sources, including apomictic, microspore
embryos, and SE accumulated protein that reacted with AGL15
antiserum. Tvorogova et al. (2019) overexpressed another TF
called MtWOX9 in Medicago truncatula and found increased
SE capacity. MtAGL15 transcript accumulation increased in
response to MtWOX9. In addition, a highly embryogenic line
called 2HA showed increased MtAGL15 transcript with SE
formation. The non-embryogenic line (A17) showed no such
increase. Ghadirzadeh-Khorzoghi et al. (2019) showed increased
PvAGL15 transcript in embryogenic explants of pistachio, while
Xu et al. (2018) showed RcAGL15 associated with protocorm-like
bodies in rose.
GENE REGULATION BY AGL15 DURING
SOMATIC EMBRYOGENESIS
AGL15 is a central regulator in SE and to understand how
it directly and indirectly controls this process, expressed
and repressed target genes were identified using chromatin
immunoprecipitation (ChIP) and transcriptome studies in
Arabidopsis and soybean (Zheng et al., 2009; Zheng and Perry,
2014). Direct targets (Arabidopsis where a genome-wide study
was performed) were found to be enriched for a cis motif called a
CArG motif that is a binding site for MADS-factors like AGL15.
The consensus CArG is CC(A/T)6GG, but AGL15 can also bind
variants of this motif with a longer A/T rich core (Tang and
Perry, 2003). Studies to determine how ectopic expression of TFs
promote SE have resulted in strategies to enhance SE from non-
transgenic tissue. Examples for AGL15 include manipulation of
gibberellin metabolism and ethylene biosynthesis and response
(discussed further below).
AGL15 interacts in a complex network with several other TFs
that promote SE, including members of the LAFL TFs, BBM, and
WUS. This was recently reviewed in Tian et al. (2020a), and some
of these interactions are highlighted in Figure 4.
Comparison of targets of AGL15 in Arabidopsis and soybean
revealed some intriguing similarities. In both species, genes
involved in stress responses were overrepresented among the
direct responsive targets. These include a set of WRKY, NAC,
Frontiers in Plant Science | www.frontiersin.org 4March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
FIGURE 3 | Phylogenetic tree of selected potential AGL15 orthologs, some of which promote SE development (please see text). Protein sequences were retrieved
from Phytozome (Goodstein et al., 2012) and submitted to the phylogenetic analysis tool NGPhylogeny.fr (Lemoine et al., 2019) with default settings. The output tree
was then uploaded to iTOL (version 6.5) (Letunic and Bork, 2019) for tree visualization. At, Arabidopsis thaliana; Os, Oryza sativa; GRMZM, Zea mays; Ca, Coffea
arabica; Sl, Solanum lycopersicum; Glyma, Glycine max; Mt, Medicago truncatula; Rc, Rosa canina; Pv, Pistacia vera; Gohir, Gossypium hirsutum.
and bZIP TFs that have been associated with dedifferentiation
(Grafi et al., 2011). A more recent ChIP-seq study for Arabidopsis
AGL15 revealed that regulatory regions corresponding to genes
encoding nearly all of these TFs are directly bound by AGL15,
which seems quite extraordinary (Paul et al., 2021). When
comparing the percentage of genes encoding these TFs bound by
AGL15 to the percentage of genes bound overall compared to the
whole genome, a significant enrichment in the dedifferentiation-
related (DD) TFs potentially directly regulated by AGL15 is
observed (Table 1). Interestingly, many of the other TFs that
can promote SE also show significant enrichment of these genes
among their direct targets (Table 1). Detailed information on
what SE TF associates with what DD TF gene is provided in
Supplementary Table 1.
A number of gene ontology (GO) categories are
overrepresented among the direct (as determined by ChIP-
chip and/or ChIP-seq) AGL15 targets that are expressed or
repressed. Select categories are shown in Figure 5. Notably,
genes involved in abiotic stress response are overrepresented,
as are hormone response, particularly hormones involved in
stress. Stresses such as those shown in Figure 5 are often used
to induce SE. Interestingly, genes associated with “chitin”
are overrepresented among AGL15 direct expressed targets
(Figure 5A). Several studies have found that lipophilic chitin
oligosaccharides, endochitinases, and arabinogalactan proteins
(AGP) have roles in stimulating SE (Van Hengel et al., 1998, 2002;
Domon et al., 2000) and also are responsive to stress (reviewed
in Grover, 2012; Mareri et al., 2019). One well-studied SE-related
chitinase is EXTRACELLULAR PROTEIN3 (EP3) from Daucus
carota (carrot), where addition of EP3 can rescue the production
of SE in a SE-defective line of carrot (Van Hengel et al., 1998). In
soybean, 35S:GmAGL15 was found to initially increase transcript
accumulation from a gene encoding a potential EP3 ortholog
in the explants. However, at later time points after culture on
D40 medium, differential transcript accumulation compared
to control was lost due to more extensive up-regulation in the
control compared to the overexpressor (Zheng and Perry, 2014).
DcEP3 is not expressed in the somatic (or zygotic) embryos, but
rather in cells surrounding the embryos and are thought to have
a nursing function, in part by cleavage of AGPs, some of which
have stimulatory and others inhibitory effects on SE (Toonen
et al., 1997; Van Hengel et al., 2001). Some AGPs have been
found to be important in the conditioned medium to promote
SE, and at least some are expressed in the non-embryogenic cells
in embryogenic cultures (reviewed in Von Arnold et al., 2002).
Interestingly, AGPs, including the subgroup of fasciclin-like
AGPs, (FLAs) were prevalent on the AGL15 direct repressed
list (Zheng et al., 2009), a list that became longer with the
ChIP-seq data (Paul et al., 2021) (Table 2). Only one (AGP14)
showed direct expression in response to AGL15 accumulation,
with one additional (FLA10) showing reduced transcript in
both comparisons.
AGL15 AND HORMONE RELATIONS IN
SOMATIC EMBRYOGENESIS
Hormones/PGRs play a crucial role in SE induction. Thus,
to understand SE, it is important to understand the signaling
processes related to both the exogenous and endogenous PGRs.
Generally, SE depends on the type and concentration of PGR
used for each culture medium. Recent studies showed that
Frontiers in Plant Science | www.frontiersin.org 5March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
FIGURE 4 | Select interactions between TFs involved in SE and control of
hormone accumulation/signaling. Not all interactions are shown in
consideration of clarity. (A) Particular gene interactions; (B) summary of tested
or presumed AGL15 effects on different hormones. Different colored ovals with
TF names indicate different families of TFs: yellow, B3 domain; blue,
CCAAT-box binding; pale pink, APETALA2 (AP2) or AP2/ERF; mint, MADS;
dark green, BZR domain; lavender, AUX/IAA domain; orange, DELLA/GRAS;
light orange, homeodomain. Gray boxes indicate hormone or other responses.
Gray hexagon, miRNA; light blue hexagon, SERK1.
different plant species, including Arabidopsis (Wójcikowska
et al., 2013), Coffea canephora (Márquez-López et al., 2018),
and Musa spp. (Awasthi et al., 2017), successfully responded
to SE induction using variable explants, conditions, and
concentrations of PGRs. Although hormones are only part of
the composition of a culture medium, it has been estimated
that design of a new tissue culture medium requires optimizing
components that can amount to 16,000 different treatments,
leading to artificial intelligence/optimization algorithms to
assist development (Hesami et al., 2020). Thus, understanding
mechanisms of SE, including how TFs that can promote SE
regulate and respond to PGRs, is important to help guide these
efforts. Here, we focus on the relation between AGL15, select
other TFs that promote SE and select PGRs, and the underlying
mechanisms that regulate different hormone signaling pathways.
A working model is shown in Figure 4. It should be noted that
hormone interactions are exceedingly complex with cooperative
or antagonistic interactions depending on the particular context,
so this model is a simplification focusing on downstream targets
of AGL15 and how they may interact to promote SE.
AGL15 and Auxin
One of the most common components in culture media for
SE is auxin, often supplied as 2,4-D, a synthetic auxin. As
mentioned above, it has been proposed that this exogenously
added PGR may act as an auxin, as a stressing agent, and/or
to induce endogenous auxin production. Several TFs that can
induce SE respond either directly or indirectly to auxin including
BBM, FUS3, LEC1, LEC2, WUS, and AGL15 (reviewed in Tian
et al., 2020a). While it does not encode a TF per se,SOMATIC
EMBRYOGENESIS RECEPTOR LIKE KINASE 1 (SERK1), is a
key marker for cells able to form somatic embryos, and, when
expressed via a strong constitutive promoter, can enhance SE
production in multiple species. SERK1 transcript accumulation
is increased by auxin and stress (for a review, please see Mendez-
Hernandez et al., 2019). Interestingly, SERK1 has been found
to interact with protein complexes that include AGL15, as
well as proteins involved in brassinosteroid signaling (Karlova
et al., 2006 and please see below). Both SERK1 and AGL15 are
positively (possibly indirectly) regulated by WUS (Busch et al.,
2010). Further studies in Arabidopsis have demonstrated the
importance of auxin in SE (Gaj et al., 2006) as well as an optimal
2,4-D treatment that can be explained, at least in part, by auxin
effect on SE-related TF genes. All of the genes in the study showed
increased and relatively stable (over the time course) transcript
accumulation at the optimal 2,4-D concentration for SE (5 µM)
(Grzybkowska et al., 2020). However, at 3 days of culture on
various media, LEC1, LEC2 (two of the previously mentioned
LAFL factors), and BBM showed a positive correlation of
transcript accumulation over the 2,4-D concentrations used,
including sub-optimal (0.1 µM) that led to shoot development,
5µM (optimal for SE), and supra-optimal (20 µM) that produced
non-embryogenic callus, whereas AGL15 and WUSCHEL (WUS)
had very little or no transcript accumulation other than at
the SE optimal 2,4-D concentration. Grzybkowska et al. (2020)
further linked this response to epigenetic regulation (discussed
below). While AGL15 is responsive to auxin (both 2,4-D and
the biologically relevant auxin indole acetic acid, IAA), it
is not an auxin early responsive gene and requires one to
two-day treatment to see a clear response (Zhu and Perry,
2005). Consistent with this is the lack of an auxin response
element in the promoter region; but auxin may act via an
ethylene responsive element (Grzybkowska et al., 2020). Putative
orthologs of AGL15 from cotton (Figure 3) are also significantly
induced by 2,4-D, and when ectopically expressed, enhance the
embryogenic potential of calli in 2,4-D containing medium (Yang
et al., 2014). While it is unclear that monocots have MADS-
factors that, when compared to Arabidopsis MADS, are most
similar to AGL15, in Zea mays AGL15-like gene expression
was correlated with early embryogenic culture initiation. For
two of the four genes encoding these maize AGL15-like genes,
transcript accumulation increased after placement on 2,4-D
Frontiers in Plant Science | www.frontiersin.org 6March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
TABLE 1 | Overrepresentation of TFs associated with dedifferentiation (DD TFs) among the genes with regions bound by TFs involved in SE.
TF # DD TFs bound Total # of genes bound Significance References
LEC1, early SE 13 3,035 0.00010 (Pelletier et al., 2017)
LEC1, late SE 17 4,890 0.00035 (Pelletier et al., 2017)
LEC1, cotyledon 22 7,216 0.00037 (Pelletier et al., 2017)
LEC2 19 2,788 5.2E-13 (Wang et al., 2020)
FUS3 4 1,218 NS (Wang and Perry, 2013)
ABI3 14 2,510 1.5E-07 (Tian et al., 2020b)
BBM 12 Top 5,000 NS (Horstman et al., 2015)
AGL15, ChIP-chip 15 3,708 7.6E-05 (Zheng et al., 2009)
AGL15, ChIP-seq 29 9,729 2.8E-05 (Paul et al., 2021)
AGL18, ChIP-seq 4 3,446 NS (Paul et al., 2021)
Chi-square test was used to compare the number of TFs associated with dedifferentiation (DD TFs; 44 total as listed in Grafi et al., 2011) bound by select regulators (TF) of embryogenesis,
compared to the total list of genes bound for each TF, using 27,334 genes total for Arabidopsis (NCBI, accessed 11/17/2021). Significance means that within the list bound by the TF,
DD TFs are overrepresented compared to expectations based on the total # of genes bound.
containing medium (Salvo et al., 2014). One of these genes is
shown in Figure 3. However, the soybean orthologs of AGL15,
(Glyma12g17721 and Glyma11g16105) did not respond to 2,4-
D, although an AGL18 ortholog (GmAGL18) was upregulated in
response to 2,4-D (Zheng and Perry, 2014).
The role of auxin in SE induction is of particular interest.
Auxin treatment is important for embryogenesis (both somatic
and zygotic); however, too high auxin concentration may hinder
the induction of embryogenic cells in the SE culture. It has
been observed that cells that are resistant to auxin in that they
maintain cell-cell adherence and do not elongate as much, may
be the particular cells capable of developing as embryos (Emons,
1994). In the Arabidopsis auxin signal transduction pathway,
the AUXIN/INDOLE-3-ACETIC ACID (AUX/IAAs) encoded
proteins bind to the Auxin Responsive Factors (ARFs), which,
in turn, are associated with DNA via auxin response elements
on auxin-responsive genes (Piya et al., 2014). AUX/IAAs
act as repressors of ARF function by keeping ARFs from
regulating gene expression until auxin perception (involving
auxin receptors TRANSPORT INHIBITOR RESPONSE1; TIR1,
and AUXIN SIGNALING F-BOX; ABFs) leads to degradation of
the IAA protein.
While auxin is important for SE, and other TFs that can induce
SE directly upregulate auxin biosynthetic genes (e.g., members
of the YUCCA gene family are induced by LEC1 and LEC2),
overexpression of YUCCAs alone cannot induce SE (reviewed in
Braybrook and Harada, 2008). It has been proposed that other
factors, including IAA30 (discussed below) and/or AGL15, may
be important for competency to respond to auxin. Part of this
“competency” could be limiting auxin response.
AGL15 controls several genes in a manner that would reduce
auxin response and/or accumulation. In soybean, we found
that 35S:GmAGL15 reduces endogenous IAA accumulation
compared to the wild type control in developing embryos (Zheng
et al., 2016). In Arabidopsis, 35S:AGL15 directly upregulated
the gene encoding an AUX/IAA transcriptional repressor,
IAA30, which may limit auxin responses and is important
for SE (Zheng et al., 2009). An iaa30 homozygous knockout
line in both Col, wt and 35S:AGL15 backgrounds showed a
significantly lower frequency of SAM SE development than in
the corresponding background with the wild type IAA30 gene
(Zheng et al., 2009). IAA30 is also a direct expressed target
of LEC2 (Braybrook and Harada, 2008). Aux/IAAs generally
contains four domains, including an N-terminal repression
domain (domain I), domain II that enables degradation of
Aux/IAAs via ubiquitin–proteasome pathway in response to
auxin, and domains III and IV are the protein-protein interaction
domains (Guilfoyle and Hagen, 2012). However, in Arabidopsis,
IAA30 lacks the domain II that is involved in binding auxin F-box
receptors in response to auxin perception. In addition, IAA20,
which is also a potentially upregulated target of AGL15, also
lacks domain II. As a result, unlike the canonical IAA proteins,
these Arabidopsis IAA proteins have a longer half-life (Sato and
Yamamoto, 2008) and overexpression of IAA30 may interrupt the
auxin responsiveness.
In addition (Gm)AGL15 (here (Gm)AGL15 is meant to
indicate similar observations for soybean and Arabidopsis
AGL15; AGL15 is meant to indicate the observation in
Arabidopsis to date) directly represses a gene encoding an
auxin receptor (Gm)TIR1 (Zheng et al., 2016) that facilitates
the degradation of AUX/IAA in response to auxin (Hayashi,
2012), and represses both (Gm)ARF6 and (Gm)ARF8 (Zheng
et al., 2016), which are reported as transcriptional activators in
the auxin signaling pathway (Piya et al., 2014). This regulatory
interaction appears to be direct for ARF6 in Arabidopsis with
binding by AGL15 observed in ChIP experiments (Zheng
et al., 2016). Both ARF6 and ARF8 are post-transcriptionally
controlled by miRNA167 and AGL15 upregulates one of the
genes encoding this microRNA, with evidence for a direct
interaction in Arabidopsis (Zheng et al., 2016). Loss-of-function
mutants of arf6 and the double arf6/arf8 showed a significant
increase in SAM SE (Zheng et al., 2016), but in other SE
systems (culture of immature zygotic embryos), there was a
reduction in SE efficiency or productivity for the arf6 and arf8
single mutants (Wójcikowska and Gaj, 2017). Thus, context
seems important.
Frontiers in Plant Science | www.frontiersin.org 7March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
FIGURE 5 | Overrepresented GO annotations among the direct (as measured
in ChIP-chip and/or ChIP-seq) and responsive (at least a 1.5-fold change at P
<0.05). Data is from the Panther Classification system accessed July 9, 2021.
(A) Directly expressed. (B) Directly repressed.
It is important to mention here that AUX/IAAS, ARFs, and
auxin receptors are multigene family proteins. Therefore, it is
feasible that other (Gm)AGL15 regulated members may promote
auxin response via other family members. For instance, there is
data supportive for direct expression of ARF5/MONOPTEROS
(MP) by AGL15 with Zheng et al. (2009) data showing a
significant reduction in mRNA in the agl15 agl18 double mutant,
although the 35S:AGL15 did not show increased accumulation
of this transcript. ARF5, when present as a loss-of-function,
had the largest effect on SE efficiency and productivity, but
overexpressing this gene did not produce any SE for the few
explants that could be obtained; again, indicating that restrictions
on auxin response may be important for SE (Wójcikowska and
Gaj, 2017). ARF3/ETTIN is repressed by AGL15, and transcript
accumulation of this ARF is negatively correlated with time
in SE culture (Wójcikowska and Gaj, 2017). Therefore, some
auxin responsive elements behave as would be believed consistent
with SE. We comment on how auxin interacts with other select
PGRs below.
AGL15 and GA
Auxin signaling leads to GA accumulation in some tissues by
activation of GA biosynthetic genes and/or GA catabolic gene
deactivation (Weiss and Ori, 2007). Therefore, it is possible that
repression of the auxin response by 35S:AGL15 may reduce
the biologically active GA, thereby promoting SE. Our prior
work corroborated this hypothesis as we found that AGL15
directly up regulates a GA catabolism gene GA2ox6 in both
Arabidopsis and soybean (Zheng et al., 2009, 2016). A reduction
of biologically active GA was also observed in 35S:AGL15
compared with the control (Wang et al., 2004). In addition, a
GA biosynthetic gene, GA3ox2 (At1g80340) was repressed by
AGL15 (Zheng et al., 2009). The accumulation of biologically
active GA was inversely correlated with SE in both Arabidopsis
and soybean (Wang et al., 2004; Zheng et al., 2016). The
addition of a GA biosynthesis inhibitor, paclobutrazol, to the
D40 medium promotes SE significantly in soybean (Zheng et al.,
2016), as well as SAM SE in Arabidopsis (Wang et al., 2004).
Interestingly, GA20ox1 (At4g25420), a gene encoding a GA
biosynthetic enzyme (Oh et al., 2014), was induced in auxin
response and directly correlated with ARF6. Thus, AGL15 also
regulates GA accumulation by controlling auxin response. In
addition, destabilization of DELLAs occurs in response to GA
(Weiss and Ori, 2007). The gibberellin-inactivated REPRESSOR
OF GA1-3 (RGA), a DELLA regulatory protein, prevents
the binding of ARF6 to DNA. GA perception leads to the
degradation of RGA, thereby enabling ARF6 to bind target genes
and cause auxin-responsive gene regulation. Another DELLA
protein, GIBBERELLIC ACID INSENSITIVE (GAI), is a putative
direct expressed target of ARF6 (Oh et al., 2014). Based on
microarray results, both the DELLA encoding genes GmRGA
and GmGAI showed increased transcript accumulation in the
35Spro:GmAGL15 on 2,4-D (Zheng et al., 2013a,b). Therefore, we
found a complex and interesting interaction between hormones
that would also include feedback mechanisms.
AGL15 and Ethylene
The effects of the gaseous hormone ethylene on plants have been
well-studied on plant growth, development, and stress responses.
The interaction between different hormonal and developmental
signals is critical in this response. According to our prior work,
(Gm)AGL15 impacts ethylene biosynthesis and perception,
which is relevant to SE (Zheng et al., 2013a). Auxin-ethylene
crosstalk reveals that they interact in a complex coregulatory
manner (Stepanova et al., 2008; Vandenbussche et al., 2012).
Auxin and ethylene impact each other’s accumulation with 2,4-
D treatment leading to an increase in ethylene accumulation
Frontiers in Plant Science | www.frontiersin.org 8March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
TABLE 2 | AGP and FLA genes that show transcript response to AGL15 and are direct targets as determined by ChIP-chip or ChIP-seq.
AGI TAIR10 ChIP-chip ChIP-seq agl15agl18/Col 35S:AGL15/Col
fold-change P-value fold-change P-value
At1g03870 FLA9 Y Y 2.26 0.03 0.41 0.16
At1g35230 AGP5 Y Y 2.60 0.00 1.01 0.96
At2g04780 FLA7 Y 2.12 0.01 0.89 0.58
At2g14890 AGP9 Y Y 1.82 0.01 0.81 0.50
At2g23130 AGP17 Y Y 2.62 0.01 0.50 0.26
At2g45470 FLA8 Y Y 2.56 0.01 0.57 0.29
At2g46330 AGP16 Y 1.79 0.00 1.13 0.26
At3g06360 AGP27 Y 0.82 0.28 0.65 0.05
At3g52370 FLA15 Y Y 0.46 0.00 0.49 0.01
At3g60900 FLA10 Y 2.20 0.00 0.76 0.44
At4g09030 AGP10 Y 1.98 0.01 0.75 0.42
At4g12730 FLA2 Y 2.05 0.02 0.39 0.10
At4g37450 AGP18 Y Y 1.66 0.02 0.41 0.06
At5g11740 AGP15 Y Y 1.60 0.01 1.06 0.73
At5g56540 AGP14 Y 0.55 0.00 1.28 0.13
At5g64310 AGP1 Y 2.06 0.00 1.15 0.68
Chip-chip is from Zheng et al. (2009), but allowing transcript change of 0.67–1.5. ChIP-seq is from Paul et al. (2021), analyzed by CLC Genomics Workbench, but also including the
three sets of data analyzed by CisGenome (Ji et al., 2008). The transcriptomic data is from Zheng et al. (2009).
(Raghavan et al., 2006; Stepanova et al., 2008; Vandenbussche
et al., 2012). Conversely, TRYPTOPHAN AMINOTRANSFERASE
RELATED2 (TAR2), that encodes an αclass of pyridoxal-50-
phosphate (PLP) dependent enzymes (Stepanova et al., 2008),
was a potential direct target of AGL15 and is expressed in
response to AGL15 (Zheng et al., 2009). TAR2 is involved in
ethylene response and encodes an enzyme involved in auxin
biosynthesis (Ma et al., 2014).
In soybean, GmAGL15 was found to increase transcript
accumulation from several ethylene biosynthetic genes
and 35S:GmAGL15 tissue has significantly higher ethylene
accumulation than the control. Genes involved in ethylene
signaling were often regulated in a consistent manner in
response to AGL15 in soybean and Arabidopsis (Zheng et al.,
2013a). For example, genes encoding potential orthologs of
MtSERF1 in Arabidopsis (one gene; At5g61590) and soybean
(two genes; Glyma20g16920 and Glyma10g24360), were found
to be directly upregulated targets of (Gm)AGL15. MtSERF1 is
aMedicago truncatula ethylene response factor subfamily B-3
of APETALA2/ETHYLENE RESPONSE FACTOR (AP2/ERF)
transcription factors that was shown to be required for
somatic embryogenesis. This gene is induced by ethylene and
responds to auxin and cytokinin (Mantiri et al., 2008). The
Arabidopsis ortholog of MtSERF1 was further demonstrated
to be relevant for SAM SE (Zheng et al., 2013a). These
studies allowed non-transgenic approaches to facilitate SE
in both Arabidopsis and soybean by manipulating ethylene
biosynthesis and perception. A precursor to ethylene, 1-
aminocyclopropane-1-carboxylic-acid (ACC), enhanced SE
in both species, whereas inhibitors of ethylene synthesis or
perception decreased SE.
Depending on the context, the ethylene-GA crosstalk
may be synergistic or antagonistic (Weiss and Ori, 2007).
Interestingly, transcript accumulation of both the soybean
and Arabidopsis SERF1 showed inverse correlation with GA,
supporting an antagonistic interaction in SE. In addition,
antagonistic interaction may also act by via stabilizing the
GA repressor DELLA proteins by ethylene signaling via EIN3
and CTR1 (Zheng et al., 2013a). Finally, Saptari and Susila
(2019) demonstrated that AGL15, FUS3, and BBM show
increased transcript accumulation in response to ethylene
via EIN3, whereas Zheng et al. (2013a), showed a positive
correlation between ethylene and GmAGL15, GmAGL18,
GmFUS3, and GmABI3 in soybean. Conversely, these genes
showed a general negative correlation with biologically active GA
(Zheng et al., 2016).
AGL15 and Brassinosteroids
While Brassinosteroids (BR) have been reported to enhance
SE (at least in combination with cytokinin: CK—Chone et al.,
2018), a recent publication reports on direct interactions between
AGL15, BR signaling, and SE (Ruan et al., 2021). This paper
documents experiments showing that BR promotes the embryo
to seedling transition, at least in part through repression of
AGL15 expression. Two positive regulators of BR signaling,
BRI1-EMS-SUPPRESSOR 1 (BES1), and BRASSINAZOLE-
RESISTANT 1 (BZR1), were found to bind directly to AGL15’s
promoter region. Furthermore, inhibition of BR biosynthesis
chemically or through mutants can promote SE development on
seedlings of the single mutant val1 that normally only shows
this phenotype in combination with val2 mutants (discussed
below), whereas dominant forms of bes1-D and bzr1-1D can
Frontiers in Plant Science | www.frontiersin.org 9March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
rescue normal seedling development. AGL15 is involved in the
production of the SEs as shown by double and triple mutants with
an agl15 loss-of-function mutant. Ruan et al. (2021) hypothesized
that BES1 and BZR1 may also directly regulate GA biosynthetic
genes, increasing bioactive GA levels and thereby removing
negative interaction of BES1/BZR1 by the DELLA proteins.
Interestingly, Ruan et al. (2021) also show that wild
type seedlings treated with BR biosynthesis inhibitors have
darker green, downward curling leaves that look similar
to seedings overexpressing AGL15 (Fernandez et al., 2000).
AGL15 binds to the promoter regions of BZR1 and BES1
(Supplementary Figure 1) and at least for BZR1, represses
transcript accumulation. Thus, there may be an intriguing
feedback loop between BR signaling, AGL15, and GA to control
the transition from seed to seedling.
AGL15 DIRECTLY BINDS GENES
ENCODING MICRO-RNAS
MicroRNA (miRNA) function is critical during embryogenesis
through various cellular, physiological, and developmental
processes. MiRNAs target transcripts from genes that include
those with a regulatory role such as TFs that can mediate stress
adaptation, developmental regulation, and hormone response
(Jones-Rhoades et al., 2006; Jin et al., 2020). MiRNAs are
single-stranded RNA molecules of 21– 24 nucleotides that post-
transcriptionally regulate gene expression (Bartel, 2009; Rubio-
Somoza and Weigel, 2011). The mechanism of miRNA action
is either by targeting mRNA for cleavage or by inhibiting
translation (Yu et al., 2017). RNA cleavage is mediated by
recognizing the nearly perfect complementary sites (Llave et al.,
2002; Tang et al., 2003) and involves numerous proteins that are
associated with the RNA-induced silencing complex (RISC).
Briefly, MIR genes encoding microRNAs are transcribed
by RNA pol II forming primary miRNAs (pri-miRNAs) with
a characteristic hairpin structure that are processed by a
complex that includes DICER-LIKE1 (DCL1) to generate
a pre-miRNA. This pre-miRNA duplex structure is further
processed, producing the mature single stranded miRNA, which
is incorporated by ARGONAUTE1 (AGO1) to form an active
RISC (Baumberger and Baulcombe, 2005). This complex is
involved in the identification of the target transcripts that are
complementary to the miRNA sequence (Rhoades et al., 2002;
Kidner and Martienssen, 2005).
There are excellent resources on the prominence of
microRNAs, and their functions on plant development (Bartel,
2009; Rubio-Somoza and Weigel, 2011; Wójcikowska et al., 2020;
Gyawali et al., 2021; Ma et al., 2021). Several studies highlight
the crucial role of miRNA in plant development, especially
focusing on embryogenesis, including SE (Armenta-Medina
et al., 2017; Liu et al., 2018; Song et al., 2019; Alves et al., 2020;
Wójcikowska et al., 2020). Here, we emphasize how AGL15, in
addition to targeting protein-encoding genes, also regulates MIR
genes. Some miRNAs repress key embryo identity genes LEC2
and FUS3 that encode products involved in embryo maturation
(Willmann et al., 2011). Other miRNAs in Arabidopsis play
roles in cell-specific gene expression programs in response
to spatial and temporal signals, to prevent precocious gene
expression, and to enable pattern formation (Nodine and Bartel,
2010; Plotnikova et al., 2019). In the early eight-cell stage of
zygotic embryogenesis, DCL1 is essential for cell differentiation.
Similarly, dcl mutants in somatic explants are unable to initiate
embryogenic induction in vitro (Wójcik and Gaj, 2016). Another
large-scale study showed the dynamics and functions of miRNA
during Arabidopsis embryogenesis (Plotnikova et al., 2019).
The group applied a high-throughput sequencing technique
to profile hundreds of miRNAs and their targets throughout
embryogenesis. The findings highlight that miRNAs dynamically
cleave and repress at least 59 transcripts, including 30 encoding
transcription factors belonging to eight different families
(Plotnikova et al., 2019). Other studies demonstrated that
mutant lines of genes encoding miR160, miR170/171, or miR319
result in incorrect divisions in the embryo and have abnormal
cotyledon development, suggesting that the corresponding
miRNA activities are required for embryo morphogenesis
(Palatnik et al., 2003; Mallory et al., 2005; Liu et al., 2010;
Takanashi et al., 2018).
Szyrajew et al. (2017), examined the expression of 190
miRNA genes during Arabidopsis somatic embryogenesis and
found that 98% of the MIR genes were active during SE,
with 64% being differentially expressed during SE. Those
differentially expressed during SE included miR156, miR157,
miR159, miR160, miR164, miR166, miR169, miR319, miR390,
miR393, miR396, and miR398. These miRNAs are well known
to have functions associated with phytohormone and stress-
related response (Szyrajew et al., 2017). One example,miR393,
regulates the expression of auxin receptors during early SE.
The authors suggest that increased miR393, resulting in reduced
auxin sensitivity by post-transcriptional regulation of production
of auxin receptors, is involved in the transition of somatic cells to
embryogenic fate in Arabidopsis (Wójcikowska and Gaj, 2016).
Recently, Nowak et al. (2020) reported how AGL15 negatively
regulates the expression of miRNA biogenesis genes through
histone acetylation to control the embryogenic reprogramming
of somatic cells in Arabidopsis. While AGL15 accumulation
showed a positive correlation with transcript accumulation
of pri-miR156h, the mature miR156h accumulation did not
agree with the pri-miRNA and in fact the agl15agl18 mutant
showed increased mature miR156h (and miR156 from other
family members) compared to 35S:AGL15. This indicated a
disconnect between production of the pri- and mature miRNA
suggesting regulation of miRNA processing components. They
found a negative correlation between AGL15 accumulation
and transcript accumulation from the miRNA biogenesis genes
DCL1,SERRATE, and HUA-ENHANCER1 (HEN1). While they
did not perform ChIP for AGL15 for these genes, DCL1 had been
identified as a direct target of AGL15 (Zheng et al., 2009) and
they did document histone acetylation states in the regulatory
regions of these genes, demonstrating increased acetylation in the
agl15agl18 mutant and decreased acetylation in the 35S:AGL15
tissue compared to control for DCL1 and SERRATE. This would
be consistent with AGL15 recruitment of histone deacetylase
complexes to repress gene expression (discussed further below)
Frontiers in Plant Science | www.frontiersin.org 10 March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
(Hill et al., 2008). Neither SERRATE nor HEN1 were found as
targets in our ChIP studies (Zheng et al., 2009; Paul et al., 2021).
Our results found that AGL15 directly regulates miR167a,
with subsequent impacts on targets ARF6 and ARF8, with
consequences for SAM SE. These results are discussed under
the section on auxin. Here we present the set of miRNA
encoding genes with regulatory regions associated with AGL15
(Table 3). All of the miRNAs mentioned in Szyrajew et al. (2017)
have members bound by AGL15. Because the miRNAs were
not represented on the microarray, we looked at response of
targets of these miRNAs as predicted by TarDB (http://www.
biosequencing.cn/TarDB/guide/guide.html) (Liu et al., 2021).
Response of the targets and annotation is provided in
Supplementary Table 2.
AGL15 AND POTENTIAL LINKS TO
EPIGENETIC REGULATION OF SOMATIC
EMBRYOGENESIS
Epigenetic modifications are crucial in development processes.
The structure of chromatin can be accessible or repressive
due to various epigenetic mechanisms like DNA methylation,
histone modifications, and chromatin remodeling. These
epigenetic mechanisms are dynamic; involved in different
plant developmental phases, and in plant adaptation in various
environmental stresses. Epigenetic mechanisms could reprogram
cell fate and could stabilize cell identity and maintain tissues
(Inácio et al., 2018; Hajheidari et al., 2019). In vitro culture
conditions produce epigenetic variation and are critical factors
during SE (De-La-Pena et al., 2015; Kumar and Van Staden,
2017). Many of the loss-of-function mutants in Arabidopsis that
produce SE on “seedlings” are mutants defective in epigenetic
processes to support the transition from embryo to seedling
(reviewed in Tian et al., 2020a). Key embryo-related TFs are
inappropriately expressed in these mutants.
DNA Methylation
One well-studied epigenetic modification is DNA methylation
and this has been found to control gene expression essential
during early somatic embryogenesis (Stricker et al., 2017;
Wójcikowska et al., 2020). DNA methylation involves the
addition of a methyl group to the 5position of the pyrimidine
ring of cytosine. In plants, methylation can occur in CpG islands
(Gruenbaum et al., 1981); CpHpG and CpHpHp (where H is
any nucleotide except G; (Feng et al., 2010). DNA methylation
differs at each stage of SE (De-La-Pena et al., 2015). Several
reports have indicated that embryogenic cultures/explants and
DNA methylation have an inverse relationship. For example,
early somatic cells of leaf explants in Coffea canephora show a
level of methylated cytosines of about 23.7%, but later stages have
higher levels of DNA methylation (Nic-Can et al., 2013). The
dedifferentiating Arabidopsis leaf protoplast has been reported
to have changes in DNA methylation (Avivi et al., 2004).
Auxin’s role in inducing epigenetic regulation of SE has been
recently reviewed (Wójcik et al., 2020). A recent study shows
hypermethylation in the promoters of TFs that are associated
with SE in response to auxin treatment (Grzybkowska et al.,
2020). They confirmed exogenous auxin substantially affected
both the expression and methylation patterns, resulting in
control of embryo-related targets LEC1, LEC2, BBM, WUS, and
AGL15, although AGL15 did not have auxin response elements
in the examined region, but did have ethylene response elements
(Grzybkowska et al., 2020). They propose that the methylation
sensitive ARF, ARF5/MP may be involved in this process.
Histone Modifications
Apart from DNA methylation, histone modifications have a
dynamic function that causes variation of gene expression that
is involved in SE. For example, the histone trimethylation mark
H3K27me3, generally a repressive mark, is removed from LEC1
loci, allowing the expression of this TF, and the expression
of BBM1 was related to the increase of both histone marks
H3K4me3 and H3K36me2 (generally marks associated with
increased transcription) using chromatin immunoprecipitation
(ChIP) assays (Nic-Can et al., 2013; Borg et al., 2020). Another
technique called immunolocalization was used in Brassica napus
to evaluate the levels of H3K9me2 that was low in microspores
before the induction of SE; however it showed increases during
later stages of SE. In contrast to methylation, it was observed that
the levels of acetylation of H3 and H4 (H3Ac and H4Ac) were
more abundant in microspores before SE induction, suggesting
it might increase transcriptional activity and regulate cellular
reprogramming and embryo development (Rodríguez-Sanz et al.,
2014). Similarly, in C. canephora, a decrease in the DNA
methylation level has been related to a decrease of H3K9me2 and
H3K27me3 during SE (Nic-Can et al., 2013).
One of the interesting groups of proteins related to histone
methylation marks that are involved in regulating plant cellular
reprogramming are the Polycomb-group (PcG). (reviewed in
Duarte-Aké et al., 2019). The PcG proteins are classified into
two complexes, POLYCOMB REPRESSIVE COMPLEX 1 and 2
(PRC1 and PRC2), which are involved in repression of genes
via histone methylation (Grossniklaus and Paro, 2014; Mozgová
et al., 2015). PRC2- mediated histone methylation represses
embryo maturation programs during vegetative development in
Arabidopsis, where cell fate could reset and SE could occur when
PRC2 depleted tissues are treated with hormone (Mozgová et al.,
2017). LEC1, LEC2, FUS3, AGL15, PLT, and WOXs, which have
regulatory roles in somatic cell differentiation and SE induction
are shown to be targeted by PRC2 in different plants (Ikeuchi
et al., 2015; Orłowska et al., 2017; Rose, 2019).
A recent study found that HSI2/VAL1 silences AGL15
to regulate the development transition from seed maturation
to vegetative phase in Arabidopsis (Chen et al., 2018).
VIVIPAROUS1/ABI3-LIKE (VAL1) and VAL2, are also known as
HIGH-LEVEL EXPRESSION OF SUGAR INDUCIBLE GENE2
(HSI2) and HSI2-LIKE1 (HSL1), respectively. VAL genes encode
B3 domain transcription factors similar to LEC2, FUS3, and
ABI3 (Suzuki et al., 2007), but also contain other conserved
domains including a plant homeodomain (PHD), a cysteine-and-
tryptophan residue containing (CW) domain, and an EAR motif
that is involved in recruiting histone deacetylation complexes.
These domains are involved in epigenetic regulation, and
Frontiers in Plant Science | www.frontiersin.org 11 March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
TABLE 3 | miRNAs that have associated regions bound by AGL15 (based on ChIP-chip and/or ChIP-seq). Included are predicted targets from TarDB (Liu et al., 2021)
and notations of involvement in SE and/or hormone function.
miRNA associated with
AGL15
Predicted targets of the miRNA that show significant response to AGL15 accumulation in
agl15 agl18 mutants and/or 35S:AGL15 10 d SAM SE (from TarDB: http://www.
biosequencing.cn/TarDB/)
Include targets involved in
hormone response; miRNA
involved in SE (Szyrajew
et al., 2017; Jin et al., 2020)
miR156 (a,c,d,e) AT1G11190,AT1G11840,AT1G48760,AT1G63080,AT2G27050,AT3G05430,
AT3G10910,AT3G17330,AT4G34990,AT4G35620,AT4G36650,AT1G71400,
AT2G34190,AT3G22160,AT5G43270,AT5G43460,AT5G50570,AT5G63850,
AT4G30790
Ethylene;Yes
miR157 (a,b,) AT3G22790,AT5G50570,AT1G30460,AT2G13290,AT4G16860,AT5G50570 Ethylene; Yes
miR158a AT1G17145,AT1G62930,AT1G65810,AT2G22300,AT2G46590,AT3G01040,
AT3G09850,AT3G12010,AT3G23690,AT4G30690, AT4G37740,AT5G35670,
AT5G38840,AT5G40340,AT5G43460,AT5G52060,AT5G53220
miR159 (a,b,) AT1G08030,AT1G21060,AT1G51780,AT3G44690,AT3G53810,AT4G12240,
AT1G04700,AT1G08030,AT1G08550,AT1G30210,AT1G35460,AT1G63130,
AT1G78610,AT3G10300,AT3G12550,AT3G28150,AT3G53160,AT4G23950,
AT5G04460, AT5G06100,AT5G09670,AT5G12100
ABA; Yes
miR160a AT1G52620,AT1G79180,AT4G18930,AT4G20430,AT5G43270 Auxin; Yes
miR164b AT1G76580,AT2G26690,AT4G28550,AT5G07680,AT5G62230 Yes
miR165a AT1G47530,AT1G58200,AT4G39150,AT5G50570,AT5G67570
miR166 (a,b,c,d,e,g) AT1G30490,AT1G61210,AT5G01030,AT1G30490,AT3G16785,AT5G01030,
AT1G30490, AT1G30490,AT1G30490,AT5G35670,AT5G65970,AT1G30490
Auxin; Yes
miR167a AT2G34530,AT3G59470,AT4G30130,AT5G02940,AT5G19950,AT5G49990 Auxin
miR168 (a,b) AT1G17760,AT2G02470,AT3G22160,AT1G48410,AT2G23130
miR169 (a,b,c,f) AT1G72830,AT2G01530,AT2G30600,AT3G27700,AT4G27710,AT4G36710,
AT5G12470,AT1G72830,AT5G38840, AT1G72830,AT5G38840,AT1G72830,
AT3G15950,AT3G61250,AT4G18890,AT5G62680
Yes
miR170 AT4G17370,AT5G16390,AT5G65970
miR172a AT1G12270,AT1G28190,AT1G71400,AT2G34190,AT2G47460,AT3G07060,
AT3G23330,AT4G16860,AT4G36650, AT5G02710,AT5G16260,AT5G17760,
AT5G63850
miR173 AT3G58780
miR319 (a,c) AT1G30210,AT1G04700,AT1G53230,AT3G06500 Yes
miR390 (a,b) AT3G04110,AT3G13430,AT5G05180,AT5G62230,AT1G55610,AT2G02470,
AT2G33730,AT5G62230
Auxin; Yes
miR393a AT1G15290,AT1G48410,AT2G46340,AT3G54280,AT4G26120 Auxin; Yes
miR394 (a,b) AT1G06580,AT1G19630,AT1G62910,AT3G07770,AT3G15840,AT3G48460,
AT3G48460,AT3G49240,AT4G20430,AT5G51100,AT5G56950,AT3G48460,
AT4G20430,AT4G20430,AT5G07870,AT5G09670,AT5G51100,AT5G56950
miR396b AT1G12820AT1G28010AT1G30210AT1G30490AT1G52500AT1G53AT1G55610,
AT1G60140,AT1G61370,AT1G62590,AT1G74790,AT1G76810,AT2G23130,
AT2G33730,AT2G40840,AT3G01370,AT3G03790,AT3G16730,AT3G16785,
AT3G17090,AT3G24310,AT3G54350,AT3G59420,AT4G19050,AT4G19440,
AT4G20440,AT4G30080,AT4G30130,AT4G32710,AT5G01030,AT5G12900,
AT5G39610,AT5G42830,AT5G46110,AT5G61440
Yes
miR398c AT1G30460, AT5G19950
miR399 (a,c) AT3G60040,AT3G62980,AT5G42880,AT1G71350,AT2G19470,AT2G33730,
AT2G33770,AT2G33770,AT3G06500,AT3G15130,AT3G59420,AT5G51220,
AT5G60890
miR400 AT1G07590,AT1G34670,AT1G48780,AT1G76870,AT2G13290,AT2G19470,
AT2G28550,AT2G43970,AT4G26680,AT5G14550,AT5G39710,AT5G41400,
AT5G53950,AT5G63490
mir403 AT4G31460,AT5G39900
miR858a AT1G50700,AT1G53230,AT1G65540,AT1G79640,AT3G10120,AT3G59470,
AT3G61570,AT5G46560,AT5G59900,AT5G66420
Frontiers in Plant Science | www.frontiersin.org 12 March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
HSI2/VAL1 recruits PRC2 components. During the transition
to seedling development, HSI2/VAL1 is required for the
transcriptional silencing of LAFL and AGL15 genes, but further
work demonstrated that while this protein interacts directly with
regulatory regions of AGL15, no such interaction was found for
the LAFL genes (Chen et al., 2018). While a binding site for B3
domains are involved in silencing AGL15, a mutation specifically
within the PHD domain interferes with H3K27me3 at AGL15 and
also prevents silencing, and this involved recruitment of PRC2
components (Veerappan et al., 2014; Chen et al., 2018).
Another mode of histone modification is acetylation where
lysine residues on the N-terminal tails of histones undergo
acetylation, which causes increased DNA accessibility for
TF binding. Histone acetyltransferases (HATs), and histone
deacetylases (HDACs) are the complexes that control the
acetylation state (Steunou et al., 2014; Lee and Grant, 2019).
HDAC6 and HDAC19 are two important deacetylases that are
involved in SE where a double knockdown generates SE on
seedlings. AGL15 has been found to recruit HDA19 via interact
with TOPLESS (TPL) and TOPLESS-RELATED PROTEIN2
(TPR2) (Causier et al., 2012), and SIN3 ASSOCIATED
POLYPEPTIDE P18 (SAP18; Hill et al., 2008).
DISCUSSION AND FUTURE DIRECTIONS
While AGL15 has a role in promoting SE, it is not to the
extent as found for some TFs, such as LEC1, LEC2, and
BBM, possibly indicating needed protein complex formation
with other proteins. However, the interactions between AGL15
and these other factors is fascinating, as is the fact that other
factors directly control AGL15 to facilitate the transition from
embryo to seedling. Many questions obviously remain. Why
does AGL15 appear to limit auxin responses and possibly
be involved in negative regulation of AGP signaling? These
observations may point to different roles of cells involved
in producing SEs. Whereas, some cells produce the somatic
embryos, others in the culture are equally important to provide
what has been referred to as “nursing” functions. Current
technologies allow marking and sorting of cells. An intriguing
avenue of investigation would be characterizing AGL15 function
in the cells that make SE compared to the support tissues. The
observations on GmEP3 at different timepoints of soybean SE
culture emphasize the need for examination of more timepoints
during the SE process. It is quite possible that genes responding
to AGL15 in 10 d SAM SE would behave very differently at
earlier timepoints. Other hormones are important for SE that
have not been examined for relation to AGL15 function. These
include BR, CK and ABA. Notably, genes in these GO categories
are overrepresented among AGL15 targets (Figure 5). Finally,
comparison of gene regulation in developmental processes in
different contexts will be important to tease out commonalities
and determine other context dependent factors. As an example,
Nowak et al. (2020) documented an inverse correlation between
AGL15 and transcripts from miRNA biogenesis genes, indicating
repression, direct or indirect of these genes by AGL15. However,
data from Zheng et al. (2009), showed decreased transcript
accumulation for DCL1, HEN1, and SERRATE in the agl15agl18
mutant (non-significant in the 35S:AGL15 compared to control),
suggesting expression of these genes. Zheng et al. (2009) used
an SAM SE system, whereas Nowak et al. (2020) used immature
zygotic embryos as explants, perhaps explaining these and
other differences.
Molecular breeding has become a significant tool to improve
crop production in a shorter time period, and efficient somatic
embryogenesis is a valuable technique in regenerating such crops.
Advances in gene editing techniques like CRISPR/Cas9 created
the potential to improve crop traits without regulatory concerns
that apply to traditional transgenics (once the Cas9 and guide
RNAs are segregated, at least in some countries; Grossman, 2019;
Eriksson et al., 2019). However, cells engineered for particular
traits must be able to exhibit pluripotency and produce a plant
to be useful. As yet, it is not well understood why some plants
regenerate well, poorly or not at all via either SE or organogenesis.
Even particular cultivars of a species can be recalcitrant to these
processes. We and others have found that AGL15 is a tool to
enhance regeneration via SE, at least in dicots. For example, plant
transformation in soybean is accelerated when overexpression of
AGL15 enhances somatic embryogenesis, which could be utilized
for producing elite cultivars. Similarly, AGL15 could provide a
vital role in scaling up production of valuable crops such as coffee
through propagation and genetic transformation (Etienne et al.,
2018; Xu et al., 2018). Other crops could benefit from examining
AGL15 expression in their species, and possibly enhancing
transformation strategies via better regeneration by SE.
DATA AVAILABILITY STATEMENT
The original contributions presented in the study are included
in the article/Supplementary Material, further inquiries can be
directed to the corresponding author/s.
AUTHOR CONTRIBUTIONS
All authors contributed to writing the manuscript. All authors
contributed to the article and approved the submitted version.
FUNDING
This work was supported by the National Science Foundation
(grant no. IOS-1656380 to SP) and by the National Institute of
Food and Agriculture, U.S. Department of Agriculture, Hatch
project (SP) under accession number 1013409.
ACKNOWLEDGMENTS
We apologize to all of our colleagues who have generated
excellent data that we could not discuss or cite due to
space constraints.
SUPPLEMENTARY MATERIAL
The Supplementary Material for this article can be found
online at: https://www.frontiersin.org/articles/10.3389/fpls.2022.
861556/full#supplementary-material
Frontiers in Plant Science | www.frontiersin.org 13 March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
REFERENCES
Alves, A., Rodrigues, A. S., and Miguel, C. (2020). “microRNAs in plant
embryogenesis, in Plant microRNAs. Concepts and Strategies in Plant Sciences,
eds C. Miguel, T. Dalmay, and I. Chaves (Cham: Springer), 99–120.
doi: 10.1007/978-3-030-35772-6_6
Armenta-Medina, A., Lepe-Soltero, D., Xiang, D., Datla, R., Abreu-Goodger,
C., and Gillmor, C. S. (2017). Arabidopsis thaliana miRNAs promote
embryo pattern formation beginning in the zygote. Dev. Biol. 431, 145–151.
doi: 10.1016/j.ydbio.2017.09.009
Avivi, Y., Morad, V., Ben-Meir, H., Zhao, J., Kashkush, K., Tzfira, T., et al.
(2004). Reorganization of specific chromosomal domains and activation of
silent genes in plant cells acquiring pluripotentiality. Dev. Dyn. 230, 12–22.
doi: 10.1002/dvdy.20006
Awasthi, P., Sharma, V., Kaur, N., Kaur, N., Pandey, P., and Tiwari, S. (2017).
Genome-wide analysis of transcription factors during somatic embryogenesis
in banana (Musa spp.) cv. Grand Naine. PLoS ONE 12, e0182242.
doi: 10.1371/journal.pone.0182242
Bartel, D. P. (2009). MicroRNAs: target recognition and regulatory functions. Cell
136, 215–233. doi: 10.1016/j.cell.2009.01.002
Baumberger, N., and Baulcombe, D. (2005). Arabidopsis ARGONAUTE1 is an
RNA slicer that selectively recruits microRNAs and short interfering RNAs.
Proc. Natl. Acad. Sci. U. S. A. 102, 11928–11933. doi: 10.1073/pnas.05054
61102
Bidabadi, S. S., and Jain, S. M. (2020). Cellular, molecular, and physiological
aspects of in vitro plant regeneration. Plants-Basel 9, 702. doi: 10.3390/plants
9060702
Borg, M., Jacob, Y., Susaki, D., Leblanc, C., Buendía, D., Axelsson, E.,
et al. (2020). Targeted reprogramming of H3K27me3 resets epigenetic
memory in plant paternal chromatin. Nature Cell Biol. 22, 621–629.
doi: 10.1038/s41556-020-0515-y
Boutilier, K., Offringa, R., Sharma, V. K., Kieft, H., Ouellet, T., Zhang, L.,
et al. (2002). Ectopic expression of BABY BOOM triggers a conversion
from vegetative to embryonic growth. Plant Cell 14, 1737–1749.
doi: 10.1105/tpc.001941
Braybrook, S. A., and Harada, J. J. (2008). LECs go crazy in embryo development.
Trends Plant Sci. 13, 624–630. doi: 10.1016/j.tplants.2008.09.008
Busch, W., Miotk, A., Ariel, F. D., Zhao, Z., Forner, J., Daum, G., et al. (2010).
Transcriptional control of a plant stem cell niche. Dev. Cell 18, 849–861.
doi: 10.1016/j.devcel.2010.03.012
Causier, B., Ashworth, M., Guo, W., and Davies, B. (2012). The TOPLESS
interactome: a framework for gene repression in Arabidopsis. Plant Physiol.
158, 423–438. doi: 10.1104/pp.111.186999
Chen, F., Zhang, X. T., Liu, X., and Zhang, L. S. (2017). Evolutionary analysis of
MIKCc-type MADS-box genes in gymnosperms and angiosperms. Front. Plant
Sci. 8, 895. doi: 10.3389/fpls.2017.00895
Chen, N., Veerappan, V., Abdelmageed, H., Kang, M., and Allen, R. D. (2018).
HSI2/VAL1 silences AGL15 to regulate the developmental transition from
seed maturation to vegetative growth in Arabidopsis. Plant Cell 30, 600–619.
doi: 10.1105/tpc.17.00655
Chone, R. M. S., Rocha, D. I., Monte-Bello, C. C., Pinheiro, H. P., Dornelas,
M. C., Haddad, C. R. B., et al. (2018). Brassinosteroid increases the
cytokinin efficiency to induce direct somatic embryogenesis in leaf explants
of Coffea arabica L. (Rubiaceae). Plant Cell Tissue Organ Cult. 135, 63–71.
doi: 10.1007/s11240-018-1443-4
Clough, S. J., and Bent, A. F. (1998). Floral dip: a simplified method for
Agrobacterium-mediated transformation of Arabidopsis thaliana.Plant J. 16,
735–743. doi: 10.1046/j.1365-313x.1998.00343.x
De-La-Pena, C., Nic-Can, G. I., Galaz-Avalos, R. M., Avilez-Montalvo, R., and
Loyola-Vargas, V. M. (2015). The role of chromatin modifications in somatic
embryogenesis in plants. Front. Plant Sci. 6, 635. doi: 10.3389/fpls.2015.
00635
Domon, J. M., Neutelings, G., Roger, D., David, A., and David, H. (2000). A basic
chitinase-like protein secreted by embryogenic tissues of Pinus caribaea acts
on arabinogalactan proteins extracted from the same cell lines. J. Plant Physiol.
156, 33–39. doi: 10.1016/S0176-1617(00)80269-2
Duarte-Aké, F., Nic-Can, G., and De-La-Peña, C. (2019). “Somatic embryogenesis:
polycomb complexes control cell-to-embryo transition, in Epigenetics in Plants
of Agronomic Importance: Fundamentals and Applications, eds R. Alvarez-
Venegas, C. De-la-Pena, and J. Casas-Mollano (Cham: Springer), 339–354.
doi: 10.1007/978-3-030-14760-0_13
Emons, M. C. (1994). Somatic embryogenesis: cell biological aspects. Acta Botanica
Neerlandica 43, 1–14. doi: 10.1111/j.1438-8677.1994.tb00729.x
Eriksson, D., Kershen, D., Nepomuceno, A., Pogson, B. J., Prieto, H.,
Purnhagen, K., et al. (2019). A comparison of the EU regulatory approach
to directed mutagenesis with that of other jurisdictions, consequences for
international trade and potential steps forward. New Phytol. 222, 1673–1684.
doi: 10.1111/nph.15627
Etienne, H., Breton, D., Breitler, J. C., Bertrand, B., Dechamp, E., Awada, R., et al.
(2018). Coffee somatic embryogenesis: how did research, experience gained and
innovations promote the commercial propagation of elite clones from the two
cultivated species? FIPS 9, 1630. doi: 10.3389/fpls.2018.01630
Fehér, A. (2015). Somatic embryogenesis—stress-induced remodeling of plant cell
fate. BBA-Gene Reg. Mech. 1849, 385–402. doi: 10.1016/j.bbagrm.2014.07.005
Fehér, A. (2019). Callus, dedifferentiation, totipotency, somatic embryogenesis:
what these terms mean in the era of molecular plant biology? Front. Plant Sci.
10, 536. doi: 10.3389/fpls.2019.00536
Feng, S., Jacobsen, S. E., and Reik, W. (2010). Epigenetic reprogramming in plant
and animal development. Science 330, 622–627. doi: 10.1126/science.1190614
Fernandez, D. E., Heck, G. R., Perry, S. E., Patterson, S. E., Bleecker, A. B., and Fang,
S.-C. (2000). The embryo MADS domain factor AGL15 acts postembryonically:
inhibition of perianth senescence and abscission via constitutive expression.
Plant Cell 12, 183–197. doi: 10.1105/tpc.12.2.183
Gaj, M. D., Trojanowska, A., Ujczak, A., Medrek, M., Koziol, A., and
Garbaciak, B. (2006). Hormone-response mutants of Arabidopsis thaliana (L.)
Heynh. impaired in somatic embryogenesis. Plant Growth Reg. 49, 183–197.
doi: 10.1007/s10725-006-9104-8
Ghadirzadeh-Khorzoghi, E., Jahanbakhshian-Davaran, Z., and Seyedi, S. M.
(2019). Direct somatic embryogenesis of drought resistance pistachio (Pistacia
vera L.) and expression analysis of somatic embryogenesis-related genes. South
Afr. J. Bot. 121, 558–567. doi: 10.1016/j.sajb.2019.01.023
Goodstein, D. M., Shu, S., Howson, R., Neupane, R., Hayes, R. D., Fazo, J., et al.
(2012). Phytozome: a comparative platform for green plant genomics. Nucl.
Acids Res. 40, D1178–D1186. doi: 10.1093/nar/gkr944
Grafi, G. (2009). The complexity of cellular dedifferentiation:
implications for regenerative medicine. Trends Biotech. 27, 329–332.
doi: 10.1016/j.tibtech.2009.02.007
Grafi, G., Chalifa-Caspi, V., Nagar, T., Plaschkes, I., Barak, S., and Ransbotyn, V.
(2011). Plant response to stress meets dedifferentiation. Planta 233, 433–438.
doi: 10.1007/s00425-011-1366-3
Gramzow, L., and Theissen, G. (2010). A hitchhiker’s guide to the MADS world of
plants. Genome Biol. 11, 214–214. doi: 10.1186/gb-2010-11-6-214
Grossman, M. R. (2019). “New plant breeding technologies: US Department
of Agriculture policy, in Functional Field of Food Law: Reconciling the
Market and Human Rights, eds A. Urazbaeva, A. Szajkowska, B. Wernaart,
N. T. Franssens, and R. S. Vaskoska (Wageningen: Wageningen Academic
Publications), 397–416. doi: 10.3920/978-90-8686-885-8_26
Grossniklaus, U., and Paro, R. (2014). Transcriptional silencing by
polycomb-group proteins. Cold Spring Harb. Perspect. Biol. 6, a019331.
doi: 10.1101/cshperspect.a019331
Grover, A. (2012). Plant chitinases: genetic diversity and physiological roles. Critic.
Rev. Plant Sci. 31, 57–73. doi: 10.1080/07352689.2011.616043
Gruenbaum, Y., Naveh-Many, T., Cedar, H., and Razin, A. (1981). Sequence
specificity of methylation in higher plant DNA. Nature 292, 860–862.
doi: 10.1038/292860a0
Grzybkowska, D., Nowak, K., and Gaj, M. D. (2020). Hypermethylation of
auxin-responsive motifs in the promoters of the transcription factor genes
accompanies the somatic embryogenesis induction in Arabidopsis. Int. J. Mol.
Sci. 21, 6849. doi: 10.3390/ijms21186849
Guan, Y., Li, S.-G., Fan, X.-F., and Su, Z.-H. (2016). Application of
somatic embryogenesis in woody plants. Front. Plant Sci. 7, 938.
doi: 10.3389/fpls.2016.00938
Frontiers in Plant Science | www.frontiersin.org 14 March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
Guilfoyle, T. J., and Hagen, G. (2012). Getting a grasp on domain III/IV responsible
for auxin response factor–IAA protein interactions. Plant Sci. 190, 82–88.
doi: 10.1016/j.plantsci.2012.04.003
Gyawali, B., Barozai, M. Y. K., and Aziz, A. N. (2021). Comparative
expression analysis of microRNAs and their targets in emerging bio-
fuel crop sweet sorghum (Sorghum bicolor L.). Plant Gene 26, 100274.
doi: 10.1016/j.plgene.2021.100274
Hajheidari, M., Koncz, C., and Bucher, M. (2019). Chromatin evolution-key
innovations underpinning morphological complexity. Front. Plant Sci. 10, 454.
doi: 10.3389/fpls.2019.00454
Harding, E. W., Tang, W., Nichols, K. W., Fernandez, D. E., and Perry, S. E. (2003).
Expression and maintenance of embryogenic potential is enhanced through
constitutive expression of AGAMOUS-Like 15.Plant Physiol. 133, 653–663.
doi: 10.1104/pp.103.023499
Hayashi, K.-I. (2012). The interaction and integration of auxin signaling
components. Plant Cell Physiol. 53, 965–975. doi: 10.1093/pcp/pcs035
Heck, G. R., Perry, S. E., Nichols, K. W., and Fernandez, D. E. (1995). AGL15,
a MADS domain protein expressed in developing embryos. Plant Cell 7,
1271–1282. doi: 10.1105/tpc.7.8.1271
Hesami, M., Naderi, R., and Tohidfar, M. (2020). Introducing a hybrid artificial
intelligence method for high-throughput modeling and optimizing plant tissue
culture processes: the establishment of a new embryogenesis medium for
chrysanthemum, as a case study. Appl. Microbiol. Biotech. 104, 10249–10263.
doi: 10.1007/s00253-020-10978-1
Hill, K., Wang, H., and Perry, S. E. (2008). A transcriptional repression motif in the
MADS factor AGL15 is involved in recruitment of histone deacetylase complex
components. Plant J. 53, 172–185. doi: 10.1111/j.1365-313X.2007.03336.x
Horstman, A., Fukuoka, H., Muino, J. M., Nitsch, L., Guo, C. H., Passarinho,
P., et al. (2015). AIL and HDG proteins act antagonistically to control cell
proliferation. Development 142, 454–464. doi: 10.1242/dev.117168
Ikeda, M., and Kamada, H. (2005). “Comparison of molecular mechanisms of
somatic and zygotic embryogenesis, in Somatic Embryogenesis, eds A. Mujib
and J. Samaj (Berlin: Springer-Verlag), 51–68. doi: 10.1007/7089_027
Ikeuchi, M., Iwase, A., Rymen, B., Harashima, H., Shibata, M., Ohnuma, M., et al.
(2015). PRC2 represses dedifferentiation of mature somatic cells in Arabidopsis.
Nat. Plants 1, 1–7. doi: 10.1038/nplants.2015.89
Ikeuchi, M., Ogawa, Y., Iwase, A., and Sugimoto, K. (2016). Plant regeneration:
cellular origins and molecular mechanisms. Development 143, 1442–1451.
doi: 10.1242/dev.134668
Inácio, V., Martins, M. T., Graça, J., and Morais-Cecílio, L. (2018). Cork oak
young and traumatic periderms show PCD typical chromatin patterns but
different chromatin-modifying genes expression. Front. Plant Sci. 9, 1194.
doi: 10.3389/fpls.2018.01194
Jha, P., and Kumar, V. (2018). BABY BOOM (BBM): a candidate transcription
factor gene in plant biotechnology. Biotech. Lett. 40, 1467–1475.
doi: 10.1007/s10529-018-2613-5
Jha, P., Ochatt, S. J., and Kumar, V. (2020). WUSCHEL: a master regulator in plant
growth signaling. Plant Cell Rep. 39, 431–444. doi: 10.1007/s00299-020-02511-5
Ji, H. K., Jiang, H., Ma, W. X., Johnson, D. S., Myers, R. M., and Wong, W. H.
(2008). An integrated software system for analyzing ChIP-chip and ChIP-seq
data. Nat. Biotech. 26, 1293–1300. doi: 10.1038/nbt.1505
Jin, L., Yarra, R., Zhou, L., Zhao, Z., and Cao, H. (2020). miRNAs as
key regulators via targeting the phytohormone signaling pathways during
somatic embryogenesis of plants. 3 Biotech 10, 495. doi: 10.1007/s13205-020-
02487-9
Jones-Rhoades, M. W., Bartel, D. P., and Bartel, B. (2006). MicroRNAs
and their regulatory roles in plants. Annu. Rev. Plant Biol. 57, 19–53.
doi: 10.1146/annurev.arplant.57.032905.105218
Karlova, R., Boeren, S., Russinova, E., Aker, J., Vervoort, J., and de Vries, S. (2006).
The Arabidopsis SOMATIC EMBRYOGENESIS RECEPTOR-LIKE KINASE1
protein complex includes BRASSINOSTEROID-INSENSITIVE1. Plant Cell 18,
626–638. doi: 10.1105/tpc.105.039412
Kaul, S., Koo, H. L., Jenkins, J., Rizzo, M., Rooney, T., Tallon, L. J., et al. (2000).
Analysis of the genome sequence of the flowering plant Arabidopsis thaliana.
Nature 408, 796–815. doi: 10.1038/35048692
Kidner, C. A., and Martienssen, R. A. (2005). The developmental role of
microRNA in plants. Curr. Opin. Plant Biol. 8, 38–44. doi: 10.1016/j.pbi.2004.
11.008
Kumar, V., and Van Staden, J. (2017). New insights into plant somatic
embryogenesis: an epigenetic view. Acta Physiol. Plant. 39, 194.
doi: 10.1007/s11738-017-2487-5
Kurczynska, E. U., Gaj, M. D., Ujczak, A., and Mazur, E. (2007).
Histological analysis of direct somatic embryogenesis in Arabidopsis
thaliana (L.) Heynh. Planta 226, 619–628. doi: 10.1007/s00425-007-
0510-6
Lee, C. Y., and Grant, P. A. (2019). “Role of histone acetylation
and acetyltransferases in gene regulation, in Toxicoepigenetics, eds
S. D. McCullough and D. C. Dolinoy (London: Elsevier), 3–30.
doi: 10.1016/B978-0-12-812433-8.00001-0
Lemoine, F., Correia, D., Lefort, V., Doppelt-Azeroual, O., Mareuil, F.,
Cohen- Boulakia, S., et al. (2019). NGPhylogeny.fr: new generation
phylogenetic services for non-specialists. Nucleic Acids Res. 47, W260–W265.
doi: 10.1093/nar/gkz303
Letunic, I., and Bork, P. (2019). Interactive Tree Of Life (iTOL) v4: recent
updates and new developments. Nucleic Acids Res. 47, W256–W259.
doi: 10.1093/nar/gkz239
Liu, H., Yu, H., Tang, G., and Huang, T. (2018). Small but powerful:
function of microRNAs in plant development. Plant Cell Rep. 37, 515–528.
doi: 10.1007/s00299-017-2246-5
Liu, J., Liu, X., Zhang, S., Liang, S., Luan, W., and Ma, X. (2021). TarDB: an online
database for plant miRNA targets and miRNA-triggered phased siRNAs. BMC
Genomics 22, 348. doi: 10.1186/s12864-021-07680-5
Liu, X., Huang, J., Wang, Y., Khanna, K., Xie, Z., Owen, H. A., et al. (2010). The
role of floral organs in carpels, an Arabidopsis loss-of-function mutation in
MicroRNA160a, in organogenesis and the mechanism regulating its expression.
Plant J. 62, 416–428. doi: 10.1111/j.1365-313X.2010.04164.x
Llave, C., Xie, Z., Kasschau, K. D., and Carrington, J. C. (2002). Cleavage of
SCARECROW-like mRNA targets directed by a class of Arabidopsis miRNA.
Science 297, 2053–2056. doi: 10.1126/science.1076311
Ma, W. Y., Li, J. J., Qu, B. Y., He, X., Zhao, X. Q., Li, B., et al. (2014). Auxin
biosynthetic gene TAR2 is involved in low nitrogen-mediated reprogramming
of root architecture in Arabidopsis. Plant J. 78, 70–79. doi: 10.1111/tpj.
12448
Ma, X., Denyer, T., Javelle, M., Feller, A., and Timmermans, M. C. (2021).
Genome-wide analysis of plant miRNA action clarifies levels of regulatory
dynamics across developmental contexts. Genome Res. 31, 811–822.
doi: 10.1101/gr.270918.120
Magnani, E., Jiménez-Gómez, J. M., Soubigou-Taconnat, L., Lepiniec, L., and
Fiume, E. (2017). Profiling the onset of somatic embryogenesis in Arabidopsis.
BMC Genomics 18, 998. doi: 10.1186/s12864-017-4391-1
Mallory, A. C., Bartel, D. P., and Bartel, B. (2005). MicroRNA-directed regulation
of Arabidopsis AUXIN RESPONSE FACTOR17 is essential for proper
development and modulates expression of early auxin response genes. Plant
Cell 17, 1360–1375. doi: 10.1105/tpc.105.031716
Mantiri, F. R., Kurdyukov, S., Lohar, D. P., Sharopova, N., Saeed, N. A., Wang,
X. D., et al. (2008). The transcription factor MtSERF1 of the ERF subfamily
identified by transcriptional profiling is required for somatic embryogenesis
induced by auxin plus cytokinin in Medicago truncatula.Plant Physiol. 146,
1622–1636. doi: 10.1104/pp.107.110379
Mareri, L., Romi, M., and Cai, G. (2019). Arabinogalactan proteins: actors or
spectators during abiotic and biotic stress in plants? Plant Biosyst. 153, 173–185.
doi: 10.1080/11263504.2018.1473525
Márquez-López, R. E., Pérez-Hernández, C., Ku-González, Á., Galaz-Ávalos, R.
M., and Loyola-Vargas, V. M. (2018). Localization and transport of indole-3-
acetic acid during somatic embryogenesis in Coffea canephora. Protoplasma
255, 695–708. doi: 10.1007/s00709-017-1181-1
Mendez-Hernandez, H. A., Ledezma-Rodriguez, M., Avilez-Montalvo, R. N.,
Juarez-Gomez, Y. L., Skeete, A., Avilez-Montalvo, J., et al. (2019). Signaling
overview of plant somatic embryogenesis. Front. Plant Sci. 10, 77.
doi: 10.3389/fpls.2019.00077
Mozgová, I., Köhler, C., and Hennig, L. (2015). Keeping the gate closed: functions
of the polycomb repressive complex PRC 2 in development. Plant J. 83,
121–132. doi: 10.1111/tpj.12828
Mozgová, I., Muñoz-Viana, R., and Hennig, L. (2017). PRC2 represses hormone-
induced somatic embryogenesis in vegetative tissue of Arabidopsis thaliana.
PLoS Genet. 13, e1006562. doi: 10.1371/journal.pgen.1006562
Frontiers in Plant Science | www.frontiersin.org 15 March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
Nic-Can, G. I., López-Torres, A., Barredo-Pool, F., Wrobel, K., Loyola-
Vargas, V. M., Rojas-Herrera, R., et al. (2013). New insights into somatic
embryogenesis: LEAFY COTYLEDON1, BABY BOOM1 and WUSCHEL-
RELATED HOMEOBOX4 are epigenetically regulated in Coffea canephora.
PLoS ONE 8, e72160. doi: 10.1371/journal.pone.0072160
Nodine, M. D., and Bartel, D. P. (2010). MicroRNAs prevent precocious gene
expression and enable pattern formation during plant embryogenesis. Genes
Dev. 24, 2678–2692. doi: 10.1101/gad.1986710
Nowak, K., Moronczyk, J., Wójcik, A., and Gaj, M. D. (2020). AGL15 controls
the embryogenic reprogramming of somatic cells in Arabidopsis through the
histone acetylation-mediated repression of the miRNA biogenesis genes. Int. J.
Mol. Sci. 21, 6733. doi: 10.3390/ijms21186733
Ochoa-Alejo, N., and Loyola-Vargas, V. M. (2016). Somatic
Embryogenesis: Fundamental Aspects and Applications. Springer.
doi: 10.1007/978-3-319-33705-0
Oh, E., Zhu, J.-Y., Bai, M.-Y., Arenhart, R. A., Sun, Y., and Wang, Z.-
Y. (2014). Cell elongation is regulated through a central circuit of
interacting transcription factors in the Arabidopsis hypocotyl. Elife 3, e03031.
doi: 10.7554/eLife.03031.025
Orłowska, A., Igielski, R., Łagowska, K., and Kepczy´
nska, E. (2017). Identification
of LEC1, L1L and polycomb repressive complex 2 genes and their
expression during the induction phase of Medicago truncatula Gaertn.
somatic embryogenesis. Plant Cell Tissue Organ Cult. 129, 119–132.
doi: 10.1007/s11240-016-1161-8
Palatnik, J. F., Allen, E., Wu, X., Schommer, C., Schwab, R., Carrington, J. C., et al.
(2003). Control of leaf morphogenesis by microRNAs. Nature 425, 257–263.
doi: 10.1038/nature01958
Palovaara, J., De Zeeuw, T., and Weijers, D. (2016). Tissue and organ initiation in
the plant embryo: a first time for everything. Ann. Rev. Cell Dev. Biol. 32, 47–75.
doi: 10.1146/annurev-cellbio-111315-124929
Paul, P., Joshi, S., Tian, R., Junior, R. D., Chakrabarti, M., and Perry,
S. E. (2021). The MADS-domain factor AGAMOUS-Like18 promotes
somatic embryogenesis. Plant Physiol. 188, 1617–1631. doi: 10.1093/plphys/
kiab553
Pelletier, J. M., Kwong, R. W., Park, S., Le, B. H., Baden, R., Cagliaria, A., et al.
(2017). LEC1 sequentially regulates the transcription of genes involved in
diverse developmental processes during seed development. Proc. Natl. Acad.
Sci. U.S.A. 114, E6710–E6719. doi: 10.1073/pnas.1707957114
Perry, S. E., Lehti, M. D., and Fernandez, D. E. (1999). The MADS-domain protein
AGAMOUS-like 15 accumulates in embryonic tissues with diverse origins.
Plant Physiol. 120, 121–129. doi: 10.1104/pp.120.1.121
Perry, S. E., Nichols, K. W., and Fernandez, D. E. (1996). The MADS domain
protein AGL15 localizes to the nucleus during early stages of seed development.
Plant Cell 8, 1977–1989. doi: 10.1105/tpc.8.11.1977
Piya, S., Shrestha, S. K., Binder, B., Stewart, C. N. Jr., and Hewezi, T. (2014).
Protein-protein interaction and gene co-expression maps of ARFs and
Aux/IAAs in Arabidopsis. Front. Plant Sci. 5, 744. doi: 10.3389/fpls.2014.
00744
Plotnikova, A., Kellner, M. J., Schon, M. A., Mosiolek, M., and Nodine, M. D.
(2019). MicroRNA dynamics and functions during Arabidopsis embryogenesis.
Plant Cell 31, 2929–2946. doi: 10.1105/tpc.19.00395
Quiroz-Figueroa, F. R., Rojas-Herrera, R., Galaz-Avalos, R. M., and Loyola-Vargas,
V. M. (2006). Embryo production through somatic embryogenesis can be used
to study cell differentiation in plants. Plant Cell Tissue Organ Cult. 86, 285.
doi: 10.1007/s11240-006-9139-6
Raghavan, C., Ong, E. K., Dalling, M. J., and Stevenson, T. W. (2006). Regulation
of genes associated with auxin, ethylene and ABA pathways by 2, 4-
dichlorophenoxyacetic acid in Arabidopsis. Funct. Integrat. Genom. 6, 60–70.
doi: 10.1007/s10142-005-0012-1
Rhoades, M. W., Reinhart, B. J., Lim, L. P., Burge, C. B., Bartel, B., and
Bartel, D. P. (2002). Prediction of plant microRNA targets. Cell 110, 513–520.
doi: 10.1016/S0092-8674(02)00863-2
Rodríguez-Sanz, H., Moreno-Romero, J., Solís, M.-T., Köhler, C., Risueño, M. C.,
and Testillano, P. S. (2014). Changes in histone methylation and acetylation
during microspore reprogramming to embryogenesis occur concomitantly
with BnHKMT and BnHAT expression and are associated with cell totipotency,
proliferation, and differentiation in Brassica napus. Cytogenet. Genome Res. 143,
209–218. doi: 10.1159/000365261
Rose, R. J. (2019). Somatic embryogenesis in the Medicago truncatula
model: cellular and molecular mechanisms. Front. Plant Sci. 10, 267.
doi: 10.3389/fpls.2019.00267
Ruan, J., Chen, H., Zhu, T., Yu, Y., Lei, Y., Yuan, L., et al. (2021). Brassinosteroids
repress the seed maturation program during the seed-to-seedling transition.
Plant Physiol. 186, 534–548. doi: 10.1093/plphys/kiab089
Rubio-Somoza, I., and Weigel, D. (2011). MicroRNA networks and
developmental plasticity in plants. Trends Plant Sci. 16, 258–264.
doi: 10.1016/j.tplants.2011.03.001
Salvo, S. A., Hirsch, C. N., Buell, C. R., Kaeppler, S. M., and Kaeppler, H. F. (2014).
Whole transcriptome profiling of maize during early somatic embryogenesis
reveals altered expression of stress factors and embryogenesis-related genes.
PLoS ONE 9, e111407. doi: 10.1371/journal.pone.0111407
Saptari, R. T., and Susila, H. (2019). Data mining study of hormone biosynthesis
gene expression reveals new aspects of somatic embryogenesis regulation.
In Vitro Cellular Dev. Biol.-Plant 55, 139–152. doi: 10.1007/s11627-01
8-9947-5
Sato, A., and Yamamoto, K. T. (2008). Overexpression of the non-
canonical Aux/IAA genes causes auxin-related aberrant phenotypes in
Arabidopsis. Physiol. Plant. 133, 397–405. doi: 10.1111/j.1399-3054.200
8.01055.x
Sharp, W. (1980). The physiology of in vitro asexual embryogenesis. Horticultural
Rev. 2, 268–310. doi: 10.1002/9781118060759.ch6
Song, X., Li, Y., Cao, X., and Qi, Y. (2019). MicroRNAs and their regulatory
roles in plant–environment interactions. Ann. Rev.Plant Biol. 70, 489–525.
doi: 10.1146/annurev-arplant-050718-100334
Stepanova, A. N., Robertson-Hoyt, J., Yun, J., Benavente, L. M., Xie, D.-
Y., DoleŽal, K., et al. (2008). TAA1-mediated auxin biosynthesis is
essential for hormone crosstalk and plant development. Cell 133, 177–191.
doi: 10.1016/j.cell.2008.01.047
Steunou, A.-L., Rossetto, D., and Côté, J. (2014). “Regulating chromatin by histone
acetylation, in Fundamentals of Chromatin, eds J. Workman and S. Abmayr
(New York, NY: Springer), 147–212. doi: 10.1007/978-1-4614-8624-4_4
Stricker, S. H., Köferle, A., and Beck, S. (2017). From profiles to function in
epigenomics. Nat. Rev. Genet. 18, 51–66. doi: 10.1038/nrg.2016.138
Su, Y. H., Tang, L. P., Zhao, X. Y., and Zhang, X. S. (2021). Plant cell totipotency:
Insights into cellular reprogramming. J. Integrative Plant Biol. 63, 228–243.
doi: 10.1111/jipb.12972
Sugimoto, K., Jiao, Y. L., and Meyerowitz, E. M. (2010). Arabidopsis regeneration
from multiple tissues occurs via a root development pathway. Dev. Cell 18,
463–471. doi: 10.1016/j.devcel.2010.02.004
Suzuki, M., Wang, H. H.-Y., and Mccarty, D. R. (2007). Repression of the LEAFY
COTYLEDON 1/B3 regulatory network in plant embryo development by
VP1/ABSCISIC ACID INSENSITIVE 3-LIKE B3 genes. Plant Physiol. 143,
902–911. doi: 10.1104/pp.106.092320
Szyrajew, K., Bielewicz, D., Dolata, J., Wójcik, A. M., Nowak, K., Szczygiel-
Sommer, A., et al. (2017). MicroRNAs are intensively regulated during
induction of somatic embryogenesis in Arabidopsis. Front. Plant Sci. 8, 18.
doi: 10.3389/fpls.2017.00018
Takanashi, H., Sumiyoshi, H., Mogi, M., Hayashi, Y., Ohnishi, T., and Tsutsumi,
N. (2018). miRNAs control HAM1 functions at the single-cell-layer level and
are essential for normal embryogenesis in Arabidopsis. Plant Mol. Biol. 96,
627–640. doi: 10.1007/s11103-018-0719-8
Tang, G., Reinhart, B. J., Bartel, D. P., and Zamore, P. D. (2003). A
biochemical framework for RNA silencing in plants. Genes Dev. 17, 49–63.
doi: 10.1101/gad.1048103
Tang, W., and Perry, S. E. (2003). Binding site selection for the plant MADS
Domain protein AGL15: an in vitro and in vivo study. J. Biol. Chem. 278,
28154–28159. doi: 10.1074/jbc.M212976200
Thakare, D., Tang, W., Hill, K., and Perry, S. E. (2008). The MADS-domain
transcriptional regulator AGAMOUS-LIKE15 promotes somatic embryo
development in Arabidopsis and soybean. Plant Physiol. 146, 1663–1672.
doi: 10.1104/pp.108.115832
Thorpe, T. A., and Stasolla, C. (2001). “Somatic embryogenesis, in Current trends
in the Embryology of Angiosperms, eds S. S. Bhojwani and W. Y. Soh (Dordrecht:
Springer), 279–336. doi: 10.1007/978-94-017-1203-3_12
Tian, R., Paul, P., Joshi, S., and Perry, S. E. (2020a). Genetic activity during early
plant embryogenesis. Biochemical J. 477, 3743–3767. doi: 10.1042/BCJ20190161
Frontiers in Plant Science | www.frontiersin.org 16 March 2022 | Volume 13 | Article 861556
Joshi et al. AGL15 and SE
Tian, R., Wang, F., Zheng, Q., Niza, V. M. A. G. E., and Downie,
A. B. (2020b). Direct and indirect targets of the Arabidopsis seed
transcription factor ABSCISIC ACID INSENSITIVE3. Plant J. 103, 1679–1694.
doi: 10.1111/tpj.14854
Toonen, M. A. J., Schmidt, E. D. L., Vankammen, A., and Devries, S.
C. (1997). Promotive and inhibitory effects of diverse arabinogalactan
proteins on Daucus carota L. somatic embryogenesis. Planta 203, 188–195.
doi: 10.1007/s004250050181
Tvorogova, V. E., Fedorova, Y. A., Potsenkovskaya, E. A., Kudriashov, A. A.,
Efremova, E. P., Kvitkovskaya, V. A., et al. (2019). The WUSCHEL-related
homeobox transcription factor MtWOX9-1 stimulates somatic embryogenesis
in Medicago truncatula.Plant Cell Tissue Organ Cult. 138, 517–527.
doi: 10.1007/s11240-019-01648-w
Van Hengel, A. J., Guzzo, F., Van Kammen, A., and De Vries, S. C. (1998).
Expression pattern of the carrot EP3 endochitinase genes in suspension cultures
and in developing seeds. Plant Physiol. 117, 43–53. doi: 10.1104/pp.117.1.43
Van Hengel, A. J., Tadesse, Z., Immerzeel, P., Schols, H., Van Kammen, A.,
and De Vries, S. C. (2001). N-acetylglucosamine and glucosamine-containing
arabinogalactan proteins control somatic embryogenesis. Plant Physiol. 125,
1880–1890. doi: 10.1104/pp.125.4.1880
Van Hengel, A. J., Van Kammen, A., and De Vries, S. C. (2002). A relationship
between seed development, Arabinogalactan-proteins (AGPs) and the AGP
mediated promotion of somatic embryogenesis. Physiol. Plant. 114, 637–644.
doi: 10.1034/j.1399-3054.2002.1140418.x
Vandenbussche, F., Vaseva, I., Vissenberg, K., and Van Der Straeten, D. (2012).
Ethylene in vegetative development: a tale with a riddle. New Phytol. 194,
895–909. doi: 10.1111/j.1469-8137.2012.04100.x
Veerappan, V., Chen, N., Reichert, A. I., and Allen, R. D. (2014). HSI2/VAL1 PHD-
like domain promotes H3K27 trimethylation to repress the expression of seed
maturation genes and complex transgenes in Arabidopsis seedlings. BMC Plant
Biol. 14, 1–19. doi: 10.1186/s12870-014-0293-4
Von Arnold, S., Sabala, I., Bozhkov, P., Dyachok, J., and Filonova, L. (2002).
Developmental pathways of somatic embryogenesis. Plant Cell Tissue Organ
Cult. 69, 233–249. doi: 10.1023/A:1015673200621
Wang, F., and Perry, S. E. (2013). Identification of direct targets of FUSCA3, a
key regulator of Arabidopsis seed development. Plant Physiol. 161, 1251–1264.
doi: 10.1104/pp.112.212282
Wang, F. X., Shang, G. D., Wu, L. Y., Xu, Z. G., Zhao, X. Y., and Wang, J. W.
(2020). Chromatin accessibility dynamics and a hierarchical transcriptional
regulatory network structure for plant somatic embryogenesis. Dev. Cell 54,
742–757. doi: 10.1016/j.devcel.2020.07.003
Wang, H., Caruso, L. V., Downie, A. B., and Perry, S. E. (2004). The embryo
MADS domain protein AGAMOUS-Like 15 directly regulates expression of a
gene encoding an enzyme involved in gibberellin metabolism. Plant Cell 16,
1206–1219. doi: 10.1105/tpc.021261
Weiss, D., and Ori, N. (2007). Mechanisms of cross talk between gibberellin and
other hormones. Plant Physiol. 144, 1240–1246. doi: 10.1104/pp.107.100370
Willmann, M. R., Mehalick, A. J., Packer, R. L., and Jenik, P. D. (2011). MicroRNAs
regulate the timing of embryo maturation in Arabidopsis. Plant Physiol. 155,
1871–1884. doi: 10.1104/pp.110.171355
Wójcik, A. M., and Gaj, M. D. (2016). miR393 contributes to the
embryogenic transition induced in vitro in Arabidopsis via the modification
of the tissue sensitivity to auxin treatment. Planta 244, 231–243.
doi: 10.1007/s00425-016-2505-7
Wójcik, A. M., Wójcikowska, B., and Gaj, M. D. (2020). Current perspectives on
the auxin-mediated genetic network that controls the induction of somatic
embryogenesis in plants. Int. J. Mol. Sci. 21, 1333. doi: 10.3390/ijms21041333
Wójcikowska, B., and Gaj, M. D. (2016). “Somatic embryogenesis in arabidopsis,
in Somatic Embryogenesis: Fundamental Aspects and Applications, eds
V. Loyola-Vargas and N. Ochoa-Alejo (Cham: Springer), 185–199.
doi: 10.1007/978-3-319-33705-0_11
Wójcikowska, B., and Gaj, M. D. (2017). Expression profiling of AUXIN
RESPONSE FACTOR genes during somatic embryogenesis induction in
Arabidopsis. Plant Cell Rep. 36, 843–858. doi: 10.1007/s00299-017-2114-3
Wójcikowska, B., Jaskola, K., Gasiorek, P., Meus, M., Nowak, K., and Gaj,
M. D. (2013). LEAFY COTYLEDON2 (LEC2) promotes embryogenic
induction in somatic tissues of Arabidopsis, via YUCCA-mediated
auxin biosynthesis. Planta 238, 425–440. doi: 10.1007/s00425-01
3-1892-2
Wójcikowska, B., Wójcik, A. M., and Gaj, M. D. (2020). Epigenetic regulation
of auxin-induced somatic embryogenesis in plants. Int. J. Mol. Sci. 21, 2307.
doi: 10.3390/ijms21072307
Xu, K. D., Liu, K., Wu, J. X., Wang, W., Zhu, Y. L., Li, C. J., et al. (2018). A MADS-
box gene associated with protocorm-like body formation in Rosa canina alters
floral organ development in Arabidopsis. Canad. J.Plant Sci. 98, 309–317.
doi: 10.1139/cjps-2017-0027
Yang, X., and Zhang, X. (2010). Regulation of somatic embryogenesis in
higher plants. Crit. Rev. Plant Sci. 29, 36–57. doi: 10.1080/073526809034
36291
Yang, Z., Li, C., Wang, Y., Zhang, C., Wu, Z., Zhang, X., et al. (2014). GhAGL15s,
preferentially expressed during somatic embryogenesis, promote embryogenic
callus formation in cotton (Gossypium hirsutum L.). Mol. Genet. Genomics 289,
873–883. doi: 10.1007/s00438-014-0856-y
Yu, Y., Jia, T., and Chen, X. (2017). The ’how’ and ’where’ of plant microRNAs.
New Phytol. 216, 1002–1017. doi: 10.1111/nph.14834
Zheng, Q., and Perry, S. E. (2014). Alterations in the transcriptome of soybean
in response to enhanced somatic embryogenesis promoted by orthologs of
AGAMOUS-Like15 and AGAMOUS-Like18. Plant Physiol. 164, 1365–1377.
doi: 10.1104/pp.113.234062
Zheng, Q., Zheng, Y., Ji, H., Burnie, W., and Perry, S. E. (2016). Gene
regulation by the AGL15 transcription factor reveals hormone interactions
in somatic embryogenesis. Plant Physiol. 172, 2374–2387. doi: 10.1104/pp.16.
00564
Zheng, Q., Zheng, Y., and Perry, S. E. (2013a). AGAMOUS-Like15 promotes
somatic embryogenesis in Arabidopsis and soybean in part by the control
of ethylene biosynthesis and response. Plant Physiol. 161, 2113–2127.
doi: 10.1104/pp.113.216275
Zheng, Q., Zheng, Y., and Perry, S. E. (2013b). Decreased GmAGL15
expression and reduced ethylene synthesis may contribute to reduced somatic
embryogenesis in a poorly embryogenic cultivar of Glycine max. Plant Signal.
Behav. 8, e25422. doi: 10.4161/psb.25422
Zheng, Y., Ren, N., Wang, H., Stromberg, A. J., and Perry, S. E. (2009). Global
identification of targets of the Arabidopsis MADS domain protein AGAMOUS-
Like15. Plant Cell 21, 2563–2577. doi: 10.1105/tpc.109.068890
Zhu, C., and Perry, S. E. (2005). Control of expression and autoregulation
of AGL15, a member of the MADS-box family. Plant J. 41, 583–594.
doi: 10.1111/j.1365-313X.2004.02320.x
Zimmerman, J. L. (1993). Somatic embryogenesis: a model for early development
in higher plants. Plant Cell 5, 1411–1423. doi: 10.2307/3869792
Conflict of Interest: The authors declare that the research was conducted in the
absence of any commercial or financial relationships that could be construed as a
potential conflict of interest.
Publisher’s Note: All claims expressed in this article are solely those of the authors
and do not necessarily represent those of their affiliated organizations, or those of
the publisher, the editors and the reviewers. Any product that may be evaluated in
this article, or claim that may be made by its manufacturer, is not guaranteed or
endorsed by the publisher.
Copyright © 2022 Joshi, Paul, Hartman and Perry. This is an open-access article
distributed under the terms of the Creative Commons Attribution License (CC BY).
The 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.
Frontiers in Plant Science | www.frontiersin.org 17 March 2022 | Volume 13 | Article 861556
... Currently, only very few genotypes, such as Coker 201/310/312/315 [4][5][6], ZM24 [7], Simian 3 [8], YZ1 [9], and Jin668 [6], exhibit high regeneration efficiency via somatic embryogenesis. Among these, Jin668 is a newly developed elite genotype and widely used in genetic engineering and genome editing [2,6,8,[10][11][12][13][14][15][16]. However, the mechanism underlying its strong regeneration ability is not understood. ...
... Together, our results indicate that auxin, cytokinin, ethylene, and wounding-related genes have different expression patterns between Jin668 and TM-1 (Fig. 8). The ethylene responsive gene AGL15 affects callus induction during somatic embryogenesis by affecting the ratio of GA/ABA [5,11]. Wound-inducing factor WIND1 plays a positive role in regulating callus induction, and this gene is up-regulated in Jin668. ...
Article
Full-text available
Background Somatic embryogenesis is a major process for plant regeneration. However, cell communication and the gene regulatory network responsible for cell reprogramming during somatic embryogenesis are still largely unclear. Recent advances in single-cell technologies enable us to explore the mechanism of plant regeneration at single-cell resolution. Results We generate a high-resolution single-cell transcriptomic landscape of hypocotyl tissue from the highly regenerable cotton genotype Jin668 and the recalcitrant TM-1. We identify nine putative cell clusters and 23 cluster-specific marker genes for both cultivars. We find that the primary vascular cell is the major cell type that undergoes cell fate transition in response to external stimulation. Further developmental trajectory and gene regulatory network analysis of these cell clusters reveals that a total of 41 hormone response-related genes, including LAX2 , LAX1 , and LOX3 , exhibit different expression patterns in the primary xylem and cambium region of Jin668 and TM-1. We also identify novel genes, including CSEF , PIS1 , AFB2 , ATHB2 , PLC2 , and PLT3 , that are involved in regeneration. We demonstrate that LAX2 , LAX1 and LOX3 play important roles in callus proliferation and plant regeneration by CRISPR/Cas9 editing and overexpression assay. Conclusions This study provides novel insights on the role of the regulatory network in cell fate transition and reprogramming during plant regeneration driven by somatic embryogenesis.
... In vitro, WOX2 overexpression was found to promote SE and organogenesis in Arabidopsis (Hassani et al. 2022), cotton (Bouchabké-Coussa et al. 2013) and coffee (Arroyo-Herrera et al. 2008). AGL15 is a member of the MADSbox transcription factor family, which plays an important role in SE regulation (reviewed by Joshi et al. 2022). When ectopically expressed, it enhances the embryogenic competency in Arabidopsis (Harding et al. 2003), cotton ) and soybean (Perry et al. 2016). ...
Article
Full-text available
Somatic embryogenesis (SE) is a process by which somatic cells reprogram, acquire totipotency and embark on embryo formation. Although SE is a valuable tool for micropropagation in many crops, it presents specific challenges in woody species due to some bottlenecks, such as loss of embryogenic competence during subcultures and the often-low somatic embryo conversion into plantlets. Hence, great interest exists in exploring the regulatory networks involved on SE. As transcription modulators, long non-coding RNAs (lncRNAs) have been found essential in a wide range of biological processes. This work aimed to identify lncRNAs related to the embryogenic competence in Solanum betaceum Cav. (tamarillo). Nanopore® long-read sequencing was conducted in cell lines with distinct cell fates and, based on their coding potential, 60 transcripts were selected as lncRNA candidates. Similar expression patterns were found among embryogenic cell lines (EC) and cells that lost their embryogenic potential (herein long-term callus, LTC), whereas non-EC (NEC) showed differentially expressed lncRNAs. Whereas lncRNAs upregulated in EC and LTC were predicted to target embryogenesis-related genes, such as AGAMOUS-like 15 and WUSCHEL-related HOMEOBOX 2, lncRNAs upregulated in NEC were predicted to target mainly genes involved in carbohydrate metabolism, cell wall formation, auxin and ethylene signalling pathways. Altogether, these results show the involvement of lncRNA in the process of somatic embryogenesis in S. betaceum, indicating that EC present a pattern of lncRNA expression that suggests its action on genes that directly modulate the morphogenic response in vitro, while in NEC the modulation through this type of RNAs seems to be predominantly reflected in genes more related to cellular physiology. Graphical Abstract
... These subgroups included FLC, AG, AGAMOUS, AGL6, GGM13, SQUA, STMADS11, DEF, and GLO, which could be important for regulating floral organ development. Additionally, subgroups, such as AGL12 [27], AGL17 [28], and TM3 [29] were involved in regulating root development, while others, such as AGL15 [30] were expressed predominantly during embryogenesis and seed development. Finally, there were specific subgroups responsible for stimulating flowering in the tips of the stems and axillary meristems. ...
Article
Full-text available
MADS-box is a key transcription factor regulating the transition to flowering and flower development. Lagerstroemia indica ‘Xiang Yun’ is a new cultivar of crape myrtle characterized by its non-fruiting nature. To study the molecular mechanism underlying the non-fruiting characteristics of ‘Xiang Yun’, 82 MADS-box genes were identified from the genome of L. indica. The physicochemical properties of these genes were examined using bioinformatics methods, and their expression as well as endogenous hormone levels at various stages of flower development were analyzed. The results showed that LiMADS genes were primarily classified into two types: type I and type II, with the majority being type II that contained an abundance of cis-acting elements in their promoters. By screening nine core proteins by predicted protein interactions and performing qRT-PCR analysis as well as in combination with transcriptome data, we found that the expression levels of most MADS genes involved in flower development were significantly lower in ‘Xiang Yun’ than in the wild type ‘Hong Ye’. Hormonal analysis indicated that ‘Xiang Yun’ had higher levels of iP, IPR, TZR, and zeatin during its early stages of flower development than ‘Hong Ye’, whereas the MeJA content was substantially lower at the late stage of flower development of ‘Hong Ye’. Finally, correlation analysis showed that JA, IAA, SA, and TZR were positively correlated with the expression levels of most type II genes. Based on these analyses, a working model for the non-fruiting ‘Xiang Yun’ was proposed. During the course of flower development, plant hormone response pathways may affect the expression of MADS genes, resulting in their low expression in flower development, which led to the abnormal development of the stamen and embryo sac and ultimately affected the fruiting process of ‘Xiang Yun’.
... Through the combination of ChIP-seq and RNA-seq studies, 283 genes were found to be induced and 472 genes were repressed in AGL15overexpressed embryogenic cultures as compared to wild types (Joshi et al., 2022a). The relationship between AGL15 and other transcription factors, hormone genes, and genes involved in epigenetic modification, suggests a more complex network interaction in embryogenesis (Joshi et al., 2022b). ...
Article
Full-text available
Transcription factors (TFs) are diverse groups of regulatory proteins. Through their specific binding domains, TFs bind to their target genes and regulate their expression, therefore TFs play important roles in various growth and developmental processes. Plant embryogenesis is a highly regulated and intricate process during which embryos arise from various sources and undergo development; it can be further divided into zygotic embryogenesis (ZE) and somatic embryogenesis (SE). TFs play a crucial role in the process of plant embryogenesis with a number of them acting as master regulators in both ZE and SE. In this review, we focus on the master TFs involved in embryogenesis such as BABY BOOM (BBM) from the APETALA2/Ethylene-Responsive Factor (AP2/ERF) family, WUSCHEL and WUSCHEL-related homeobox (WOX) from the homeobox family, LEAFY COTYLEDON 2 (LEC2) from the B3 family, AGAMOUS-Like 15 (AGL15) from the MADS family and LEAFY COTYLEDON 1 (LEC1) from the Nuclear Factor Y (NF-Y) family. We aim to present the recent progress pertaining to the diverse roles these master TFs play in both ZE and SE in Arabidopsis, as well as other plant species including crops. We also discuss future perspectives in this context.
... There are several extensive reviews on SE with specific focus on exogenous factors and their regulatory mechanisms, including SE genetic and epigenetic modes (Bednarek and Orłowska 2020;Gulzar et al. 2020;Joshi et al. 2022;Kumar and van Staden 2017;Sivanesan et al. 2022;Us-Camas et al. 2014;Wang et al. 2022;Wojcikowska et al. 2020). However, the molecular mechanism of the transition of somatic cells into SEs remain poorly understood. ...
Article
Full-text available
Key message This review discusses the epigenetic changes during somatic embryo (SE) development, highlights the genes and miRNAs involved in the transition of somatic cells into SEs as a result of epigenetic changes, and draws insights on biotechnological opportunities to study SE development. Abstract Somatic embryogenesis from somatic cells occurs in a series of steps. The transition of somatic cells into somatic embryos (SEs) is the most critical step under genetic and epigenetic regulations. Major regulatory genes such as SERK, WUS, BBM, FUS3/FUSA3, AGL15, and PKL, control SE steps and development by turning on and off other regulatory genes. Gene transcription profiles of somatic cells during SE development is the result of epigenetic changes, such as DNA and histone protein modifications, that control and decide the fate of SE formation. Depending on the type of somatic cells and the treatment with plant growth regulators, epigenetic changes take place dynamically. Either hypermethylation or hypomethylation of SE-related genes promotes the transition of somatic cells. For example, the reduced levels of DNA methylation of SERK and WUS promotes SE initiation. Histone modifications also promote SE induction by regulating SE-related genes in somatic cells. In addition, miRNAs contribute to the various stages of SE by regulating the expression of auxin signaling pathway genes (TIR1, AFB2, ARF6, and ARF8), transcription factors (CUC1 and CUC2), and growth-regulating factors (GRFs) involved in SE formation. These epigenetic and miRNA functions are unique and have the potential to regenerate bipolar structures from somatic cells when a pluripotent state is induced. However, an integrated overview of the key regulators involved in SE development and downstream processes is lacking. Therefore, this review discusses epigenetic modifications involved in SE development, SE-related genes and miRNAs associated with epigenetics, and common cis-regulatory elements in the promoters of SE-related genes. Finally, we highlight future biotechnological opportunities to alter epigenetic pathways using the genome editing tool and to study the transition mechanism of somatic cells.
... Camellia sinensis is an essential cash crop 37 , but researches into the breeding and propagation of superior tea varieties has been relatively slow due to the limitations of its own characteristics 38 . Somatic embryogenesis is a key factor in plant regeneration, and tea somatic embryogenesis occurs in both direct and indirect ways. ...
Article
Full-text available
The high frequency, stable somatic embryo system of tea has still not been established due to the limitations of its own characteristics and therefore severely restricts the genetic research and breeding process of tea plants. In this study, the transcriptome was used to illustrate the mechanisms of gene expression regulation in the somatic embryogenesis of tea plants. The number of DEGs for the (IS intermediate stage)_PS (preliminary stage), ES (embryoid stage)_IS and ES_PS stages were 109, 2848 and 1697, respectively. The enrichment analysis showed that carbohydrate metabolic processes were considerably enriched at the ES_IS stage and performed a key role in somatic embryogenesis, while enhanced light capture in photosystem I could provide the material basis for carbohydrates. The pathway analysis showed that the enriched pathways in IS_PS process were far less than those in ES_IS or ES_PS, and the photosynthesis and photosynthetic antenna protein pathway of DEGs in ES_IS or ES_PS stage were notably enriched and up-regulated. The key photosynthesis and photosynthesis antenna protein pathways and the Lhcb1 gene were discovered in tea plants somatic embryogenesis. These results were of great significance to clarify the mechanism of somatic embryogenesis and the breeding research of tea plants.
... The molecular evidence of miR156mediated regulation of starch accumulation during somatic embryogenesis has also been recently reported in citrus by Feng et al. (2022). Furthermore, Joshi et al. (2022) reviewed the involvement of MCM1, AGAMOUS, DEFICINS and SRF, serum response factor (MADS)-domain transcription factor AGAMOUS-Like 15 (AGL15) in somatic embryo development by influencing functions such as transcription factors, hormone signaling, and epigenetic regulation. Our result on better germination of SE 2 embryos correlates well with the above findings and confirms the need for the above factors for the somatic embryogenesis pathway in plant species. ...
Article
Full-text available
Kinnow (Citrus nobilis Lour. × Citrus deliciosa Ten.) needs to be genetically improved for traits such as seedlessness using biotechnological tools. Indirect somatic embryogenesis (ISE) protocols have been reported for citrus improvement. However, its use is restricted due to frequent occurrences of somaclonal variation and low recovery of plantlets. Direct somatic embryogenesis (DSE) using nucellus culture has played a significant role in apomictic fruit crops. However, its application in citrus is limited due to the injury caused to tissues during isolation. Optimization of the explant developmental stage, explant preparation method, and modification in the in vitro culture techniques can play a vital role in overcoming the limitation. The present investigation deals with a modified in ovulo nucellus culture technique after the concurrent exclusion of preexisting embryos. The ovule developmental events were examined in immature fruits at different stages of fruit growth (stages I–VII). The ovules of stage III fruits (>21–25 mm in diameter) were found appropriate for in ovulo nucellus culture. Optimized ovule size induced somatic embryos at the micropylar cut end on induction medium containing Driver and Kuniyuki Walnut (DKW) basal medium with kinetin (KIN) 5.0 mg L⁻¹ and malt extract (ME) 1,000 mg L⁻¹. Simultaneously, the same medium supported the maturation of somatic embryos. The matured embryos from the above medium gave robust germination with bipolar conversion on Murashige and Tucker (MT) medium + gibberellic acid (GA3) 2.0 mg L⁻¹ + ά-naphthaleneacetic acid (NAA) 0.5 mg L⁻¹ + spermidine 100 mg L⁻¹ + coconut water (CW) 10% (v/v). The bipolar germinated seedlings established well upon preconditioning in a plant bio regulator (PBR)-free liquid medium under the light. Consequently, a cent percent survival of emblings was achieved on a potting medium containing cocopeat:vermiculite:perlite (2:1:1). Histological studies confirmed the single nucellus cell origin of somatic embryos by undergoing normal developmental events. Eight polymorphic Inter Simple Sequence Repeats (ISSR) markers confirmed the genetic stability of acclimatized emblings. Since the protocol can induce rapid single-cell origin of genetically stable in vitro regenerants in high frequency, it has potential for the induction of solid mutants, besides crop improvement, mass multiplication, gene editing, and virus elimination in Kinnow mandarin.
Article
Full-text available
Somatic embryogenesis (SE) is a process by which an embryo is derived from somatic tissue. Transcription factors (TFs) have been identified that control this process. One such TF that promotes SE is AGAMOUS-like 15 (AGL15). Prior work has shown that AGL15 can both induce and repress gene expression. One way this type of dual function TF works is via protein interactions, so a yeast 2-hybrid (Y2H) screen was undertaken. One intriguing protein with which AGL15 interacted in Y2H was LBD40. LBD40 encodes a LATERAL ORGAN BOUNDARIES (LOB)-domain TF that is unique to plants and is primarily expressed during seed development. Here, we confirm the AGL15-LBD40 interaction by quantitative assays and in planta co-immunoprecipation. We also document a role for LBD40, and the closely related protein LBD41, in supporting SE. To determine downstream genes potentially controlled by LBD40, chromatin immunoprecipitation followed by high throughput sequencing (ChIP-seq) was used. More than 400 binding regions for LBD40 were consistently found genome-wide. To determine genes responsive to LBD40/41 accumulation, RNA-seq analysis of transcriptomes of wild-type control and loss-of-function lbd40/lbd41 was performed. Combining these datasets provides insight into genes directly and indirectly controlled by these LOB domain TFs. The gene ontology (GO) enrichment analysis of these regulated genes showed an overrepresentation of biological processes that are associated with SE, further indicating the importance of LBD40 in SE. This work provides insight into SE, a poorly understood, but essential process to generate transgenic plants to meet agricultural demands or test gene function. This manuscript reports on experiments to understand the role that LDB40, a TF, plays in support of SE by investigating genes directly and indirectly controlled by LBD40 and examining physical and genetic interactions with other TFs active in SE. We uncover targets of LBD40 and an interacting TF of the MADS family and investigate targets involvement in SE.
Article
The requirement of light on somatic embryogenesis (SE) has been documented in many species; however, no mechanism of action has been elucidated. Using Arabidopsis SE as a model, the effect of red light (660 nm) during the induction phase corresponding to the formation of the embryogenic tissue was examined. Analyses of several phytochrome mutants revealed that red light signaling, conducive to SE, was mediated by PHYTOCHROME E (PHYE). Both phyE and darkness were sufficient to repress the formation of somatic embryos and reduced the expression of CONSTITUTIVE PHOTOMORPHIC DWARF 3 (CPD3), a rate limiting step in brassinosteroid (BR) biosynthesis, as well as AGAMOUS LIKE 15 (AGL15), a key inducer of many SE genes. We further integrated BR signaling and nitric oxide (NO) with PHYE by demonstrating that applications of both compounds to phyE explants and WT explants cultured in the dark partially restored AGL15 expression. These results demonstrate that SE induction by red light operates via PHYE through BR signaling and NO required to induce AGL15.
Article
Full-text available
Some of the hormone crosstalk and transcription factors (TFs) involved in wound-induced organ regeneration have been extensively studied in the model plant Arabidopsis thaliana. In previous work, we established Solanum lycopersicum “Micro-Tom” explants without the addition of exogenous hormones as a model to investigate wound-induced de novo organ formation. The current working model indicates that cell reprogramming and founder cell activation requires spatial and temporal regulation of auxin-to-cytokinin (CK) gradients in the apical and basal regions of the hypocotyl combined with extensive metabolic reprogramming of some cells in the apical region. In this work, we extended our transcriptomic analysis to identify some of the gene regulatory networks involved in wound-induced organ regeneration in tomato. Our results highlight a functional conservation of key TF modules whose function is conserved during de novo organ formation in plants, which will serve as a valuable resource for future studies.
Article
Full-text available
AGAMOUS-Like 18 (AGL18) is a MADS domain transcription factor that is structurally related to AGAMOUS-Like 15 (AGL15). Here we show that, like AGL15, AGL18 can promote somatic embryogenesis (SE) when ectopically expressed in Arabidopsis (Arabidopsis thaliana). Based on loss-of-function mutants, AGL15 and AGL18 have redundant functions in developmental processes such as SE. To understand the nature of this redundancy, we undertook a number of studies to look at the interaction between these factors. We studied the genome-wide direct targets of AGL18 to characterize its roles at the molecular level using chromatin immunoprecipitation (ChIP)-SEQ combined with RNA-SEQ. The results demonstrated that AGL18 binds to thousands of sites in the genome. Comparison of ChIP-SEQ data for AGL15 and AGL18 revealed substantial numbers of genes bound by both AGL15 and AGL18, but there were also differences. GO analysis revealed that target genes were enriched for seed, embryo, and reproductive development as well as hormone and stress responses. The results also demonstrated that AGL15 and AGL18 interact in a complex regulatory loop, where AGL15 inhibited transcript accumulation of AGL18, while AGL18 increased AGL15 transcript accumulation. Co-immunoprecipitation revealed an interaction between AGL18 and AGL15 in somatic embryo tissue. The binding and expression analyses revealed a complex crosstalk and interactions among embryo transcription factors and their target genes. In addition, our study also revealed that phosphorylation of AGL18 and AGL15 was crucial for the promotion of SE.
Article
Full-text available
Background In plants, microRNAs (miRNAs) are pivotal regulators of plant development and stress responses. Different computational tools and web servers have been developed for plant miRNA target prediction; however, in silico prediction normally contains false positive results. In addition, many plant miRNA target prediction servers lack information for miRNA-triggered phased small interfering RNAs (phasiRNAs). Creating a comprehensive and relatively high-confidence plant miRNA target database is much needed. Results Here, we report TarDB, an online database that collects three categories of relatively high-confidence plant miRNA targets: (i) cross-species conserved miRNA targets; (ii) degradome/PARE (Parallel Analysis of RNA Ends) sequencing supported miRNA targets; (iii) miRNA-triggered phasiRNA loci. TarDB provides a user-friendly interface that enables users to easily search, browse and retrieve miRNA targets and miRNA initiated phasiRNAs in a broad variety of plants. TarDB has a comprehensive collection of reliable plant miRNA targets containing previously unreported miRNA targets and miRNA-triggered phasiRNAs even in the well-studied model species. Most of these novel miRNA targets are relevant to lineage-specific or species-specific miRNAs. TarDB data is freely available at http://www.biosequencing.cn/TarDB . Conclusions In summary, TarDB serves as a useful web resource for exploring relatively high-confidence miRNA targets and miRNA-triggered phasiRNAs in plants.
Article
Full-text available
Development of complex organisms requires the delicate and dynamic spatiotemporal regulation of gene expression. Central to this are microRNAs (miRNAs). These mobile small RNAs offer specificity in conveying positional information and versatility in patterning the outcomes of gene expression. However, the parameters that shape miRNA output during development are still to be clarified. Here, we address this question on a genome-wide scale, using the maize shoot apex as a model. We show that patterns and levels of miRNA accumulation are largely determined at the transcriptional level, but are finessed post-transcriptionally in a tissue-dependent manner. The stem cell environments of the shoot apical meristem and vasculature appear particularly liable to this. Tissue-specific effects are also apparent at the level of target repression, with target cleavage products in the vasculature exceeding those of other tissues. Our results argue against a clearance mode of regulation purely at the level of transcript cleavage, leading us to propose that transcript cleavage provides a baseline level of target repression, onto which miRNA-driven translational repression can act to toggle the mode of target regulation between clearance and rheostat. Our data show how the inherent complexities of miRNA pathways allow the accumulation and activity of these small RNAs to be tailored in space and time to bring about the gene expression versatility needed during development.
Article
Full-text available
In flowering plants, repression of the seed maturation program is essential for the transition from the seed to the vegetative phase, but the underlying mechanisms remain poorly understood. The B3-domain protein VIVIPAROUS1/ABSCISIC ACID INSENSITIVE3-LIKE 1 (VAL1) is involved in repressing the seed maturation program. Here we uncovered a molecular network triggered by the plant hormone brassinosteroid (BR) that inhibits the seed maturation program during the seed-to-seedling transition in Arabidopsis (Arabidopsis thaliana). val1-2 mutant seedlings treated with a BR biosynthesis inhibitor form embryonic structures, whereas BR signaling gain-of-function mutations rescue the embryonic structure trait. Furthermore, the BR-activated transcription factors BRI1-EMS-SUPPRESSOR 1 (BES1) and BRASSINAZOLE-RESISTANT 1 (BZR1) bind directly to the promoter of AGAMOUS-LIKE15 (AGL15), which encodes a transcription factor involved in activating the seed maturation program, and suppress its expression. Genetic analysis indicated that BR signaling is epistatic to AGL15 and represses the seed maturation program by downregulating AGL15. Finally, we showed that the BR-mediated pathway functions synergistically with the VAL1/2-mediated pathway to ensure the full repression of the seed maturation program. Together, our work uncovered a mechanism underlying the suppression of the seed maturation program, shedding light on how BR promotes seedling growth.
Article
Full-text available
Data-driven models in a combination of optimization algorithms could be beneficial methods for predicting and optimizing in vitro culture processes. This study was aimed at modeling and optimizing a new embryogenesis medium for chrysanthemum. Three individual data-driven models, including multi-layer perceptron (MLP), adaptive neuro-fuzzy inference system (ANFIS), and support vector regression (SVR), were developed for callogenesis rate (CR), embryogenesis rate (ER), and somatic embryo number (SEN). Consequently, the best obtained results were used in the fusion process by a bagging method. For medium reformulation, effects of eight ionic macronutrients on CR, ER, and SEN and effects of four vitamins on SEN were evaluated using data fusion (DF)–non-dominated sorting genetic algorithm-II (NSGA-II) and DF-genetic algorithm (GA), respectively. Results showed that DF models with the highest R² had superb performance in comparison with all other individual models. According to DF-NSGAII, the highest ER and SEN can be obtained from the medium containing 14.27 mM NH4⁺, 38.92 mM NO3⁻, 22.79 mM K⁺, 5.08 mM Cl⁻, 3.34 mM Ca²⁺, 1.67 mM Mg²⁺, 2.17 mM SO4²⁻, and 1.44 mM H2PO4⁻. Based on the DF-GA model, the maximum SEN can be obtained from a medium containing 0.61 μM thiamine, 5.93 μM nicotinic acid, 0.25 μM biotin, and 0.26 μM riboflavin. The efficiency of the established-optimized medium was experimentally compared to Murashige and Skoog medium (MS) for embryogenesis of five chrysanthemum cultivars, and results indicated the efficiency of optimized medium over MS medium. Key points • MLP, SVR, and ANFIS were fused by a bagging method to develop a data fusion model. • NSGA-II and GA were linked to the data fusion model for establishing and optimizing a new embryogenesis medium. • The new culture medium (HNT) had better efficiency than MS medium.
Article
Full-text available
Seeds are essential for human civilization, so understanding the molecular events underpinning seed development and the zygotic embryo it contains is important. In addition, the approach of somatic embryogenesis is a critical propagation and regeneration strategy to increase desirable genotypes, to develop new genetically modified plants to meet agricultural challenges, and at a basic science level, to test gene function. We briefly review some of the transcription factors (TFs) involved in establishing primary and apicalmeristems during zygotic embryogenesis, as well as TFs necessary and/or sufficient to drive somatic embryo programs. We focus on the model plant Arabidopsis for which many tools are available, and review as well as speculate about comparisons and contrasts between zygotic and somatic embryo processes.
Article
Full-text available
The auxin-induced embryogenic reprogramming of plant somatic cells is associated with extensive modulation of the gene expression in which epigenetic modifications, including DNA methylation, seem to play a crucial role. However, the function of DNA methylation, including the role of auxin in epigenetic regulation of the SE-controlling genes, remains poorly understood. Hence, in the present study, we analysed the expression and methylation of the TF genes that play a critical regulatory role during SE induction (LEC1, LEC2, BBM, WUS and AGL15) in auxin-treated explants of Arabidopsis. The results showed that auxin treatment substantially affected both the expression and methylation patterns of the SE-involved TF genes in a concentration-dependent manner. The auxin treatment differentially modulated the methylation of the promoter (P) and gene body (GB) sequences of the SE-involved genes. Relevantly, the SE-effective auxin treatment (5.0 µM of 2,4-D) was associated with the stable hypermethylation of the P regions of the SE-involved genes and a significantly higher methylation of the P than the GB fragments was a characteristic feature of the embryogenic culture. The presence of auxin-responsive (AuxRE) motifs in the hypermethylated P regions suggests that auxin might substantially contribute to the DNA methylation-mediated control of the SE-involved genes.
Article
Full-text available
The embryogenic transition of somatic cells requires an extensive reprogramming of the cell transcriptome. Relevantly, the extensive modulation of the genes that have a regulatory function, in particular the genes encoding the transcription factors (TFs) and miRNAs, have been indicated as controlling somatic embryogenesis (SE) that is induced in vitro in the somatic cells of plants. Identifying the regulatory relationships between the TFs and miRNAs during SE induction is of central importance for understanding the complex regulatory interplay that fine-tunes a cell transcriptome during the embryogenic transition. Hence, here, we analysed the regulatory relationships between AGL15 (AGAMOUS-LIKE 15) TF and miR156 in an embryogenic culture of Arabidopsis. Both AGL15 and miR156 control SE induction and AGL15 has been reported to target the MIR156 genes in planta. The results showed that AGL15 contributes to the regulation of miR156 in an embryogenic culture at two levels that involve the activation of the MIR156 transcription and the containment of the abundance of mature miR156 by repressing the miRNA biogenesis genes DCL1 (DICER-LIKE1), SERRATE and HEN1 (HUA-ENHANCER1). To repress the miRNA biogenesis genes AGL15 seems to co-operate with the TOPLESS co-repressors (TPL and TPR1-4), which are components of the SIN3/HDAC silencing complex. The impact of TSA (trichostatin A), an inhibitor of the HDAC histone deacetylases, on the expression of the miRNA biogenesis genes together with the ChIP results implies that histone deacetylation is involved in the AGL15-mediated repression of miRNA processing. The results indicate that HDAC6 and HDAC19 histone deacetylases might co-operate with AGL15 in silencing the complex that controls the abundance of miR156 during embryogenic induction. This study provides new evidence about the histone acetylation-mediated control of the miRNA pathways during the embryogenic reprogramming of plant somatic cells and the essential role of AGL15 in this regulatory mechanism.
Article
Sweet sorghum (Sorghum bicolor L) is a C4 plant, and it recently has emerged as a model feedstock for biofuel production. MicroRNAs (miRNAs) are known as key regulators in different biological processes in plant including sugar accumulation. Despite bioinformatics analyses for the discovery of miRNAs and prediction of targets in recent years, little information is available on the expression of mature candidate miRNAs and biological validation of their targets in sweet sorghum. This study aims to gain an insight on the conserved and novel miRNAs as well as their potential targets associated with biofuel related traits, with the help of next generation RNA sequencing (RNA-seq) followed by RT-qPCR validation. In order to identify biofuel related miRNAs and their potential targets in sweet sorghum, small RNA libraries were constructed from four varieties, i.e., Achi Turi, Dale, Dasht Local, and Topper 76-6 after being grown under controlled conditions, and sequenced by Illumina Hiseq 2000 platform. Sequencing yield from 12 libraries, 3 from each variety i.e., leaf vegetative stage, leaf reproductive stage and stem reproductive stage, ranged from 19 to 48 million reads. Mapping with Sorghum bicolor ensemble genome while aligned with miRBase-release 21, 276 conserved and novel biofuel relevant miRNAs were predicted in all varieties. Target prediction by employing psRNA Target tool indicated 475 biofuel relevant targets for identified conserved and novel miRNAs. We performed the RT-qPCR analysis on selected 36 candidate mature miRNAs along with their 18 targets. Potential biofuel relevant targets identified in our study included the sucrose phosphate synthase (SPS), sucrose synthase (SUS), sucrose transporters (SUT) and invertase (INV) enzymes, which are believed to play key role in sucrose synthesis, accumulation and transport. Altogether, 19 predicted miRNAs and their 15 target mRNAs showed significant change in expression under the three growth stages of four varieties. Significant down-regulation at reproductive stage was observed with Sbi-miR160, Sbi-miR164, Sbi-miR166 and Sbi-miR168 in most of the varieties while significant upregulation was observed for Sbi-miR156a and Sbi-miR156e during vegetative stages. This study showed a positive as well as negative correlations between levels of miRNA expressions and that of their target mRNAs. These findings, both computational and experimental, provide valuable insights on the role of miRNAs and their targets during sugar accumulation stages. The identification of these significantly expressed miRNAs and their targets will help to explore the molecular mechanisms for sugar accumulation in sweet sorghum.
Article
Somatic embryogenesis is the regeneration of embryos from the somatic cell via dedifferentiation and redifferentiation without the occurrence of fertilization. A complex network of genes regulates the somatic embryogenesis process. Especially, microRNAs (miRNAs) have emerged as key regulators by affecting phytohormone biosynthesis, transport and signal transduction pathways. miRNAs are small, non-coding small RNA regulatory molecules involved in various developmental processes including somatic embryogenesis. Several types of miRNAs such as miR156, miR157, miR 159, miR 160, miR165, miR166, miR167, miR390, miR393 and miR396 have been reported to intricate in regulating somatic embryogenesis via targeting the phytohormone signaling pathways. Here we review current research progress on the miRNA-mediated regulation involved in somatic embryogenesis via regulating auxin, ethylene, abscisic acid and cytokinin signaling pathways. Further, we also discussed the possible role of other phytohormone signaling pathways such as gibberellins, jasmonates, nitric oxide, polyamines and brassinosteroids. Finally, we conclude by discussing the expression of miRNAs and their targets involved in somatic embryogenesis and possible regulatory mechanisms cross talk with phytohormones during somatic embryogenesis.