ArticlePDF AvailableLiterature Review

Abstract

Flowering plants produce flowers and one of the most complex floral structures is the pistil or the gynoecium. All the floral organs differentiate from the floral meristem. Various reviews exist on molecular mechanisms controlling reproductive development, but most focus on a short time window and there has been no recent review on the complete developmental time frame of gynoecium and fruit formation. Here, we highlight recent discoveries, including the players, interactions and mechanisms that govern gynoecium and fruit development in Arabidopsis. We also present the currently known gene regulatory networks from gynoecium initiation until fruit maturation.
REVIEW
Gynoecium and fruit development in Arabidopsis
Humberto Herrera-Ubaldo and Stefan de Folter*
ABSTRACT
Flowering plants produce flowers and one of the most complex floral
structures is the pistil or the gynoecium. All the floral organs
differentiate from the floral meristem. Various reviews exist on
molecular mechanisms controlling reproductive development, but
most focus on a short time window and there has been no recent
review on the complete developmental time frame of gynoecium and
fruit formation. Here, we highlight recent discoveries, including the
players, interactions and mechanisms that govern gynoecium and
fruit development in Arabidopsis. We also present the currently
known gene regulatory networks from gynoecium initiation until fruit
maturation.
KEY WORDS: Gynoecium, Fruit, Gene regulatory networks,
Transcription factors, Hormones
Introduction
Multicellular life on Earth is inconceivable without plants. Our
planet harbors close to 400,000 vascular plant species, of which
approximately 94% are seed plants (Govaerts, 2001; Willis, 2017),
which produce and reproduce themselves via seed. The vast
majority of seed plants are angiosperms, which produce flowers,
such as Arabidopsis thaliana of the Brassicaceae family (Fig. 1A).
Flower and seed formation are adaptive advantages of plant sexual
reproduction, which requires male and female gametes. In general,
flowers have four different types of floral organs placed in four
whorls, from outside to inside: sepals, petals, stamens and carpels
(see Glossary, Box 1). The stamen, also called the androecium, is
the malepart of the flower that produces the male gametophyte
from which gametes differentiate. The femalepart of the flower is
formed by one or more carpels. In Arabidopsis, two congenitally
fused carpels form the pistil, which is also called gynoecium (see
Glossary, Box 1). Ovules are formed inside the gynoecium and the
female gametophyte, which produces the female gametes, develops
within each ovule (see Glossary, Box 1). Upon fertilization, seed
development begins and, in most plant species, the gynoecium turns
into a fruit (reviewed by Alvarez-Buylla et al., 2010; Dresselhaus
et al., 2016).
The identification of AGAMOUS (AG) as a transcription factor
controlling male and female reproductive organ development in
plants (androecium and gynoecium, respectively), marked the
beginning of a journey towards the understanding of the molecular
aspects of flower formation (Bowman et al., 1989; Irish, 2017;
Yanofsky et al., 1990). During three decades of research, many
master regulators, such as SEPALLATAs (SEPs), CRABS
CLAW (CRC), SPATULA (SPT), SHATTERPROOFs (SHPs),
FRUITFULL (FUL) and SEEDSTICK (STK), among many others,
have emerged as crucial regulators of reproductive development,
especially of gynoecium development (reviewed by Alvarez-Buylla
et al., 2010; Bowman et al., 1999; Ferrándiz et al., 2010; Reyes-
Olalde et al., 2013; Roeder and Yanofsky, 2006; Simonini and
Østergaard, 2019; Zúñiga-Mayo et al., 2019). Besides the
identification of more transcription factors involved in gynoecium
development, information on genes acting downstream of them
have also been discovered (reviewed by Pajoro et al., 2014).
Together, these discoveries have opened the road for charting gene
regulatory networks (GRNs). In addition, several datasets from
chromatin immunoprecipitation followed by sequencing (ChIP-seq)
experiments, have allowed genome-wide analysis of binding events
for many master regulators that participate in the transition to
reproduction and flower development (Chen et al., 2018). Now,
the fine-tuning aspects of flower development are starting to be
revealed.
Over the years, many review articles on flower and fruit
development have been published (Alvarez-Buylla et al., 2010;
Ballester and Ferrándiz, 2017; Bowman et al., 1999; Ferrándiz et al.,
2010; Marsch-Martínez and de Folter, 2016; Reyes-Olalde and de
Folter, 2019; Reyes-Olalde et al., 2013; Roeder and Yanofsky,
2006; Sehra and Franks, 2015; Simonini and Østergaard, 2019;
Smyth et al., 1990; Zúñiga-Mayo et al., 2019). However, there is
currently no recent review that includes all the current GRNs known
to date for gynoecium and fruit development. In 2015, we proposed
approaches and tools to construct a comprehensive GRN for
gynoecium development (Chávez Montes et al., 2015). Therefore,
this Review focuses on recent discoveries and the integration of the
GRNs that guide the development of Arabidopsis, starting from
gynoecium initiation until fruit maturation.
From a floral meristem to gynoecium initiation (stages 1-6)
Floral meristem
Plants possess the ability to form new organs continuously
(reviewed by Gaillochet and Lohmann, 2015; Sablowski, 2007).
When environmental and endogenous genetic cues are met, the
shoot apical meristem (SAM; see Glossary, Box 1) transitions to an
inflorescence meristem (IM; see Glossary, Box 1) (Fig. 1B,E),
marking the beginning of the reproductive phase of the plant
(reviewed by Andrés and Coupland, 2012). At the flanks of the IM,
auxin induces differentiation to give rise to flower primordia, each
with a floral meristem (FM; see Glossary, Box 1). In Arabidopsis,
the most recently formed flower primordium is referred to as
stage 1(Smyth et al., 1990). When the next flower primordium is
formed, the previous one is called stage 2, and so on. At floral stage
3, the flower primordium increases in size and sepal primordia
become visible (Fig. 1B,C,F). The FM gives rise to the different
floral organs present in a mature flower (Fig. 1D,F) (reviewed by
Denay et al., 2017). During the next stages, sepals continue to grow;
during stage 5, petal and stamen primordia appear, whereas stage 6
is characterized by the complete coverage by the sepals of the
FM and internal flower primordia, including the gynoecium
Unidad de Genomica Avanzada (UGA-Langebio), Centro de Investigacion y de
Estudios Avanzados del Instituto Politecnico Nacional (CINVESTAV-IPN), Km. 9.6
Libramiento Norte, Carretera Irapuato-Leon, Irapuato 36824, Guanajuato, Mexico.
*Author for correspondence (stefan.defolter@cinvestav.mx)
H.H., 0000-0002-5408-4022; S.d., 0000-0003-4363-7274
1
© 2022. Published by The Company of Biologists Ltd
|
Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
primordium (see Glossary, Box 1), which becomes visible at stage 6
(Fig. 1F,G) (Alvarez-Buylla et al., 2010; Denay et al., 2017; Smyth
et al., 1990).
Floral meristem maintenance and termination
In general, the genes that regulate meristem maintenance in the
SAM and the FM are shared (reviewed by Chang et al., 2020;
Gaillochet and Lohmann, 2015). The role of the WUSCHEL-
CLAVATA3 (WUS-CLV3) circuit in plant stem cell niche
maintenance is key (Brand et al., 2000; Schoof et al., 2000). The
activity of WUS, together with LEAFY (LFY) and others,
converges on the regulation of AG, which starts to be expressed at
stage 3 in the FM (Fig. 2A). In turn, AG starts to repress WUS and
activates several genes that negatively regulate WUS expression
(Fig. 2).
The meristematic activity of the FM ends when the gynoecium
primordium is formed at stage 6; however, the programs controlling
FM termination begin at stage 3. The molecular mechanisms
underlying FM termination have recently been reviewed (Chang
et al., 2020; Lee et al., 2019; Shang et al., 2019; Xu et al., 2019;
Zúñiga-Mayo et al., 2019). The GRNs that control FM termination
(Fig. 2) and gynoecium initiation (Fig. 2B) contain at least 15
transcription factors and various other proteins. AG plays a leading
role in both these processes, as illustrated in the ABC and quartet
model for floral organ formation (reviewed by Coen and
Meyerowitz, 1991; Theißen and Saedler, 2001). AG is therefore
observed in the GRNs as the main hub (Fig. 2A). AG activates
pathways related to auxin and cytokinin signaling Maoiléidigh
et al., 2018; Yamaguchi et al., 2017). Cytokinin signaling, in turn,
also controls AG (Gómez-Felipe et al., 2021; Rong et al., 2018).
The effects of AG in the control of FM termination and gynoecium
development must be tightly controlled to perform those functions
properly. Indeed, the mechanisms that control AG expression
are diverse, involving transcription factors, microRNAs, histone
deacetylase complexes and RNA-binding complexes (reviewed by
Pelayo et al., 2021). Interactions with other proteins and the
formation of protein complexes provide additional layers of control
of AG activity (Box 2). For example, the interaction with SEP3, via
the formation of dimers or tetramers, controls different biological
processes (Hugouvieux et al., 2018; Lai et al., 2020).
New nodes in the GRNs
Over the past few years, some new nodes have been included in the
networks that underlie gynoecium initiation. AINTEGUMENTA
(ANT), which has reported roles in the establishment of
flower primordia and the development of floral organs, has
now been identified to function at very early stages of flower
formation (stages 3-6), and the roles of ANT are shared with
AINTEGUMENTA-LIKE 6 (AIL6) (Krizek et al., 2020, 2021). At
stage 3, both proteins directly regulate AG, as well as other genes,
such as MONOPTEROS (MP) and REVOLUTA (REV), which
are related to meristem activity regulation and new primordia
formation (Krizek et al., 2021) (Fig. 2A). At stage 6, ANT positively
regulates SPT, an important gene for tissue development in
the medial domain, as described below (Fig. 2B) (Krizek et al.,
2020).
Epigenetic silencing of AG is illustrated by the participation of
POLYCOMB-GROUP (PcG) complexes with CURLY LEAF
(CLF), EMBRYONIC FLOWERING 1 (EMF1) and EMF2. In a
similar way, APETALA 2 (AP2) recruits TOPLESS (TPL) and
HISTONE DEACETYLASE 19 (HDA19) to repress AG (Fig. 2A)
(Pelayo et al., 2021). Now, some evidence points to the involvement
of PcG in silencing of WUS (Sun et al., 2019) and the modification
of chromatin in the control of auxin biosynthesis by AG and CRC
(Fig. 2) (Yamaguchi et al., 2018).
KNUCKLES (KNU), a C2H2 zinc-finger transcription factor,
also functions during FM termination in addition to its role in
repressing WUS in the FM (Fig. 2B). A recent analysis of its
expression pattern has revealed the presence of KNU in the CLV3
expression domain. The extended functions of KNU include
suppression of the expression of several floral meristem
regulators, such as CLV1 and CLV3, at stage 6 (Kwas
niewska
et al., 2021; Shang et al., 2021).
The transcription factor ETTIN (ETT), also known as AUXIN
RESPONSE FACTOR 3 (ARF3), is involved in gynoecium
patterning (Heisler et al., 2001). An additional function of
ETT is to inhibit cytokinin biosynthesis by repressing
ISOPENTENYLTRANSFERASE (IPT) and LONELY GUY (LOG)
genes, and the gene encoding the cytokinin receptor
ARABIDOPSIS HISTIDINE KINASE 4 (AHK4) during stages 5
Box 1. Glossary
Carpel. The female reproductive structure of a flower, consisting of an
ovary, a stigma and a style.
Carpel margin meristem (CMM). A meristematic region in the medial
domain, a zone located where the carpel margins fused. The CMM is the
origin of the medial tissues: placenta, ovules, septum, style and stigma.
Clearly visible at stages 7 and 8.
Chalaza. The basal part of an ovule opposite to the micropyle; where the
integuments, the nucellus and the funiculus are joined.
Floral meristem (FM). The progenitor of all the flower organs.
Fruit dehiscence. The opening of a mature fruit to release the seeds.
Funiculus. The stalk that attaches an ovule or seed to the placenta.
Gynoecium. The female part of a flower. In Arabidopsis, the gynoecium
consists of two congenitally fused carpels.
Gynoecium primordium. A dome-shaped group of cells, first visible at
stage 6.
Inflorescence meristem (IM). The region at the tip of the growing shoot
containing meristematic cells that generates the flowers.
Lateral domain. The two regions that are on either side of the medial
domain, corresponding to the carpel walls or valves. Visible from stage 7
onwards.
Medial domain. In Arabidopsis, the two carpels are fused vertically at
their margins, and these fused margins correspond to the medial domain
of the gynoecium. Visible from stage 7 onwards (oval-shaped hollow
tube).
Nucellus. The central part of the ovule that contains the embryo sac.
Ovule. A structure that contains the female reproductive cells, after
fertilization ovules become seeds. Ovule primordia arise at stage 9.
Placenta. The region within the ovary to which the ovules and seeds are
attached.
Replum. Located at the outer parts of the septum in the medial domain.
The valves are attached to them.
Septum. In Arabidopsis, there is a false septum that divides the two
carpels. Contains the transmitting tract.
Shoot apical meristem (SAM). The region at the tip of the growing shoot
containing meristematic cells that generates the aerial organs (leaves
and branches).
Silique. A type of dry fruit that has two fused carpels in Arabidopsis,
characteristic of the Brassicaceae family.
Stigma. Specialized epidermal cells located at the top of the gynoecium,
which capture the pollen grains, allowing germination and the first steps
of pollen tube growth.
Style. The tissue that connects the stigma with the ovary.
Transmitting tract. Specialized tissue derived from PCD in the style and
some septum cells in the ovary. Possesses ECM that provides nutrients,
guidance and support to the growing pollen tubes.
Valves. The outer tissue of the ovary; the carpel walls.
2
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
and 6. Through these mechanisms, ETT contributes to FM
termination (Fig. 2B) (Zhang et al., 2018).
Conversely, cytokinin also activates transcription factors important
for gynoecium initiation and patterning, which starts around stage 6
(Fig. 2B). For example, the master regulator AG is regulated by
cytokinin via the Arabidopsis response regulators type-B (ARRs)
(Gómez-Felipe et al., 2021; Rong et al., 2018). Furthermore, other
transcription factors that act in parallel downstream of AG are CRC,
SHP2 and SPT, which are important for gynoecium initiation and
patterning, as well as being positively regulated by ARR type-B
transcription factors (Fig. 2B) (Gómez-Felipe et al., 2021).
Gene expression: additional nodes in the GRNs?
Many transcription factors involved in gynoecium development
(Reyes-Olalde et al., 2013) are also expressed during gynoecium
initiation. In total, over 60 genes are expressed at stage 4 and over 65
genes at stage 6 (Table S1) (Herrera-Ubaldo et al., 2018 preprint;
Jiao and Meyerowitz, 2010). Current studies are revealing novel
roles for well-known regulators; additionally, previously unreported
regulators and interactions are continuously being discovered.
Therefore, we expect that these GRNs will expand in the future.
This expansion of GRNs also applies to the GRNs described for the
subsequent stages of gynoecium and fruit development.
Gynoecium domains, medial and lateral (stages 7 and 8)
The transition from a gynoecium primordium into two, well-defined
domains (lateral and medial; see Glossary, Box 1) marks the
beginning of patterning and the formation of internal tissues
(Fig. 1G; Fig. 3). During stages 7 and 8, the activation and
maintenance of the carpel margin meristem (CMM; see Glossary,
Box 1) is crucial for the subsequent development of specialized
structures: in the lateral domains, valves (see Glossary, Box 1)
develop to protect the ovules, and seeds are subsequently formed in
the medial domain. Additional tissues, such as the septum and the
transmitting tract (see Glossary, Box 1), also form in the medial
domain to facilitate fertilization (Fig. 1G).
The process of gynoecium differentiation into medial and lateral
domains requires several transcription factors and hormonal signals
(Reyes-Olalde and de Folter, 2019; Reyes-Olalde et al., 2013). One
important regulatory module involves the integration of auxin and
cytokinin signaling with the activity of the transcription factors
HECs and SPT in the medial domain around floral stages 7 and 8
(Müller et al., 2017; Reyes-Olalde et al., 2017; Schuster et al., 2015)
(Fig. 3C). To summarize, auxin signaling is present in the lateral
domains and cytokinin signaling in the medial domain in the
CMM. In the medial domain, cytokinin signaling induces auxin
biosynthesis and auxin transporters to create an auxin flux to
transport auxin to the lateral domain. Auxin-induced proteins in the
lateral domain repress cytokinin signaling genes to limit them to the
medial domain, such as the cytokinin signaling inhibitor
ARABIDOPSIS HISTIDINE PHOSPHOTRANSFER PROTEIN
6 (AHP6) and the auxin response factor ETT (Heisler et al., 2001;
Reyes-Olalde et al., 2017). The created auxin flux is important for
the apical-basal growth of the gynoecium. Cytokinin biosynthesis in
Stage 6
Stage 7 Stage 8 Stage 9 Stage 10 Stage 11 Stage 12
Gynoecium
primordium Lateral domain
Medial domain
Carpel margin
meristem (CMM)
Val v es
You n g
replum
Ovule
primordium
Medial
region Ovules
Septum
Repl um
Transmitting tract
Val v es
EF
35
12
IM
G Gynoecium development
C Floral budB Inflorescence
Stage 3
Stage 4
Stage 6
Stage 2
Stage 5
Stage 4
Sepal
Petal
Stage 6 Stage 13Stage 4Stage 3
Sepals
Peta ls
Stamen
Gynoecium
Key
tage 4
D Mature flower
Stamen
Gynoecium
A A. thaliana
Fig. 1. Overview of gynoecium development in Arabidopsis.(A-F) Once the reproductive phase has initiated, the plant Arabidopsis thaliana (Col-0)
(A) produces an inflorescence with an inflorescence meristem (IM) at the apical tip (top view; B), and the IM produces on its flanks floral buds with floral meristems
(FMs) (B,C,E). The FM gives rise to all the floral organs present in a mature flower: sepals, petals and stamens, with the gynoecium at the center (D,F).
(E) Schematic of an IM and the initial stages of flower development (longitudinal view). Numbers indicate different stages of FMs. At floral stage 3, sepal primordia
become visible and at stage 6 the primordia of petal, stamens and gynoecium are visible. (G) Transverse gynoecia sections showing key tissues and regions
during gynoecium development at stages 6-12. Scale bars: 1 cm (A); 25 µm (B); 40 µm (C); 200 µm (D). Images in B and C are taken from Zuniga-Mayo et al.
(2019). Image in D is from Zuniga-Mayo et al. (2012).
3
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
the CMM is assumed to be regulated by SHOOT MERISTEMLESS
(STM), as has been shown for other meristems (Jasinski et al.,
2005; Yanai et al., 2005). STM is a KNOTTED (KNAT)1-LIKE
homeodomain transcription factor important for CMM formation
and activity (Scofield et al., 2007). The STM gene is activated by the
CUP-SHAPED COTYLEDON proteins (CUCs) (Kamiuchi et al.,
2014). A positive regulatory loop between STM and CUC genes has
been reported in the SAM and could be present in the CMM as well
(Spinelli et al., 2011). The CUC transcription factors seem to also be
important for SPT expression in the basal part of the gynoecium
(Nahar et al., 2012).
The current GRN that controls activities in the CMM has been
derived from functional studies of genes and reporter lines
(reviewed by Reyes-Olalde and de Folter, 2019; Reyes-Olalde
et al., 2013). The study and integration of additional regulators is
currently in progress; there are more than 50 transcription factors
related to gynoecium development with reported expression in this
region (Table S1) (Herrera-Ubaldo et al., 2018 preprint). The
coordination of biochemical and genetic processes underlying
meristematic activity in the CMM, and its further differentiation,
probably involves the participation of many proteins and pathways
that are yet to be characterized (Kivivirta et al., 2021; Villarino et al.,
2016; Wynn et al., 2011).
Ovule initiation and patterning (stages 9 and 10)
One of the most important functions of the gynoecium is to provide
protection to the developing ovules. The determination of ovule
identity occurs at the flanks of the CMM. Ovule primordia are
formed at stage 9 by periclinal cell divisions within the epidermal
tissue of the placenta (see Glossary, Box 1) (Fig. 1G; Fig. 4). Like
previous examples, the GRNs involve the coordinated action of
several transcription factors and hormonal signaling pathways that
control the main aspects of ovule development: the establishment of
boundaries and primordia (Fig. 4A), the control of ovule initiation
and number, and ovule patterning (Fig. 4B,C) (Barro-Trastoy et al.,
2020b; Cucinotta et al., 2014, 2020).
Regulation of ovule initiation
The intricate mechanisms that guide ovule initiation are far from
fully understood, but advances have been made (Barro-Trastoy
et al., 2020b; Cucinotta et al., 2020). Indeed, the identification of
enzymes and proteins that act at different levels outside the GRN has
shed light on fine aspects of the molecular mechanisms, and opens
new questions and directions regarding our understanding of ovule
development. For example, factors have been recently identified that
control the spacing of ovules: two secreted peptides and their
ERECTA (ER) family receptor kinases coordinate regular ovule
primordia initiation coupled to fruit growth from the placenta and
carpel walls (Kawamoto et al., 2020).
Processes that occur at the cell-wall level have also been
discovered. A recent study has reported on the involvement of
placenta-expressed genes encoding CELL WALL INVERTASE
(CWIN) 2 and 4 as positive regulators of ovule initiation. CWINs
hydrolyze sucrose into glucose and fructose, and have additional
roles in sugar signaling. Specific suppression of the activity of these
enzymes leads to a significant reduction in ovule number, as well as
ovule and seed abortion. Based on transcriptome analysis in
CWIN2- and CWIN4-silenced lines, it has been shown that the
ovule-identity gene STK, as well as auxin signaling-related genes
and genes encoding hexose transporters are downregulated (Liao
et al., 2020).
Also at the membrane level, the localization of PIN auxin
transporters in the placenta epidermis is crucial for ovule initiation
(Fig. 2A). The distribution of PIN1 and auxin fluxes is very
dynamic in the placenta and affects not only one ovule primordium,
but also the neighboring ovule primordium. Analysis of marker
lines and auxin response has revealed that ovules initiate
asynchronously and follow this trend through later stages (Yu
et al., 2020).
A Stage 3: FM maintenance and termination
B Stage 6: Gynoecium initiation
HEN1 DCL1
LFY+WUS
TPL+HDA19+AP2
ATX 1
ANT/AIL6
PcG RBL SQN ULT
HUA1/2
HEN2/4
PAN
AG
Proliferation
Auxin
distribution
WUS CLV3
YUC1/4 SUP
AG
KNU
CLV3
ANT
SP
T
ETT
WUS AHK4 LOG
3/4/7 STM IPT3/5/7
type-A
ARRs ARR1/10/12
SHP2
CRC
Other
type-B
ARRs
DRNL
AHP6
PIN7
PIN3 TRN2
Cytokinin
signaling
Auxin
maxima
Gynoecium
development
YUC1/4
miR172
Fig. 2. Floral meristem termination and gynoecium establishment (stages
3-6). (A) A gene regulatory network for floral meristem (FM) maintenance and
termination at stage 3. (B) AGAMOUS (AG) orchestrates gynoecium identity
establishment by integrating hormonal signals with the activation/repression of
transcription factors (Xu et al., 2019; Zuniga-Mayo et al., 2019). Dashed lines
indicate indirect regulation. Colored words indicate different hormone-related
processes: auxin signaling (blue); cytokinin signaling (green). AHK4,
ARABIDOPSIS HISTIDINE PROTEIN KINASE 4; AHP6, HISTIDINE
PHOSPHOTRANSFER PROTEIN 6; AIL6, AINTEGUMENTA-LIKE6; ANT,
AINTEGUMENTA; AP2, APETALA2; ARRs, ARABIDOPSIS RESPONSE
REGULATORs; ARR1/10/12, RESPONSE REGULATOR 1, 10, 12; ATX1,
ARABIDOPSIS HOMOLOG OF TRITHORAX 1; CLV3, CLAVATA3; CRC,
CRABS CLAW; DCL1, DICER LIKE1; DRNL, DORNROSCHEN-LIKE; ETT,
ETTIN; HEN1/2/4, HUA ENHANCER 1, 2, 4; HDA19, HISTONE
DEACETYLASE 19; HUA1/2, HUA1, 2 (At5g23150); IPT3/5/7,
ISOPENTENYLTRANSFERASE 3, 5, 7; KNU, KNUCKLES; LFY, LEAFY;
LOG3/4/7, LONELY GUY 3, 4, 7; miR172, microRNA172; PAN, PERIANTHIA;
PcG, Polycomb-group; PIN3/7, PIN-FORMED 3, 7; RBL, REBELOTE; SHP2,
SHATTERPROOF 2; SPT, SPATULA; SQN, SQUINT; STM, SHOOT
MERISTEMLESS; SUP, SUPERMAN; TPL, TOPLESS; TRN2, TORNADO 2;
ULT, ULTRAPETALA; WUS, WUSCHEL; YUC1/4, YUCCA 1, 4.
4
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
Ovule number
On average around 60 ovules are formed inside the gynoecium
(reviewed by Barro-Trastoy et al., 2020b; Cucinotta et al., 2014,
2020; Yuan and Kessler, 2019). However, ovule numbers can range
between 40 and 80, depending on the accession of Arabidopsis
(Yuan and Kessler, 2019). Some modules of the GRN that control
ovule number can also control gynoecium size (reviewed by
Cucinotta et al., 2020). A recent study has uncovered a previously
unidentified positive regulator of ovule number: NEW
ENHANCER OF ROOT DWARFISM1 (NERD1) (Yuan and
Kessler, 2019).
Transcription factors, such as MP, ANT and CUCs, are
interconnected with hormone homeostasis mechanisms, mainly
with auxins, cytokinins, brassinosteroids and gibberellins, to
regulate ovule number (Fig. 4A) (reviewed by Barro-Trastoy
et al., 2020b; Cucinotta et al., 2020; Qadir et al., 2021). As
mentioned above, auxin positively regulates ovule primordia
initiation and cytokinin signaling positively regulates ovule
number (Bartrina et al., 2011; Cerbantez-Bueno et al., 2020;
Cucinotta et al., 2018; Galbiati et al., 2013; Reyes-Olalde et al.,
2017; Zúñiga-Mayo et al., 2018). It has now been shown that
cytokinin metabolism in the epidermis of the placenta is also
important for ovule number (Werner et al., 2021). Furthermore,
brassinosteroids and gibberellins positively and negatively regulate
ovule number, respectively (Barro-Trastoy et al., 2020a;
Huang et al., 2013; Nole-Wilson et al., 2010; Gomez et al., 2018,
2019).
Ovule patterning
After the establishment of ovule identity and ovule number
definition (Fig. 4A), ovule patterning takes place, which involves
differentiation of the funiculus, chalaza and nucellus (see Glossary,
Box 1) at stage 10 (Fig. 4B), and continues with the formation of the
inner and outer integuments at stage 11 (Fig. 4C). Various recent
reviews on these stages are available (Barro-Trastoy et al., 2020b;
Pinto et al., 2019), and an in-depth discussion of this topic goes
beyond the scope of this article. However, one important
transcription factor, STK, is the master regulator of ovule identity
(Favaro et al., 2003; Pinyopich et al., 2003), although this gene has
many additional functions, as described below. Various other genes
involved in ovule patterning have additional functions during
gynoecium development.
Gynoecium patterning (stages 11 and 12)
During stages 11 and 12, gynoecium patterning continues to form
all the tissues, and the gynoecium attains the final shape suitable for
pollination (Fig. 1D). During this period, there is active patterning
in the main axes: the abaxial-adaxial with funiculus, chalaza
and nucellus (summarized in Fig. 5); the medial-lateral with the
valves, valve margins, replum (see Glossary, Box 1), septum and
transmitting tract (Fig. 5A); and the apical-basal with stigma (see
Glossary, Box 1), style (see Glossary, Box 1), ovary and gynophore
(Fig. 5B) (Chávez Montes et al., 2015; Deb et al., 2018; Marsch-
Martínez and de Folter, 2016; Simonini and Østergaard, 2019;
Zúñiga-Mayo et al., 2019). Most of the best-characterized regulators
in these GRNs are transcription factors that affect tissue
differentiation.
Medial-lateral patterning
In a mature gynoecium, the medial-lateral axis is composed of the
valves (the carpel walls), which protect the developing ovules. After
pollination, the valves are attached to the replum (Fig. 1F,G;
Fig. 5A). The formation of this axis (valves-replum-valves) requires
the concerted action of medial factors, such as BP, RPL,
WUSCHEL-RELATED HOMEOBOX 13 (WOX13) and/or NO
TRASMITTING TRACT (NTT), which possess meristematic-
related functions and repress the action of lateral factors, such as
FILAMENTOUS FLOWER/YABBY3 (FIL/YAB3) and the
ASYMMETRIC LEAVES 1/2 (AS1/2) (Alonso-Cantabrana et al.,
2007; Dinneny et al., 2005; Marsch-Martínez et al., 2014; Romera-
Branchat et al., 2013). Additionally, the so-called boundary
factors(CUCs and KNAT2/6) participate in between, marking
the division of the meristematic and lateral organ functions
(Fig. 5A) (reviewed by Ballester and Ferrándiz, 2017; Simonini
and Østergaard, 2019).
Inside the ovary, transmitting tract formation is controlled by
NTT (Crawford et al., 2007), the basic helix-loop-helix (bHLH)
proteins CESTA/HALF FILLED (CES/HAF), BRASSINOSTEROID
ENHANCED EXPRESSION 1 and 3 (BEE1, BEE3) (Crawford
and Yanofsky, 2011), HECs (Gremski et al., 2007), SPT
Box 2. Protein interactions: a combinatorial model for
gynoecium patterning?
The concerted action of transcription factors (MADS-box and
APETALA 2), to guide the establishment of floral organ identity is
elegantly represented by the ABCDEmodel (Theißen and Saedler,
2001). For gynoecium patterning, the involvement of numerous
transcription factor families surpasses those for floral organ
specification. The crucial action and effect of transcription factors may
suggest a combinatorial model for gynoecium patterning; some crucial
transcription factors [i.e. SEUSS (SEU) and LEUNIG (LUG)] may
dimerize to act as a chassis for higher-order complexes in the early
stages (Azhakanandam et al., 2008). In terms of protein interactions, the
best-studied tissues are the stigma and style, where transcription factors,
such as INDEHISCENT (IND), SPATULA (SPT) and NGATHA (NGA)
participate (Ballester et al., 2021; Simonini et al., 2018).
We have made efforts to find such protein complexes. The
identification of physical interactions between several transcription
factors in combination with gene expression and functional information
is useful to represent the networks and interaction dynamics during the
development of the gynoecium (Herrera-Ubaldo et al., 2018 preprint).
Figure shows protein-protein interaction networks at two early stages of
gynoecium development.
The involvement of thesetranscription factors is just the beginning of the
story;previous workshave extendedregulatorynetworks beyondthe range
of transcription factors alone.For example, AGAMOUS (AG) interacts with
histone deacetylases and other proteins (Smaczniak et al., 2012).
Stage 7: Lateral domain
Stage 7: Medial domain
Stage 6:
Gynoecium primordium
5
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
(Heisler et al., 2001) and STK (Di Marzo et al., 2020b; Herrera-
Ubaldo et al., 2019) (Fig. 5A). In addition, several of these factors
(NTT, SPT and STK) control the development of other tissues in the
gynoecium that are related to transmitting tract formation; they
control programmed cell death (PCD) and the deposition of
extracellular matrix (ECM) (reviewed by Crawford and Yanofsky,
2008; Pereira et al., 2021).
Some processes occurring between stages 11 and 12 highlight the
importance of cell wall modifications in medial-lateral gynoecium
patterning. The leading role of ETT in gynoecium morphology is
clear; ETT is known to participate, in coordination with auxin input,
in the guidance of early patterning events (Fig. 2B; Fig. 3) (Sessions
et al., 1997; Simonini et al., 2016). Now, an additional pathway of
ETT participation has been identified, the regulation of pectin
methylesterase (PME) activity in the valves (Andres-Robin et al.,
2018). PME activity in valve cell walls increases the levels of
demethylesterified pectins, allowing the reduction of cell wall
stiffness and the elongation of the valves.
Inside the gynoecium, other modifications need to occur in the
medial domain to allow the formation of the transmitting tract.
Recently, NTT and STK have been shown to participate in the
modulation of the cell wall composition by regulating a gene
encoding a putative mannanase enzyme, as well as other genes
encoding proteins involved in lipid biosynthesis and transport
(Herrera-Ubaldo et al., 2019).
Apical-basal patterning
Major changes must be coordinated to achieve the distinct apical-
basal features (stigma, style and ovary) observed in a mature
gynoecium. The ovary is a bilateral structure in contrast to the
style, which exhibits radial symmetry (Fig. 1F,G). During floral
stage 12, the style differentiates and the bilateral-to-radial transition
is controlled by the modulation of auxin flux, as well as cytokinin
sensitivity (Fig. 5B) (Carabelli et al., 2021; Moubayidin and
Ostergaard, 2014). First, the transcription factors SPT and
INDEHISCENT (IND) promote apolar localization of PIN
proteins (auxin efflux transporters normally localized to cell poles
to create polar auxin transport; Moubayidin and Ostergaard, 2014).
SPT and HECs then control the expression of the adaxial-identity
genes HOMEOBOX ARABIDOPSIS THALIANA 3 (HAT3) and
ARABIDOPSIS THALIANA HOMEOBOX 4 (ATHB4), which
orchestrate the final steps of the radicalization process (Fig. 5B)
(Carabelli et al., 2021). In another study focusing on style
development using genetics and protein-protein interaction
experiments, it has been suggested that ETT, IND,
BREVIPEDICELLUS (BP), REPLUMLESS (RPL) and SEUSS
(SEU) work synergistically to regulate style morphology (Fig. 5B)
(Simonini et al., 2018).
Sequential activation and function of transcription factors is
required for stigma development: the NGATHA (NGA) and
HECATE (HEC) protein families cooperatively regulate stigma
A Stage 7 B Stage 8 C
Medial
domain
Lateral
domain
Lateral
domain
STM IPTs
AHP6
DRNL
MP
ANT
HEC1
ETT
ARR-A
Auxin
TAA 1
Auxin
Auxin
flux
PIN1
CUCs
PINs
Cytokinin
SPT ARR-B
Lateral domain
Medial domain
Carpel margin meristem
Key
Fig. 3. Carpel margin meristem activation and maintenance (stages 7 and 8). (A) Schematic of a young gynoecium at stage 7, when two different domains
lateral (green) and medial (yellow) are visible. (B) At stage 8, the carpel margin meristem (red) becomes specified. (C) The activity of the CMM is maintained by
the activity of the bHLH transcription factors SPATULA (SPT) and HECATE 1 (HEC1), via the activation of auxin synthesis and transport. Genes related to
cytokinin signaling are also involved (Reyes-Olalde and de Folter,2019). Upper panels represent longitudinal sections of the dashed lines in the bottom panels of
the gynoecium images. Dashed lines in the regulatory network indicate indirect regulation, or interactions found in other tissues that are likely occurring in the
gynoecium. Colored words indicate different hormone-related processes: auxin signaling (blue); cytokinin signaling (green). Purple lines in B and C indicate the
regions in which the processes occur. AHP6, HISTIDINE PHOSPHOTRANSFER PROTEIN 6; ANT, AINTEGUMENTA; ARRs, ARABIDOPSIS RESPONSE
REGULATORs; CUCs, CUP-SHAPED COTYLEDONs; DRNL, DORNROSCHEN-LIKE; ETT, ETTIN; HEC1, HECATE 1; IPTs,
ISOPENTENYLTRANSFERASEs; MP, MONOPTEROS; PIN1, PIN-FORMED 1; SPT, SPATULA; STM, SHOOT MERISTEMLESS; TAA1, TRYPTOPHAN
AMINOTRANSFERASE OF ARABIDOPSIS 1.
6
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
development through the activation of IND, which later interacts
with NGA (and probably HECs) to activate SPT. SPT is then
probably integrated into the NGA-IND-HEC complex to activate
target genes (Ballester et al., 2021). Recently, three angiosperm-
specific regulators of the style and stigma have been identified
called STIGMA AND STYLE STYLIST 1-3 (SSS1-3), which
belong to a previously unreported family. They are expressed in the
apical tissues of the gynoecium and act downstream of the NGA
transcription factors to fine-tune the development of the stigma and
style (Fig. 5B) (Li et al., 2020).
Pollination and fertilization (stage 13)
Pollination and fertilization involve a continuous and active
communication between female and male tissues over several
steps: pollen hydration, pollen germination, pollen tube growth,
pollen tube attraction to the ovule, pollen tube reception, sperm cell
delivery and gamete activation (reviewed by Cascallares et al.,
2020).
As discussed earlier, the formation of the transmitting tract
includes regulation of PCD and ECM deposition (reviewed by
Pereira et al., 2021), as well as other changes in cell wall
composition in the medial region (Herrera-Ubaldo and de Folter,
2018). The ECM provides nutrition, adhesion and guidance during
pollen tube growth. During fertilization, the transmitting tract
secrets chemical signals, such as arabinogalactan-proteins (AGPs),
whereas other signals, such as LURE proteins, derive from the
ovules to attract pollen tubes (reviewed by Johnson et al., 2019;
Pereira et al., 2021).
Checkpoints during the pollination process control the
acceptance or rejection of the pollen prior to fertilization
(reviewed by Dresselhaus et al., 2016) and various small peptides
and receptors are involved in these processes (reviewed by Zhang
et al., 2021). Recently, some new discoveries have been made
towards the understanding of communication and signaling
cascades during pollination. One such discovery is the existence
of an autocrine signaling pathway acting at the surface of stigmatic
papillae, inducing reactive oxygen species (ROS) production; ROS
levels are reduced upon pollination owing to an antagonistic peptide
from the pollen coat, allowing pollen hydration and germination
(Liu et al., 2021; Zhou et al., 2021). Another discovery also involves
Nucellus
Chalaza
Funiculus
C Stage 11B Stage 10A Stage 9
Ovule
primordium
Boundary
SPL
REV
INO
SUP
BEL1
PHB
PHV
CNA
ETT+ATS
Auxin
UCN
Brassinosteroids
Gibberellins
GID1
PHB
miR166
INO
BRI1
BZR1
WUS
ATS+GAI
RGA
STK
PIN1
Cytokinin
WUS
AHK2/3/4
WUS
SPL
ANT
BEL1
PIN1
Auxin
maximum
CKX3/5
GibberellinsCytokinin
SAUR8/10/12
CYP85A2
DET2
BRI1
Brassinosteroids
BIN2
BZR1
ARGOS
ANT
CUC1/2
PIN1
miR164
MP
YUC1
Auxin
YUC4
UGTs
AHK2/3/4
CRF2
GAI
RGA
RGL2
GID1
Ovule number
Ovule primordium, ovuleTransmitting tract
Val v e
SeptumRepl um
Key
Fig. 4. Ovule identity specification and patterning (stages 9-11). (A-C) The beginning of ovule development (determination of number and position) is
controlled by a complex network that involves the action of at least four hormones and several transcription factors. The regulatory networks controlling ovule
primordia initiation (A), patterning (B) and morphogenesis (C) (Barro-Trastoy et al., 2020b; Cucinotta et al., 2020) are summarized. Colored words indicate
different hormone-related processes: auxin signaling (blue); cytokinin signaling (green); gibberellins (purple); brassinosteroids (orange). AHK2/3/4,
ARABIDOPSIS HISTIDINE PROTEIN KINASE 2, 3, 4; ANT, AINTEGUMENTA; ARGOS, AUXIN-REGULATED GENE INVOLVED IN ORGAN SIZE; ATS,
ABERRANT TESTA SHAPE; BEL1, BELL 1; BIN2, BRASSINOSTEROID-INSENSITIVE 2; BRI1, BRASSINOSTEROID INSENSITIVE 1; BZR1,
BRASSINAZOLE RESISTANT 1; CKX3/5, CYTOKININ OXIDASE/DEHYDROGENASE 3, 5; CNA, CORONA; CRF2, CYTOKININ RESPONSE FACTOR 2;
CUC1/2, CUP-SHAPED COTYLEDON 1, 2; CYP85A2, cytochrome p450 enzyme; DET2, ATDET2; ETT, ETTIN; GAI, GIBBERELLIC ACID INSENSITIVE;
GID1, GA INSENSITIVE DWARF; INO, INNER NO OUTER; miR164, microRNA164; miR166, microRNA166; MP, MONOPTEROS; PHB, PHABULOSA; PHV,
PHAVOLUTA; PIN1, PIN-FORMED 1; REV, REVOLUTA; RGA, REPRESSOR OF GA; RGL2, RGA-LIKE 2; SAUR8/10/12, SAUR-like auxin-responsive protein
family; SPL, SPOROCYTELESS; STK, SEEDSTICK; SUP, SUPERMAN; UCN, UNICORN; UGTs, UDP-glucosyl transferases; WUS, WUSCHEL; YUC1/4,
YUCCA 1, 4.
7
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
the receptor kinase FERONIA (FER) in the control of male-
female interactions. FER, located at the entrance to the female
gametophyte, mediates de-esterified pectins and the production of
nitric oxide (NO), which is necessary for when the first pollen tube
arrives at the ovule; NO accumulation stops the attraction of further
pollen tubes (Duan et al., 2020). The process of pollination, and the
subsequent events, can only occur in a specific time window,
delimited by stigmatic receptivity. Two transcription factors, KIRA1
(KIR1) and ORESARA1 (ORE1), belonging to the NAC family
promote PCD in the stigma cells, thereby controlling stigma
longevity (Gao et al., 2018).
Seed and fruit development (stages 14-20)
After fertilization, the gynoecium becomes a fruit, and its
development in Arabidopsis continues over 12 days to finally
release fully formed, fertile seeds. During the final stages of fruit
development, the fruit grows, the seeds mature and developmental
programs prepare for fruit dehiscence (see Glossary, Box 1) and
seed dispersal (Alvarez-Buylla et al., 2010; Ballester and Ferrándiz,
2017; Roeder and Yanofsky, 2006). After fertilization (stage 13),
the fruit grows mainly longitudinally to reach its maximum size by
stage 17, followed by the start of senescence of the fruit (stage 18),
and finally fruit dehiscence and seed release (stage 20) (Fig. 6A-C).
Longitudinal growth
The involvement of cytokinin in fruit elongation has recently been
reported: phenotypic alterations observed in fruit length of the stk
mutant, which produces short fruits, resemble those of fruits of the
ckx7 mutant (encoding for the enzyme CYTOKININ OXIDASE/
DEHYDROGENASE 7), in which cytokinin degradation is affected
(Di Marzo et al., 2020a). Therefore, cytokinin has a negative effect
on fruit elongation. This additional function of STK also integrates
pathways that control fruit size because STK positively regulates
CKX7 expression in the fruit and indirectly controls FUL, which is
A Stage 12: Medial-lateral patterning B Stage 12: Apical-basal patterning
ARF6/8SPT+HECs STY NTT
BEEs+HAF
Tra n sm it ti n g tr a ct f or m at io n
Repl um Septum Va lve OvuleTransmitting tract Stigma StyleGynophore
Key
AS1/2 KNAT2/6
NTT
ARF4 ETT
KAN AP2
BP
JAG FIL YAB3
FUL CUC1/2RPL
STM
SHP1/2
Replu m
Val ve
margin
Val ve s LL SL
SPT
IND ALC
ETT
HEC1+SPT
Cytokinin
TCP15 AHP6
Auxin
Medial region
development
Stigma
and
style
formation
ARF6/8
CRC
SSSs
HAT3,
ATH B 4 SPT+HECs
Cytokinin
SHP1/2
ANT+FIL
SEU+LUG
Auxin
maxima PINs
PID
(P)
Auxin YUCs SHI/STYs
NGAs
ETT,
SPT+IND
HECs SPLs TCP15
Fig. 5. Medial-lateral and apical-basal gynoecium patterning fertilization (stage 12). (A) In the outer part of the gynoecium, lateral and medial factors
cooperate or compete to define the valves, valve margins and replum. The valve margin is divided in the lignification (LL) and separation layer (SL). The formation
of the transmitting tract is coordinated by several transcription factors. Dashed line represents the transverse section depicted in B. (B) In the apical-basal axis,
stigma and style formation is guided by regulation that converges in the production of auxins. Cooperation between transcription factors and hormones
participates in maintaining the ovary (Alonso-Cantabrana et al., 2007; Deb et al., 2018; Marsch-Martı
nez and de Folter, 2016; Simonini and Østergaard, 2019).
Colored words indicate different hormone-related-processes: auxin signaling (blue); cytokinin signaling (green). (P)indicates that PID phosphorylates PIN
proteins. AHP6, HISTIDINE PHOSPHOTRANSFER PROTEIN 6; ALC, ALCATRAZ; ANT, AINTEGUMENTA; AP2, APETALA 2; ARF4/6/8, AUXIN RESPONSE
FACTOR 4, 6, 8; AS1/2, ASYMMETRIC LEAVES 1, 2; ATHB4, ARABIDOPSIS THALIANA HOMEOBOX 4; BEEs, BR-ENHANCED EXPRESSIONS; BP,
BREVIPEDICELLUS; CRC, CRABS CLAW; CUC1/2, CUP-SHAPED COTYLEDON 1, 2; ETT, ETTIN; FIL, FILAMENTOUS FLOWER; FUL, FRUITFULL; HAF,
HALF FILLED; HAT3, HOMEOBOX ARABIDOPSIS THALIANA 3; HECs, HECATEs; IND, INDEHISCENT; JAG, JAGGED; KAN, KANADI; KNAT2/6,
KNOTTED-LIKE FROM ARABIDOPSIS THALIANA 2, 6; LUG, LEUNIG; NGAs, NGATHAs; NTT, NO TRANSMITTING TRACT; PID, PINOID; PINs, PIN-
FORMED; RPL, REPLUMLESS; SEU, SEUSS; SHI, SHORT INTERNODES; SHP1/2, SHATTERPROOF 1, 2; SPLs, SQUAMOSA PROMOTER BINDING
PROTEIN-LIKEs; SPT, SPATULA; SSSs, STIGMA AND STYLE STYLISTs; STM, SHOOT MERISTEMLESS; STYs, STYLISHs; TCP15, TEOSINTE
BRANCHED, CYCLOIDEA/PCF 15; YAB3, YABBY 3; YUCs, YUCCAs.
8
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
involved in a pathway with miR172 and AP2 (Fig. 6E) (Di Marzo
et al., 2020a; Ripoll et al., 2015). Additionally, STK regulates fruit
growth by controlling the expression of α-XYLOSIDASE 1 (XYL1),
which is involved in the modification of xyloglucans (XyGs),
which form a component of cell walls. Indeed, XyG modifications
contribute to the mechanical properties of the cell wall, affecting its
extensibility and thus cell growth. The activity of XYL1 is also
required in the growing seeds in order for them to achieve their final
size (Di Marzo et al., 2021).
The involvement of hormones in fruit development and
maturation goes beyond just auxin and cytokinin. Gibberellins
induce fruit growth in Arabidopsis (Dorcey et al., 2009) and
regulate the formation of the separation layer in the valve margins
(Fig. 6D,F) (Arnaud et al., 2010). The role of hormones, such as
abscisic acid, ethylene and brassinosteroids, are better described for
fleshy fruit development (reviewed by Li et al., 2021; McAtee et al.,
2013; Sotelo-Silveira et al., 2014).
Related to longitudinal growth of the silique (see Glossary,
Box 1) or fruit, a recent study has reported a spatiotemporal map of
fruit growth at the cellular level that captures quantitative data to
measure cell division and cell expansion. Analysis of the ntt mutant
affected in seed-set revealed that final fruit length correlates with the
number of seeds (Ripoll et al., 2019).
Preparation for dehiscence
Early studies on fruit development have described the initial models
for fruit patterning related to fruit opening (dehiscence or pod
shattering), highlighting the key roles of the transcription factors
FUL, SHPs, RPL, ALC and IND (Ballester and Ferrándiz, 2017;
Liljegren et al., 2004). The dehiscence zone, located on the valve
margins, are specialized cell layers that allow valve detachment.
SHP proteins control the expression of IND, which guides the
formation of the lignification layer (Fig. 6D,F). Additionally, SHPs
control ALCATRAZ (ALC), which guides the formation of the
separation layer (reviewed by Ballester and Ferrándiz, 2017).
Moreover, FUL controls the formation of a lignified layer in the
valves (Liljegren et al., 2004).
Another role of STK is the control of lignification in the
funiculus, which is necessary for seed abscission (Balanzà et al.,
2016). In the funiculus, STK, together with the co-repressor SEU,
represses the expression of HEC3,SPT and ALC (Fig. 6F). A rather
similar process occurs in valve margin lignification, whereby SHPs
regulate IND,SPT and ALC, but the regulation is inverted (i.e. SHPs
activate gene expression) (Fig. 5A; Fig. 6F).
Important roles of hormones have been uncovered as well. IND
regulates the expression of PINOID (PID) and WAG2 to control
auxin transport outside the separation layer in the valve margins,
creating a so-called auxin minimumthat is important for dehiscence
zone formation (Sorefan et al., 2009). However, further studies
regarding the participation of auxin in the control of dehiscence,
using reporter lines to track auxin distribution have revealed that, at
early fruit-development stages (stage 14-16), auxin is present at the
valve margin for specifying the dehiscence zone (van Gelderen et al.,
2016). Perhaps the auxin minimumat later fruit stages (e.g. stage
17B) is still functionally relevant for cell separation (reviewed by
Ballester and Ferrándiz, 2017; de Folter, 2016). The involvement of
cytokinin signaling during fruit development has also been studied.
The valve margins display cytokinin signaling, which is lost in
mutants that are indehiscent (ind and shp1 shp2). When cytokinin is
applied to these mutants, fruit dehiscence is restored (Marsch-
Martínez et al., 2012). Furthermore, replum development is also
dependent on the presence of cytokinin; with less cytokinin or
Repl um
A Fruit development
B Stage 13 C Stage 17B
Ovule
Septum
Val v e
Seed
Septum
Repl um
Val v e
Val v e
margin
Stage
13 14 15 16 17A 17B 18 19 20
Lignified valve layer Lignified margin layer Separation layer
SeptumVal v e Replum
Key
EFG Dehiscence
ARF6/8 FUL STK
miR172C IND CKX7
Cytokinin
degradation
AP2
cis-Zeatin
trans-Zeatin
Cell
elongation
Fruit
elongation
STK+SEU
HEC3
SPT
ALC
Seed abscission
zone
IND
SHPs
SPT
ALC
Val ve
margins
SHPs
IND
GA3ox1
Gibberellins Cytokinin
Auxin
depletion
PIN3
WAG 2
PID
Val ve
margins
D
Fig. 6. Fruit growth and dehiscence (stages 14-20). (A) After fertilization, the
gynoecium becomes a fruit. During stages 14-19, the fruit grows and prepares to
release the seeds. (B-D) Internally, the gynoecium-to-fruit transition involves
changes in tissue conformation. One of the most relevant isthe formation of the
seeds, as well as the differentiation of the lignification (LL) and separation (SL)
layers in the valve margins prior to dehiscence. D shows a detailed view of the
boxed area in C. (E-G) At the molecular level, fruit elongation relies on cell
elongation (E) and dehiscence (F), which are controlled by transcription factors
with hormonal input. (G) The specification of the seed abscission zone and the
valve margins share someregulatory elements, such as SEEDSTICK (STK) and
SHATTERPROOFs (SHPs). Dashed lines indicate indirect regulation. Colored
words indicate different hormone-related processes: auxin signaling (blue);
cytokinin signaling (green); gibberellins (purple) (Balanzaet al., 2016; Di Marzo
et al., 2020a; Marsch-Martı
nez and de Folter, 2016; van Gelderen et al., 2016).
ALC, ALCATRAZ; AP2, APETALA2; ARF6/8, AUXIN RESPONSE FACTOR 6,
8; CKX7, CYTOKININ OXIDASE/DEHYDROGENASE 7; FUL, FRUITFULL;
GA3ox1, GIBBERELLIN 3-beta-dioxygenase 1; HEC3, HECATE 3;
IND, INDEHISCENT; miR172C, microRNA172C; PID, PINOID; PIN3,
PIN-FORMED3; SPT, SPATULA; SEU, SEUSS; WAG2, WAG2.
9
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
cytokinin signaling, the width of the replum is reduced and vice versa
(Marsch-Martínez et al., 2012; Reyes-Olalde et al., 2017). In a double
Arabidopsis ntt rpl mutant that lacks the replum, replum formation is
restored upon cytokinin application, suggesting that cytokinin
signaling acts downstream of NTT and RPL (Zúñiga-Mayo et al.,
2018). In addition, increased replum width has also been observed
when NTT is overexpressed (Marsch-Martínez et al., 2014).
In addition to the hormonal and genetic control of dehiscence, a
recent study adds the importance of environmental input.
Temperature affects the expression of IND; higher temperature
causes chromatin modifications at the IND locus, which results in
increased IND expression that accelerates valve margin
development (Fig. 6G) (Li et al., 2018).
Gene expression: additional nodes in the GRNs?
The concerted action of GRNs during these stages will ultimately
guide fruit growth and shape. Although a lot of information is
available, further players are likely still to be discovered. A recent
study has conducted transcriptome analysis to reveal gene
expression dynamics during fruit growth and maturation,
including silique samples, from 3, 6, 9 and 12 days after
pollination (Mizzotti et al., 2018). This dataset is supported by
another study, which has analyzed the pre-fertilization stages: carpel
initiation (stage 5), elongation of carpel walls (stage 9), gynoecia
during female meiosis (stage 11), and gynoecia before anthesis
(stage 12) (Kivivirta et al., 2021). These two studies (and others,
e.g. Klepikova et al., 2016; Villarino et al., 2016; Wynn et al., 2011)
now provide a broad gene expression atlas for the complete
developmental window from gynoecium-fruit development. In
both works, samples included complete gynoecia and fruits, so
additional work is required to resolve gene expression further or
dissect transcriptomic data at the tissue- and single-cell level.
Conclusions
The current knowledge on gynoecium and fruit development is
expanding. We have given an overview of the GRNs known to date.
In the future, it is likely that these GRNs will expand because we
know that many more transcription factors are expressed at the
specific stages discussed (Table S1). Furthermore, many recent
works have provided new datasets and information related to
gynoecium formation, such as transcriptomics during development
(Kivivirta et al., 2021; Mizzotti et al., 2018) or at specific times or
scenarios (Krizek et al., 2021; Liao et al., 2020; Martínez-Fernández
et al., 2020; Yu et al., 2020); variation by genome-wide association
studies (Yuan and Kessler, 2019) or quantitative trait locus analysis
(Kawamoto et al., 2020); and protein-protein interactions (Herrera-
Ubaldo et al., 2018 preprint). The presented GRNs are based on data
of Arabidopsis it is likely that these GRNs have variations and/or
rewiring that mayexplain existing diversity in gynoecia and fruits of
other species (Box 3).
Current technologies allow the study of morphogenesis in
Arabidopsis in unprecedented detail. Live imaging, lineage
tracking and geometric analysis, in combination with single-cell
transcriptome profiling, are changing how we study plants and
paving the way to study development in four dimensions. Some
important advances have been made during early flower
development, such as the construction of a 4D atlas that integrates
cell growth quantification and gene expression data (Refahi et al.,
2021); a 3D gene expression atlas of the FM based on the spatial
reconstruction of single-nucleus RNA-sequencing data (Neumann
et al., 2021 preprint); and a study of sepal (Zhu et al., 2020) and
stamen growth (Silveira et al., 2021). In the case of ovule formation,
the generation of a 3D reference atlas (Vijayan et al., 2021) or the
analysis of the relationship between organ geometry and cell fate
(Hernandez-Lagana et al., 2021) provide comprehensive views of
ovule development.
These works are related to reproductive structures, but the field
can also take advantage of broader studies performed at the whole-
plant level. For example, the protein interactome for hormone-
related proteins (Altmann et al., 2020) or the Arabidopsis proteome
with samples from whole flowers and fruits (as well as carpels,
seeds, valves and septum samples; Mergner et al., 2020), provide
useful resources and valuable information to be integrated.
Additionally, a recent dataset has facilitated the study of protein
complexes at the pan-plant level (McWhite et al., 2020), and
another allows the comparison of transcriptomic programs
across land plants (Julca et al., 2021). Finally, the Plant Cell Atlas
initiative will provide insight into the content and processes taking
place in each cell type in the plant (Plant Cell Atlas Consortium
et al., 2021; Rhee et al., 2019). The integration of all these data inthe
Box 3. Evolution of gene regulatory networks (GRNs)
directing gynoecium development
The study of the origin of the carpel at the morphological level is an active
topic of research (Endress, 2019; Endress and Doyle, 2015; Endress
and Igersheim, 2000; Scutt et al., 2006). Some evidence points to the
tube-like (ascidiate) or leaf-like (plicate) origin of the carpel. The
information from Arabidopsis is the reference for comparative studies.
The identification of key regulators (as individuals or families) and their
conservation across the tree of life and its evolution (Pfannebecker et al.,
2017a,b) has allowed the tracing of the history of the main regulators
(Becker, 2020; Ferrándiz and Fourquin, 2014).
Current work is focused on deciphering the origin of regulators and
their connections in GRNs that guide the variation in the shape and
function of fruits. Two basic aspects of GRN evolution are: (1) the rise of
nodes and connections (see figure, green dots and lines); and
(2) rewiring of the edges (blue lines). Furthermore, variation in cell
division rates or cell expansion may explain shape diversity in related
species (Dong et al., 2020; Łangowski et al., 2016).
The construction of genetic and protein interaction networks that
retrieves proteins of different clades gives an idea of how regulatory
networks gain complexity. Furthermore, information from other fruit
species, such as Solanaceae (Ortiz-Ramírez et al., 2018) and others
(Gomariz-Fernández et al., 2017; Pabón-Mora et al., 2014; Simonini
et al., 2018), could give insights into gynoecium evolution.
Initial network
New nodes and
new interactions
Gene duplication and divergence
10
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
coming years will reveal many hidden aspects of morphogenesis
during gynoecium and fruit formation and will achieve a detailed
picture of the mechanisms involved, at a similar level to other
systems, such as plant embryo patterning (Harnvanichvech et al.,
2021).
Acknowledgements
We apologize to researchers in the field whose work has not been cited owing to
space constraints. We thank Andrea Gomez-Felipe and Victor Zuniga-Mayo for
providing images for Fig. 1B,C and Fig. 1D, respectively. We also thank the two
anonymous reviewers for their input.
Competing interests
The authors declare no competing or financial interests.
Funding
Work in the S.d.F. laboratory is financed by the Mexican National Council of Science
and Technology (Consejo Nacional de Ciencia y Tecnologı
a; CONACYT) grants
(CB-2012-177739, INFR-2015-253504, FC-2015-2/1061 and CB-2017-2018-A1-S-
1012). H.H.-U. acknowledges a postdoc fellowship from CONACYT. S.d.F.
acknowledges the Fundacion Marcos Moshinsky, and also acknowledges
participation in the European Union Horizon 2020 H2020-MSCA-RISE-2020
EVOfruland project (101007738) and H2020-MSCA-RISE-2019 MAD project
(872417).
References
Alonso-Cantabrana, H., Ripoll, J. J., Ochando, I., Vera, A., Ferrandiz, C. and
Martı
nez-Laborda, A. (2007). Common regulatory networks in leaf and fruit
patterning revealed by mutations in the Arabidopsis ASYMMETRIC LEAVES1
gene. Development 134, 2663-2671. doi:10.1242/dev.02864
Altmann, M., Altmann, S., Rodriguez, P. A., Weller, B., Elorduy Vergara, L.,
Palme, J., Marı
n-de la Rosa, N., Sauer, M., Wenig, M., Villaecija-Aguilar, J. A.
et al. (2020). Extensive signal integration by the phytohormone protein network.
Nature 583, 271-276. doi:10.1038/s41586-020-2460-0
Alvarez-Buylla, E. R., Benı
tez, M., Corvera-Poire, A., Chaos Cador, A., de
Folter, S., Gamboa de Buen, A., Garay-Arroyo, A., Garcı
a-Ponce, B., Jaimes-
Miranda, F., Perez-Ruiz, R. V. et al. (2010). Flower development. Arabidopsis
Book 8, e0127. doi:10.1199/tab.0127
Andres, F. and Coupland, G. (2012). The genetic basis of flowering responses to
seasonal cues. Nat. Rev. Genet. 13, 627-639. doi:10.1038/nrg3291
Andres-Robin, A., Reymond, M. C., Dupire, A., Battu, V., Dubrulle, N.,
Mouille, G., Lefebvre, V., Pelloux, J., Boudaoud, A., Traas, J. et al. (2018).
Evidence for the regulation of gynoecium morphogenesis by ETTIN via cell wall
dynamics. Plant Physiol. 178, 1222-1232. doi:10.1104/pp.18.00745
Arnaud, N., Girin, T., Sorefan, K., Fuentes, S., Wood, T. A., Lawrenson, T.,
Sablowski, R. and Østergaard, L. (2010). Gibberellins control fruit patterning in
Arabidopsis thaliana. Genes Dev. 24, 2127-2132. doi:10.1101/gad.593410
Azhakanandam, S., Nole-Wilson, S., Bao, F. and Franks, R. G. (2008). SEUSS
and AINTEGUMENTA mediate patterning and ovule initiation during gynoecium
medial domain development. Plant Physiol. 146, 1165-1181. doi:10.1104/pp.107.
114751
Balanza, V., Roig-Villanova, I., Di Marzo, M., Masiero, S. and Colombo, L.
(2016). Seed abscission and fruit dehiscence required for seed dispersal rely on
similar genetic networks. Development 143, 3372-3381. doi:10.1242/dev.135202
Ballester, P. and Ferrandiz, C. (2017). Shattering fruits: variations on a dehiscent
theme. Curr. Opin. Plant Biol. 35, 68-75. doi:10.1016/j.pbi.2016.11.008
Ballester, P., Martı
nez-Godoy, M. A., Ezquerro, M., Navarrete-Gomez, M.,
Trigueros, M., Rodrı
guez-Concepcion, M. and Ferrandiz, C. (2021). A
transcriptional complex of NGATHA and bHLH transcription factors directs
stigma development in Arabidopsis. Plant Cell 33, 3645-3657. doi:10.1093/plcell/
koab236
Barro-Trastoy, D., Carrera, E., Banos, J., Palau-Rodrı
guez, J., Ruiz-Rivero, O.,
Tornero, P., Alonso, J. M., Lopez-Dı
az, I., Gomez, M. D. and Perez-
Amador, M. A. (2020a). Regulation of ovule initiation by gibberellins and
brassinosteroids in tomato and Arabidopsis: two plant species, two molecular
mechanisms. Plant J. 102, 1026-1041. doi:10.1111/tpj.14684
Barro-Trastoy, D., Dolores Gomez, M., Tornero, P. and Perez-Amador, M. A.
(2020b). On the way to ovules: the hormonal regulation of ovule development.
CRC Crit. Rev. Plant Sci. 39, 431-456. doi:10.1080/07352689.2020.1820203
Bartrina, I., Otto, E., Strnad, M., Werner, T. and Schmu
lling, T. (2011). Cytokinin
regulates the activity of reproductive meristems, flower organ size, ovule
formation, and thus seed yield in Arabidopsis thaliana. Plant Cell 23, 69-80.
doi:10.1105/tpc.110.079079
Becker, A. (2020). A molecular update on the origin of the carpel. Curr. Opin. Plant
Biol. 53, 15-22. doi:10.1016/j.pbi.2019.08.009
Bowman, J. L., Smyth, D. R. and Meyerowitz, E. M. (1989). Genes directing flower
development in Arabidopsis. Plant Cell 1, 37-52. doi:10.1105/tpc.1.1.37
Bowman, J. L., Baum, S. F., Eshed, Y., Putterill, J. and Alvarez, J. (1999). 4
molecular genetics of gynoecium development in arabidopsis. In Current Topics in
Developmental Biology, Vol. 45, pp. 155-205. Elsevier.
Brand, U., Fletcher, J. C., Hobe, M., Meyerowitz, E. M. and Simon, R. (2000).
Dependence of stem cell fate in Arabidopsis on a feedback loop regulated by
CLV3 activity. Science 289, 617-619. doi:10.1126/science.289.5479.617
Carabelli, M., Turchi, L., Morelli, G., Østergaard, L., Ruberti, I. and
Moubayidin, L. (2021). Coordination of biradial-to-radial symmetry and tissue
polarity by HD-ZIP II proteins. Nat. Commun. 12, 4321. doi:10.1038/s41467-021-
24550-6
Cascallares, M., Setzes, N., Marchetti, F., Lopez, G. A., Distefano, A. M.,
Cainzos, M., Zabaleta, E. and Pagnussat, G. C. (2020). A complex journey: cell
wall remodeling, interactions, and integrity during pollen tube growth. Front. Plant
Sci. 11, 599247. doi:10.3389/fpls.2020.599247
Cerbantez-Bueno, V. E., Zuniga-Mayo, V. M., Reyes-Olalde, J. I., Lozano-
Sotomayor, P., Herrera-Ubaldo, H., Marsch-Martinez, N. and de Folter, S.
(2020). Redundant and Non-redundant functions of the AHK Cytokinin receptors
during gynoecium development. Front. Plant Sci. 11, 568277. doi:10.3389/fpls.
2020.568277
Chang, W., Guo, Y.,Zhang, H., Liu, X. and Guo, L. (2020). Same actor in different
stages: genes in shoot apical meristem maintenance and floral meristem
determinacy in arabidopsis. Front. Ecol. Evol. 8, 89. doi:10.3389/fevo.2020.00089
Chavez Montes, R. A., Herrera-Ubaldo, H., Serwatowska, J. and de Folter, S.
(2015). Towards a comprehensive and dynamic gynoecium gene regulatory
network. Curr. Plant Biol. 3-4, 3-12. doi:10.1016/j.cpb.2015.08.002
Chen, D., Yan, W., Fu, L.-Y. and Kaufmann, K. (2018). Architecture of gene
regulatory networks controlling flower development in Arabidopsis thaliana. Nat.
Commun. 9, 4534. doi:10.1038/s41467-018-06772-3
Coen, E. S. and Meyerowitz, E. M. (1991). The war of the whorls: genetic
interactions controlling flower development. Nature 353, 31-37. doi:10.1038/
353031a0
Crawford, B. C. W. and Yanofsky, M. F. (2008). The formation and function of the
female reproductive tract in flowering plants. Curr. Biol. 18, R972-R978.
doi:10.1016/j.cub.2008.08.010
Crawford, B. C. W. and Yanofsky, M. F. (2011). HALF FILLED promotes
reproductive tract development and fertilization efficiency in Arabidopsis
thaliana. Development 138, 2999-3009. doi:10.1242/dev.067793
Crawford, B. C. W., Ditta, G. and Yanofsky, M. F. (2007). The NTT gene is required
for transmitting-tract development in carpels of Arabidopsis thaliana. Curr. Biol.
17, 1101-1108. doi:10.1016/j.cub.2007.05.079
Cucinotta, M., Colombo, L. and Roig-Villanova, I. (2014). Ovule development, a
new model for lateral organ formation. Front. Plant Sci. 5, 117. doi:10.3389/fpls.
2014.00117
Cucinotta, M., Manrique, S., Cuesta, C., Benkova, E., Novak, O. and
Colombo, L. (2018). CUP-SHAPED COTYLEDON1 (CUC1) and CUC2
regulate cytokinin homeostasis to determine ovule number in Arabidopsis.
J. Exp. Bot. 69, 5169-5176. doi:10.1093/jxb/ery281
Cucinotta, M., Di Marzo, M., Guazzotti, A., de Folter, S., Kater, M. M. and
Colombo, L. (2020). Gynoecium size and ovule number are interconnected traits
that impact seed yield. J. Exp. Bot. 71, 2479-2489. doi:10.1093/jxb/eraa050
Deb, J., Bland, H. M. and Østergaard, L. (2018). Developmental cartography:
coordination via hormonal and genetic interactions during gynoecium formation.
Curr. Opin. Plant Biol. 41, 54-60. doi:10.1016/j.pbi.2017.09.004
Denay, G., Chahtane, H., Tichtinsky, G. and Parcy, F. (2017). A flower is born: an
update on Arabidopsis floral meristem formation. Curr. Opin. Plant Biol. 35, 15-22.
doi:10.1016/j.pbi.2016.09.003
de Folter, S. (2016). Auxin is required for valve margin patterning in arabidopsis
after all. Mol. Plant 9, 768-770. doi:10.1016/j.molp.2016.05.005
Dinneny, J. R., Weigel, D. and Yanofsky, M. F. (2005). A genetic framework for fruit
patterning in Arabidopsis thaliana. Development 132, 4687-4696. doi:10.1242/
dev.02062
Di Marzo, M., Herrera-Ubaldo, H., Caporali, E., Novak, O., Strnad, M.,
Balanza, V., Ezquer, I., Mendes, M. A., de Folter, S. and Colombo, L.
(2020a). SEEDSTICK controls arabidopsis fruit size by regulating cytokinin levels
and FRUITFULL. Cell Rep. 30, 2846-2857.e3. doi:10.1016/j.celrep.2020.01.101
Di Marzo, M., Roig-Villanova, I., Zanchetti, E., Caselli, F., Gregis, V.,Bardetti, P.,
Chiara, M., Guazzotti, A., Caporali, E., Mendes, M. A. et al. (2020b). MADS-Box
and bHLH Transcription Factors Coordinate Transmitting Tract Development in
Arabidopsis thaliana. Front. Plant Sci. 11, 526. doi:10.3389/fpls.2020.00526
Di Marzo, M., Viana, V. E., Banfi, C., Cassina, V., Corti, R., Herrera-Ubaldo, H.,
Babolin, N., Guazzotti, A., Kiegle, E., Gregis, V. et al. (2021). Cell wall
modifications by α-XYLOSIDASE1 are required for the control of seed and fruit
size. J. Exp. Bot. erab514. doi:10.1093/jxb/erab514
Dong, Y., Majda, M., S
imura, J., Horvath, R., Srivastava, A. K., Łangowski, Ł.,
Eldridge, T., Stacey, N., Slotte, T., Sadanandom, A. et al. (2020).
HEARTBREAK Controls Post-translational Modification of INDEHISCENT to
Regulate Fruit Morphology in Capsella. Curr. Biol. 30, 3880-3888.e5. doi:10.1016/
j.cub.2020.07.055
11
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
Dorcey, E., Urbez, C., Blazquez, M. A., Carbonell, J. and Perez-Amador, M. A.
(2009). Fertilization-dependent auxin response in ovules triggers fruit
development through the modulation of gibberellin metabolism in Arabidopsis.
Plant J. 58, 318-332. doi:10.1111/j.1365-313X.2008.03781.x
Dresselhaus, T., Sprunck, S. and Wessel, G. M. (2016). Fertilization mechanisms
in flowering plants. Curr. Biol. 26, R125-R139. doi:10.1016/j.cub.2015.12.032
Duan, Q., Liu, M.-C. J., Kita, D., Jordan, S. S., Yeh, F.-L. J., Yvon, R.,
Carpenter, H., Federico, A. N., Garcia-Valencia, L. E., Eyles, S. J. et al. (2020).
FERONIA controls pectin- and nitric oxide-mediated male-female interaction.
Nature 579, 561-566. doi:10.1038/s41586-020-2106-2
Endress, P. K. (2019). The morphological relationship between carpels and ovules
in angiosperms: pitfalls of morphological interpretation. Botan. J. Linn. Soc. 189,
201-227. doi:10.1093/botlinnean/boy083
Endress, P. K. and Doyle, J. A. (2015). Ancestral traits and specializations in the
flowers of the basal grade of living angiosperms. Taxon 64, 1093-1116.
doi:10.1002/646.1
Endress, P. K. and Igersheim, A. (2000). Gynoecium structure and evolution in
basal angiosperms. Int. J. Plant Sci. 161, S211-S213. doi:10.1086/317572
Favaro, R., Pinyopich, A., Battaglia, R., Kooiker, M., Borghi, L., Ditta, G.,
Yanofsky, M. F., Kater, M. M. and Colombo, L. (2003). MADS-box protein
complexes control carpel and ovule development in Arabidopsis. Plant Cell 15,
2603-2611. doi:10.1105/tpc.015123
Ferrandiz, C. and Fourquin, C. (2014). Role of the FUL-SHP network in the
evolution of fruit morphology and function. J. Exp. Bot. 65, 4505-4513.
doi:10.1093/jxb/ert479
Ferrandiz, C., Fourquin, C., Prunet, N., Scutt, C. P., Sundberg, E., Trehin, C. and
Vialette-Guiraud, A. C. M. (2010). Carpel Development, pp. 1-73. Elsevier.
Gaillochet, C. and Lohmann, J. U. (2015). The never-ending story: from
pluripotency to plant developmental plasticity. Development 142, 2237-2249.
doi:10.1242/dev.117614
Galbiati, F., Sinha Roy, D., Simonini, S., Cucinotta, M., Ceccato, L., Cuesta, C.,
Simaskova, M., Benkova, E., Kamiuchi, Y.,Aida , M. et al. (2013). An integrative
model of the control of ovule primordia formation. Plant J. 76, 446-455.
doi:10.1111/tpj.12309
Gomariz-Fernandez, A., Sanchez-Gerschon, V., Fourquin, C. and Ferrandiz, C.
(2017). The role of SHI/STY/SRS genes in organ growth and carpel development
is conserved in the distant eudicot species arabidopsis thaliana and nicotiana
benthamiana. Front. Plant Sci. 8, 814. doi:10.3389/fpls.2017.00814
Gao, Z., Daneva, A., Salanenka, Y., Van Durme, M., Huysmans, M., Lin, Z., De
Winter, F., Vanneste, S., Karimi, M., Van de Velde, J. et al. (2018). KIRA1 and
ORESARA1 terminate flower receptivity by promoting cell death in the stigma of
Arabidopsis. Nat. Plants 4, 365-375. doi:10.1038/s41477-018-0160-7
Gomez, M. D., Barro-Trastoy, D., Escoms, E., Saura-Sanchez, M., Sanchez, I.,
Briones-Moreno, A., Vera-Sirera, F., Carrera, E., Ripoll, J.-J., Yanofsky, M. F.
et al. (2018). Gibberellins negatively modulate ovule number in plants.
Development 145, dev163865. doi:10.1242/dev.163865
Gomez, M. D., Fuster-Almunia, C., Ocana-Cuesta, J., Alonso, J. M. and Perez-
Amador, M. A. (2019). RGL2 controls flower development, ovule number and
fertility in Arabidopsis. Plant Sci. 281, 82-92. doi:10.1016/j.plantsci.2019.01.014
Gomez-Felipe, A., Kierzkowski, D. and de Folter, S. (2021). The Relationship
between AGAMOUS and Cytokinin Signaling in the Establishment of Carpeloid
Features. Plants 10, 827. doi:10.3390/plants10050827
Govaerts, R. (2001). How many species of seed plants are there? Taxon 50,
1085-1090. doi:10.2307/1224723
Gremski, K., Ditta, G. and Yanofsky, M. F. (2007). The HECATE genes regulate
female reproductive tract development in Arabidopsis thaliana. Development 134,
3593-3601. doi:10.1242/dev.011510
Harnvanichvech, Y., Gorelova, V., Sprakel, J. and Weijers, D. (2021). The
Arabidopsis embryo as a quantifiable model for studying pattern formation. Quant
Plant Bio. 2, e3. doi:10.1017/qpb.2021.3
Heisler, M. G., Atkinson, A., Bylstra, Y. H., Walsh, R. and Smyth, D. R. (2001).
SPATULA, a gene that controls development of carpel margin tissues in
Arabidopsis, encodes a bHLH protein. Development 128, 1089-1098.
doi:10.1242/dev.128.7.1089
Hernandez-Lagana, E., Mosca, G., Mendocilla-Sato, E., Pires, N., Frey, A.,
Giraldo-Fonseca, A., Michaud, C., Grossniklaus, U., Hamant, O., Godin, C.
et al. (2021). Organ geometry channels reproductive cell fate in the Arabidopsis
ovule primordium. eLife 10, e66031. doi:10.7554/eLife.66031
Herrera-Ubaldo, H. and de Folter, S. (2018). Exploring cell wall composition and
modifications during the development of the gynoecium medial domain in
arabidopsis. Front. Plant Sci. 9, 454. doi:10.3389/fpls.2018.00454
Herrera-Ubaldo, H., Campos, S. E., Luna Garcia, V.,Zunig a-Mayo, V. M., Armas-
Caballero, G., DeLuna, A., Marsch-Martinez, N. and de Folter, S. (2018). An
interaction map of transcription factors controlling gynoecium development in
Arabidopsis. BioRxiv. doi:10.1101/500736
Herrera-Ubaldo, H., Lozano-Sotomayor, P., Ezquer, I., Di Marzo, M., Chavez
Montes, R. A., Gomez-Felipe, A., Pablo-Villa, J., Diaz-Ramirez, D.,
Ballester, P., Ferrandiz, C. et al. (2019). New roles of NO TRANSMITTING
TRACT and SEEDSTICK during medial domain development in Arabidopsis
fruits. Development 146, dev172395. doi:10.1242/dev.172395
Huang, H.-Y., Jiang, W.-B., Hu, Y.-W., Wu, P., Zhu, J.-Y., Liang, W.-Q., Wang, Z.-
Y. and Lin, W.-H. (2013). BR signal influences Arabidopsis ovule and seed
number through regulating related genes expression by BZR1. Mol. Plant 6,
456-469. doi:10.1093/mp/sss070
Hugouvieux, V., Silva, C. S., Jourdain, A., Stigliani, A., Charras, Q., Conn, V.,
Conn, S. J., Carles, C. C., Parcy, F. and Zubieta, C. (2018). Tetramerization of
MADS family transcription factors SEPALLATA3 and AGAMOUS is required for
floral meristem determinacy in Arabidopsis. Nucleic Acids Res. 46, 4966-4977.
doi:10.1093/nar/gky205
Irish, V. (2017). The ABC model of floral development. Curr. Biol. 27, R887-R890.
doi:10.1016/j.cub.2017.03.045
Jasinski, S., Piazza, P., Craft, J., Hay, A., Woolley, L., Rieu, I., Phillips, A.,
Hedden, P. and Tsiantis, M. (2005). KNOX action in Arabidopsis is mediated by
coordinate regulation of cytokinin and gibberellin activities. Curr. Biol. 15,
1560-1565. doi:10.1016/j.cub.2005.07.023
Jiao, Y. and Meyerowitz, E. M. (2010). Cell-type specific analysis of translating
RNAs in developing flowers reveals new levels of control. Mol. Syst. Biol. 6, 419.
doi:10.1038/msb.2010.76
Johnson, M. A., Harper, J. F. and Palanivelu, R. (2019). A Fruitful Journey: Pollen
Tube Navigation from Germination to Fertilization. Annu. Rev. Plant Biol. 70,
809-837. doi:10.1146/annurev-arplant-050718-100133
JoseRipoll, J., Bailey, L. J., Mai, Q.-A., Wu, S. L., Hon, C. T., Chapman, E. J.,
Ditta, G. S., Estelle, M. and Yanofsky, M. F. (2015). microRNA regulation of fruit
growth. Nat. Plants 1, 15036. doi:10.1038/nplants.2015.36
Julca, I., Ferrari, C., Flores-Tornero, M., Proost, S., Lindner, A.-C.,
Hackenberg, D., Steinbachova, L., Michaelidis, C., Gomes Pereira, S.,
Misra, C. S. et al. (2021). Comparative transcriptomic analysis reveals
conserved programmes underpinning organogenesis and reproduction in land
plants. Nat. Plants 7, 1143-1159. doi:10.1038/s41477-021-00958-2
Kamiuchi, Y., Yamamoto, K., Furutani, M., Tasaka, M. and Aida, M. (2014). The
CUC1 and CUC2 genes promote carpel margin meristem formation during
Arabidopsis gynoecium development. Front. Plant Sci. 5, 165. doi:10.3389/fpls.
2014.00165
Kawamoto, N., Del Carpio, D. P., Hofmann, A., Mizuta, Y., Kurihara, D.,
Higashiyama, T., Uchida, N., Torii, K. U., Colombo, L., Groth, G. et al. (2020). A
peptide pair coordinates regular ovule initiation patterns with seed number and
fruit size. Curr. Biol. 30, 4352-4361.e4. doi:10.1016/j.cub.2020.08.050
Kivivirta, K. I., Herbert, D., Roessner, C., de Folter, S., Marsch-Martinez, N. and
Becker, A. (2021). Transcriptome analysis of gynoecium morphogenesis
uncovers the chronology of gene regulatory network activity. Plant Physiol. 185,
1076-1090. doi:10.1093/plphys/kiaa090
Klepikova, A. V., Kasianov, A. S., Gerasimov, E. S., Logacheva, M. D. and
Penin, A. A. (2016) A high resolution map of the Arabidopsis thaliana
developmental transcriptome based on RNA-seq profiling. Plant J. 88,
1058-1070. doi:10.1111/tpj.13312
Krizek, B. A., Blakley, I. C., Ho, Y.-Y., Freese, N. and Loraine, A. E. (2020). The
Arabidopsis transcription factor AINTEGUMENTA orchestrates patterning genes
and auxin signaling in the establishment of floral growth and form. Plant J. 103,
752-768. doi:10.1111/tpj.14769
Krizek, B. A., Bantle, A. T., Heflin, J. M., Han, H., Freese, N. H. and Loraine, A. E.
(2021). AINTEGUMENTA and AINTEGUMENTA-LIKE6 directly regulate floral
homeotic, growth, and vascular development genes in young Arabidopsis flowers.
J. Exp. Bot. 72, 5478-5493. doi:10.1093/jxb/erab223
Kwasniewska, K., Breathnach, C., Fitzsimons, C., Goslin, K., Thomson, B.,
Beegan, J., Finocchio, A., Prunet, N., ÓMaoileidigh, D. S. and Wellmer, F.
(2021). Expression of KNUCKLES in the stem cell domain is required for its
function in the control of floral meristem activity in arabidopsis.Front. Plant Sci. 12,
704351. doi:10.3389/fpls.2021.704351
Lai, X., Stigliani, A., Lucas, J., Hugouvieux, V., Parcy, F. and Zubieta, C. (2020).
Genome-wide binding of SEPALLATA3 and AGAMOUS complexes determined
by sequential DNA-affinity purification sequencing. Nucleic Acids Res. 48,
9637-9648. doi:10.1093/nar/gkaa729
Łangowski, Ł., Stacey, N. and Østergaard, L. (2016). Diversification of fruitshape
in the Brassicaceae family. Plant Reprod. 29, 149-163. doi:10.1007/s00497-016-
0278-6
Lee, Z. H., Hirakawa, T., Yamaguchi, N. and Ito, T. (2019). The roles of plant
hormones and their interactions with regulatory genes in determining meristem
activity. Int. J. Mol. Sci. 20, 4065. doi:10.3390/ijms20164065
Liao, S., Wang, L., Li, J. and Ruan, Y.-L. (2020). Cellwall invertase is essential for
ovule development through sugar signaling rather than provision of carbon
nutrients. Plant Physiol. 183, 1126-1144. doi:10.1104/pp.20.00400
Liljegren, S. J., Roeder, A. H. K., Kempin, S. A., Gremski, K., Østergaard, L.,
Guimil, S., Reyes, D. K. and Yanofsky,M. F. (2004). Control of fruit patterning in
Arabidopsis by INDEHISCENT. Cell 116, 843-853. doi:10.1016/S0092-
8674(04)00217-X
Liu, C., Shen, L., Xiao, Y., Vyshedsky, D., Peng, C., Sun, X., Liu, Z., Cheng, L.,
Zhang, H., Han, Z. et al. (2021). Pollen PCP-B peptides unlock a stigma peptide-
receptor kinase gating mechanism for pollination. Science 372, 171-175.
doi:10.1126/science.abc6107
12
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
Li, X.-R., Deb, J., Kumar, S. V. and Østergaard, L. (2018). Temperature modulates
tissue-specification program to control fruit dehiscence in brassicaceae. Mol.
Plant 11, 598-606. doi:10.1016/j.molp.2018.01.003
Li, W., Huang, X., Zou, J., Wu, J., Jiao, H., Peng, X. and Sun, M.-X. (2020). Three
STIGMA AND STYLE STYLISTs pattern the fine architectures of apical
gynoecium and are critical for male gametophyte-pistil interaction. Curr. Biol.
30, 4780-4788.e5. doi:10.1016/j.cub.2020.09.006
Li, S., Chen, K. and Grierson, D. (2021). Molecular and hormonal mechanisms
regulating fleshy fruit ripening. Cells 10, 1136. doi:10.3390/cells10051136
Marsch-Martı
nez, N. and de Folter, S. (2016). Hormonal control of the
development of the gynoecium. Curr. Opin. Plant Biol. 29, 104-114.
doi:10.1016/j.pbi.2015.12.006
Marsch-Martı
nez, N., Ramos-Cruz, D., Irepan Reyes-Olalde, J., Lozano-
Sotomayor, P., Zuniga-Mayo, V. M. and de Folter, S. (2012). The role of
cytokinin during Arabidopsis gynoecia and fruit morphogenesis and patterning.
Plant J. 72, 222-234. doi:10.1111/j.1365-313X.2012.05062.x
Marsch-Martı
nez, N., Zuniga-Mayo, V. M., Herrera-Ubaldo, H.,
Ouwerkerk, P. B. F., Pablo-Villa, J., Lozano-Sotomayor, P., Greco, R.,
Ballester, P., Balanza, V., Kuijt, S. J. H. et al. (2014). The NTT transcription
factor promotes replum development in Arabidopsis fruits. Plant J. 80, 69-81.
doi:10.1111/tpj.12617
Martı
nez-Fernandez, I., Menezes de Moura, S., Alves-Ferreira, M., Ferrandiz, C.
and Balanza,V.(2020). Identification of players controlling meristem arrest
downstream of the FRUITFULL-APETALA2 pathway. Plant Physiol. 184,
945-959. doi:10.1104/pp.20.00800
McAtee, P., Karim, S., Schaffer, R. and David, K. (2013). A dynamic interplay
between phytohormones is required for fruit development, maturation, and
ripening. Front. Plant Sci. 4, 79. doi:10.3389/fpls.2013.00079
McWhite, C. D., Papoulas, O., Drew, K., Cox, R. M., June, V., Dong, O. X.,
Kwon, T., Wan, C., Salmi, M. L., Roux, S. J. et al. (2020). A Pan-plant protein
complex map reveals deep conservation and novel assemblies. Cell 181,
460-474.e14. doi:10.1016/j.cell.2020.02.049
Mergner, J., Frejno, M., List, M., Papacek, M., Chen, X., Chaudhary, A.,
Samaras, P., Richter, S., Shikata, H., Messerer, M. et al. (2020). Mass-
spectrometry-based draft of the Arabidopsis proteome. Nature 579, 409-414.
doi:10.1038/s41586-020-2094-2
Mizzotti, C., Rotasperti, L., Moretto, M., Tadini, L., Resentini, F., Galliani, B. M.,
Galbiati, M., Engelen, K., Pesaresi, P. and Masiero, S. (2018). Time-course
transcriptome analysis of arabidopsis siliques discloses genes essential for fruit
development and maturation. Plant Physiol. 178, 1249-1268. doi:10.1104/pp.18.
00727
Moubayidin, L. and Ostergaard, L. (2014). Dynamic control of auxin distribution
imposes a bilateral-to-radial symmetry switch during gynoecium development.
Curr. Biol. 24, 2743-2748. doi:10.1016/j.cub.2014.09.080
Mu
ller, C. J., Larsson, E., Spı
chal, L. and Sundberg, E. (2017). Cytokinin-Auxin
crosstalk in the gynoecial primordium ensures correct domain patterning. Plant
Physiol. 175, 1144-1157. doi:10.1104/pp.17.00805
Nahar, M. A.-U., Ishida, T., Smyth, D. R., Tasaka, M. and Aida, M. (2012).
Interactions of CUP-SHAPED COTYLEDON and SPATULA genes control carpel
margin development in Arabidopsis thaliana. Plant Cell Physiol. 53, 1134-1143.
doi:10.1093/pcp/pcs057
Neumann, M., Xu, X., Smaczniak, C., Schumacher, J., Yan, W., Greb, T.,
Bluthgen, N., Jonsson, H., Traas, J., Kaufmann, K. et al. (2021). A 3D gene
expression atlas of the floral meristem based on spatial reconstruction of single
nucleus RNA sequencing data. BioRxiv. doi:10.1101/2021.06.30.450319
Nole-Wilson, S., Azhakanandam, S. and Franks, R. G. (2010). Polar auxin
transport together with aintegumenta and revoluta coordinate early Arabidopsis
gynoecium development. Dev. Biol. 346, 181-195. doi:10.1016/j.ydbio.2010.07.
016
ÓMaoileidigh, D. S., Stewart, D., Zheng, B., Coupland, G. and Wellmer, F.
(2018). Floral homeotic proteins modulate the genetic program for leaf
development to suppress trichome formation in flowers. Development 145,
dev157784. doi:10.1242/dev.157784
Ortiz-Ramı
rez, C. I., Plata-Arboleda, S. and Pabon-Mora, N. (2018). Evolution of
genes associated with gynoecium patterning and fruit development in
Solanaceae. Ann. Bot. 121, 1211-1230. doi:10.1093/aob/mcy007
Pabon-Mora, N., Wong, G. K.-S. and Ambrose, B. A. (2014). Evolution of fruit
development genes in flowering plants. Front. Plant Sci. 5, 300. doi:10.3389/fpls.
2014.00300
Pajoro, A., Biewers, S., Dougali, E., Leal Valentim, F., Mendes, M. A., Porri, A.,
Coupland, G., Van de Peer, Y., van Dijk, A. D. J., Colombo, L. et al. (2014). The
(r)evolution of gene regulatory networks controlling Arabidopsis plant
reproduction: a two-decade history. J. Exp. Bot. 65, 4731-4745. doi:10.1093/
jxb/eru233
Pelayo, M. A., Yamaguchi, N. and Ito, T. (2021). One factor, many systems: the
floral homeotic protein AGAMOUS and its epigenetic regulatory mechanisms.
Curr. Opin. Plant Biol. 61, 102009. doi:10.1016/j.pbi.2021.102009
Pereira, A. M., Moreira,D., Coimbra, S. and Masiero, S. (2021). Paving the way for
fertilization: the role of the transmitting tract. Int. J. Mol. Sci. 22, 2603. doi:10.3390/
ijms22052603
Pfannebecker, K. C., Lange, M., Rupp, O. and Becker, A. (2017a). An
evolutionary framework for carpel developmental control genes. Mol. Biol. Evol.
34, 330-348. doi:10.1093/molbev/msw229
Pfannebecker, K. C., Lange, M., Rupp, O. and Becker, A. (2017b). Seed plant-
specific gene lineages involved in carpel development. Mol. Biol. Evol. 34,
925-942. doi:10.1093/molbev/msw297
Pinto, S. C., Mendes, M. A., Coimbra, S. and Tucker, M. R. (2019). Revisiting the
female germline and its expanding toolbox. Trends Plant Sci. 24, 455-467.
doi:10.1016/j.tplants.2019.02.003
Pinyopich, A., Ditta, G. S., Savidge, B., Liljegren, S. J., Baumann, E.,
Wisman, E. and Yanofsky, M. F. (2003). Assessing the redundancy of MADS-
box genes during carpel and ovule development. Nature 424, 85-88. doi:10.1038/
nature01741
Plant Cell Atlas Consortium,Jha, S. G., Borowsky, A. T., Cole, B. J.,
Fahlgren, N., Farmer, A., Huang, S.-S. C., Karia, P., Libault, M.,
Provart, N. J. et al. (2021). Vision, challenges and opportunities for a plant cell
atlas. eLife 10, e66877. doi:10.7554/eLife.66877
Qadir, M., Wang, X., Shah, S. R. U., Zhou, X.-R., Shi, J. and Wang, H. (2021).
Molecular network for regulation of ovule number in plants. Int. J. Mol. Sci. 22,
12965. doi:10.3390/ijms222312965
Refahi, Y., Zardilis, A., Michelin, G., Wightman, R., Leggio, B., Legrand, J.,
Faure, E., Vachez, L., Armezzani, A., Risson, A.-E. et al. (2021). A multiscale
analysis of early flower development in Arabidopsis provides an integrated view of
molecular regulation and growth control. Dev. Cell 56, 540-556.e8. doi:10.1016/
j.devcel.2021.01.019
Reyes-Olalde, J. I. and de Folter, S. (2019). Control of stem cell activity in the
carpel margin meristem (CMM) in Arabidopsis. Plant Reprod. 32, 123-136.
doi:10.1007/s00497-018-00359-0
Reyes-Olalde, J. I., Zuniga-Mayo, V. M., Chavez Montes, R. A., Marsch-
Martı
nez, N. and de Folter, S. (2013). Inside the gynoecium: at the carpel margin.
Trends Plant Sci. 18, 644-655. doi:10.1016/j.tplants.2013.08.002
Reyes-Olalde, J. I., Zuniga-Mayo, V. M., Serwatowska, J., Chavez Montes, R. A.,
Lozano-Sotomayor, P., Herrera-Ubaldo, H., Gonzalez-Aguilera, K. L.,
Ballester, P., Ripoll, J. J., Ezquer, I. et al. (2017). The bHLH transcription
factor SPATULA enables cytokinin signaling, and both activate auxin biosynthesis
and transport genes at the medial domain of the gynoecium. PLoS Genet. 13,
e1006726. doi:10.1371/journal.pgen.1006726
Rhee, S. Y., Birnbaum, K. D. and Ehrhardt, D. W. (2019). Towards building a plant
cell atlas. Trends Plant Sci. 24, 303-310. doi:10.1016/j.tplants.2019.01.006
Ripoll, J., Bailey, L. J., Mai, Q.-A., Wu, S. L., Hon, C. T., Chapman, E. J., Ditta, G.
S., Estelle, M. and Yanofsky, M. F. (2015). microRNA regulation of fruit growth.
Nat. Plants 1, 15036. doi:10.1038/nplants.2015.36
Ripoll, J.-J., Zhu, M., Brocke, S., Hon, C. T., Yanofsky, M. F., Boudaoud, A. and
Roeder, A. H. K. (2019). Growth dynamics of the Arabidopsis fruit is mediated by
cell expansion. Proc. Natl. Acad. Sci. U.S.A. 116, 25333-25342. doi:10.1073/
pnas.1914096116
Roeder, A. H. K. and Yanofsky, M. F. (2006). Fruit development in Arabidopsis.
Arabidopsis Book 4, e0075.
Romera-Branchat, M., Ripoll, J. J., Yanofsky, M. F. and Pelaz, S. (2013). The
WOX13 homeobox gene promotes replum formation in the Arabidopsis thaliana
fruit. Plant J. 73, 37-49. doi:10.1111/tpj.12010
Rong, X. F., Sang, Y. L., Wang, L., Meng, W. J., Zou, C. H., Dong, Y. X., Bie, X. M.,
Cheng, Z. J. and Zhang, X. S. (2018). Type-B ARRs control carpel regeneration
through mediating AGAMOUS expression in arabidopsis. Plant Cell Physiol. 59,
756-764. doi:10.1093/pcp/pcx187
Sablowski, R. (2007). Flowering and determinacy in Arabidopsis. J. Exp. Bot. 58,
899-907. doi:10.1093/jxb/erm002
Schoof, H., Lenhard, M., Haecker, A., Mayer, K. F., Ju
rgens, G. and Laux, T.
(2000). The stem cell population of Arabidopsis shoot meristems in maintained by
a regulatory loop between the CLAVATA and WUSCHEL genes. Cell 100,
635-644. doi:10.1016/S0092-8674(00)80700-X
Schuster, C., Gaillochet, C. and Lohmann, J. U. (2015). Arabidopsis HECATE
genes function in phytohormone control during gynoecium development.
Development 142, 3343-3350. doi:10.1242/dev.120444
Scofield, S., Dewitte, W. and Murray, J. A. H. (2007). The KNOX gene SHOOT
MERISTEMLESS is required for the development of reproductive meristematic
tissues in Arabidopsis. Plant J. 50, 767-781. doi:10.1111/j.1365-313X.2007.
03095.x
Scutt, C. P., Vinauger-Douard, M., Fourquin, C., Finet, C. and Dumas, C. (2006).
An evolutionary perspective on the regulation of carpel development. J. Exp. Bot.
57, 2143-2152. doi:10.1093/jxb/erj188
Sehra, B. and Franks, R. G. (2015). Auxin and cytokinin act during gynoecial
patterning and the development of ovules from the meristematic medial domain.
Wiley Interdiscip. Rev. Dev. Biol. 4, 555-571. doi:10.1002/wdev.193
Sessions, A., Nemhauser, J. L., McColl, A., Roe, J. L., Feldmann, K. A. and
Zambryski, P. C. (1997). ETTIN patterns the Arabidopsis floral meristem and
reproductive organs. Development 124, 4481-4491. doi:10.1242/dev.124.22.
4481
13
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
Shang, E., Ito, T. and Sun, B. (2019). Control of floral stem cell activity in
Arabidopsis. Plant Signal. Behav. 14, 1659706. doi:10.1080/15592324.2019.
1659706
Shang, E., Wang, X., Li, T., Guo, F., Ito, T. and Sun, B. (2021). Robust control of
floral meristem determinacy by position-specific multifunctions of KNUCKLES.
Proc. Natl. Acad. Sci. U.S.A. 118, e2102826118. doi:10.1073/pnas.2102826118
Silveira, S. R., Le Gloanec, C., Gomez-Felipe, A., Routier-Kierzkowska, A.-L.
and Kierzkowski, D. (2021). Live-imaging provides an atlas of cellular growth
dynamics in the stamen. Plant Physiol. kiab363. doi:10.1093/plphys/kiab363
Simonini, S. and Østergaard, L. (2019). Female reproductive organ formation: a
multitasking endeavor. Curr. Top. Dev. Biol. 131, 337-371. doi:10.1016/bs.ctdb.
2018.10.004
Simonini, S., Deb, J., Moubayidin, L., Stephenson, P., Valluru, M., Freire-
Rios, A., Sorefan, K., Weijers, D., Friml, J. and Østergaard, L. (2016). A
noncanonical auxin-sensing mechanism is required for organ morphogenesis in
Arabidopsis. Genes Dev. 30, 2286-2296. doi:10.1101/gad.285361.116
Simonini, S., Stephenson, P. and Østergaard, L. (2018). A molecular framework
controlling style morphology in Brassicaceae. Development 145, dev158105.
doi:10.1242/dev.158105
Smaczniak, C., Immink, R. G. H., Muino, J. M., Blanvillain, R., Busscher, M.,
Busscher-Lange, J., Dinh, Q. D. P., Liu, S., Westphal, A. H., Boeren, S. et al.
(2012). Characterization of MADS-domain transcription factor complexes in
Arabidopsis flower development. Proc. Natl. Acad. Sci. U.S.A. 109, 1560-1565.
doi:10.1073/pnas.1112871109
Smyth, D. R., Bowman, J. L. and Meyerowitz, E. M. (1990). Early flower
development in Arabidopsis. Plant Cell 2, 755-767. doi:10.1105/tpc.2.8.755
Sorefan, K., Girin, T., Liljegren, S. J., Ljung, K., Robles, P., Galvan-
Ampudia, C. S., Offringa, R., Friml, J., Yanofsky, M. F. and Østergaard, L.
(2009). A regulated auxin minimum is required for seed dispersal in Arabidopsis.
Nature 459, 583-586. doi:10.1038/nature07875
Sotelo-Silveira, M., Marsch-Martı
nez, N. and de Folter, S. (2014). Unraveling the
signal scenario of fruit set. Planta 239, 1147-1158. doi:10.1007/s00425-014-
2057-7
Spinelli, S. V., Martin, A. P., Viola, I. L., Gonzalez, D. H. and Palatnik, J. F. (2011).
A mechanistic link between STM and CUC1 during Arabidopsis development.
Plant Physiol. 156, 1894-1904. doi:10.1104/pp.111.177709
Sun, B., Zhou, Y., Cai, J., Shang, E., Yamaguchi, N., Xiao, J., Looi, L.-S.,
Wee, W.-Y., Gao, X., Wagner, D. et al. (2019). Integration of transcriptional
repression and polycomb-mediated silencing of WUSCHEL in floral meristems.
Plant Cell 31, 1488-1505. doi:10.1105/tpc.18.00450
Theißen, G. and Saedler, H. (2001). Floral quartets. Nature 409, 469-471.
doi:10.1038/35054172
van Gelderen, K., van Rongen, M., Liu, A., Otten, A. and Offringa, R. (2016). An
INDEHISCENT-controlled auxin response specifies the separation layer in early
Arabidopsis fruit. Mol. Plant 9, 857-869. doi:10.1016/j.molp.2016.03.005
Vijayan, A., Tofanelli, R., Strauss, S., Cerrone, L., Wolny, A., Strohmeier, J.,
Kreshuk, A., Hamprecht, F. A., Smith, R. S. and Schneitz, K. (2021). A digital
3D reference atlas reveals cellular growth patterns shaping the Arabidopsis ovule.
eLife 10, e63262. doi:10.7554/eLife.63262
Villarino, G. H., Hu, Q., Manrique, S., Flores-Vergara, M., Sehra, B., Robles, L.,
Brumos, J., Stepanova, A. N., Colombo, L., Sundberg, E. et al. (2016).
Transcriptomic signature of the SHATTERPROOF2 expression domain reveals
the meristematic nature of arabidopsis gynoecial medial domain. Plant Physiol.
171, 42-61. doi:10.1104/pp.15.01845
Werner, S., Bartrina, I., Novak, O., Strnad, M., Werner, T. and Schmu
lling, T.
(2021). The cytokinin status of the epidermis regulates aspects of vegetative and
reproductive development in Arabidopsis thaliana. Front. Plant Sci. 12, 613488.
doi:10.3389/fpls.2021.613488
Willis, K. J. (2017). State of the worlds plants, 2017.
Wynn, A. N., Rueschhoff, E. E. and Franks, R. G. (2011). Transcriptomic
characterization of a synergistic genetic interaction during carpel margin meristem
development in Arabidopsis thaliana. PLoS ONE 6, e26231. doi:10.1371/journal.
pone.0026231
Xu, Y., Yamaguchi, N., Gan, E.-S. and Ito, T. (2019). When to stop: an update on
molecular mechanisms of floral meristem termination. J. Exp. Bot. 70, 1711-1718.
doi:10.1093/jxb/erz048
Yamaguchi, N., Huang, J., Xu, Y., Tanoi, K. and Ito, T. (2017). Fine-tuning of auxin
homeostasis governs the transition from floral stem cell maintenance to
gynoecium formation. Nat. Commun. 8, 1125. doi:10.1038/s41467-017-01252-6
Yamaguchi, N., Huang, J., Tatsumi, Y., Abe, M., Sugano, S. S., Kojima, M.,
Takebayashi, Y., Kiba, T., Yokoyama, R., Nishitani, K. et al. (2018). Chromatin-
mediated feed-forward auxin biosynthesis in floral meristem determinacy. Nat.
Commun. 9, 5290. doi:10.1038/s41467-018-07763-0
Yanai, O., Shani, E., Dolezal, K., Tarkowski, P., Sablowski, R., Sandberg, G.,
Samach, A. and Ori, N. (2005). Arabidopsis KNOXI proteins activate cytokinin
biosynthesis. Curr. Biol. 15, 1566-1571. doi:10.1016/j.cub.2005.07.060
Yanofsky, M. F., Ma, H., Bowman, J. L., Drews, G. N., Feldmann, K. A. and
Meyerowitz, E. M. (1990). The protein encoded by the Arabidopsis homeotic
gene agamous resembles transcription factors. Nature 346, 35-39. doi:10.1038/
346035a0
Yu, S.-X., Zhou, L.-W., Hu, L.-Q., Jiang, Y.-T., Zhang, Y.-J., Feng, S.-L., Jiao, Y.,
Xu, L. and Lin, W.-H. (2020). Asynchrony of ovule primordia initiation in
Arabidopsis.Development 147, dev196618. doi:10.1242/dev.196618
Yuan, J. and Kessler, S. A. (2019). A genome-wide association study reveals a
novel regulator of ovule number and fertility in Arabidopsis thaliana. PLoS Genet.
15, e1007934. doi:10.1371/journal.pgen.1007934
Zhang, K., Wang, R., Zi, H., Li, Y., Cao, X., Li, D., Guo, L., Tong, J., Pan, Y.,
Jiao, Y. et al. (2018). AUXIN RESPONSE FACTOR3 regulates floral meristem
determinacy by repressing cytokinin biosynthesis and signaling. Plant Cell 30,
324-346. doi:10.1105/tpc.17.00705
Zhang, J., Yue, L., Wu, X., Liu, H. and Wang, W. (2021). Function of small peptides
during male-female crosstalk in plants. Front. Plant Sci. 12, 671196. doi:10.3389/
fpls.2021.671196
Zhou, L.-Z., Qu, L.-J. and Dresselhaus, T. (2021). Stigmatic ROS: regulator of
compatible pollen tube perception? Trends Plant Sci. 26, 993-995. doi:10.1016/j.
tplants.2021.06.013
Zhu, M., Chen, W., Mirabet, V., Hong, L., Bovio, S., Strauss, S., Schwarz, E. M.,
Tsugawa, S., Wang, Z., Smith, R. S. et al. (2020). Robust organ size requires
robust timing of initiation orchestrated by focused auxin and cytokinin signalling.
Nat. Plants 6, 686-698. doi:10.1038/s41477-020-0666-7
Zuniga-Mayo, V. M., Banos-Bayardo, C. R., Dı
az-Ramı
rez, D., Marsch-
Martı
nez, N. and de Folter, S. (2018). Conserved and novel responses to
cytokinin treatments during flower and fruit development in Brassica napus and
Arabidopsis thaliana. Sci. Rep. 8, 6836. doi:10.1038/s41598-018-25017-3
Zuniga-Mayo, V. M., Gomez-Felipe, A., Herrera-Ubaldo, H. and de Folter, S.
(2019). Gynoecium development: networks in Arabidopsis and beyond. J. Exp.
Bot. 70, 1447-1460. doi:10.1093/jxb/erz026
Zuniga-Mayo, V. M., Marsch-Martı
nez, N. and de Folter, S. (2012). JAIBA, a class-
II HD-ZIP transcription factor involved in the regulation of meristematic activity,
and important for correct gynoecium and fruit development in Arabidopsis. Plant J.
71, 314-326. doi:10.1111/j.1365-313X.2012.04990.x
14
REVIEW Development (2022) 149, dev200120. doi:10.1242/dev.200120
DEVELOPMENT
... During the development of the medial domain, the CMM gives rise to the carpel marginal tissues, which include the placenta, ovules, septum, transmitting tract, style, and stigma, while in the lateral domains, the formation of the valve margins and valves takes place. All these tissues and structures are critical to making up a mature gynoecium ready for fertilization, where a fine action of genes and hormones is required (Balanzá et al., 2006;Ferrándiz et al., 2010;Herrera-Ubaldo & de Folter, 2022;Reyes-Olalde et al., 2013;Roeder & Yanofsky, 2006;Simonini & Østergaard, 2019;Zúñiga-Mayo et al., 2019). ...
... Many years of research focused on identifying regulators involved in gynoecium development have resulted in close to a centenary of genes, many of them coding for transcription factors (Ferrándiz et al., 2010;Herrera-Ubaldo & de Folter, 2022;Reyes-Olalde et al., 2013;Roeder & Yanofsky, 2006;Simonini & Østergaard, (Crawford et al., 2007;Crawford & Yanofsky, 2011;Gremski et al., 2007;Heisler et al., 2001;Kamiuchi et al., 2014) (Bowman & Smyth, 1999;Dinneny et al., 2005;Gu et al., 1998). In addition, experimental evidence of expression patterns of reporters and genes (such as those involved in biosynthesis, transport, and transcriptional response) related to cytokinin and auxin pathways (e.g., Marsch-Martínez et al., 2012b;Moubayidin & Østergaard, 2014;Müller et al., 2017;Reyes-Olalde et al., 2017) indicates that hormones are playing a key role during carpel marginal tissue development. ...
... ;https://doi.org/10.1101https://doi.org/10. /2023 Auxin and cytokinin homeostasis in the medial and lateral domains of the gynoecium Plant hormones, especially auxin and cytokinin, are important for gynoecium development (Deb et al., 2018;Herrera-Ubaldo & de Folter, 2022;Larsson et al., 2013;Marsch-Martínez et al., 2012a;Marsch-Martínez & de Folter, 2016;Robert et al., 2015;Sehra & Franks, 2015). It is known that hormone homeostasis is a complex process that involves hormone biosynthesis, transport, response, and degradation pathways. ...
Preprint
Full-text available
Angiosperms are characterized by the formation of flowers, and in their inner floral whorl, one or various gynoecia are produced. These female reproductive structures are responsible for fruit and seed production, thus ensuring the reproductive competence of angiosperms. In Arabidopsis thaliana, the gynoecium is composed of two fused carpels with different tissues that need to develop and differentiate to consolidate a mature gynoecium and thus the reproductive competence of Arabidopsis. For these reasons, they have become the object of study for floral and fruit development. However, due to the complexity of the gynoecium, specific spatio-temporal tissues expression patterns are still scarce. In this study, we used precise laser-assisted microdissection and high-throughput RNA sequencing to describe the transcriptional profiles of the medial and lateral domain tissues of the Arabidopsis gynoecium. We provide evidence that the method used is reliable and that, in addition to corroborating gene expression patterns of previously reported regulators of these tissues, we found genes whose expression dynamics point to being involved in cytokinin and auxin homeostasis and in cell cycle progression. Furthermore, based on differential gene expression analyses, we functionally characterized several genes and found that they are involved in gynoecium development. This new resource is available via the Arabidopsis eFP browser and will serve the community in future studies on developmental and reproductive biology.
... In addition, hormones need different kinds of proteins to carry out their biosynthesis, signaling and transport, furthermore, most hormones are regulated by feedback mechanisms. Altogether, flower development is a stable and robust process [4][5][6][7]. ...
... The gynoecium, together with the stamens, is crucial for plant sexual reproduction. After fertilization, the gynoecium develops into a fruit, protecting the seeds that develop inside it and releasing them at maturation [3,4,6,7,82,83]. ...
... Throughout its development different tissues are differentiated along three developmental axes (see Fig. 3b). The abaxialadaxial axis refers to the identity of cell types that develop on the outer or inner side of the gynoecium (ovary), respectively; the medial-lateral axis that includes the valves (lateral domain), and the carpel margin meristem (CMM) and the other tissues generated from it, which constitute the medial domain; and the apical-basal axis composed, from top to bottom of stigma, style, ovary and gynophore [3,4,6,7,82,84]. ...
Article
Sexual reproduction requires the participation of two gametes, female and male. In angiosperms, gametes develop in specialized organs, pollen (containing the male gametes) develops in the stamens, and the ovule (containing the female gamete) develops in the gynoecium. In Arabidopsis thaliana, the female and male sexual organs are found within the same structure called flower, surrounded by the perianth, which is composed of petals and sepals. During flower development, different organs emerge in an established order and throughout their development distinct tissues within each organ are differentiated. All this requires the coordination and synchronization of several biological processes. To achieve this, hormones and genes work together. These components can interact at different levels generating hormonal interplay and both positive and negative feedback loops, which in turn, gives robustness, stability, and flexibility to flower development. Here, we summarize the progress made on elucidating the role of different hormonal pathways during flower development in Arabidopsis thaliana.
... The CMM expansion also implies the development of four placentae in the gynoecium and each one is related to ovule formation [2,3,5,7,30]. The placentae enlargement and ovule primordia initiation are regulated by brassinosteroids (BRs) and CKs through the transcriptional regulation of genes involved in ovule identity and initiation [31]. ...
... The diagram shows the different gynoecium tissues, as well as the molecular and genetic regulators mentioned in the text. Note, more genes are involved in the development of each tissue, for a complete overview of involved regulators see the recent review [5]. complex upregulates IND and SPT, which together regulate targets involved in stigma differentiation [52*]. ...
Article
Angiosperms are the most successful group of land plants. This success is mainly due to the gynoecium, the innermost whorl of the flower. In Arabidopsis, the gynoecium is a syncarpic structure formed by two congenitally fused carpels. At the fusion edges of the carpels, the carpel margin meristem forms. This quasi-meristem is important for medial-tissue development, including the ovules. After the double fertilization, both the seeds and fruit begin to develop. Due to the importance of seeds and fruits as major food sources worldwide, it has been an important task for the scientific community to study gynoecium development. In this review, we present the most recent advances in Arabidopsis gynoecium patterning, as well as some questions that remain unanswered.
... The gynoecium, the female reproductive structure of the flower, is composed in Arabidopsis thaliana of two carpels fused with the repla and topped with the style and stigma (Fig. 1c) 16,17 . Gynoecium patterning is established early during development and is tightly controlled by transcription factors and signaling molecules 18 . ...
Article
Full-text available
Morphogenesis requires the coordination of cellular behaviors along developmental axes. In plants, gradients of growth and differentiation are typically established along a single longitudinal primordium axis to control global organ shape. Yet, it remains unclear how these gradients are locally adjusted to regulate the formation of complex organs that consist of diverse tissue types. Here we combine quantitative live imaging at cellular resolution with genetics, and chemical treatments to understand the formation of Arabidopsis thaliana female reproductive organ (gynoecium). We show that, contrary to other aerial organs, gynoecium shape is determined by two orthogonal, time-shifted differentiation gradients. An early mediolateral gradient controls valve morphogenesis while a late, longitudinal gradient regulates style differentiation. Local, tissue-dependent action of these gradients serves to fine-tune the common developmental program governing organ morphogenesis to ensure the specialized function of the gynoecium.
... Fruits develop primarily from the fourth whorl organ, the ovary, or associated parts of the flower [94]. Scientists have therefore identified several mutations in the genes controlling floral or reproductive development that lead to parthenocarpy by screening and analyzing the plant mutants capable of natural parthenocarpy. ...
Article
Full-text available
Parthenocarpy is an important agricultural trait that not only produces seedless fruits, but also increases the rate of the fruit set under adverse environmental conditions. The study of parthenocarpy in Cucurbitaceae crops has considerable implications for cultivar improvement. This article provides a comprehensive review of relevant studies on the parthenocarpic traits of several major Cucurbitaceae crops and offers a perspective on future developments and research directions.
Article
Since the discovery of brassinolide in the pollen of rapeseed, brassinosteroids (BRs) have consistently been associated with reproductive traits. However, compared to what is known for how BRs shape vegetative development, the understanding of how these hormones regulate reproductive traits is comparatively still lacking. Nevertheless, there is now considerable evidence that BRs regulate almost all aspects of reproduction, from ovule and pollen formation to seed and fruit development. Here, we review the current body of knowledge on how BRs regulate reproductive processes in plants and what is known about how these pathways are transduced at the molecular level. We also discuss how the manipulation of BR biosynthesis and signaling can be a promising avenue for improving crop traits that rely on efficient reproduction. We thus propose that BRs hold an untapped potential for plant breeding, which could contribute to attaining food security in the coming years.
Article
Full-text available
Tomato (Solanum lycopersicum) is a globally cultivated crop with great economic value. The exocarp determines the appearance of tomato fruit and protects it from various biotic and abiotic challenges at both pre-harvest and post-harvest stages. However, no tomato exocarp-specific promoter is currently available, which hinders exocarp-based genetic engineering. Here, we identified by RNA sequencing and reverse transcription-quantitative PCR analyses that the tomato gene SlPR10 (PATHOGENESIS RELATED 10) was abundantly and predominantly expressed in the exocarp. A fluorescent reporter expressed by a 2,087-bp SlPR10 promoter (pSlPR10) was mainly detected in the exocarp of transgenic tomato plants of both Ailsa Craig and Micro-Tom cultivars. This promoter was further utilized for transgenic expression of SlANT1 and SlMYB31 in tomato, which are master regulators of anthocyanin and cuticular wax biosynthesis, respectively. pSlPR10-driven SlANT1 expression resulted in anthocyanin accumulation in the exocarp, conferring gray mold resistance and extended shelf life to the fruit, while SlMYB31 expression led to waxy thickening in the fruit skin, delaying water loss and also extending fruit shelf life. Intriguingly, pSlPR10 and two other weaker tomato exocarp-preferential promoters exhibited coincided expression specificities in the gynophore of transgenic Arabidopsis (Arabidopsis thaliana) plants, providing not only an inkling of evolutionary homology between tomato exocarp and Arabidopsis gynophore but also useful promoters for studying gynophore biology in Arabidopsis. Collectively, this work reports a desirable promoter enabling targeted gene expression in tomato exocarp and Arabidopsis gynophore and demonstrates its usefulness in genetic improvement of tomato fruit quality.
Article
Full-text available
As seed precursors, ovules are fundamental organs during the plant life cycle. Decades of morphological and molecular study have allowed for the elucidation of the complex and intricate genetic network regulating ovule development. Ovule and seed number is highly dependent on the number of ovule primordia that are determined from the placenta during early pistil development. Ovule initiation is positively regulated by the plant hormones auxins, cytokinins, and brassinosteroids, as well as negatively regulated by gibberellins. Each hormone does not act independently; multiple points of hormonal crosstalk occur to coordinately regulate ovule primordia initiation. In this review, we highlight the roles of these hormones and their interactions in the genetic and hormonal network co-regulating ovule initiation in Arabidopsis.
Article
Full-text available
Angiosperms are characterized by the formation of flowers, and in their inner floral whorl, one or various gynoecia are produced. These female reproductive structures are responsible for fruit and seed production, thus ensuring the reproductive competence of angiosperms. In Arabidopsis (Arabidopsis thaliana), the gynoecium is composed of two fused carpels with different tissues that need to develop and differentiate to form a mature gynoecium and thus the reproductive competence of Arabidopsis. For these reasons, they have become the object of study for floral and fruit development. However, due to the complexity of the gynoecium, specific spatio-temporal tissue expression patterns are still scarce. In this study, we used precise laser-assisted microdissection and high-throughput RNA sequencing to describe the transcriptional profiles of the medial and lateral domain tissues of the Arabidopsis gynoecium. We provide evidence that the method used is reliable and that, in addition to corroborating gene expression patterns of previously reported regulators of these tissues, we found genes whose expression dynamics point to being involved in cytokinin and auxin homeostasis and in cell cycle progression. Furthermore, based on differential gene expression analyses, we functionally characterized several genes and found that they are involved in gynoecium development. This resource is available via the Arabidopsis eFP browser and will serve the community in future studies on developmental and reproductive biology.
Article
The appearance of the flower marks a key event in the evolutionary history of plants. Among the four types of floral organs, the gynoecium represents the major adaptive advantage of the flower. The gynoecium is an enclosing structure that protects and facilitates the fertilisation of the ovules, which then mature as seeds. Upon fertilisation, in many species, the gynoecium itself eventually becomes the fruit, which contributes to the dispersal of the seeds. However, despite its importance and the recent advances in our understanding of the genetic regulatory network (GRN) guiding early gynoecium development, many questions remain to be resolved regarding the extent of the conservation of the molecular mechanisms for gynoecium development among different taxa, and how these mechanisms give origin and diversification to the gynoecium. In this review, we compile the existing knowledge about the evolution, development and molecular mechanisms involved in the origin and evolution of the gynoecium.
Article
Full-text available
In seed-bearing plants, the ovule (“small egg”) is the organ within the gynoecium that develops into a seed after fertilization. The gynoecium located in the inner compartment of the flower turns into a fruit. The number of ovules in the ovary determines the upper limit or the potential of seed number per fruit in plants, greatly affecting the final seed yield. Ovule number is an important adaptive characteristics for plant evolution and an agronomic trait for crop improvement. Therefore, understanding the mechanism and pathways of ovule number regulation becomes a significant research aspect in plant science. This review summarizes the ovule number regulators and their regulatory mechanisms and pathways. Specially, we construct the first integrated molecular network for ovule number regulation, in which phytohormones played a central role, followed by transcription factors, enzymes, other protein and micro-RNA. Of them, AUX, BR and CK are positive regulator of ovule number, whereas GA acts negatively on it. Interestingly, many ovule number regulators have conserved functions across several plant taxa, which should be the targets of genetic improvement via breeding or gene editing. Many ovule number regulators identified to date are involved in the diverse biological process, such as ovule primordia formation, ovule initiation, patterning, and morphogenesis. The relations between ovule number and related characteristics/traits especially of gynoecium/fruit size, ovule fertility, and final seed number, as well as upcoming research questions, are also discussed. In summary, this review provides a general overview of the present finding in ovule number regulation, which represents a more comprehensive and further cognition on it.
Article
Full-text available
The stigma is an angiosperm-specific tissue that is essential for pollination. In the last two decades, several transcription factors with key roles in stigma development in Arabidopsis thaliana have been identified. However, genetic analyses have thus far been unable to unravel the precise regulatory interactions among these transcription factors or the molecular basis for their selective roles in different spatial and temporal domains. Here, we show that the NGATHA (NGA) and HECATE (HEC) transcription factors, which are involved in different developmental processes but are both essential for stigma development, require each other to perform this function. This relationship is likely mediated by their physical interaction in the apical gynoecium. NGA/HEC transcription factors subsequently upregulate INDEHISCENT (IND) and SPATULA and are indispensable for the binding of IND to some of its targets to allow stigma differentiation. Our findings support a non-hierarchical regulatory scenario in which the combinatorial action of different transcription factors provides exquisite temporal and spatial specificity of their developmental outputs.
Article
Full-text available
With growing populations and pressing environmental problems, future economies will be increasingly plant-based. Now is the time to reimagine plant science as a critical component of fundamental science, agriculture, environmental stewardship, energy, technology and healthcare. This effort requires a conceptual and technological framework to identify and map all cell types, and to comprehensively annotate the localization and organization of molecules at cellular and tissue levels. This framework, called the Plant Cell Atlas (PCA), will be critical for understanding and engineering plant development, physiology and environmental responses. A workshop was convened to discuss the purpose and utility of such an initiative, resulting in a roadmap that acknowledges the current knowledge gaps and technical challenges, but also underscores how the PCA initiative can help to overcome them.
Article
Full-text available
In the model plant Arabidopsis thaliana, the zinc-finger transcription factor KNUCKLES (KNU) plays an important role in the termination of floral meristem activity, a process that is crucial for preventing the overgrowth of flowers. The KNU gene is activated in floral meristems by the floral organ identity factor AGAMOUS (AG), and it has been shown that both AG and KNU act in floral meristem control by directly repressing the stem cell regulator WUSCHEL (WUS), which leads to a loss of stem cell activity. When we re-examined the expression pattern of KNU in floral meristems, we found that KNU is expressed throughout the center of floral meristems, which includes, but is considerably broader than the WUS expression domain. We therefore hypothesized that KNU may have additional functions in the control of floral meristem activity. To test this, we employed a gene perturbation approach and knocked down KNU activity at different times and in different domains of the floral meristem. In these experiments we found that early expression in the stem cell domain, which is characterized by the expression of the key meristem regulatory gene CLAVATA3 (CLV3), is crucial for the establishment of KNU expression. The results of additional genetic and molecular analyses suggest that KNU represses floral meristem activity to a large extent by acting on CLV3. Thus, KNU might need to suppress the expression of several meristem regulators to terminate floral meristem activity efficiently.
Article
Full-text available
Symmetry establishment is a critical process in the development of multicellular organs and requires careful coordination of polarity axes while cells actively divide within tissues. Formation of the apical style in the Arabidopsis gynoecium involves a bilateral-to-radial symmetry transition, a stepwise process underpinned by the dynamic distribution of the plant morphogen auxin. Here we show that SPATULA (SPT) and the HECATE (HEC) bHLH proteins mediate the final step in the style radialisation process and synergistically control the expression of adaxial-identity genes, HOMEOBOX ARABIDOPSIS THALIANA 3 ( HAT3 ) and ARABIDOPSIS THALIANA HOMEOBOX 4 ( ATHB4 ). HAT3/ATHB4 module drives radialisation of the apical style by promoting basal-to-apical auxin flow and via a negative feedback mechanism that finetune auxin distribution through repression of SPT expression and cytokinin sensitivity. Thus, this work reveals the molecular basis of axes-coordination and hormonal cross-talk during the sequential steps of symmetry transition in the Arabidopsis style.
Article
Full-text available
The appearance of plant organs mediated the explosive radiation of land plants, which shaped the biosphere and allowed the establishment of terrestrial animal life. The evolution of organs and immobile gametes required the coordinated acquisition of novel gene functions, the co-option of existing genes and the development of novel regulatory programmes. However, no large-scale analyses of genomic and transcriptomic data have been performed for land plants. To remedy this, we generated gene expression atlases for various organs and gametes of ten plant species comprising bryophytes, vascular plants, gymno-sperms and flowering plants. A comparative analysis of the atlases identified hundreds of organ- and gamete-specific ortho-groups and revealed that most of the specific transcriptomes are significantly conserved. Interestingly, our results suggest that co-option of existing genes is the main mechanism for evolving new organs. In contrast to female gametes, male gametes showed a high number and conservation of specific genes, which indicates that male reproduction is highly specialized. The expression atlas capturing pollen development revealed numerous transcription factors and kinases essential for pollen biogenesis and function.
Article
With growing populations and pressing environmental problems, future economies will be increasingly plant-based. Now is the time to reimagine plant science as a critical component of fundamental science, agriculture, environmental stewardship, energy, technology and healthcare. This effort requires a conceptual and technological framework to identify and map all cell types, and to comprehensively annotate the localization and organization of molecules at cellular and tissue levels. This framework, called the Plant Cell Atlas (PCA), will be critical for understanding and engineering plant development, physiology and environmental responses. A workshop was convened to discuss the purpose and utility of such an initiative, resulting in a roadmap that acknowledges the current knowledge gaps and technical challenges, and underscores how the PCA initiative can help to overcome them.
Article
Cell wall modifications are of pivotal importance during plant development. Among cell wall components, xyloglucans are the major hemicellulose polysaccharide in primary cell walls of dicots and non graminaceous monocots. They can connect the cellulose microfibril surface to affect cell wall mechanical properties. Changes in xyloglucan structure are known to play an important role regulating cell growth. Therefore, the degradation of xyloglucan is an important modification that alters the cell wall. The α-XYLOSIDASE1 (XYL1) gene encodes the only α-xylosidase acting on xyloglucans in Arabidopsis thaliana. Here, we show that mutation of XYL1 strongly influences seed size, seed germination, and fruit elongation. We found that the expression of XYL1 is directly regulated in developing seeds and fruit by the MADS-box transcription factor SEEDSTICK (STK). We demonstrate that XYL1 complements the stk smaller seed phenotype. Finally, by Atomic Force Microscopy (AFM), we investigate the role of XYL1 activity in maintaining cell stiffness and growth, confirming the importance of cell wall modulation in shaping organs.
Article
Significance In floral meristem (FM), the transcription factor (TF) WUS and the secreted peptide CLV3 form a feedback loop to maintain robust stem cell activities. We found that the TF KNU is a temporally regulated repressor of WUS . However, how the spatial expression patterns of CLV3 and WUS are fully extinguished during floral organogenesis is unknown. Here, we show that KNU has position-specific multifunctions that achieve termination of robust FM. Besides repressing WUS in the organizing center, in the central zone, KNU directly represses CLV3 and inhibits WUS from sustaining CLV3 . Furthermore, KNU represses CLV1 and disrupts WUS–WUS and WUS–HAM1 interactions. Thus, KNU plays an essential role to promote FM determinacy for the proper formation of floral reproductive organs.
Article
Development of multicellular organisms is a complex process involving precise coordination of growth among individual cells. Understanding organogenesis requires measurements of cellular behaviors over space and time. In plants, such a quantitative approach has been successfully used to dissect organ development in both leaves and external floral organs such as sepals. However, the observation of floral reproductive organs is hampered as they develop inside tightly closed floral buds, and are therefore difficult to access for imaging. We developed a confocal time-lapse imaging method, applied here to Arabidopsis (Arabidopsis thaliana), that allows full quantitative characterization of the development of stamens, the male reproductive organs. Our lineage tracing reveals the early specification of the filament and the anther. Formation of the anther lobes is associated with a temporal increase of growth at the lobe surface that correlates with intensive growth of the developing locule. Filament development is very dynamic and passes through three distinct phases: (1) initial intense, anisotropic growth and high cell proliferation; (2) restriction of growth and proliferation to the filament proximal region; (3) resumption of intense and anisotropic growth, displaced to the distal portion of the filament, without cell proliferation. This quantitative atlas of cellular growth dynamics provides a solid framework for future studies into stamen development.