Next Article in Journal
Spatio-Temporal Changes of Extracellular Matrix (ECM) Stiffness in the Development of the Leech Hirudo verbana
Next Article in Special Issue
Genome-Wide Identification of Homeodomain Leucine Zipper (HD-ZIP) Transcription Factor, Expression Analysis, and Protein Interaction of HD-ZIP IV in Oil Palm Somatic Embryogenesis
Previous Article in Journal
Atraric Acid Ameliorates Hyperpigmentation through the Downregulation of the PKA/CREB/MITF Signaling Pathway
Previous Article in Special Issue
Reducing Pre- and Post-Treatments in Cryopreservation Protocol and Testing Storage at −80 °C for Norway Spruce Embryogenic Cultures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Induction of Somatic Embryogenesis in Plants: Different Players and Focus on WUSCHEL and WUS-RELATED HOMEOBOX (WOX) Transcription Factors

Department of Agriculture Food and Environment, University of Pisa, Via del Borghetto 80, 56124 Pisa, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(24), 15950; https://doi.org/10.3390/ijms232415950
Submission received: 4 November 2022 / Revised: 6 December 2022 / Accepted: 8 December 2022 / Published: 15 December 2022

Abstract

:
In plants, other cells can express totipotency in addition to the zygote, thus resulting in embryo differentiation; this appears evident in apomictic and epiphyllous plants. According to Haberlandt’s theory, all plant cells can regenerate a complete plant if the nucleus and the membrane system are intact. In fact, under in vitro conditions, ectopic embryos and adventitious shoots can develop from many organs of the mature plant body. We are beginning to understand how determination processes are regulated and how cell specialization occurs. However, we still need to unravel the mechanisms whereby a cell interprets its position, decides its fate, and communicates it to others. The induction of somatic embryogenesis might be based on a plant growth regulator signal (auxin) to determine an appropriate cellular environment and other factors, including stress and ectopic expression of embryo or meristem identity transcription factors (TFs). Still, we are far from having a complete view of the regulatory genes, their target genes, and their action hierarchy. As in animals, epigenetic reprogramming also plays an essential role in re-establishing the competence of differentiated cells to undergo somatic embryogenesis. Herein, we describe the functions of WUSCHEL-RELATED HOMEOBOX (WOX) transcription factors in regulating the differentiation–dedifferentiation cell process and in the developmental phase of in vitro regenerated adventitious structures.

1. Introduction

Higher plants retain the capacity of unlimited growth that relies on the mitotic activity of cells located in the apical and root meristems, which continuously divide to renew themselves and produce cells for organ formation [1]. Plant developmental processes are very flexible compared to those described in animals. Hydra, planarians, and echinoderms preserve interstitial cells and neoblasts (Figure 1), which can replace differentiated cells lost after injury [2].
The observation that genes critical for proper embryonic development in higher animals, such as those coding for T-domain transcription factors or involved in Wnt/Wingless signaling, are expressed during de novo Hydra head regeneration has led to important insights into the molecular basis of animal self-organization [3]. Furthermore, some differentiated cells can switch their fate to acquire a new one (trans-differentiation): in axolotls (Figure 1), neural cells can trans-differentiate into muscle and cartilage [4], and zebrafish can regenerate their hearts fully, even after 20% of the ventricle is removed [5,6]. In mammals, trans-differentiation can involve liver cells [7]. Nevertheless, in the animal kingdom, only the zygote can produce a mature organism that requires the ability to generate all the cells of the body, as well as to organize them into a specific temporal and spatial sequence, that is, to undergo a coordinated process of development. In vertebrates, virgin birth is the development of an embryo from a female gamete (parthenogenesis), an unusual process [8,9,10,11]; by contrast, clonal reproduction is more frequently observed in several invertebrates [12,13,14]. Therefore, totipotency, in this strict sense, is established by the ability of an isolated cell to produce a fertile, adult individual [15].
In plants, the zygote is not the only cell that can develop into an embryo. Zygotic embryogenesis depends on fertilization, but apomictic embryos produced in certain species provide clear evidence of in vivo embryo development without sex [16,17,18,19,20,21]. In epiphyllous species (Figure 1), even somatic cells of vegetative organs can spontaneously form in vivo ectopic embryos [19,22,23,24,25,26,27,28]. For a long time, the first hints about how plant cells might regulate totipotency came from empirical approaches based on in vitro culture methods. In 1902, Haberlandt proposed the concept of plant cell totipotency [29], hypothesizing that entire plants could be generated from somatic cells. Experimental evidence supporting this hypothesis was lacking until 1958, when Steward et al. [30] demonstrated that segments of differentiated secondary phloem tissue from carrot could regenerate whole plants, thus highlighting the remarkable totipotent potential of somatic cells. This result was shortly confirmed [31].
The differentiation process leads to a range of determinate cellular fates, which are functional to several organs of the mature organism. The functions are performed efficiently and to the benefit of the whole organism, but at the price that the specialized cells have only limited parts of their genome open for transcription. Nevertheless, cell commitment is reversible, and new developmental patterns can arise as a response to external or internal factors [32,33,34,35,36,37,38,39,40,41]. The recurrent ontogenesis presumes that each differentiated cell can restore the morphogenetic potential of the zygote. This feature is evident during in vitro regeneration when the positional signals are strongly altered, and the differentiated cells can suddenly change identities and follow a gradual process of de-differentiation [38,42]. Thus, differentiated cells can completely revert to the meristematic state or, in some cases, to the somatic embryogenic one (totipotency) [39]. The mechanisms allowing the conservation of the stem cell niche and the change of cellular fate are only partially understood. Nevertheless, it has been demonstrated that somatic embryogenesis can be achieved by ectopic overexpression of some genes, which is crucial for the formation and maintenance of the shoot apical meristem (SAM) and embryo development [43,44,45]. Among them are those encoding the homeodomain transcription factors (TFs) WUSCHEL (WUS), AINTEGUMENTA-LIKE (AIL) AP2/ERF-domain TF PLETHORA4/BABY BOOM (PLT4/BBM) or PLT5/EMBRYO MAKER (PLT5/EMK), MADS-box TF AGAMOUS-LIKE15 (AGL15), NF-Y (nuclear factor of the Y box) TF LEAFY COTYLEDON1 (LEC1), B3 TF LEC2 and FUSCA3 (FUS3), MYB TF MYB118 and RWP-RK DOMAIN-CONTAINING4 (RKD4)/GROUNDED (GRD), class I KNOX homeodomain SHOOT MERISTEMLESS1 (STM1), and the receptor kinase CLAVATA1 (CLV1) [46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65]. In addition, a large number of other TFs that are differentially activated during somatic embryogenesis have been identified through time-course analyses (e.g., Somatic Embryogenesis Receptor Kinase (SERK), CUP-SHAPED COTYLEDON1 (CUC1) and CUC2, STM-Like [66,67,68,69]. The broad range of regulatory factors involved in acquiring embryogenic competence suggests that numerous independent and interrelated pathways are functional in this process. Whether and how these genes work together to promote embryogenic cell formation is unknown [70]. It is likely that epigenetic changes can decide the cell fate, and that they constitute a layer of control superimposed on the activity of the plethora of TFs which are involved in these processes [71]. For example, several reports have suggested that epigenetic mechanisms may have an essential role in cellular de-differentiation, changing cell fate, and in plant totipotency [36,39,45,72,73,74].
Although much is known about the patterning of the shoot meristem during embryogenesis, there is little understanding of patterning that must occur during de novo induction of plant tissues in culture. For example, by overexpressing WUS, the formation of somatic embryos (SEs) from zygotic embryos and vegetative tissue was enhanced. These SEs can develop normally and germinate into typical plants [70]. However, how WUS regulates the identity of totipotent cells directed toward SE initiation and development needs to be better understood. In this review, we discuss the role of WUS-related homeobox (WOX) TFs in the induction of cell totipotency as well as the regulation of somatic embryogenesis, also considering the interactions of WOX with other TFs, plant growth regulators (PGRs), and epigenetic signals.

2. WUS-Related Homeobox (WOX) Genes

2.1. Nomenclature and Molecular Characteristic Motifs

Genes codifying for homeobox proteins contain a widely conserved domain of 180 bp, which encodes for a globular homeodomain (HD) of approximately 60 amino acids usually involved in DNA binding [75,76]. Likely, the first homeobox genes could have appeared during early eukaryote evolution, probably deriving from a helix-turn-helix (HLH) TF [77]. In plants, the first homeobox gene was identified in the early 1990s. Vollbrecht et al. [78] resumed the study of an old maize mutant named Knotted-1 (Kn-1) in which clusters of cells located along the lateral leaf veins continued to divide, forming characteristic growths known as knots [79].
The WUS-related homeobox (WOX) genes encode for a class of plant-specific homeobox TFs, which perform several key functions in plant development processes. These include, for example, organization and embryonic development, maintenance of stem cells, and formation of various organs [1,80,81,82,83,84,85,86]. These functions may be related to the increase in cell division and/or to the prevention of early cell differentiation. In Arabidopsis, 15 different WOX genes were identified. An analysis obtained by a phylogenetic tree divided the WOX TFs into three main clades: WUS, INTERMEDIATE, and ANCIENT [87]. The specific function of each WOX gene is determined by spatiotemporal expression patterns and by the interaction of their products with other proteins [88]. For example, WUS acts mainly as a repressor in stem cell populations in shoot meristems (SAM) formation, but becomes an activator when involved in the regulation of the AGAMOUS (AG) gene in flower development [80,87,89,90,91,92].
In WOX sequences, the helix-loop-helix-turn-helix (HLHTH) motif is characterized by three α-helices joined by a short loop and a short turn (Figure 2A,B).
The second helix binds DNA through hydrogen bonds and hydrophobic interactions. In particular, specific regions of the protein chains establish a bond with the exposed bases and with thymine methyl groups within the major DNA groove. Analysis of the tertiary structure of sunflower (Helianthus annuus) HaWUS proteins revealed highly conserved structures compared to the homeodomain of other species (Figure 2C) [84,88,93].
In the carboxy-terminal region, members of the WUS clade also contain the WUS-box motif T-LX-LFPXX (where T stands for Threonine, L for Leucine, F for Phenylalanine, P for Proline, and X for any amino acid), which distinguishes them from other homeobox TFs (Figure 2A). The WUS-box motif is essential for WUS functions, regulating the stem cell population of SAM and controlling flower development [88,89]. In addition, members of the WUS clade possess a motif of two amino acids (Threonine and Leucine) at the beginning of the WUS-box, whereas the WOX proteins belonging to the other clades show amino acid variations in the beginning position. Some WOX proteins contain a series of amino acids between the HD and the WUS-box that could potentially act as an activation domain, and/or show a carboxy-terminal ERF-associated Amphiphilic Repression (EAR) domain (Figure 2A) [88,93,94,95,96,97,98,99,100,101,102,103]. The EAR motif structure identified in Arabidopsis is [LVI]-X-[LVI]-X-[LVI] (where V represents Valine). This motif has been detected for all three major WOX clades members [87]. The C-terminal EAR domains in WUS, WOX5, and WOX7 regulate the repression activity in vegetative and floral meristems [89]. This can be partially mediated by interaction with other proteins, such as TOPLESS, which interacts with WUS through the EAR motif [89,104,105]. In Oryza sativa, WOX11 and WOX3 do not possess the EAR motif; nevertheless, they can act as repressors. For instance, WOX11 directly represses the transcription of the Two-component response regulator OsRR2 (RR2) gene, which encodes for a TF that acts as a negative regulator of cytokinin (CK) signaling [106]. In rice, WOX3 can block the gene involved in synthesizing the YABBY3 (YAB3) TF during leaf development, controlling the abaxial cell fate [89,107]. In addition, a simple L-X-L motif has been detected in all Arabidopsis WOX proteins, except for WOX8 and WOX10 [87]. Most members of the WOX family work as TFs; however, no Nuclear Localization Signals (NLS) have been predicted for any of them. Nevertheless, WUS, WOX6 [108], and WOX11 [106] are located inside the nucleus. This nuclear localization could indicate NLS motifs which have not yet been detected, or the interaction with other proteins that support NLS function.

2.2. In Vivo Roles of WOX Genes in Plants: Developmental Aspects and Stress Response

In Arabidopsis, WUS and WOX5 genes maintain stem cell functioning in SAM cells and root apical meristem (RAM) [109]. In particular, expression of WUS begins in the 16-cell stage embryo in two inner apical cells, and maintains a tightly restricted pattern throughout embryogenesis [80,110,111]. WOX5 is also expressed in the early stages of lateral root and cotyledon development [112]. In the case of WUS inactivation, totipotent cells, whose fate is regulated by signals deriving from the Organization Center (OC) of the SAM, underwent differentiation in both Arabidopsis and Antirrhinum majus [110,113]. In Arabidopsis, maize, and rice, WUS also controls the development of ovules and anthers [114,115]. wox5 mutants of Arabidopsis showed that totipotent cells of columella root undergo differentiation processes [109]. WOX3 of Arabidopsis and the orthologous NARROW SHEATH 1 (NS1) and NS2 of Zea mays regulate the recruitment of initial organ cells in response to signals from peripheral regions of meristems, promoting cell proliferation [116]. In Arabidopsis, WOX6 prevents premature differentiation during the formation of the integuments that envelop the embryonic sac and the egg cell [108]. A further role of WOX6 has been identified through the isolation of the hos9-1 mutant, characterized by slower growth and late flowering, with an accentuated sensitivity to low temperatures that leads the plant to undergo extensive freezing phenomena [117]. In Arabidopsis, WOX2 is regulated by WOX8 and WOX9, and is required to form the SAM during embryonic development [118]. WOX1, WOX3, WOX5, WOX8, and WOX9 are redundantly expressed together with WOX2 during SAM formation [118]. In the early stages of plant development, WOX2 and WOX8 are co-expressed in the zygote. However, during embryonic development, the expression of WOX2 is restricted in the apical zone, while WOX8 and WOX9 are restricted to the basal region [119]. In fact, the products of the WOX8 and WOX9 genes are required in large quantities in Arabidopsis to develop the micropylar region of the embryo in suspensor and hypophysis organization. Therefore, the molecular system controlled by WOX genes establishes the central axis of the embryo and regulates the localized response to auxin by interacting with the auxin transporter PIN-FORMED1 (PIN1) [88]. WOX9 is also strategic for SAM maintenance [120] and cell division activity during embryonic and post-embryonic development in Arabidopsis [121], Solanum esculentum [122], and Petunia hybrida [123]. In Arabidopsis, the genes of the ANCIENT clade WOX13 and WOX14 are expressed in primary roots, lateral roots, and floral organs, precluding premature cell differentiation [124]. The development of an adaptable root system is essential for firm crop production under variable environments. In rice plants, two lateral roots have been described: S-type (short and thin) and L-type (long, thick, and capable of further branching). Recently, it was evidenced that two WOX genes have opposing roles in controlling LR primordium (LRP) size in this crop [125].
In a postulated model that regards the SAM, vegetative stem cell homeostasis is controlled by the interaction between WUS and the CLV genes, which are required to rapidly downregulate WUS in apical daughter cells after cell division [126,127,128,129,130]. WUS promotes the expression of CLV3, which, via activation of the signaling pathway dependent by the CLV1/CLV2 protein complex, acts on WUS transcription and limits the size of the WUS-expressing OC [126,131]. The ability of the protein WUS to migrate to adjacent cells and activate its negative regulator is unique to plant stem cell niches [83]. An additional negative feedback mechanism, implying both WUS and AG genes, is responsible for the maintenance of floral stem cells. In particular, WUS activates the transcription of the AG gene, which is responsible for the identity of the flower organs. Later, AG represses the WUS transcription with negative feedback [132]. In Arabidopsis roots, the signal peptide CLAVATA3/EMBRYO SURROUNDING REGION40 (CLE40) and ARABIDOPSIS CRINKLY4 kinase (ACR4) (a member of the receptor-like kinase family CRINKLY4) have been identified as regulators of WOX5 activity, with a negative loop mechanism similar to that involving WUS and CLV [133]. It has also been shown that WUS can repress the transcription of some Type-A Arabidopsis Response Regulators (ARR-A) genes, which encode negative regulators of signaling mechanisms originating from the CK [134]. ARR-A proteins probably compete for phosphorylation with positive ARR-B regulators [81].
Recently, the role of WOX genes has been extended in bulbil differentiation, a vegetative propagation strategy of Lilium lancifolium [135]. LlWOX9 and LlWOX11 are positive regulators of bulbil formation. Cytokinin type-B LlRRs can bind to the promoters of LlWOX9 and LlWOX11 to promote their transcription. In addition, LlWOX11 can enhance cytokinin signaling by inhibiting the transcription of type-A LlRR9 [135]. In addition, the asexual propagation in three Crassulaceae species showed the concomitant expression of WUS with the shoot induction phase on leaves [136].
In plants, tissue repair is essential to counteract damage-associated stress; moreover, repair processes are required in grafting events. Recently, the involvement of WOX13, an ancient member of the WOX family, in callus formation and organ adhesion in A. thaliana has been reported [137]. After wounding, WOX13 expression was rapidly induced, and this response was partly dependent on the activity of WOUND-INDUCED DEDIFFERENTIATION 1 (WIND1). Interestingly, WOX13 directly upregulated WIND2 and WIND3 to reinforce cellular reprogramming and organ regeneration [137]. To elucidate the role in callus formation and tissue repair of WOX13, a schematic diagram was reported by Ikeuchi et al. [137].
Recently, constructive research has finally interpreted the molecular mechanism of meristematic cell resistance to viruses, an essential phenomenon in the health restoration of crops [138,139]. WUS responds to cucumber mosaic virus (CMV) infection and blocks virus accumulation in the central and peripheral zones of the meristem. In particular, WUS hinders viral protein synthesis by repressing the expression of plant S-adenosyl-L-methionine–dependent methyltransferases with a role in ribosomal RNA processing and ribosome stability [138].
Environmental factors can significantly influence the development and final yield of many crops. To cope with environmental stresses, plants have evolved various defense mechanisms. At the gene level, one of the most effective mechanisms is the regulation of specific genes encoding TFs [140]. Among these, it has been shown that WOX genes also play a key role in response to multiple types of stress. For example, in rice, overexpression of the OsWOX13 gene under the control of the rab21 promoter increases the tolerance to drought [141]. In Jatropha curcas, water deprivation strongly reduces the expression of both JcWOX5 and JcWOX6 genes, while intense saline stress induces the expression of JcWOX1 and JcWOX8 [140]. In Gossypium hirsutum, over half of WOX genes show no significant response to heat, cold, salinity, or drought. By contrast, the GhWOX10_Dt, GhWOX13a_At/Dt, and GhWOX13b_At/Dt genes are strongly induced under multiple stress [94]. In the hybrid Populus alba × P. glandulosa, the PagWOX11/12a gene is predominantly expressed in roots under drought [142]. In Camellia sinensis, the lack of water activates CsWOX13 and CsWOX15, but inhibits the expression of CsWOX1 and CsWOX9. Tea plants subjected to cold conditions showed reduced expression of both CsWOX9 and CsWOX14. Furthermore, the CsWOX5, CsWOX3, and CsWOX2 genes are significantly regulated by exogenous treatments with ethylene, abscisic acid, methyl-jasmonate, and gibberellic acid, respectively [143]. In Brassica napus, 58 WOX genes have been identified and characterized. Analysis of their expression levels showed that four members belonging to the WOX4 sub-clade (BnWOX10, BnWOX50, BnWOX44, and BnWOX18) are activated by abiotic stress or PGR treatments [95].

3. WOX Genes Activity during Embryogenesis and Interaction with Plant Growth Regulators

3.1. Zygotic Embryogenesis

Different embryogenic processes share a common PGR control that essentially relies on the action of auxin, abscisic acid (ABA), cytokinins (CKs), and gibberellins (GAs). In particular, the establishment of auxin synthesis and polar auxin transport (PAT) are key steps in embryo development [144]. During early zygotic embryogenesis of Arabidopsis, auxin accumulates dynamically at specific positions that correlate with developmental cell fate. In such a way, the developmental decision is tied to both auxin transport regulators and components of response machinery [145]. Members of the PIN FORMED (PIN) family of auxin efflux facilitators are the key players in determining the direction of intercellular auxin flow [146,147,148]. After zygotic division, PIN7 localizes at the upper side of basal cell descendants and reverses orientation at the globular stage. The localization of PIN1 becomes polarized from the midglobular stage onward. These localizations correlate with an auxin maximum in the apical cell at the early stages of embryogenesis, as well as relocation of this maximum to the basal pole at the globular stage. From the embryo’s late-globular stage onward, PIN4 is detected along the surface of the hypophysis and at the lower side of the adjacent suspensor cell [146]. The polar localization of PIN1 is disrupted in the gnom mutant embryos that do not undergo coordinated development [149]. GNOM is required for PIN1 recycling between the plasma membrane and endosomes [145]. In extreme cases, loss of multiple PIN family members phenocopies gnom embryos. When the endocytosis-dependent mechanism of PIN polarity generation is altered, auxin response in apical embryo regions increases, leading to a cell fate change from cotyledon to root [147]. In A. thaliana, specific dynamics of WOX gene expression identified cell fate decisions during zygotic embryogenesis. For example, WOX2 and WOX8, initially coexpressed in the zygote, act as complementary cell fate regulators in the apical and basal lineage, respectively, during the apical–basal axis formation. WOX2 expression in the apical lineage is an early and important downstream function of WOX8/WOX9 activity, and the WOX transcriptional machinery is linked with the establishment of a localized auxin response in the proembryo [118]. Moreover, the expression of WOX8 is independent of the axis patterning signal auxin; by contrast, with the redundant gene WOX9, it is activated in the zygote, its basal daughter cell, and the hypophysis by the zinc-finger transcription factor WRKY2 [150]. It is interesting to note that an ancestral role of WOX in seed plant embryo development has been demonstrated, thus corroborating the proposed connection between PAT, PIN-FORMED (PIN), and WOX in regulating embryo patterning in seed plants [151].
Analysis of the expression pattern of WOX genes in maize during zygotic embryogenesis suggested that apical and basal cell lineages and cell fate determination in grasses may involve different players, or occur at a later embryonic stage compared to A. thaliana [152].
In Picea abies zygotic embryogenesis, Zhu et al. [153] revealed that PaWOX2 plays a more crucial role during early embryogenesis (i.e., protoderm formation and suspensor expansion) than during late embryogenesis. This essential function was evident because the downregulation of PaWOX2 at the beginning of embryo development caused a significant decrease in the yield of mature embryos. By contrast, the downregulation of PaWOX2 after late embryos were formed did not affect further embryo development or maturation [153].

3.2. Somatic Embryogenesis

Somatic cell reprogramming and transition to embryogenic cell competence also require high concentrations of auxin (primarily 2,4-dichlorophenoxyacetic acid (2,4-D), which is a synthetic auxin-like plant growth regulator), which presumably acts to trigger a signaling cascade in the modulation of numerous SE-associated TF genes [34,39,144]. A transient increase in endogenous auxin (IAA) levels has been observed in various systems [154]. For example, 2,4-D must be removed from the culture medium to induce SE formation from the embryonic callus. The removal of 2,4-D activates the expression of YUCCA2 (YUC2) and YUC4, which encode auxin biosynthetic enzymes, resulting in increased endogenous IAA levels [155,156]. In turn, the different expression of the genes encoding the core components of the auxin-signaling pathway, the AUXIN/INDOLE-3-ACETIC ACIDs (Aux/IAAs), and AUXIN RESPONSE FACTORs (ARFs), was demonstrated to accompany SE induction. In particular, a crucial role of asymmetric auxin distribution has been established for de novo morphogenesis in in vitro SEs using live imaging analysis of PIN gene expression [157].
In Arabidopsis, the results of Su et al. [155] established that an auxin gradient and PIN1-mediated PAT are essential for WUS induction and somatic embryogenesis. Su et al. [155] also showed that other regulatory genes of zygotic embryogenesis were upregulated during SE development. Nonetheless, WUS expression was identified within the embryonic callus when SEs could not be morphologically recognized. In addition, Su et al. [158] demonstrated that the expression patterns of several regulatory genes critical for RAM formation were correlated with the establishment of the embryonic root meristem during somatic embryogenesis in Arabidopsis. Notably, early expression of WOX5 and WUS genes was induced, and nearly overlapped within the embryonic callus when SEs could not be identified morphologically. In addition, cytokinin response signals were detected in specific regions correlated with induced WOX5 expression and subsequent SE formation. Overexpression of ARR7 and ARR15 (feedback repressors of cytokinin signaling) disturbed RAM initiation and SE induction. These results provided information regarding auxin and cytokinin-regulated apical-basal polarity formation of the shoot–root axis during somatic embryogenesis [158].
In Medicago truncatula, a comparative analysis of the expression of WOX and PIN genes in ovules, and of the course of somatic embryogenesis, was performed [159]. MtWOX11-like and MtPIN10 showed an increased expression level in ovules, and were activated during SE development. By contrast, in other WOX and PIN genes, the expression level was low in ovules and did not show transcription activation associated with somatic embryogenesis. These results confirmed that the exact regulatory mechanisms could control the early stages of somatic and zygotic embryogenesis [159].
In Picea abies, PaWOX2 expression was highest during embryogenesis at the earliest stages of development. However, no activity was detected in non-embryogenic cell culture, indicating that PaWOX2 plays a fundamental role in early SE development and can be used as a possible marker for embryogenic potential [160]. It has also been shown that the PAT inhibitor NPA impairs embryo morphology and increases the expression of PIN1 during Picea abies SE development [161]. The results, in Gymnosperms as well, strengthen the proposed connection between PAT and WOX in the regulation of embryo patterning in seed plants [161,162]. The close link between PGRs and WOX genes is expected to induce somatic embryogenesis. In Triticum aestivum, TaWOX5 was primarily expressed in the root and calli induced by auxin and cytokinin, indicating that TaWOX5 may be related to root formation or development, and is associated with regulation of PGRs in somatic embryogenesis [163]. In the interspecific hybrid Liquidambar styraciflua × Liquidambar formosana, auxin, cytokinin signal transduction, and biosynthesis-related genes were significantly expressed during somatic embryogenesis. Among these, there were many auxin signal transduction genes. In particular, small auxin up RNA (SUAR) family genes were predominant [164]. In embryogenic callus, the contents of many auxin-related genes, such as ARF17, ARF18, and AUX1, were significantly higher than in non-embryogenic callus. In particular, Qi et al. [164] found that PIN1-like, PIN2, AUX1, ARFs, GH3, and SAUR were dramatically upregulated in embryogenic callus and downregulated during SE development, indicating that the auxin-responsive genes played a critical role mainly in the transition from vegetative to embryogenic competent cells. Analogously, WOX9 was specifically expressed in embryogenic callus and downregulated during SE development [164]. By contrast, WOX11 was highly expressed throughout SE phases and almost not expressed in non-embryogenic callus or vegetative organs. Hence, these genes might play an essential role in the somatic embryogenesis of hybrid sweetgum, but at a different stage of SE initiation and development.
Time course experiments have often been designed to identify TFs involved in the induction and development of SEs [66,67,68]. In Lactuca sativa, the expression LsWUS1L was almost absent in the early stages of in vitro culture (1–4 days) in the presence of cytokinin and auxin (with a 4:1 ratio), but it gradually increased during the following days (7–12) of culture [69]. Since the onset of the organization of adventitious structures was observed from seven to twelve days, the time-dependent increase in the transcript level of LsWUS1L could be related to the development of buds and/or SEs, as well as the initial organization of SAMs. By contrast, LsWUS2L showed a highly significant increase of transcripts up to four days of in vitro culture (Figure 3).
The sudden increase in the transcriptional activity of LsWUS2L appears remarkable because, simultaneously, the histological analysis showed that the groups of poorly vacuolated cells, with large nuclei and rich in cytoplasm, began to be present in the explants (Figure 3). These cells subsequently organized themselves into SEs and adventitious buds. Therefore, these results suggest that LsWUS2L could be involved in the induction phase of embryogenic competence [69].

3.3. WOX Genes Activation and Shoot Regeneration

For Arabidopsis, Gordon et al. [165] induced new shoot meristems from cultured root explants. These authors characterized early patterning during the de novo development of the Arabidopsis shoot meristems using fluorescent reporters of known genes and proteins required for SAM development and maintenance. They demonstrated that a small number of progenitor cells initiate the development of new shoot meristems through stereotypical stages of reporter expression and activity of CUC2, WUS, PIN1, STM, FILAMENTOUS FLOWER (FIL), REVOLUTA (REV), ARABIDOPSIS THALIANA MERISTEM L1 LAYER (ATML1), and CLV3. These genes, also interacting with PGRs (cytokinins and auxins), participate in various stages of initiation and coordinated development of the adventitious meristems. In particular, a functional requirement for WUS activity during de novo shoot meristem initiation was detected, demonstrating the interaction of WUS with cytokinins [165]. This relationship was confirmed by Dai et al. [166], which demonstrated, by chromatin immunoprecipitation (ChIP) and transient activation assays, that in plant cells cultured in vitro, ARR12 binds to the promoter of WUS. Therefore, during shoot regeneration, ARR12 acts as a molecular link between cytokinin signaling and the expression of WUS.
Regarding in vitro regeneration methods, likely one of the most common is the induction of adventitious buds by an indirect inductive process involving an intermediate callus phase. During the induction of this process, the auxin–cytokinin crosstalk was frequently crucial. Recently, Zhai and Xu [167] reported in Arabidopsis that the tissue structure after callus induction was comparable to that of the root primordium or apical meristem. Interestingly, authors identified the middle cell layer characterized by a quiescent center-like transcriptional identity, and at this site, tissues exhibited the ability to regenerate organs. In particular, WOX5 directly interacts with PLETHORA1 and 2 to promote the TRYPTOPHAN AMINOTRANSFERASE OF ARABIDOPSIS1 expression for endogenous auxin production. WOX5 also interacts with the B-type ARABIDOPSIS RESPONSE REGULATOR12 (ARR12) and represses A-type ARRs to break the negative feedback loop in cytokinin signaling. Therefore, auxin production and the enhancement of cytokinin sensitivity were required for pluripotency acquisition in the middle cell layer of the callus for organ regeneration [167].
To exemplify the studies in which WOX genes are involved in in vitro morphogenesis, a selection is shown in Table 1.

4. Overexpression of Genes Encoding Transcription Factors to Increase Transformation Rate

Several reports have strongly suggested that the co-expression of morphogenetic genes, as well as BBM and WUS, stimulates the growth of the embryogenic callus, resulting in improved transformation frequencies with Agrobacterium tumefaciens in transformation-recalcitrant monocot species [174,175,176,177,178,179]. Moreover, WUS overexpression has been reported to enhance somatic embryogenesis in dicot and gymnosperm species, such as N. tabacum [180], Coffea canephora [181], Picea glauca [182], and Medicago truncatula [65]. To examine the effect of the ectopic expression of three Arabidopsis WOX2, WOX8, and WOX9 genes on the regenerative competency of tissues and cells cultured in vitro, Kyo et al. [183] developed a transgenic variety of Nicotiana tabacum in which these genes were under the transcriptional control of a chemical-inducible expression system. The authors observed remarkable regeneration of plantlets only in explants derived from the hybrids possessing two transgenes, namely WOX2 combined with WOX8 or WOX9, but found no regeneration in the segments derived from their parental lines, proving that not all genes of the WOX family act in the same way. Capsicum chinense varieties are well known for their unique flavors, and many have exceptional heat. The hottest peppers in the world are members of this species [184]. Unfortunately, C. chinense is a recalcitrant species for in vitro morphogenesis, and the development of new biotechnological tools (NBTs) has been hindered. Nevertheless, an in vitro transformation method was successfully obtained in this species via Agrobacterium tumefaciens co-cultivation with a system that expresses the heterologous gene WUS from Arabidopsis [185]. Similarly, in Gossypium hirsutum, somatic embryogenesis was significantly improved in AtWUS-overexpressing calli compared to control explants [186,187].
More recently, Kadri et al. [188] investigated the effect of the heterologous Arabidopsis WUS gene overexpression, under the control of the jasmonate-responsive vsp1 promoter, on the morphogenetic responses of Medicago truncatula explants. WUS expression in leaf explants increased callogenesis and embryogenesis without growth regulators. Similarly, WUS expression enhanced the embryogenic potential of hairy root fragments.
However, especially in maize, overexpression of BBM and WUS can compromise the quality of regenerated plants, leading to male and/or female sterility. Agrobacterium-mediated T-DNA transformation involves transient T-DNA gene expression within 36–48 h of bacterium infection, followed by stable T-DNA integration into the plant genome [189,190]. This has been demonstrated using CRE-mediated excision of a genomic locus flanked by homologous loxP sites, without stable integration of the T-DNA harboring the CRE recombinase gene [191]. Therefore, to increase the production of fertile T0 plants, Wang et al. [192] used an inducible site-specific recombinase (Cre) to excise morphogenetic genes (i.e., WUS2 and/or BBM) after transformation, but before regeneration. In addition, the use of developmentally regulated promoters, such as Ole, Glb1, End2, and Ltp2, to drive Cre enabled the excision of morphogenetic genes in early embryo development and produced excised events at a rate of 25–100%.
Several other methods have been developed to exclude the effect of the stable integration of BBM and WUS. For example, Hoerster et al. [193] determined that the expression of Zm-WUS2 alone, driven by the maize Pltp promoter (Zm-Pltppro), was sufficient to induce rapid SE formation from the scutella of immature maize zygotic embryos. Notably, the authors demonstrated that co-infecting with two strains of Agrobacterium, one with a WUS2 expression cassette and the other with a combination of selectable and visual marker cassettes, transformed T0 plants that contained only a single copy of the selectable marker T-DNA, without the integration of WUS2. Furthermore, the process was optimized by varying the ratio of the two Agrobacterium strains and by modulating WUS2 expression to enable high-frequency recovery of selectable marker-containing T0 plants, without the WUS2 gene.
The increase in transformation efficiency in cereal crops is a prominent goal in plant breeding through biotechnological approaches, and recently, Wang et al. [194] have obtained interesting results. These authors demonstrated that overexpression of the wheat gene TaWOX5 dramatically increases the transformation frequency of wheat, as well as five other cereal species; the significant result was obtained in different genotypes, and had no apparent deleterious effects after regeneration [194].

5. Molecular Players in Plant Totipotency

So far, the nature of totipotency remains a mystery in some aspects [73,195,196]. As reported above, some data suggest that in higher plants, in addition to WUS/WOX genes, the ectopic upregulation of many other TFs (e.g., LEC1, LEC2, BBM, FUS3, and AGL15) is involved in the induction of embryo formation from somatic cells [47,49,51,52,197]. One common feature of these genes is their role in promoting embryogenic cell competence in the absence of auxin or stress-related treatments, usually used to establish embryogenic programs on in vitro systems [34,73]. However, the variety of the aforementioned regulatory factors also suggests that numerous independent and/or interrelated pathways can lead to the acquisition of embryogenic competence [35,73]. At the same time, WUS overexpression has been associated with somatic embryogenesis [70] and the ectopic proliferation of flower meristems [198]. On the other hand, WUS overexpression appears to repress the LEC1 gene [199], whereas the overexpression of AGL15 upregulates SERK1; furthermore, when LEC2 is ectopically expressed, an increase in AGL15 can be observed. Whether and how these genes work together to promote embryogenic cell formation is not yet wholly known [200]. Recently, some regulatory networks have been proposed [73,196,201], and some results suggest that the ectopic expression of these TFs is not necessarily linked to the switch of cell fate towards the same embryogenic competence [196]. The expression of LEC1, LEC2, WUS, and AGL15 genes in the egg cell or zygote is lacking. Currently, the hypothesis that their overexpression acts indirectly, causing some stress or responses that lead to change cell fate, is still open. In addition, the cellular context in which a gene is expressed appears to be important in determining the ability of a cell to become totipotent [196]. For example, WUS functions non-cell-autonomously to maintain pluripotent stem cell identity in the meristem niche, whereas it appears to function cell autonomously when ectopically expressed in a range of somatic cells that become totipotent embryogenic cells [70,110].
Notably, in apomictic Cenchrus ciliaris, the BBM-Like (BBML) gene is naturally expressed in the egg cell [202], and expression of BBML under its promoter or an egg-cell-expressed promoter can induce parthenogenesis in sexual grasses and tobacco [197,203,204]. In maize and rice, BBML genes are expressed in the zygote shortly after fertilization, and in the sperm cell before fertilization [205,206]. In Arabidopsis, Chen et al. [207] showed that BBM is expressed transiently in the chalazal region of the ovule and seed, in the embryo starting from the zygote stage, and in the first few dividing endosperm cells. These data suggest early BBM functions in the embryo and broader functions during seed development. In addition, ectopic BBM expression in the egg cell of Arabidopsis and in the dicot crops Brassica napus and Solanum lycopersicon is sufficient to bypass the fertilization requirement for embryo development [207]. Therefore, it was demonstrated that ectopic BBM expression in the egg cell is sufficient to induce haploid embryo development. These data suggest that BBM TF is at the apex of a transcriptional network that promotes asexual embryo development [207]. The BBML gene in the egg cell is suitable for inducing parthenogenesis in rice [208]. However, to obtain diploid embryos, the misexpression of BBML is insufficient. When genome editing to substitute mitosis for meiosis (MiMe) [209,210] is combined with the expression of rice BBM1 in the egg cell, then a clonal progeny can be obtained that retains genome-wide parental heterozygosity [204]. Moreover, Xie et al. [211], via CRISPR/Cas9, deactivated the genes OsPO11-1, OsREC8, OsOSD1, and OsMATL to create apomictic rice plants. The resulting quadruple mutant, Apomictic Offspring Production (AOP), showed a transformation from meiosis to mitosis and produced clonal diploid gametes. However, consistently with previously reports, mutation of OsMATL gave rise to low fertility and low haploid or apomictic induction rates [212,213].
It should be clear that from the examples reported, members of the WOX TF family play a pivotal role in the acquisition of embryogenic competence, and they control various stages of SE development. However, these data are insufficient to explain the achievement of embryogenic totipotency. It cannot be the activation of a single gene that reprograms the fate of a differentiated cell towards acquiring all the degrees of freedom representative of the zygotic cell. During differentiation, a less specialized cell type gradually transforms into a more specialized one, which is constrained to a stable morphology, structure, and function [214]. Moreover, the differentiation process reduces the cells’ self-renewal ability and embryogenic totipotency [35,73,196].

6. An Epigenetic Hierarchical Network Reprograms Differentiated Cells to Pluripotent/Totipotent Cells

In the classical view of development, the differentiated state of a cell was believed to be terminal and irreversible. This concept has been best exemplified by Conrad H. Waddington’s description of the epigenetic landscape [215]. In many cases, the complex PGR and gene regulation, which leads a zygote to differentiate the various cell types of an organism, is only partially understood. Nevertheless, it is much more challenging to understand the reverse path. The differentiation is a process of differential gene activation and de-activation that cannot involve either gene mutations or stress conditions. Thus, it would not be easy to reconcile such a phenomenon which directly affects de-differentiation towards the reacquisition of embryogenic totipotency. Various stressful conditions are known to be inductive to somatic embryogenesis [34,73]. Perhaps we should ask ourselves what differentiates zygotes from all other plant cells. Why is a somatic cell subjected to severe stress conditions, devastating PGR conditions, or the overexpression of many TFs capable of inducing differentiated cells to reacquire potentialities typical of the zygote? The zygote undergoes extensive self-renewal and can differentiate along multiple cell lineages. Differentiation requires long-lasting changes in gene expression, with differential gene inactivation or activation during the process. In plants and animals, epigenetic mechanisms are essential to controlling the heritable cellular memory of gene expression during development [214,216,217,218,219]. Therefore, it is plausible to hypothesize that activating analogous mechanisms is necessary for a differentiated cell to regain embryogenic totipotency [73,196]. Problems arise here. The degree of differentiation of cells grown in vitro can be very significant. We do not know, even superficially, the epigenetic landscapes traveled by the cells in their differentiation [220,221]. Therefore, the stimuli that we supply to cells, or, more commonly, to organs and tissues, to reacquire embryogenic totipotency can only be empirical (e.g., large amounts of auxin, stress conditions, and overexpression of TFs). We should understand how these treatments intersect with epigenetic changes by redirecting differentiated somatic cells to lose their determination state. All of this is to achieve a condition that allows for the development of SE, in some cases, in the absence of exogenous PGR treatments [24,25,26,27,28,70].
It has been highlighted that no spatial regulation is active in the control of the embryogenic cell pool compared to the pool of stem cells in the meristems, where signals from neighboring cells assure stemness [222,223]. According to this hypothesis, embryogenic cells exhibit relevant structural differences concerning meristematic stem cells, dealing with an increased cell wall thickening, a reduced number of plasmodesmata, and a specific nuclear chromatin architecture [222]. Thus, it is likely that the physical or physiological isolation of somatic cells from their immediate surroundings is coupled with the reprogramming of gene expression profiles, leading to a cell fate switch [73,224].
In this context, the de-differentiation event results from the precise choreography of genes whose transcription is temporally and spatially controlled. Several studies have demonstrated the essential role of chromatin structure in controlling gene transcription [214,225]. As a consequence, chromatin remodeling, which can be accomplished by multiple mechanisms such as histone post-transcriptional modifications and changes in DNA methylation [226], is necessary to couple with cellular de-differentiation and switching of cell fate [36,39,73,196,227]. Thus, it is likely that epigenetic mechanisms control cell fate through superimposed activity on the TFs and PGRs, and may have a role in regulating cell totipotency in plants [71]. In carrot plants, the LEC1 promoter region showed a reduced level of DNA methylation during somatic embryogenesis, followed by an increase during the transition from embryonic to vegetative growth [228]. In addition, if hypermethylation of a region of 5′-LEC1 promoter was induced by RNA-directed DNA methylation (RdDM), the gene expression was reduced in the embryogenic cells, indicating a negative correlation between DNA methylation and LEC1 expression [228].
Characterizing the PICKLE (PKL) gene has strengthened this hypothesis [229,230,231]. The mutant pkl of Arabidopsis shows swollen primary roots. Its phenotype is characterized by the postembryonic expression of embryo-specific markers (LEC1, LEC2, and FUS3) in the primary root, and by the ectopic proliferation of SEs. PKL encodes for an SWI/SNF member, a putative Chromodomain Helicase DNA-binding protein 3 (CHD3). In roots, PKL-mediated downregulation of LEC genes depends on the activation of Polycomb group (PcG) genes [232,233]. In eukaryotes, the general mechanisms of PcG protein functions are well-preserved, although, in plants, different players appeared during evolution [234]. PcG and Trithorax (TrxG) complexes control the expression of developmental regulator genes through modifications of the chromatin status, e.g., the deposition of the repressive H3K27me3 and the activator H3K4me3, respectively [233,235,236,237,238,239]. PcG proteins exist in multiprotein complexes; polycomb repressive complex 2 (PRC2) is a crucial regulator of epigenetic states which, via its catalytic subunit E(z), shuts off gene expression by trimethylating Lys27 of histone H3, resulting in H3K27me3 [240]. In contrast, PRC1, via its chromodomain-containing subunit polycomb (Pc), binds H3K27me3 generated by PRC2, resulting in a stable silencing chromatin state [241]. In the pkl mutant, the reduced PRC1 protein activity can lead to cell de-differentiation and callus-like tissue formation. Mutations in Arabidopsis PRC1-like ring-finger protein genes AtRING1a and AtRING1b lead to the de-repression of embryonic traits during vegetative growth, and induce the ectopic expression of several key regulatory genes involved in embryogenesis and stem cell activity [242]. PRC2 composition is conserved from humans to plants, but the function of PRC2 during the early stages of plant life has not yet been fully defined. PRC2 is involved in repressing embryo maturation programs during the establishment of vegetative development in Arabidopsis [243,244,245]. The absence of PRC2 can induce somatic cells to become totipotent, thereby leading to somatic embryogenesis. Thus, in contrast to mammals, where PRC2 proteins are required to maintain pluripotency and prevent cell differentiation, in plants, PRC2 proteins are required to promote cell differentiation by suppressing embryonic development [246,247]. Mutations in genes coding for PRC2 proteins resulted in mutants for fertilization-independent endosperm (fie) or seed (fis) formation, respectively, suggesting that the embryogenic program is repressed by chromatin-based gene silencing, and becomes released in response to fertilization [248]. Furthermore, as in animals, the plant zygote is transcriptionally relatively quiescent, and maternally stored products can sustain the first cell divisions; by contrast, downregulation of transcription is deleterious to endosperm development. The transient zygotic quiescence, resulting from endosperm-embryo interactions, is probably required for extensive genome reprogramming after fertilization [249].
WUS regulation can also be subjected to epigenetic mechanisms [250,251,252]. Based on the observation that the gene AtGCN5, coding a histone acetyltransferase, is required to control floral meristem through the WUS/AG pathway, it has been supposed that the histone acetylation possibly restricts the WUS expression domain within the floral meristem, turning on a WUS repressor [253]. Acetylation, a post-translational modification affecting H3 and H4 histones, is performed by highly specific enzymes identified in plants and animals [254,255]. The histone modification system interplays and crosstalks with DNA methylation status, affecting the degree and location of histone post-translational modifications [256,257]. In the promoter region, the methylation of DNA directly lays down the silencing of transcription [258]. In animals, DNA methylation concerns cytosines placed in the dinucleotide CpG, but in plants, the CpHpG and CpHpHpG patterns can also be involved. DNA methylation, catalyzed by DNA methyltransferases (DNMTs), can be reversed by the Ten-Eleven Translocation (TET) proteins in mammalian [259] and the Demeter (Dme) family of DNA glycosylases in plants [260]. In plants, DNA methylation can be directed by small RNAs (RdDM) using two plant-specific RNA polymerases-PolIV (NRPD1) and PolV (NRPE1), and by de novo DNA methylase DRM1/2 “Domain Rearranged Methyltransferase 1/2” [261]. DNA methylations in both CG and CHG sites are achieved by METHYLTRANSFERASE1 (MET1) and CHROMOMETHYLASE3 (CMT3), respectively [262,263,264]. Maintaining methylation in non-CG sites of heterochromatic TEs also requires the activity of CMT2 [265]. Lysine 9 methylation of histone H3 (H3K9me2) is mediated by the SET domain of SU(VAR)3–9 HOMOLOGUE 4/KRYPTONITE (SUVH4/KYP), SUVH5, and SUVH6 proteins, which promotes the binding of CMT3 to chromatin and maintains methylation in CHG sites [263,266]. Furthermore, the activity of DECREASE IN DNA METHYLATION 1 (DDM1), a chromatin-remodeling factor, is required both to modify the chromatin conformation and to maintain the DNA methylation patterns across diverse plant species. This is required for RdDM to maintain mCHH islands [265,267,268].
In sunflower, the involvement of histone H3 methylation in HaWUS regulation is also suggested by the presence of several putative binding motives; e.g., PROMO software finds DNA-binding sequences for the ALFIN1-like protein implied in the switch from the H3K4me3-associated active to the H3K27me3-associated repressive transcription state of seed developmental genes, promoting seed germination [84,269]. In A. thaliana, a histone acetylase, which is coded by AtGCN5 and is involved in both long-term and short-term dynamic transcriptional regulations [233,270], limits the domain of WUS expression within the floral meristem, turning on a WUS repressor through the WUS/AG pathway [253]. Moreover, in the termination of floral stem cell maintenance in A. thaliana, AG represses WUS expression and recruits PcG proteins to deposit H3K27me3 [91] in a still-unknown way.
Based on chromatin accessibility dynamics, Wang et al. [196] showed that the developmental stage of in vitro cultured tissue (i.e., non-germinated seeds) is at the top of the regulatory hierarchy that governs SE initiation. This finding explains why post-embryonic somatic tissues (i.e., germinated seedlings) are resistant to reprogramming for somatic embryogenesis. Wang et al. [196] suggested that the cellular status of the juvenile phase is less amenable to reshaping the chromatin status of the gene loci determining totipotency. It was also hypothesized that histone 2A monoubiquitination at lysine 119 (H2Aub) marking of ABI3/FUS3/LEC2 leads to their initial repression, further maintained by PcG-mediated H3K27me3, and might contribute to the loss of reprogramming competence in the somatic cells after seed germination [196]. In this model, LEC1, BBM, and auxin form a feed-forward loop to reinforce cell fate transition. LEC2 acts at the cell totipotent gene network’s output node by activating early embryonic development genes, such as WOX2 and WOX3 [196].
In a vision that involves epigenetic mechanisms for controlling cellular totipotency, a key role is also played by micro RNAs (miRNAs or miRs). MiRNAs are the most characterized class of non-coding RNAs, and are engaged in many cellular processes, including cell differentiation, development, and homeostasis [271]. MiRNAs are 21–24-nt-long single-stranded nucleic acids that function post-transcriptionally to regulate gene expression [271,272].
MiRNAs have been implicated in maintaining the pluripotency of mammalian cells [273]. In plants, several types of miRNAs involved in somatic embryogenesis have been reported. For example, miR156, miR162, miR166a, miR167, miR168, miR171a/b, miR171c, miR393, miR397, and miR398 play a very active role during various stages of somatic embryogenesis [274,275,276,277]. Genome-wide analysis of the somatic embryo transcriptome in Arabidopsis indicated that numerous miRNAs are differentially expressed during somatic embryogenesis by the extensive modulation of transcription factor gene expression [66,272,278]. In particular, there are some indications of their role in maintaining pluripotency [73,279]. Among them, miR160 interacts with ARF10 to inhibit callus initiation and shoot regeneration. At the same time, miR156, which targets SQUAMOSA promoter-binding protein-like (SPL) TFs, contributes to a decline in the explants’ capacity to regenerate shoots [280,281,282]. MiR160, targeting ARF10, ARF16, and ARF17, is required to develop many organs, particularly in embryos [283]. Therefore, MIR160a loss-of-function mutants exhibit various defects during embryogenesis [284]. By contrast, transgenic expression of a miR160-resistant form of ARF10 was associated with a high level of shoot regeneration. This transgenic line also showed an elevated expression level of shoot meristem-specific genes CL3, CUC1 and -2, and WUS [280].
In Arabidopsis, miR165/166 and miR160 contribute to the induction of somatic embryogenesis associated with the LEC2-controlled auxin response that occurs in somatic cells during the embryogenic transition [275]. MiR167 functions in somatic embryogenesis by regulating the expression of its target auxin response genes [285]. Thus, specific miRNAs target the genes that govern the transition from differentiated to totipotent cells. The differential expression of miRNAs during somatic embryogenesis has also been observed in other plant species, such as Oryza sativa [286] and Zea mays [287].
ARGONAUTE (AGO) proteins bind with mature miRNAs to guide the riboprotein complex to its target mRNA [288,289]. In vitro, AGO10 inhibited shoot regeneration via repressing de novo SAM formation [286]. In fact, in in vitro cultured explants of the loss-of-function mutant ago10, a much larger number of SAMs was formed, and within these, the stem cell marker genes WUS, CLV3, and STM were all strongly expressed [290]. AGO10 repressed the accumulation of the miR165/166, thereby upregulating HD-ZIP III genes. The overproduction of miR166 was shown to promote shoot regeneration. At the same time, the absence of miR165/166 information resulted in a blockage to shoot regeneration, and only a partial rescue of the phenotype of the ago10 mutant [290]. Notably, explants derived from the men1 mutant (an overproducer of miR166a) regenerated shoots more readily than WT explants, while explants derived from the loss-of-function of the miR165/166 line MIM165/166 were less productive [290]; these observations suggest that miR165/166 acted to promote adventitious shoots.
Thus, epigenetic mechanisms appear to be essential upstream factors in determining embryogenic capacity; during the induction of somatic embryogenesis, the remodeling of chromatin results in the release of the embryogenic program otherwise repressed by chromatin-based silencing mechanisms. Consequently, variations in the expression profile of genes directly involved in the epigenetic regulation of cell transcriptome could have a primary role in initiating an embryogenic program.

7. Conclusions and Prospects

In conclusion, as underlined by Jha et al. [88], WOX genes are master regulators of different aspects of plant biology and in Figure 4 a simplified overview is reported. We focused on the involvement of these genes in the in vitro regeneration of plants and, more generally, in the manifestation of cellular totipotency.
In particular, somatic embryogenesis is a fascinating phenomenon in plant biology, and is useful for vegetative propagation and plant regeneration in genome editing methods.
From the pioneering studies conducted with in vitro experiments, information about the nature of the factors underlying the induction processes has dramatically increased. The pathway to developing SE from somatic cells is not always the same, and, as suggested by Fehér [35], the claim to identify a key trigger valid for all somatic embryogenic systems will be challenging to achieve. In several plants, differential expression of specific genes is observed in vitro during the induction of SE; however, to date, the hierarchical relationship between the genes found to be differentially expressed is only partially decoded, depending on the analyzed species. In Arabidopsis, an updated hierarchical mechanism for SE in vitro has recently been suggested [45]. Regarding this, at the top is a permissive chromatin environment to allow the reactivation of genes encoding cell totipotency-related TFs with a concurrent auxin stimulation. Hence, at the successive step, direct activation of early embryonic development genes such as WOX2 and WOX3 occurs [45]. Concerning the modification of the epigenetic landscape, one of the more critical genes is PRC2 [245,291]. At the same time, regarding the totipotency-related TFs, a position of absolute prominence should undoubtedly be assigned to BBM [43].
What about WUS in the induction of SE? The somatic embryogenesis is unaffected in the wus mutant of Arabidopsis, although malformations in the shoot meristem were recognized [80]. Furthermore, numerous studies have shown that WUS expression is often induced during the induction of SEs, and overexpression is associated with a more significant SE differentiation in some species [88,292]. In addition, the concomitant overexpression of BBM and WUS in monocots promotes SE induction, with a consequent increase in Agrobacterium-mediated transformation percentage [174]. However, the involvement of the WUS gene in regeneration processes in vitro cannot be confined to the embryogenic pathway, since its involvement in the differentiation of adventitious buds was repeatedly observed. Indeed, during the induction of shoots in vitro from hypocotyl explants in Arabidopsis, transcripts of the WUS gene mark the shoot progenitor region of cells at an extremely early stage [62]. It is interesting to point out that for the ectopic WUS expression in a restricted group of cells, a sequence of factors is required: exogenous cytokinin supply to remove a repressive condition of epigenetic nature, and then the action of ARRs genes and binding with microRNA165/6-targeted HD-ZIP III TFs [62]. Cytokinin signaling and a permissive epigenetic environment are also required in de novo activation of WUS expression during axillary meristem initiation in vivo [293]. To date, no such detailed chain of molecular events has described the role of the WUS gene in the induction of somatic embryogenesis. Zuo et al. [70] have obtained strong evidence regarding the importance of WUS in the vegetative-to-embryonic transition in Arabidopsis. However, the authors have not characterized the WUS-expressing cells at the morphological and functional levels, and it is still unclear whether that group of cells represents a functional organizing center [70]. On the other hand, WUS represses LEC1 expression, suggesting that WUS cannot activate the embryo identity pathway [70].
Furthermore, it should be noted that in vivo, WUS and WOX genes are also implicated in vegetative propagation, such as epiphyllous plantlets and the differentiation of bulbils [135,136]. In particular, somatic embryogenesis in vivo is a remarkable phenomenon, but, unfortunately, it has not been studied in detail. It might be interesting to learn, in more depth, the role of WOX genes in epiphylly, since only a few genes have been investigated at present [26,294,295,296,297]. In this regard, the recent study by Jácome-Blásquez et al. [298] appeared to increase knowledge regarding the genes involved in the ectopic differentiation of plantlets in vivo. In addition, it might also be of interest to address the molecular characterization of somatic embryogenesis in other plant species with a spontaneous ability to develop in vivo ectopic embryos, such as Malaxis paludosa [22]. We are still far from outlining a model that links, in a hierarchical relationship, the crucial factors for natural SE.

Author Contributions

Conceptualization, M.F., G.U. and C.P.; investigation, C.P. and M.F.; writing—original draft preparation, C.P. and M.F.; writing—review and editing, M.F., G.U. and C.P.; visualization, M.F., G.U. and C.P.; supervision, M.F., G.U. and C.P.; funding acquisition, C.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AGL15MADS-box TF AGAMOUS-LIKE15
AILAINTEGUMENTA-LIKE
CLV1CLAVATA1
CUC1CUP-SHAPED COTYLEDON1
EARERF-associated Amphiphilic Repression
FUS3FUSCA3
GRDGROUNDED
HBWOX homeodomain
HLHTHhelix-loop-helix-turn-helix
Kn-1Knotted-1
LEC1NF-Y (nuclear factor of the Y box) TF LEAFY COTYLEDON1
NLSNuclear Localization Signals
PLT4/BBMAP2/ERF-domain TF PLETHORA4/BABY BOOM
PLT5/EMKPLT5/EMBRYOMAKER
PGRPlant growth regulator
RAMRoot Apical Meristem
RKD4RWP-RK DOMAIN-CONTAINING4
SAMShoot Apical Meristem
SESomatic embryo
SERKSomatic Embryogenesis Receptor Kinase
STM1class I KNOX homeodomain SHOOT MERISTEMLESS1
TFtranscription factor
WOXWUSCHEL-related homeobox genes
WUSWUSCHEL

References

  1. Basile, A.; Fambrini, M.; Pugliesi, C. The vascular plants: Open system of growth. Dev. Genes Evol. 2017, 227, 129–157. [Google Scholar] [CrossRef] [PubMed]
  2. Birnbaum, K.D.; Sanchéz-Alvarado, A.S. Slicing across kingdoms: Regeneration in plants and animals. Cell 2008, 132, 697–710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Technau, U.; von Laue, C.C.; Rentzsch, F.; Luft, S.; Hobmayer, B.; Bode, H.R.; Holstein, T.W. Parameters of self-organization in Hydra aggregates. Proc. Natl. Acad. Sci. USA 2000, 97, 12127–12131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Echeverri, K.; Tanaka, E.M. Ectoderm to mesoderm lineage switching during axolotl tail regeneration. Science 2002, 298, 1993–1996. [Google Scholar] [CrossRef] [PubMed]
  5. Poss, K.D.; Wilson, L.G.; Keating, M.T. Heart regeneration in zebrafish. Science 2002, 298, 2188–2190. [Google Scholar] [CrossRef]
  6. Beffagna, G. Zebrafish as a smart model to understand regeneration after heart injury: How fish could help humans. Front. Cardiovasc. Med. 2019, 6, 107. [Google Scholar] [CrossRef] [Green Version]
  7. Taub, R. Liver regeneration: From myth to mechanism. Nat. Rev. Mol. Cell Biol. 2004, 5, 836–847. [Google Scholar] [CrossRef]
  8. Rougier, N.; Werb, Z. Minireview: Parthenogenesis in mammalian. Mol. Reprod. Dev. 2001, 59, 468–474. [Google Scholar] [CrossRef]
  9. Lampert, K.P. Facultative parthenogenesis in vertebrates: Reproductive error or chance? Sex. Dev. 2008, 2, 290–301. [Google Scholar] [CrossRef]
  10. Ramachandran, R.; McDaniel, C.D. Parthenogenesis in birds: A review. Reproduction 2018, 155, R245–R257. [Google Scholar] [CrossRef]
  11. Fujita, M.K.; Singhal, S.; Brunes, T.O.; Maldonado, J.A. Evolutionary dynamics and consequences in parthenogenesis in vertebrates. Annu. Rev. Ecol. Evol. Syst. 2020, 51, 191–214. [Google Scholar] [CrossRef]
  12. Heimpel, G.E.; de Boer, J.G. Sex determination in the hymenoptera. Ann. Rev. Entomol. 2008, 53, 209–230. [Google Scholar] [CrossRef] [PubMed]
  13. Tvedte, E.S.; Logsdon, J.M., Jr.; Forbes, A.A. Sex loss in insects: Causes of asexuality and consequences for genomes. Curr. Opin. Insect Sci. 2019, 31, 77–83. [Google Scholar] [CrossRef] [PubMed]
  14. Picard, M.A.L.; Viscoso, B.; Bertrand, S.; Escriva, H. Diversity of modes of reproduction and sex determination systems in invertebrates, and the putative contribution of genetic conflict. Genes 2021, 12, 1136. [Google Scholar] [CrossRef]
  15. Condic, M.L. Totipotency: What it is and what it is not. Stem Cells Dev. 2014, 23, 796–812. [Google Scholar] [CrossRef] [Green Version]
  16. Asker, S.E.; Jerling, L. Apomixis in Plants, 1st ed.; CRC Press: London, UK, 1992. [Google Scholar]
  17. Bicknell, R.A.; Koltunow, A.M. Understanding apomixis: Recent advances and remaining conundrums. Plant Cell 2004, 16 (Suppl. S1), S228–S245. [Google Scholar] [CrossRef] [Green Version]
  18. Conner, J.A.; Goel, S.; Gunawan, G.; Cordonnier-Pratt, M.-C.; Johnson, V.E.; Liang, C.; Wang, H.; Pratt, L.H.; Mullet, J.E.; DeBarry, J.; et al. Sequence analysis of bacterial artificial chromosome clones from the apospory-specific genomic region of Pennisetum and Cenchrus. Plant Physiol. 2008, 147, 1396–1411. [Google Scholar] [CrossRef] [Green Version]
  19. Hand, M.L.; de Vries, S.; Koltunow, A.M.G. A comparison of in vitro and in vivo asexual embryogenesis. Methods Mol. Biol. 2016, 1359, 3–23. [Google Scholar]
  20. Barcaccia, G.; Palumbo, F.; Sgorbati, S.; Albertini, E.; Pupilli, F. A Reappraisal of the evolutionary and developmental pathway of apomixis and its genetic control in angiosperms. Genes 2020, 11, 859. [Google Scholar] [CrossRef]
  21. Hojsgaard, D. Molecular basis of apomixis in plants. Genes 2021, 12, 576. [Google Scholar] [CrossRef]
  22. Taylor, R.L. The foliar embryos of Malaxis paludosa. Can. J. Bot. 1967, 45, 1553–1556. [Google Scholar] [CrossRef]
  23. Dickinson, T.A. Epiphylly in angiosperms. Bot. Rev. 1978, 44, 181–232. [Google Scholar] [CrossRef]
  24. Fambrini, M.; Cionini, G.; Bianchi, R.; Pugliesi, C. Epiphylly in a variant of Helianthus annuus × H. tuberosus induced by in vitro tissue culture. Int. J. Plant Sci. 2000, 161, 13–22. [Google Scholar] [CrossRef] [PubMed]
  25. Fambrini, M.; Durante, C.; Cionini, G.; Geri, C.; Giorgetti, L.; Michelotti, V.; Salvini, M.; Pugliesi, C. Characterization of LEAFY COTYLEDON1-LIKE gene in Helianthus annuus and its relationship with zygotic and somatic embryogenesis. Dev. Genes Evol. 2006, 216, 253–264. [Google Scholar] [CrossRef] [PubMed]
  26. Garcês, H.M.P.; Champagne, C.E.M.; Townsley, B.T.; Park, S.; Malhó, R.; Pedroso, M.C.; Harada, J.J.; Sinha, N.R. Evolution of asexual reproduction in leaves of the genus Kalanchoë. Proc. Natl. Acad. Sci. USA 2007, 104, 15578–15583. [Google Scholar] [CrossRef] [Green Version]
  27. Chiappetta, A.; Michelotti, V.; Fambrini, M.; Bruno, L.; Salvini, M.; Petrarulo, M.; Azmi, A.; Van Onckelen, H.; Pugliesi, C.; Bitonti, M.B. Zeatin accumulation and misexpression of a class I knox gene are intimately linked in the epiphyllous response of the interspecific hybrid EMB-2 (Helianthus annuus × H. tuberosus). Planta 2006, 223, 917–931. [Google Scholar] [CrossRef]
  28. Chiappetta, A.; Fambrini, M.; Petrarulo, M.; Rapparini, F.; Michelotti, V.; Bruno, L.; Greco, M.; Baraldi, M.; Salvini, M.; Pugliesi, C.; et al. Ectopic expression of LEAFY COTYLEDON-LIKE gene and localized auxin accumulation mark embryogenic competence in epiphyllous plants of Helianthus annuus × H. tuberosus. Ann. Bot. 2009, 103, 735–747. [Google Scholar] [CrossRef] [Green Version]
  29. Haberlandt, G. Culturversuche mit isolierten pflanzenzellen. In Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften; Mathematisch-Naturwissenschaftliche Classe; Springer: Wien, Austria, 1902; Volume 111, pp. 69–92. [Google Scholar]
  30. Steward, F.C.; Mapes, M.O.; Mears, K. Growth and organized development of cultured cells. II. Organization in cultures grown from freely suspended cells. Am. J. Bot. 1958, 45, 705–708. [Google Scholar] [CrossRef]
  31. Reinert, J. Über die kontrolle der morphogenese und die induktion von adventive embryonen an gewebekulturen aus karotten. Planta 1959, 53, 318–333. [Google Scholar] [CrossRef]
  32. Fehér, A.; Pasternak, T.; Dudits, D. Transition of somatic plant cells to an embryogenic state. Plant Cell Tissue Organ Cult. 2003, 74, 201–228. [Google Scholar] [CrossRef]
  33. Fehér, A. Why somatic plant cells start to form embryos? In Somatic Embryogenesis, Plant Cell Monographs; Mujib, A., Samaj, J., Eds.; Springer-Verlag: Berlin/Heidelberg, Germany, 2005; Volume 2, pp. 85–101. [Google Scholar]
  34. Fehér, A. Somatic embryogenesis-stress-induced remodeling of plant cell fate. Biochim. Biophys. Acta Gene Regul. Mech. 2015, 1849, 385–402. [Google Scholar] [CrossRef] [PubMed]
  35. Fehér, A. Callus, dedifferentiation, totipotency, somatic embryogenesis: What these terms mean in the era of molecular plant biology? Front. Plant Sci. 2019, 10, 536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Grafi, G. How cells dedifferentiate: A lesson from plants. Dev. Biol. 2004, 268, 1–6. [Google Scholar] [CrossRef] [PubMed]
  37. Elhiti, M.; Huang, S.; Mira, M.M.; Hill, R.D.; Stasolla, C. Redirecting cell fate during in vitro embryogenesis: Phytoglobins as molecular switches. Front. Plant Sci. 2018, 9, 1477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Efroni, I.A. Conceptual framework for cell identity transition in plants. Plant Cell Physiol. 2018, 59, 696–706. [Google Scholar] [CrossRef] [Green Version]
  39. Su, Y.H.; Tang, L.P.; Yu, X.; Zang, X.S. Plant cell totipotency: Insights into cellular reprogramming. J. Integr. Plant Biol. 2020, 63, 228–243. [Google Scholar] [CrossRef]
  40. Eshed-Williams, L. Genetics of shoot meristem and shoot regeneration. Annu. Rev. Genet. 2021, 55, 661–681. [Google Scholar] [CrossRef]
  41. Long, Y.; Yang, Y.; Shen, Y. New insights into tissue culture plant-regeneration mechanisms. Front. Plant Sci. 2022, 13, 926752. [Google Scholar] [CrossRef]
  42. Drapek, C.; Sparks, E.E.; Benfey, P.N. Uncovering gene regulatory networks controlling plant cell differentiation. Trends Genet. 2017, 33, 529–539. [Google Scholar] [CrossRef]
  43. Horstman, A.; Bemer, M.; Boutilier, K.A. Transcriptional view on somatic embryogenesis. Regeneration 2017, 4, 201–216. [Google Scholar] [CrossRef]
  44. Salaün, C.; Lepiniec, L.; Dubreucq, B. Genetic and molecular control of somatic embryogenesis. Plants 2021, 10, 1467. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, F.X.; Shang, G.D.; Wang, J.W. Towards a hierarchical gene regulatory network underlying somatic embryogenesis. Trends Plant Sci. 2022, 27, 1206–1217. [Google Scholar] [CrossRef] [PubMed]
  46. Sinha, N.R.; Williams, R.E.; Hake, S. Overexpression of the maize homeo box gene, KNOTTED-1, causes a switch from determinate to indeterminate cell fates. Genes Dev. 1993, 7, 787–795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Lotan, T.; Ohto, M.; Yee, K.M.; West, M.A.; Lo, R.; Kwong, R.W.; Yamagishi, K.; Fischer, R.L.; Goldberg, R.B.; Harada, J.J. Arabidopsis LEAFY COTYLEDON1 is sufficient to induce embryo development in vegetative cells. Cell 1998, 93, 1195–1205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Hecht, V.; Vielle-Calzada, J.P.; Hartog, M.V.; Schmidt, E.D.; Boutilier, K.; Grossniklaus, U.; de Vries, S.C. The Arabidopsis SOMATIC EMBRYOGENESIS RECEPTOR KINASE 1 gene is expressed in developing ovules and embryos and enhances embryogenic competence in culture. Plant Physiol. 2001, 127, 803–816. [Google Scholar] [CrossRef]
  49. Stone, S.L.; Kwong, L.W.; Yee, K.M.; Pelletier, J.; Lepiniec, L.; Fischer, R.L.; Goldberg, R.B.; Harada, J.J. LEAFY COTYLEDON2 encodes a B3 domain transcription factor that induces embryo development. Proc. Natl. Acad. Sci. USA 2001, 98, 11806–11811. [Google Scholar] [CrossRef] [Green Version]
  50. Stone, S.L.; Braybrook, S.A.; Paula, S.L.; Kwong, L.W.; Meuser, J.; Pelletier, J.; Hsieh, T.-F.; Fischer, R.L.; Goldberg, R.B.; Harada, J.J. Arabidopsis LEAFY COTYLEDON2 induces maturation traits and auxin activity: Implications for somatic embryogenesis. Proc. Natl. Acad. Sci. USA 2008, 105, 3151–3156. [Google Scholar] [CrossRef] [Green Version]
  51. Boutilier, K.; Offringa, R.; Sharma, V.K.; Kieft, H.; Ouellet, T.; Zhang, L.; Hattori, J.; Liu, C.M.; van Lammeren, A.A.M.; Miki, B.L.A.; et al. Ectopic expression of BABY BOOM triggers a conversion from vegetative to embryonic growth. Plant Cell 2002, 14, 1737–1749. [Google Scholar] [CrossRef] [Green Version]
  52. Harding, E.W.; Tang, W.N.; Nichols, K.W.; Fernandez, D.E.; Perry, S.E. Expression and maintenance of embryogenic potential is enhanced through constitutive expression of AGAMOUS-LIKE 15. Plant Physiol. 2003, 133, 653–663. [Google Scholar] [CrossRef] [Green Version]
  53. Kwong, R.W.; Bui, A.Q.; Lee, H.; Kwong, L.W.; Fischer, R.L.; Goldberg, R.B.; Harada, J.J. LEAFY COTYLEDON1-LIKE defines a class of regulators essential for embryo development. Plant Cell 2003, 15, 5–18. [Google Scholar] [CrossRef] [Green Version]
  54. Gallois, J.-L.; Nora, F.R.; Mizukami, Y.; Sablowski, R. WUSCHEL induces shoot stem cell activity and developmental plasticity in the root meristem. Genes Dev. 2004, 18, 375–380. [Google Scholar] [CrossRef] [PubMed]
  55. Braybrook, S.A.; Harada, J.J. LECs go crazy in embryo development. Trends Plant Sci. 2008, 13, 624–630. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, X.; Niu, Q.-W.; Teng, C.; Li, C.; Mu, J.; Chua, N.-H.; Zuo, J. Overexpression of PGA37/MYB118 and MYB115 promotes vegetative-to-embryonic transition in Arabidopsis. Cell Res. 2009, 19, 224–235. [Google Scholar] [CrossRef] [Green Version]
  57. Tsuwamoto, R.; Yokoi, S.; Takahata, Y. Arabidopsis EMBRYOMAKER encoding an AP2 domain transcription factor plays a key role in developmental change from vegetative to embryonic phase. Plant Mol. Biol. 2010, 73, 481–492. [Google Scholar] [CrossRef] [PubMed]
  58. Gambino, G.; Minuto, M.; Boccacci, P.; Perrone, I.; Vallania, R.; Gribaudo, I. Characterization of expression dynamics of WOX homeodomain transcription factors during somatic embryogenesis in Vitis vinifera. J. Exp. Bot. 2011, 62, 1089–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Waki, T.; Hiki, T.; Watanabe, R.; Hashimoto, T.; Nakajima, K. The Arabidopsis RWP-RK protein RKD4 triggers gene expression and pattern formation in early embryogenesis. Curr. Biol. 2011, 9, 1277–1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Elhiti, M.A.; Stasolla, C.; Wang, A. Molecular regulation of plant somatic embryogenesis. Vitr. Cell. Dev. Biol. Plant 2013, 49, 631–642. [Google Scholar] [CrossRef]
  61. Kareem, A.; Durgaprasad, K.; Sugimoto, K.; Du, Y.; Pulianmackal, A.J.; Trivedi, Z.B.; Abhayadev, P.V.; Pinon, V.; Meyerowitz, E.M.; Scheres, B.; et al. PLETHORA genes control regeneration by a two-step mechanism. Curr. Biol. 2015, 25, 1017–1030. [Google Scholar] [CrossRef] [Green Version]
  62. Zhang, T.Q.; Lian, H.; Zhou, C.M.; Xu, L.; Jiao, Y.; Wang, J.W. A two-step model for de novo activation of WUSCHEL during plant shoot regeneration. Plant Cell 2017, 29, 1073–1087. [Google Scholar] [CrossRef] [Green Version]
  63. Ikeuchi, M.; Shibata, M.; Rymen, B.; Iwase, A.; Bågman, A.M.; Watt, L.; Coleman, D.; Favero, D.S.; Takahashi, T.; Ahmert, S.E.; et al. A gene regulatory network for cellular reprogramming in plant regeneration. Plant Cell Physiol. 2018, 59, 770–782. [Google Scholar] [CrossRef]
  64. Durgaprasad, K.; Roy, M.V.; Venugopal, A.; Rareem, A.; Raj, K.; Willemsen, V.; Mähönen, A.P.; Scheres, B.; Prasad, K. Gradient expression of transcription factor imposes a boundary on organ regeneration potential in plants. Cell Rep. 2019, 29, 453–463. [Google Scholar] [CrossRef] [PubMed]
  65. Tvorogova, V.E.; Fedorova, Y.A.; Potsenkovskaya, E.A.; Kudriashov, A.A.; Efremova, E.P.; Kvitkovskaya, V.A.; Wolabu, T.W.; Zhang, F.; Tadege, M.; Lutova, L.A. The WUSCHEL-related homeobox transcription factor MtWOX9-1 stimulates somatic embryogenesis in Medicago truncatula. Plant Cell Tissue Organ Cult. 2019, 138, 517–527. [Google Scholar] [CrossRef]
  66. Gliwicka, M.; Nowak, K.; Balazadeh, S.; Mueller-Roeber, B.; Gaj, M.D. Extensive modulation of the transcription factor transcriptome during somatic embryogenesis in Arabidopsis thaliana. PLoS ONE 2013, 8, e69261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Wickramasuriya, A.M.; Dunwell, J.M. Global scale transcriptome analysis of Arabidopsis embryogenesis in vitro. BMC Gen. 2015, 16, 301. [Google Scholar] [CrossRef] [Green Version]
  68. Szczygieł-Sommer, A.; Gaj, M.D. The miR396-GRF regulatory module controls the embryogenic response in Arabidopsis via an auxin-related pathway. Int. J. Mol. Sci. 2019, 20, 5221. [Google Scholar] [CrossRef] [Green Version]
  69. Farina, P.; Fambrini, M.; Pugliesi, C.; Viviani, A. Expression of homeobox genes during in vitro culture of Lactuca sativa. Plant Biosyst. 2021, 155, 609–621. [Google Scholar] [CrossRef]
  70. Zuo, J.; Niu, Q.W.; Frugis, G.; Chua, N.H. The WUSCHEL gene promotes vegetative-to-embryonic transition in Arabidopsis. Plant J. 2002, 30, 349–359. [Google Scholar] [CrossRef]
  71. Costa, S.; Shaw, P. ‘Open minded’ cells: How cells can change fate. Trends Cell Biol. 2007, 17, 101–106. [Google Scholar] [CrossRef]
  72. Yamamoto, N.; Kobayashi, H.; Togashi, T.; Mori, Y.; Kikuchi, K.; Kuriyama, K.; Tokuji, Y. Formation of embryogenic cell clumps from carrot epidermal cells is suppressed by 5-azacytidine, a DNA methylation inhibitor. J. Plant Physiol. 2005, 162, 47–54. [Google Scholar] [CrossRef]
  73. Wójcikowska, B.; Wójcik, A.M.; Gaj, M.D. Epigenetic regulation of auxin-induced somatic embryogenesis in plants. Int. J. Mol. Sci. 2020, 21, 2307. [Google Scholar] [CrossRef] [Green Version]
  74. Aflaki, F.; Gutzat, R.; Mozgová, I. Chromatin during plant regeneration: Opening towards root identity? Curr. Opin. Plant Biol. 2022, 69, 102265. [Google Scholar] [CrossRef] [PubMed]
  75. Gehring, W.J.; Hiromi, Y. Homeotic genes and the homeobox. Annu. Rev. Genet. 1986, 20, 147–173. [Google Scholar] [CrossRef] [PubMed]
  76. Bürglin, T.R.; Affolter, M. Homeodomain proteins: An update. Chromosoma 2016, 125, 497–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Fischer, F.; Grigolon, G.; Benner, C.; Ristow, M. Evolutionary conserved transcription factors as regulators of longevity and targets for geroprotection. Physiol. Rev. 2022, 102, 1449–1494. [Google Scholar] [CrossRef] [PubMed]
  78. Vollbrecht, E.; Veit, B.; Sinha, N.; Hake, S. The developmental gene Knotted-1 is a member of a maize homeobox gene family. Nature 1991, 350, 241–243. [Google Scholar] [CrossRef]
  79. Bryan, A.A.; Sass, J.A. Heritable characters in maize: 51-“Knotted Leaf”. J. Hered. 1941, 32, 342–346. [Google Scholar] [CrossRef]
  80. Laux, T.; Mayer, K.F.; Berger, J.; Jürgens, G. The WUSCHEL gene is required for shoot and floral meristem integrity in Arabidopsis. Development 1996, 122, 87–96. [Google Scholar] [CrossRef]
  81. Leibfried, A.; To, J.P.; Busch, W.; Stehling, S.; Kehle, A.; Demar, M.; Kieber, J.J.; Lohmann, J.U. WUSCHEL controls meristem function by direct regulation of cytokinin-inducible response regulators. Nature 2005, 438, 1172–1175. [Google Scholar] [CrossRef]
  82. Yadav, R.K.; Girke, T.; Pasala, S.; Xie, M.; Reddy, G.V. Gene expression map of the Arabidopsis shoot apical meristem stem cell niche. Proc. Natl. Acad. Sci. USA 2009, 106, 4941–4946. [Google Scholar] [CrossRef] [Green Version]
  83. Yadav, R.K.; Perales, M.; Gruel, J.; Girke, T.; Jönsson, H.; Reddy, G.V. WUSCHEL protein movement mediates stem cell homeostasis in the Arabidopsis shoot apex. Genes Dev. 2011, 25, 2025–2030. [Google Scholar] [CrossRef] [Green Version]
  84. Salvini, M.; Fambrini, M.; Giorgietti, L.; Pugliesi, C. Molecular aspects of zygotic embryogenesis in sunflower (Helianthus annuus L.): Correlation of positive histone marks with HaWUS expression and putative link HaWUS/HaL1L. Planta 2016, 243, 199–215. [Google Scholar] [CrossRef] [PubMed]
  85. Soyars, C.L.; James, S.R.; Nimchuk, Z.L. Ready, aim, shoot: Stem cell regulation of the shoot apical meristem. Curr. Opin. Plant Biol. 2016, 29, 163–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Singh, A.; Yadav, C.B.; Tabassum, N.; Bajpeyee, A.K.; Verma, V. Stem cell niche: Dynamic neighbor of stem cells. Eur. J. Cell Biol. 2019, 98, 65–73. [Google Scholar] [CrossRef]
  87. van der Graaff, E.; Laux, T.; Rensing, S.A. The WUS homeobox containing (WOX) protein family. Genome Biol. 2009, 10, 248. [Google Scholar] [CrossRef] [PubMed]
  88. Jha, P.; Ochatt, S.J.; Kumar, V. WUSCHEL: A master regulator in plant growth signaling. Plant Cell Rep. 2020, 39, 431–444. [Google Scholar] [CrossRef] [PubMed]
  89. Ikeda, M.; Mitsuda, N.; Ohme-Takagi, M. Arabidopsis WUSCHEL is a bifunctional transcription factor that acts as a repressor in stem cell regulation and as an activator in floral patterning. Plant Cell 2009, 21, 3493–3505. [Google Scholar] [CrossRef] [Green Version]
  90. Liu, X.; Kim, Y.J.; Müller, R.; Yumul, R.E.; Liu, C.; Pan, Y.; Cao, X.; Goodrich, J.; Chen, X. AGAMOUS terminates floral stem cell maintenance in Arabidopsis by directly repressing WUSCHEL through recruitment of Polycomb Group proteins. Plant Cell 2011, 23, 3654–3670. [Google Scholar] [CrossRef] [Green Version]
  91. Liu, X.; Dinh, T.T.; Li, D.; Shi, B.; Li, Y.; Cao, X.; Guo, L.; Pan, Y.; Jiao, Y.; Chen, X. AUXIN RESPONSE FACTOR 3 integrates the functions of AGAMOUS and APETALA2 in floral meristem determinacy. Plant J. 2014, 80, 629–641. [Google Scholar] [CrossRef] [Green Version]
  92. Costanzo, E.; Trehin, C.; Vandenbussche, M. The role of WOX genes in flower development. Ann. Bot. 2014, 114, 1545–1553. [Google Scholar] [CrossRef] [Green Version]
  93. Li, H.; Li, C.; Wang, Y.; Qin, X.; Meng, L.; Sun, X. Genome-wide analysis of the WOX transcription factor genes in Dendrobium catenatum Lindl. Genes 2022, 13, 1481. [Google Scholar] [CrossRef]
  94. Yang, Z.; Qian, G.; Qin, W.; Yang, Z.; Li, F. Genome-wide analysis of WOX genes in upland cotton and their expression pattern under different stresses. BMC Plant Biol. 2017, 17, 113. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, M.-M.; Liu, M.-M.; Ran, F.; Guo, P.-C.; Ke, Y.-Z.; Wu, Y.-W.; Wen, J.; Li, P.-F.; Li, J.-N.; Du, H. Global analysis of WOX transcription factor gene family in Brassica napus reveals their stress- and hormone-responsive patterns. Int. J. Mol. Sci. 2018, 19, 3470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Li, X.; Hamyat, M.; Liu, C.; Salman, A.; Gao, X.; Guo, C.; Wang, Y.; Guo, Y. Identification and characterization of the WOX family genes in five Solanaceae species reveal their conserved roles in peptide signaling. Genes 2018, 9, 260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Hao, Q.; Zhang, L.; Yang, Y.; Shan, Z.; Zhou, X.A. Genome-wide analysis of the WOX gene family and function exploration of GmWOX18 in soybean. Plants 2019, 8, 215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Han, N.; Tang, R.; Chen, X.; Xu, Z.; Ren, Z.; Wang, L. Genome-wide identification and characterization of WOX genes in Cucumis sativus. Genome 2021, 64, 761–776. [Google Scholar] [CrossRef]
  99. Rathour, M.; Sharma, A.; Kaur, A.; Upadhyay, S.K. Genome-wide characterization and expression and co-expression analysis suggested diverse functions of WOX genes in bread wheat. Heliyon 2021, 6, e05762. [Google Scholar] [CrossRef] [PubMed]
  100. Feng, C.; Zou, S.; Gao, P.; Wang, Z. In silico identification, characterization expression profile of WUSCHEL-Related Homeobox (WOX) gene family in two species of kiwifruit. PeerJ 2021, 9, e12348. [Google Scholar] [CrossRef]
  101. Akbulut, S.E.; Okay, A.; Aksoy, T.; Aras, E.S.; Büyük, İ. The genome-wide characterization of WOX gene family in Phaseolus vulgaris L. during salt stress. Physiol. Mol. Biol. Plants 2022, 28, 1297–1309. [Google Scholar] [CrossRef]
  102. Wang, Z.; Cai, Q.; Xia, H.; Han, B.; Li, M.; Wang, Y.; Zhu, M.; Jiao, C.; Wang, D.; Zhu, J.; et al. Genome-wide identification and comparative analysis of WOX genes in four Euphorbiaceae species and their expression patterns in Jatropha curcas. Front. Genet. 2022, 13, 878554. [Google Scholar] [CrossRef]
  103. Yang, T.; Gao, T.; Wang, C.; Wang, X.; Chen, C.; Tian, M.; Yang, W. In silico genome wide identification and expression analysis of the WUSCHEL-related homeobox gene family in Medicago sativa. Genom. Inform. 2022, 20, e19. [Google Scholar] [CrossRef]
  104. Long, J.A.; Ohno, C.; Smith, Z.R.; Meyerowitz, E.M. TOPLESS regulates apical embryonic fate in Arabidopsis. Science 2006, 312, 1520–1523. [Google Scholar] [CrossRef] [PubMed]
  105. Paponov, I.A.; Tale, W.; Lang, D.; Paponov, M.; Reski, R.; Rensing, S.A.; Palme, K. The evolution of nuclear auxin signalling. BMC Evol. Biol. 2009, 9, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Zhao, Y.; Hu, Y.; Dai, M.; Huang, L.; Zhou, D.-X. The WUSCHEL-related homeobox gene WOX11 is required to activate shoot-borne crown root development in rice. Plant Cell 2009, 21, 736–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Dai, M.; Hu, Y.; Zhao, Y.; Liu, H.; Zhou, D.X. A WUSCHEL-LIKE HOMEOBOX gene represses a YABBY gene expression required for rice leaf development. Plant Physiol. 2007, 144, 380–390. [Google Scholar] [CrossRef] [Green Version]
  108. Park, S.O.; Zheng, Z.; Oppenheimer, D.G.; Hauser, B.A. The PRETTY FEW SEEDS2 gene encodes ad Arabidopsis homeodomain protein that regulates ovule development. Development 2005, 132, 841–849. [Google Scholar] [CrossRef] [Green Version]
  109. Sarkar, A.K.; Luijten, M.; Miyashima, S.; Lenhard, M.; Hashimoto, T.; Nakajima, K.; Scheres, B.; Heidstra, R.; Laux, T. Conserved factors regulate signalling in Arabidopsis thaliana shoot and root stem cell organizers. Nature 2007, 446, 811–814. [Google Scholar] [CrossRef]
  110. Mayer, K.F.; Schoof, H.; Haecker, A.; Lenhard, M.; Jürgens, G.; Laux, T. Role of WUSCHEL in regulating stem cell fate in the Arabidopsis shoot meristem. Cell 1998, 95, 805–815. [Google Scholar] [CrossRef] [Green Version]
  111. Baürle, I.; Laux, T. Regulation of WUSCHEL transcription in the stem cell niche of the Arabidopsis shoot meristem. Plant Cell 2000, 17, 2271–2280. [Google Scholar] [CrossRef] [Green Version]
  112. Ditengou, F.A.; Teale, W.D.; Kochersperger, P.; Flittner, K.A.; Kneuper, I.; van der Graaff, E.; Nziengui, H.; Pinosa, F.; Li, X.; Nitschke, T.; et al. Mechanical induction of lateral root initiation in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2008, 105, 18818–18823. [Google Scholar] [CrossRef] [Green Version]
  113. Kieffer, M.; Stern, Y.; Cook, H.; Clerici, E.; Maulbetsch, C.; Laux, T.; Davies, B. Analysis of the transcription factor WUSCHEL and its functional homologue in Antirrhinum reveals a potential mechanism for their roles in meristem maintenance. Plant Cell 2006, 18, 560–573. [Google Scholar] [CrossRef] [Green Version]
  114. Nardmann, J.; Werr, W. The shoot stem cell niche in angiosperms: Expression patterns of WUS orthologues in rice and maize imply major modifications in the course of mono- and dicot evolution. Mol. Biol. Evol. 2006, 23, 2492–2504. [Google Scholar] [CrossRef] [PubMed]
  115. Deyhle, F.; Sarkar, A.K.; Tucker, E.J.; Laux, T. WUSCHEL regulates cell differentiation during anther development. Dev. Biol. 2007, 302, 154–159. [Google Scholar] [CrossRef] [PubMed]
  116. Shimizu, R.; Ji, J.; Kelsey, E.; Ohtsu, K.; Schnable, P.S.; Scanlon, M.J. Tissue specificity and evolution of meristematic WOX3 function. Plant Physiol. 2009, 149, 841–850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Zhu, J.; Shi, H.; Lee, B.-H.; Damsz, B.; Cheng, S.; Stirm, V.; Zhu, J.-K.; Hasegawa, P.M.; Bressan, R.A. An Arabidopsis homeodomain transcription factor gene, HOS9, mediates cold tolerance through a CBF-independent pathway. Proc. Natl. Acad. Sci. USA 2004, 101, 9873–9878. [Google Scholar] [CrossRef] [Green Version]
  118. Breuninger, H.; Rikirsch, E.; Hermann, M.; Ueda, M.; Laux, T. Differential expression of WOX genes mediates apical-basal axis formation in the Arabidopsis embryo. Dev. Cell 2008, 14, 867–876. [Google Scholar] [CrossRef] [Green Version]
  119. Peret, B.; De Rybel, B.; Casimiro, I.; Benková, E.; Swarup, R.; Laplaze, L.; Beeckman, T.; Bennett, M.J. Arabidopsis lateral root development: An emerging story. Trends Plant Sci. 2009, 14, 399–408. [Google Scholar] [CrossRef]
  120. Wu, X.; Dabi, T.; Weigel, D. Requirement of homeobox gene STIMPY/WOX9 for Arabidopsis meristem growth and maintenance. Curr. Biol. 2005, 15, 436–440. [Google Scholar] [CrossRef] [Green Version]
  121. Wu, X.; Chory, J.; Weigel, D. Combinations of WOX activities regulate tissue proliferation during Arabidopsis embryonic development. Dev. Biol. 2007, 309, 306–316. [Google Scholar] [CrossRef] [Green Version]
  122. Lippman, Z.B.; Cohen, O.; Alvarez, J.P.; Abu-Abied, M.; Pekker, I.; Paran, I.; Eshed, Y.; Zamir, D. The making of a compound inflorescence in tomato and related nightshades. PLoS Biol. 2008, 6, e288. [Google Scholar] [CrossRef]
  123. Rebocho, A.B.; Bliek, M.; Kusters, E.; Castel, R.; Procissi, A.; Roobeek, I.; Souer, E.; Koes, R. Role of EVERGREEN in the development of the cymose petunia inflorescence. Dev. Cell 2008, 15, 437–447. [Google Scholar] [CrossRef] [Green Version]
  124. Deveaux, Y.; Toffano-Nioche, C.; Claisse, G.; Thareau, V.; Morin, H.; Laufs, P.; Moreau, H.; Kreis, M.; Lecharny, A. Genes of the most conserved WOX clade in plants affect root and flower development in Arabidopsis. BMC Evol. Biol. 2008, 8, 291. [Google Scholar] [CrossRef] [PubMed]
  125. Kawai, T.; Shibata, K.; Akahoshi, R.; Nishiuchi, S.; Takahashi, H.; Nakazono, M.; Kojima, T.; Nosaka-Takahashi, M.; Sato, Y.; Toyoda, A.; et al. WUSCHEL-related homeobox family genes in rice control lateral root primordium size. Proc. Natl. Acad. Sci. USA 2022, 119, e2101846119. [Google Scholar] [CrossRef] [PubMed]
  126. Schoof, H.; Lenhard, M.; Haecker, A.; Mayer, K.F.; Jürgens, G.; Laux, T. The stem cell population of Arabidopsis shoot meristems is maintained by a regulatory loop between the CLAVATA and WUSCHEL genes. Cell 2000, 100, 635–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Fletcher, J.C. The CLV-WUS stem cell signaling pathway: A roadmap to crop yield optimization. Plants 2018, 7, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Dao, T.Q.; Weksler, N.; Liu, H.M.; Leiboff, S.; Fletcher, J.C. Interactive CLV3, CLE16, and CLE17 signaling mediates stem cell homeostasis in the Arabidopsis shoot apical meristem. Development 2022, 149, dev.200787. [Google Scholar] [CrossRef] [PubMed]
  129. Nicolas, A.; Laufs, P. Meristem initiation and de novo stem cell formation. Front. Plant Sci. 2022, 13, 891228. [Google Scholar] [CrossRef] [PubMed]
  130. Shimotohno, A. Illuminating the molecular mechanisms underlying shoot apical meristem homeostasis in plants. Plant Biotechnol. 2022, 39, 19–28. [Google Scholar] [CrossRef] [PubMed]
  131. Brand, U.; Fletcher, J.C.; Hobe, M.; Meyerowitz, E.M.; Simon, R. Dependence of stem cell fate in Arabidopsis on a feedback loop regulated by CLV3 activity. Science 2000, 289, 617–619. [Google Scholar] [CrossRef]
  132. Lenhard, M.; Bohnert, A.; Jürgens, G.; Laux, T. Termination of stem cell maintenance in Arabidopsis floral meristems by interactions between WUSCHEL and AGAMOUS. Cell 2001, 105, 805–814. [Google Scholar] [CrossRef] [Green Version]
  133. Stahl, Y.; Wink, R.H.; Ingram, G.C.; Simon, R. A signaling module controlling the stem cell niche in Arabidopsis root meristems. Curr. Biol. 2009, 19, 909–914. [Google Scholar] [CrossRef] [Green Version]
  134. Gordon, S.P.; Chickarmane, V.S.; Ohno, C.; Meyerowitz, E.M. Multiple feedback loops through cytokinin signaling control stem cell number within the Arabidopsis shoot meristem. Proc. Natl. Acad. Sci. USA 2009, 106, 16529–16534. [Google Scholar] [CrossRef] [PubMed]
  135. He, G.; Cao, Y.; Song, M.; Bi, M.; Tang, Y.; Xu, L.; Ming, J.; Yang, P. WUSCHEL-related homeobox genes cooperate with cytokinin to promote bulbil formation in Lilium lancifolium. Plant Physiol. 2022, 190, 387–402. [Google Scholar] [CrossRef] [PubMed]
  136. Guo, J.; Liu, H.; He, Y.; Cui, X.; Du, X.; Zhu, J. Origination of asexual plantlets in three species of Crassulaceae. Protoplasma 2015, 252, 591–603. [Google Scholar] [CrossRef] [PubMed]
  137. Ikeuchi, M.; Iwase, A.; Ito, T.; Tanaka, H.; Favero, D.S.; Kawamura, A.; Sakamoto, S.; Wakazaki, M.; Tameshige, T.; Fujii, H.; et al. Wound-inducible WUSCHEL-RELATED HOMEOBOX 13 is required for callus growth and organ reconnection. Plant Physiol. 2022, 188, 425–441. [Google Scholar] [CrossRef]
  138. Wu, H.; Qu, X.; Dong, Z.; Luo, L.; Shao, C.; Forner, J.; Lohmann, J.U.; Su, M.; Xu, M.; Liu, X.; et al. WUSCHEL triggers innate antiviral immunity in plant stem cells. Science 2020, 370, 227–231. [Google Scholar] [CrossRef]
  139. Delseny, M. Innate antiviral immunity in plant apical meristem explained. C. R. Biol. 2021, 343, 5–6. [Google Scholar] [CrossRef]
  140. Tang, Y.; Li, H.; Guan, Y.; Li, S.; Xun, C.; Dong, Y.; Huo, R.; Guo, Y.; Bao, X.; Pei, E.; et al. Genome-wide identification of the physic nut WUSCHEL-related homeobox gene family and functional analysis of the abiotic stress responsive gene JcWOX5. Front. Genet. 2020, 11, 670. [Google Scholar] [CrossRef]
  141. Minh-Thu, P.-T.; Kim, J.S.; Chae, S.; Jun, K.M.; Lee, G.-S.; Kim, D.-E.; Cheong, J.-J.; Song, S.I.; Nahm, B.H.; Kim, Y.-K. A WUSCHEL homeobox transcription factor, OsWOX13, enhances drought tolerance and triggers early flowering in rice. Mol. Cells 2018, 41, 781–798. [Google Scholar]
  142. Wang, L.-Q.; Li, Z.; Wen, S.-S.; Wang, J.-N.; Zhao, A.-T.; Lu, M.-Z. WUSCHEL-related homeobox gene PagWOX11/12a responds to drought stress by enhancing root elongation and biomass growth in poplar. J. Exp. Bot. 2020, 71, 1503–1513. [Google Scholar] [CrossRef]
  143. Wang, P.; Guo, Y.; Chen, X.; Zheng, X.; Sun, Y.; Yang, J.; Ye, N. Genome-wide identification of WOX genes and their expression patterns under different hormone and abiotic stress treatments in tea plant (Camellia sinensis). Trees 2019, 33, 1129–1142. [Google Scholar] [CrossRef]
  144. Winnicki, K. The winner takes it all: Auxin–the main player during plant embryogenesis. Cells 2020, 9, 606. [Google Scholar] [CrossRef] [Green Version]
  145. Geldner, N.; Anders, N.; Wolters, H.; Keicher, J.; Kornberger, W.; Muller, P.; Delbarre, A.; Ueda, T.; Nakano, A.; Jürgens, G. The Arabidopsis GNOM ARF-GEF mediates endosomal recycling, auxin transport, and auxin-dependent plant growth. Cell 2003, 112, 219–230. [Google Scholar] [CrossRef]
  146. Friml, J.; Vieten, A.; Sauer, A.; Weijers, M.; Schwarz, H.; Hamann, T.; Offringa, R.; Jürgens, G. Efflux-dependent auxin gradients establish the apical–basal axis of Arabidopsis. Nature 2003, 426, 147–153. [Google Scholar] [CrossRef]
  147. Dhonukshe, P.; Tanaka, H.; Goh, T.; Ebine, K.; Mähönen, A.P.; Prasad, K.; Blilou, I.; Geldner, N.; Xu, J.; Uemura, T.; et al. Generation of cell polarity in plants links endocytosis, auxin distribution and cell fate decisions. Nature 2008, 456, 962–966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Ung, K.L.; Winkler, M.; Schulz, L.; Kolb, M.; Janacek, D.P.; Dedic, E.; Stokes, D.L.; Hammes, U.Z.; Pederson, B.P. Structures and mechanism of the plant PIN-FORMED auxin transporter. Nature 2022, 609, 605–610. [Google Scholar] [CrossRef] [PubMed]
  149. Steinmann, T.; Geldner, N.; Grebe, M.; Mangold, S.; Jackson, C.L.; Paris, S.; Gälweiler, L.; Palme, K.; Jurgens, G. Coordinated polar localization of auxin efflux carrier PIN1 by GNOM ARF GEF. Science 1999, 286, 316–318. [Google Scholar] [CrossRef] [PubMed]
  150. Ueda, M.; Zhang, Z.; Laux, T. Transcriptional activation of Arabidopsis axis patterning genes WOX8/9 links zygote polarity to embryo development. Development 2011, 20, 264–270. [Google Scholar]
  151. Haecker, A.; Groβ-Hardt, R.; Geiges, B.; Sarkar, A.; Breuninger, H.; Hermann, M.; Laux, T. Expression dynamics of WOX genes mark cell fate decisions during early embryonic patterning in Arabidopsis thaliana. Development 2004, 131, 657–668. [Google Scholar] [CrossRef] [Green Version]
  152. Nardmann, J.; Zimmermann, R.; Durantini, D.; Kranz, E.; Werr, W. WOX gene phylogeny in Poaceae: A comparative approach addressing leaf and embryo development. Mol. Biol. Evol. 2007, 24, 2474–2484. [Google Scholar] [CrossRef]
  153. Zhu, T.; Moschou, P.N.; Alvarez, J.M.; Sohlberg, J.J.; von Arnold, S. WUSCHEL-RELATED HOMEOBOX 2 is important for protoderm and suspensor development in the gymnosperm Norway spruce. BMC Plant Biol. 2016, 16, 19. [Google Scholar] [CrossRef] [Green Version]
  154. Jiménez, V.M. Involvement of plant hormones and plant growth regulators on in vitro somatic embryogenesis. Plant Growth Regul. 2005, 47, 91–110. [Google Scholar] [CrossRef]
  155. Su, Y.; Zhao, X.Y.; Liu, Y.B.; Zhang, C.; O’Neill, S.D.; Zhang, X.S. Auxin-induced WUS expression is essential for embryonic stem cell renewal during somatic embryogenesis in Arabidopsis. Plant J. 2009, 59, 448–460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Wójcikowska, B.; Jaskóła, K.; Gąsiorek, P.; Meus, M.; Nowak, K.; Gaj, M.D. LEAFY COTYLEDON2 (LEC2) promotes embryogenic induction in somatic tissues of Arabidopsis, via YUCCA-mediated auxin biosynthesis. Planta 2013, 238, 425–440. [Google Scholar] [CrossRef]
  157. Zhao, X.Y.; Su, Y.H.; Cheng, Z.J.; Zhang, X.S. Cell fate switch during in vitro plant organogenesis. J. Integr. Plant Biol. 2008, 50, 816–824. [Google Scholar] [CrossRef]
  158. Su, Y.H.; Liu, Y.B.; Bai, B.; Zhang, X.S. Establishment of embryonic shoot-root axis is involved in auxin and cytokinin response during Arabidopsis somatic embryogenesis. Front. Plant Sci. 2015, 5, 792. [Google Scholar] [CrossRef] [Green Version]
  159. Tvorogova, V.E.; Lebedeva, M.A.; Lutova, L.A. Expression of WOX and PIN genes during somatic and zygotic embryogenesis in Medicago truncatula. Russ. J. Genet. 2015, 51, 1189–1198. [Google Scholar] [CrossRef]
  160. Palovaara, J.; Hakman, I. Conifer WOX-related homeodomain transcription factors, developmental consideration and expression dynamic of WOX2 during Picea abies somatic embryogenesis. Plant Mol. Biol. 2008, 66, 533–549. [Google Scholar] [CrossRef]
  161. Palovaara, J.; Hallberg, H.; Stasolla, C.; Hakman, I. Comparative expression pattern analysis of WUSCHEL-related homeobox 2 (WOX2) and WOX8/9 in developing seeds and somatic embryos of the gymnosperm Picea abies. New Phytol. 2010, 188, 122–135. [Google Scholar] [CrossRef]
  162. Palovaara, J.; Hakman, I. WOX2 and polar auxin transport during spruce embryo axis formation. Plant Signal. Behav. 2009, 4, 153–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Zhao, S.; Jiang, Q.-T.; Ma, J.; Zhang, X.-W.; Zhao, Q.-Z.; Wang, X.-Y.; Wang, C.-S.; Cao, X.; Lu, Z.-X.; Zheng, Y.-L.; et al. Characterization and expression analysis of WOX5 genes from wheat and its relatives. Gene 2014, 537, 63–69. [Google Scholar] [CrossRef]
  164. Qi, S.; Zhao, R.; Yan, Y.; Huang, C.; Li, H.; Chen, S.; Zhang, T.; Kong, L.; Zhao, J.; Zhang, J. Global transcriptome and coexpression network analyses reveal new insights into somatic embryogenesis in hybrid sweetgum (Liquidambar styraciflua × Liquidambar formosana). Front. Plant Sci. 2021, 12, 751866. [Google Scholar] [CrossRef] [PubMed]
  165. Gordon, S.P.; Heisler, M.G.; Reddy, G.V.; Ohno, C.; Das, P.; Meyerowitz, E.M. Pattern formation during de novo assembly of the Arabidopsis shoot meristem. Development 2007, 134, 3539–3548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Dai, X.; Liu, Z.; Qiao, M.; Li, J.; Li, S.; Xiang, F. ARR12 promotes de novo shoot regeneration in Arabidopsis thaliana via activation of WUSCHEL expression. J. Integr. Plant Biol. 2017, 59, 747–758. [Google Scholar] [CrossRef] [PubMed]
  167. Zhai, N.; Xu, L. Pluripotency acquisition in the middle cell layer of callus is required for organ regeneration. Nat. Plants 2021, 7, 1453–1460. [Google Scholar] [CrossRef] [PubMed]
  168. Klimaszewska, K.; Overton, C.; Tewart, D.; Rutledge, R.G. Initiation of somatic embryos and regeneration of plants from primordial shoots of 10-year-old somatic white spruce and expression profiles of 11 genes followed during the tissue culture process. Planta 2011, 233, 635–647. [Google Scholar] [CrossRef]
  169. Rupps, A.; Raschke, J.; Rümmler, M.; Linke, B.; Zoglauer, K. Identification of putative homologs of Larix decidua to BABYBOOM (BBM), LEAFY COTYLEDON1 (LEC1), WUSCHEL-related HOMEOBOX2 (WOX2) and SOMATIC EMBRYOGENESIS RECEPTOR like KINASE (SERK) during somatic embryogenesis. Planta 2016, 243, 473–488. [Google Scholar] [CrossRef]
  170. Hassani, S.B.; Trontin, J.-F.; Raschke, J.; Zoglauer, K.; Rupps, A. Constitutive overexpression of a conifer WOX2 homolog affects somatic embryo development in Pinus pinaster and promotes somatic embryogenesis and organogenesis in Arabidopsis seedlings. Front. Plant Sci. 2022, 13, 838421. [Google Scholar] [CrossRef]
  171. Orłowska, A.; Kępczyńska, E. Identification of Polycomb Repressive Complex1, Trithorax group genes and their simultaneous expression with WUSCHEL, WUSCHEL-related Homeobox5 and SHOOT MERISTEMLESS during the induction phase of somatic embryogenesis in Medicago truncatula Gaertn. Plant Cell Tissue Organ Cult. 2018, 134, 345–356. [Google Scholar] [CrossRef] [Green Version]
  172. Chen, S.K.; Kurdyukov, S.; Kereszt, A.; Wang, X.D.; Gresshoff, P.M.; Rose, R.J. The association of homeobox gene expression with stem cell formation and morphogenesis in cultured M. truncatula. Planta 2009, 230, 827–840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Izuno, A.; Maruyama, T.E.; Ueno, S.; Ujino-Ihara, T.; Moriguchi, Y. Genotype and transcriptome effects on somatic embryogenesis in Cryptomeria japonica. PLoS ONE 2020, 15, e0244634. [Google Scholar] [CrossRef] [PubMed]
  174. Lowe, K.; Wu, E.; Wang, N.; Hoerster, G.; Hastings, C.; Cho, M.J.; Scelonge, C.; Lenderts, B.; Chamberlin, M.; Cushatt, J.; et al. Morphogenic regulators Baby boom and Wuschel improve monocot transformation. Plant Cell 2016, 28, 1998–2015. [Google Scholar] [CrossRef] [Green Version]
  175. Anand, A.; Bass, S.H.; Wu, E.; Wang, N.; Mcbride, K.E.; Annaluru, N.; Miller, M.; Hua, M.; Jones, T.J. An improved ternary vector system for Agrobacterium-mediated rapid maize transformation. Plant Mol. Biol. 2018, 97, 187–200. [Google Scholar] [CrossRef] [Green Version]
  176. Mookkan, M.; Nelson-Vasilchik, K.; Hague, J.; Zhang, Z.J.; Kausch, A.P. Selectable marker independent transformation of recalcitrant maize inbred B73 and sorghum P898012 mediated by morphogenic regulators BABY BOOM and WUSCHEL2. Plant Cell Rep. 2017, 36, 1477–1491. [Google Scholar] [CrossRef] [PubMed]
  177. Luo, G.; Palmgren, M. GRF-GIF chimeras boost plant regeneration. Trends Plant Sci. 2021, 26, 201–204. [Google Scholar] [CrossRef] [PubMed]
  178. Maren, N.A.; Duan, H.; Da, K.; Yencho, G.C.; Ranney, T.G.; Liu, W. Genotype-independent plant transformation. Hortic. Res. 2022, 9, uhac047. [Google Scholar] [CrossRef] [PubMed]
  179. Lian, Z.; Nguyen, C.D.; Liu, L.; Wang, G.; Chen, J.; Wang, S.; Yi, G.; Wilson, S.; Ozias-Akins, P.; Gong, H.; et al. Application of developmental regulators to improve in planta or in vitro transformation in plants. Plant Biotechnol. J. 2022, 20, 1622–1635. [Google Scholar] [CrossRef]
  180. Rashid, S.Z.; Yamaji, N.; Kyo, M. Shoot formation from root tip region: A developmental alteration by WUS in transgenic tobacco. Plant Cell Rep. 2007, 26, 1449–1455. [Google Scholar] [CrossRef]
  181. Arroyo-Herrera, A.; Ku-Gonzalez, A.; Canche-Moo, R.; Quiroz-Figueroa, F.R.; Loyola-Vargas, V.M.; Rodriguez-Zapata, L.C.; D’Hondt, C.B.; Suárez-Solís, V.M.; Castaño, E. Expression of WUSCHEL in Coffea canephora causes ectopic morphogenesis and increases somatic embryogenesis. Plant Cell Tissue Organ Cult. 2008, 94, 171–180. [Google Scholar] [CrossRef]
  182. Klimaszewska, K.; Pelletier, G.; Overton, C.; Stewart, D.; Rutledge, R.G. Hormonally regulated overexpression of Arabidopsis WUS and conifer LEC1 (CHAP3A) in transgenic white spruce: Implications for somatic embryo development and somatic seedling growth. Plant Cell Rep. 2010, 29, 723–734. [Google Scholar] [CrossRef]
  183. Kyo, M.; Maida, K.; Nishioka, Y.; Matsui, K. Coexpression of WUSCHEL related homeobox (WOX) 2 with WOX8 or WOX9 promotes regeneration from leaf segments and free cells in Nicotiana tabacum L. Plant Biotechnol. 2018, 35, 23–30. [Google Scholar] [CrossRef] [Green Version]
  184. Bianchi, P.A.; Almeida da Silva, L.R.; da Silva Alencar, A.A.; Diniz Santos, P.H.A.; Pimenta, S.; Pombo Sudré, C.; Erpen-Dalla Corte, L.; Azeredo Conçalves, L.S.; Rodrigues, R. Biomorphological characterization of Brazilian Capsicum Chinense Jacq. germplasm. Agronomy 2020, 10, 447. [Google Scholar] [CrossRef] [Green Version]
  185. Solís-Ramos, L.Y.; González-Estrada, T.; Nahuath-Dzib, S.; Zapata-Rodriguez, L.C.; Castaño, E. Overexpression of WUSCHEL in C. chinense causes ectopic morphogenesis. Plant Cell Tissue Organ Cult. 2009, 96, 279–287. [Google Scholar] [CrossRef]
  186. Bouchabké-Coussa, O.; Obellianne, M.; Linderme, D.; Montes, E.; Maia-Grondard, A.; Vilaine, F.; Pannetier, C. Wuschel overexpression promotes somatic embryogenesis and induces organogenesis in cotton (Gossypium hirsutum L.) tissues cultured in vitro. Plant Cell Rep. 2013, 32, 675–686. [Google Scholar] [CrossRef] [PubMed]
  187. Zheng, W.; Zhang, X.; Yang, Z.; Wu, J.; Li, F.; Duan, L.; Liu, C.; Lu, L.; Zhang, C.; Li, F. AtWuschel promotes formation of the embryogenic callus in Gossypium hirsutum. PLoS ONE 2014, 9, e87502. [Google Scholar] [CrossRef] [Green Version]
  188. Kadri, A.; De March, G.G.; Guerineau, F.; Cosson, V.; Ratet, P. WUSCHEL overexpression promotes callogenesis and somatic embryogenesis in Medicago truncatula Gaertn. Plants 2021, 10, 715. [Google Scholar] [CrossRef] [PubMed]
  189. Yoshioka, Y.; Takahashi, Y.; Matsuoka, K.; Nakamura, K.; Koizumi, J.; Kojima, M.; Machida, Y. Transient gene expression in plant cells mediated by Agrobacterium tumefaciens: Application for the analysis of virulence loci. Plant Cell Physiol. 1996, 37, 782–789. [Google Scholar] [CrossRef] [Green Version]
  190. Gelvin, S.B. Agrobacterium-mediated plant transformation: The biology behind the “gene-jockeying” tool. Microbiol. Mol. Biol. Rev. 2003, 67, 16–37. [Google Scholar] [CrossRef] [Green Version]
  191. De Buck, S.; De Wilde, C.; Van Montagu, M.; Depicker, A. Determination of the T-DNA transfer and the T-DNA integration frequencies upon cocultivation of Arabidopsis thaliana root explants. Mol. Plant Microbe Inter. 2000, 13, 658–665. [Google Scholar] [CrossRef] [Green Version]
  192. Wang, N.; Arling, M.; Hoerster, G.; Ryan, L.; Wu, E.; Lowe, K.; Gordon-Kamm, W.; Jones, T.J.; Chilchoat, N.D.; Anand, A. An efficient gene excision system in maize. Front. Plant Sci. 2020, 11, 1298. [Google Scholar] [CrossRef]
  193. Hoerster, G.; Wang, N.; Ryan, L.; Wu, E.; Anand, A.; McBride, K.; Lowe, K.; Jones, T.; Gordon-Kamm, B. Use of non-integrating Zm-Wus2 vectors to enhance maize transformation. Vitr. Cell. Dev. Biol. Plant 2020, 56, 265–279. [Google Scholar] [CrossRef] [Green Version]
  194. Wang, K.; Shi, L.; Liang, X.; Zhao, P.; Wang, W.; Liu, J.; Chang, Y.; Hiei, Y.; Yanagihara, C.; Du, L.; et al. The gene TaWOX5 overcomes genotype dependency in wheat genetic transformation. Nat. Plants 2022, 8, 110–117. [Google Scholar] [CrossRef]
  195. Vogel, G. Breakthrough of the year. Reprogramming cells. Science 2008, 322, 1766–1767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Wang, F.-X.; Shang, G.-D.; Wu, L.-Y.; Xu, Z.-G.; Zhao, X.-Y.; Wang, J.-W. Chromatin accessibility dynamics and a hierarchical transcriptional regulatory network structure for plant somatic embryogenesis. Dev. Cell 2020, 54, 742–757. [Google Scholar] [CrossRef] [PubMed]
  197. Zhang, Z.; Conner, J.; Guo, Y.; Ozias-Akins, P. Haploidy in tobacco induced by PsASGR-BBML transgenes via parthenogenesis. Genes 2020, 11, 1072. [Google Scholar] [CrossRef]
  198. Xu, Y.-Y.; Wang, X.-M.; Li, J.; Li, J.-H.; Wu, J.-S.; Walker, J.C.; Xu, Z.-H.; Chong, K. Activation of the WUS gene induces ectopic initiation of floral meristems on mature stem surface in Arabidopsis thaliana. Plant Mol. Biol. 2005, 57, 773–784. [Google Scholar] [CrossRef] [PubMed]
  199. Salvini, M.; Sani, E.; Fambrini, M.; Pistelli, L.; Pucciariello, C.; Pugliesi, C. Molecular analysis of a sunflower gene encoding an homologous of the B subunit of a CAAT binding factor. Mol. Biol. Rep. 2012, 39, 6449–6465. [Google Scholar] [CrossRef]
  200. Verdeil, J.-L.; Alemanno, L.; Niemenak, N.; Tranbarger, T.J. Pluripotent versus totipotent plant stem cells: Dependence versus autonomy? Trends Plant Sci. 2007, 12, 245–252. [Google Scholar] [CrossRef]
  201. Bull, T.; Michelmore, R. Molecular determinants of in vitro plant regeneration: Prospects for enhanced manipulation of lettuce (Lactuca sativa L.). Front. Plants Sci. 2022, 13, 888425. [Google Scholar] [CrossRef]
  202. Ke, Y.; Podio, M.; Conner, J.; Ozias-Akins, P. Single-cell transcriptome profiling of buffelgrass (Cenchrus ciliaris) eggs unveils apomictic parthenogenesis signatures. Sci. Rep. 2021, 11, 9880. [Google Scholar] [CrossRef]
  203. Conner, J.A.; Mookkan, M.; Huo, H.; Chae, K.; Ozias-Akins, P. A parthenogenesis gene of apomict origin elicits embryo formation from unfertilized eggs in a sexual plant. Proc. Natl. Acad. Sci. USA 2015, 112, 11205–11210. [Google Scholar] [CrossRef] [Green Version]
  204. Conner, J.A.; Podio, M.; Ozias-Akins, P. Haploid embryo production in rice and maize induced by PsASGR-BBML transgenes. Plant Reprod. 2017, 30, 41–52. [Google Scholar] [CrossRef]
  205. Anderson, S.N.; Johnson, C.S.; Chesnut, J.; Khanday, I.; Woodhouse, M.; Li, C.; Conrad, L.J.; Russell, S.D.; Sundaresan, V. The zygotic transition is initiated in unicellular plant zygotes with asymmetric activation of parental genomes. Dev. Cell 2017, 43, 349–358. [Google Scholar] [CrossRef] [PubMed]
  206. Chen, J.; Strieder, N.; Krohn, N.G.; Cyprys, P.; Sprunck, S.; Engelmann, J.C.; Dresselhaus, T. Zygotic genome activation occurs shortly after fertilization in maize. Plant Cell 2017, 29, 2106–2125. [Google Scholar] [CrossRef] [PubMed]
  207. Chen, B.; Maas, L.; Fiueiredo, D.; Zhong, Y.; Reis, R.; Li, M.; Horstman, A.; Riksen, T.; Weemen, M.; Liu, H.; et al. BABY BOOM regulates early embryo and endosperm development. Proc. Natl. Acad. Sci. USA 2022, 119, e2201761119. [Google Scholar] [CrossRef] [PubMed]
  208. Khanday, I.; Skinner, D.; Yang, B.; Mercier, R.; Sundaresan, V. A male-expressed rice embryogenic trigger redirected for asexual propagation through seeds. Nature 2019, 565, 91–95. [Google Scholar] [CrossRef] [PubMed]
  209. d’Erfurth, I.; Jolivet, S.; Froger, N.; Catrice, O.; Novatchkova, M.; Mercier, R. Turning meiosis into mitosis. PLoS Biol. 2009, 7, e1000124. [Google Scholar] [CrossRef] [Green Version]
  210. Mieulet, D.; Jolivet, S.; Rivard, M.; Cromer, L.; Vernet, A.; Mayonove, P.; Pereira, L.; Droc, G.; Courtois, B.; Guiderdoni, E.; et al. Turning rice meiosis into mitosis. Cell Res. 2016, 26, 1242–1254. [Google Scholar] [CrossRef]
  211. Xie, E.; Li, Y.; Tang, D.; Lv, Y.; Shen, Y.; Cheng, Z. A strategy for generating rice apomixis by gene editing. J. Integr. Plant Biol. 2019, 61, 911–916. [Google Scholar] [CrossRef] [Green Version]
  212. Yao, L.; Zhang, Y.; Liu, C.; Liu, Y.; Wang, Y.; Liang, D.; Liu, J.; Sahoo, G.; Kelliher, T. OsMATL mutation induces haploid seed formation in indica rice. Nat. Plants 2018, 4, 530–533. [Google Scholar] [CrossRef]
  213. Wang, C.; Liu, Q.; Shen, Y.; Hua, Y.; Wang, J.; Lin, J.; Wu, M.; Sun, T.; Cheng, Z.; Mercier, R.; et al. Clonal seeds from hybrid rice by simultaneous genome engineering of meiosis and fertilization genes. Nat. Biotechnol. 2019, 37, 283–286. [Google Scholar] [CrossRef]
  214. Xu, M.; Hu, T.; Smith, M.R.; Poethig, R.S. Epigenetic regulation of vegetative phase change in Arabidopsis. Plant Cell 2016, 28, 28–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Waddington, C.H. The Strategy of the Genes: A Discussion of Some Aspects of Theoretical Biology; Allen & Unwin: London, UK, 1957. [Google Scholar]
  216. Wu, H.; Sun, Y.E. Epigenetic regulation of stem cell differentiation. Pediatr. Res. 2006, 59, 21–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Goldberg, A.D.; Allis, C.D.; Bernstein, E. Epigenetics: A landscape lakes shape. Cell 2007, 128, 635–638. [Google Scholar] [CrossRef]
  218. Enikoff, S.; Greally, J.M. Epigenetic, cellular memory and gene regulation. Curr. Biol. 2016, 14, R644–R648. [Google Scholar] [CrossRef] [Green Version]
  219. To, T.K.; Kakutami, T. Crosstalk among pathways to generate DNA methylome. Curr. Opin. Plant Biol. 2022, 68, 102248. [Google Scholar] [CrossRef]
  220. Ptashne, M. Epigenetics: Core misconcept. Proc. Natl. Acad. Sci. USA 2013, 110, 7101–7103. [Google Scholar] [CrossRef] [Green Version]
  221. Ptashne, M. Faddish stuff: Epigenetics and the inheritance of acquired characteristics. FASEB J. 2013, 27, 1–2. [Google Scholar] [CrossRef] [Green Version]
  222. Veit, B. Stem cell signaling networks in plants. Plant Mol. Biol. 2006, 60, 793–810. [Google Scholar] [CrossRef]
  223. Sena, G.; Wang, X.; Liu, H.-Y.; Hofhuis, H.; Birnbaum, D. Organ regeneration does not require a functional stem cell niche in plants. Nature 2009, 457, 1150–1153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Choi, Y.E.; Soh, W.Y. Enhanced somatic single embryo formation by plasmolyzing pretreatment from cultures of ginseng cotyledons. Plant Sci. 1997, 130, 197–206. [Google Scholar] [CrossRef]
  225. Reyes, J.C. Chromatin modifiers that control plant development. Curr. Opin. Plant Biol. 2006, 9, 21–27. [Google Scholar] [CrossRef]
  226. Law, J.A.; Jacobsen, S.E. Molecular biology. Dynamic DNA methylation. Science 2009, 323, 1568–1569. [Google Scholar] [CrossRef] [PubMed]
  227. Loschiavo, F.; Pitto, L.; Giuliano, G.; Torti, G.; Nuti-Ronchi, V.; Marazziti, D.; Vergara, R.; Orselli, S.; Terzi, M. DNA methylation of embryogenic cell cultures and its variation as caused by mutation, differentiation, hormones and hypomethylating drugs. Theor. Appl. Genet. 1989, 77, 325–331. [Google Scholar] [CrossRef]
  228. Shibukava, T.; Yazawa, K.; Kikuchi, A.; Kamada, H. Possible involvement of DNA methylation on expression regulation of carrot LEC1 gene in its 50-upstream region. Gene 2009, 437, 22–31. [Google Scholar] [CrossRef] [PubMed]
  229. Ogas, J.; Cheng, J.-C.; Sung, Z.R.; Somerville, C. Cellular differentiation regulated by gibberellin in the Arabidopsis thaliana pickle mutant. Science 1997, 277, 91–94. [Google Scholar] [CrossRef] [PubMed]
  230. Ogas, J.; Kaufmann, S.; Henderson, J.; Somerville, C. PICKLE is a CHD3 chromatin-remodeling factor that regulates the transition from embryonic to vegetative development in Arabidopsis. Proc. Natl. Acad. Sci. USA 1999, 96, 13839–13844. [Google Scholar] [CrossRef] [Green Version]
  231. Henderson, J.T.; Li, H.-C.; Rider, S.D.; Mordhorst, A.P.; Romero-Severson, A.; Cheng, J.-C.; Robey, J.; Sung, Z.R.; de Vries, S.C.; Ogas, J. PICKLE acts throughout the plant to repress expression of embryonic traits and may play a role in gibberellin-dependent responses. Plant Physiol. 2004, 134, 995–1005. [Google Scholar] [CrossRef] [Green Version]
  232. Zhang, H.; Rider, S.D., Jr.; Henderson, J.T.; Fountain, M.; Chuang, K.; Kandachar, V.; Simons, A.; Edenberg, H.J.; Romero-Severson, J.; Muir, W.M.; et al. The CHD3 remodeler PICKLE promotes trimethylation of histone H3 lysine 27. J. Biol. Chem. 2008, 283, 22637–22648. [Google Scholar] [CrossRef] [Green Version]
  233. Aichinger, E.; Villar, C.B.R.; Farrona, S.; Reyes, J.C.; Henning, L.; Köhler, C. CHD3 proteins and polycomb group proteins antagonistically determine cell identity in Arabidopsis. PloS Genet. 2009, 5, e1000605. [Google Scholar] [CrossRef] [Green Version]
  234. Hennig, L.; Derkacheva, M. Diversity of Polycomb group complexes in plants: Same rules, different players? Trends Genet. 2009, 25, 414–423. [Google Scholar] [CrossRef]
  235. Berr, A.; Shafiq, S.; Shen, W.H. Histone modifications in transcriptional activation during plant development. Biochim. Biophys. Acta 2011, 1809, 567–576. [Google Scholar] [CrossRef] [PubMed]
  236. Derkacheva, M.; Hennig, L. Variations on a theme: Polycomb group proteins in plants. J. Exp. Bot. 2014, 65, 2769–2784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. de la Paz Sanchez, M.; Aceves-García, P.; Petrone, E.; Steckenborn, S.; Vega-León, R.; Álvarez-Buylla, E.R.; Garay-Arroyo, A.; García-Ponce, B. The impact of Polycomb group (PcG) and Trithorax group (TrxG) epigenetic factors in plant plasticity. New Phytol. 2015, 208, 684–694. [Google Scholar] [CrossRef]
  238. Mozgova, I.; Hennig, L. The Polycomb group protein regulatory network. Annu. Rev. Plant Biol. 2015, 66, 269–296. [Google Scholar] [CrossRef] [PubMed]
  239. Xu, F.; Kuo, T.; Rosli, T.; Liu, M.S.; Wu, L.; Chen, L.F.O.; Fletcher, J.C.; Sung, Z.R.; Pu, L. Trithorax group proteins act together with a polycomb group protein to maintain Trithorax chromatin integrity for epigenetic silencing during seed germination in Arabidopsis. Mol. Plants 2020, 11, 659–677. [Google Scholar] [CrossRef] [Green Version]
  240. Fischer, S.; Weber, L.M.; Liefke, R. Evolutionary adaptation of the Polycomb repressive complex 2. Epigenet. Chromatin 2022, 15, 7. [Google Scholar] [CrossRef]
  241. Schuettengruber, B.; Chourrout, D.; Vervoort, M.; Leblanc, B.; Cavalli, G. Genome regulation by polycomb and trithorax proteins. Cell 2007, 128, 735–745. [Google Scholar] [CrossRef] [Green Version]
  242. Chen, D.; Molitor, A.; Liu, C.; Shen, W.-S. The Arabidopsis PRC1-like ring-finger proteins are necessary for repression of embryonic traits during vegetative growth. Cell Res. 2010, 20, 1332–1344. [Google Scholar] [CrossRef] [Green Version]
  243. Bouyer, D.; Roudier, F.; Heese, M.; Anderson, E.D.; Gey, D.; Novack, M.K.; Goodrich, J.; Renou, J.-P.; Grini, P.E.; Colot, V.; et al. Polycomb repressive complex 2 controls the embryo-to-seedling phase transition. PLoS Genet. 2011, 7, e1002014. [Google Scholar] [CrossRef] [Green Version]
  244. Ikeuchi, M.; Iwase, A.; Rymen, B.; Harashima, H.; Shibata, M.; Ohnuma, M.; Breuer, C.; Morao, A.K.; de Lucas, M.; De Veylder, L.; et al. PRC2 represses dedifferentiation of mature somatic cells in Arabidopsis. Nat. Plants 2015, 1, 15089. [Google Scholar] [CrossRef] [Green Version]
  245. Mozgová, I.; Muñoz-Viana, R.; Hennig, L. PRC2 Represses hormone-induced somatic embryogenesis in vegetative tissue of Arabidopsis thaliana. PLoS Genet. 2017, 13, e1006562. [Google Scholar] [CrossRef] [Green Version]
  246. Guitton, A.E.; Berger, F. Loss of function of MULTICOPY SUPPRESSOR OF IRA 1 produces nonviable parthenogenetic embryos in Arabidopsis. Curr. Biol. 2005, 15, 750–754. [Google Scholar] [CrossRef]
  247. Schubert, D.; Clarenz, O.; Goodrich, J. Epigenetic control of plant development by Polycomb-group proteins. Curr. Opin. Plant Biol. 2005, 8, 553–561. [Google Scholar] [CrossRef] [PubMed]
  248. Baroux, C.; Pien, S.; Grossniklaus, U. Chromatin modification and remodeling during early seed development. Curr. Opin. Gen. Dev. 2007, 17, 473–479. [Google Scholar] [CrossRef] [PubMed]
  249. Pillot, M.; Baroux, C.; Vazquez, M.A.; Autran, D.; Leblanc, O.; Vielle-Calzada, J.P.; Grossniklaus, U.; Grimanelli, D. Embryo and endosperm inherit distinct chromatin and transcriptional states from the female gametes in Arabidopsis. Plant Cell 2010, 22, 307–320. [Google Scholar] [CrossRef] [Green Version]
  250. Iwasaki, M.; Paszkowski, J. Epigenetic memory in plants. EMBO J. 2014, 33, 1987–1998. [Google Scholar] [CrossRef] [Green Version]
  251. Kim, D.H.; Su, S. Polycomb-mediated gene silencing in Arabidopsis thaliana. Mol. Cells 2014, 37, 841–850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Steffen, P.A.; Ringrose, L. What are memories made of? How polycomb and trithorax proteins mediate epigenetic memory. Nat. Rev. Mol. Cell Biol. 2014, 15, 340–356. [Google Scholar] [CrossRef]
  253. Bertrand, C.; Bergounioux, C.; Domenichini, S.; Delarue, M.; Zhou, D.-X. Arabidopsis histone acetyltransferase AtGCN5 regulates the floral meristem activity through the WUSCHEL/AGAMOUS pathway. J. Biol. Chem. 2003, 278, 28246–28251. [Google Scholar] [CrossRef] [Green Version]
  254. Finley, A.; Copeland, R.A. Small molecule control of chromatin remodeling. Chem. Biol. 2014, 18, 1196–1210. [Google Scholar] [CrossRef] [Green Version]
  255. Wang, Z.; Cao, H.; Chen, F.; Liu, Y. The roles of histone acetylation in seed performance and plant development. Plant Physiol. Biochem. 2014, 84, 125–133. [Google Scholar] [CrossRef] [PubMed]
  256. Collings, C.K.; Waddell, P.J.; Anderson, J.N. Effects of DNA methylation on nucleosome stability. Nucleic Acids Res. 2013, 41, 2918–2931. [Google Scholar] [CrossRef] [Green Version]
  257. Rivera, C.; Gurard-Levin, Z.A.; Almouzni, G.; Loyola, A. Histone lysine methylation and chromatin replication. Biochim. Biophys. Acta 2014, 1839, 1433–1439. [Google Scholar] [CrossRef]
  258. Gallego-Bartolomé, J. DNA methylation in plants: Mechanisms and tools for targeted manipulation. New Phytol. 2020, 227, 38–44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Wu, S.C.; Zhang, Y. Active DNA demethylation: Many roads lead to Rome. Nat. Rev. Mol. Cell Biol. 2010, 11, 607–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Zhu, J.-K. Active DNA demethylation mediated by DNA glycosylases. Annu. Rev. Genet. 2009, 43, 143–166. [Google Scholar] [CrossRef] [Green Version]
  261. Le, T.N.; Miyazaki, Y.; Takuno, S.; Saze, H. Epigenetic regulation of intragenic transposable elements impacts gene transcription in Arabidopsis thaliana. Nucleic Acid Res. 2015, 43, 3911–3921. [Google Scholar] [CrossRef] [Green Version]
  262. Lisch, D. Epigenetic regulation of transposable elements in plants. Annu. Rev. Plant Biol. 2009, 60, 43–66. [Google Scholar] [CrossRef] [Green Version]
  263. Law, J.A.; Jacobsen, S.E. Establishing, maintaining and modifying DNA methylation patterns in plants and animals. Nat. Rev. Genet. 2010, 11, 204–220. [Google Scholar] [CrossRef]
  264. Stroud, H.; Greenberg, M.V.; Feng, S.; Bernatavichute, Y.V.; Jacobsen, S.E. Comprehensive analysis of silencing mutants reveals complex regulation of the Arabidopsis methylome. Cell 2013, 152, 352–364. [Google Scholar] [CrossRef] [Green Version]
  265. Zemach, A.; Kim, M.Y.; Hsieh, P.H.; Coleman-Derr, D.; Eshed-Williams, L.; Thao, K.; Harmer, S.L.; Zilberman, D. The Arabidopsis nucleosome remodeler DDM1 allows DNA methyltransferases to access H1-containing heterochromatin. Cell 2013, 153, 193–205. [Google Scholar] [CrossRef] [Green Version]
  266. Du, Z.; Li, H.; Wei, Q.; Zhao, X.; Wang, C.; Zhu, Q.; Yi, X.; Xu, W.; Liu, X.S.; Jin, W.; et al. Genome-wide analysis of histone modifications: H3K4me2, H3K4me3, H3K9ac, and H3K27ac in Oryza sativa L. Japonica. Mol. Plant 2013, 6, 1463–1472. [Google Scholar] [CrossRef] [Green Version]
  267. Stroud, H.; Do, T.; Du, J.; Zhong, X.; Feng, S.; Johnson, L.; Patel, D.J.; Jacobsen, S.E. Non-CG methylation patterns shape the epigenetic landscape in Arabidopsis. Nat. Struct. Mol. Biol. 2014, 21, 64–72. [Google Scholar] [CrossRef]
  268. Long, J.; Liu, J.; Xia, A.; Springer, N.M.; He, Y. Maize decrease in DNA methylation 1 targets RNA-directed DNA methylation on active chromatin. Plant Cell 2021, 33, 2183–2196. [Google Scholar] [CrossRef] [PubMed]
  269. Molitor, A.M.; Bu, Z.; Yu, Y.; Shen, W.H. Arabidopsis AL PHD-PRC1 complexes promote seed germination through H3K4me3-to-H3K27me3 chromatin state switch in repression of seed developmental genes. PLoS Genet. 2014, 10, e1004091. [Google Scholar] [CrossRef] [Green Version]
  270. Benhamed, M.; Martin-Magniette, M.-L.; Taconnat, L.; Bitton, F.; Servet, C.; De Clercq, R.; De Meyer, B.; Buysschaert, C.; Rombauts, S.; Villarroel, R.; et al. Genome-scale Arabidopsis promoter array identifies targets of the histone acetyltransferase GCN5. Plant J. 2008, 56, 493–504. [Google Scholar] [CrossRef] [Green Version]
  271. Tomasello, L.; Distefano, R.; Nigita, G.; Croce, C.M. The microRNA family gets wider: The IsomiRs classification and role. Front. Cell Dev. Biol. 2021, 9, 668648. [Google Scholar] [CrossRef]
  272. Jones-Rhoades, M.M.; Bartel, D.P.; Bartel, B. MicroRNAs and their regulatory roles in plants. Annu. Rev. Plant Biol. 2006, 57, 19–53. [Google Scholar] [CrossRef] [PubMed]
  273. Mallanna, A.K.; Rizzino, A. Emerging roles of microRNAs in the control of embryonic stem cells and the generation of induced pluripotent stem cells. Dev. Biol. 2010, 344, 16–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Wójcik, A.M.; Gaj, M.D. miR393 contributes to the embryogenic transition induced in vitro in Arabidopsis via the modification of the tissue sensitivity to auxin treatment. Planta 2016, 244, 231–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Wójcik, A.M.; Nodine, M.D.; Gaj, M.D. MiR160 and miR166/165 contribute to the LEC2-mediated auxin response involved in the somatic embryogenesis induction in Arabidopsis. Front. Plant Sci. 2017, 8, 2024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Long, J.; Liu, C.; Feng, M.; Liu, Y.; Wu, X.; Guo, W. miR156-SPL module regulates somatic embryogenesis induction in citrus callus. J. Exp. Bot. 2018, 69, 2979–2993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Siddiqui, Z.H.; Abbas, Z.H.; Ansari, M.W.; Khan, M.N. The role of miRNA in somatic embryogenesis. Genomics 2019, 111, 1026–1033. [Google Scholar] [CrossRef]
  278. Szyrajew, K.; Bielewicz, D.; Dolata, J.; WóJcik, A.M.; Novak, K.; Szczygieł-Sommer, A.; Szweykowska-Kulinska, Z.; Jarmolowski, A.; Gaj, M.D. MicroRNAs are intensively regulated during induction of somatic embryogenesis in Arabidopsis. Front. Plant Sci. 2017, 8, 18. [Google Scholar] [CrossRef] [Green Version]
  279. Zhang, L.; Xiang, Y.; Chen, S.; Shi, M.; Jiang, X.; He, Z.; Gao, S. Mechanisms of microRNA biogenesis and stability control in plants. Front. Plant Sci. 2022, 13, 844149. [Google Scholar] [CrossRef] [PubMed]
  280. Qiao, M.; Zhao, Z.; Song, Y.; Liu, Z.; Cao, L.; Yu, Y.; Li, S.; Xiang, F. Proper regeneration from in vitro cultured Arabidopsis thaliana requires the microRNA-directed action of an auxin response factor. Plant J. 2012, 71, 14–22. [Google Scholar] [CrossRef]
  281. Zhang, S.; Liu, Y.; Yu, B. New insights into pri-miRNA processing and accumulation in plants. Wiley Interdiscip. Rev. RNA 2015, 6, 533–545. [Google Scholar] [CrossRef]
  282. Liu, Z.; Li, J.; Wang, L.; Li, Q.; Lu, Q.; Yu, Y.; Li, S.; Bai, M.-Y.; Hu, Y.; Xiang, F. Repression of callus initiation by the miRNA-directed interaction of auxin-cytokinin in Arabidopsis thaliana. Plant J. 2016, 87, 391–402. [Google Scholar] [CrossRef]
  283. Liu, P.P.; Montgomery, T.A.; Fahlgren, N.; Kasschau, K.D.; Nonogaki, H.; Carrington, J.C. Repression of AUXIN RESPONSE FACTOR10 by microRNA160 is critical for seed germination and post-germination stages. Plant J. 2007, 52, 133–146. [Google Scholar] [CrossRef]
  284. Liu, X.; Huang, J.; Wang, Y.; Khanna, K.; Xie, Z.; Owen, H.A.; Zhao, D. The role of floral organs in carpels, an Arabidopsis loss-of-function mutation in microRNA160a, in organogenesis and the mechanism regulating its expression. Plant J. 2010, 62, 416–428. [Google Scholar] [CrossRef] [PubMed]
  285. Su, Y.H.; Liu, Y.B.; Zhou, C.; Li, X.M.; Zhang, X.S. The microRNA167 controls somatic embryogenesis in Arabidopsis through regulating its target genes ARF6 and ARF8. Plant Cell Tissue Organ Cult. 2016, 124, 405–417. [Google Scholar] [CrossRef]
  286. Chen, C.J.; Liu, Q.; Zhang, Y.C.; Qu, L.H.; Chen, Y.Q.; Gautheret, D. Genome-wide discovery and analysis of microRNAs and other small RNAs from rice embryogenic callus. RNA Biol. 2011, 8, 538–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Chávez-Hernández, E.C.; Alejandri-Ramírez, N.D.; Juárez-González, V.T.; Dinkova, T.D. Maize miRNA and target regulation in response to hormone depletion and light exposure during somatic embryogenesis. Front. Plant Sci. 2015, 6, 555. [Google Scholar] [CrossRef] [PubMed]
  288. Mallory, A.C.; Vaucheret, H. ARGONAUTE 1 homeostasis invokes the coordinate action of the microRNA and siRNA pathways. EMBO Rep. 2009, 10, 521–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Mallory, A.C.; Vaucheret, H. Form, function, and regulation of ARGONAUTE proteins. Plant Cell 2010, 22, 3879–3889. [Google Scholar] [CrossRef] [Green Version]
  290. Xue, T.; Dai, X.; Wang, R.; Wang, J.; Liu, Z.; Xiang, F. ARGONAUTE10 inhibits in vitro shoot regeneration via repression of miR165/166 in Arabidopsis thaliana. Plant Cell Physiol. 2017, 58, 1789–1800. [Google Scholar] [CrossRef] [Green Version]
  291. Nakamura, M.; Batista, R.A.; Köhler, C.; Hennig, L. Polycomb Repressive Complex 2-mediated histone modification H3K27me3 is associated with embryogenic potential in Norway spruce. J. Exp. Bot. 2020, 71, 6366–6378. [Google Scholar] [CrossRef]
  292. Sivanesan, I.; Nayeem, S.; Venkidasamy, B.; Kuppuraj, S.P.; Rn, C.; Samynathan, R. Genetic and epigenetic modes of the regulation of somatic embryogenesis: A review. Biol. Futur. 2022, 73, 259–277. [Google Scholar] [CrossRef]
  293. Wang, J.; Tian, C.; Zhang, C.; Shi, B.; Cao, X.; Zhang, T.-Q.; Zhao, Z.; Wang, J.-W.; Jiao, Y. Cytokinin signaling activates WUSCHEL expression during axillary meristem initiation. Plant Cell 2017, 29, 1373–1387. [Google Scholar] [CrossRef] [Green Version]
  294. Garcês, H.M.P.; Koenig, D.; Townsley, B.T.; Kim, M.; Sinha, N.R. Truncation of LEAFY COTYLEDON1 protein is required for asexual reproduction in Kalanchoë daigremontiana. Plant Physiol. 2014, 156, 196–206. [Google Scholar] [CrossRef] [Green Version]
  295. Laura, M.; Borghi, C.; Regis, C.; Cassetti, A.; Allavena, A. Ectopic expression of Kxhkn5 in the viviparous species Kalanchoe × Houghtonii induces a novel pattern of epiphyll development. Transgenic Res. 2013, 22, 59–74. [Google Scholar] [CrossRef]
  296. Zhu, C.; Wang, L.; Liu, C.; Zeng, H.; Wang, H. Over-expression of KdSOC1 gene affected plantlet morphogenesis in Kalanchoe daigremontiana. Sci. Rep. 2017, 7, 5629. [Google Scholar] [CrossRef] [PubMed]
  297. McCready, K.; Spencer, V.; Jácome-Blásquez, F.; Burnett, J.; Viveros Sánchez, I.M.; Riches, Z.; Kim, M. TARGET of RAPAMYCIN is essential for asexual vegetative reproduction in Kalanchoë. Plant Physiol. 2022, 189, 248–263. [Google Scholar] [CrossRef] [PubMed]
  298. Jácome-Blásquez, F.; Ooi, J.P.; Zeef, L.; Kim, M. Comparative transcriptome analysis of two Kalanchoë species during plantlet formation. Plants 2022, 11, 1643. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Animals able to regenerate lost body organs in vivo: (A) Hydra vulgare, an aquatic invertebrate (©Warren Photographic, www.warrenphotographic.co.uk/05836-common-hydra-budding; accessed on 11 October 2022). (B) Mexican axolotl (Ambystoma mexicanum), an aquatic salamander (©Warren Photographic, www.warrenphotographic.co.uk/00108-albino-axolotl; accessed on 11 October 2022). Examples of plants with natural epiphylly: (C) Kalanchoe laetivirens. (D) Malaxis paludosa (Michael Dodd’s photo, Open University, Milton Keynes, UK). The white arrowheads indicate the propagules. (E) Hieracium alpinum, a species of Asteraceae with an apomictic triploid cytotype in vivo (Mauro Felicioli’s photo, www.actaplantarum.org/flora/flora_info.php?id=506593; accessed on 11 October 2022). (F) An unusual example of a plant with epiphyllous embryos induced in vitro: the interspecific hybrid EMB-2 (Helianthus annuus x H. tuberosus). The white arrowheads indicate the somatic embryos.
Figure 1. Animals able to regenerate lost body organs in vivo: (A) Hydra vulgare, an aquatic invertebrate (©Warren Photographic, www.warrenphotographic.co.uk/05836-common-hydra-budding; accessed on 11 October 2022). (B) Mexican axolotl (Ambystoma mexicanum), an aquatic salamander (©Warren Photographic, www.warrenphotographic.co.uk/00108-albino-axolotl; accessed on 11 October 2022). Examples of plants with natural epiphylly: (C) Kalanchoe laetivirens. (D) Malaxis paludosa (Michael Dodd’s photo, Open University, Milton Keynes, UK). The white arrowheads indicate the propagules. (E) Hieracium alpinum, a species of Asteraceae with an apomictic triploid cytotype in vivo (Mauro Felicioli’s photo, www.actaplantarum.org/flora/flora_info.php?id=506593; accessed on 11 October 2022). (F) An unusual example of a plant with epiphyllous embryos induced in vitro: the interspecific hybrid EMB-2 (Helianthus annuus x H. tuberosus). The white arrowheads indicate the somatic embryos.
Ijms 23 15950 g001
Figure 2. Analysis of WUSCHEL transcription factor from sunflower (HaWUS) (GenBank accession HE616565.1). (A). Amino acid sequence of HaWUS (293 amino acid residues). The homeodomain is in bold and green characters; the WUS-box motif (TLPLFPXX) is underlined, and the conserved amino acid residues are in bold character. The EAR-like motif is in light blue character, the conserved amino acid residues are in bold character, and the acidic region is in red character [84]. (B). Within the homeodomain, the helix-loop-helix-turn-helix (HLHTH) motifs are indicated. Red dots evidence completely conserved residues. (C). Modeling of the three-dimensional structure of HaWUS in the homeodomain. The predicted modeling was generated by GENO3D (http://geno3d-pbil.ibcp.fr; accessed on 10 January 2015) and displayed by Deep View/Swiss Pdb-viewer v.4.1 (http://www.expasy.org/spdbv/; accessed on 10 January 2015).
Figure 2. Analysis of WUSCHEL transcription factor from sunflower (HaWUS) (GenBank accession HE616565.1). (A). Amino acid sequence of HaWUS (293 amino acid residues). The homeodomain is in bold and green characters; the WUS-box motif (TLPLFPXX) is underlined, and the conserved amino acid residues are in bold character. The EAR-like motif is in light blue character, the conserved amino acid residues are in bold character, and the acidic region is in red character [84]. (B). Within the homeodomain, the helix-loop-helix-turn-helix (HLHTH) motifs are indicated. Red dots evidence completely conserved residues. (C). Modeling of the three-dimensional structure of HaWUS in the homeodomain. The predicted modeling was generated by GENO3D (http://geno3d-pbil.ibcp.fr; accessed on 10 January 2015) and displayed by Deep View/Swiss Pdb-viewer v.4.1 (http://www.expasy.org/spdbv/; accessed on 10 January 2015).
Ijms 23 15950 g002
Figure 3. In vitro regeneration of Lactuca sativa cv. ‘Romana’ on a regeneration medium with 2.2 μM of 6-benzylaminopurine and 0.54 μM of α-naphthalenacetic acid; time-dependent analysis of WUSCHEL2-LIKE (LsWUS2L) expression levels. (AC): Histological analysis of cotyledon explants cultivated in vitro. Explants sectioned after 4 [T4 (A)], 7 [T7 (B)], and 12 [T12 (C)] days of culture on the regeneration medium. (A) Longitudinal section of a callus area characterized by small and poorly vacuolated cells located in a marginal region (black arrowheads). (B). Longitudinal section of an explant with initial stages of adventitious structures (black arrowheads). (C). Longitudinal section of an explant with early-stage somatic embryos (se). Scale bars: 0.7 mm (A), 0.19 mm (B), 0.28 mm (C). Binocular microscope photos of cotyledon explants in regeneration. (D), (E) Explants after 7 days (T7) of culture on regeneration medium: regeneration events occurred with the appearance of round-shaped structures (rss, white arrowheads) characterized by a regular profile. (F) Explants with massive formation of well-organized regenerated plantlets (white arrowheads). Scale bars: 0.85 mm (D), 1.02 mm (E), 8.7 mm (F). (G) Time-course analysis of the expression of WUSCHEL2-LIKE (LsWUS2L) during in vitro regeneration from cotyledon explants of lettuce. The times of in vitro culture analyzed were: T0 (0 days), T4 (4 days), T7 (7 days), and T12 (12 days). The relative transcriptional values were calculated using T12 as the reference sample. Data are means ± standard deviation (SD) (n = 4). Different letters indicate statistically significant differences between genotypes (one-way ANOVA and Tukey’s test: p < 0.05) [69].
Figure 3. In vitro regeneration of Lactuca sativa cv. ‘Romana’ on a regeneration medium with 2.2 μM of 6-benzylaminopurine and 0.54 μM of α-naphthalenacetic acid; time-dependent analysis of WUSCHEL2-LIKE (LsWUS2L) expression levels. (AC): Histological analysis of cotyledon explants cultivated in vitro. Explants sectioned after 4 [T4 (A)], 7 [T7 (B)], and 12 [T12 (C)] days of culture on the regeneration medium. (A) Longitudinal section of a callus area characterized by small and poorly vacuolated cells located in a marginal region (black arrowheads). (B). Longitudinal section of an explant with initial stages of adventitious structures (black arrowheads). (C). Longitudinal section of an explant with early-stage somatic embryos (se). Scale bars: 0.7 mm (A), 0.19 mm (B), 0.28 mm (C). Binocular microscope photos of cotyledon explants in regeneration. (D), (E) Explants after 7 days (T7) of culture on regeneration medium: regeneration events occurred with the appearance of round-shaped structures (rss, white arrowheads) characterized by a regular profile. (F) Explants with massive formation of well-organized regenerated plantlets (white arrowheads). Scale bars: 0.85 mm (D), 1.02 mm (E), 8.7 mm (F). (G) Time-course analysis of the expression of WUSCHEL2-LIKE (LsWUS2L) during in vitro regeneration from cotyledon explants of lettuce. The times of in vitro culture analyzed were: T0 (0 days), T4 (4 days), T7 (7 days), and T12 (12 days). The relative transcriptional values were calculated using T12 as the reference sample. Data are means ± standard deviation (SD) (n = 4). Different letters indicate statistically significant differences between genotypes (one-way ANOVA and Tukey’s test: p < 0.05) [69].
Ijms 23 15950 g003
Figure 4. Schematic representation of some major roles of WOX genes in vivo and in vitro in plants. The diagram is simplified, and bibliographic references are partial.
Figure 4. Schematic representation of some major roles of WOX genes in vivo and in vitro in plants. The diagram is simplified, and bibliographic references are partial.
Ijms 23 15950 g004
Table 1. Involvement of WUS and WOX genes in somatic embryogenesis and organogenesis in vitro, observed in different herbaceous and tree plants.
Table 1. Involvement of WUS and WOX genes in somatic embryogenesis and organogenesis in vitro, observed in different herbaceous and tree plants.
GeneSpeciesMolecular Event AnalyzedProcess InvolvedReference
AtWUSArabidopsis thalianaOverexpressionPGR independent somatic embryo development[70]
AtWUSArabidopsis thalianaOverexpressionRegeneration in PGR-free medium[62]
AtWUS and AtWOX5Arabidopsis thalianaStart of expressionIn cells of callus, previous events of ectopic morphogenesis[158]
PgWOX2Picea glaucaStart of expressionDuring the early stage of somatic embryogenesis[168]
LdWOX2Larix deciduaStart of expressionDuring the early stage of somatic embryogenesis[169]
PpWOX2Pinus pinasterHigh expressionDuring somatic embryogenesis proliferation[170]
PpWOX2Pinus pinasterOverexpressionNegative effect on the maturation of somatic embryos[170]
VvWOX2 and VvWOX9Vitis viniferaStart of expressionDuring the early stage of somatic embryogenesis[58]
VvWOX3 and VvWOX11Vitis viniferaHigh expression levelsSpecific stages of somatic embryos (torpedo and cotyledonary)[58]
MtWOX11-likeMedicago truncatulaStart of expressionDuring the early stage of somatic embryogenesis[159]
MtWUS and
MtWOX5
Medigago truncatulaStart of expressionIn calli with different embryogenic competence[171]
MtWUSMedicago truncatulaExpression cytokinin-dependentDuring the induction of totipotent stem cells[172]
CjWUSCryptomeria japonicaUpregulation of expressionIn callus with a high rate of embryogenesis[173]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fambrini, M.; Usai, G.; Pugliesi, C. Induction of Somatic Embryogenesis in Plants: Different Players and Focus on WUSCHEL and WUS-RELATED HOMEOBOX (WOX) Transcription Factors. Int. J. Mol. Sci. 2022, 23, 15950. https://doi.org/10.3390/ijms232415950

AMA Style

Fambrini M, Usai G, Pugliesi C. Induction of Somatic Embryogenesis in Plants: Different Players and Focus on WUSCHEL and WUS-RELATED HOMEOBOX (WOX) Transcription Factors. International Journal of Molecular Sciences. 2022; 23(24):15950. https://doi.org/10.3390/ijms232415950

Chicago/Turabian Style

Fambrini, Marco, Gabriele Usai, and Claudio Pugliesi. 2022. "Induction of Somatic Embryogenesis in Plants: Different Players and Focus on WUSCHEL and WUS-RELATED HOMEOBOX (WOX) Transcription Factors" International Journal of Molecular Sciences 23, no. 24: 15950. https://doi.org/10.3390/ijms232415950

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

Article Metrics

Back to TopTop