Next Article in Journal
Genome-Wide Identification and Expression Analysis of 1-Aminocyclopropane-1-Carboxylate Synthase (ACS) Gene Family in Chenopodium quinoa
Previous Article in Journal
Phytochemical Profiles and In Vitro Immunomodulatory Activities of Extracts Obtained from Limonium gmelinii Using Different Extraction Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unlocking Nature’s Clock: CRISPR Technology in Flowering Time Engineering

by
Ashkan Hodaei
and
Stefaan P. O. Werbrouck
*
Laboratory for Applied In Vitro Plant Biotechnology, Department of Plants and Crops, Faculty of Bioscience Engineering, Ghent University, 9000 Ghent, Belgium
*
Author to whom correspondence should be addressed.
Plants 2023, 12(23), 4020; https://doi.org/10.3390/plants12234020
Submission received: 17 October 2023 / Revised: 24 November 2023 / Accepted: 27 November 2023 / Published: 29 November 2023

Abstract

:
Flowering is a crucial process in the life cycle of most plants as it is essential for the reproductive success and genetic diversity of the species. There are situations in which breeders want to expedite, delay, or prevent flowering, for example, to shorten or prolong vegetative growth, to prevent unwanted pollination, to reduce the risk of diseases or pests, or to modify the plant’s phenotypes. This review aims to provide an overview of the current state of knowledge to use CRISPR/Cas9, a powerful genome-editing technology to modify specific DNA sequences related to flowering induction. We discuss the underlying molecular mechanisms governing the regulation of the photoperiod, autonomous, vernalization, hormonal, sugar, aging, and temperature signal pathways regulating the flowering time. In addition, we are investigating the most effective strategies for nominating target genes. Furthermore, we have collected a dataset showing successful applications of CRISPR technology to accelerate flowering in several plant species from 2015 up to date. Finally, we explore the opportunities and challenges of using the potential of CRISPR technology in flowering time engineering.

1. Flowering Time Matters

The induction of flowering is a key process in the life cycle of mature angiosperms because it marks the transition from vegetative growth to reproductive development. During this process, shoot apical meristems (SAM) undergo several changes in their metabolism, morphology, and gene expression.
Natural selection exerts selective pressure on plants that exhibit synchronized flowering, eliminating individuals that deviate from this optimal timing. Early flowering carries risks, such as being susceptibility to damage from late frosts, and the insufficient availability of pollinators and other flowering individuals for effective pollination. Conversely, delayed flowering can lead to inhospitable conditions for seed maturation or dispersal, failure to complete the seed set before mortality from late season frost or drought, or the production of progeny in unfavorable growing environments [1,2,3,4,5].
In addition to natural selection, as part of the domestication process, early farmers made conscious or unconscious choices regarding the flowering time in light of the harvest time and yield practices. These decisions focused on strategically managing workloads over different time frames, optimizing the use of labor, and maximizing overall productivity [6,7]. Today, the flowering time is one of the main goals of breeding and genetic engineering. Longer vegetative growth leads to the development of robust vegetative organs, which in turn facilitates the development of strong reproductive organs. This, particularly in cereals such as rice, significantly increases the seed yield, quality, and nutrient accumulation [8]. In addition, precocious bolting, characterized by the early onset of flowering and seed production, generally causes some vegetables to become unusable. Addressing this problem has significant economic benefits, including extending the vegetative growing season and thus increasing yield [9,10]. In fruit trees, reducing the juvenile phase is interesting for the early production of a plantation. Breeding for late or early seasonal flowering will extend the harvest time. Alternate or biennial bearing prevents regular yields and should be avoided as it affects the yield quantity and quality [11,12,13,14]. In addition, the development of fast-flowering plant lines holds great potential for speed breeding researchers, who aim to minimize the overall duration of the breeding process, which includes various time-consuming stages of crossing, selection, and testing involved in generating new plant varieties, and can extend the timeline for developing a new variety to one or two decades [15,16].
Global climate change causes heat waves, extreme cold, changing temperatures and rainfall patterns, and disrupts the phenology of many plant species and the development of their pollinators, leading to anomalous fertility and a suboptimal fruit set [17,18,19,20,21,22,23,24,25,26]. A strategy could be to actively change the flowering time of existing elite cultivars.
The recent pioneering technique of CRISPR offers a revolutionary solution to the above challenges. To effectively select the genes to be edited, a solid understanding of the relevant pathways and molecular mechanisms is essential. In the following chapter, we will take a closer look at these mechanisms and pathways.

2. Molecular Mechanisms Regulating Flowering Time in Arabidopsis

A series of experiments beginning in 1865 led to the introduction of the florigen hypothesis, which postulated the existence of a substance that could be transferred from the leaf to the shoot and induce flowering. This hypothesis led to widespread research efforts to elucidate the nature and existence of florigen [27,28]. Florigen is currently considered synonymous with the flowering locus T (FT) gene, which comprises the FT mRNA, its protein, or both [28,29,30,31]. The study of FT presents challenges due to the presence of numerous paralogs, orthologs, and homologs that exhibit similar or antagonistic functional behaviors, as well as their close sequence similarity with other proteins and interactions with different pathways. For example, in the antagonistic interaction between FT and TERMINAL FLOWER 1 (TFL1), a single amino acid turns a repressor (TFL1) into an activator of flowering (FT) [32].
The generally accepted schematic of physiological pathways proposed by Corbesier and Coupland (2005) [33] to control flowering includes four pathways: the photoperiodic, autonomous, vernalization, and gibberellin pathways. However, an updated scheme based on data collected in the Flowering Interactive Database [FLOR-ID] [34] introduces three additional pathways: sugar, aging, and temperature. Here, we present a revised figure with a simplified overview in Figure 1. Below, we explain briefly key signals or each pathway:

2.1. Photoperiod Pathway

The transcription factor, known as CONSTANS (CO), plays a central role in this pathway. Its role is to enhance FT expression by forming a complex with NUCLEAR FACTORY Y subunits (NF-Ys), TGACG MOTIF- BINDING FACTOR 4 (TGA4), or ASYMMETRIC LEAVES 1 (AS1). The activation of this pathway is primarily controlled by the circadian clock, orchestrated by GIGANTEA (GI), FLOWERING BHLHs (FBHs), RED AND FAR-RED INSENSITIVE 2 (RFI2), REPRESSOR OF UV-B PHOTOMORPHOGENESIS 2 (RUP2), and CONSTITUTIVE PHOTOMORPHOGENIC 1 (COP1). In addition, non-circadian regulators such as the MULTICOPY SUPPRESSOR OF IRA1 (MSI1), DAY NEUTRAL FLOWERING (DNF), and MEDIATOR25 (MED25) also play a role in the initiation of this pathway. All these elements can either positively or negatively influence the expression of CO. CO activity is also modulated by PHYTOCHROME A (PHYA), which is more sensitive to far-red light, and PHYTOCHROME B (PHYB), which responds better to red light. While PHYA acts as a positive regulator, PHYB exerts a negative control on CO. CO is regulated in transcriptional and translational levels. On short days (SDs), although the mRNA levels are high, the translation into protein does not take place. This is because COP1 and PHYB act as negative regulators during the translation process. On the other hand, on long days (LDs), GI and PHYA stimulate the translation of CO mRNA. The presence of CO protein peaks during the last hours of daylight and decreases as darkness begins [35,36,37,38,39,40,41,42,43,44,45,46].

2.2. Autonomous Pathway

In the autonomous pathway, the FLOWERING LOCUS C (FLC) serves as the major player, forming a complex with the SHORT VEGETATIVE PHASE (SVP) to negatively regulate FT in leaves and the SUPRESSOR OF OVEREXPRESSION OF CO1 (SOC1) in SAM.
The autonomous pathway is the most crowded of the flowering pathways, involving approximately 115 genes, mostly multifunctional, that control diverse processes including the cell cycle and DNA replication, chromatin modification, transcriptional regulation, the control of protein stability, and the processing of mRNA and microRNA. Notably, FLOWERING CONTROL LOCUS A (FCA) induces a reduction in FLC mRNA levels, FLOWERING LOCUS D (FLD) represses FLC by facilitating histone H3 Lys-4 demethylation at the FLC site, and FLOWERING LOCUS KH (FLK) represses FLC through post-transcriptional modification. It is also important to note that biotic stress, which affects flowering, is also considered as a part of the autonomous pathway. For instance, a single mutation in CADMIUM SENSITIVE 2 (CAD2), which is responsible for the biotic stress response, shows a delayed flowering phenotype under long days in Arabidopsis [47,48,49].

2.3. Vernalization Pathway

The vernalization pathway has FLC as its central component, similar to the autonomous pathway. Before the cold phase, six protein complexes (FRIGIDA, COMPASS, RAD6-BRE1, RAF1, SWA1, and FACT) are more active and positively regulate FLC in the transcriptional level. After cold exposure, three different protein complexes (PRC2, PRC1-like, and HDAC) increase their role in repressing FLC by not allowing RNA Polymerase to attach to the FLC locus. The prolonged exposure to cold shifts more cells to suppress FLC, increasing FT signaling from the leaves to SAM, and increases SOC1 activity in the SAM and, therefore, activates the meristem identity genes (MIs) to facilitate the transition from the vegetative to the reproductive phase [50,51,52,53,54].

2.4. Hormonal Pathway

The hormonal pathway, also known as the gibberellin pathway, involves the activity of cytokinins and gibberellins. Cytokinin affects the TWIN SISTER OF FT (TSF), while GAs affect the regulation of FT. Cytokinin, in a putative mechanism, positively regulates the transcription of TSF both in leaves and in SAM. TSF can be transferred from the leaf to the SAM and forms a complex there with a transcription factor, FD, and acts directly on MIs, making this the only pathway that can affect flowering independently of FT and SOC1.
In addition, GA4 boosts GID proteins, resulting in reduced DELLA proteins, which subsequently downregulate FT in leaves and MIs in SAM. In essence, GA4 positively controls FT in leaves and MIs in SAM, but negatively regulates SOC1 via its interaction with SQUAMOSA PROMOTER BINDING PROTEIN-LIKE 9 (SPL9).
Gibberellin signaling affects flowering by interfering with established endogenous and environmental flowering pathways, as well as by interacting with various phytohormone signaling pathways. In addition, cytokinins (CKs), abscisic acid (ABA), jasmonic acid (JA), ethylene (ET), brassinosteroids (BRs), and auxin also interact with DELLA proteins and, therefore, affect MIs. In conclusion, although this pathway has been traditionally referred to as the “Gibberellic pathway”, it would be more accurate to call it the “Hormonal pathway” [55,56,57,58,59].

2.5. Sugar Pathway

Trehalose-6-phosphate (T6P) is the main actor of this pathway which positively regulates FT in leaves and SOC1 in SAM by affecting the MIR156-SPL9 pathway. This signaling molecule (T6P) is the product of URACIL-DIPHOSPHATE GLUCOSE (UDPG) and GLUCOSE 6-PHOSPHATE (G6P) catalyzed by TREHALOSE-6-PHOSPHATE SYNTHASE 1 (TPS1). It is highly dependent on photosynthesis and reaches its peak concentration just before darkness under LD conditions, gradually decreasing with the onset of darkness [60,61,62,63,64,65]. In addition to being affected by the sugar pathway, FT also has an effect on sugar transport by activating sugar transporters such as SWEET10 [66].

2.6. Aging Pathway

The aging pathway, delays flowering in the juvenile phase and promotes it in the adult phase, even in the absence of external triggers. In particular, the miR156, SPL transcription factors (SPLs), miR172, SCHLAFMUTZE (SMZ), and TARGET OF EARLY ACTIVATION TAGGED 1 (TOE1) are key players in this ageing pathway, driving the transition from the juvenile to adult phase and then promoting flowering. MiRNA156 is more abundant in young plants and decreases as they mature, while miRNA172 shows the opposite pattern, increasing with age. The SMZ and TOE1 exert a repressive influence on the FT gene in the leaves and APETALA1 (AP1) in the SAM, affecting the expression of MIs and, consequently, controlling the initiation of flowering [67,68,69,70,71,72,73,74].

2.7. Temperature Pathway

The ambient temperature pathway influences FT through various components. At high temperatures, PHYTOCHROME INTERACTING FACTOR 4 (PIF4) and FLOWERING CONTROL LOCUS A (FCA) enhance FT. Conversely, at low temperatures, the SHORT VEGETATIVE PHASE (SVP) negatively regulates FT and SOC1 through a complex with FLC (a product of the vernalization pathway) [75,76,77,78,79,80,81,82,83,84,85,86].

2.8. Interconnectedness between Pathways

It is important to note that these pathways interact with each other, and it is difficult to completely distinguish between them. For example, miR172 is influenced by SPLs (“aging pathway”), FCA (“temperature pathway”), and GI (“Photoperiodic pathway”), demonstrating the interconnectedness of these pathways. Or in another example, the production of T6P, which is considered as the key player of the sugar pathway, is obviously dependent on photosynthesis and, therefore, the photoperiodic pathway.

2.9. Transportation of FT

FT and its homologous counterparts, known as FT-likes (FTLs), are synthesized in leaf tissues via the above-mentioned biochemical pathways. These molecules are then transported by specific transporters to the apical meristems in the shoots. The process of FT migration, a 22 kD protein with 175 amino acids, can be divided into two steps: first, it must be exported from companion cells into sieve elements, and second, it must cross sieve elements to reach the SAM.
The involvement of the transmembrane region proteins FT-INTERACTING PROTEIN1 (FTIP1) and QUIRKY (QKY) in the initial export step has been identified. To reach the SAM, another protein called SODIUM POTASSIUM ROOTDEFECTIVE 1 (NaKR1) with a heavy-metal-associated (HMA) domain plays a crucial role. The activity of NaKR1 is positively regulated by the product of the photoperiodic pathway, CO, in the model plant Arabidopsis [30,87,88,89].
In rice, the long-distance transporter for FT is the tetratricopeptide repeat 075 (TPR075) protein, while FTIP1 and FTIP9 are involved in the translocation of FT from companion cells to sieve elements [90].
Taken together, the FTIPs appear to play a critical role in the short-range transport of FT. Given recent research suggesting the involvement of different proteins in the long-distance transport of FT, further investigation is needed to determine whether the ortholog of TPR075 has a similar function in dicots, whether the ortholog of NaKR1 plays a role in monocots, and whether other heavy-metal-associated domain proteins or tetratricopeptide repeat proteins may also play a role.
In summary, our understanding of florigen transport remains largely incomplete, highlighting the need for comprehensive studies in a wide range of plant species to develop a more comprehensive understanding.

3. Optimal Gene Targeting Strategy for Flowering Time Engineering

3.1. Altering Flowering Time

Identifying the most effective gene(s) for either delaying or accelerating the flowering process depends on variables such as the species, cultivar, and growth environment, such as the plant’s response to the day length. The first step is to identify the predominant pathway(s) involved. For example, in the case of cabbage, where vernalization is a key determinant, genes such as BraFLC2, BraFLC3, AGL19s, and AGL24s associated with the vernalization pathway have shown remarkable success in breeding trials [91,92].
However, as a comprehensive strategy, targeting key proteins that are central to the overall flowering mechanism is a viable option. In genes such as FT, TFL1, and SOC1, as detailed in Table 1, 20% of efforts were by knock outing these genes. FT affects the expression of approximately 3652 genes [93]. However, from a breeding perspective, it does not appear to pose a significant challenge to the breeding objectives, except for its impact on seed dormancy [94,95]. In addition, FT affects the expression of some sugar transporters, suggesting a potential influence on other sugar-related traits. However, there is a paucity of research investigating the relationship between FT and traits such as fruit flavor, which warrants further investigation. Putting all this together, the modulation of the FT function is unlikely to result in major abnormalities. Conversely, the manipulation of TFL1, a flowering repressor, could be challenging, as some studies suggest that TFL1 mutations can lead to the aberrant development of floral structures [95,96,97,98,99]. To achieve the precise control of flowering, it may be prudent to explore alternative regulators of the FT gene, taking into account the primary pathway in your specific plant that effects greater FT expression. Alternatively, the modification of FT(s)’s Cis-regulatory elements can be used to precisely control the timing and level of expression too.
In addition to TFL1, in Arabidopsis, AGAMOUS-LIKE 12 (AGL12) displays late flowering under LD, but also the short root phenotype [100]. The overexpression of C-REPEAT/DRE BINDING FACTOR 1 (CBF1) results in late flowering under LD, but also a reduction in the freezing tolerance [101]. CURVY 1 (CVY1) single mutant flowers early under SD; however, as a side effect, altered trichome development and an increased number of siliques were observed [102]. AGAMOUS-LIKE 19 (AGL19) single mutant shows slightly late flowering under a short day, but also reduced the sensibility to vernalization [103].
The concept of interfering with transporters has also become a subject of debate. First, it is likely that NaKR1 in Arabidopsis or TPR075 in rice are not the only transporters of the FT protein. In addition, FT mRNA has the ability to migrate as well. Furthermore, NaKR1 also serves as a transporter for a wide range of sugars [89]. Mutations in NaKR1 could lead to variations in the concentration of both primary and secondary metabolites (Figure 2).
From a holistic perspective, a breeding project focusing on flowering time requires us to consider the various interrelated pathways highlighted in this article, including signaling proteins, miRNAs, transcription regulators, transporters, etc. By recognizing the importance of these key components, we can gain a thorough understanding of the complex mechanisms that regulate flowering time and, therefore, successful breeding.
As many of the proteins involved in flowering have multiple functions and diverse effects, one strategy for regulating their activity is to place greater emphasis on manipulating cis-regulatory elements (Figure 3). This approach allows us not only to fine-tune the level of gene expression, but also to precisely control the timing and location of expression. This is exactly what happened during the domestication process. To illustrate, in Brassica napus, a mutation in the transcription factor BnFLC.A10 was key to successfully altering the timing of flowering [104]. Another example is found in maize, where mutations in the transcription factors CCT and Vgt1/Rap2.7 led to significant adjustments in the flowering time [105,106].
As recent advancements, in grapefruit plants, CsLOB1 knockout mutants by CRISPR showed a delayed flowering time compared to WT [107]. The technique of dCas9 SunTag actively generated DNA methylation at the FWA promoter region, resulting in promoter silencing and the early flowering phenotype in Arabidopsis [108]. In Arabidopsis, the MS2-p300 CRISPR/dCas9 system, which incorporates H3K27 acetyltransferase as an effector domain and is linked to a nuclear-targeted MS2, has been used to modify the promoter region of the Flowering locus T (FT) gene. This resulted in a substantial two-fold increase in H3K27 acetylation within the FT promoter, which subsequently led to a significant alteration in the flowering time [109].

3.2. Stop Flowering

The idea of non-flowering angiosperms seems unlikely from a botanical point of view. However, in horticulture, postponing the flowering process could lead to a state of non-flowering cultivars. The disruption of FT or its transporter alone is not sufficient to achieve this result. This is due to the probability that FT is not solely dependent on a single transporter, and also because FT is not the only signaling molecule transmitted from the leaves to the shoot apical meristem (SAM). In fact, there are at least three known molecules—TSF, T6P, and AGL17—that also transduce the signal. What is more, certain pathways in the SAM can function independently of the leaves, ultimately influencing the expression of MI genes and thus the development of floral structures.
Of particular note is the structural similarity between FT and TFL1, two proteins with opposing functions. By exploiting these similarities, we can change specific amino acids, such as substituting His-88 in TFL1 and Tyr-85 in FT [32]. This simple amino acid change, with only a single nucleotide difference between His and Tyr, can switch their roles between repressor and activator. This insight holds considerable promise for targeted CRISPR breeding programs aimed at manipulating the flowering time, for example, by strategically designing gRNA sequences to introduce such mutations using base or prime editing.

4. CRISPR-Mediated Modulation of Flowering Time in Literature

Since 2015, there has been a striking trend in the scientific literature regarding the manipulation of the flowering time using CRISPR technology. From then until now (mid-2023), there have been numerous publications investigating different gene families. These studies cover a wide range of plant species, with a particular focus on altering the flowering time. A thorough review was conducted of 103 peer-reviewed research publications, which are listed in Table 1. Table 1 provides a comprehensive overview of genes with diverse functionalities that contribute to successful flowering time engineering. These genes include transcriptional regulators, non-coding RNAs, enzymes, signaling molecules, chromatin modifiers, epigenetic modifiers, and many more (Figure 4).
Table 1. Non-exhaustive overview of studies of flowering time using CRISPR.103 peer-reviewed research papers listed by the modified plant. In the table, “KO” represents knockouts, while “OE” represents overexpression. The terms “early”, “late”, or “No effect” indicate whether the time of flowering is earlier or later, or no difference compared to the wild type.
Table 1. Non-exhaustive overview of studies of flowering time using CRISPR.103 peer-reviewed research papers listed by the modified plant. In the table, “KO” represents knockouts, while “OE” represents overexpression. The terms “early”, “late”, or “No effect” indicate whether the time of flowering is earlier or later, or no difference compared to the wild type.
Author(s)YearDirected Gene(s)MechanismPlantResults in Flowering
Torre et al. [110]2022AaFRAT1KOAlpine cressEarly
Charrier et al. [111]2019MdTFL1.1, PcTFL1.1KOApple, PearEarly
Liu et al. [112]2019TFL1, AP1, SVPKOArabidopsisAbnormal flower development
Ning et al. [113]2015NACsKOArabidopsisEarly
Nobusawa et al. [114]2022AMP1KOArabidopsisEarly
Capovilla et al. [115]2017FLM-β KOArabidopsisEarly
Branchat et al. [116]2020FDP, fdKOArabidopsisEarly, late
Lian et al. [117]2021MIR172sKO and OEArabidopsisEarly, late, no effect
Yan et al. [118]2017KHZ1 and KHZ2KO and OEArabidopsisLate and early
Hyun et al. [119]2015FT and SPLKOArabidopsisLate
Hou et al. [120]2019AtMIR396KOArabidopsisLate
Yao et al. [121]2019mir167aKOArabidopsisLate
Huang et al. [122]2019OsNCED5OEArabidopsisLate
Wang et al. [123]2021RBP45DKO ArabidopsisLate
Zhao et al. [124]2022CIS1KOArabidopsisLate
Yang et al. [125]2023AtAGL79KOArabidopsisLate
Pyott et al. [126]2016eIF(iso)4EKOArabidopsisNo effect
Soto et al. [127]2022FT2KOAspenNo report
Qin et al. [128]2019FTL9KOBrachypodiumLate
Jeong et al. [92]2019BraFLC2, BraFLC3 KOCabbageEarly
Jung et al. [129]2021BrSOC1KO CabbageEarly
Hong et al. [130]2021BrVRN1KO CabbageLate
Shin et al. [131]2023BrLFYKOCabbageLate
Lee et al. [132]2023BrFT1 and BrFT2KOCabbageLate
Shin et al. [91]2022AGL19s, AGL24sKOCabbageLate
Park et al. [133]2019GIKO CabbageNo report
Bellec et al. [134]202215 genesKOCamelinaEarly
Jiang et al. [135]2018BnaSDG8.A, BnaSDG8.CKOCanolaEarly
Sriboon et al. [99]2020BnaC03.TFL1KOCanolaEarly
Guo et al. [136]2022BnaCOL9KOCanolaEarly
Ahmar et al. [137]2022BnaSVPsKOCanolaEarly
Zhou et al. [138]2022BnaSVP, BnaSEP1KOCanolaEarly, no effect
Odipio et al. [139]2018TFL1-likeKOCassavaEarly
Bull et al. [140]2018AtFTEctopic expressionCassavaEarly
Liu et al. [141]2023CiTFL1a, CiTFL1bKOChrysantemumEarly
Huang et al. [142]2017ZmCCT9KOCornEarly
Li et al. [143]2020ZmPHYC1, ZmPHYC2KOCornEarly
Takahashi et al. [144]2022GtFT2KOGentianLate
Ying et al. [145]2022BdRFSKO and OEBrachypodiumEarly, late
Sheng et al. [146]2021YSL3KOBrachypodium Late
Herath et al. [147]2022AcBFT2KOKiwiNo effect
Gasic et al. [148]2019AcCEN4, AcCENKOKiwiEarly
Varkonyi et al. [149]2021CEN, CEN4, SyGlKOKiwiEarly
Choi et al. [10]2022SOC1KOLettuceLate
Singer et al. [150]2021MsSPL8KOLucerneEarly
Galindo-Sotomonte et al. [151]2023MSAD_264347KO LucerneLate
Wolabu et al. [152]2023MsFTa1KO LucerneLate
Shibuya et al. [153]2018EPHEMERAL1KOMorning GloryDelay in petal aging
Andre et al. [154]2022FT2bOEPopulusEarly
Elorriaga et al. [155]2018PLFY, PAG1, PAG2KOPopulusNo report
Lebedeva et al. [156]2022StLFYKOPotatoLate, non-flowering
Li et al. [157]2017Hd2, Hd4, Hd5KO RiceEarly
Brambilla et al. [158]2017hbf1, hbf2KORiceEarly
Zhou et al. [159]2018Ghd8KORiceEarly
Cui et al. [160]2019se14KORiceEarly
Wang et al. [161]2020OsGhd7KORiceEarly
Karthika et al. [162]2021MSH2KORiceEarly
Leon et al. [163]2021OsGA20ox2KORiceEarly
Sun et al. [164]2022qHD5KORiceEarly
Yin et al. [165]2023HBP1KORiceEarly
Sedeek et al. [166]2023Hd2, Hd4, Hd5KORiceEarly
Guo et al. [167]2022OsFTL4KORiceEarly
Sun et al. [168]2021OsLHYKORiceEarly, late
Zhang et al. [169]2020OsCCTsKORiceEarly, late, no effect
Cui Y et al. [170]202114 genesKORiceEarly, late, no effect
Yasui et al. [171]2017MADS3KORiceEarly/Late
Wu et al. [172]2020Ehd1KORiceLate
Li et al. [173]2021OsLHYKORiceLate
Liu et al. [174]2021OsHd2KORiceLate
Zhang et al. [175]2022ga3ox-2KORiceLate
Xu et al. [176]2023OsLUXKORiceLate
Zhang et al. [177]2022ospil12-1 and ospil12-2KORiceLate
Andrade et al. [178]2022LUX, ELF3KORiceNon-flowering
Dai et al. [179]2021HbFT1-2, HbTFL1-3KO Rubber TreeNo report
Wang et al. [180]2022SiPHYCKOSetariaEarly
Zhu et al. [181]2022spp1 KOSetariaNo effect
Char et al. [182]2019SbFTKOSorghumLate
Han et al. [183]2019E1KOSoyEarly
Wang et al. [184]2020Gmprr37KOSoyEarly
Wang et al. [185]2020GmNMHC5OESoyEarly
Zhaobo Li et al. [186]2021LNK2KOSoyEarly
Wan et al. [187]2022E1KOSoyEarly
Zhai et al. [188]2022E1KO and OESoyEarly, late
Cai et al. [189]2018GmFT2aKOSoyLate
Wang et al. [190]2019GmLCLa1-4KOSoyLate
Cong Li et al. [191]2020GmPRR3bH6KOSoyLate
Chen et al. [192]2020GmAP1KOSoyLate
Zhao et al. A [193]2022GmPHYAsKOSoyLate
Schmidt et al. [194]2020NtFT5KOTobaccoNon-flowering
Soyk et al. [195]2017SP5GKOTomatoEarly
Lemmon et al. [196]2018SP5GKOTomatoEarly
Li et al. [197]2018SP and SP5GKOTomatoEarly
Hu et al. [198]2022SlDOF9sKOTomatoEarly
Moreira et al. [199]2022SP3C KO and OETomatoEarly, late
Xu et al. [200]2016S1BOPKOTomatoNo effect
Lin et al. [201]2021SlMIR172c, SlMIR172d KOTomatoNo report
Kwon et al. [202]2019SP5G, SP, SlERKOTomatoEarly
Gupta et al. [203]2022TaSPL13KOWheatEarly
Sun et al. [204]2023TaTFL1-5KOWheatEarly
Chen et al. [205]2022FT-D1KOWheatLate
Errum et al. [206]2023TaPpdKOWheatLate
Huiyun Liu et al. [207]2020TaAQ and TaDqKOWheatNo report
The visual representation in Figure 5A shows the percentage distribution of plant species to which CRISPR technology has been applied. This distribution is likely influenced by the ease of transformation and use of model organisms. It is noteworthy that this approach is highly promising and can be extrapolated to various other commercial crops. Altering the flowering time provides a versatile means of addressing a wide range of cultivation, harvest, and post-harvest challenges.
The use of CRISPR technology for gene knockout, particularly in flowering projects, has seen sustained growth (Figure 5B). Further breakthroughs are expected as researchers refine and expand their understanding, leading to more sophisticated applications in the future.

5. Perspectives and Challenges of CRISPR-Mediated Flowering Time Engineering

Flowering-related breeding projects sometimes aim to completely inhibit flowering, while in other cases, a delay or advance of days to weeks is desired. To achieve these goals, the selection of the right pathway and protein to engineer is critical. In addition, the location of the mutation on the gene and the type of mutation affects the flowering time, which can vary from species to species based on the genotype or environmental factors.
About 300 genes are directly or indirectly involved in the flowering process [31]. To obtain the desired breeding results, it might be necessary to test several genes or gene combinations. However, the limited capacity of generating gene-edited plants slows down breeding. This slow pace hinders the exploration of many candidate genes as potential targets. In this regard, the need to increase the efficiency of transformation, regeneration, and gene editing processes is, therefore, essential to save time and explore a larger gene pool.
CRISPR technology is a leading innovation in plant breeding, offering precision, speed, and flexibility. The rapid generation of plants with desired traits surpasses the time limits of traditional breeding methods, while the scalability of CRISPR allows multiple target genes to be edited in parallel. However, to fully achieve CRISPR’s potential, it is increasingly recognized that it must be strategically combined with complementary techniques.
In this context, using inventive techniques and approaches to increase efficiency and save time is critical to the success of molecular breeding projects. Recent advances include improvements in delivery using improved Agrobacterium strains [208,209], viral delivery [210], improvements in regeneration using morphogenic regulators [211,212], and new approaches to use multiplex gene editing to study many genes simultaneously [213,214].
Through such comprehensive and interconnected approaches, breeders can gain new insights into the intricate web of regulatory networks that regulate the flowering time. This new knowledge can guide targeted breeding initiatives aimed at producing crop varieties with improved flowering traits.

Author Contributions

The conceptualization and design of the review was a collaborative effort. A.H. conducted a comprehensive literature search and analysis, synthesized key findings and trends, and took the lead in drafting the manuscript. S.P.O.W. provided essential insights that enhanced the depth and scope of the review. Both authors actively participated in the editing of the manuscript and approved the final version submitted for publication. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

I would like to express my sincere gratitude to Laurens Pauwels for his invaluable contributions to the transformation gene editing section of this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gaudinier, A.; Blackman, B.K. Evolutionary processes from the perspective of flowering time diversity. New Phytol. 2020, 225, 1883–1898. [Google Scholar] [CrossRef]
  2. Anderson, J.T.; Willis, J.H.; Mitchell-Olds, T. Evolutionary genetics of plant adaptation. Trends Genet. 2011, 27, 258–266. [Google Scholar] [CrossRef]
  3. Thomson, J.D. Flowering phenology, fruiting success and progressive deterioration of pollination in an early-flowering geophyte. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2010, 365, 3187–3199. [Google Scholar] [CrossRef]
  4. Kehrberger, S.; Holzschuh, A. How does timing of flowering affect competition for pollinators, flower visitation and seed set in an early spring grassland plant? Sci. Rep. 2019, 9, 15593. [Google Scholar] [CrossRef]
  5. Osnato, M. Evolution of flowering time genes in rice: From the paleolithic to the anthropocene. Plant Cell Environ. 2023, 46, 1046–1059. [Google Scholar] [CrossRef]
  6. Cockram, J.; Jones, H.; Leigh, F.J.; O’Sullivan, D.; Powell, W.; Laurie, D.A.; Greenland, A.J. Control of flowering time in temperate cereals: Genes, domestication, and sustainable productivity. J. Exp. Bot. 2007, 58, 1231–1244. [Google Scholar] [CrossRef]
  7. Jung, C.; Müller, A.E. Flowering time control and applications in plant breeding. Trends Plant Sci. 2009, 14, 563–573. [Google Scholar] [CrossRef]
  8. Wang, J.; Wu, F.; Zhu, S.; Xu, Y.; Cheng, Z.; Wang, J.; Li, C.; Sheng, P.; Zhang, H.; Cai, M.; et al. Overexpression of OsMYB1R1–VP64 fusion protein increases grain yield in rice by delaying flowering time. FEBS Lett. 2016, 590, 3385–3396. [Google Scholar] [CrossRef]
  9. Chen, C.; Huang, W.; Hou, K.; Wu, W. Bolting, an Important Process in Plant Development, Two Types in Plants. J. Plant Biol. 2019, 62, 161–169. [Google Scholar] [CrossRef]
  10. Choi, S.H.; Ahn, W.S.; Jie, E.Y.; Cho, H.-S.; Kim, S.W. Development of late-bolting plants by CRISPR/Cas9-mediated genome editing from mesophyll protoplasts of lettuce. Plant Cell Rep. 2022, 41, 1627–1630. [Google Scholar] [CrossRef]
  11. Tominaga, A.; Ito, A.; Sugiura, T.; Yamane, H. How Is Global Warming Affecting Fruit Tree Blooming? “Flowering (Dormancy) Disorder” in Japanese Pear (Pyrus pyrifolia) as a Case Study. Front. Plant Sci. 2021, 12, 787638. [Google Scholar] [CrossRef] [PubMed]
  12. Fadón, E.; Herrera, S.; Guerrero, B.I.; Guerra, M.E.; Rodrigo, J. Chilling and Heat Requirements of Temperate Stone Fruit Trees (Prunus sp.). Agronomy 2020, 10, 409. [Google Scholar] [CrossRef]
  13. Dong, X.; Jiang, X.; Kuang, G.; Wang, Q.; Zhong, M.; Jin, D.; Hu, J. Genetic control of flowering time in woody plants: Roses as an emerging model. Plant Divers 2017, 39, 104–110. [Google Scholar] [CrossRef]
  14. Haberman, A.; Bakhshian, O.; Cerezo-Medina, S.; Paltiel, J.; Adler, C.; Ben-Ari, G.; Mercado, J.A.; Pliego-Alfaro, F.; Lavee, S.; Samach, A. A possible role for flowering locus T-encoding genes in interpreting environmental and internal cues affecting olive (Olea europaea L.) flower induction. Plant Cell Environ. 2017, 40, 1263–1280. [Google Scholar] [CrossRef]
  15. Ahmar, S.; Gill, R.A.; Jung, K.-H.; Faheem, A.; Qasim, M.U.; Mubeen, M.; Zhou, W. Conventional and Molecular Techniques from Simple Breeding to Speed Breeding in Crop Plants: Recent Advances and Future Outlook. Int. J. Mol. Sci. 2020, 21, 2590. [Google Scholar] [CrossRef]
  16. Abdul Fiyaz, R.; Ajay, B.C.; Ramya, K.T.; Aravind Kumar, J.; Sundaram, R.M.; Subba Rao, L.V. Speed Breeding: Methods and Applications. In Accelerated Plant Breeding; Gosal, S., Wani, S., Eds.; Springer: Cham, Switzerland, 2020; Volume 1. [Google Scholar] [CrossRef]
  17. Rafferty, N.E.; Diez, J.M.; Bertelsen, C.D. Changing Climate Drives Divergent and Nonlinear Shifts in Flowering Phenology across Elevations. Curr. Biol. 2020, 30, 432–441.e3. [Google Scholar] [CrossRef]
  18. Anderson, J.T.; Inouye, D.W.; McKinney, A.M.; Colautti, R.I.; Mitchell-Olds, T. Phenotypic plasticity and adaptive evolution contribute to advancing flowering phenology in response to climate change. Proc. R. Soc. B Biol. Sci. 2012, 279, 3843–3852. [Google Scholar] [CrossRef]
  19. Siegmund, J.F.; Wiedermann, M.; Donges, J.F.; Donner, R.V. Impact of temperature and precipitation extremes on the flowering dates of four German wildlife shrub species. Biogeosciences 2016, 13, 5541–5555. [Google Scholar] [CrossRef]
  20. Mohandass, D.; Zhao, J.-L.; Xia, Y.-M.; Campbell, M.J.; Li, Q.-J. Increasing temperature causes flowering onset time changes of alpine ginger Roscoea in the Central Himalayas. J. Asia-Pac. Biodivers 2015, 8, 191–198. [Google Scholar] [CrossRef]
  21. Lambert, A.M.; Miller-Rushing, A.J.; Inouye, D.W. Changes in snowmelt date and summer precipitation affect the flowering phenology of Erythronium grandiflorum (glacier lily; Liliaceae). Am. J. Bot. 2010, 97, 1431–1437. [Google Scholar] [CrossRef]
  22. Menzel, A.; Sparks, T.H.; Estrella, N.; Koch, E.; Aasa, A.; Ahas, R.; Alm-Kübler, K.; Bissolli, P.; Braslavská, O.; Briede, A.; et al. European phenological response to climate change matches the warming pattern. Glob. Chang. Biol. 2006, 12, 1969–1976. [Google Scholar] [CrossRef]
  23. Craufurd, P.Q.; Wheeler, T.R. Climate change and the flowering time of annual crops. J. Exp. Bot. 2009, 60, 2529–2539. [Google Scholar] [CrossRef]
  24. Hu, Q.; Weiss, A.; Feng, S.; Baenziger, P.S. Earlier winter wheat heading dates and warmer spring in the U.S. Great Plains. Agric. For. Meteorol. 2005, 135, 284–290. [Google Scholar] [CrossRef]
  25. Drogoudi, P.; Kazantzis, K.; Blanke, M. Climate change effects on cherry flowering in northern Greece. Acta Hortic. 2017, 1162, 45–50. [Google Scholar] [CrossRef]
  26. Fitter, A.H.; Fitter, R.S.R. Rapid Changes in Flowering Time in British Plants. Science 2002, 296, 1689–1691. [Google Scholar] [CrossRef]
  27. Garner, W.W.; Allard, H.A. Effect of the relative length of day and night and other environmental factors on the growth and reproduction of plants. Mon. Wea. Rev. 1920, 48, 415. [Google Scholar] [CrossRef]
  28. Zeevaart, J.A. Florigen Coming of Age after 70 Years. Plant Cell 2006, 18, 1783–1789. [Google Scholar] [CrossRef] [PubMed]
  29. Huang, T.; Böhlenius, H.; Eriksson, S.; Parcy, F.; Nilsson, O. The mRNA of the Arabidopsis gene FT moves from leaf to shoot apex and induces flowering. Science 2005, 309, 1694–1696. [Google Scholar] [CrossRef]
  30. Corbesier, L.; Vincent, C.; Jang, S.; Fornara, F.; Fan, Q.; Searle, I.; Giakountis, A.; Farrona, S.; Gissot, L.; Turnbull, C.; et al. FT protein movement contributes to long-distance signaling in floral induction of Arabidopsis. Science 2007, 316, 1030–1033. [Google Scholar] [CrossRef]
  31. Yu, Z.; Chen, W.; Wang, Y.; Zhang, P.; Shi, N.; Hong, Y. Mobile Flowering Locus T RNA—Biological Relevance and Biotechnological Potential. Front. Plant Sci. 2022, 12, 792192. [Google Scholar] [CrossRef] [PubMed]
  32. Hanzawa, Y.; Money, T.; Bradley, D. A single amino acid converts a repressor to an activator of flowering. Proc. Natl. Acad. Sci. USA 2005, 102, 7748–7753. [Google Scholar] [CrossRef] [PubMed]
  33. Corbesier, L.; Coupland, G. Photoperiodic flowering of Arabidopsis: Integrating genetic and physiological approaches to characterization of the floral stimulus. Plant Cell Environ. 2004, 28, 54–66. [Google Scholar] [CrossRef]
  34. Bouché, F.; Lobet, G.; Tocquin, P.; Périlleux, C. FLOR-ID: An interactive database of flowering-time gene networks in Arabidopsis thaliana. Nucleic Acids Res. 2016, 44, D1167–D1171. [Google Scholar] [CrossRef] [PubMed]
  35. Takagi, H.; Hempton, A.K.; Imaizumi, T. Photoperiodic flowering in Arabidopsis: Multilayered regulatory mechanisms of CONSTANS and the florigen FLOWERING LOCUS T. Plant Commun. 2023, 4, 100552. [Google Scholar] [CrossRef]
  36. Wenkel, S.; Turck, F.; Singer, K.; Gissot, L.; Le Gourrierec, J.; Samach, A.; Coupland, G. CONSTANS and the CCAAT box binding complex share a functionally important domain and interact to regulate flowering of Arabidopsis. Plant Cell 2006, 18, 2971–2984. [Google Scholar] [CrossRef]
  37. Shim, J.S.; Kubota, A.; Imaizumi, T. Circadian Clock and Photoperiodic Flowering in Arabidopsis: CONSTANS Is a Hub for Signal Integration. Plant Physiol. 2017, 173, 5–15. [Google Scholar] [CrossRef] [PubMed]
  38. Ito, S.; Song, Y.H.; Josephson-Day, A.R.; Miller, R.J.; Breton, G.; Olmstead, R.G.; Imaizumi, T. FLOWERING BHLH transcriptional activators control expression of the photoperiodic flowering regulator CONSTANS in Arabidopsis. Proc. Natl. Acad. Sci. USA 2012, 109, 3582–3587. [Google Scholar] [CrossRef]
  39. Mockler, T.; Yang, H.; Yu, X.; Parikh, D.; Cheng, Y.-C.; Dolan, S.; Lin, C. Regulation of photoperiodic flowering by Arabidopsis photoreceptors. Proc. Natl. Acad. Sci. USA 2003, 100, 2140–2145. [Google Scholar] [CrossRef] [PubMed]
  40. Leivar, P.; Monte, E.; Al-Sady, B.; Carle, C.; Storer, A.; Alonso, J.M.; Ecker, J.R.; Quail, P.H. The Arabidopsis phytochrome-interacting factor PIF7, together with PIF3 and PIF4, regulates responses to prolonged red light by modulating phyB levels. Plant Cell 2008, 20, 337–352. [Google Scholar] [CrossRef]
  41. Sawa, M.; Kay, S.A. GIGANTEA directly activates Flowering Locus T in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2011, 108, 11698–11703. [Google Scholar] [CrossRef]
  42. Brandoli, C.; Petri, C.; Egea-Cortines, M.; Weiss, J. Gigantea: Uncovering New Functions in Flower Development. Genes 2020, 11, 1142. [Google Scholar] [CrossRef]
  43. Rival, P.; Press, M.O.; Bale, J.; Grancharova, T.; Undurraga, S.F.; Queitsch, C. The conserved PFT1 tandem repeat is crucial for proper flowering in Arabidopsis thaliana. Genetics 2014, 198, 747–754. [Google Scholar] [CrossRef]
  44. Steinbach, Y.; Hennig, L. Arabidopsis MSI1 functions in photoperiodic flowering time control. Front. Plant Sci. 2014, 5, 77. [Google Scholar] [CrossRef]
  45. Morris, K.; Jackson, S.P. DAY NEUTRAL FLOWERING does not act through GIGANTEA and FKF1 to regulate CONSTANS expression and flowering time. Plant Signal Behav. 2010, 5, 1105–1107. [Google Scholar] [CrossRef]
  46. Valverde, F. CONSTANS and the evolutionary origin of photoperiodic timing of flowering. J. Exp. Bot. 2011, 62, 2453–2463. [Google Scholar] [CrossRef]
  47. Ogawa, K.I.; Tasaka, Y.; Mino, M.; Tanaka, Y.; Iwabuchi, M. Association of Glutathione with Flowering in Arabidopsis thaliana. Plant Cell Physiol. 2001, 42, 524–530. [Google Scholar] [CrossRef]
  48. Kyung, J.; Jeon, M.; Lee, I. Recent advances in the chromatin-based mechanism of FLOWERING LOCUS C repression through autonomous pathway genes. Front. Plant Sci. 2022, 13, 964931. [Google Scholar] [CrossRef] [PubMed]
  49. Cheng, J.-Z.; Zhou, Y.-P.; Lv, T.-X.; Xie, C.-P.; Tian, C.-E. Research progress on the autonomous flowering time pathway in Arabidopsis. Physiol. Mol. Biol. Plants 2017, 23, 477–485. [Google Scholar] [CrossRef] [PubMed]
  50. Kim, D.-H. Current understanding of flowering pathways in plants: Focusing on the vernalization pathway in Arabidopsis and several vegetable crop plants. Hortic. Environ. Biotechnol. 2020, 61, 209–227. [Google Scholar] [CrossRef]
  51. Duncan, S.; Holm, S.; Questa, J.; Irwin, J.; Grant, A.; Dean, C. Seasonal shift in timing of vernalization as an adaptation to extreme winter. eLife 2015, 4, e06620. [Google Scholar] [CrossRef]
  52. Surkova, S.Y.; Samsonova, M.G. Mechanisms of Vernalization-Induced Flowering in Legumes. Int. J. Mol. Sci. 2022, 23, 9889. [Google Scholar] [CrossRef]
  53. Luo, X.; He, Y. Experiencing winter for spring flowering: A molecular epigenetic perspective on vernalization. J. Integr. Plant Biol. 2020, 62, 104–117. [Google Scholar] [CrossRef] [PubMed]
  54. Preston, J.C.; Fjellheim, S. Flowering time runs hot and cold. Plant Physiol. 2022, 190, 5–18. [Google Scholar] [CrossRef] [PubMed]
  55. Bao, S.; Hua, C.; Shen, L.; Yu, H. New insights into gibberellin signaling in regulating flowering in Arabidopsis. J. Integr. Plant Biol. 2020, 62, 118–131. [Google Scholar] [CrossRef] [PubMed]
  56. Zhao, L.; Li, X.; Chen, W.; Xu, Z.; Chen, M.; Wang, H.; Yu, D. The emerging role of jasmonate in the control of flowering time. J. Exp. Bot. 2022, 73, 11–21. [Google Scholar] [CrossRef]
  57. Izawa, T. What is going on with the hormonal control of flowering in plants? Plant J. 2021, 105, 431–445. [Google Scholar] [CrossRef] [PubMed]
  58. Davière, J.-M.; Achard, P. Gibberellin signaling in plants. Development 2013, 140, 1147–1151. [Google Scholar] [CrossRef] [PubMed]
  59. Conti, L. Hormonal control of the floral transition: Can one catch them all? Dev. Biol. 2017, 430, 288–301. [Google Scholar] [CrossRef]
  60. Cho, L.-H.; Pasriga, R.; Yoon, J.; Jeon, J.-S.; An, G. Roles of Sugars in Controlling Flowering Time. J. Plant Biol. 2018, 61, 121–130. [Google Scholar] [CrossRef]
  61. Moghaddam, M.R.B.; Ende, W.V.D. Sugars, the clock and transition to flowering. Front. Plant Sci. 2013, 4, 22. [Google Scholar] [CrossRef]
  62. Roldán, M.; Gómez-Mena, C.; Ruiz-García, L.; Salinas, J.; Martínez-Zapater, J.M. Sucrose availability on the aerial part of the plant promotes morphogenesis and flowering of Arabidopsis in the dark. Plant J. 1999, 20, 581–590. [Google Scholar] [CrossRef]
  63. El-Lithy, M.E.; Reymond, M.; Stich, B.; Koornneef, M.; Vreugdenhil, D. Relation among plant growth, carbohydrates and flowering time in the Arabidopsis Landsberg erecta × Kondara recombinant inbred line population. Plant Cell Environ. 2010, 33, 1369–1382. [Google Scholar] [CrossRef] [PubMed]
  64. Wahl, V.; Ponnu, J.; Schlereth, A.; Arrivault, S.; Langenecker, T.; Franke, A.; Feil, R.; Lunn, J.E.; Stitt, M.; Schmid, M. Regulation of flowering by trehalose-6-phosphate signaling in Arabidopsis thaliana. Science 2013, 339, 704–707. [Google Scholar] [CrossRef] [PubMed]
  65. Ohto, M.-A.; Onai, K.; Furukawa, Y.; Aoki, E.; Araki, T.; Nakamura, K. Effects of Sugar on Vegetative Development and Floral Transition in Arabidopsis. Plant Physiol. 2001, 127, 252–261. [Google Scholar] [CrossRef] [PubMed]
  66. Andrés, F.; Kinoshita, A.; Kalluri, N.; Fernández, V.; Falavigna, V.S.; Cruz, T.M.D.; Jang, S.; Chiba, Y.; Seo, M.; Mettler-Altmann, T.; et al. The sugar transporter SWEET10 acts downstream of FLOWERING LOCUS T during floral transition of Arabidopsis thaliana. BMC Plant Biol. 2020, 20, 53. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, J.-W.; Schwab, R.; Czech, B.; Mica, E.; Weigel, D. Dual Effects of miR156-Targeted SPL Genes and CYP78A5/KLUH on Plastochron Length and Organ Size in Arabidopsis thaliana. Plant Cell 2008, 20, 1231–1243. [Google Scholar] [CrossRef]
  68. Cui, L.; Shan, J.; Shi, M.; Gao, J.; Lin, H. The miR156-SPL9-DFR pathway coordinates the relationship between development and abiotic stress tolerance in plants. Plant J. 2014, 80, 1108–1117. [Google Scholar] [CrossRef] [PubMed]
  69. Shikata, M.; Koyama, T.; Mitsuda, N.; Ohme-Takagi, M. Arabidopsis SBP-box genes SPL10, SPL11 and SPL2 control morphological change in association with shoot maturation in the reproductive phase. Plant Cell Physiol. 2009, 50, 2133–2145. [Google Scholar] [CrossRef]
  70. Schwab, R.; Palatnik, J.F.; Riester, M.; Schommer, C.; Schmid, M.; Weigel, D. Specific effects of microRNAs on the plant transcriptome. Dev. Cell 2005, 8, 517–527. [Google Scholar] [CrossRef]
  71. Gandikota, M.; Birkenbihl, R.P.; Höhmann, S.; Cardon, G.H.; Saedler, H.; Huijser, P. The miRNA156/157 recognition element in the 3′ UTR of the Arabidopsis SBP box gene SPL3 prevents early flowering by translational inhibition in seedlings. Plant J. 2007, 49, 683–693. [Google Scholar] [CrossRef]
  72. Stief, A.; Altmann, S.; Hoffmann, K.; Pant, B.D.; Scheible, W.-R.; Bäurle, I. Arabidopsis miR156 Regulates Tolerance to Recurring Environmental Stress through SPL Transcription Factors. Plant Cell 2014, 26, 1792–1807. [Google Scholar] [CrossRef] [PubMed]
  73. Yu, S.; Cao, L.; Zhou, C.-M.; Zhang, T.-Q.; Lian, H.; Sun, Y.; Wu, J.; Huang, J.; Wang, G.; Wang, J.-W. Sugar is an endogenous cue for juvenile-to-adult phase transition in plants. eLife 2013, 2, e00269. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, J.-W. Regulation of flowering time by the miR156-mediated age pathway. J. Exp. Bot. 2014, 65, 4723–4730. [Google Scholar] [CrossRef] [PubMed]
  75. Franklin, K.A.; Lee, S.H.; Patel, D.; Kumar, S.V.; Spartz, A.K.; Gu, C.; Ye, S.; Yu, P.; Breen, G.; Cohen, J.D.; et al. Phytochrome-interacting factor 4 (PIF4) regulates auxin biosynthesis at high temperature. Proc. Natl. Acad. Sci. USA 2011, 108, 20231–20235. [Google Scholar] [CrossRef] [PubMed]
  76. Oh, E.; Zhu, J.-Y.; Wang, Z.-Y. Interaction between BZR1 and PIF4 integrates brassinosteroid and environmental responses. Nat. Cell Biol. 2012, 14, 802–809. [Google Scholar] [CrossRef]
  77. Thines, B.C.; Youn, Y.; Duarte, M.I.; Harmon, F.G. The time of day effects of warm temperature on flowering time involve PIF4 and PIF5. J. Exp. Bot. 2014, 65, 1141–1151. [Google Scholar] [CrossRef]
  78. Bäurle, I.; Smith, L.; Baulcombe, D.C.; Dean, C. Widespread role for the flowering-time regulators FCA and FPA in RNA-mediated chromatin silencing. Science 2007, 318, 109–112. [Google Scholar] [CrossRef]
  79. Sonmez, C.; Bäurle, I.; Magusin, A.; Dreos, R.; Laubinger, S.; Weigel, D.; Dean, C. RNA 3′ processing functions of Arabidopsis FCA and FPA limit intergenic transcription. Proc. Natl. Acad. Sci. USA 2011, 108, 8508–8513. [Google Scholar] [CrossRef]
  80. Jung, J.-H.; Seo, P.J.; Ahn, J.H.; Park, C.-M. Arabidopsis RNA-binding protein FCA regulates microRNA172 processing in thermosensory flowering. J. Biol. Chem. 2012, 287, 16007–16016. [Google Scholar] [CrossRef]
  81. Lee, H.-J.; Jung, J.-H.; Llorca, L.C.; Kim, S.-G.; Lee, S.; Baldwin, I.T.; Park, C.-M. FCA mediates thermal adaptation of stem growth by attenuating auxin action in Arabidopsis. Nat. Commun. 2014, 5, 5473. [Google Scholar] [CrossRef]
  82. Mateos, J.L.; Madrigal, P.; Tsuda, K.; Rawat, V.; Richter, R.; Romera-Branchat, M.; Fornara, F.; Schneeberger, K.; Krajewski, P.; Coupland, G. Combinatorial activities of SHORT VEGETATIVE PHASE and FLOWERING LOCUS C define distinct modes of flowering regulation in Arabidopsis. Genome Biol. 2015, 16, 31. [Google Scholar] [CrossRef] [PubMed]
  83. Andrés, F.; Porri, A.; Torti, S.; Mateos, J.; Romera-Branchat, M.; García-Martínez, J.L.; Fornara, F.; Gregis, V.; Kater, M.M.; Coupland, G. SHORT VEGETATIVE PHASE reduces gibberellin biosynthesis at the Arabidopsis shoot apex to regulate the floral transition. Proc. Natl. Acad. Sci. USA 2014, 111, E2760–E2769. [Google Scholar] [CrossRef]
  84. Capovilla, G.; Schmid, M.; Posé, D. Control of flowering by ambient temperature. J. Exp. Bot. 2015, 66, 59–69. [Google Scholar] [CrossRef] [PubMed]
  85. Mathieu, J.; Yant, L.J.; Mürdter, F.; Küttner, F.; Schmid, M. Repression of Flowering by the miR172 Target SMZ. PLoS Biol. 2009, 7, e1000148. [Google Scholar] [CrossRef] [PubMed]
  86. Posé, D.; Verhage, L.; Ott, F.; Yant, L.; Mathieu, J.; Angenent, G.C.; Immink, R.G.H.; Schmid, M. Temperature-dependent regulation of flowering by antagonistic FLM variants. Nature 2013, 503, 414–417. [Google Scholar] [CrossRef] [PubMed]
  87. Liu, L.; Liu, C.; Hou, X.; Xi, W.; Shen, L.; Tao, Z.; Wang, Y.; Yu, H. FTIP1 is an essential regulator required for florigen transport. PLoS Biol. 2012, 10, e1001313. [Google Scholar] [CrossRef]
  88. Zhu, Y.; Liu, L.; Shen, L.; Yu, H. NaKR1 regulates long-distance movement of FLOWERING LOCUS T in Arabidopsis. Nat. Plants 2016, 2, 16075. [Google Scholar] [CrossRef]
  89. Tian, H.; Baxter, I.R.; Lahner, B.; Reinders, A.; Salt, D.E.; Ward, J.M. Arabidopsis NPCC6/NaKR1 Is a phloem mobile metal binding protein necessary for phloem function and root meristem maintenance. Plant Cell 2010, 22, 3963–3979. [Google Scholar] [CrossRef]
  90. Zhang, L.; Zhang, F.; Zhou, X.; Poh, T.X.; Xie, L.; Shen, J.; Yang, L.; Song, S.; Yu, H.; Chen, Y. The tetratricopeptide repeat protein OsTPR075 promotes heading by regulating florigen transport in rice. Plant Cell 2022, 34, 3632–3646. [Google Scholar] [CrossRef]
  91. Shin, Y.-H.; Lee, H.-M.; Park, Y.-D. CRISPR/Cas9-Mediated Editing of AGAMOUS-like Genes Results in a Late-Bolting Phenotype in Chinese Cabbage (Brassica rapa ssp. pekinensis). Int. J. Mol. Sci. 2022, 23, 15009. [Google Scholar] [CrossRef]
  92. Jeong, S.Y.; Ahn, H.; Ryu, J.; Oh, Y.; Sivanandhan, G.; Won, K.-H.; Park, Y.D.; Kim, J.-S.; Kim, H.; Lim, Y.P.; et al. Generation of early-flowering Chinese cabbage (Brassica rapa spp. pekinensis) through CRISPR/Cas9-mediated genome editing. Plant Biotechnol. Rep. 2019, 13, 491–499. [Google Scholar] [CrossRef]
  93. Duplat-Bermúdez, L.; Ruiz-Medrano, R.; Landsman, D.; Mariño-Ramírez, L.; Xoconostle-Cázares, B. Transcriptomic analysis of Arabidopsis overexpressing flowering locus T driven by a meristem-specific promoter that induces early flowering. Gene 2016, 587, 120–131. [Google Scholar] [CrossRef]
  94. Chen, M.; Penfield, S. Feedback regulation of COOLAIR expression controls seed dormancy and flowering time. Science 2018, 360, 1014–1017. [Google Scholar] [CrossRef]
  95. Auge, G.A.; Penfield, S.; Donohue, K. Pleiotropy in developmental regulation by flowering-pathway genes: Is it an evolutionary constraint? New Phytol. 2019, 224, 55–70. [Google Scholar] [CrossRef]
  96. Li, C.; Fu, Q.; Niu, L.; Luo, L.; Chen, J.; Xu, Z.-F. Three TFL1 homologues regulate floral initiation in the biofuel plant Jatropha curcas. Sci. Rep. 2017, 7, srep43090. [Google Scholar] [CrossRef]
  97. Zhang, S.; Hu, W.; Wang, L.; Lin, C.; Cong, B.; Sun, C.; Luo, D. TFL1/CEN-like genes control intercalary meristem activity and phase transition in rice. Plant Sci. 2005, 168, 1393–1408. [Google Scholar] [CrossRef]
  98. Gao, Y.; Gao, Y.; Wu, Z.; Bu, X.; Fan, M.; Zhang, Q. Characterization of TEMINAL FLOWER1 homologs CmTFL1c gene from Chrysanthemum morifolium. Plant Mol. Biol. 2019, 99, 587–601. [Google Scholar] [CrossRef]
  99. Sriboon, S.; Li, H.; Guo, C.; Senkhamwong, T.; Dai, C.; Liu, K. Knock-out of TERMINAL FLOWER 1 genes altered flowering time and plant architecture in Brassica napus. BMC Genet. 2020, 21, 52. [Google Scholar] [CrossRef] [PubMed]
  100. Tapia-López, R.; García-Ponce, B.; Dubrovsky, J.G.; Garay-Arroyo, A.; Pérez-Ruíz, R.V.; Kim, S.-H.; Acevedo, F.; Pelaz, S.; Alvarez-Buylla, E.R. An AGAMOUS-Related MADS-Box Gene, XAL1 (AGL12), regulates root meristem cell proliferation and flowering transition in Arabidopsis. Plant Physiol. 2008, 146, 1182–1192. [Google Scholar] [CrossRef] [PubMed]
  101. Novillo, F.; Medina, J.; Salinas, J. Arabidopsis CBF1 and CBF3 have a different function than CBF2 in cold acclimation and define different gene classes in the CBF regulon. Proc. Natl. Acad. Sci. USA 2007, 104, 21002–21007. [Google Scholar] [CrossRef] [PubMed]
  102. Gachomo, E.W.; Baptiste, L.J.; Kefela, T.; Saidel, W.M.; Kotchoni, S.O. The Arabidopsis CURVY1 (CVY1) gene encoding a novel receptor-like protein kinase regulates cell morphogenesis, flowering time and seed production. BMC Plant Biol. 2014, 14, 221. [Google Scholar] [CrossRef]
  103. Schönrock, N.; Bouveret, R.; Leroy, O.; Borghi, L.; Köhler, C.; Gruissem, W.; Hennig, L. Polycomb-group proteins repressthe floral activator AGL19 in the FLC-independent vernalization pathway. Minerva Anestesiol. 2006, 20, 1667–1678. [Google Scholar] [CrossRef] [PubMed]
  104. Hou, J.; Long, Y.; Raman, H.; Zou, X.; Wang, J.; Dai, S.; Xiao, Q.; Li, C.; Fan, L.; Liu, B.; et al. A Tourist-like MITE insertion in the upstream region of the BnFLC.A10 gene is associated with vernalization requirement in rapeseed (Brassica napus L.). BMC Plant Biol. 2012, 12, 238. [Google Scholar] [CrossRef] [PubMed]
  105. Yang, Q.; Li, Z.; Li, W.; Ku, L.; Wang, C.; Ye, J.; Li, K.; Yang, N.; Li, Y.; Zhong, T.; et al. CACTA-like transposable element in ZmCCT attenuated photoperiod sensitivity and accelerated the postdomestication spread of maize. Proc. Natl. Acad. Sci. USA 2013, 110, 16969–16974. [Google Scholar] [CrossRef] [PubMed]
  106. Salvi, S.; Sponza, G.; Morgante, M.; Tomes, D.; Niu, X.; Fengler, K.A.; Meeley, R.; Ananiev, E.V.; Svitashev, S.; Bruggemann, E.; et al. Conserved noncoding genomic sequences associated with a flowering-time quantitative trait locus in maize. Proc. Natl. Acad. Sci. USA 2007, 104, 11376–11381. [Google Scholar] [CrossRef] [PubMed]
  107. Jia, H.; Zhang, Y.; Orbović, V.; Xu, J.; White, F.F.; Jones, J.B.; Wang, N. Genome editing of the disease susceptibility gene CsLOB1 in citrus confers resistance to citrus canker. Plant Biotechnol. J. 2017, 15, 817–823. [Google Scholar] [CrossRef] [PubMed]
  108. Papikian, A.; Liu, W.; Gallego-Bartolomé, J.; Jacobsen, S.E. Site-specific manipulation of Arabidopsis loci using CRISPR-Cas9 SunTag systems. Nat. Commun. 2019, 10, 729. [Google Scholar] [CrossRef] [PubMed]
  109. Lee, J.E.; Neumann, M.; Duro, D.I.; Schmid, M. CRISPR-based tools for targeted transcriptional and epigenetic regulation in plants. PLoS ONE 2019, 14, e0222778. [Google Scholar] [CrossRef] [PubMed]
  110. de la Torre, N.V.; Vayssières, A.; Obeng-Hinneh, E.; Neumann, U.; Zhou, Y.; Lázaro, A.; Roggen, A.; Sun, H.; Stolze, S.C.; Nakagami, H.; et al. FLOWERING REPRESSOR AAA+ ATPase 1 is a novel regulator of perennial flowering in Arabis alpina. New Phytol. 2022, 236, 729–744. [Google Scholar] [CrossRef]
  111. Charrier, A.; Vergne, E.; Dousset, N.; Richer, A.; Petiteau, A.; Chevreau, E. Efficient Targeted Mutagenesis in Apple and First Time Edition of Pear Using the CRISPR-Cas9 System. Front. Plant Sci. 2019, 10, 40. [Google Scholar] [CrossRef]
  112. Liu, Y.; Gao, Y.; Gao, Y.; Zhang, Q. Targeted deletion of floral development genes in Arabidopsis with CRISPR/Cas9 using the RNA endoribonuclease Csy4 processing system. Hortic. Res. 2019, 6, 99. [Google Scholar] [CrossRef]
  113. Ning, Y.-Q.; Ma, Z.-Y.; Huang, H.-W.; Mo, H.; Zhao, T.-T.; Li, L.; Cai, T.; Chen, S.; Ma, L.; He, X.-J. Two novel NAC transcription factors regulate gene expression and flowering time by associating with the histone demethylase JMJ14. Nucleic Acids Res. 2015, 43, 1469–1484. [Google Scholar] [CrossRef]
  114. Nobusawa, T.; Yamatani, H.; Kusaba, M. Early flowering phenotype of the Arabidopsis altered meristem program1 mutant is dependent on the FLOWERING LOCUS T-mediated pathway. Plant Biotechnol. 2022, 39, 317–321. [Google Scholar] [CrossRef]
  115. Capovilla, G.; Symeonidi, E.; Wu, R.; Schmid, M. Contribution of major FLM isoforms to temperature-dependent flowering in Arabidopsis thaliana. J. Exp. Bot. 2017, 68, 5117–5127. [Google Scholar] [CrossRef]
  116. Romera-Branchat, M.; Severing, E.; Pocard, C.; Ohr, H.; Vincent, C.; Née, G.; Martinez-Gallegos, R.; Jang, S.; Andrés, F.; Madrigal, P.; et al. Functional Divergence of the Arabidopsis Florigen-Interacting bZIP Transcription Factors FD and FDP. Cell Rep. 2020, 31, 107717. [Google Scholar] [CrossRef]
  117. Lian, H.; Wang, L.; Ma, N.; Zhou, C.-M.; Han, L.; Zhang, T.-Q.; Wang, J.-W. Redundant and specific roles of individual MIR172 genes in plant development. PLoS Biol. 2021, 19, e3001044. [Google Scholar] [CrossRef]
  118. Yan, Z.; Jia, J.; Yan, X.; Shi, H.; Han, Y. Arabidopsis KHZ1 and KHZ2, two novel non-tandem CCCH zinc-finger and K-homolog domain proteins, have redundant roles in the regulation of flowering and senescence. Plant Mol. Biol. 2017, 95, 549–565. [Google Scholar] [CrossRef]
  119. Hyun, Y.; Kim, J.; Cho, S.W.; Choi, Y.; Kim, J.-S.; Coupland, G. Site-directed mutagenesis in Arabidopsis thaliana using dividing tissue-targeted RGEN of the CRISPR/Cas system to generate heritable null alleles. Planta 2015, 241, 271–284. [Google Scholar] [CrossRef]
  120. Hou, N.; Cao, Y.; Li, F.; Yuan, W.; Bian, H.; Wang, J.; Zhu, M.; Han, N. Epigenetic regulation of miR396 expression by SWR1-C and the effect of miR396 on leaf growth and developmental phase transition in Arabidopsis. J. Exp. Bot. 2019, 70, 5217–5229. [Google Scholar] [CrossRef]
  121. Yao, X.; Chen, J.; Zhou, J.; Yu, H.; Ge, C.; Zhang, M.; Gao, X.; Dai, X.; Yang, Z.-N.; Zhao, Y. An Essential Role for miRNA167 in Maternal Control of Embryonic and Seed Development. Plant Physiol. 2019, 180, 453–464. [Google Scholar] [CrossRef]
  122. Huang, Y.; Jiao, Y.; Xie, N.; Guo, Y.; Zhang, F.; Xiang, Z.; Wang, R.; Wang, F.; Gao, Q.; Tian, L.; et al. OsNCED5, a 9-cis-epoxycarotenoid dioxygenase gene, regulates salt and water stress tolerance and leaf senescence in rice. Plant Sci. 2019, 287, 110188. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, L.; Xu, D.; Scharf, K.; Frank, W.; Leister, D.; Kleine, T. The RNA-binding protein RBP45D of Arabidopsis promotes transgene silencing and flowering time. Plant J. 2022, 109, 1397–1415. [Google Scholar] [CrossRef] [PubMed]
  124. Zhao, Z.; Dent, C.; Liang, H.; Lv, J.; Shang, G.; Liu, Y.; Feng, F.; Wang, F.; Pang, J.; Li, X.; et al. CRY2 interacts with CIS1 to regulate thermosensory flowering via FLM alternative splicing. Nat. Commun. 2022, 13, 7045. [Google Scholar] [CrossRef] [PubMed]
  125. Yang, H.; Zhang, P.; Guo, D.; Wang, N.; Lin, H.; Wang, X.; Niu, L. Transcriptional repressor AGL79 positively regulates flowering time in Arabidopsis. J. Plant Physiol. 2023, 285, 153985. [Google Scholar] [CrossRef] [PubMed]
  126. Pyott, D.E.; Sheehan, E.; Molnar, A. Engineering of CRISPR/Cas9-mediated potyvirus resistance in transgene-free Arabidopsis plants. Mol. Plant Pathol. 2016, 17, 1276–1288. [Google Scholar] [CrossRef] [PubMed]
  127. Gómez-Soto, D.; Allona, I.; Perales, M. FLOWERING LOCUS T2 Promotes Shoot Apex Development and Restricts Internode Elongation via the 13-Hydroxylation Gibberellin Biosynthesis Pathway in Poplar. Front. Plant Sci. 2022, 12, 814195. [Google Scholar] [CrossRef] [PubMed]
  128. Qin, Z.; Bai, Y.; Muhammad, S.; Wu, X.; Deng, P.; Wu, J.; An, H.; Wu, L. Divergent roles of FT-like 9 in flowering transition under different day lengths in Brachypodium distachyon. Nat. Commun. 2019, 10, 812. [Google Scholar] [CrossRef]
  129. Jung, H.; Lee, A.; Jo, S.H.; Park, H.J.; Jung, W.Y.; Kim, H.-S.; Lee, H.-J.; Jeong, S.-G.; Kim, Y.-S.; Cho, H.S. Nitrogen Signaling Genes and SOC1 Determine the Flowering Time in a Reciprocal Negative Feedback Loop in Chinese Cabbage (Brassica rapa L.) Based on CRISPR/Cas9-Mediated Mutagenesis of Multiple BrSOC1 Homologs. Int. J. Mol. Sci. 2021, 22, 4631. [Google Scholar] [CrossRef]
  130. Hong, J.K.; Suh, E.J.; Park, S.R.; Park, J.; Lee, Y.-H. Multiplex CRISPR/Cas9 Mutagenesis of BrVRN1 Delays Flowering Time in Chinese Cabbage (Brassica rapa L. ssp. pekinensis). Agriculture 2021, 11, 1286. [Google Scholar] [CrossRef]
  131. Shin, Y.-H.; Park, Y.-D. CRISPR/Cas9-Mediated Mutagenesis of BrLEAFY Delays the Bolting Time in Chinese Cabbage (Brassica rapa L. ssp. pekinensis). Int. J. Mol. Sci. 2022, 24, 541. [Google Scholar] [CrossRef]
  132. Lee, A.; Jung, H.; Park, H.J.; Jo, S.H.; Jung, M.; Kim, Y.-S.; Cho, H.S. Their C-termini divide Brassica rapa FT-like proteins into FD-interacting and FD-independent proteins that have different effects on the floral transition. Front. Plant Sci. 2023, 13, 1091563. [Google Scholar] [CrossRef]
  133. Park, S.-C.; Park, S.; Jeong, Y.J.; Lee, S.B.; Pyun, J.W.; Kim, S.; Kim, T.H.; Kim, S.W.; Jeong, J.C.; Kim, C.Y. DNA-free mutagenesis of GIGANTEA in Brassica oleracea var. capitata using CRISPR/Cas9 ribonucleoprotein complexes. Plant Biotechnol. Rep. 2019, 13, 483–489. [Google Scholar] [CrossRef]
  134. Bellec, Y.; Guyon-Debast, A.; François, T.; Gissot, L.; Biot, E.; Nogué, F.; Faure, J.-D.; Tepfer, M. New Flowering and Architecture Traits Mediated by Multiplex CRISPR-Cas9 Gene Editing in Hexaploid Camelina sativa. Agronomy 2022, 12, 1873. [Google Scholar] [CrossRef]
  135. Jiang, L.; Li, D.; Jin, L.; Ruan, Y.; Shen, W.-H.; Liu, C. Histone lysine methyltransferases BnaSDG8.A and BnaSDG8.C are involved in the floral transition in Brassica napus. Plant J. 2018, 95, 672–685. [Google Scholar] [CrossRef] [PubMed]
  136. Guo, J.; Zeng, L.; Chen, H.; Ma, C.; Tu, J.; Shen, J.; Wen, J.; Fu, T.; Yi, B. CRISPR/Cas9-Mediated Targeted Mutagenesis of BnaCOL9 Advances the Flowering Time of Brassica napus L. Int. J. Mol. Sci. 2022, 23, 14944. [Google Scholar] [CrossRef] [PubMed]
  137. Ahmar, S.; Zhai, Y.; Huang, H.; Yu, K.; Khan, M.H.U.; Shahid, M.; Samad, R.A.; Khan, S.U.; Amoo, O.; Fan, C.; et al. Development of mutants with varying flowering times by targeted editing of multiple SVP gene copies in Brassica napus L. Crop J. 2022, 10, 67–74. [Google Scholar] [CrossRef]
  138. Zhou, E.; Zhang, Y.; Wang, H.; Jia, Z.; Wang, X.; Wen, J.; Shen, J.; Fu, T.; Yi, B. Identification and Characterization of the MIKC-Type MADS-Box Gene Family in Brassica napus and Its Role in Floral Transition. Int. J. Mol. Sci. 2022, 23, 4289. [Google Scholar] [CrossRef]
  139. Odipio, J.; Alicai, T.; Nusinow, D.; Bart, R.; Taylor, N. CRISPR/Cas9-mediated disruption of multiple TFL1-like floral repressors activates flowering in cassava. In Vitro Cellular & Developmental Biology-Animal; Springer: New York, NY, USA, 2018; Volume 54, p. S47. [Google Scholar]
  140. Bull, S.E.; Seung, D.; Chanez, C.; Mehta, D.; Kuon, J.-E.; Truernit, E.; Hochmuth, A.; Zurkirchen, I.; Zeeman, S.C.; Gruissem, W.; et al. Accelerated ex situ breeding of GBSS- and PTST1-edited cassava for modified starch. Sci. Adv. 2018, 4, eaat6086. [Google Scholar] [CrossRef] [PubMed]
  141. Liu, L.; Xue, Y.; Luo, J.; Han, M.; Liu, X.; Jiang, T.; Zhao, Y.; Xu, Y.; Ma, C. Developing a UV–visible reporter-assisted CRISPR/Cas9 gene editing system to alter flowering time in Chrysanthemum indicum. Plant Biotechnol. J. 2023, 21, 1519–1521. [Google Scholar] [CrossRef]
  142. Huang, C.; Sun, H.; Xu, D.; Chen, Q.; Liang, Y.; Wang, X.; Xu, G.; Tian, J.; Wang, C.; Li, D.; et al. ZmCCT9 enhances maize adaptation to higher latitudes. Proc. Natl. Acad. Sci. USA 2018, 115, E334–E341. [Google Scholar] [CrossRef]
  143. Li, Q.; Wu, G.; Zhao, Y.; Wang, B.; Zhao, B.; Kong, D.; Wei, H.; Chen, C.; Wang, H. CRISPR/Cas9-mediated knockout and overexpression studies reveal a role of maize phytochrome C in regulating flowering time and plant height. Plant Biotechnol. J. 2020, 18, 2520–2532. [Google Scholar] [CrossRef]
  144. Takahashi, H.; Nishihara, M.; Yoshida, C.; Itoh, K. Gentian FLOWERING LOCUS T orthologs regulate phase transitions: Floral induction and endodormancy release. Plant Physiol. 2022, 188, 1887–1899. [Google Scholar] [CrossRef] [PubMed]
  145. Ying, S.; Scheible, W.-R.; Lundquist, P.K. A stress-inducible protein regulates drought tolerance and flowering time in Brachypodium and Arabidopsis. Plant Physiol. 2022, 191, 643–659. [Google Scholar] [CrossRef] [PubMed]
  146. Sheng, H.; Jiang, Y.; Rahmati, M.; Chia, J.-C.; Dokuchayeva, T.; Kavulych, Y.; Zavodna, T.-O.; Mendoza, P.N.; Huang, R.; Smieshka, L.M.; et al. YSL3-mediated copper distribution is required for fertility, seed size and protein accumulation in Brachypodium. Plant Physiol. 2021, 186, 655–676. [Google Scholar] [CrossRef] [PubMed]
  147. Herath, D.; Voogd, C.; Mayo-Smith, M.; Yang, B.; Allan, A.C.; Putterill, J.; Varkonyi-Gasic, E. CRISPR-Cas9 -mediated mutagenesis of kiwifruit BFT genes results in an evergrowing but not early flowering phenotype. Plant Biotechnol. J. 2022, 20, 2064–2076. [Google Scholar] [CrossRef] [PubMed]
  148. Varkonyi-Gasic, E.; Wang, T.; Voogd, C.; Jeon, S.; Drummond, R.S.M.; Gleave, A.P.; Allan, A.C. Mutagenesis of kiwifruit CENTRORADIALIS-like genes transforms a climbing woody perennial with long juvenility and axillary flowering into a compact plant with rapid terminal flowering. Plant Biotechnol. J. 2019, 17, 869–880. [Google Scholar] [CrossRef]
  149. Varkonyi-Gasic, E.; Wang, T.; Cooney, J.; Jeon, S.; Voogd, C.; Douglas, M.J.; Pilkington, S.M.; Akagi, T.; Allan, A.C. Shy Girl, a kiwifruit suppressor of feminization, restricts gynoecium development via regulation of cytokinin metabolism and signalling. New Phytol. 2021, 230, 1461–1475. [Google Scholar] [CrossRef]
  150. Singer, S.D.; Hughes, K.B.; Subedi, U.; Dhariwal, G.K.; Kader, K.; Acharya, S.; Chen, G.; Hannoufa, A. The CRISPR/Cas9-Mediated Modulation of SQUAMOSA PROMOTER-BINDING PROTEIN-LIKE 8 in Alfalfa Leads to Distinct Phenotypic Outcomes. Front. Plant Sci. 2022, 12, 774146. [Google Scholar] [CrossRef]
  151. Galindo-Sotomonte, L.; Jozefkowicz, C.; Gómez, C.; Stritzler, M.; Frare, R.; Bottero, E.; Tajima, H.; Blumwald, E.; Ayub, N.; Soto, G. CRISPR/Cas9-mediated knockout of a polyester synthase-like gene delays flowering time in alfalfa. Plant Cell Rep. 2023, 42, 953–956. [Google Scholar] [CrossRef]
  152. Wolabu, T.W.; Mahmood, K.; Jerez, I.T.; Cong, L.; Yun, J.; Udvardi, M.; Tadege, M.; Wang, Z.; Wen, J. Multiplex CRISPR/Cas9-mediated mutagenesis of alfalfa FLOWERING LOCUS Ta1 (MsFTa1) leads to delayed flowering time with improved forage biomass yield and quality. Plant Biotechnol. J. 2023, 21, 1383–1392. [Google Scholar] [CrossRef]
  153. Shibuya, K.; Watanabe, K.; Ono, M. CRISPR/Cas9-mediated mutagenesis of the EPHEMERAL1 locus that regulates petal senescence in Japanese morning glory. Plant Physiol. Biochem. 2018, 131, 53–57. [Google Scholar] [CrossRef]
  154. André, D.; Marcon, A.; Lee, K.C.; Goretti, D.; Zhang, B.; Delhomme, N.; Schmid, M.; Nilsson, O. FLOWERING LOCUS T paralogs control the annual growth cycle in Populus trees. Curr. Biol. 2022, 32, 2988–2996.e4. [Google Scholar] [CrossRef] [PubMed]
  155. Elorriaga, E.; Klocko, A.L.; Ma, C.; Strauss, S.H. Variation in Mutation Spectra Among CRISPR/Cas9 Mutagenized Poplars. Front. Plant Sci. 2018, 9, 594. [Google Scholar] [CrossRef] [PubMed]
  156. Lebedeva, M.; Komakhin, R.; Konovalova, L.; Ivanova, L.; Taranov, V.; Monakhova, Y.; Babakov, A.; Klepikova, A.; Zlobin, N. Development of potato (Solanum tuberosum L.) plants with StLEAFY knockout. Planta 2022, 256, 116. [Google Scholar] [CrossRef] [PubMed]
  157. Li, X.; Zhou, W.; Ren, Y.; Tian, X.; Lv, T.; Wang, Z.; Fang, J.; Chu, C.; Yang, J.; Bu, Q. High-efficiency breeding of early-maturing rice cultivars via CRISPR/Cas9-mediated genome editing. J. Genet. Genom. 2017, 44, 175–178. [Google Scholar] [CrossRef]
  158. Brambilla, V.; Martignago, D.; Goretti, D.; Cerise, M.; Somssich, M.; de Rosa, M.; Galbiati, F.; Shrestha, R.; Lazzaro, F.; Simon, R.; et al. Antagonistic Transcription Factor Complexes Modulate the Floral Transition in Rice. Plant Cell 2017, 29, 2801–2816. [Google Scholar] [CrossRef] [PubMed]
  159. Zhou, X.; Jiang, G.; Yang, L.; Qiu, L.; He, P.; Nong, C.; Wang, Y.; He, Y.; Xing, Y. Gene diagnosis and targeted breeding for blast-resistant Kongyu 131 without changing regional adaptability. J. Genet. Genom. 2018, 45, 539–547. [Google Scholar] [CrossRef] [PubMed]
  160. Cui, Y.; Zhu, M.; Xu, Z.; Xu, Q. Assessment of the effect of ten heading time genes on reproductive transition and yield components in rice using a CRISPR/Cas9 system. Theor. Appl. Genet. 2019, 132, 1887–1896. [Google Scholar] [CrossRef]
  161. Wang, G.; Wang, C.; Lu, G.; Wang, W.; Mao, G.; Habben, J.E.; Song, C.; Wang, J.; Chen, J.; Gao, Y.; et al. Knockouts of a late flowering gene via CRISPR–Cas9 confer early maturity in rice at multiple field locations. Plant Mol. Biol. 2020, 104, 137–150. [Google Scholar] [CrossRef]
  162. Chandrashekar, B.K.; Ag, S.; Makarla, U.; Ramu, V.S. Disruption in the DNA Mismatch Repair Gene MSH2 by CRISPR-Cas9 in Indica Rice Can Create Genetic Variability. J. Agric. Food Chem. 2021, 69, 4144–4152. [Google Scholar] [CrossRef]
  163. Andrew-Peter-Leon, M.T.; Selvaraj, R.; Kumar, K.K.; Muthamilarasan, M.; Yasin, J.K.; Pillai, M.A. Loss of Function of OsFBX267 and OsGA20ox2 in Rice Promotes Early Maturing and Semi-Dwarfism in γ-Irradiated IWP and Genome-Edited Pusa Basmati-1. Front. Plant Sci. 2021, 12, 714066. [Google Scholar] [CrossRef] [PubMed]
  164. Sun, B.; Xue, P.; Wen, X.-X.; Gong, K.; Wang, B.-F.; Xu, P.; Lin, Z.-C.; Peng, Z.-Q.; Fu, J.-L.; Yu, P.; et al. qHD5 encodes an AP2 factor that suppresses rice heading by down-regulating Ehd2 expression. Plant Sci. 2022, 324, 111446. [Google Scholar] [CrossRef] [PubMed]
  165. Yin, Y.; Yan, Z.; Guan, J.; Huo, Y.; Wang, T.; Li, T.; Cui, Z.; Ma, W.; Wang, X.; Chen, W. Two interacting basic helix-loop-helix transcription factors control flowering time in rice. Plant Physiol. 2023, 192, 205–221. [Google Scholar] [CrossRef] [PubMed]
  166. Sedeek, K.; Zuccolo, A.; Fornasiero, A.; Weber, A.M.; Sanikommu, K.; Sampathkumar, S.; Rivera, L.F.; Butt, H.; Mussurova, S.; Alhabsi, A.; et al. Multi-omics resources for targeted agronomic improvement of pigmented rice. Nat. Food 2023, 4, 366–371. [Google Scholar] [CrossRef] [PubMed]
  167. Gu, H.; Zhang, K.; Chen, J.; Gull, S.; Chen, C.; Hou, Y.; Li, X.; Miao, J.; Zhou, Y.; Liang, G. OsFTL4, an FT-like Gene, Regulates Flowering Time and Drought Tolerance in Rice (Oryza sativa L.). Rice 2022, 15, 47. [Google Scholar] [CrossRef] [PubMed]
  168. Sun, C.; Zhang, K.; Zhou, Y.; Xiang, L.; He, C.; Zhong, C.; Li, K.; Wang, Q.; Yang, C.; Wang, Q.; et al. Dual function of clock component OsLHY sets critical day length for photoperiodic flowering in rice. Plant Biotechnol. J. 2021, 19, 1644–1657. [Google Scholar] [CrossRef]
  169. Zhang, J.; Fan, X.; Hu, Y.; Zhou, X.; He, Q.; Liang, L.; Xing, Y. Global analysis of CCT family knockout mutants identifies four genes involved in regulating heading date in rice. J. Integr. Plant Biol. 2021, 63, 913–923. [Google Scholar] [CrossRef]
  170. Cui, Y.; Xu, Z.; Xu, Q. Elucidation of the relationship between yield and heading date using CRISPR/Cas9 system-induced mutation in the flowering pathway across a large latitudinal gradient. Mol. Breed. 2021, 41, 23. [Google Scholar] [CrossRef]
  171. Yasui, Y.; Tanaka, W.; Sakamoto, T.; Kurata, T.; Hirano, H.-Y. Genetic Enhancer Analysis Reveals that FLORAL ORGAN NUMBER2 and OsMADS3 Co-operatively Regulate Maintenance and Determinacy of the Flower Meristem in Rice. Plant Cell Physiol. 2017, 58, 893–903. [Google Scholar] [CrossRef]
  172. Wu, M.; Liu, H.; Lin, Y.; Chen, J.; Fu, Y.; Luo, J.; Zhang, Z.; Liang, K.; Chen, S.; Wang, F. In-Frame and Frame-Shift Editing of the Ehd1 Gene to Develop Japonica Rice with Prolonged Basic Vegetative Growth Periods. Front. Plant Sci. 2020, 11, 307. [Google Scholar] [CrossRef]
  173. Li, C.; Liu, X.-J.; Yan, Y.; Shah, A.; Liu, Z.; Tao, R.-F.; Yue, E.-K.; Duan, M.-H.; Xu, J. OsLHY regulates photoperiodic flowering through the unique pathways under long-day conditions in rice (Oryza sativa L.). Authorea 2021. [Google Scholar] [CrossRef]
  174. Liu, X.; Liu, H.; Zhang, Y.; He, M.; Li, R.; Meng, W.; Wang, Z.; Li, X.; Bu, Q. Fine-tuning Flowering Time via Genome Editing of Upstream Open Reading Frames of Heading Date 2 in Rice. Rice 2021, 14, 69. [Google Scholar] [CrossRef] [PubMed]
  175. Zhang, S.; Deng, L.; Cheng, R.; Hu, J.; Wu, C. RID1 sets rice heading date by balancing its binding with SLR1 and SDG722. J. Integr. Plant Biol. 2022, 64, 149–165. [Google Scholar] [CrossRef] [PubMed]
  176. Xu, P.; Zhang, Y.; Wen, X.; Yang, Q.; Liu, L.; Hao, S.; Li, J.; Wu, Z.; Shah, L.; Sohail, A.; et al. The clock component OsLUX regulates rice heading through recruiting OsELF3-1 and OsELF4s to repress Hd1 and Ghd7. J. Adv. Res. 2023, 48, 17–31. [Google Scholar] [CrossRef] [PubMed]
  177. Zhang, S.; Deng, L.; Zhao, L.; Wu, C. Genome-wide binding analysis of transcription factor Rice Indeterminate 1 reveals a complex network controlling rice floral transition. J. Integr. Plant Biol. 2022, 64, 1690–1705. [Google Scholar] [CrossRef]
  178. Andrade, L.; Lu, Y.; Cordeiro, A.; Costa, J.M.F.; Wigge, P.A.; Saibo, N.J.M.; Jaeger, K.E. The evening complex integrates photoperiod signals to control flowering in rice. Proc. Natl. Acad. Sci. USA 2022, 119, e2122582119. [Google Scholar] [CrossRef]
  179. Dai, X.; Yang, X.; Wang, C.; Fan, Y.; Xin, S.; Hua, Y.; Wang, K.; Huang, H. CRISPR/Cas9-mediated genome editing in Hevea brasiliensis. Ind. Crops Prod. 2021, 164, 113418. [Google Scholar] [CrossRef]
  180. Wang, H.; Jia, G.; Zhang, N.; Zhi, H.; Xing, L.; Zhang, H.; Sui, Y.; Tang, S.; Li, M.; Zhang, H.; et al. Domestication-associated PHYTOCHROME C is a flowering time repressor and a key factor determining Setaria as a short-day plant. New Phytol. 2022, 236, 1809–1823. [Google Scholar] [CrossRef]
  181. Zhu, C.; Box, M.S.; Thiruppathi, D.; Hu, H.; Yu, Y.; Martin, C.; Doust, A.N.; McSteen, P.; Kellogg, E.A. Pleiotropic and nonredundant effects of an auxin importer in Setaria and maize. Plant Physiol. 2022, 189, 715–734. [Google Scholar] [CrossRef]
  182. Char, S.N.; Wei, J.; Mu, Q.; Li, X.; Zhang, Z.J.; Yu, J.; Yang, B. An Agrobacterium-delivered CRISPR/Cas9 system for targeted mutagenesis in sorghum. Plant Biotechnol. J. 2020, 18, 319–321. [Google Scholar] [CrossRef]
  183. Han, J.; Guo, B.; Guo, Y.; Zhang, B.; Wang, X.; Qiu, L.-J. Creation of Early Flowering Germplasm of Soybean by CRISPR/Cas9 Technology. Front. Plant Sci. 2019, 10, 1446. [Google Scholar] [CrossRef]
  184. Wang, L.; Sun, S.; Wu, T.; Liu, L.; Sun, X.; Cai, Y.; Li, J.; Jia, H.; Yuan, S.; Chen, L.; et al. Natural variation and CRISPR/Cas9-mediated mutation in GmPRR37 affect photoperiodic flowering and contribute to regional adaptation of soybean. Plant Biotechnol. J. 2020, 18, 1869–1881. [Google Scholar] [CrossRef]
  185. Wang, W.; Wang, Z.; Hou, W.; Chen, L.; Jiang, B.; Liu, W.; Feng, Y.; Wu, C. GmNMHC5, A Neoteric Positive Transcription Factor of Flowering and Maturity in Soybean. Plants 2020, 9, 792. [Google Scholar] [CrossRef]
  186. Li, Z.; Cheng, Q.; Gan, Z.; Hou, Z.; Zhang, Y.; Li, Y.; Li, H.; Nan, H.; Yang, C.; Chen, L.; et al. Multiplex CRISPR/Cas9-mediated knockout of soybean LNK2 advances flowering time. Crop J. 2021, 9, 767–776. [Google Scholar] [CrossRef]
  187. Wan, Z.; Liu, Y.; Guo, D.; Fan, R.; Liu, Y.; Xu, K.; Zhu, J.; Quan, L.; Lu, W.; Bai, X.; et al. CRISPR/Cas9-mediated targeted mutation of the E1 decreases photoperiod sensitivity, alters stem growth habits, and decreases branch number in soybean. Front. Plant Sci. 2022, 13, 1066820. [Google Scholar] [CrossRef] [PubMed]
  188. Zhai, H.; Wan, Z.; Jiao, S.; Zhou, J.; Xu, K.; Nan, H.; Liu, Y.; Xiong, S.; Fan, R.; Zhu, J.; et al. GmMDE genes bridge the maturity gene E1 and florigens in photoperiodic regulation of flowering in soybean. Plant Physiol. 2022, 189, 1021–1036. [Google Scholar] [CrossRef]
  189. Cai, Y.; Chen, L.; Liu, X.; Guo, C.; Sun, S.; Wu, C.; Jiang, B.; Han, T.; Hou, W. CRISPR/Cas9-mediated targeted mutagenesis of GmFT2a delays flowering time in soya bean. Plant Biotechnol. J. 2018, 16, 176–185. [Google Scholar] [CrossRef]
  190. Wang, Y.; Yuan, L.; Su, T.; Wang, Q.; Gao, Y.; Zhang, S.; Jia, Q.; Yu, G.; Fu, Y.; Cheng, Q.; et al. Light- and temperature-entrainable circadian clock in soybean development. Plant Cell Environ. 2020, 43, 637–648. [Google Scholar] [CrossRef] [PubMed]
  191. Li, C.; Li, Y.-H.; Li, Y.; Lu, H.; Hong, H.; Tian, Y.; Li, H.; Zhao, T.; Zhou, X.; Liu, J.; et al. A Domestication-Associated Gene GmPRR3b Regulates the Circadian Clock and Flowering Time in Soybean. Mol. Plant 2020, 13, 745–759. [Google Scholar] [CrossRef]
  192. Chen, L.; Nan, H.; Kong, L.; Yue, L.; Yang, H.; Zhao, Q.; Fang, C.; Li, H.; Cheng, Q.; Lu, S.; et al. Soybean AP1 homologs control flowering time and plant height. J. Integr. Plant Biol. 2020, 62, 1868–1879. [Google Scholar] [CrossRef]
  193. Zhao, F.; Lyu, X.; Ji, R.; Liu, J.; Zhao, T.; Li, H.; Liu, B.; Pei, Y. CRISPR/Cas9-engineered mutation to identify the roles of phytochromes in regulating photomorphogenesis and flowering time in soybean. Crop J. 2022, 10, 1654–1664. [Google Scholar] [CrossRef]
  194. Schmidt, F.J.; Zimmermann, M.M.; Wiedmann, D.R.; Lichtenauer, S.; Grundmann, L.; Muth, J.; Twyman, R.M.; Prüfer, D.; Noll, G.A. The Major Floral Promoter NtFT5 in Tobacco (Nicotiana tabacum) Is a Promising Target for Crop Improvement. Front. Plant Sci. 2020, 10, 1666. [Google Scholar] [CrossRef] [PubMed]
  195. Soyk, S.; Müller, N.A.; Park, S.J.; Schmalenbach, I.; Jiang, K.; Hayama, R.; Zhang, L.; Van Eck, J.; Jiménez-Gómez, J.M.; Lippman, Z.B. Variation in the flowering gene SELF PRUNING 5G promotes day-neutrality and early yield in tomato. Nat. Genet. 2017, 49, 162–168. [Google Scholar] [CrossRef] [PubMed]
  196. Lemmon, Z.H.; Reem, N.T.; Dalrymple, J.; Soyk, S.; Swartwood, K.E.; Rodriguez-Leal, D.; Van Eck, J.; Lippman, Z.B. Rapid improvement of domestication traits in an orphan crop by genome editing. Nat. Plants 2018, 4, 766–770. [Google Scholar] [CrossRef] [PubMed]
  197. Li, T.; Yang, X.; Yu, Y.; Si, X.; Zhai, X.; Zhang, H.; Dong, W.; Gao, C.; Xu, C. Domestication of wild tomato is accelerated by genome editing. Nat. Biotechnol. 2018, 36, 1160–1163. [Google Scholar] [CrossRef] [PubMed]
  198. Hu, G.; Wang, K.; Huang, B.; Mila, I.; Frasse, P.; Maza, E.; Djari, A.; Hernould, M.; Zouine, M.; Li, Z.; et al. The auxin-responsive transcription factor SlDOF9 regulates inflorescence and flower development in tomato. Nat. Plants 2022, 8, 419–433. [Google Scholar] [CrossRef] [PubMed]
  199. Moreira, J.d.R.; Quiñones, A.; Lira, B.S.; Robledo, J.M.; Curtin, S.J.; Vicente, M.H.; Ribeiro, D.M.; Ryngajllo, M.; Jiménez-Gómez, J.M.; Peres, L.E.P.; et al. SELF PRUNING 3C is a flowering repressor that modulates seed germination, root architecture, and drought responses. J. Exp. Bot. 2022, 73, 6226–6240. [Google Scholar] [CrossRef]
  200. Xu, C.; Park, S.J.; Van Eck, J.; Lippman, Z.B. Control of inflorescence architecture in tomato by BTB/POZ transcriptional regulators. Minerva Anestesiol. 2016, 30, 2048–2061. [Google Scholar] [CrossRef]
  201. Lin, W.; Gupta, S.K.; Arazi, T.; Spitzer-Rimon, B. MIR172d Is Required for Floral Organ Identity and Number in Tomato. Int. J. Mol. Sci. 2021, 22, 4659. [Google Scholar] [CrossRef]
  202. Kwon, C.-T.; Heo, J.; Lemmon, Z.H.; Capua, Y.; Hutton, S.F.; Van Eck, J.; Park, S.J.; Lippman, Z.B. Rapid customization of Solanaceae fruit crops for urban agriculture. Nat. Biotechnol. 2020, 38, 182–188. [Google Scholar] [CrossRef]
  203. Gupta, A.; Hua, L.; Zhang, Z.; Yang, B.; Li, W. CRISPR-induced miRNA156-recognition element mutations in TaSPL13 improve multiple agronomic traits in wheat. Plant Biotechnol. J. 2023, 21, 536–548. [Google Scholar] [CrossRef]
  204. Sun, J.; Bie, X.M.; Chu, X.L.; Wang, N.; Zhang, X.S.; Gao, X.-Q. Genome-edited TaTFL1-5 mutation decreases tiller and spikelet numbers in common wheat. Front. Plant Sci. 2023, 14, 1142779. [Google Scholar] [CrossRef] [PubMed]
  205. Chen, Z.; Ke, W.; He, F.; Chai, L.; Cheng, X.; Xu, H.; Wang, X.; Du, D.; Zhao, Y.; Chen, X.; et al. A single nucleotide deletion in the third exon of FT-D1 increases the spikelet number and delays heading date in wheat (Triticum aestivum L.). Plant Biotechnol. J. 2022, 20, 920–933. [Google Scholar] [CrossRef] [PubMed]
  206. Errum, A.; Rehman, N.; Uzair, M.; Inam, S.; Ali, G.M.; Khan, M.R. CRISPR/Cas9 editing of wheat Ppd-1 gene homoeologs alters spike architecture and grain morphometric traits. Funct. Integr. Genom. 2023, 23, 66. [Google Scholar] [CrossRef]
  207. Liu, H.; Wang, K.; Tang, H.; Gong, Q.; Du, L.; Pei, X.; Ye, X. CRISPR/Cas9 editing of wheat TaQ genes alters spike morphogenesis and grain threshability. J. Genet. Genom. 2020, 47, 563–575. [Google Scholar] [CrossRef] [PubMed]
  208. 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] [PubMed]
  209. Kang, M.; Lee, K.; Finley, T.; Chappell, H.; Veena, V.; Wang, K. An Improved Agrobacterium-Mediated Transformation and Genome-Editing Method for Maize Inbred B104 Using a Ternary Vector System and Immature Embryos. Front. Plant Sci. 2022, 13, 860971. [Google Scholar] [CrossRef]
  210. Uranga, M.; Daròs, J. Tools and targets: The dual role of plant viruses in CRISPR–Cas genome editing. Plant Genome 2023, 16, e20220. [Google Scholar] [CrossRef]
  211. Maher, M.F.; Nasti, R.A.; Vollbrecht, M.; Starker, C.G.; Clark, M.D.; Voytas, D.F. Plant gene editing through de novo induction of meristems. Nat. Biotechnol. 2020, 38, 84–89. [Google Scholar] [CrossRef]
  212. Lee, K.; Wang, K. Strategies for genotype-flexible plant transformation. Curr. Opin. Biotechnol. 2023, 79, 102848. [Google Scholar] [CrossRef]
  213. Impens, L.; Lorenzo, C.D.; Vandeputte, W.; Wytynck, P.; Debray, K.; Haeghebaert, J.; Herwegh, D.; Jacobs, T.B.; Ruttink, T.; Nelissen, H.; et al. Combining multiplex gene editing and doubled haploid technology in maize. New Phytol. 2023, 239, 1521–1532. [Google Scholar] [CrossRef] [PubMed]
  214. Gaillochet, C.; Develtere, W.; Jacobs, T.B. CRISPR screens in plants: Approaches, guidelines, and future prospects. Plant Cell 2021, 33, 794–813. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A simplified schematic representation of the physiological pathways involved in flowering, including photoperiod, autonomous, vernalization, hormonal, temperature, aging, and sugar. Each pathway has one major player which in turn regulates FT. These pathways, except for photoperiod, play a role not only in leaves, but also in shoot meristems. Flowering Locus T (FT), known as florigen, plays a central role as it travels long distances from leaf to the shoot apical meristem (SAM). Arrows indicate positive regulation, while blunt-ended arrows indicate negative regulation. Abbreviations: SVP = SHORT VEGETATIVE PHASE, FLC = FLOWERING LOCUS C, SOC1 = SUPPRESSOR OF OVEREXPRESION OF CO1, FT = FLOWERING LOCUS T, GA = GIBBERELLIC ACID, SPL = SQUAMOSA PROMOTER BINDING PROTEIN-LIKE, T6P = TREHALOSE-6-PHOSPHATE, CO = CONSTANS, CK = CYTOKININ.
Figure 1. A simplified schematic representation of the physiological pathways involved in flowering, including photoperiod, autonomous, vernalization, hormonal, temperature, aging, and sugar. Each pathway has one major player which in turn regulates FT. These pathways, except for photoperiod, play a role not only in leaves, but also in shoot meristems. Flowering Locus T (FT), known as florigen, plays a central role as it travels long distances from leaf to the shoot apical meristem (SAM). Arrows indicate positive regulation, while blunt-ended arrows indicate negative regulation. Abbreviations: SVP = SHORT VEGETATIVE PHASE, FLC = FLOWERING LOCUS C, SOC1 = SUPPRESSOR OF OVEREXPRESION OF CO1, FT = FLOWERING LOCUS T, GA = GIBBERELLIC ACID, SPL = SQUAMOSA PROMOTER BINDING PROTEIN-LIKE, T6P = TREHALOSE-6-PHOSPHATE, CO = CONSTANS, CK = CYTOKININ.
Plants 12 04020 g001
Figure 2. The transport of FT from the leaf to the SAM involves a number of steps. First, FT is synthesized in the leaf through the pathways shown in Figure 1 (A). Next, FTIP1 facilitates the transfer of FT from CC to SE (B). Finally, NaKR1 is responsible for the long-distance transport of FT into the SAM (C). The scissor symbols in the diagram indicate potential CRISPR knockout sites that could be used to disrupt FT transport. Abbreviations: FT = FLOWERING LOCUS T, SAM = SHOOT APICAL MERISTEM, FTIP1 = FT-INTERACTION PROTEIN 1, CC = Companion cell, SE = Sieve element, NaKR1 = SODIUM POTASSIUM ROOT DEFECTIVE 1 (Created with BioRender.com).
Figure 2. The transport of FT from the leaf to the SAM involves a number of steps. First, FT is synthesized in the leaf through the pathways shown in Figure 1 (A). Next, FTIP1 facilitates the transfer of FT from CC to SE (B). Finally, NaKR1 is responsible for the long-distance transport of FT into the SAM (C). The scissor symbols in the diagram indicate potential CRISPR knockout sites that could be used to disrupt FT transport. Abbreviations: FT = FLOWERING LOCUS T, SAM = SHOOT APICAL MERISTEM, FTIP1 = FT-INTERACTION PROTEIN 1, CC = Companion cell, SE = Sieve element, NaKR1 = SODIUM POTASSIUM ROOT DEFECTIVE 1 (Created with BioRender.com).
Plants 12 04020 g002
Figure 3. Showing the effect of cis-regulatory elements in controlling gene expression at the transcriptional level in CRISPR-based flowering time engineering. By highlighting key regulators such as FRIGIDA and PRC-1-like complexes, the figure illustrates their ability to fine-tune FLC expression either positively or negatively in transcriptional regulation. Green arrow indicates as positive regulation, while red blunt-ended arrow indicates as negative regulation. Abbreviations: SUF4 = SUPRESSOR OF FRIGIDA 4, CBP20 = CAP-BINDING PROTEIN 20, FLX = FLC EXPRESSOR, FES1 = FRIGIDA ESSENTIAL 1, CBP80 = CAP BINDING PROTEIN 80, FRL2 = FRIGIDA LIKE 2, FRL1 = FRIGIDA LIKE 1, FLL4 = FLOWERING LOCUS C EXPRESSOR-LIKE 4, EFS = EARLY FLOWERING IN SHORT DAYS, BMI1A = DREB2A-INTERACTING PROTEIN 2, BMI1B = DREB2A-INTERACTING PROTEIN 1, BMI1C = BMI1C, VRN1 = REDUCED VERNALIZATION RESPONSE 1, EMF1 = EMBRYONIC FLOWER 1, LHP1 = LIKE HETEROCHROMATIN PROTEIN 1, LIF2 = LHP1-INTERACTING FACTOR [31] (Created with BioRender.com).
Figure 3. Showing the effect of cis-regulatory elements in controlling gene expression at the transcriptional level in CRISPR-based flowering time engineering. By highlighting key regulators such as FRIGIDA and PRC-1-like complexes, the figure illustrates their ability to fine-tune FLC expression either positively or negatively in transcriptional regulation. Green arrow indicates as positive regulation, while red blunt-ended arrow indicates as negative regulation. Abbreviations: SUF4 = SUPRESSOR OF FRIGIDA 4, CBP20 = CAP-BINDING PROTEIN 20, FLX = FLC EXPRESSOR, FES1 = FRIGIDA ESSENTIAL 1, CBP80 = CAP BINDING PROTEIN 80, FRL2 = FRIGIDA LIKE 2, FRL1 = FRIGIDA LIKE 1, FLL4 = FLOWERING LOCUS C EXPRESSOR-LIKE 4, EFS = EARLY FLOWERING IN SHORT DAYS, BMI1A = DREB2A-INTERACTING PROTEIN 2, BMI1B = DREB2A-INTERACTING PROTEIN 1, BMI1C = BMI1C, VRN1 = REDUCED VERNALIZATION RESPONSE 1, EMF1 = EMBRYONIC FLOWER 1, LHP1 = LIKE HETEROCHROMATIN PROTEIN 1, LIF2 = LHP1-INTERACTING FACTOR [31] (Created with BioRender.com).
Plants 12 04020 g003
Figure 4. Various CRISPR-Cas approaches to modulate flowering time: (A) Cas cleavage-induced amino acid changes resulting in non-functional protein. (B) Cas cleavage resulting in stop codon insertion, terminating transcription, and preventing mRNA and, therefore, protein production. (C) Disruption of gene transcription by targeting elements involved in transcription, such as transcription factors, promoters, and upstream enhancer regions. The scissor symbols in the diagram indicate potential CRISPR knockout targets (D) Genes that regulate chromatin remodeling act as activators or inhibitors, affecting transcription by controlling chromatin structure (Created with BioRender.com).
Figure 4. Various CRISPR-Cas approaches to modulate flowering time: (A) Cas cleavage-induced amino acid changes resulting in non-functional protein. (B) Cas cleavage resulting in stop codon insertion, terminating transcription, and preventing mRNA and, therefore, protein production. (C) Disruption of gene transcription by targeting elements involved in transcription, such as transcription factors, promoters, and upstream enhancer regions. The scissor symbols in the diagram indicate potential CRISPR knockout targets (D) Genes that regulate chromatin remodeling act as activators or inhibitors, affecting transcription by controlling chromatin structure (Created with BioRender.com).
Plants 12 04020 g004
Figure 5. (A) Peer-reviewed publications on CRISPR-based flowering time engineering by crop (2015–mid-2023). (B) Annual number of indexed articles on CRISPR-mediated flowering time engineering. (A,B) are derived from the information presented in Table 1, which represents a statistical population that provides insight into the general state of the field.
Figure 5. (A) Peer-reviewed publications on CRISPR-based flowering time engineering by crop (2015–mid-2023). (B) Annual number of indexed articles on CRISPR-mediated flowering time engineering. (A,B) are derived from the information presented in Table 1, which represents a statistical population that provides insight into the general state of the field.
Plants 12 04020 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hodaei, A.; Werbrouck, S.P.O. Unlocking Nature’s Clock: CRISPR Technology in Flowering Time Engineering. Plants 2023, 12, 4020. https://doi.org/10.3390/plants12234020

AMA Style

Hodaei A, Werbrouck SPO. Unlocking Nature’s Clock: CRISPR Technology in Flowering Time Engineering. Plants. 2023; 12(23):4020. https://doi.org/10.3390/plants12234020

Chicago/Turabian Style

Hodaei, Ashkan, and Stefaan P. O. Werbrouck. 2023. "Unlocking Nature’s Clock: CRISPR Technology in Flowering Time Engineering" Plants 12, no. 23: 4020. https://doi.org/10.3390/plants12234020

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