Next Article in Journal
Evaluation of Soybean Plant Introductions for Traits that can Improve Emergence under Varied Soil Moisture Levels
Next Article in Special Issue
Low Lignin Mutants and Reduction of Lignin Content in Grasses for Increased Utilisation of Lignocellulose
Previous Article in Journal
Modelling Nitrogen Uptake in Plants and Phytoplankton: Advantages of Integrating Flexibility into the Spatial and Temporal Dynamics of Nitrate Absorption
Previous Article in Special Issue
Genomic Selection in Cereal Breeding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unraveling Molecular and Genetic Studies of Wheat (Triticum aestivum L.) Resistance against Factors Causing Pre-Harvest Sprouting

1
College of Agronomy, Anhui Agricultural University, Hefei 230036, China
2
Key Laboratory of Wheat Biology and Genetic Improvement on South Yellow and Huai River Valley, Ministry of Agriculture, Hefei 230036, China
3
National Engineering Laboratory for Crop Stress Resistance Breeding, Hefei 230036, China
4
Anhui Key Laboratory of Crop Biology, Hefei 230036, China
5
School of Life Sciences, Anhui Agricultural University, Hefei 230036, China
*
Author to whom correspondence should be addressed.
Agronomy 2019, 9(3), 117; https://doi.org/10.3390/agronomy9030117
Submission received: 10 December 2018 / Revised: 6 February 2019 / Accepted: 26 February 2019 / Published: 1 March 2019

Abstract

:
Pre-harvest sprouting (PHS) is one of the most important factors having adverse effects on yield and grain quality all over the world, particularly in wet harvest conditions. PHS is controlled by both genetic and environmental factors and the interaction of these factors. Breeding varieties with high PHS resistance have important implications for reducing yield loss and improving grain quality. The rapid advancements in the wheat genomic database along with transcriptomic and proteomic technologies have broadened our knowledge for understanding the regulatory mechanism of PHS resistance at transcriptomic and post-transcriptomic levels. In this review, we have described in detail the recent advancements on factors influencing PHS resistance, including grain color, seed dormancy, α-amylase activity, plant hormones (especially abscisic acid and gibberellin), and QTL/genes, which are useful for mining new PHS-resistant genes and developing new molecular markers for multi-gene pyramiding breeding of wheat PHS resistance, and understanding the complicated regulatory mechanism of PHS resistance.

1. Introduction

Pre-harvest sprouting (PHS) refers to the germination of grains in mature cereal spikes before harvest under continuous wet weather conditions [1]. PHS has adverse impacts on wheat quality and yield [2,3] and reduces baking quality of dough by making it porous, sticky, and off-color. The price of sprouted grain is decreased by 20–50% and is unacceptable for human food if it contains more than 4% sprouted grains [4]. The decreased bread and noodle quality is due to increased activity of lipases, amylases, and proteases, enzymes which degrade lipids, starch, and proteins in sprouting grains [5,6]. Global yield and quality losses due to PHS have a financial impact estimated at $1 billion annually [7]. PHS occurred frequently in many major wheat producing areas of the world, including China, USA, Japan, Canada, Australia, and also in Europe [8]. In China, PHS is a major problem, especially in the northern spring wheat region, Yangtze River Valley, and northeastern spring wheat region which are characterized by heavy rainfall and high humidity before harvest [9]. In recent years, it has also become a serious problem in the Yellow and Huai Valleys’ wheat region due to climate changes. Therefore, improving PHS resistance is a major breeding objective to mitigate the risk of PHS and increase the production of high-quality wheat.
PHS resistance is associated with several developmental, physiological, and morphological features of the spike and seed, which includes seed coat (pericarp) color and permeability, seed dormancy, α-amylase activity, and levels of plant growth hormones (abscisic acid, gibberellin and auxin) [1,10,11,12,13,14,15,16,17,18]. Other factors, such as waxiness, hairiness, ear morphology, and germination-inhibitory compounds produced in bracts surrounding the grains have also been linked with PHS resistance [19,20]. Among them, seed dormancy is the major genetic factor controlling PHS resistance, therefore, much attention has been paid to understand the molecular mechanism of seed dormancy as a means to improve PHS resistance in wheat breeding programs.
PHS resistance is a typical quantitative trait controlled by numerous QTL/genes. Many quantitative trait loci (QTL) have been identified for PHS resistance in wheat [1,14,18,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37]. Several candidate genes for PHS resistance have also identified, including TaSdr, TaPHS1, TaMFT, TaVp-1, Tamyb10, and TaMKK3-A [38,39,40,41,42,43,44,45,46]. These QTL/genes are valuable for gene pyramiding in breeding programs. However, the regulatory mechanisms of PHS remain unclear, which is why progress in improving wheat PHS resistance is limited.
To understand the regulatory mechanism of PHS resistance and provide valuable information for developing PHS resistant wheat varieties, this review summarizes recent advances of several major factors affecting PHS resistance, including grain color, seed dormancy, α-amylase activity, and plant growth hormones.

1.1. Grain Color

Grain color (GC) is an important genetic factor affecting the brightness of flour and is also associated with seed dormancy and PHS resistance. It is controlled by the R-1 gene series distally located on long arms of chromosomes 3A, 3B, and 3D [47]. Dominant R-1 alleles confer red grain color and are denoted by R-A1b, R-B1b, and R-D1b whereas the recessive alleles contribute white grain color and are named as R-A1a, R-B1a, and R-D1a, respectively. For dominant R-A1b, R-B1b, and R-D1b alleles, only one allele is enough for red color, while redness increases in a gene dosage-dependent manner [48]. The R genes act as transcriptional activators of flavonoid synthesis genes and are positioned in the same region as Myb-type transcription factor loci (Tamyb10-A1, Tamyb10-B1, and Tamyb10-D1) [49]. Himi et al. [40] confirmed the three Tamyb10-1 genes on chromosomes 3AL, 3BL, and 3DL as candidate genes underlying the R-1 loci for wheat grain color.
The red pigment in the testa of plant grains is composed of catechin, and proanthocyanidins (PA) that are produced in the flavonoid biosynthesis pathway and synthesized by different enzymes such as dihydroflavonol-4-reductase (DFR), chalcone flavanone isomerase (CHI), flavanone 3-hydroxylase (F3H), and chalcone synthase (CHS) [50,51,52] (Figure 1). These enzymes are expressed only in immature red grains and are almost completely repressed in the grains of white wheat [49]. The above Myb-type Tamyb10-1 transcription factors control anthocyanin production and the red pigment of wheat grain by up-regulating the structural genes encoding DFR, CHI, F3H, and CHS in the flavonoid biosynthesis pathway.
In general, red-grained genotypes are more resistant to PHS compared to white-grained genotypes [53,54]. Himi et al. [53] observed the effect of R genes on grain dormancy by using near-isogenic red grained ANK lines and white grained mutant (EMS-AUS) lines and found that the level of dormancy conferred by R genes decreased rapidly in ANK lines during the after-ripening stage whereas reduction in the white grained mutant (EMS-AUS) line was not large indicating that R genes might play a minor role in seed dormancy. Groos et al. [1] detected four QTL for both PHS resistance and GC using a recombinant inbred line (RIL) population from a cross between Renan (red-grained) and Récital (white-grained). Three of these QTLs were close to R genes, and one was mapped on chromosome 5AS. Lin et al. [55] reported the genetic architecture of GC and PHS and genetic relationship of these two traits in a panel of 185 U.S. elite breeding lines and cultivars using a genome-wide association study (GWAS). These results showed that GC genes (Tamyb10-A1 and Tamyb10-D1) had a significant effect on PHS resistance, but Tamyb10-B1 was significant only for GC and not for PHS resistance. In addition, a novel QTL for GC was also identified on chromosome 1B. Zhou et al. [37] identified three main QTLs for PHS resistance by GWAS, including a novel locus on chromosome 5D and two loci co-located with Tamyb10-1 genes on chromosomes 3A and 3D. Furthermore, 32 GC-related QTLs (GCR-QTL) were also detected, and a strong correlation was observed between the number of GCR-QTL and seed germination rate. The above results imply that GC is significantly associated with PHS resistance, and might be controlled jointly by many QTLs in addition to he Tamyb10-1 gene. Of these, some QTLs are for both GC and PHS resistance; others are for GC only and not for PHS resistance. Therefore, it should be possible to breed PHS-resistant white wheat by using the gene-editing technology known as CRISPR/Cas9 to alter the GC-related genes keeping in view the other dormancy-related QTLs besides those provided by the R-1 genes of the red grained parent used for such editing.

1.2. Seed Dormancy

Dormancy is the inhibition of germination of morphologically ripe and healthy seeds even under optimum conditions of light, moisture, and temperature [56,57]. Initiation and maintenance of dormancy is affected by both genetic and environmental factors [58]. Dormancy is regarded as a major genetic component of PHS resistance [59,60,61]. Seed dormancy in wheat is a complex phenomenon and can be divided into seed coat-imposed and embryo-imposed dormancy [62,63]. Seed coat inhibitory compounds are associated with seed coat-based dormancy [53], whereas crosstalk of phytohormones, such as abscisic acid (ABA), gibberellin (GA), and auxin, are involved in embryo-imposed dormancy [64,65]. Seed coat-imposed dormancy in particular is involved in the seed survival mechanism of several species [66]. The seed coat exerts its germination-restrictive action by its mechanical resistance to radicle protrusion or being impermeable to water and/or oxygen. These properties are positively correlated with seed coat color due to phenolic compounds in diverse species. In wheat, red-grained genotypes exhibit a wide range of seed dormancy and are more resistant to PHS because they contained dominant alleles in their trigenic series, whereas white-grained cultivars lack seed dormancy at maturity and are susceptible to PHS [63,67,68,69].
It is widely known that abscisic acid (ABA) is the major mediator for seed dormancy because it plays a significant role in inducing and maintaining dormancy during seed development as well as in imbibed seeds [70,71]. Many genes, like TaPHS1 (a TaMFT-like gene), TaCYP707A1, and TaDOG1, have been identified for seed dormancy and are also involved in ABA synthesis and its signal transduction [41,43,72,73,74]. Until now, TaPHS1/TaMFT, TaSdr, PM19-A1/A2, and TaMKK3-A are the cloned genes involved in controlling seed dormancy and PHS resistance in wheat. TaMFT (Mother of FT and TFL1) is a homologue of the Arabidopsis MFT gene which controls seed dormancy and also regulates ABA and GA signal transduction. These studies indicated that wheat and Arabidopsis share the same regulatory mechanism of seed dormancy [41,43,72]. An SNP in the promoter region (at position –222) of TaMFT has identified which may increase MFT expression and likely contributes to increase seed dormancy [41]. Another gene TaPHS1 (homolog of wheat MFT gene) involved in the regulation of seed dormancy and PHS resistance was identified on wheat chromosome 3A while the mutations at +646 and +666 positions of the coding region of TaPHS1 gene resulted in PHS susceptibility [42,43]. It has been reported that Sdr gene plays an intermediate role in inhibiting germination and promoting dormancy in rice [75]. In wheat, three TaSdr genes TaSdr-A1, TaSdr-B1, and TaSdr-D1 have been cloned and are involved in seed dormancy, among them; TaSdr-B1 on chromosome 2B was observed to play a vital role in regulating seed dormancy [46,76]. Barrero et al. [44] identified two candidate genes PM19-A1 and PM19-A2 which positively regulate seed dormancy. They also demonstrated that PM19-A1 highly expressed in dormant genotypes during grain maturation while PM19-A2 showed sequence variations between non-dormant and dormant genotypes. In wheat, another gene MKK3-A (mitogen-activated protein kinase kinase 3), also called TaMKK3-A, has been identified on chromosome 4AL as a candidate gene of the Phs-A1 locus which is associated with the length of seed dormancy [45]. Despite the multi-genic control of seed dormancy, a few major loci, including Phs-A1 on chromosome 4AL and TaPHS1/TaMFT on chromosome 3A, have also shown to involve in sprouting resistance and account for a significant proportion of natural variations in diverse mapping populations [77]. Based on the strong effect, Phs-A1 has been identified in at least 15 multi-parent and bi-parental mapping populations developed from diverse germplasm originated in the United Kingdom, Australia, China, Japan, Mexico, Europe, Canada, and Africa [44,78,79,80,81]. Shorinola et al. [77] studied the physiological evaluation of Phs-A1 during seed after-ripening and observed that it delayed the rate of loss in dormancy when plants were grown across a wide temperature range of 13–22 °C. In another study, Shorinola et al. [82] conducted a haplotype analysis of the Phs-A1 locus and found that TaMKK3-A, but not PM19, was the causal gene underlying variation in sprouting associated with Phs-A1 in diverse Asian, North American, European, and Australian germplasm.
In addition to the PHS-resistant genes identified in wheat, maize could also act as a model to improve PHS resistance in wheat, e.g., the maize viviparous-1 (Vp1) gene. McCarty et al. [83] reported that Vp1 gene encodes a transcription factor that plays a significant role in the regulation of late embryogenesis in maize and late embryo development in bread wheat. TaVp1 genes were extensively studied in wheat and were linked with seed dormancy and PHS resistance [38,39,61,84,85,86,87,88,89]. The TaVp1 genes were mapped about 30 cM from R loci on homologous group 3Lchromosomes [86]. Six TaVp-1B alleles, TaVp-1Ba, TaVp-1Bb, TaVp-1Bc, TaVp-1Bd, TaVp-1Be, and TaVp-1B were identified in wheat [38,39,89,90]. Based on this allelic variation, the STS marker (Vp1B3) was developed for seed dormancy and it was observed that TaVp-1Bb and TaVp-1Bc alleles were linked with higher PHS tolerance [38]. Another STS marker (Vp1A3) was also developed for PHS tolerance by observing greater PHS resistance in various combinations of allelic variations, like TaVp-1Agm/TaVp-1Ba, TaVp-1Agm/TaVp-1Bb, TaVp-1Aam/TaVp-1Bb, and TaVp-1Aim/TaVp-1Bb [61]. Moreover, genetic analyses identified other genes such as DOG1 involved in seed germination and dormancy [91,92,93,94]. The DELAY OF GERMINATION1 (DOG1) gene was first identified in Arabidopsis as a major QTL involved in increased seed dormancy [95]. The length of seed dormancy was estimated by the amount of expression of DOG1 protein in freshly-harvested seeds, which indicated that the DOG1 gene is a timer for the release from dormancy [96]. In a recent study, Nishimura et al. [97] demonstrated that the DOG1 and AHG1 genes interact with multiple environmental factors as well as the PYL/RCAR ABA receptor-mediated regulatory system to establish an important regulatory mechanism for control of seed dormancy and germination. Recent advances in genome sequencing and whole genome assembly of hexaploid wheat will trigger progress in identifying more seed dormancy and PHS resistance genes [98,99,100,101,102].

1.3. α-Amylase Activity

The α-amylase (amy) enzyme is involved in many plant physiological processes such as cold tolerance and germination rate and can hydrolyze α-1,4 -glycosidic bonds in saccharides [103,104]. The expression of amy enzyme is strictly controlled by the phytohormones ABA and gibberellin. ABA inhibits the amy expression during grain development. However, in genetic defect wheat, a high level of high pI amy genes could be expressed, resulting in poor grain quality during late grain development which is referred to as late maturity α-amylase (LMA) [105]. An elevated level of GA promotes amy expression during seed germination [106]. Alpha-amylase activity and PHS resistance are associated with each other possibly due to the fact that increased α-amylase activity upon water absorption promotes seed germination [107,108]. A remarkable difference was found in α-amylase activity between PHS-resistant and -susceptible cultivars in wheat [108]. Of three PHS traits, falling number (FN) [109] was found to indirectly measure the α-amylase enzyme activity that degrades starch in germinating seeds and is an important factor in quantifying PHS [110]. Breakdown of starch due to increased α-amylase activity results in a decreased FN value and is an indirect sign of low seed dormancy and low PHS resistance. Four isozymes of α-amylase affecting PHS have been identified in wheat, including malt-α-amylase (α-amylase-1) on homologous chromosomes 6, green-α-amylase (α-amylase-2) on homologous chromosomes 7, α-amylase-3 on homologous chromosomes 5, and α-amylase-4 has two members on homologous chromosomes 2 and 3 [111,112]. The wheat B genome contains genes for α-amylase-1 and α-amylase-2 among all the three genomes. Promoters of α-amylase-1 gene contains GA responsive complex that consists of a GA-responsive element (CAATAAA), pyrimidine box (CCTTTT), and TATCCAT/C box [112]. GA3 seemed to be involved in regulation of expression level of α-amylase-1 and α-amylase-2 [113]. The α-amylase-1 activity seemed to be significantly correlated with seed dormancy and contributed about 84% to seed germination [111].
In addition to α-amylase variation, α-amylase subtilisin inhibitors (ASI) were also identified in wheat, rice, rye, and barley by limiting α-amylase activity to restrain seed germination [114,115]. Moreover, ten ASI isomerides were identified through monoclonal antibody immune imprinting and isoelectric focusing electrophoresis techniques [116,117]. Yuan et al. [118] reported that PHS tolerance can be increased by reducing the α-amylase activity through combing α-amylase-1 and ASI complex. The α-amylase quantity and activity is very low in dormant seeds and increases after seed germination, therefore, it is necessary to identify the regulatory factors interacting with α-amylase, which can contribute to understand the complicated molecular mechanism of α-amylase regulating PHS tolerance.

1.4. Plant Growth Hormones

Previous studies have described the significance of plant hormones in metabolic and signaling aspects and their probable role in the maintenance and release of dormancy in seeds of cereal crops [43,119,120]. Among plant growth hormones, abscisic acid (ABA) and gibberellin (GA) play important roles in regulation of dormancy and germination, ABA induces dormancy and GA stimulates seed germination [121,122]. A change in balance between ABA and GA levels in seed constitutes a regulatory mechanism that results in maintenance or release of seed dormancy [120,123]. Several studies have reported the regulatory mechanisms of other hormones like ethylene, jasmonate, brassinosteroids, and auxin in controlling seed dormancy, germination and PHS resistance [43,121,124,125]. Environmental factors, such as light and temperature, also affect the dormancy and germination by disturbing the balance between ABA and GA levels in cereal crops [126,127].
Numerous mutants have been developed to understand the regulatory role of plant growth hormones in seed germination and dormancy. Recent advances in the genomics of cereal crops have led to identify many genes involved in metabolic and signaling pathways of plant hormones for regulating seed germination and dormancy. The levels of plant growth hormones are noticeably different in PHS resistant and susceptible varieties; therefore, PHS resistance can be improved by identifying more genes involved in the expression and regulation of plant growth hormones.

1.4.1. Abscisic Acid

Abscisic acid (ABA) is an essential hormone that promotes seed dormancy, seed maturation and tolerance to desiccation [128]. Dormant wheat ABA levels increases by up to 2.5-fold during imbibition but remains unchanged in non-dormant grains [129]. ABA level in seeds/tissues is regulated by its synthesis and catabolism [130]. ABA biosynthesis is catalyzed by numerous enzymes like NCED (9-cis-epoxycarotenoid dioxygenase) that acts as a key regulator of ABA biosynthesis during seed maturation. During ABA biosynthesis, oxidative cleavage of violaxanthin and 9-cis-neoxanthin by NCED is regarded as rate-limiting step [131], whereas ABA catabolism is triggered by ABA 8′-hydroxylase enzyme (ABA8′OH) encoded by CYP707A genes that induce ABA hydroxylation at the 8′ position [132,133,134] (Figure 2). Therefore, the NCED and CYP707A genes play important roles in germination and dormancy by controlling the ABA level in seeds.
Seed development in wheat is characterized by two peaks of ABA accumulation that occur during the mid and late maturation phases. The first peak arises around 25 days after pollination (DAP) while the second peak arises around 35 DAP and extended up to 40 DAP in dormant wheat seeds [135,136], indicating the significance of ABA in inducing embryo dormancy [137]. Moreover, imbibed dormant wheat seeds have shown 3.8-fold higher expression of TaNCED2 than non-dormant imbibed seed, while non-dormant seeds exhibit 2.5-fold higher expression of TaABA8′OH1 (a wheat homolog of CYP707A) than dormant seeds in both imbibed and dry conditions [138]. Mutational analysis of the two homologs of TaABA8′OH1 (TaABA8′OH1A and TaABA8′OH1D) showed an increase in embryonic ABA contents during mid and late stages (40–60 DAA) of seed development resulting in a higher level of seed dormancy [139] and highlighting the importance of higher embryonic ABA levels in inducing seed dormancy during the seed maturation phase in wheat.
ABA has been involved in the regulation of several seed developmental processes like deposition of storage reserves and primary dormancy induction that are evident from the observation of ABA mutants or deficient plants in maize and Arabidopsis [140]. A large number of mutants with reduced ability of synthesizing ABA have been developed in various crops, like the aba1 mutant in Nicotiana plumbaginifolia, viviparous (Vp) mutants Vp5, Vp7, Vp10/Vp13, Vp14, and Vp15 in maize; aba1, aba2, and aba3 mutants in Arabidopsis, and sit, flc, and not mutants in tomato [91,131]. Several knock-out mutations are available for most wheat genes that provide an invaluable resource for characterizing the gene function. The resource of Targeting Induced Local Lesions In Genome (TILLING) mutants, like Kronos (tetraploid) and Cadenza (hexaploid), have been developed in wheat. The exome sequences of 1535 Kronos and 1200 Cadenza mutants have been resequenced using Illumina next-generation sequencing that can be used to screen for mutations in pre-harvest sprouting and dormancy related genes [141].
ABA biosynthetic mutants failed to induce seed dormancy and revealed a wilty vegetative phenotype, e.g., the aba1 mutant in Arabidopsis and aba2 mutant in tobacco were not able to produce zeaxanthin epoxidase (ZEP), the first identified ABA biosynthetic enzyme [142]. Another ABA-deficient mutant aba4 was identified in Arabidopsis during a screening of paclobutrazol resistance germination and has known to be impaired in neoxanthin synthase (NSY) enzyme [143]. The vp14 (viviparous14) mutant in maize and the notabilis mutant in tomato have shown impairment for NCED, which acts as a catalyst for oxidative cleavage of 9’-cis neoxanthin and/or xanthophylls, 9-cis-violaxanthin, and produces xanthoxin, as shown in Figure 2 [144,145]. The vp10 and vp15 mutants in maize, sitiens and flacca mutant in tomato and aba2 and aba3 mutants in Arabidopsis have also shown the impairment in later steps during ABA biosynthetic pathway in the cytosol [146,147].
The role of ABA in seed dormancy of wheat has already been described [16,43]. Nambara et al. [130] reported three core components of ABA signaling in seeds such as protein phosphatase 2Cs (PP2Cs), SNF1-related protein kinase2s (SnRK2s) and pyrabactin resistance/pyrabactin like/regulatory components of ABA receptors (PYR/PYL/RCAR), as shown in Figure 2. ABA forms a complex by binding with its receptor PYR/PYL/RCAR which then interacts with PP2Cs to inhibit its function. The PP2Cs negatively regulate ABA signaling by repressing the SnRK2s activity, which is a positive regulator of downstream targets. Inhibition of PP2Cs causes de-repression of SnRK2s, which in turn phosphorylates and activates down-stream transcriptional factors including ABI3 (B3 type protein), ABI4 (AP2 type transcription factor), ABI5 (abscisic acid insensitive 5), and ABFs (bZIP-type transcription factors). These transcriptional factors are important for the expressional regulation of ABA-responsive genes of seeds [130]. In the absence of ABA, PP2Cs becomes activated and, in turn, dephosphorylate and deactivate the SnRK2s. The molecular components involved in the ABA signaling pathway seem to be conserved in seeds of both monocot and dicot species [148].
The current understanding of signaling elements like ABA that control seed dormancy and germination mainly results from genetic analysis. In wheat, QTL and mutational analysis have revealed the importance of ABA sensitivity in regulating seed dormancy [149,150]. Dormant wheat seeds show more ABA sensitivity than non-dormant seeds [151,152]. Vp1 was the first gene cloned in maize against ABA response [83]. Expression of Vp1 in wheat embryos was positively correlated with ABA sensitivity and degree of seed dormancy [87,153]. Splicing of the Vp1 gene in wheat and rice counterpart resulted in susceptibility to PHS in both species [88,154]. ABI3 is the ortholog of Vp1 in Arabidopsis and the seeds containing Vp1 or ABI3 alleles exhibited similar phenotypes including ABA insensitivity, desiccation intolerance, and premature activation of the shoot apical meristem [155]. Mutational analysis of the ABI4 and ABI5 loci in Arabidopsis showed similar quantitative effects as ABI3 on ABA sensitivity and seed development, but ABI3 null mutations were more destructive than ABI4 and ABI5 [156]. These studies may help to explain the role of ABA in inducing wheat seed dormancy and to understand the molecular mechanisms underlying the regulation of ABA metabolism in inducing dormancy during seed maturation.

1.4.2. Gibberellin

Gibberellin (GA) is another major plant hormone that plays an important role in regulation of seed dormancy and germination [64]. GA breaks the seed dormancy and promotes germination by balancing the primitive endogenous inhibitors [157]. It also regulates the expression of α-amylase synthesis genes involved in seed germination and hydrolyzing the starch in the endosperm. In Arabidopsis, leafy cotyledon 2 (lec2) and fusca3 (fus3) could up-regulate GA activity resulting in germination of seeds before maturity [158,159].
The bioactive concentration of GA in plants is regulated by the balance between its synthesis and inactivation, that are mainly controlled by the genes GA2ox (encoding GA 2-oxidase), GA3ox (encoding GA 3-oxidase), and GA20ox (encoding GA 20-oxidase), respectively [160] (Figure 3). Many genes encoding these enzymes have been identified in a range of crop species including wheat, rice, and barley [160,161] and their expression plays significant roles in dormancy and germination by regulating the GA level in seeds.
Variations at the transcriptional level of these genes orthologs due to after-ripening and in non-dormant and dormant cereal crop seeds indicated the role of GA in regulating the seed germination and dormancy. For instance, dormancy loss in imbibed after-ripened barley and wheat seeds has shown to be linked with increased expression of the TaGA3ox and TaGA20ox genes and a higher level of bioactive GA1 [43,126,162]. Moreover, transcriptional activation of GA20ox gene orthologs induced an increase in the level of GA4 in non-dormant embryos of sorghum seeds, whereas up-regulation of GA2ox gene orthologs led to a decrease in the GA4 level [163]. Mutational studies of these genes will provide further information regarding the molecular mechanisms of GA in regulating seed germination and dormancy. Genetic studies in rice have identified other candidate genes, such as OsGA20ox2 and OsGA2ox3, responsible for regulating seed germination [164,165], while mutational analysis of OsGA20ox2 showed greater dormancy due to reduction in the GA levels in seeds [164].
GA signals in plants are perceived by the soluble receptor protein gene GID1 (Gibberellin insensitive dwarf 1), that was first mapped in rice. Mutational analysis of GID1 in rice showed repression of α-amylase production and had no inhibitory effect on seed germination [166]. Orthologs of GID1 protein have also been identified in wheat but further characterization of functions analysis of GID1 orthologs in seed dormancy is required [167].
DELLA proteins in plants are another important element of GA signaling pathway, which function as a GA repressor and are broken down by ubiquitination induced by GA [168]. GA binds with GIDI and triggers the formation of the GA-GID1-DELLA complex which then interacts with F-box protein (the principal component of SCFSLY 1/GID2 E3 ubiquitin ligase) to degrade the DELLA protein through the ubiquitin-26S proteasome pathway [169,170,171] (Figure 3). In Arabidopsis, five DELLA proteins including RGL1 (RGA like1), RGL2 (RGA like2), RGL3 (RGA like3), GA1 (GA insensitive), and RGA (Repressor of GA1) were identified [172], among them RGL2 is known to be an important seed germination repressor [173]. The DELLA proteins in cereals such as RHT (reduced height) in wheat, SLN1 (slender1) in barley and SLR1 (slender rice1) in rice are transcribed by single genes [174,175,176]. Chandler [177] observed that DELLA mutant seeds of barley were non-dormant and exhibited higher α-amylase activity in the aleurone layer.
In addition to DELLA proteins, other GA-regulated MYB transcriptional regulators (GAMYB) also play a significant role in the GA signaling pathway in aleurone cells of cereal crops [178]. In cereal aleurone, GAMYB triggers the transcriptional activation of GA and regulates hydrolytic enzymes especially α-amylase by directly binding to GA responsive elements in promoter regions [178]. The GAMYB function in cereal aleurone cells was repressed by another GA downstream signaling component named KGM1 (KINASE ASSOCIATED WITH GAMYB1) [179]. Mutational analysis of GAMYB orthologs in rice showed repression of α-amylase gene expression but had no effect on seed germination [180]. In wheat, whole seed transcriptional analysis showed no transcriptional differences in orthologs of GID1, RHT, GAMYB, and KGM1 represented in the GeneChip Wheat Genome Array between after-ripened and dormant seeds. However, transcription of GA responsive genes encoding cell wall hydrolases and amalyses are induced in response to after-ripening [43]. These results might indicate that wheat seeds responsiveness to GA is controlled by post-transcriptional mechanisms or functions without these GA signaling elements. Although these studies demonstrate the role of GAMYB in germination and dormancy, but need further studies to identify and characterize more genes that interact with GAMYB and increase our understanding about the role of downstream GA signaling elements in controlling germination and seed dormancy.

1.4.3. Other Plant Hormones

Previous studies have described the importance of some other plant hormones, like ethylene, jasmonate (JA), auxin, and brassinosteroid (BR), in regulating seed dormancy and PHS resistance [121,124,125]. Transcriptomic analysis of dormant wheat seeds showed that imbibition triggered variations in expression level of several orthologous genes encoding key enzymes involved in ethylene, JA, auxin, and BR biosynthesis and their metabolic pathways due to after-ripening, indicating roles in regulating seed germination and dormancy [43,119]. Jacobsen et al. [181] reported that JA promotes dormancy release in dormant wheat seeds and perform antagonistically to ABA. Transcriptional activation of biosynthetic gene orthologs, such as KAT3 (3-ketoacyl coenzyme a thiolase3), LOX5 (lipoxygenase5), and AOS (allene oxide synthase) involved in JA biosynthesis were detected in imbibed after-ripened seeds. The production of a huge amount of jasmonate-isoleucine along with other orthologs of the biosynthetic pathway during after-ripened seed imbibition revealed the role of JA in controlling wheat seed dormancy. It has been reported that methyl jasmonate decreased the level of seed dormancy in wheat, but its role was regulated by variation in expression levels of the ABA8´OH and NCED1 genes and ABA concentration [181]. Xu et al. [182] studied the role of JA and other hormones in the stratification of wheat dormant seeds and found that JA formation is necessary for seed germination induced by cold stratification. They also concluded that an increase in JA synthesis promoted a reduction in ABA concentration in cold-stratified grain embryos. However, the mechanism of JA in release of seed dormancy to allow germination is still not clear.
Transcriptomic analysis of after-ripened dormant wheat seeds revealed imbibition-mediated activation of BR biosynthetic ortholog genes such as DET2 (De-etiolated 2) and DWF4 (Dwarf 4) [119]. Transcriptional activation of these BR biosynthetic and signaling components in imbibed after-ripened wheat seeds have shown to be associated with transcriptional stimulation of BR responsive orthologs, such as BBE (BR enhanced expression) and PRE (paclobutrazol resistance), that control cell elongation, a process essential for seed germination [183,184]. These results indicated that BR plays a significant role in controlling seed dormancy and germination in wheat.
In Arabidopsis, previous studies have revealed that BR plays a significant role in the regulation of seed dormancy along with ABA and also increased the synthesis of ethylene, which has a regulatory role in seed dormancy of monocot species, such as wild oat [185,186,187]. During ethylene biosynthesis, BR mediates post-transcriptional activation of the ethylene biosynthetic enzymes such as ACC (aminocyclopropane-1- carboxylic acid) synthase (ACS) which acts as a catalyst during the first step [188]. In addition to ACS, ethylene biosynthesis is also mediated by another enzyme ACO (ACC oxidase). In wheat, ACO orthologs are involved in up-regulation in whole imbibed after-ripened seeds compared to dormant seeds [119]. Transcriptional activation of these enzymes and other ethylene receptor orthologs in wheat, such as ERS1 (ethylene response sensor 1), have revealed the role of ethylene in controlling seed germination and dormancy. In other cereal crops like rice and barley, ethylene promotes germination in non-dormant seed but is not involved in the loss of dormancy [189,190]. At present, the role of ethylene in regulating seed dormancy is not clear and needs further investigation of its regulatory role in seed dormancy in different crops.
Recent studies have revealed the role of auxin in maintaining seed dormancy. For instance, exogenous application of auxin increased seed dormancy in wheat [191,192] and Arabidopsis [193,194] through ABA activation. Liu et al. [43] studied the temporal expression patterns of metabolic and signaling genes of ABA, GA, IAA and jasmonate in both dormant and after-ripening dry and imbibed wheat seeds and observed that after-ripening mediated developmental switch from dormancy to germination seems to be linked with declines in seed sensitivity to ABA and IAA and repression of auxin signaling. Mutational analysis of wheat ERA8 (Enhance Response To ABA8) mutant showed an increase in dormancy due to increased level of embryonic jasmonate and aleurone IAA [195]. Metabolomic profiling of two water imbibed wheat cultivars Sukang (dormant) and Baegjoong (non-dormant) detected variable amounts of many auxin-related compounds in the 48h samples of Baejoong and found that indoleacetate abundance was not changed in the Sokang sample but showed an abrupt reduction in Baegjoong at 48h water imbibition. Three catabolites of IAA including indole-3-carboxylate were also detected that showed similar of IAA at 48 h but with two other metabolites, such as indoleacetyl-aspartate and 2-oxindole-3-acetate, showed much higher levels at 48h in Baegjoong [196]. It also regulates several plant growths and developmental processes through the auxin signaling pathway mediated by aux/indole-3-acetic acid (IAA), transport inhibitor response1 (TIR1)/additional F box protein (AFB), and auxin response factors (ARFs) [197,198]. The ARFs are responsible for regulating the expression of a large number of auxin-responsive genes by binding with promoters of auxin response elements (AuxREs) [199,200]. Auxin inhibits seed germination and promotes dormancy through ABA-mediated response by regulating seed specific signaling components of ABA like ABI3, ABI4, and ABI5. Among them, ABI3 is the only transcription factor involved in the regulation of seed dormancy [201,202,203]. Liu et al. [65] observed that auxin and ARF10/16 are involved in the regulation of ABI3 expression which in turn inhibits seed germination and promotes seed dormancy in after-imbibed seeds. The function of ARF10 and ARF16 as positive regulators during the ABA signaling pathway contributes in developing a map of integrated hormone signaling during plant growth and development [204,205]. Auxin-induced seed dormancy seems to be an evolutionary mechanism that inhibits seed germination during unfavorable conditions and might be important for conservation of diversity and evolution in seed plant species [65]. Whether auxin is directly involved in seed dormancy is not clear, and its mechanism in controlling seed dormancy is also still unknown.

1.5. Environmental Factors Affecting PHS Resistance

Environmental factors such as rainfall, temperature and high relative humidity during the grain filling and maturation stages play an important role in the regulation of dormancy and sprouting in wheat. Temperature is one of the most important environmental factors for maintaining dormancy during seed development and for inducing dormancy during seed imbibition [206]. During seed development, low temperatures between 10 and 15 °C can induce high and prolonged dormancy while low temperature during germination breaks dormancy of freshly-harvested wheat seeds [207,208]. At low temperatures the MOTHER OF FT AND TFL1 (MFT) gene is involved in enhancing the dormancy during seed development in wheat [41]. It has also been reported that rainfall 10–20 days before harvest causes little or no sprouting but may influence the crop to be highly susceptible for sprouting at later rainfall [209]. Lunn et al. [210] studied the relationship between mean temperature and period of seed dormancy during the grain development and found that shorter dormancy periods occur after high mean temperatures.
Environmental factors such as temperature fluctuations, salinity and seed moisture content seem to promote ABA synthesis in plants with consequent effects on seed dormancy [211]. Footitt et al. [212] reported that the depth of seed dormancy and gene expression patterns were linked with seasonal variations in soil temperature. They also explained that ABA signaling was linked with deep dormancy during winter while its repression was linked with relief from dormancy during spring. ABA-signaling repression was concurrent with increased DELLA repression of germination. During winter, the expression of NCED6 (ABA synthesis) and GA2ox2 (GA catabolism) genes were found to be increased resulting in increased seed dormancy due to the decrease in soil temperature whereas, during spring, the endogenous ABA biosynthesis was found to be decreased while the expression of CYP707A2 (ABA catabolism) and GA3ox1 (GA synthesis) genes was increased resulting in declined seed dormancy in Arabidopsis [212]. In another study, Kashiwakura et al. [162] used two PHS-tolerant varieties, OS38 (highly dormant) and Gifu-komugi (Gifu, moderately dormant) to characterize the mechanisms of both dormancy maintenance and breakage at low temperatures. They observed that Gifu grains were germinated after imbibition at 15 °C whereas OS38 grains remained dormant. Imbibition of Gifu grains at low temperature caused a reduction in ABA levels in dormant embryos primarily due to the expression of TaABA8′OH1 and TaABA8′OH2 (ABA catabolism) and TaGA3ox2 (GA synthesis) genes resulting in increased GA levels. On the other hand, imbibition of extremely dormant OS38 grains at a low temperature increased ABA levels by inducing the expression of TaNCED (ABA biosynthesis) gene and suppressing TaGA3ox2 and TaABA8′OH2 genes. In a recent study, Izydorczyk et al. [127] observed a delay in germination of non-dormant imbibed wheat seeds under supra-optimal and suboptimal temperatures which was due to the expression of ABA signaling genes; ABI3, ABI5, PYL5, and SnRK2 in the embryo tissues resulted in enhanced ABA sensitivity. These studies explained the role of ABA and GA in dormancy and germination but needs further genetic studies to understand the physiological role of metabolic genes of ABA and GA in the regulation of seed dormancy in cereal crops.

1.6. QTL/genes Identified for PHS Resistance

The genetics of PHS resistance is controlled by both epistatic and additive effects which are affected by environmental conditions. The interaction between QTL epistasis (Q × Q) and the environment (Q × E, Q × Q × E) for PHS resistance has been studied to understand the complex genetic structure of QTL [213,214,215]. In wheat, PHS resistance is controlled jointly by multiple QTLs located an almost 21 chromosomes (1A, 1B, 2A, 2B, 2D, 3A, 3B, 3D, 4A, 4B, 5B, 5D, 6A, 6B, 6D, 7A, 7B and 7D) [1,14,18,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,216,217,218,219,220] (Table 1).
In addition, several candidate genes for PHS resistance have also been identified based on comparative genomics or transcriptomic analysis, such as TaSdr-A1 and TaSdr-B1 on 2AS and 2BS [46,76], TaPHS1 and TaMFT on 3AS [16,41,42], TaVp-1 and Tamyb10 on group 3 chromosomes [38,39,40,84,86], and PM19-A1/A2 and TaMKK3-A on 4AL chromosome [44,45] (Table 2).
The interaction of genes with different PHS resistant QTL is different. QTL positioned on chromosome 4A may interact with R gene controlling red seed color to affect the PHS tolerance. Another QTL for PHS resistance was identified on wheat chromosome 5D independently of an R gene [29]. PHS is a typical quantitative trait controlled by multiple QTL and genes. A major QTL was mapped on the chromosome 4A by using various mapping populations which controlled about 40% of the phenotypic variation in PHS resistance in wheat [3,27,227]. These studies were conducted using SSR markers for the construction of genetic map and QTL mapping. Due to cost-effective and rapid innovations in sequencing technologies, thousands of molecular markers especially SNPs (such as wheat 820K, 660K, and 90K arrays) have been developed in wheat, which are useful for fine-mapping of QTLs and for cloning of candidate genes in the target regions. Moreover, recent advances in genome sequencing and whole genome assembly of hexaploid wheat will provide the bases for rapid identification of various PHS resistance genes [98,99,100,101,102].
PHS resistance is controlled by genotype, environment and their interaction [58]. The use of molecular markers for marker-assisted selection (MAS) could be helpful for direct identification of favorable or deleterious alleles in diverse groups of genotypes [228]. Iyer-Pascuzzi and McCouch [229] reported that MAS can be used for indirect selection of desired traits with considerable reduction in cost and time. Many gene-specific markers, such as SSRs (Xgwm15, Xgwm894, and Xgwm937), STMS markers (wmc104, Xwmc397, and Xwmc468), and STS markers (Vp1-B2 and Vp1B3), were developed for the Vp1 gene and can be used for identification of PHS resistance in diverse genotypes [15,27,34,38,230,231,232,233]. Ogbonnaya et al. [27] found that Xgwm894 and Xgwm937 markers are significantly associated with PHS resistance and could be used for improving PHS resistance in wheat breeding programs. Liu et al. [16] developed an SNP marker named TaPHS1-SNP1 that can be used as diagnostic marker for identifying the resistance allele of TaPHS1 in breeding. Based on SNP flanking sequences on chromosomes 3B, 4A, 7B, and 7D, 18 KASP markers were developed that can be used for PHS resistance in future genetic studies and might also be useful for evaluating the PHS in breeding as well as germplasm materials [80]. In a study, Rasheed et al. [234] developed five KASP based assays of functional markers for four genes, including SDR_SNP for TaSdr-B1, TaMFT-1617R for TaMFT-A1, TaMFT-721J for TaMFT-A1, Vp1B1-83_IND for TaVp-1B, and Vp1B1-193_IND for TaVp-1B. These four genes may have different pathways to induce PHS tolerance; therefore, these KASP assays could provide an excellent opportunity to combine beneficial alleles for PHS tolerance in breeding programs. Wang et al. [235] developed STS (sequence tagged site) marker Tamyb10D for the Tamyb10D1 gene and showed that it can be used as an efficient marker for evaluating the depth of seed dormancy in bread wheat genotypes. Moreover, the CAPS (cleaved amplified polymorphism sequence) marker Sdr2A has also been developed which is positioned on 2.9 cM intervals between Xgwm95 and Xgwm372 markers and can be used for identifying PHS resistant genotypes [46]. With the advancements in sequencing technologies, more than 124,000 gene loci have been annotated, [98,101], which provides a huge base for identifying more genes for PHS resistance and for the development of functional markers linked with PHS resistance, that can be used for developing PHS resistant varieties during wheat molecular breeding programs.

2. Conclusions and Future Prospects

PHS is a complex trait and determined by various endogenous and exogenous factors. Development of PHS resistant varieties is desirable in wheat growing areas especially having long wet weather conditions during harvest. Only a few PHS-resistant cultivars are commercially available in the field, and the grain quality of those cultivars needs to be improved. Therefore, selecting and inserting new resources could also be helpful in developing PHS resistant cultivars.
Understanding the genetics of various factors affecting PHS resistance is necessary for improving PHS resistance in wheat cultivars. The combination of both genetic and genomic technologies should be used to deeply study the temporal and spatial transcription of the genes involved directly or indirectly in controlling PHS resistance. Genomic and post-genomic data will broaden our knowledge about various factors affecting PHS resistance.
Construction of mutant libraries is important for future studies. Map-based cloning and mutant analysis of the genes underlying PHS resistance will provide new insights in improving PHS resistance of crop species. It is also necessary to use the available crop genome database that will trigger the progress in this field. Furthermore, rapid advancements in molecular technologies, like next-generation sequencing (NGS) technologies [236], and ongoing chromosomal-based and wheat whole genome sequencing projects (International Wheat Genome Sequencing Consortium, IWGSC) [98,99,100,101,102] will provide new opportunities for identification and functional analysis of the candidate genes controlling PHS resistance.

Author Contributions

A.A. and H.-P.Z. wrote the manuscript. S.W.S. and L.S. drew all the figures in the manuscript. J.C. and H.J. drew all the tables in the manuscript. C.C. and C.M. reviewed and edited the manuscript.

Funding

National Natural Science Foundation of China: 31401372, National Key Research and Development Program: 2017YFD0100703, China Agriculture Research System: CARS-03

Acknowledgments

This work was supported by National key research and development program “Creation of new breeding material with stress resistance and new wheat variety breeding” (2017YFD0100703); grants from the National Natural Science Foundation of China (31401372); China Agriculture Research System (CARS-03); Wheat Genetics and Breeding Research Platform Innovation Team of Anhui’s University, the Promotion Project of High Education of Anhui Province “Team construction of high level teacher of crop discipline”; Jiangsu Collaborative Innovation Center for Modern Crop Production (JCIC-MCP); the Agriculture Research System of Anhui Province (AHCYTX-02); and the Introduced Leading Talent Research Team for Universities in Anhui Province.

Conflicts of Interest

The authors declare that they have no conflict of interests.

References

  1. Groos, C.; Gay, G.; Perretant, M.-R.; Gervais, L.; Bernard, M.; Dedryver, F.; Charmet, G. Study of the relationship between pre-harvest sprouting and grain color by quantitative trait loci analysis in a white × red grain bread-wheat cross. Theor. Appl. Genet. 2002, 104, 39–47. [Google Scholar] [CrossRef] [PubMed]
  2. Kulwal, P.; Ishikawa, G.; Benscher, D.; Feng, Z.; Yu, L.X.; Jadhav, A.; Mehetre, S.; Sorrells, M.E. Association mapping for pre-harvest sprouting resistance in white winter wheat. Theor. Appl. Genet. 2012, 125, 793–805. [Google Scholar] [CrossRef] [PubMed]
  3. Mares, D.; Mrva, K.; Cheong, J.; Williams, K.; Watson, B.; Storlie, E.; Sutherland, M.; Zou, Y. A QTL located on chromosome 4A associated with dormancy in white- and red-grained wheats of diverse origin. Theor. Appl. Genet. 2005, 111, 1357–1364. [Google Scholar] [CrossRef] [PubMed]
  4. Sorrells, M.; Sherman, J. Facts: Pre—Harvest Sprouting. MAS Wheat, 2007. Available online: https://maswheat.ucdavis.edu/ (accessed on 2 January 2019).
  5. Andreoli, C.; Bassoi, M.C.; Brunetta, D. Genetic control of seed dormancy and pre-harvest sprouting in wheat. Sci. Agric. 2006, 63, 564–566. [Google Scholar] [CrossRef] [Green Version]
  6. Simsek, S.; Ohm, J.B.; Lu, H.; Rugg, M.; Berzonsky, W.; Alamri, M.S.; Mergoum, M. Effect of pre-Harvest sprouting on physicochemical properties of starch in wheat. Foods 2014, 3, 194–207. [Google Scholar] [CrossRef] [PubMed]
  7. Bewley, J.D.; Black, M.; Halmer, P. The encyclopedia of seeds science. In Technology and Uses; CABI Publishing: Oxfordshire, UK, 2006; p. 528. [Google Scholar]
  8. Rajjou, L.; Duval, M.; Gallardo, K.; Catusse, J.; Bally, J.; Job, C.; Job, D. Seed germination and vigor. Annu. Rev. Plant Biol. 2012, 63, 507–533. [Google Scholar] [CrossRef] [PubMed]
  9. Xiao, S.-H.; Zhang, X.-Y.; Yan, C.-S.; Lin, H. Germplasm improvement for preharvest sprouting resistance in Chinese white-grained wheat: An overview of the current strategy. Euphytica 2002, 126, 35–38. [Google Scholar] [CrossRef]
  10. Kruger, J.E. Biochemistry of preharvest sprouting in cereals. In Preharvest Field Sprouting in Cereals; Derera, N.F., Ed.; CRC Press, Inc.: Boca Raton, FL, USA, 1989; pp. 61–84. [Google Scholar]
  11. Wahl, T.I.; O’Rourke, A.D. The economics of sprout damage in wheat. In Preharvest Field Sprouting in Cereals; Walker-Simmonds, M.K., Ried, J.L., Eds.; CRC Press: Boca Raton, FL, USA; American Association of Cereal Chemists: St. Paul, MN, USA, 1993; pp. 10–17. [Google Scholar]
  12. Lan, X.; Zheng, Y.; Ren, X.; Liu, D.; Wei, Y.; Yan, Z. Utilization of preharvest sprouting tolerance gene of synthetic wheat RSP. J. Plant Genet. Resour. 2005, 6, 204–209. [Google Scholar]
  13. Lin, R.; Horsley, R.D.; Schwarz, P.B. Associations between caryopsis dormancy, α-amylase activity, and pre-harvest sprouting in barley. J. Cereal Sci. 2008, 48, 446–456. [Google Scholar] [CrossRef]
  14. Munkvold, J.D.; Tanaka, J.; Benscher, D.; Sorrells, M.E. Mapping quantitative trait loci for preharvest sprouting resistance in white wheat. Theor. Appl. Genet. 2009, 119, 1223–1235. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, J.H.; Yu, Y.X.; Cheng, J.; Tan, X.L.; Shen, W.P. Study on the Pre-harvest Sprouting Tolerance in Triticum aestivum ssp. Yunnanense King. J. Triticeae Crops 2011, 31, 747–752. [Google Scholar]
  16. Liu, S.; Sehgal, S.K.; Li, J.; Lin, M.; Trick, H.N.; Yu, J.; Gill, B.S.; Bai, G. Cloning and characterization of a critical regulator for preharvest sprouting in wheat. Genetics 2013, 195, 263–273. [Google Scholar] [CrossRef] [PubMed]
  17. Gao, F.; Ayele, B.T. Functional genomics of seed dormancy in wheat: Advances and prospects. Front. Plant Sci. 2014, 5, 458. [Google Scholar] [CrossRef] [PubMed]
  18. Mares, D.J.; Mrva, K. Wheat grain preharvest sprouting and late maturity alpha-amylase. Planta 2014, 240, 1167–1178. [Google Scholar] [CrossRef] [PubMed]
  19. King, R.W.; von Wettstein-Knowles, P. Epicuticular waxes and regulation of ear wetting and pre-harvest sprouting in barley and wheat. Euphytica 2000, 112, 157–166. [Google Scholar] [CrossRef]
  20. Gatford, K.T.; Eastwood, R.F.; Halloran, G.M. Germination inhibitors in bracts surrounding the grain of Triticum tauschii. Funct. Plant Biol. 2002, 29, 881–890. [Google Scholar] [CrossRef]
  21. Kato, K.; Nakamura, W.; Tabiki, T.; Miura, H.; Sawada, S. Detection of loci controlling seed dormancy on group 4 chromosomes of wheat and comparative mapping with rice and barley genomes. Theor. Appl. Genet. 2001, 102, 980–985. [Google Scholar] [CrossRef]
  22. Flintham, J.; Adlam, R.; Bassoi, M.; Holdsworth, M.; Gale, M. Mapping genes for resistance to sprouting damage in wheat. Euphytica 2002, 126, 39–45. [Google Scholar] [CrossRef]
  23. Osa, M.; Kato, K.; Mori, M.; Shindo, C.; Torada, A.; Miura, H. Mapping QTLs for seed dormancy and the Vp1 homologue on chromosome 3A in wheat. Theor. Appl. Genet. 2003, 106, 1491–1496. [Google Scholar] [CrossRef] [PubMed]
  24. Kulwal, P.L.; Kumar, N.; Gaur, A.; Khurana, P.; Khurana, J.P.; Tyagi, A.K.; Balyan, H.S.; Gupta, P.K. Mapping of a major QTL for pre-harvest sprouting tolerance on chromosome 3A in bread wheat. Theor. Appl. Genet. 2005, 111, 1052–1059. [Google Scholar] [CrossRef] [PubMed]
  25. Mori, M.; Uchino, N.; Chono, M.; Kato, K.; Miura, H. Mapping QTLs for grain dormancy on wheat chromosome 3A and the group 4 chromosomes, and their combined effect. Theor. Appl. Genet. 2005, 110, 1315–1323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kottearachchi, N.S.; Uchino, N.; Kato, K.; Miura, H. Increased grain dormancy in white-grained wheat by introgression of preharvest sprouting tolerance QTLs. Euphytica 2006, 152, 421–428. [Google Scholar] [CrossRef] [Green Version]
  27. Ogbonnaya, F.C.; Imtiaz, M.; Ye, G.; Hearnden, P.R.; Hernandez, E.; Eastwood, R.F.; van Ginkel, M.; Shorter, S.C.; Winchester, J.M. Genetic and QTL analyses of seed dormancy and preharvest sprouting resistance in the wheat germplasm CN10955. Theor. Appl. Genet. 2008, 116, 891–902. [Google Scholar] [CrossRef] [PubMed]
  28. Ren, X.B.; Lan, X.J.; Liu, D.C.; Wang, J.L.; Zheng, Y.L. Mapping QTLs for pre-harvest sprouting tolerance on chromosome 2D in a synthetic hexaploid wheat x common wheat cross. J. Appl. Genet. 2008, 49, 333–341. [Google Scholar] [PubMed]
  29. Fofana, B.; Humphreys, D.G.; Rasul, G.; Cloutier, S.; Brûlé-Babel, A.; Woods, S.; Lukow, O.M.; Somers, D.J. Mapping quantitative trait loci controlling pre-harvest sprouting resistance in a red × white seeded spring wheat cross. Euphytica 2009, 165, 509–521. [Google Scholar] [CrossRef]
  30. Kumar, A.; Kumar, J.; Singh, R.; Garg, T.; Chhuneja, P.; Balyan, H.S.; Gupta, P.K. QTL analysis for grain colour and pre-harvest sprouting in bread wheat. Plant Sci. 2009, 177, 114–122. [Google Scholar] [CrossRef]
  31. Mares, D.; Rathjen, J.; Mrva, K.; Cheong, J. Genetic and environmental control of dormancy in white-grained wheat (Triticum aestivum L.). Euphytica 2009, 168, 311–318. [Google Scholar] [CrossRef]
  32. Mohan, A.; Kulwal, P.; Singh, R.; Kumar, V.; Mir, R.R.; Kumar, J.; Prasad, M.; Balyan, H.S.; Gupta, P.K. Genome-wide QTL analysis for pre-harvest sprouting tolerance in bread wheat. Euphytica 2009, 168, 319–329. [Google Scholar] [CrossRef]
  33. Zhu, Z.L.; Tian, B.; Liu, B.; Xie, Q.G.; Tian, J.C. Quantitative Trait Loci Analysis for Pre-harvest Sprouting Using Intact Spikes in Wheat (Triticum aestivum L.). Shandong Agric. Sci. 2010, 6, 19–23. [Google Scholar]
  34. Zhang, H.P.; Feng, J.M.; Chang, C. Investigation of main loci contributing to strong seed dormancy of Chinese wheat landrace. J. Agric. Biotechnol. 2011, 19, 270–277. [Google Scholar]
  35. Knox, R.E.; Clarke, F.R.; Clarke, J.M.; Fox, S.L.; DePauw, R.M.; Singh, A.K. Enhancing the identification of genetic loci and transgressive segregants for preharvest sprouting resistance in a durum wheat population. Euphytica 2012, 186, 193–206. [Google Scholar] [CrossRef]
  36. Kumar, S.; Knox, R.E.; Clarke, F.R.; Pozniak, C.J.; DePauw, R.M.; Cuthbert, R.D.; Fox, S. Maximizing the identification of QTL for pre-harvest sprouting resistance using seed dormancy measures in a white-grained hexaploid wheat population. Euphytica 2015, 205, 287–309. [Google Scholar] [CrossRef] [Green Version]
  37. Zhou, Y.; Tang, H.; Cheng, M.-P.; Dankwa, K.O.; Chen, Z.-X.; Li, Z.-Y.; Gao, S.; Liu, Y.-X.; Jiang, Q.-T.; Lan, X.-J.; et al. Genome-Wide Association Study for Pre-harvest Sprouting Resistance in a Large Germplasm Collection of Chinese Wheat Landraces. Front. Plant Sci. 2017, 8, 401. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, Y.; Zhao, X.L.; Xia, L.Q.; Chen, X.M.; Xia, X.C.; Yu, Z.; He, Z.H.; Roder, M. Development and validation of a Viviparous-1 STS marker for pre-harvest sprouting tolerance in Chinese wheats. Theor. Appl. Genet. 2007, 115, 971–980. [Google Scholar] [CrossRef] [PubMed]
  39. Chang, C.; Feng, J.M.; Si, H.Q.; Yin, B.; Zhang, H.P.; Ma, C.X. Validating a novel allele of viviparous-1 (Vp-1Bf) associated with high seed dormancy of Chinese wheat landrace, Wanxianbaimaizi. Mol. Breed. 2010, 25, 517–525. [Google Scholar] [CrossRef]
  40. Himi, E.; Maekawa, M.; Miura, H.; Noda, K. Development of PCR markers for Tamyb10 related to R-1, red grain color gene in wheat. Theor. Appl. Genet. 2011, 122, 1561–1576. [Google Scholar] [CrossRef] [PubMed]
  41. Nakamura, S.; Abe, F.; Kawahigashi, H.; Nakazono, K.; Tagiri, A.; Matsumoto, T.; Utsugi, S.; Ogawa, T.; Handa, H.; Ishida, H.; et al. Wheat Homolog of MOTHER OF FT AND TFL1 Acts in the Regulation of Germination. Plant Cell 2011, 23, 3215–3229. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, S.; Sehgal, S.K.; Lin, M.; Li, J.; Trick, H.N.; Gill, B.S.; Bai, G. Independent mis-splicing mutations in TaPHS1 causing loss of preharvest sprouting (PHS) resistance during wheat domestication. New Phytol. 2015, 208, 928–935. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, A.; Gao, F.; Kanno, Y.; Jordan, M.C.; Kamiya, Y.; Seo, M.; Ayele, B.T. Regulation of Wheat Seed Dormancy by After-Ripening Is Mediated by Specific Transcriptional Switches That Induce Changes in Seed Hormone Metabolism and Signaling. PLoS ONE 2013, 8, e56570. [Google Scholar] [CrossRef] [PubMed]
  44. Barrero, J.M.; Cavanagh, C.; Verbyla, K.L.; Tibbits, J.F.G.; Verbyla, A.P.; Huang, B.E.; Rosewarne, G.M.; Stephen, S.; Wang, P.; Whan, A.; et al. Transcriptomic analysis of wheat near-isogenic lines identifies PM19-A1 and A2 as candidates for a major dormancy QTL. Genome Biol. 2015, 16, 93. [Google Scholar] [CrossRef] [PubMed]
  45. Torada, A.; Koike, M.; Ogawa, T.; Takenouchi, Y.; Tadamura, K.; Wu, J.; Matsumoto, T.; Kawaura, K.; Ogihara, Y. A Causal Gene for Seed Dormancy on Wheat Chromosome 4A Encodes a MAP Kinase Kinase. Curr. Biol. 2016, 26, 782–787. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, Y.; Xia, X.; He, Z. The seed dormancy allele TaSdr-A1a associated with pre-harvest sprouting tolerance is mainly present in Chinese wheat landraces. Theor. Appl. Genet. 2017, 130, 81–89. [Google Scholar] [CrossRef] [PubMed]
  47. Metzger, R.J.; Silbaugh, B.A. Location of genes for seed coat color in hexaploid wheat (Triticum aestivum L.). Crop Sci. 1970, 10, 495–496. [Google Scholar] [CrossRef]
  48. Wang, D.; Dowell, F.E.; Lacey, R.E. Predicting the Number of Dominant R Alleles in Single Wheat Kernels Using Visible and Near-Infrared Reflectance Spectra. Cereal Chem. 1999, 76, 6–8. [Google Scholar] [CrossRef]
  49. Himi, E.; Noda, K. Red grain colour gene (R) of wheat is a Myb-type transcription factor. Euphytica 2005, 143, 239–242. [Google Scholar] [CrossRef]
  50. Chopra, S.; Athma, P.; Peterson, T. Alleles of the maize P gene with distinct tissue specificities encode Myb-homologous proteins with C-terminal replacements. Plant Cell 1996, 8, 1149–1158. [Google Scholar] [CrossRef] [PubMed]
  51. Mol, J.; Grotewold, E.; Koes, R. How genes paint flowers and seeds. Trends Plant Sci. 1998, 3, 212–217. [Google Scholar] [CrossRef]
  52. Kohyama, N.; Chono, M.; Nakagawa, H.; Matsuo, Y.; Ono, H.; Matsunaka, H. Flavonoid compounds related to seed coat color of wheat. Biosci. Biotechnol. Biochem. 2017, 81, 2112–2118. [Google Scholar] [CrossRef] [PubMed]
  53. Himi, E.; Mares, D.J.; Yanagisawa, A.; Noda, K. Effect of grain colour gene (R) on grain dormancy and sensitivity of the embryo to abscisic acid (ABA) in wheat. J. Exp. Bot. 2002, 53, 1569–1574. [Google Scholar] [CrossRef] [PubMed]
  54. Warner, R.L.; Kudrna, D.A.; Spaeth, S.C.; Jones, S.S. Dormancy in white-grain mutants of Chinese Spring wheat (Triticum aestivum L.). Seed Sci. Res. 2007, 10, 51–60. [Google Scholar] [CrossRef]
  55. Lin, M.; Zhang, D.; Liu, S.; Zhang, G.; Yu, J.; Fritz, A.K.; Bai, G. Genome-wide association analysis on pre-harvest sprouting resistance and grain color in U.S. winter wheat. BMC Genom. 2016, 17, 794. [Google Scholar] [CrossRef] [PubMed]
  56. Chouard, P. Vernalization and its Relations to Dormancy. Annu. Rev. Plant Physiol. 1960, 11, 191–238. [Google Scholar] [CrossRef]
  57. Belderok, B. Seed dormancy problems in cereals. Field Crop Abstr. 1968, 21, 203–211. [Google Scholar]
  58. Marzougui, S.; Sugimoto, K.; Yamanouchi, U.; Shimono, M.; Hoshino, T.; Hori, K.; Kobayashi, M.; Ishiyama, K.; Yano, M. Mapping and characterization of seed dormancy QTLs using chromosome segment substitution lines in rice. Theor. Appl. Genet. 2012, 124, 893–902. [Google Scholar] [CrossRef] [PubMed]
  59. Gubler, F.; Millar, A.A.; Jacobsen, J.V. Dormancy release, ABA and pre-harvest sprouting. Curr. Opin. Plant Biol. 2005, 8, 183–187. [Google Scholar] [CrossRef] [PubMed]
  60. Sun, Y.W.; Jones, H.D.; Yang, Y.; Dreisigacker, S.; Li, S.M.; Chen, X.M.; Shewry, P.R.; Xia, L.Q. Haplotype analysis of Viviparous-1 gene in CIMMYT elite bread wheat germplasm. Euphytica 2012, 186, 25–43. [Google Scholar] [CrossRef]
  61. Yang, Y.; Zhang, C.L.; Liu, S.X.; Sun, Y.Q.; Meng, J.Y.; Xia, L.Q. Characterization of the rich haplotypes of Viviparous-1A in Chinese wheats and development of a novel sequence-tagged site marker for pre-harvest sprouting resistance. Mol. Breed. 2014, 33, 75–88. [Google Scholar] [CrossRef]
  62. Belderok, B. Physiological-biochemical aspects of dormancy in wheat. Cereal Res. Commun. 1976, 4, 133–137. [Google Scholar]
  63. Freed, R.D.; Everson, E.H.; Ringlund, K.; Gullord, M. Seed coat color in wheat and the relationship to seed dormancy at maturity. Cereal Res. Commun. 1976, 4, 147–149. [Google Scholar]
  64. Finch-Savage, W.E.; Leubner-Metzger, G. Seed dormancy and the control of germination. New Phytol. 2006, 171, 501–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Liu, X.; Zhang, H.; Zhao, Y.; Feng, Z.; Li, Q.; Yang, H.-Q.; Luan, S.; Li, J.; He, Z.-H. Auxin controls seed dormancy through stimulation of abscisic acid signaling by inducing ARF-mediated ABI3 activation in Arabidopsis. Proc. Natl. Acad. Sci. USA 2013, 110, 15485–15490. [Google Scholar] [CrossRef] [PubMed]
  66. Kelly, K.M.; Van Staden, J.; Bell, W.E. Seed coat structure and dormancy. Plant Growth Regul. 1992, 11, 201–209. [Google Scholar] [CrossRef]
  67. Gfeller, F.; Svejda, F. Inheritance of post-harvest seed dormancy and kernal colour in spring wheat lines. Can. J. Plant Sci. 1960, 40, 1–6. [Google Scholar] [CrossRef]
  68. Everson, E. Varietal variation for dormancy in mature wheat. Q. Bull. Mich. St. Univ. Agric. Exp. Stn. 1961, 43, 820–829. [Google Scholar]
  69. McEwan, J.M. The sprouting reaction of stocks with single genes for red grain colour derived from hilgendorf 61 wheat. Cereal Res. Commun. 1980, 8, 261–264. [Google Scholar]
  70. Jacobsen, J.V.; Pearce, D.W.; Poole, A.T.; Pharis, R.P.; Mander, L.N. Abscisic acid, phaseic acid and gibberellin contents associated with dormancy and germination in barley. Physiol. Plant. 2002, 115, 428–441. [Google Scholar] [CrossRef] [PubMed]
  71. Koornneef, M.; Bentsink, L.; Hilhorst, H. Seed dormancy and germination. Curr. Opin. Plant Biol. 2002, 5, 33–36. [Google Scholar] [CrossRef] [Green Version]
  72. Lei, L.; Zhu, X.; Wang, S.; Zhu, M.; Carver, B.F.; Yan, L. TaMFT-A1 Is Associated with Seed Germination Sensitive to Temperature in Winter Wheat. PLoS ONE 2013, 8, e73330. [Google Scholar] [CrossRef] [PubMed]
  73. Ashikawa, I.; Mori, M.; Nakamura, S.; Abe, F. A transgenic approach to controlling wheat seed dormancy level by using Triticeae DOG1-like genes. Transgen. Res. 2014, 23, 621–629. [Google Scholar] [CrossRef] [PubMed]
  74. Murphey, M.; Kovach, K.; Elnacash, T.; He, H.; Bentsink, L.; Donohue, K. DOG1-imposed dormancy mediates germination responses to temperature cues. Environ. Exp. Bot. 2015, 112, 33–43. [Google Scholar] [CrossRef]
  75. Sugimoto, K.; Takeuchi, Y.; Ebana, K.; Miyao, A.; Hirochika, H.; Hara, N.; Ishiyama, K.; Kobayashi, M.; Ban, Y.; Hattori, T.; et al. Molecular cloning of Sdr4, a regulator involved in seed dormancy and domestication of rice. Proc. Natl. Acad. Sci. USA 2010, 107, 5792–5797. [Google Scholar] [CrossRef] [PubMed]
  76. Zhang, Y.; Miao, X.; Xia, X.; He, Z. Cloning of seed dormancy genes (TaSdr) associated with tolerance to pre-harvest sprouting in common wheat and development of a functional marker. Theor. Appl. Genet. 2014, 127, 855–866. [Google Scholar] [CrossRef] [PubMed]
  77. Shorinola, O.; Bird, N.; Simmonds, J.; Berry, S.; Henriksson, T.; Jack, P.; Werner, P.; Gerjets, T.; Scholefield, D.; Balcárková, B.; et al. The wheat Phs-A1 pre-harvest sprouting resistance locus delays the rate of seed dormancy loss and maps 0.3 cM distal to the PM-19 genes in UK germplasm. J. Exp. Bot. 2016, 67, 4169–4178. [Google Scholar] [CrossRef] [PubMed]
  78. Ogbonnaya, F.C.; Imtiaz, M.; Depauw, R.M. Haplotype diversity of preharvest sprouting QTLs in wheat. Genome Biol. 2007, 50, 107–118. [Google Scholar] [CrossRef] [PubMed]
  79. Torada, A.; Koike, M.; Ikeguchi, S.; Tsutsui, I. Mapping of a major locus controlling seed dormancy using backcrossed progenies in wheat (Triticum aestivum L.). Genome 2008, 51, 426–432. [Google Scholar] [CrossRef] [PubMed]
  80. Cabral, A.L.; Jordan, M.C.; McCartney, C.A.; You, F.M.; Humphreys, D.G.; MacLachlan, R.; Pozniak, C.J. Identification of candidate genes, regions and markers for pre-harvest sprouting resistance in wheat (Triticum aestivum L.). BMC Plant Biol. 2014, 14, 340. [Google Scholar] [CrossRef] [PubMed]
  81. Sydenham, S.L.; Barnard, A. Targeted Haplotype Comparisons between South African Wheat Cultivars Appear Predictive of Pre-harvest Sprouting Tolerance. Front. Plant Sci. 2018, 9, 63. [Google Scholar] [CrossRef] [PubMed]
  82. Shorinola, O.; Balcárková, B.; Hyles, J.; Tibbits, J.F.G.; Hayden, M.J.; Holušova, K.; Valárik, M.; Distelfeld, A.; Torada, A.; Barrero, J.M.; et al. Haplotype Analysis of the Pre-harvest Sprouting Resistance Locus Phs-A1 Reveals a Causal Role of TaMKK3-A in Global Germplasm. Front. Plant Sci. 2017, 8, 1555. [Google Scholar] [CrossRef] [PubMed]
  83. McCarty, D.R.; Hattori, T.; Carson, C.B.; Vasil, V.; Lazar, M.; Vasil, I.K. The Viviparous-1 developmental gene of maize encodes a novel transcriptional activator. Cell 1991, 66, 895–905. [Google Scholar] [CrossRef]
  84. Chang, C.; Zhang, H.-P.; Zhao, Q.-X.; Feng, J.-M.; Si, H.-Q.; Lu, J.; Ma, C.-X. Rich allelic variations of Viviparous-1A and their associations with seed dormancy/pre-harvest sprouting of common wheat. Euphytica 2011, 179, 343–353. [Google Scholar] [CrossRef]
  85. Hattori, T.; Vasil, V.; Rosenkrans, L.; Hannah, L.C.; McCarty, D.R.; Vasil, I.K. The Viviparous-1 gene and abscisic acid activate the C1 regulatory gene for anthocyanin biosynthesis during seed maturation in maize. Genes Dev. 1992, 6, 609–618. [Google Scholar] [CrossRef] [PubMed]
  86. Bailey, P.C.; McKibbin, R.S.; Lenton, J.R.; Holdsworth, M.J.; Flintham, J.E.; Gale, M.D. Genetic map locations for orthologous Vp1 genes in wheat and rice. Theor. Appl. Genet. 1999, 98, 281–284. [Google Scholar] [CrossRef]
  87. Nakamura, S.; Toyama, T. Isolation of a VP1 homologue from wheat and analysis of its expression in embryos of dormant and non-dormant cultivars. J. Exp. Bot. 2001, 52, 875–876. [Google Scholar] [CrossRef] [PubMed]
  88. McKibbin, R.S.; Wilkinson, M.D.; Bailey, P.C.; Flintham, J.E.; Andrew, L.M.; Lazzeri, P.A.; Gale, M.D.; Lenton, J.R.; Holdsworth, M.J. Transcripts of Vp-1 homeologues are misspliced in modern wheat and ancestral species. Proc. Natl. Acad. Sci. USA 2002, 99, 10203–10208. [Google Scholar] [CrossRef] [PubMed]
  89. Xia, L.Q.; Ganal, M.W.; Shewry, P.R.; He, Z.H.; Yang, Y.; Röder, M.S. Exploiting the diversity of Viviparous-1 gene associated with pre-harvest sprouting tolerance in European wheat varieties. Euphytica 2008, 159, 411–417. [Google Scholar] [CrossRef]
  90. Yang, Y.; Chen, X.; He, Z.; Röder, M.; Xia, L. Distribution of Vp-1 alleles in Chinese white-grained landraces, historical and current wheat cultivars. Cereal Res. Commun. 2009, 37, 169–177. [Google Scholar] [CrossRef]
  91. Bentsink, L.; Koornneef, M. Seed Dormancy and Germination. Am. Soc. Plant Biol. 2008, 6, e0119. [Google Scholar] [CrossRef] [PubMed]
  92. Graeber, K.; Nakabayashi, K.; Miatton, E.; Leubner-Metzger, G.; Soppe, W.J. Molecular mechanisms of seed dormancy. Plant Cell Environ. 2012, 35, 1769–1786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Finkelstein, R. Abscisic acid synthesis and response. Arab Book 2013, 11, e0166. [Google Scholar] [CrossRef] [PubMed]
  94. Nonogaki, H. Seed dormancy and germination-emerging mechanisms and new hypotheses. Front. Plant Sci. 2014, 5, 233. [Google Scholar] [CrossRef] [PubMed]
  95. Bentsink, L.; Jowett, J.; Hanhart, C.J.; Koornneef, M. Cloning of DOG1, a quantitative trait locus controlling seed dormancy in Arabidopsis. Proc. Natl Acad. Sci. USA 2006, 103, 17042–17047. [Google Scholar] [CrossRef] [PubMed]
  96. Nakabayashi, K.; Bartsch, M.; Xiang, Y.; Miatton, E.; Pellengahr, S.; Yano, R.; Seo, M.; Soppe, W.J. The time required for dormancy release in Arabidopsis is determined by DELAY OF GERMINATION1 protein levels in freshly harvested seeds. Plant Cell 2012, 24, 2826–2838. [Google Scholar] [CrossRef] [PubMed]
  97. Nishimura, N.; Tsuchiya, W.; Moresco, J.J.; Hayashi, Y.; Satoh, K.; Kaiwa, N.; Irisa, T.; Kinoshita, T.; Schroeder, J.I.; Yates, J.R.; et al. Control of seed dormancy and germination by DOG1-AHG1 PP2C phosphatase complex via binding to heme. Nat. Commun. 2018, 9, 2132. [Google Scholar] [CrossRef] [PubMed]
  98. International Wheat Genome Sequencing Consortium (IWGSC). A chromosome-based draft sequence of the hexaploid bread wheat (Triticum aestivum) genome. Science 2014, 345, 1251788. [Google Scholar] [CrossRef] [PubMed]
  99. Chapman, J.A.; Mascher, M.; Buluç, A.; Barry, K.; Georganas, E.; Session, A.; Strnadova, V.; Jenkins, J.; Sehgal, S.; Oliker, L.; et al. A whole-genome shotgun approach for assembling and anchoring the hexaploid bread wheat genome. Genome Biol. 2015, 16, 26. [Google Scholar] [CrossRef] [PubMed]
  100. Clavijo, B.J.; Venturini, L.; Schudoma, C.; Accinelli, G.G.; Kaithakottil, G.; Wright, J.; Borrill, P.; Kettleborough, G.; Heavens, D.; Chapman, H.; et al. An improved assembly and annotation of the allohexaploid wheat genome identifies complete families of agronomic genes and provides genomic evidence for chromosomal translocations. Genome Res. 2017, 27, 885–896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. International Wheat Genome Sequencing Consortium (IWGSC); Appels, R.; Eversole, K.; Feuillet, C.; Keller, B.; Rogers, J.; Stein, N.; Pozniak, C.J.; Choulet, F.; Distelfeld, A.; et al. Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science 2018, 361, eaar7191. [Google Scholar] [CrossRef] [PubMed]
  102. Ramírez-González, R.H.; Borrill, P.; Lang, D.; Harrington, S.A.; Brinton, J.; Venturini, L.; Davey, M.; Jacobs, J.; van Ex, F.; Pasha, A.; et al. The transcriptional landscape of polyploid wheat. Science 2018, 361, eaar6089. [Google Scholar] [CrossRef] [PubMed]
  103. Autio, K.; Simoinen, T.; Suortti, T.; Salmenkallio-Marttila, M.; Lassila, K.; Wilhelmson, A. Structural and Enzymic Changes in Germinated Barley and Rye. J. Inst. Brew. 2001, 107, 19–25. [Google Scholar] [CrossRef] [Green Version]
  104. Masojć, P.; Milczarski, P. Relationship between QTLs for preharvest sprouting and alpha-amylase activity in rye grain. Mol. Breed. 2009, 23, 75–84. [Google Scholar] [CrossRef]
  105. Barrero, J.M.; Mrva, K.; Talbot, M.J.; White, R.G.; Taylor, J.; Gubler, F.; Mares, D.J. Genetic, hormonal and physiological analysis of late maturity alpha-amylase (LMA) in wheat. Plant Physiol. 2013, 161, 1265–1277. [Google Scholar] [CrossRef] [PubMed]
  106. Woodger, F.; Jacobsen, J.V.; Gubler, F. Gibberellin action in germinated cereal grains. In Plant Hormones: Biosynthesis, Signal Transduction, Action; Davies, P.J., Ed.; Springer: Dordrecht, The Netherlands, 2010; pp. 221–240. [Google Scholar]
  107. Wu, Y.; Hu, H.; Wang, G.; Zhang, Y.; Ji, J. Relationship between alpha amylase activity and resistance of pre-harvest sprouting in spring wheat. J. Jilin Agric. Univ. 2002, 24, 22–25. [Google Scholar]
  108. Wang, X.G.; Ren, J.P.; Yin, J. The mechanism on wheat pre-harvest resistant sprouting. China Agric. Sci. 2008, 24, 243–250. [Google Scholar]
  109. Paterson, A.H.; Sorrells, M.E.; Obendorf, R.L. Methods of evalution for pre-harvest sprouting resistance in wheat breeding programs. Can. J. Plant Sci. 1989, 69, 681–689. [Google Scholar] [CrossRef]
  110. Humphreys, D.G.; Noll, J. Methods for characterization of preharvest sprouting resistance in a wheat breeding program. Euphytica 2002, 126, 61–65. [Google Scholar] [CrossRef]
  111. Gale, M.D.; Ainsworth, C.C. The relationship between alpha-amylase species found in developing and germinating wheat grain. Biochem. Genet. 1984, 22, 1031–1036. [Google Scholar] [CrossRef] [PubMed]
  112. Zhang, Q.; Li, C. Comparisons of Copy Number, Genomic Structure, and Conserved Motifs for α-Amylase Genes from Barley, Rice, and Wheat. Front. Plant Sci. 2017, 8, 1727. [Google Scholar] [CrossRef] [PubMed]
  113. Marchylo, B.A.; Kruger, J.E.; Macgregor, A.W. Production of multiple forms of α-amylase in germinated, incubated, whole, de-embryonated wheat kernels. Cereal Chem. 1983, 61, 305–310. [Google Scholar]
  114. Mundy, J.; Hejgaard, J.; Svendsen, I. Characterization of a bifunctional wheat inhibitor of endogenous α-amylase and subtilisin. FEBS Lett. 1984, 167, 210–214. [Google Scholar] [CrossRef] [Green Version]
  115. Henry, R.J.; Battershell, V.G.; Brennan, P.S.; Oono, K. Control of wheat α-amylase using inhibitors from cereals. J. Sci. Food Agric. 1992, 58, 281–284. [Google Scholar] [CrossRef]
  116. Macgregor, A.W.; Marchylo, B.A.; Kruger, J.E. Multiple α-amylase components in germinated cereal grains determined by isoelectric focusing and chromatofocusing. Cereal Chem. 1988, 65, 326–333. [Google Scholar]
  117. Masojc, P.; Zawistowski, J.; Howes, N.K.; Aung, T.; Gale, M.D. Polymorphism and chromosomal location of endogenous alpha-amylase inhibitor genes in common wheat. Theor. Appl. Genet. 1993, 85, 1043–1048. [Google Scholar] [CrossRef] [PubMed]
  118. Yuan, Y.P.; Chen, X.; Xiao, S.H.; Zhang, W. Extraction and identification of barley α-amylase/subtilisin inhibitor. J. Triticeae Crops 2005, 25, 40–43. [Google Scholar]
  119. Chitnis, V.R.; Gao, F.; Yao, Z.; Jordan, M.C.; Park, S.; Ayele, B.T. After-ripening induced transcriptional changes of hormonal genes in wheat seeds: The cases of brassinosteroids, ethylene, cytokinin and salicylic acid. PLoS ONE 2014, 9, e87543. [Google Scholar] [CrossRef] [PubMed]
  120. Shu, K.; Liu, X.D.; Xie, Q.; He, Z.H. Two faces of one seed: Hormonal regulation of dormancy and germination. Mol. Plant 2016, 9, 34–45. [Google Scholar] [CrossRef] [PubMed]
  121. Kucera, B.; Cohn, M.A.; Leubner-Metzger, G. Plant hormone interactions during seed dormancy release and germination. Seed Sci. Res. 2005, 15, 281–307. [Google Scholar] [CrossRef]
  122. Finkelstein, R.; Reeves, W.; Ariizumi, T.; Steber, C. Molecular aspects of seed dormancy. Ann. Rev. Plant Biol. 2008, 59, 387–415. [Google Scholar] [CrossRef] [PubMed]
  123. Finch-Savage, W.E.; Footitt, S. Seed dormancy cycling and the regulation of dormancy mechanisms to time germination in variable field environments. J. Exp. Bot. 2017, 68, 843–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Matilla, A.J.; Matilla-Vázquez, M.A. Involvement of ethylene in seed physiology. Plant Sci. 2008, 175, 87–97. [Google Scholar] [CrossRef]
  125. Linkies, A.; Leubner-Metzger, G. Beyond gibberellins and abscisic acid: How ethylene and jasmonates control seed germination. Plant Cell Rep. 2012, 31, 253–270. [Google Scholar] [CrossRef] [PubMed]
  126. Gubler, F.; Hughes, T.; Waterhouse, P.; Jacobsen, J. Regulation of dormancy in barley by blue light and after-ripening: Effects on abscisic acid and gibberellin metabolism. Plant Physiol. 2008, 147, 886–896. [Google Scholar] [CrossRef] [PubMed]
  127. Lzydorczyk, C.; Nguyen, T.N.; Jo, S.; Son, S.; Tuan, P.A.; Ayele, B.T. Spatiotemporal modulation of abscisic acid and gibberellin metabolism and signaling mediates the effects of suboptimal and supraoptimal temperatures on seed germination in wheat (Triticum aestivum L.). Plant Cell Environ. 2017, 41, 1022–1037. [Google Scholar] [CrossRef] [PubMed]
  128. Rodriguez, M.V.; Mendiondo, G.M.; Maskin, L.; Gudesblat, G.E.; Iusem, N.D.; Benech-Arnold, R.L. Expression of ABA signalling genes and ABI5 protein levels in imbibed Sorghum bicolor caryopses with contrasting dormancy and at different developmental stages. Ann. Bot. 2009, 104, 975–985. [Google Scholar] [CrossRef] [PubMed]
  129. Ried, J.L.; Walker-Simmons, M.K. Synthesis of abscisic Acid-responsive, heat-stable proteins in embryonic axes of dormant wheat grain. Plant Physiol. 1990, 93, 662–667. [Google Scholar] [CrossRef] [PubMed]
  130. Nambara, E.; Okamoto, M.; Tatematsu, K.; Yano, R.; Seo, M.; Kamiya, Y. Abscisic acid and the control of seed dormancy and germination. Seed Sci. Res. 2010, 20, 55–67. [Google Scholar] [CrossRef]
  131. Schwartz, S.H.; Qin, X.; Zeevaart, J.A. Elucidation of the indirect pathway of abscisic acid biosynthesis by mutants, genes, and enzymes. Plant Physiol. 2003, 131, 1591–1601. [Google Scholar] [CrossRef] [PubMed]
  132. Cutler, A.J.; Krochko, J.E. Formation and breakdown of ABA. Trends Plant Sci. 1999, 4, 472–478. [Google Scholar] [CrossRef]
  133. Saito, S.; Hirai, N.; Matsumoto, C.; Ohigashi, H.; Ohta, D.; Sakata, K.; Mizutani, M. Arabidopsis CYP707As encode (+)-abscisic acid 8′-hydroxylase, a key enzyme in the oxidative catabolism of abscisic acid. Plant Physiol. 2004, 134, 1439–1449. [Google Scholar] [CrossRef] [PubMed]
  134. Kushiro, T.; Okamoto, M.; Nakabayashi, K.; Yamagishi, K.; Kitamura, S.; Asami, T.; Hirai, N.; Koshiba, T.; Kamiya, Y.; Nambara, E. The Arabidopsis cytochrome P450 CYP707A encodes ABA 8′-hydroxylases: Key enzymes in ABA catabolism. EMBO J. 2004, 23, 1647–1656. [Google Scholar] [CrossRef] [PubMed]
  135. Suzuki, T.; Matsuura, T.; Kawakami, N.; Noda, K. Accumulation and leakage of abscisic acid during embryo development and seed dormancy in wheat. J. Plant Growth Regul. 2000, 30, 253–260. [Google Scholar] [CrossRef]
  136. Chono, M.; Honda, I.; Shinoda, S.; Kushiro, T.; Kamiya, Y.; Nambara, E.; Kawakami, N.; Kaneko, S.; Watanabe, Y. Field studies on the regulation of abscisic acid content and germinability during grain development of barley: Molecular and chemical analysis of pre-harvest sprouting. J. Exp. Bot. 2006, 57, 2421–2434. [Google Scholar] [CrossRef] [PubMed]
  137. Garello, G.; Le Page-Degivry, M.T. Evidence for the role of abscisic acid in the genetic and environmental control of dormancy in wheat (Triticum aestivum L.). Seed Sci. Res. 1999, 9, 219–226. [Google Scholar] [CrossRef]
  138. Son, S.; Chitnis, V.R.; Liu, A.; Gao, F.; Nguyen, T.N.; Ayele, B.T. Abscisic acid metabolic genes of wheat (Triticum aestivum L.): Identification and insights into their functionality in seed dormancy and dehydration tolerance. Planta 2016, 244, 429–447. [Google Scholar] [CrossRef] [PubMed]
  139. Chono, M.; Matsunak, H.; Seki, M.; Fujita, M.; Kiribuchi-Otobe, C.; Oda, S.; Kojima, H.; Kobayashi, D.; Kawakami, N. Isolation of a wheat (Triticum aestivum L.) mutant in ABA 8′-hydroxylase gene: Effect of reduced ABA catabolism on germination inhibition under field condition. Breed. Sci. 2013, 63, 104–115. [Google Scholar] [CrossRef] [PubMed]
  140. Holdsworth, M.; Kurup, S.; Mkibbin, R. Molecular and genetic mechanisms regulating the transition from embryo development to germination. Trends Plant Sci. 1999, 4, 275–280. [Google Scholar] [CrossRef]
  141. Ksenia, V.K.; Vasquez-Gross, H.A.; Howell, T.; Bailey, P.; Paraiso, F.; Clissold, L.; Simmonds, J.; Ramirez-Gonzalez, R.H.; Wang, X.; Borrill, P.; et al. Uncovering hidden variation in polyploid wheat. Proc. Natl. Acad. Sci. USA 2017, 114, 913–921. [Google Scholar] [CrossRef]
  142. Marin, E.; Nussaume, L.; Quesada, A.; Gonneau, M.; Sotta, B.; Hugueney, P.; Frey, A.; Marion-Poll, A. Molecular identification of zeaxanthin epoxidase of Nicotiana plumbaginifolia, a gene involved in abscisic acid biosynthesis and corresponding to the ABA locus of Arabidopsis thaliana. EMBO J. 1996, 15, 2331–2342. [Google Scholar] [CrossRef] [PubMed]
  143. North, H.M.; De Almeida, A.; Boutin, J.P.; Frey, A.; To, A.; Botran, L.; Sotta, B.; Marion-Poll, A. The Arabidopsis ABA-deficient mutant aba4 demonstrates that the major route for stress-induced ABA accumulation is via neoxanthin isomers. Plant J. 2007, 50, 810–824. [Google Scholar] [CrossRef] [PubMed]
  144. Tan, B.C.; Schwartz, S.H.; Zeevaart, J.A.; McCarty, D.R. Genetic control of abscisic acid biosynthesis in maize. Proc. Natl. Acad. Sci. USA 1997, 94, 12235–12240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Burbidge, A.; Grieve, T.M.; Jackson, A.; Thompson, A.; McCarty, D.R.; Taylor, I.B. Characterization of the ABA-deficient tomato mutant notabilis and its relationship with maize Vp14. Plant J. 1999, 17, 427–431. [Google Scholar] [CrossRef] [PubMed]
  146. Schwartz, S.H.; Leon-Kloosterziel, K.M.; Koornneef, M.; Zeevaart, J.A. Biochemical characterization of the aba2 and aba3 mutants in Arabidopsis thaliana. Plant Physiol. 1997, 114, 161–166. [Google Scholar] [CrossRef] [PubMed]
  147. Sagi, M.; Scazzocchio, C.; Fluhr, R. The absence of molybdenum cofactor sulfuration is the primary cause of the flacca phenotype in tomato plants. Plant J. 2002, 31, 305–317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Kim, H.; Hwang, H.; Hong, J.W.; Lee, Y.N.; Ahn, I.P.; Yoon, I.S.; Yoo, S.D.; Lee, S.; Lee, S.C.; Kim, B.G. A rice orthologue of the ABA receptor, OsPYL/RCAR5, is a positive regulator of the ABA signal transduction pathway in seed germination and early seedling growth. J. Exp. Bot. 2012, 63, 1013–1024. [Google Scholar] [CrossRef] [PubMed]
  149. Noda, K.; Matsuura, T.; Maekawa, M.; Taketa, S. Chromosomes responsible for sensitivity of embryo to abscisic acid and dormancy in wheat. Euphytica 2002, 123, 203–209. [Google Scholar] [CrossRef]
  150. Schramm, E.C.; Nelson, S.K.; Kidwell, K.K.; Steber, C.M. Increased ABA sensitivity results in higher seed dormancy in soft white spring wheat cultivar ‘Zak’. Theor. Appl. Genet. 2013, 126, 791–803. [Google Scholar] [CrossRef] [PubMed]
  151. Morris, C.F.; Moffatt, J.M.; Sears, R.G.; Paulsen, G.M. Seed dormancy and responses of caryopses, embryos, and calli to abscisic Acid in wheat. Plant Physiol. 1989, 90, 643–647. [Google Scholar] [CrossRef] [PubMed]
  152. Corbineau, F.B.A.; Come, D. Changes in sensitivity to abscisic acid of the developing and maturing embryo of two wheat cultivars with different sprouting susceptibility. Isr. J. Plant Sci. 2000, 48, 189–197. [Google Scholar] [CrossRef]
  153. De Laethauwer, S.; Reheul, D.; De Riek, J.; Haesaert, G. Vp1 expression profiles during kernel development in six genotypes of wheat, triticale and rye. Euphytica 2012, 188, 61–70. [Google Scholar] [CrossRef]
  154. Fan, J.; Niu, X.; Wang, Y.; Ren, G.; Zhuo, T.; Yang, Y.; Lu, B.R.; Liu, Y. Short, direct repeats (SDRs)-mediated post-transcriptional processing of a transcription factor gene OsVP1 in rice (Oryza sativa). J. Exp. Bot. 2007, 58, 3811–3817. [Google Scholar] [CrossRef] [PubMed]
  155. Giraudat, J.; Hauge, B.M.; Valon, C.; Smalle, J.; Parcy, F.; Goodman, H.M. Isolation of the Arabidopsis ABI3 gene by positional cloning. Plant Cell 1992, 4, 1251–1261. [Google Scholar] [CrossRef] [PubMed]
  156. Finkelstein, R.R.; Gampala, S.S.L.; Rock, C.D. Abscisic Acid Signaling in Seeds and Seedlings. Plant Cell 2002, 14, s15–s45. [Google Scholar] [CrossRef] [PubMed]
  157. McCrate, A.J.; Nielsen, M.T.; Paulsen, G.M.; Heyne, E.G. Relationship between sprouting in wheat and embryo response to endogenous inhibition. Euphytica 1982, 31, 193–200. [Google Scholar] [CrossRef]
  158. Curaba, J.; Moritz, T.; Blervaque, R.; Parcy, F.; Raz, V.; Herzog, M.; Vachon, G. AtGA3ox2, a Key Gene Responsible for Bioactive Gibberellin Biosynthesis, Is Regulated during Embryogenesis by LEAFY COTYLEDON2 and FUSCA3 in Arabidopsis. Plant Physiol. 2004, 136, 3660–3669. [Google Scholar] [CrossRef] [PubMed]
  159. Lu, Q.S.; Paz, J.D.; Pathmanathan, A.; Chiu, R.S.; Tsai, A.Y.; Gazzarrini, S. The C-terminal domain of FUSCA3 negatively regulates mRNA and protein levels, and mediates sensitivity to the hormones abscisic acid and gibberellic acid in Arabidopsis. Plant J. 2010, 64, 100–113. [Google Scholar] [CrossRef] [PubMed]
  160. Yamaguchi, S. Gibberellin metabolism and its regulation. Ann. Rev. Plant Biol. 2008, 59, 225–251. [Google Scholar] [CrossRef] [PubMed]
  161. Pearce, S.; Huttly, A.K.; Prosser, I.M.; Li, Y.D.; Vaughan, S.P.; Gallova, B.; Patil, A.; Coghill, J.A.; Dubcovsky, J.; Hedden, P.; et al. Heterologous expression and transcript analysis of gibberellin biosynthetic genes of grasses reveals novel functionality in the GA3ox family. BMC Plant Biol. 2015, 15, 130. [Google Scholar] [CrossRef] [PubMed]
  162. Kashiwakura, Y.-I.; Kobayashi, D.; Jikumaru, Y.; Takebayashi, Y.; Nambara, E.; Seo, M.; Kamiya, Y.; Kushiro, T.; Kawakami, N. Highly Sprouting-Tolerant Wheat Grain Exhibits Extreme Dormancy and Cold Imbibition-Resistant Accumulation of Abscisic Acid. Plant Cell Physiol. 2016, 57, 715–732. [Google Scholar] [CrossRef] [PubMed]
  163. Rodriguez, M.V.; Mendiondo, G.M.; Cantoro, R.; Auge, G.A.; Luna, V.; Masciarelli, O.; Benech-Arnold, R.L. Expression of seed dormancy in grain sorghum lines with contrasting pre-harvest sprouting behavior involves differential regulation of gibberellin metabolism genes. Plant Cell Physiol. 2012, 53, 64–80. [Google Scholar] [CrossRef] [PubMed]
  164. Ye, H.; Feng, J.; Zhang, L.; Zhang, J.; Mispan, M.S.; Cao, Z.; Beighley, D.H.; Yang, J.; Gu, X. Map-based cloning of Seed Dormancy1-2 identified a gibberellin synthesis gene regulating the development of endosperm-imposed dormancy in rice. Plant Physiol. 2015, 169, 2152–2165. [Google Scholar] [CrossRef] [PubMed]
  165. Magwa, R.A.; Zhao, H.; Xing, Y. Genome-wide association mapping revealed a diverse genetic basis of seed dormancy across subpopulations in rice (Oryza sativa L.). BMC Genet. 2016, 17, 28. [Google Scholar] [CrossRef] [PubMed]
  166. Ueguchi-Tanaka, M.; Ashikari, M.; Nakajima, M.; Itoh, H.; Katoh, E.; Kobayashi, M.; Chow, T.Y.; Hsing, Y.I.; Kitano, H.; Yamaguchi, I.; et al. GIBBERELLIN INSENSITIVE DWARF1 encodes a soluble receptor for gibberellin. Nature 2005, 437, 693–698. [Google Scholar] [CrossRef] [PubMed]
  167. Li, A.; Yang, W.; Li, S.; Liu, D.; Guo, X.; Sun, J.; Zhang, A. Molecular characterization of three GIBBERELLIN-INSENSITIVE DWARF1 homologous genes in hexaploid wheat. J. Plant Physiol. 2013, 170, 432–443. [Google Scholar] [CrossRef] [PubMed]
  168. Sun, T.P. The molecular mechanism and evolution of the GA-GID1-DELLA signaling module in plants. Curr. Biol. 2011, 21, 338–345. [Google Scholar] [CrossRef] [PubMed]
  169. Sasaki, A.; Itoh, H.; Gomi, K.; Ueguchi-Tanaka, M.; Ishiyama, K.; Kobayashi, M.; Jeong, D.H.; An, G.; Kitano, H.; Ashikari, M.; et al. Accumulation of phosphorylated repressor for gibberellin signaling in an F-box mutant. Science 2003, 299, 1896–1898. [Google Scholar] [CrossRef] [PubMed]
  170. McGinnis, K.M.; Thomas, S.G.; Soule, J.D.; Strader, L.C.; Zale, J.M.; Sun, T.P.; Steber, C.M. The Arabidopsis SLEEPY1 gene encodes a putative F-box subunit of an SCF E3 ubiquitin ligase. Plant Cell 2003, 15, 1120–1130. [Google Scholar] [CrossRef] [PubMed]
  171. Murase, K.; Hirano, Y.; Sun, T.P.; Hakoshima, T. Gibberellin-induced DELLA recognition by the gibberellin receptor GID1. Nature 2008, 456, 459–463. [Google Scholar] [CrossRef] [PubMed]
  172. Daviere, J.M.; Achard, P. Gibberellin signaling in plants. Development 2013, 140, 1147–1151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Tyler, L.; Thomas, S.G.; Hu, J.; Dill, A.; Alonso, J.M.; Ecker, J.R.; Sun, T.P. Della proteins and gibberellin-regulated seed germination and floral development in Arabidopsis. Plant Physiol. 2004, 135, 1008–1019. [Google Scholar] [CrossRef] [PubMed]
  174. Peng, J.; Richards, D.E.; Hartley, N.M.; Murphy, G.P.; Devos, K.M.; Flintham, J.E.; Beales, J.; Fish, L.J.; Worland, A.J.; Pelica, F.; et al. Green revolution’ genes encode mutant gibberellin response modulators. Nature 1999, 400, 256–261. [Google Scholar] [CrossRef] [PubMed]
  175. Ikeda, A.; Ueguchi-Tanaka, M.; Sonoda, Y.; Kitano, H.; Koshioka, M.; Futsuhara, Y.; Matsuoka, M.; Yamaguchi, J. Slender rice, a constitutive gibberellin response mutant, is caused by a null mutation of the SLR1 gene, an ortholog of the height-regulating gene GAI/RGA/RHT/D8. Plant Cell 2001, 13, 999–1010. [Google Scholar] [CrossRef] [PubMed]
  176. Chandler, P.M.; Marion-Poll, A.; Ellis, M.; Gubler, F. Mutants at the Slender1 locus of barley cv Himalaya. Molecular and physiological characterization. Plant Physiol. 2002, 129, 181–190. [Google Scholar] [CrossRef] [PubMed]
  177. Chandler, P.M. Hormonal regulation of gene expression in the “slender” mutant of barley (Hordeum vulgare L.). Planta 1988, 175, 115–120. [Google Scholar] [CrossRef] [PubMed]
  178. Gubler, F.; Raventos, D.; Keys, M.; Watts, R.; Mundy, J.; Jacobsen, J.V. Target genes and regulatory domains of the GAMYB transcriptional activator in cereal aleurone. Plant J. 1999, 17, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Woodger, F.J.; Gubler, F.; Pogson, B.J.; Jacobsen, J.V. A Mak-like kinase is a repressor of GAMYB in barley aleurone. Plant J. 2003, 33, 707–717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Kaneko, M.; Inukai, Y.; Ueguchi-Tanaka, M.; Itoh, H.; Izawa, T.; Kobayashi, Y.; Hattori, T.; Miyao, A.; Hirochika, H.; Ashikari, M.; et al. Loss-of-function mutations of the rice GAMYB gene impair alpha-amylase expression in aleurone and flower development. Plant Cell 2004, 16, 33–44. [Google Scholar] [CrossRef] [PubMed]
  181. Jacobsen, J.V.; Barrero, J.M.; Hughes, T.; Julkowska, M.; Taylor, J.M.; Xu, Q.; Gubler, F. Roles for blue light, jasmonate and nitric oxide in the regulation of dormancy and germination in wheat grain (Triticum aestivum L.). Planta 2013, 238, 121–138. [Google Scholar] [CrossRef] [PubMed]
  182. Xu, Q.; Truong, T.T.; Barrero, J.M.; Jacobsen, J.V.; Hocart, C.H.; Gubler, F. A role for jasmonates in the release of dormancy by cold stratification in wheat. J. Exp. Bot. 2016, 67, 3497–3508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Friedrichsen, D.M.; Nemhauser, J.; Muramitsu, T.; Maloof, J.N.; Alonso, J.; Ecker, J.R.; Furuya, M.; Chory, J. Three redundant brassinosteroid early response genes encode putative bHLH transcription factors required for normal growth. Genetics 2002, 162, 1445–1456. [Google Scholar] [PubMed]
  184. Zhang, L.Y.; Bai, M.Y.; Wu, J.; Zhu, J.Y.; Wang, H.; Zhang, Z.; Wang, W.; Sun, Y.; Zhao, J.; Sun, X.; et al. Antagonistic HLH/bHLH transcription factors mediate brassinosteroid regulation of cell elongation and plant development in rice and Arabidopsis. Plant Cell 2009, 21, 3767–3780. [Google Scholar] [CrossRef] [PubMed]
  185. Adkins, S.W.; Ross, J.D. Studies in Wild Oat Seed Dormancy: The role of ethylene in dormancy breakage and germination of wild oat seeds (Avena fatua L.). Plant Physiol. 1981, 67, 358–362. [Google Scholar] [CrossRef] [PubMed]
  186. Steber, C.M.; McCourt, P. A role for brassinosteroids in germination in Arabidopsis. Plant Physiol. 2001, 125, 763–769. [Google Scholar] [CrossRef] [PubMed]
  187. Divi, U.K.; Krishna, P. Overexpression of the brassinosteroid biosynthetic gene AtDWF4 in Arabidopsis seeds overcomes abscisic acid-induced inhibition of germination and increases cold tolerance in transgenic seedlings. J. Plant Growth Regul. 2010, 29, 385–393. [Google Scholar] [CrossRef]
  188. Hansen, M.; Chae, H.S.; Kieber, J.J. Regulation of ACS protein stability by cytokinin and brassinosteroid. Plant J. 2009, 57, 606–614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Locke, J.M.; Bryce, J.H.; Morris, P.C. Contrasting effects of ethylene perception and biosynthesis inhibitors on germination and seedling growth of barley (Hordeum vulgare L.). J. Exp. Bot. 2000, 51, 1843–1849. [Google Scholar] [CrossRef] [PubMed]
  190. Gianinetti, A.; Laarhoven, L.J.; Persijn, S.T.; Harren, F.J.; Petruzzelli, L. Ethylene production is associated with germination but not seed dormancy in red rice. Ann. Bot. 2007, 99, 735–745. [Google Scholar] [CrossRef] [PubMed]
  191. Morris, C.F.; Mueller, D.D.; Faubion, J.M.; Paulsen, G.M. Identification of l-Tryptophan as an Endogenous Inhibitor of Embryo Germination in White Wheat. Plant Physiol. 1988, 88, 435–440. [Google Scholar] [CrossRef] [PubMed]
  192. Ramaih, S.; Guedira, M.; Paulsen, G.M. Relationship of indoleacetic acid and tryptophan to dormancy and preharvest sprouting of wheat. Funct. Plant Biol. 2003, 30, 939–945. [Google Scholar] [CrossRef]
  193. Brady, S.M.; Sarkar, S.F.; Bonetta, D.; McCourt, P. The ABSCISIC ACID INSENSITIVE 3 (ABI3) gene is modulated by farnesylation and is involved in auxin signaling and lateral root development in Arabidopsis. Plant J. 2003, 34, 67–75. [Google Scholar] [CrossRef] [PubMed]
  194. Liu, P.P.; Montgomery, T.A.; Fahlgren, N.; Kasschau, K.D.; Nonogaki, H.; Carrington, J.C. Repression of AUXIN RESPONSE FACTOR10 by microRNA160 is critical for seed germination and post-germination stages. Plant J. 2007, 52, 133–146. [Google Scholar] [CrossRef] [PubMed]
  195. Martinez, S.A.; Tuttle, K.M.; Takebayashi, Y.; Seo, M.; Campbell, K.G.; Steber, C.M. The wheat ABA hypersensitive ERA8 mutant is associated with increased preharvest sprouting tolerance and altered hormone accumulation. Euphytica 2016, 212, 229–245. [Google Scholar] [CrossRef]
  196. Das, A.; Kim, D.; Khadka, P.; Rakwal, R.; Rohila, J.S. Unraveling Key Metabolomic Alterations in Wheat Embryos Derived from Freshly Harvested and Water-Imbibed Seeds of Two Wheat Cultivars with Contrasting Dormancy Status. Front. Plant Sci. 2017, 8, 1203. [Google Scholar] [CrossRef] [PubMed]
  197. Chapman, E.J.; Estelle, M. Mechanism of auxin-regulated gene expression in plants. Ann. Rev. Genet. 2009, 43, 265–285. [Google Scholar] [CrossRef] [PubMed]
  198. Vanneste, S.; Friml, J. Auxin: A trigger for change in plant development. Cell 2009, 136, 1005–1016. [Google Scholar] [CrossRef] [PubMed]
  199. Guilfoyle, T.J.; Hagen, G. Auxin response factors. Curr. Opin. Plant Biol. 2007, 10, 453–460. [Google Scholar] [CrossRef] [PubMed]
  200. Mockaitis, K.; Estelle, M. Auxin receptors and plant development: A new signaling paradigm. Ann. Rev. Cell Dev. Biol. 2008, 24, 55–80. [Google Scholar] [CrossRef] [PubMed]
  201. Ooms, J.; Leon-Kloosterziel, K.M.; Bartels, D.; Koornneef, M.; Karssen, C.M. Acquisition of Desiccation Tolerance and Longevity in Seeds of Arabidopsis thaliana (A Comparative Study Using Abscisic Acid-Insensitive abi3 Mutants). Plant Physiol. 1993, 102, 1185–1191. [Google Scholar] [CrossRef] [PubMed]
  202. Finkelstein, R.R. Mutations at two new Arabidopsis ABA response loci are similar to the abi3 mutations. Plant J. 1994, 5, 765–771. [Google Scholar] [CrossRef]
  203. Brocard-Gifford, I.M.; Lynch, T.J.; Finkelstein, R.R. Regulatory Networks in Seeds Integrating Developmental, Abscisic Acid, Sugar, and Light Signaling. Plant Physiol. 2003, 131, 78–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Nemhauser, J.L.; Hong, F.; Chory, J. Different plant hormones regulate similar processes through largely nonoverlapping transcriptional responses. Cell 2006, 126, 467–475. [Google Scholar] [CrossRef] [PubMed]
  205. Santner, A.; Estelle, M. Recent advances and emerging trends in plant hormone signalling. Nature 2009, 459, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  206. Kendall, S.; Penfield, S. Maternal and zygotic temperature signalling in the control of seed dormancy and germination. Seed Sci. Res. 2012, 22, S23–S29. [Google Scholar] [CrossRef]
  207. Reddy, L.V.; Metzger, R.J.; Ching, T.M. Effect of Temperature on Seed Dormancy of Wheat. Crop Sci. 1985, 25, 455–458. [Google Scholar] [CrossRef]
  208. Nyachiro, J.; Clarke, F.R.; Depauw, R.; Knox, R.; Armstrong, K.C. Temperature effects on seed germination and expression of seed dormancy in wheat. Euphytica 2002, 126, 123–127. [Google Scholar] [CrossRef]
  209. Mares, D. Pre-harvest sprouting in wheat. I. Influence of cultivar, rainfall and temperature during grain ripening. Aust. J. Agric. Res. 1993, 44, 1259–1272. [Google Scholar] [CrossRef]
  210. Lunn, G.D.; Kettlewell, P.; Major, B.J.; Scott, R.K. Variation in dormancy duration of the U.K. wheat cultivar Hornet due to environmental conditions during grain development. Euphytica 2002, 126, 89–97. [Google Scholar] [CrossRef]
  211. Biddulph, T.; Mares, D.; Plummer, J.A.; Setter, T. Drought and high temperature increases preharvest sprouting tolerance in a genotype without grain dormancy. Euphytica 2005, 143, 277–283. [Google Scholar] [CrossRef]
  212. Footitt, S.; Douterelo-Soler, I.; Clay, H.; Finch-Savage, W.E. Dormancy cycling in Arabidopsis seeds is controlled by seasonally distinct hormone-signaling pathways. Proc. Natl. Acad. Sci. USA 2011, 108, 20236–20241. [Google Scholar] [CrossRef] [PubMed]
  213. Kulwal, P.L.; Singh, R.; Balyan, H.S.; Gupta, P.K. Genetic basis of pre-harvest sprouting tolerance using single-locus and two-locus QTL analyses in bread wheat. Funct. Integr. Genom. 2004, 4, 94–101. [Google Scholar] [CrossRef] [PubMed]
  214. Imtiaz, M.; Ogbonnaya, F.C.; Oman, J.; van Ginkel, M. Characterization of quantitative trait loci controlling genetic variation for preharvest sprouting in synthetic backcross-derived wheat lines. Genetics 2008, 178, 1725–1736. [Google Scholar] [CrossRef] [PubMed]
  215. Liu, S.; Bai, G. Dissection and fine mapping of a major QTL for preharvest sprouting resistance in white wheat Rio Blanco. Theor. Appl. Genet. 2010, 121, 1395–1404. [Google Scholar] [CrossRef] [PubMed]
  216. Flintham, J. Different genetic components control coat-imposed and embryo-imposed dormancy in wheat. Seed Sci. Res. 2000, 10, 43–50. [Google Scholar] [CrossRef]
  217. Lohwasser, U.; Röder, M.S.; Börner, A. QTL mapping of the domestication traits pre-harvest sprouting and dormancy in wheat (Triticum aestivum L.). Euphytica 2005, 143, 247–249. [Google Scholar] [CrossRef]
  218. Chao, S.; Xu, S.; Elias, E.; Faris, J.; Sorrells, M. Identification of Chromosome Locations of Genes Affecting Preharvest Sprouting and Seed Dormancy Using Chromosome Substitution Lines in Tetraploid Wheat (Triticum turgidum L.). Crop Sci. 2010, 50, 1180–1187. [Google Scholar] [CrossRef] [Green Version]
  219. Lohwasser, U.; Rehman, M.A.; Börner, A. Discovery of loci determining pre-harvest sprouting and dormancy in wheat and barley applying segregation and association mapping. Biol. Plant 2013, 57, 663–674. [Google Scholar] [CrossRef]
  220. Borner, A.; Nagel, M.; Agacka-Moldoch, M.; Gierke, P.U.; Oberforster, M.; Albrecht, T.; Mohler, V. QTL analysis of falling number and seed longevity in wheat (Triticum aestivum L.). J. Appl. Genet. 2018, 59, 35–42. [Google Scholar] [CrossRef] [PubMed]
  221. Liu, S.; Cai, S.; Graybosch, R.; Chen, C.; Bai, G. Quantitative trait loci for resistance to pre-harvest sprouting in US hard white winter wheat Rio Blanco. Theor. Appl. Genet. 2008, 117, 691–699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Singh, A.K.; Knox, R.E.; Clarke, J.M.; Clarke, F.R.; Singh, A.; DePauw, R.M.; Cuthbert, R.D. Genetics of pre-harvest sprouting resistance in a cross of Canadian adapted durum wheat genotypes. Mol. Breed. 2014, 33, 919–929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Lin, M.; Cai, S.; Wang, S.; Liu, S.; Zhang, G.; Bai, G. Genotyping-by-sequencing (GBS) identified SNP tightly linked to QTL for pre-harvest sprouting resistance. Theor. Appl. Genet. 2015, 128, 1385–1395. [Google Scholar] [CrossRef] [PubMed]
  224. Cao, L.; Hayashi, K.; Tokui, M.; Mori, M.; Miura, H.; Onishi, K. Detection of QTLs for traits associated with pre-harvest sprouting resistance in bread wheat (Triticum aestivum L.). Breed. Sci. 2016, 66, 260–270. [Google Scholar] [CrossRef] [PubMed]
  225. Dale, Z.; Jie, H.; Luyu, H.; Cancan, Z.; Yun, Z.; Yarui, S.; Suoping, L. An Advanced Backcross Population through Synthetic Octaploid Wheat as a “Bridge”: Development and QTL Detection for Seed Dormancy. Front. Plant Sci. 2017, 8, 2123. [Google Scholar] [CrossRef] [PubMed]
  226. Lin, Y.; Liu, S.; Liu, Y.; Liu, Y.; Chen, G.; Xu, J.; Deng, M.; Jiang, Q.; Wei, Y.; Lu, Y.; et al. Genome-wide association study of pre-harvest sprouting resistance in Chinese wheat founder parents. Genet. Mol. Biol. 2017, 40, 620–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Torada, A.; Ikeguchi, S.; Koike, M. Mapping and validation of PCR-based markers associated with a major QTL for seed dormancy in wheat. Euphytica 2005, 143, 251–255. [Google Scholar] [CrossRef]
  228. Lazo, G.R.; Chao, S.; Hummel, D.D.; Edwards, H.; Crossman, C.C.; Lui, N.; Matthews, D.E.; Carollo, V.L.; Hane, D.L.; You, F.M.; et al. Development of an expressed sequence tag (EST) resource for wheat (Triticum aestivum L.): EST generation, unigene analysis, probe selection and bioinformatics for a 16,000-locus bin-delineated map. Genetics 2004, 168, 585–593. [Google Scholar] [CrossRef] [PubMed]
  229. Iyer-Pascuzzi, A.S.; McCouch, S.R. Functional markers for xa5-mediated resistance in rice (Oryza sativa, L.). Mol. Breed. 2007, 19, 291–296. [Google Scholar] [CrossRef]
  230. Yang, Y.; Zhao, X.-L.; Zhang, Y.; Chen, X.-M.; He, Z.-H.; Yu, Z.; Xia, L.-Q. Evaluation and Validation of Four Molecular Markers Associated with Pre-harvest Sprouting Tolerance in Chinese Wheat. Acta Agron. Sin. 2008, 34, 17–24. [Google Scholar] [CrossRef]
  231. Guo, F.; Liang, W.; Fan, Q.; Huang, C.; Gao, Q.; Li, G. Distribution and evolution of allelic variation of Vp1B3 in Shandong wheat. J. Triticeae Crops 2009, 29, 575–578. [Google Scholar]
  232. Xia, L.Q.; Yang, Y.; Ma, Y.Z.; Chen, X.M.; He, Z.H.; Röder, M.S.; Jones, H.D.; Shewry, P.R. What can the Viviparous-1 gene tell us about wheat pre-harvest sprouting? Euphytica 2009, 168, 385–394. [Google Scholar] [CrossRef]
  233. Zhao, B.; Wan, Y.X.; Wang, R. Screening of wheat cultivar resources with pre-harvest sprouting resistance. J. Anhui Agric. Sci. 2010, 38, 8900–8902. [Google Scholar]
  234. Rasheed, A.; Wen, W.; Gao, F.; Zhai, S.; Jin, H.; Liu, J.; Guo, Q.; Zhang, Y.; Dreisigacker, S.; Xia, X.; et al. Development and validation of KASP assays for genes underpinning key economic traits in bread wheat. Theor. Appl. Genet. 2016, 129, 1843–1860. [Google Scholar] [CrossRef] [PubMed]
  235. Wang, Y.; Wang, X.L.; Meng, J.Y.; Zhang, Y.J.; He, Z.H.; Yang, Y. Characterization of Tamyb10 allelic variants and development of STS marker for pre-harvest sprouting resistance in Chinese bread wheat. Mol. Breed. 2016, 36, 148. [Google Scholar] [CrossRef] [PubMed]
  236. Brenchley, R.; Spannagl, M.; Pfeifer, M.; Barker, G.L.A.; D’Amore, R.; Allen, A.M.; McKenzie, N.; Kramer, M.; Kerhornou, A.; Bolser, D.; et al. Analysis of the bread wheat genome using whole-genome shotgun sequencing. Nature 2012, 491, 705. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of flavonoid biosynthesis pathway in plants. Enzymes are shown in blue while intermediates are shown in black. End products are indicated in colored shapes. Dotted arrows represent multiple steps. CHS, chalcone synthase; CHI, chalcone isomerase; F3H, flavanone 3-hydroxylase; DFR, dihydroflavonol 4-reductase; ANS, anthocyanidin synthase; UFGT, UDP-glucose flavonoid 3-O glucosyltransferase; FLS, flavonol synthase; LAR, leucoanthocyanidin reductase; ANR, anthocyanidin reductase.
Figure 1. Schematic representation of flavonoid biosynthesis pathway in plants. Enzymes are shown in blue while intermediates are shown in black. End products are indicated in colored shapes. Dotted arrows represent multiple steps. CHS, chalcone synthase; CHI, chalcone isomerase; F3H, flavanone 3-hydroxylase; DFR, dihydroflavonol 4-reductase; ANS, anthocyanidin synthase; UFGT, UDP-glucose flavonoid 3-O glucosyltransferase; FLS, flavonol synthase; LAR, leucoanthocyanidin reductase; ANR, anthocyanidin reductase.
Agronomy 09 00117 g001
Figure 2. Schematic representation of ABA metabolism and signaling pathway in plants. Enzymes are shown in blue while intermediates are shown in black. The end product is indicated in the yellow circule; ZEP, zeaxanthin epoxidase; VDE, violaxanthin de-epoxidase; NSY, neoxanthin synthase; NCED, 9-cisepoxycarotenoid dioxygenase; XD, xanthoxin dehydrogenase; AAO3, abscisic aldehyde oxidase; PYR/PYL/RCAR, pyrabactin resistance/pyrabactin-like/regulatory components of ABA receptors; PP2C, protein phosphatase 2C; SnRK2, SNF1-related protein kinase2; ABI3, abscisic acid insensitive 3; ABI4, abscisic acid insensitive 4; ABI5, abscisic acid insensitive 5; VP1, viviparous 1; ABF, ABRE binding factor.
Figure 2. Schematic representation of ABA metabolism and signaling pathway in plants. Enzymes are shown in blue while intermediates are shown in black. The end product is indicated in the yellow circule; ZEP, zeaxanthin epoxidase; VDE, violaxanthin de-epoxidase; NSY, neoxanthin synthase; NCED, 9-cisepoxycarotenoid dioxygenase; XD, xanthoxin dehydrogenase; AAO3, abscisic aldehyde oxidase; PYR/PYL/RCAR, pyrabactin resistance/pyrabactin-like/regulatory components of ABA receptors; PP2C, protein phosphatase 2C; SnRK2, SNF1-related protein kinase2; ABI3, abscisic acid insensitive 3; ABI4, abscisic acid insensitive 4; ABI5, abscisic acid insensitive 5; VP1, viviparous 1; ABF, ABRE binding factor.
Agronomy 09 00117 g002
Figure 3. Schematic representation of GA metabolism and signaling pathway in plants. Enzymes are shown in blue while intermediates are shown in black. Multiple arrows represent multiple steps. GGDP, geranylgeraryl diphosphate; CPS, ent- ent- Copalyl diphosphate synthase; KS, ent-Kaurene synthase; KO, ent-Kaurene oxidase; KAO, ent- Kaurenoic acid oxidase; GA20ox,GA20 oxidase; GA3ox,GA3 oxidase; GA2ox,GA2 oxidase; GID1, gibberellin insensitive dwarf 1; GID2, gibberellin insensitive dwarf 2; SLN, slender1 in barley; SLR1, slender rice1; RHT, reduced height; GAMYB, GA regulated MYB transcriptional regulator; KGM, kinase associated with GAMYB.
Figure 3. Schematic representation of GA metabolism and signaling pathway in plants. Enzymes are shown in blue while intermediates are shown in black. Multiple arrows represent multiple steps. GGDP, geranylgeraryl diphosphate; CPS, ent- ent- Copalyl diphosphate synthase; KS, ent-Kaurene synthase; KO, ent-Kaurene oxidase; KAO, ent- Kaurenoic acid oxidase; GA20ox,GA20 oxidase; GA3ox,GA3 oxidase; GA2ox,GA2 oxidase; GID1, gibberellin insensitive dwarf 1; GID2, gibberellin insensitive dwarf 2; SLN, slender1 in barley; SLR1, slender rice1; RHT, reduced height; GAMYB, GA regulated MYB transcriptional regulator; KGM, kinase associated with GAMYB.
Agronomy 09 00117 g003
Table 1. QTLs for PHS and related traits identified in wheat.
Table 1. QTLs for PHS and related traits identified in wheat.
TraitQTLChromosomeNearest MarkerQTL nameMaterialReference
PHS and GC53AL
3BL
3DL
5AS
Xffb293
Xgwm403, Xbcd131
Xgwm3
Xbcd1871
-RILs[1]
PHS and GC31BS
4BL
7AS
Xpsp3000
Xpsp3030-Xpsp3078
Xpsp3050
-RILs[22]
PHS and SD33A
3A
3A
Xpsr394-Xgwm5
Xcdo345
Xcdo345-Xbcd141
taVp1
QPhs.ocs-3A.1
QPhs.ocs-3A.2
RILs[23]
PHS and SD24AL
3AL
Xksuf8a-Xbcd402b
Xpsr903b-XATPased
-RILs[217]
PHS13ALXwmc153-Xgwm155QPhs.ccsu3A.1RILs[24]
SD14AXgwm397-Xgwm269-Xbarc170-DHLs[3]
PHS and SD13ALXbarc310-Xbcd907QPhs.ocs-3A.1RILs[25]
PHS and SD13AS
3AS
Xbarc310
Xbarc321
QPhs-3AS
QPhs.pseru-3AS
RILs[26]
PHS 32B
2B
Xdup398-Xbarc54
Xbarc105-Xbarc334
QPhs.pseru-2B.1
QPhs.pseru-2B.2
RILs [221]
PHS12DSXgwm261-Xgwm484Qphs.sau-2DF2 and F6[28]
PHS 42B
2D
3D
6D
Xbarc55-Xwmc474
Xwmc111-WxPt-999
7Xbarc1161
Xcfd37-Xbarc196
QPhs.cnl-2B.1
QPhs.cnl-2D.1
QPhs.cnl-3D.1
QPhs.cnl-6D.1
DHLs [14]
PHS 32AL
3AL
3BL
Xgwm1045-Xgwm296
Xgwm153-Xgwm155
Xgwm1005–Xgwm980
QPhs.ccsu-2A.5
QPhs.ccsu-3A.1
QPhs.ccsu-3B.6
RILs [32]
PHS and GC 53B
3D
3A
5D
3D
Xbarc77–Xwmc30
7Xwmc552–Xwmc533
Xcfa2193–Xwmc594
Xgwm469–Xcfd10
Xwmc11–Xcfd223
QGi.crc-3B
QGi.crc-3D
QSi.crc-3A
QSi.crc-5D
QCl.crc-3D
DHLs [29]
PHS and SD13BLXwmc527-Xgwm77-DHLs[31]
PHS15DXCFD40-XBARC1097qPhs5D.1DHLs[33]
PHS and SD 52A
2B
3A
4A
7B
521-2A
521-2B
521-3A
521-4A
521-7B
-Single chromosome
substitution lines
[218]
PHS 31A
2A
7B
Xwmc611-Xwmc333
Xgwm515-Xgwm425
Xgwm297-Xwmc532
QPhsd.spa.-1A.1
QPhsd.spa.-2A.1
QPhsd.spa.-7B.1
RILs [35]
PHS 43B
4A
7B
7D
19 SNPs flanking the QTL
12 SNPs flanking the QTL
10 SNPs flanking the QTL
04 SNPs flanking the QTL
QSi.crc-3B
QGi.crc-4A
QSi.crc-7B
QFn.crc-7D
DHLs [80]
PHS 51A
1B
5B
7A
7B
wPt-6274
Xwmc191
wPt-6910-wPt-7400
Xcfa2174
Xwmc606
QPhs.spa-1A
QPhs.spa-1B
QPhs.spa-5B
QPhs.spa-7A
QPhs.spa-7B
DHLs [222]
PHS and SD12BXwmc477-Xbarc55Sdr2BRILs[76]
PHS and SD14A wsnp_Ex_c66324_64493429 - CD9202984A-1RILs[44]
PHS and SD 44A
4B
5A
5B
GBS212432-GBS10994
7Xbarc20-Xwmc238
TTM_199619-TTM_1259
7Xbarc346-2-TTM_62137_50
Qphs.pseru-4A.1
Qphs.pseru-4B.1
Qphs.pseru-5A.1
Qphs.pseru-5B.1
RILs[223]
PHS and GC 63AL
3AL
3AL
3DL
3DL
1A/1D/3A/5B
Xwmc559-1
Tamyb10-A1-66
Tamyb10-A1-74
BS00067163_51
Tamyb10-D1-93
Xbarc148
- RILs [55]
PHS 63A
4A
1B
7B
4A
6B
TaMFT
cfa2256
Xbarc181
UCW99
cfa2256
Xwmc397
QDor-3A
QDor-4A
QDor-1B
QHt-7B
QAwn-4A
QAwn-6B
RILs [224]
PHS and SD12AXgwm95-Xgwm372Sdr2ARILs[46]
PHS and SD
PHS
PHS
3
2
3
2D
3D
3D
1B
1B
3A
3D
5D
Xwmc503
Xcfd22
Vp1-4
tPt-7980
wPt-645
7AX-111578083
3 DArT-seq and 5 SNPs
AX-109028892
QDor-2D
QDor-3D
TaVp1
-
QTL1
QTL2
QTL3
Back crosspopulation
86 Chinesegermplasm
717 Chinese wheat landraces
[225]
[226]
[37]
PHS 51A
4D
5A
5D
7B
wPt-6654-wPt-7030
wPt-0710-Rht-D1
gwm186-P7560-439
P7551-267-wmc574
P7455-236-P7553-711
- RIL [220]
Note: PHS-pre-harvest sprouting; DHLs-doubled haploid lines; RILs-recombinant inbred lines; GC-grain color; SD-seed dormancy.
Table 2. Genes for PHS and related traits identified in wheat.
Table 2. Genes for PHS and related traits identified in wheat.
Wheat GeneChromosomesGene FunctionHomologs/Orthologs GeneExperimental MethodologyReferences
TaSdr-A12AS SD Rice OsSdr4
orthologs
Comparative genomics approach [46]
TaSdr-B12BS SD Rice OsSdr4
orthologs
homologous
cloning approach
[76]
TaMFT3AS SD Wheat TaMFT homologTranscriptomic approach [41]
TaPHS13AS SD Wheat MFT homologcomparative fine mapping and map-based cloning [16,42]
TaVp-1Group 3 Chromosomes SD and PHS Maize Vp1 and rice OsVp1 orthologsGenomic southern analysis[38,39,84,86]
Tamyb10
PM19-A1/A2
TaMKK3-A
Group 3 Chromosomes
4AL
4AL
GC
SD
SD
Arabidopsis TT2 and Rice OsMYB3 orthologs
-
-
Cloning approach
Transcriptomic approach
Map-based
approach
[40]
[44]
[45]
Note: SD-seed dormancy; PHS-pre-harvest sprouting; GC-grain color.

Share and Cite

MDPI and ACS Style

Ali, A.; Cao, J.; Jiang, H.; Chang, C.; Zhang, H.-P.; Sheikh, S.W.; Shah, L.; Ma, C. Unraveling Molecular and Genetic Studies of Wheat (Triticum aestivum L.) Resistance against Factors Causing Pre-Harvest Sprouting. Agronomy 2019, 9, 117. https://doi.org/10.3390/agronomy9030117

AMA Style

Ali A, Cao J, Jiang H, Chang C, Zhang H-P, Sheikh SW, Shah L, Ma C. Unraveling Molecular and Genetic Studies of Wheat (Triticum aestivum L.) Resistance against Factors Causing Pre-Harvest Sprouting. Agronomy. 2019; 9(3):117. https://doi.org/10.3390/agronomy9030117

Chicago/Turabian Style

Ali, Ahmad, Jiajia Cao, Hao Jiang, Cheng Chang, Hai-Ping Zhang, Salma Waheed Sheikh, Liaqat Shah, and Chuanxi Ma. 2019. "Unraveling Molecular and Genetic Studies of Wheat (Triticum aestivum L.) Resistance against Factors Causing Pre-Harvest Sprouting" Agronomy 9, no. 3: 117. https://doi.org/10.3390/agronomy9030117

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