Next Article in Journal
Nutrient Metabolism Pathways Analysis and Key Candidate Genes Identification Corresponding to Cadmium Stress in Buckwheat through Multiomics Analysis
Previous Article in Journal
Association of Single Nucleotide Polymorphisms of Cytokine Genes with Depression, Schizophrenia and Bipolar Disorder
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Advances to Combat Different Biotic and Abiotic Stresses in Linseed (Linum usitatissimum L.): A Comprehensive Review

1
Department of Genetics and Plant Breeding, College of Agriculture, Rajmata Vijayaraje Scindia Krishi Vishwa Vidyalaya, Gwalior 474002, India
2
Department of Plant Molecular Biology and Biotechnology, College of Agriculture, Rajmata Vijayaraje Scindia Krishi Vishwa Vidyalaya, Gwalior 474002, India
3
Directorate of Research Services, Jawaharlal Nehru Krishi Vishwa Vidyalaya, Jabalpur 482004, India
4
All India Coordinated Research Project on Linseed, Jawaharlal Nehru Krishi Vishwa Vidyalaya, Regional Agricultural Research Station, Sagar 470001, India
*
Authors to whom correspondence should be addressed.
Genes 2023, 14(7), 1461; https://doi.org/10.3390/genes14071461
Submission received: 12 June 2023 / Revised: 11 July 2023 / Accepted: 14 July 2023 / Published: 17 July 2023
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
Flax, or linseed, is considered a “superfood”, which means that it is a food with diverse health benefits and potentially useful bioactive ingredients. It is a multi-purpose crop that is prized for its seed oil, fibre, nutraceutical, and probiotic qualities. It is suited to various habitats and agro-ecological conditions. Numerous abiotic and biotic stressors that can either have a direct or indirect impact on plant health are experienced by flax plants as a result of changing environmental circumstances. Research on the impact of various stresses and their possible ameliorators is prompted by such expectations. By inducing the loss of specific alleles and using a limited number of selected varieties, modern breeding techniques have decreased the overall genetic variability required for climate-smart agriculture. However, gene banks have well-managed collectionns of landraces, wild linseed accessions, and auxiliary Linum species that serve as an important source of novel alleles. In the past, flax-breeding techniques were prioritised, preserving high yield with other essential traits. Applications of molecular markers in modern breeding have made it easy to identify quantitative trait loci (QTLs) for various agronomic characteristics. The genetic diversity of linseed species and the evaluation of their tolerance to abiotic stresses, including drought, salinity, heavy metal tolerance, and temperature, as well as resistance to biotic stress factors, viz., rust, wilt, powdery mildew, and alternaria blight, despite addressing various morphotypes and the value of linseed as a supplement, are the primary topics of this review.

1. Introduction

Flaxseed (Linum usitatissimum L.) is one of the oldest cropsand has been farmed for a long time. In Latin, its name translates to “extremely beneficial” crop, presumably because it is a source of three main bioactive components, i.e., α-linolenic acid, lignans, and dietary fibre [1]. This important crop is an annual herb that produces golden-yellow- to reddish-brown-coloured small, flat seeds. In the beginning, the crop was introduced to the US, predominantly to create textile fibres [2]. The stem creates robust, long-lasting, high-quality fibres, which is perhaps one of the reasons that it was named flaxseed. Flax produced from this crop was primarily utilised in the production of textiles (linen) and paper industries. The name linseed is generally used in terms of oil production and other byproducts, like animal feed [3,4].
L. usitatissimum is the only cultivated species among the roughly 200 species of the genus Linum, and it is farmed for the essential oils in its seeds and for the fibre in the stem that it produces [5,6]. Linseed domestication started with the selection of traits and more effective self-fertilisation, while Mansby et al. [7] claimed that out-crossing was of greater value. Most linseed variants are created by crossing within the gene pool of L. usitatissimum [8,9].
The importance of L. usitatissimum as a multifunctional crop is demonstrated by the fact that various plant parts have various economic uses. Flax fibres, which have two to three times the strength in comparison to cotton fibres, and both fibres can be combined to make linen [10]. The two most popular varieties of flax are brown and golden (or yellow). Both contain equivalent amounts of short-chain fatty acids withhigh nutritional profiles. Whole flaxseed meal, powdered flaxseed, flaxseed oil extracted from the seeds, and partially defatted flaxseed meal are the four most frequent ways to consume flaxseed [11]. Flax “milk” is a new product in the market and establishing itself because it is free from lactose, along with cholesterol. It is a big source of omega fatty acids, fibre, and protein. Flax milk may be a fantastic substitute for dairy milk. It is great for individuals who are intolerant to gluten, nuts, or soy, and it is healthier than almond milk [12,13,14,15,16]. Because of the possible health advantages, particularly those linked to its biochemically active elements, flaxseed has gained significant interest in the area of dietary and health research over the last 20 years [17]. Flaxseeds have been used traditionally in many Ayurvedic preparations since ancient times [17]. In Ayurveda, they are used for “vata”- and “pitta”-related disorders [12]. It is also essential in processes such as virechana (pharmacologic laxative therapy). Hippocrates used flax to treat abdominal pain [18]. Flaxseed is an excellent dietary supplement that deserves more research [13]. On average, flaxseed contains about 30–40% oil, is high in protein at 20–25%, and contains 20–28% total fibre and minerals [1]. Seed oil is nutritionally superior because of its high amounts of α-linolenic acid, omega-3 fatty acid, omega-6 fatty acids, and vitamins A, B, D, and E [15]. These small seeds are believed to aid digestion and lower the risk of cardiovascular disease and especially type-2 diabetes [16]. However, some people reject flaxseed as a dietary supplement because of concerns about anti-nutritional factors [10,19].
For a breeding programme to be successful, diversity is a key component [20,21,22]. Initially, genetic variation evaluations relied on morphological and biochemical indicators, such as isozymes [23,24]. Nevertheless, phenotypic characteristics are labour- and time-intensive, as well as sensitive to environmental factors. Biochemical markers may also not be useful for variation analysis among crop genotypes due to their alteration with environmental conditions [25]. DNA-based markers are stretches of nucleotides and are not influenced by environmental conditions [26]. Random amplified polymorphic DNA (RAPD), inter simple sequence repeat (ISSR), simple sequence repeat (SSR), amplified fragment length polymorphism (AFLP), sequence-tagged site (STS), and single-nucleotide polymorphism (SNP) are among the various molecular markers being used for agricultural plant genetic variation analysis [27]. Except for SNPs, other markers do not require in-depth genomic information and are simpler, less costly, and less labour-intensive than other DNA marker approaches [10]. Some studies are available on applications of markers for diversity analysis in linseed. Such markers include isozymes [28], RAPD [29,30,31,32,33], AFLP [34,35,36,37], ISSR [38,39,40], and SSR [41,42].
The basis of linseed genetic improvement and cultivar development is the conventional breeding technique, which has resulted in the release of new cultivars with long-lasting disease resistance, improved yield stability, and better agronomic performance [43]. However, molecular breeding [44,45] has also been applied for this purpose. A restricted genetic base was employed to develop a Canadian linseed variety [46,47,48]. The absence of related species for incorporating novel genes has hampered yield and quality innovations, hindering linseed competitiveness [11].
In the years to come, climate change may put enormous pressure on the breeding society to breed varieties [49] having adjustableproperties underadverse climatic conditions. For any crop to be improved or adjusted to changing environmental conditions or market demands, access to genetic variability is a must, and the availability of diverse germplasm lines of the targeted crop plays a significant role in it [10,50,51,52]. A variety of germplasms, including wild relatives, landraces, commercial, extinct, and elite varieties, pure and breeding lines, mutants, polyploids, and hybrids, are found in gene bank collections [11] and can be exploited for the improvement of the crop. There are 46,513 linseed/flax accessions documented in global germplasm collections, of which L. bienne (the wild ancestor of farmed flax) is only intermittently represented in gene banks [53]. Breeders can use these plant genetic resources of linseed for the improvement of linseed crops with potential climate adaptations [54,55]. Following the preliminary assessments of flax germplasm collections, subsequent evaluations for different characteristics associated with tolerance to different biotic and abiotic stress factors were carried out [56,57], alongside the present concentration on the germplasm screening of single-gene attributes, such as disease resistance [58]. As a result, the extensive diversity found in linseed germplasm collections, along with data on characterisation and assessment, is a potential source that is worth using in breeding for abiotic and biotic stresses. The current review focuses on all the advancements made in the flax research programme to overcome these stresses.

2. Abiotic Stress

To achieve agricultural sustainability, it is critical to develop and nurture products that are tolerant to rising abiotic pressures caused by climate variability [49]. Numerous abiotic and biotic stressors that constantly threaten plants have an impact on their outputs. Through intricate endogenous signalling networks and numerous modifications, the plant reacts to these stressors. Under these environmental circumstances, the plant’s output and reproductive success are determined by the interactions between these networks. Linseed, like other crops, if subjected to a variety of abiotic stresses, might have reduce yields [59]. The different abiotic stresses influencing linseed are summarised in the following subsections.

2.1. Drought Tolerance

Drought is considered one ofthe most pervasive and harmful abiotic factors that impact agricultural output. Drought, on average, affects crop growth, production, and quality by more than 50% [60,61,62,63]. By 2050, nearly fifty percent of land that is usable for agriculture is anticipated to have disastrous effects on plant development [64]. Soil moisture deficiency can have a substantial impact on the flax output capacity, oil quantity and fatty acid arrangement, and fibre quality indicators [65,66,67].
According to Hu and Xiong [68], drought causes normal metabolism to be disrupted by decreased leaf growth, oxidative damage, increased membrane lipid peroxidation, leaf senescence, and abscission. Flax can tolerate drought conditions more effectively than numerous other oil and food crops due to its hardiness [69]; however, flax plants lose a lot of water due to their high transpiration coefficients, which range from 787 to 1093 [70]. To generate optimal yields, fibre flax requires at least 600–650 mm of yearly precipitation, with at least 110–150 mm falling during the vegetative season [71,72]. Since drought is an unpredictable and irregular environmental event, genotypes with high yield potential and drought tolerance may be selected [10].
To overcome these limitations and enhance flax yields, research activities based on conventional breeding and later transgenic approaches were initiated [73]. Understanding the adaptation processes and identifying the underlying genes, markers, and QTLs might help to improve the genetics and production of linseed in semi-arid and arid areas because drought tolerance is a complicated polygenic trait [49]. A limited number of studies have been published identifying the accessions of flax that are resistant to drought [74,75,76] and genome-wide analyses of drought-induced gene expression [72]. Linseed has a shallow root system as compared to other oilseed crops, such as safflower, rapeseed, and sunflower. Therefore, understanding the root system architecture is crucial for improving flax’s ability to absorb water. Many crops, including rice, wheat, and maize, have recently demonstrated the significance of specific root characteristics for efficient water and nutrient absorption under water-stress conditions [77,78,79,80], but knowledge is still somewhat limited in flax. Among different morpho-physiological characteristics, plant height, biomass, seed colour, lignin content, seed yield, leaf absolute water content, and leaf relative water content have been found to be associated with drought tolerance in linseed [81,82].
Plants have many adaptations for surviving in conditions of drought and/or water shortage [83,84]. It is crucial to comprehend these processes in order to create agricultural plants that are resistant to such circumstances [85]. Quéro et al. [86] tried to investigate the processes underlying b-aminobutyric acid (BABA)-induced drought adaptation in plants. BABA causes a reorganisation of the solute content in flax leaves that results in an increased accumulation of proline and nonstructural carbohydrates, as well as a drop in inorganic solutes, according to metabolomic and ionomic profiling of the leaves [59]. In one of the previous studies, it was discovered that BABA therapy caused alterations that made flax plants more resistant to drought stress [86]. Under various irrigation regimes, Ansari et al. [87] investigated the relationship between the crop and mycorrhizal fungi. In a recent study conducted by Liu et al. [11], it was discovered that, in contrast to various plant species, flax mycorrhizal fungus may promote development under both stress and non-stress conditions.
Shivaraj et al. [59] explained the effects of various amounts of titanium dioxide (TiO2) at the nanoscale (10–25 nm) on flax development under stress and non-stress circumstances. It was discovered that TiO2 at modest concentrations facilitated the development of flax in water-scarce environments. Additionally, flax plants treated with nano-TiO2 at a concentration of 100 mg L−1 contained higher seed oil and protein contents. As a result, it has been concluded that the application of nano-TiO2 particles at modest concentrations can reduce the harm caused by drought stress to flax plants as well as increase drought tolerance. Crops that are resistant to duress are developed using both forward and backward genetic methods. In the same setting, the drought-responsive element binding protein 2A (DREB2A) gene was inserted into the plant genome to create a flax cv. Blanka cell line that is drought-tolerant [73].

2.2. Salinity Tolerance

Numerous variables, such as obscene irrigation, rock weathering, little precipitation, ion exchange, high surface evaporation, and poor cultivation practices, have contributed to the exponential increase in soil salinity in recent years [88,89,90]. According to estimates from Shrivastava and Kumar [91], greater than fifty percent of arable land will be salinised by around 2050. Currently, salty conditions impact around 33% of irrigated land and 20% of all cultivated land. According to Dubey et al. [88], soil salinity–alkalinity in flax causes delayed germination and emergence, reduced seedling survival, erratic crop development, and decreased yield. In a few investigations, flax germplasm was tested for resistance to salinity–alkalinity stressors, and salinity-tolerant lines were found based on biomass, germination, seedling traits, and the K+/Na+ ratio [92,93,94,95,96]. Wu et al. [97] discovered genes in flax that increase the root length, improve membrane damage, and alter the ion distribution to confer salt tolerance. Flax may be cultivated on agricultural terrain where other crops cannot since it can withstand pH levels up to 9.
All agricultural plants face genuine threats from rising soil salinity and alkalinity, among which alkaline-salt stress is more harmful than neutral-salt stress. Considering this, understanding the processes that control plant resistance to saline–alkaline stress has emerged as a hotly debated area of plant science. The impact of these circumstances on the sprouting of the flax plant was examined by Guo et al. [98]. In three experimental circumstances, they discovered that germination decreased with a rise in the ionic concentrations for each of the 10 common flax genotypes they examined. As anticipated, it was discovered that low-concentration treatments with neutral salt and alkaline salt had little impact on germination in all varieties [11].
All varieties of plants responded the most negatively to alkaline-salt stress, and higher concentrations of this treatment completely prevented germination. When five linseed genotypes were developed after being treated with neutral salt and alkaline salt, it was discovered that they had greater salt tolerance [98]. Digital gene expression in flax plants under alkaline-salt stress, neutral-salt stress, and alkaline stress was examined by Yu et al. [99]. In their experiment, it was discovered that under neutral-salt stress, carbohydrate metabolism was impacted, while under alkaline-salt stress, photosynthesis and thereaction to biotic stimuli were badly compromised. In some of the studies, the differential expression of important factors was implicated in responses to abiotic stressors, like WRKY, abscisic acid (ABA), ion channels, and mitogen-activated protein kinase (MAPKK) [10,59]. More of these differentially expressed genes were triggered by alkaline-salt stress than by alkaline stress or neutral-salt stress, indicating that a greater number of genes are involved in regulating the alkaline-salt stress pathway.
Salinity stress has a molecular foundation that is still poorly known. Therefore, Yu et al. [100] used deep sequencing to analyse small RNAs and the degradome in samples that had been exposed to three stress conditions in order to understand the genetic basis of salinity tolerance in flax. Small RNA target genes that control reactions to cues were discovered to be induced. Using transcriptome analysis and degradation genome sequencing, 29 showed opposite expression patterns. The development of climate-smart flax is progressingforall stressors, and 243 miRNA–target combinations have been identified. Two miRNAs, miR398a and miR530, have been reported to be linked to salt stress tolerance in flax. Furthermore, in a separate experiment, Wei et al. [101] discovered that flax seedlings of HIZ019, YOI254, and Tianxin3 had higher salt tolerance because of some important biochemical alterations in flax seedlings under salt stress.

2.3. Heavy Metal Stress Tolerance

Heavy metal buildup in vegetation is caused by heavy metal contamination of the earth. Heavy metals in the dietary chain are consequently biomagnified. One such heavy metal that can be detrimental to both plants and animals, even in minute quantities, is cadmium. Plants have evolved specialised regulatory mechanisms that allow them to bind metals to compounds like phytochelatins and metallothionein to create complexes (metallothionein group III). Numerous studies have demonstrated that linen clothing tolerates cadmium pollution well. Even though the toxic effect of cadmium restricts its use for both food and medicine, the use of metal-accumulating plants for the phytoremediation of contaminated soils opens a novel and hopeful path towards improving the resilience of their cultivars and variants to Cd stress [59]. Over the past ten years, extensive research has been conducted on the mechanisms that drivecadmium tolerance in plants. High Cd concentrations reduced root development and, to different degrees, increased the generation of membrane permeability, hydrogen peroxide (H2O2), protein oxidation, and lipid peroxides. Additionally, it was discovered that there had been a substantial change in the effectiveness of antioxidant and scavenging enzymes.
The impact of salicylic acid on the antioxidant defence system in flax plants was examined by Belkadhi et al. [102]. Salicylic acid, on the other hand, was discovered to lessen the toxic effects of Cd on the lipid composition of membranes, the antioxidant system, and root development [11]. Belkadhi et al. [103] discovered that a preliminary treatment with salicylic acid preferentially safeguarded plastidial lipids due to greater amounts of polyunsaturated fatty acids in flax. These findings imply that under Cd stress, salicylic acid-pretreated flax plantlets had increased membrane integrity. Salicylic acid boosts root antioxidant mechanisms in flax. Kaplan et al. [104] investigated the effects of cadmium stress on the fatty acid composition of flax in the presence of mycorrhizal fungi. They discovered that seeds from plants cultivated in mycorrhizal fungi contained more unsaturated (18:1, 18:2, and 18:3) fatty acids overall. At 15 ppm of Cd, these impacts became more obvious (the amounts of 18:1, 18:2, and 18:3 were increased by 169, 370, and 150%, respectively). These findings indicate that once the level of Cd in seeds hits a certain level, this heavy metal increases the effectiveness of the enzymes that control the transformation of saturated fatty acids into unsaturated fatty acids.
Although zinc (Zn) is a crucial ingredient needed for the healthy growth and development of plants, too much of it can also be harmful. Significant attempts have been made to comprehend the tolerance of plants to zinc in flax. Grant et al. [105] discovered that flax cultivars are better at transferring Zn through shoots to seeds than Cd. The use of plant species for phytoremediation is determined by the transfer of metal ions to seeds. Plants are also using a variety of biotechnological techniques to limit the absorption or transfer of the metal ions in plant cells. Smykalova et al. [106] investigated how flax varieties could collect and move Ca and Zn in an in vitro culture made from hypocotyl tissues. The flax varieties discovered by Smykalova et al. [106] are those that are tolerant to zinc, and increasing the amount of zinc and decreasing the amount of cadmium in the kernel will be crucial for increasing the nutritional value of linseed.
The toxic effects of heavy metals like copper (Cu), zinc (Zn), cadmium (cd), nickel (Ni), lead (Pb), chromium (Cr), arsenic (As), and cobalt (Co) on seed germination in various varieties of flax have been assessed in another important research conducted by Soudek et al. [107]. Various heavy metals manifested their harmful effects in the following order: As > Cu > Cd > Co > Cr > Ni > Pb > Zn. In fact, a lot of these studies point to the effective use of flax’s potential for phytoremediation, especially when cultivated for fibre production. The growing of flax on heavy-metal-contaminated soils may be made easier with a thorough grasp of the mechanisms underlying heavy metal tolerance [59].

2.4. Cold Stress Tolerance

Many cool-season perennial crops’ outputs are influenced by two significant physiological processes: vernalisation and photoperiodism. Darapuneni et al. [108] examined the flowering response of various flax genotypes under two photoperiod and vernalisation conditions in growth chamber research. The findings imply that the photoperiod, vernalisation, and genetics significantly influence early blooming in flax. For frigid climate regions like the Upper Midwest of the US and Canada, the early-blooming trait is more important. However, owing to the hot spring and summer temperatures in some places, such as Texas, flax is produced in the autumn. Flax varieties demonstrated genotypic interactions with the photoperiod and vernalisation.
In particular, Texas flax genotypes of the winter variety were susceptible to both vernalisation and blooming photoperiods. The flowering times of most other spring-grown flax genotypes were unchanged by the vernalisation treatments, whereas Texas genotypes postponed anthesis for 7 days or longer in unvernalised seedlings [10]. The majority of other genotypes were unaffected by circadian rhythms (photoperiodism)in vernalised seedlings, while Texas cultivars delayed anthesis for a period of twelve days or longer under vernalised and short-day conditions. Screening for vernalisation and photoperiodic sensitivity in Texas genotypes, as well as the introgression of these characteristics into newly adopted spring-grown genotypes, is required for the development of high-yielding flax genotypes for production sites in the southern Great Plains [59].

2.5. Heat Tolerance

Particularly in tropical and subtropical climates, heat stress negatively impacts physiological processes, development (Figure 1), growth, and yield [109]. According to several studies [110,111,112,113], a prolonged duration of heat stress (40 °C for 5–7 days) in relation to flowering duration may have a substantial effect on flax’s boll development, pollen viability, pollen production, flowering, seed oil quality, quantity, and seed set. Flax fibre does not need high temperatures. Flax grows best and produces the highest-quality fibre under circumstances that are damp, overcast, and moderately chilly (18–20 °C). High temperatures, especially terminal heat, which is limiting for flax growth, cause poor adaptation of superior fibre linseed genotypes to hotter climates. Although there have not been many studies on how higher temperatures affect flax development, physiological processes, and yields [113,114], the molecular dissection is also unclear.

3. Biotic Stresses

Biotic stresses are responsible for major yield losses [115]; fungal infections, a few viruses, and a phytoplasma are the principal causes of flax biotic stresses. Neither bacteria nor nematodes are known to cause major stress in flax. All varieties of flax (L. usitatissimum L.), such assolin flax (low linolenic acid/vegetable oil flax), linseed (industrial oilseed flax), and fibre flax, are susceptible to fungi. The varieties of each species of flax may react differently to certain diseases or races of the same pathogen. In the parts of the globe where flax is grown, flax diseases vary in frequency, severity, and relevance from one location to another. The most feasible and effective method for engineering disease resistance in crop species is through breeding. Early attempts to enhance flax focused on breeding disease resilience. The morpho-physiological symptoms after biotic stresses in linseed are the yellowing of leaves, a reduced number of seeds, premature ripening (leading to yield loss), and early plant death [116]. Three types of resistance have been found according to the number of genes that influence inheritance: Resistance can be controlled by one gene (monogenic), a few genes (oligogenic), or several genes (polygenic) [117].

3.1. Rust Resistance

Flax-growing places around the globe have documented cases of rust. Melampsora lini (Ehrenb.) Desm., a fungal organism, is the causal agent of rust in flax crops. Rusts have the potential to be the most destructiveplant diseases, and flax rust is no exception [118]. Early infections and fast disease development, which are made more likely by meteorological conditions, can totally decimate flax plants and result in a significant reduction in the production and quality of both fibre and seeds. The main difficulty in producing resistant cultivars is the regular emergence of new races [119] of the pathogen. The availability of genetic resistance sources for most flax-breeding programmes globally has been credited with the disease’s rather good global management.
Bright orange and powdery pustules that appear on leaves, stems, bolls, and other aerial plant parts are the primary indications of rust in flax. On leaves and bolls, these pustules are often round, although they are elongated on the stems. Depending on the interplay between a person’s ethnicity and genetic makeup, chlorotic or necrotic zones may develop surrounding the pustules. Severe outbreaks result in the leaves drying out and withering, severely defoliating the plants. Numerous urediniospores are produced by rust pustules, and they easily detach and travel long distances in the wind. The black telia pustules that develop from the maturing plants’ orange pustules (uredia) create hardy teliospores that can survive the winter [120].
In contrast to many other rusts that need different hosts, flax rust is an autoecious rust, meaning that the fungus may go through all four phases of its life cycle—pycnia, aecia, uredia, and telia—on the flax plant (Figure 2). Field inspectors seldom ever see the pycnia and aecia, which typically occur on cotyledons and lower leaves early in the season. The stages of pycnia and aecia are crucial for the fungus to complete its sexual life cycle and for the emergence of new races. The uredia are this fungus’ third developmental stage. Since uredia develop cycles of urediniospores that might lead to new infections with each cycle during the growing season, the uredial stage is the most harmful to the crop. When the fourth stage reaches maturity, it generates tough overwintering teliospores that can endure inclement weather [120].

3.2. Fusarium Wilt Resistance

A typical flax biotic stress is Fusarium wilt, which is brought on by Fusarium oxysporum. The fungus kills plants shortly after they appear, while later infestations result in the wilting and yellowing of foliage [121]. Agrawal et al. [122] demonstrated that flax wilt resistance is acquired by recessive alleles based on F2 segregation ratios among nine intervarietal crosses comprising the flax wilt-resilient cultivars R552 and RLC6, along with four susceptible genotypes. Spielmeyer et al. [123] investigated the transfer of Fusarium wilt resistance in a recombinant double-haploid (DH) population produced from F2 seeds of a wilt-resistant (twinning Linola TM) and a wilt-prone (Australian flax cultivar Glenelg) line. The partition of two separate genes with additive effects, which accounted for 38 and 26% of the genetic diversity for Fusarium wilt resistance, was considered responsiblefor the phenotypic variance.
Spielmeyer et al. [124] used an amplified fragment length polymorphism (AFLP) marker to assign QTLs to two linkage groups. The flax varieties Mogilevsky-2, Torzhoksky 85, Kievsky, Kalininsky, Rodnik, Nika, K-65 Niva, T-17, and Ustiensky were discovered by Portyankin and Karachan [125] to be resistant to Fusarium wilt. Flax accessions from East Asia and North and South America had greater than average and mean resistance to Fusarium wilt, respectively, but flax accessions from Europe and the Indian subcontinent had lower than average resistance. This variation in resistance to Fusarium wilt was documented by Diederichsen et al. [126].

3.3. Alternaria Blight Resistance

Most of the world’s flax-growing regions have documented flax disorders linked to infection by Alternaria spp. [127]. Most reports come from India [128] and European countries where the crop is produced. The disease damages leaves and flowers as well as seedlings, producing seedling blight. In India, yield losses of up to 90% have been documented [129], and in the UK, yield losses of up to 35% have been documented [130].
According to Vloutoglou [131], the three most common species on flax are Alternaria infectoria Simmons, A. linicola, and A. alternata. Additional species include A. linicola Skolko and Groves, which is endemic to India and loves temperatures ranging from 26 to 33 degrees Celsius, as well as humidity [132]. In flax waste on and in the soil, A. linicola persists as thick-walled chlamydospores [131]. Flax seedlings are infected by early-season conidia. Infected seeds can spread the fungus, which results in seedling blight [131]. The seedling blight or brown rot caused by the seed-borne inoculum is the first sign of A. linicola. Stunted seedlings that have dark red lesions on their hypocotyls and cotyledons collapse completely in one to two weeks. A rich reservoir of inoculum is created by the diseased and dying seedlings, which spreads to infect healthy plants. On diseased leaves, dark brown lesions develop and typically combine to cover the whole leaf, which then becomes chlorotic and then dies [120].

3.4. Powdery Mildew Resistance

Powdery mildew, another significant flax disease caused by the fungus Oidium lini, was first observed in Western Canada in 1997 [133]. Ashry et al. [134] examined flax cultivars for resilience to powdery mildew. They discovered that choosing flax as a fibre variety leads to a rise in powdery mildew incidence. Choosing seed-type flax has the opposite effect, lowering resilience to powdery mildew. When Rashid and Duguid [135] investigated the genetics of resistance to powdery mildew in flax plants, they discovered that resistance to the fungus is caused by a single dominant gene (PM1). The output of flax crops is also known to be impacted by the Pasmo illness. In Western Canada, Septoria linicola (Speg.) Garassini has been reported to cause Pasmo disease and is common and pervasive on flax [136]. Most varieties of flax in Canada are vulnerable, and little is known about how S. linicola isolates are pathogenically dangerous. High-humidity and high-temperature regions show rapid illness progression. S. linicola thrives on agricultural leftovers that have been infected and left in the field from one season to the next.

4. Flax Genetic Resources

4.1. International

Dillman [46] underlined the importance of collecting germplasm during initial linseed breeding. Further, Zhuchenko and Rozhmina [137] described the use of landraces for fibre flax breeding. Over 48,000 flax accessions are thought to be present in ex situ collections globally, with 10,000 of them possibly being unique. The range of genetic variety maintained in the separate flax national collections (national inventory) of each unique European country, as well as the breeding techniques utilised, are essential for successful breeding activity. The European Cooperative Programme for Crop Genetic Resources (ECP/GR) network adopted the first step towards generating an inventory of the European flax gene pool at an informal meeting of flax-gathering owners within the recently built the Industrial Crops and Potato Network, among the most recent European System of Cooperative Research Networks in Agriculture (ESCORENA), held in Prague on 7 and 8 December 2001 [138].
The 2001 inventory of European flax genetic resources revealed the status. The collection of Bulgarian flax is housed in two institutions. A total 945 accessions were kept at Sadovo in 2001. Of the 450 studied accessions, 23% were linseed, 29% were flax, and 46% were of an intermediate kind. Old and landrace cultivars provided 58% of the total, followed by modern cultivars (15%) and breeder’s lines (14%). There were 12 wild species of Linum in the collection other than L. usitatissimum, namely, L. austriacum, L. bienne, L. flavum, L. punctatum, L. grandiflorium, L. perenne, L. setaceum, L. strictum, L. trigynum, L. humile, L. altaicum, and L. viscosum [139]. Moreover, the AgroBioInstitut (ABI) in Kostinbrod maintained 283 accessions. Outof these, 178 came from IgreSadovo, 29 were from other Bulgarian institutes, 71 were from abroad, and 5 were self-bred specimens. These accessions came from Europe, Asia, Africa, and America, respectively. The collection is made up of 54% cultivars, 27% landrace and primitive cultivars, 15% breeding lines, 1% wild forms, and 2.5% genetic stocks. By plant type, the collection was constituted by 31% fibre flax, 35% linseed, 32% combination type, and 2% other types. According to the International Union for the Protection of New Varieties of Plants [140,141], the International Flax Database (IFDB) descriptors [142,143], and other species descriptors, 259 accessions for 15 morphological traits, 5 biological characteristics, and 4 yield traits were described [144,145].
The assessment, description, and characterisation systems for flax germplasm vary depending on the needs of breeders in each country, whether it is in Europe, the USA, or Canada. To exploit the genetic diversity of flax for breeding purposes, a good, unique approach for evaluation, description, and characterisation is necessary, which should be standard at least across European nations. This is because different evaluation and characterisation techniques are employed in various European genebanks as well as outside. Thus, the IFDB may be viewed as the first input for this goal [143]. The IFDB was initially overseen and managed by Agritec Ltd. (Viale delle Terme di Caracalla, 00153 Roma (Italy)) starting in1993 in the framework of FAO ESCORENA Flax and Other Base Plants Network (FAO FOBPN) [146] and then, over time, the framework of IPGRI Coordination Group Network for Sugar, Starch and Fibre Crops (CGNSSFC), currently namedSugar, Starch, Fibre Crops & Aromatic Plants Network (CGN SSFC&APN) at Bioversity International. In theACCESS IFDB framework, it currently contains passport information for 8385 accessions of 11 collections from 10 nations. It is estimated that 37% of the accessions are unique based on a study of the provided data. It also showed that there were significant variations in the collections’ description fill rates. Table 1 displays the findings of current germplasm characterisation and assessment research carried out at the Plant Gene Resources of Canada (PGRC), Canadian National Seed Gene Bank [147].

4.2. National

Cultivated forms have been reported to be rich in diversity in the Indian region. A total of 2627 accessions, comprising 2495 indigenous, 51 released, and 132 foreign varieties, were preserved in long-term and medium-term storage at the National Bureau of Plant Genetic Resources (NBPGR), New Delhi, India. According to the information shown below, the Central Variety Release Committee (CVRC) and States Variety Release Committee have released and notified 51 improved varieties for general cultivation in various agro-climatic conditions. Table 2 lists the different varieties of linseed grown in different climatic regimes of India [147]. Table 3 contains list of linseed varieties that are tolerant to various abiotic and biotic stresses.

5. Breeding through Conventional Methods

Collections of flax and linseed genetic resources are researched for breeding purposes, and the selection of the first practical parental patterns is commonly debated. According to Tadesse et al. [148], who conducted research on crossings between high-yielding and low-yielding flax components, it may be beneficial to hybridise low- and high-yielding flax lines in order to increase genetic diversity and produce high-yielding flax lines. According to the findings of Pavelek [149], crosses of high-yielding flax components may be productive when the genetic variability of the parental lines is substantial. The largest linseed growers and producers in the world are India, the USA, and Canada. Consequently, increasing seed production, increasing oil content and quality, snowballing resilience to diseases and pests, and developing early-maturing accessions are the key breeding objectives.
Today’s breeding efforts are mostly focused on traits that stabilise yields, such as resistance to fungal infections, and qualitatively novel contents of materials. Mutagenesis [150,151], genomic approaches [152], and other methodsare used to develop new breeding techniques; for example, anther culture, mutation breeding, molecular-marker-based analysis [153], and techniques of transformation can be employed to find a solution when the genetic variability’s options have been exhausted and are no longer available. For breeding fibre flax, the most crucial factors are fibre content and resistance to Fusarium and scorch, whereas for linseed, the most crucial factors are seed production, fatty acid composition, and resistance to Fusarium and rust [154].

5.1. Interspecific Hybridisation

Although many interspecific crossings have been made, only those involving species with a similar number of chromosomes have so far proved beneficial. The F1 offspring of hybrids between L. usitatissimum and a few other species with n = 15 chromosomes are robust and fruitful, and they have been exploited to introduce certain desired traits, such as resistance to linseed rust.
Through modifications to the pollination of the cultivated species, interspecific hybridisation can also increase genetic variability. Some Linum species, including L. grandiflorum, L. anstriacum, and L. perenne, are incompatible with one another. Numerous attempts were made to cross the cultivated linseed (L. usitatissimum) with the self-incompatible species without success. When L. usitatissimum was employed as the female parent in crosses, abortive seeds that were dried up and unable to germinate were obtained [155]. The embryo of the cross between L. usitatissimum and L. grandiflorum degenerates after 7 days due to somatoplast sterility, according to embryological investigations. By cultivating the embryos of the cross and transmitting the incompatibility mechanism to the cultured species, attempts are being made to raise the hybrid. Induced self-incompatibility would force cross-pollination and raise genetic variation in farmed plants [147].

5.2. Hybrid Linseed

Knowledge of floral morphology and pollination methods is necessary to produce hybrid varieties with vigour or heterosis effects. Flax heterosis has ranged from 40% [156] to 231% [157]. The development of flax and linseed hybrids is less common than that of sorghum, maize, sugar beet, onions, or pear millet. The primary cause is that the male sterile flax flower has a tiny corolla that does not open to facilitate cross-pollination. Several researchers have investigated the issue of the male sterile flax, including Kumar and Singh [158,159], Dubey and Singh [160,161], and Kumar [162]. Particularly in India, several researchers have looked at the compatibility of flax or linseed kinds in diallel crosses, allowing for the assessment of the heterosis impact. Pant and Mishra [163] found positive and substantial heterosis for seed yield per plant at 40 crossings and fibre yield at 13 crosses. This impact frequently results from dominance, non-allelic interaction, or the combined influence of additive genes.
Trait selection is influenced by the relationship between the genes expressed by either the general combining ability (GCA) or the specialised combining ability (SCA). In contrast to the nonadditive effects of SCA, which were confirmed by Sood et al. [164] for number of seeds per capsule and seed weight, the GCA of flax and linseed is related to traits like the number of capsules per plant, plant height, the weight of capsules per plant, and seed yield. There were occasionally both additive and nonadditive gene effects seen [165].

5.3. Mutation Breeding

Despite Tammes reporting a mutation in linseed as early as 1925, nothing is known about the nature, induction, and use of induced mutations in flax. Mutagenesis in linseed has been attempted using both physical [166,167] and chemical mutagens [168,169]. Deshpande [170] identified flax seedlings with low levels of chlorophyll in the NP-12 normal group. Levan [171] discovered three families segregating for chlorophyll mutations in diploidplants but not in tetraploid plants following X-irradiation. Three different linseed varieties, Hira, Mukta, and Neelum, were exposed to radiation of10, 20, 30, and 40 kR by Rai and Das [167].
Along with various kinds of chlorophyll mutations, an increase in mutation frequency was seen with increasing irradiation dosages. The observedcholorophyll mutants showed delayed blooming, shorter plants, and fewer capsules per plant. In contrast to normal plants, these mutants had a much higher number of non-bearing tillers per plant. Sharov [172] reported mutants with good yield and a 52–56% increase in resistance to F. oxysporum following chemical mutagenic therapy. After irradiating the variety Neelum, George and Nayar [166] acquired a dwarf mutant (TL-1), which developed 30 days quicker than the typical plant. It is interesting to note that the mutants had larger 1000-seed weights and lighter-coloured oils with higher oil contents.
In Australia and New Zealand, three new linseed cultivars, known as “Wallaga”, “Argyle”, and “Eyre”, have been introduced under the generic name “Linola”. Golden-yellow seeds are seen in “Linola” varieties. “Eyre” is a result of a single-plant selection made in the F4 of the cross “Glenelg”/CPI 84495//4*”Zero”, from which an F8 bulk was obtained. The EMS (ethyl methane sulphonate) mutagenesis of the Australian linseed cultivar “Glenelg” and the recombination of two mutant genes produced the low-linolenic acid genotype “Zero” [173].
Eight Indian states, namely, Madhya Pradesh, Bihar, Gujarat, Delhi, Punjab, Rajasthan, Uttar Pradesh, and West Bengal, are focused on linseed breeding [174]. India’s linseed breeding industry is well established and well organised. The Indian Central Oilseeds Committee (ICOC) increased its linseed research activities in 1947, and then in 1958, the project “Intensification of Regional Research on Cotton, Oilseeds and Millets” (PIRCOM) took over. The All India Coordinated Research Project on Oilseeds (AICRPO) was launched by the Indian Council of Agricultural Research in 1968 [174] utilising international methods.

6. Advanced Methods in Linseed Breeding

6.1. Molecular Diversity

Flax genetic diversity can be assessed using a variety of methods, including physical traits, isozymes, and molecular markers. While isozymes are scarce, morphological characteristics are frequently more quantitative and reliant on their surroundings. The prevalence and distinctiveness of DNA markers are some of their benefits. They are helpful for identifying varieties and assessing DNA diversity. To examine the genetic variation in flax, various molecular markers, such as restriction fragment length polymorphism (RFLP), random amplified polymorphic DNA (RAPD), amplified fragment length polymorphism (AFLP), and simple sequence repeat (SSR), have been developed.
By using RAPD markers, Fu [29] evaluated the geographic trends of flax variability in the global collection of farmed flax. A total of 149 RAPD markers were measured for each of the 2727 flax accessions, which came from 63 different nations and 1 unidentified group. The accessions were divided into 12 main areas, which explained 8.2% of the RAPD variation. The most varied accessions came from East Asia and European areas, but the most distinctive accessions came from the Indian peninsula and Africa. Genetically, accessions from the West Asia area were less related to those from the Indian peninsula and more related to those from Africa. These discoveries are crucial for comprehending the domestication of flax and are also helpful for categorising the intraspecific variety of cultivated flax, identifying the fundamental subgroup of the flax collection and looking into novel gene sources for flax improvement.
Cloutier et al. [175] conducted the first comprehensive study on the production and analysis of a considerable number of SSR markers in flax. The genetic relationship among 635 genotypes was discovered by using 275 variable EST-SSRs. The pedigree connection between the accessions was associated with subclusters within the major clusters. The expressed sequence tags-signature sequence tags (EST-SSTs) created here signal the beginning of SSR markers in flax on a wide scale. They could be applied to the creation of physical and genetic maps, studies of genetic variation, association mapping, and cultivar identification.
Several molecular methods are presently available to identify germplasm at the DNA level. Due to the large number of markers that can be produced per analysis, AFLPs are typically regarded as being comparatively strong in germplasm analysis when compared to alternative genetic marker systems. In the research by Van Treuren et al. [35], duplicate germplasm in the flax collection of the Netherlands’ Center for Genetic Resources was identified using AFLPs. ANOVA was used to compare accessions in pairs in order to find duplicate genotypes. The 29 accessions of breeder’s lines could be reduced to 14 by stepwise bulking accessions until all surviving accessions were noticeably distinct. This bulking strategy only had a negligibly tiny negative impact on the among-population component of variance, demonstrating a 2.6% decrease. This outcome is addressed in connection to enhancing collection management effectiveness.

6.2. Association Mapping

Association mapping studies are crucial throughout animal and plant research because they aid in the discovery of genes that influence phenotypic variation and create the framework for the recognition of genes that play a vital role in what is seen as variation. Genome-wide association studies (GWASs) have become a more common practice in the quickly developing genomic age. The early accessibility of complete genome sequencing data in humans appears to have contributed to the broad application of GWASs throughout human research before it was adopted by plant science. Several plant genomes have been mapped as a result of the significant advances in sequencing technology over the past ten years. The availability of plant whole-genome sequences gave researchers the chance to perform effective GWASs [59].
Traditionally, biparental crosses, which are very simple to create, have been used to identify quantitative trait loci (QTLs) in plants. Biparental hybrids do have some restrictions. Since the community is descended from two parents, it only works with genetic variance within the parental lines, which is the most important restriction. Additionally, biparental QTL mapping has poor precision. A GWAS, in comparison, has several advantages over QTL mapping, including high resolution and the use of genetic variation present in readily available germplasm resources. The genetic makeup of the targeted material, phenotypic diversity, and the quantity and distribution of molecular markers have a significant impact on the effectiveness of GWAS.
Single-nucleotide polymorphism (SNP) markers have all the characteristics needed for high-throughput genotyping and are widely available and evenly dispersed throughout the complete genome. The majority of GWASson plants are conducted on important species like rice, maize, and wheat, as well as model species like Arabidopsis. Limited attempts have been made to perform GWASs on flax compared to the main food products and model species. Since flax is an oilseed grain, its breeding programme focuses mainly on increasing the seed oil content and total output. Secured output is heavily reliant on climatic factors that exert various biotic and abiotic pressures [59].
Plant resistance against stresses is governed by a sophisticated polygenetic system. GWAS is a cost-effective and effective way to analyse such intricate genomic control. For the recognition of significantly associated loci and subsequent candidate gene identification, an in-depth knowledge of the level of linkage disequilibrium (LD), the profile of LD variations across the entire genome at substantially associated loci, and functional annotations of predicted gene models is required [59].

6.3. Extent of Linkage Disequilibrium

The degree of LD differs significantly between various chromosomes and loci on the same chromosome, and it is not consistent across plant species. Population drift, selection, admixture, inbreeding, and taming are some of the factors that influence LD. The rate of recombination, transversion, translocation, and chromosomal duplications all play a significant role in the LD variation at various genetic locations. Numerous plant taxa have been the subject of in-depth research on LD decays. Overall, species have been found to cross-pollinate more frequently and experience very little LD degradation. For instance, it has been found that LD decays occur within 0.3–2 kb [176,177]. However, self-pollinated species like soybean have shown evidence of LD decay after extremely lengthy distances [178]. According to Sonah et al. [178], distinct LD decays at r2 = 0.2 were observed in soybean across various chromosomes, spanning from 250 kb to 2.5 Mb on average. Similarly, the flax genome exhibits comparatively delayed LD decays relative to cross-pollinated species since it is a self-pollinated plant species. Soto-Cerda et al. [179] used genetic information from 448 microsatellite markers and 407 flax accessions from around the world in a genome-wide analysis. According to Soto-Cerda et al. [180], LD declined across the entire genome within an average of 1.5 cM. Like the previous studies, [181] carried out two additional genome-wide investigations using 460 SSRs for 390 accessions and 112 SSRs for 407 accessions, respectively, and found LD decay up to 1.8 cM. Aside from the availability of the whole-genome sequence, insufficient attempts have been made to carry out an effective LD analysis and consequent GWAS.

6.4. Genetic Loci Identified by GWAS

One of the main goals of flax breeding programmes is to increase the amount of linolenic acid (LIN) in the plant. It is difficult for breeders to increase LIN content without influencing other fatty acids (FAs) because LIN content has negative correlations with linoleic acid (LIO), stearic acid (STE), palmitic acid (PAL), and oleic acid (OLE) [182]. Some flax accessions with high LIN were discovered through careful evaluation of a wide range of germplasm, but they are not well suited to cropping systems [183,184]. Identification of the genetic regions controlling FA composition is crucial in this respect. Nine loci that are strongly linked to five seed quality characteristics have been found in flax. For the iodine index, LG8 linkage group (LG), LG5 linkage group (LIN), total oil content, and LG9 linkage group (LG), the loci with the largest impacts have been described. The markers on LG2 (Lu2046) and LG6 (Lu2555), however, described about 8 and 4% of the phenotypic variance in PAL and OLE, respectively. The research was unable to find substantially linked loci for PAL and OLE. The reason that there is a negative association between LIN and other FAs may be explained by several markers associated with various FAs that were discovered to be colocalised in this research. It is interesting to note that Soto-Cerda et al. [181] found a locus linked to genes expected to be involved in FA biosynthesis.
For instance, it was discovered that the acyl-CoA: diacylglycerol acyltransferase A (dgatA) gene is located near the strongly linked locus on LG3. Earlier, a biparental quantitative trait locus (QTL) mapping strategy was used to identify flax dgatA. As a result, the loci found in flax using the GWAS method appear promising.

6.5. Molecular Mapping of QTLs

Genetic mapping has been used widely in both animal and plant species since the finding of DNA-based identifiers. After Botstein et al. [185] released the initial report of restriction fragment length polymorphism (RFLP), genetic mapping efforts increased, but they were hampered by the lack of an effective study. Early in the 1990s, advances in mapping analysis led to the first genetic map spanning the complete set of chromosomes, becoming available and being reported in humans and some plant species. The first RFLP-marker-based linkage map was created in 1988 and used for QTL mapping in plants [186]. The first inheritance studies and genetic mapping of flax were conducted in 1930 by Henry, who looked at how rust protection was passed down in the Bombay and Ottawa 770B varieties. These mapping initiatives, however, were based on physical traits. The first genetic linkage map of flax was created using RFLP and RAPD markers following the development of DNA-based molecular marker methods [187]. The 1000 cM genetic linkage map created was made up of 15 unique linkage groups with ninety-four evenly spaced markers. The F2 mapping population created from the hybrid between CI1303 and Stormont Cirrus was used in the research. Multiple investigations have since discovered QTLs for various flax characteristics. Cloutier et al. [175] constructed a genetic linkage map utilising114 SSR markers and 5 SNPs in a DH population of 78 individuals obtained from a crossing of SP2047 (yellow-seeded withlittle linolenic acid) and UGG5-5 (brown-seeded withelevated linolenic acid). For LIO, LIN, and the iodine value (IOD), two significant QTLs were found for each, and one major QTL was found for palmitic acid. The QTL QPal.crc-LG9 explained 42% of the trait variance in palmitic acid.
In related research, Kumar et al. [188] used a recombinant inbred line (RIL) population of 243 individuals from a hybrid between cultivars CDC Bethune and Macbeth to create a genetic map using 329 SNPs and 362 SSRs. For a total of 14 distinct characteristics, 20 QTLs were found in their study. There were three QTLs for STE and OLE, two QTLs for iodine content and LIO, and two QTLs for seed protein, oil content, thousand-seed weight, cell wall, straw weight, output, seeds per boll, and days to maturity. Candidate genes for characteristics related to yield component traits, cell wall synthesis, fibre production, and fatty acid biosynthesis were discovered through the analysis of the QTL.

7. Integrated Omics Approach

Among the major components of omics are transcriptomics (gene regulation and protein identification and effects, expression profiling), metabolomics (pathway and intermediates, metabolite profiling, regulation), phenomics (automated study and analysis of phenotypic and physiological effects), transcriptomics (gene regulation and expression profiling), proteomics (protein identification and effects), and ionomics. According to in-depth studies, various omics mechanisms and their combinations (Figure 3) are crucial for comprehending plant systems biology [189,190,191].
Omics-aided technologies have been utilised to conduct stress tolerance research on rice [192], soybean [193], and flax [194]. In contrast to other crops, however, comparatively fewer attempts have been made to use the available genetic and genomic tools for flax development. In accordance with Akhmetshina et al. [195] and Shivaraj et al. [59], who investigated the application of omics-assisted breeding and high-throughput sequencing technologies for the advancement of climate-smart flax, advanced tools such as genome-wide association studies (GWASs) and genomic selection in conjunction with various omic technologies offer an opportunity to increase the precision of flax improvement and plant selection.
In order to introduce weather-resilient characteristics into flax cultivars for long-term productivity, an extensive strategy incorporating a variety of technologies can significantly simplify the process [196].

7.1. Genomics

The genetic diversity of flax was evaluated in the early years of the twenty-first century using a variety of molecular markers [187,197,198,199,200]. The release of the flax whole-genome transcriptome provided a significant advantage in the creation of genomic tools [201]. The information from reduced representation sequencing and whole-genome re-sequencing has since been successfully applied to comprehend crop diversification, construct linkage maps, recognise markers, and discover QTLs in flax. Using the genotyping-by-sequencing (GBS) method, 258,873 SNPs spread across fifteen linseed chromosomes have been reported [188].
Recent studies have used genome-wide association studies (GWASs) to find potential SNPs for various traits, including improved stress tolerance indices, yield, and other traits in different species, including sorghum [202,203,204], rice [205], and sesame [206,207]. Although the AQP gene family improves drought resilience, multidrug and toxic compound extrusion (MATE) mediates the reaction to abiotic stresses [208]. However, research success in flax has beenless rapid than in cereals [209,210]. He et al. [211] used GBS to find 258,873 SNPs from the Canadian flax accession (core collection) as part of a GWAS to discover genetic areas linked to Pasmo resistance. Of the 500 potential QTLs found, 45 covered 85 resistance genes. Additionally, two potential genes (Lus10031043 and Lus10020016) for flax protection against this pathogen were found to be similar based on orthology with Arabidiopsis thaliana genes. In recent years, You et al. [212] examined 447 flax accessions for resistance to powdery mildew over 5–8 years from three sites, encompassing 372 samples from the primary collection and 75 mating lines.
By estimating the genetic potential rather than looking for specific QTLs, genomic selection (GS) is a breeding technique that overcomes the restrictions of marker-aided selection (MAS) for pace breeding (Figure 4). GS has the capacity to resolve all variations in the genetics of complex traits, as opposed to the traditional agricultural breeding technique, which is slow in concentrating on the complicated and barely heritable quantitative variables. Because of this, it is becoming clear that using molecular genetic markers to create brand-new models based on markers for genetic assessment is a promising method [196]. The precise phenotyping of a selected diverse collection of genotypes (for the purpose of breeding) in various conditions is required in order to develop a statistical model (GS model) that can be used to calculate genomic-predicted breeding values in the breeding population. As it addresses the minor effects of QTLs, the GS technique has many benefits over traditional and marker-assisted breeding [213].
Because it combines all known molecular markers with phenotypic information, GS has been promoted as the most effective technique for determining genetic values for selection [214,215]. An enhanced genetic gain/unit period was observed during the breeding cycle in flax GS experiments [216,217]. Table 4 and Table 5 represent various QTLs/QTNs linked to different stresses in flax. For QTL mapping with GS model optimisation, they employed three biparental populations created by mating high-yielding with high- and low-ALA acid content lines. He et al. (2019) [211] developed one of the most accurate genetic prediction models for resistance to plant diseases for the genetic resilience of flax to Septoria linicola. You et al. [212] used 447 flax inclusions as an experimental population to create the most current model for forecasting the evaluations for powdery mildew across five years at three separate sites. The 349 QTNs discovered through GWAS described ninety-six percent of the variance in powdery mildew, demonstrating the model’s strong predictive power and promise for use in genomic prediction.
More comprehensive GS studies are anticipated soon due to the enhanced genetic techniques with genomic resources for linseed, which could help in the release of new cultivars suited to requirements. Due to greater costs than MAS, the more widespread use of GS is currently still a problem. Budhlakoti et al. [216] reviewed the present state of genomic selection research in crop plants and the prospects for its effective application in the making of weather-resilient crops. They stressed that research on genetic makeup under heat with drought stress can invariably speed up the production of stress-resilient cultivars through genomic selection [196].

7.2. Transcriptomics

The discovery of important genes engaged in the stress tolerance process is aided by transcriptome profiling, which offers a thorough overview of gene expression and regulation. Depending on the abundance of generated genomic materials and plant variety, different methods are used to investigate the transcriptome, including spotted microarrays, expressed sequence tags, Affymetrix Gene Chips, sequencing combined with suppression subtractive hybridisation, and RNA sequencing. The advancement of next-generation sequencing technologies and RNA sequencing has made them the most efficient, affordable, and high-throughput transcriptomic techniques [196].
Numerous transcriptomic studies have been conducted to date to determine the impact of salt and drought on oilseed products like sesame [207], flax [220], jatropha [221], sunflower [222], and soybean [223]. The sequencing of the linseed genome and the presence of genetic maps [182,195,201] set the groundwork for numerous transcriptomic studies and the discovery of genes underpinning characteristics of economic and agronomic significance. High-throughput sequencing was used to examine how flax responded to the alterations in the levels of salt and pH [224,225,226], drought [227], nutrient deficiency [228,229], and metal stress [230].
Dash et al. [227] performed a transcriptome analysis on the relatively drought-tolerant flax cultivar of India, i.e., T-397, and expression profiling assisted in identifying markers for the identification of drought-resistant flax varieties. Shivaraj et al. [231] showed the increased expression of integral membrane proteins, primarily aquaporins, using transcriptome analysis data, which improved the knowledge of their physiological function. In a transgenic strain of flax cv. Blanka, the overexpression of the DREB2A gene confers drought resistance [73]. Like this, only a small number of flax genes have been identified and functionally characterised under extreme temperature stress. According to Saha et al. [111], 34 potential HSF genes from the flax genome were identified across the entire genome.
Under conditions of heat stress, different expression patterns are bestowed by HSF, along with NAC domain TFs. By lengthening roots, reducing membrane injury, and improving ion transport, two salt-tolerant genes that are related to Arabidopsis Senescence-Associated Gene 29 (SAG29) may improve salt tolerance. The flax transcriptome reaction to acidic soil revealed genes with altered expression patterns [232]. Through transcriptomics, the reaction of flax to an unfavourable zinc shortage and soil acidity identified genes incorporated in cell wall biogenesis photosynthesis and ion transfer [224]. Several pathogen-related dominant genes were induced to counteract the negative impacts of low Zn concentration. Melnikova et al. [228] found 96 conservative microRNA homologs from 21 families, and they also described the function of 7 microRNAs (miR408, miR395, miR168, miR398, miR169, miR399, and lus-miR-N1) in the control of metabolism and gene expression in plants during nutritional stress.
As a defence strategy, plants’ transcriptional and translational profiles are also altered by pathogen assault, which causes the implementation of various metabolic pathways and genes. In their analysis of the gene reaction during the early phases of Fusarium infection, Kostyn et al. [233] found 47 genes responsible for antioxidant biosynthesis and phenylpropanoid pathways in flax. Hundreds of genes with differential expression in reaction to early pathogenesis were found in the prevalent Canadian cv. CDC Bethune transcriptome of flax immune to Fusarium wilt and a susceptible variety, Lutea. Several important genes, including those involved in secondary metabolism, the stimulation of pathogenesis-related (PR) interactions, and the formation of lignin, have higher transcript abundance among these [196]. The transcriptome of two susceptible and two resistant BC2F5 populations, as well as four cultivars of fibrous flax, revealed the enhanced expression of many genes involved in the defence response, such as PR protein-encoding genes, ROS production genes, and genes related to cell wall formation [234].
Boba et al. [235] recently found that upon Fusarium infection in flax, upregulation of the terpenoid pathway, resulting in higher ABA content, initiates the fast plant response, and PR genes, particularly chitinase and 1,3-glucanase, play an important role in resistance. In previous research, Wróbel-Kwiatkowska et al. [236] found that transgenic flax plants overexpressing the 1,3-glucanase gene displayed reduced vulnerability to this pathogen. In reaction to infection by Fusarium oxysporumlini, it was discovered that the resistant flax cultivar’s transcriptomic response was faster and more effective, enabling the translation ofa greater number of activated and repressed genes [237].
The NCBI Sequence Read Archive and NCBI Gene Expression Omnibus libraries contain the transcriptome sequences, and gene expression data for flax. L. usitatissimum microRNA sequences are provided in the “miRbase” database, along with their main and secondary structures and locations within the flax genome. In the field of flax transcriptomics, there are more studies published (Table 6) on tolerance to abiotic stressors than on resistance to biotic stresses, which is significant and might be a result of targeted characteristics under breeding programmes for areas. Most of the transcriptomic research only looked at one or two genotypes; however, for comparative analysis and gene annotation, a larger number of varied genotypes should be examined [196].

7.3. Metabolomics

Interpreting the mechanisms of plant biology requires integrating metabolomics because metabolites, which are the consequences of gene expression, provide a detailed picture of an organism’s biological and physiological condition [243,244,245]. However, as discussed by Hall et al. [246], metabolomic research is not without its difficulties because the extent of the plant metabolome is unclear. Plants have a widervariety of metabolites than other species. A wide range of statistical channels, involving mass spectrometry (MS)- and nuclear magnetic resonance (NMR)-based techniques like CE-MS (capillary electrophoresis–mass spectrometry), GC-MS (Gas Chromatography–Mass Spectrometry), FTIR, and LC-MS (Liquid Chromatography–Mass Spectrometry), have been used to identify and quantify awide variety of primary and secondary metabolites produced in response to abiotic and biotic stresses. NMR is usually used to identify moderate-to-high-abundance metabolites, and it requires little sample preparation.
Additionally, new improvements in NMR superconducting magnet field intensity have led to better spectral precision and detection sensitivity. In contrast to NMR, modern MS-based methods produce analyses of complicated plant metabolite mixtures with greater sensitivity. A review paper by Ibáez et al. [247] provides a summary of the most contemporary literature on direct ionisation methods used in the metabolomics of food. It is well known that metabolites like lignans, particularly omega-3 fatty acids and polyunsaturated fatty acids (PUFA), are likely to induce a positive impact on nourishment and the avoidance of some diseases [248,249]. In order to choose flaxseed types with a superior nutrient composition, a tool based on NMR metabolomics was developed [250]. Lignans and omega-3 fatty acids, two bioactive substances of importance for human health, are significant components of flaxseed. Oilseeds have been found to accumulate a wide range of secondary metabolites, such as sugars, inorganic ions, glycine betaine, and proline, in addition to the metabolomics of nutritive compounds, to promote plant adaptation to abiotic stress [251].
BABA, for example, has been shown to play a part in promoting drought resistance in a variety of plants, including Arabidopsis [252], apple [253], rice [254], spring wheat [255], potato [256], and tomato [257]. More osmoprotectants, anthocyanins, and proline were accumulated when PR1, PR2, and PR5 genes were overexpressed in Arabidopsis, while trehalose production made tobacco more drought-tolerant [258,259]. In reaction to water stress, BABA in flax increases proline and nonstructural carbs while decreasing aspartate concentration and inorganic solutes [260]. Flax [261], along with rice, was shown to have relatively higher concentrations of proline and glycinebetaine under salinity stress [262].
Total soluble sugars, compatible solutes, and total protein contents, like proline and betaine, were found to increase with increasing salinity in flax genotypes under PEG-induced water stress and saline–alkaline environments, suggesting that they may play a role in adjusting to osmotic stress [263]. Under salt and/or osmotic stress, different levels of both wild-type and PLR-RNAi transgenic flax, lipid peroxidation, and a distinct metabolic profile of MDA have been noted [261,264]. As a defence strategy, pathogen assault also causes changes in several secondary metabolites, including lignin, flavonoids, polyamines, catecholamines, terpenoids, tannins, phenyl propanoic acids, and phenolics, which are produced as a result of the translational profile of plants. Many rice studies related to metabolomics have been conducted to identify important metabolites and routes in reaction to different types of environmental distress [265]. The purpose of these studies was to comprehend how the model crop rice induced defence mechanisms involving effector-prompted immunity (ETI) and PAMP-triggered immunity (PTI) in response to pathogens. As a result, stress causes significant alterations in a plant’s metabolic makeup (Figure 5), and full metabolite profiling may offer important insights into processes for stress tolerance [196].

7.4. Proteomics

Proteomics is the real-time analysis of the molecular and functional properties of every protein in a living thing. It combines different bioinformatic techniques with two-dimensional (2-D) gel electrophoresis, Western blotting, mass spectrometry (MS), matrix-assisted laser desorption ionisation–time of flight (MALDI TOF), and ELISA [266,267]. Recent advances in proteomics have decreased protein evaluation mistakes and opened new avenues for high-throughput proteome studies. Most of the proteomic research has been undertaken on the entire genomes of soybean, wheat, barley, rice, and potatoes. Reports on proteomic research on oilseeds include those on flax [268,269], sunflower [270], and Indian mustard [271].
Proteome analyses have shown that tolerant plants have consistently increased levels of stress-reactive chains of amino acids compared to their sensitive counterparts, which include transcriptional regulators like Myb protein, B-Peru-like protein, and SWIB/MDM2 protein involved in anthocyanin biosynthesis [272,273]. In comparison to their sensitive counterparts, drought-tolerant wheat and barley types were found to have higher levels of proteins, such as several chaperons (HSP70, HSP90, and cyclophilin A), glutathione-S-transferase (GST), and lipoxygenase (LOX) [274]. Another study found that drought [275], low temperature, and salt [276] caused reductions in RubisCo, enzymes like phosphoglycerokinase (PGK), and transketolase phosphoribulokinase (PRK) in wheat [277]. An increase in numerous minor heat shock proteins, as well as HSP82 from the HSP90 family, was found inside the endosperm of growing wheat seeds exposed to a heat phase [278].
Proteomic investigation and physiological responses in two drought-resistant maize cultivars demonstrated that HSP plays a substantial role in protecting plants from drought stress in a manner [279]. Recently, Halder et al. [280] summarised the proteomic research on salinity, drought stress tolerance, and root system architecture carried out in the previous ten years. They also examined the function of proteomics for abiotic stress tolerance in wheat. It is useful to characterise the proteome of tissues with pathogen infections using proteomic studies of biotic stress. Numerous potent metabolites that are responsible for resilience have been identified in a thorough analysis of the worldwide proteomics studies looking into the reactions of rice to biotic stress [265]. In flax, this area has recently begun to be studied [269]. Some of the examples are given in the Table 7.

7.5. Ionomics

Ionomics is the study of the elemental composition of non-metals, metals, and metalloids in various plant species, with a focus on high-throughput detection and quantification [196]. Different plant varieties’ ionomic compositions may be determined using high-throughput methods like mass spectrometry with inductive and inductively coupled plasma–atomic emission spectroscopy. It plays a crucial part in helping us comprehend the various elements, their chemical and physical makeup, and how they affect plant metabolism, nourishment, and biochemistry. Due to the diversity of soil types and various edaphic variables affecting growth, plants have developed a range of element uptake skills [286]. Additionally, transit, environmental stress, root uptake capacity, and element abundance all have an impact on the ionomic makeup of a crop.
The quantity of accumulated Na+ and metabolites implicated in glycolysis and the tricarboxylic acid (TCA) cycle were significantly negatively correlated, according to the ionome of wild and cultivated barley that was exposed to various degrees of salt tolerance [287]. Studies on wheat [288], as well as other grasses like Aneurolepidium chinense [289], Setaria viridis [290], and flax, revealed that the concentration of Na+ rises with increasing alkalinity stress because plants store significant amounts of Na+ in their vacuoles to lower cell water potential. K+, Na+, Ca2+, and Cl were the major inorganic ions engaged in osmotic adjustment under PEG-induced water stress in flax, improving drought resilience [291]. In the caseof flax, there were no appreciable variations between the impacts of salt and alkalinity stress on the concentrations of Na+ and K+ in shoots [292]. This indicates that, unlike other plants like wheat and Chloris virgata, where the K+ concentration of shoots was found to be reduced under alkaline stress, flax shoots may have a different adaptive mechanism to alkaline stress. Another study on flax revealed that under salt stress, the amounts of NO3 in flax shoots dropped, while the uptake of Cl and H2PO4 increased significantly. Therefore, knowing the elemental makeup can help with understanding the process of stress endurance. Ionomics research in flax has not received much notice [196].

7.6. Phenomics

High-throughput analysis is used to research phenomics, which is the group of phenotypes combining GxPxE relationships under environmental circumstances [190]. As a result, the phenotype offers the strongest link between surroundings and plant genetics. While phenotypic characterisation has advanced more slowly over the past ten years due to improvements in sequencing technologies, this has limited the discovery of quantitative characteristics, especially those linked to stress resistance [293].
Phenotyping in reaction to abiotic stress remains a significant task because of the complex biosynthetic processes that handle the response of plants to external stressors [294]. Because genomic approaches like GWAS, GS, and QTL rely on high-throughput phenotyping for the improvement of specific traits, the significance of accurate phenotyping has increased in the postgenomic age [295]. The most promise exists for plant breeding when phenomics is used in conjunction with other omic methods. As a result, non-invasive technologies such as magnetic resonance imaging (MRI), lidar (which includes RBG digital imaging) to evaluate growth parameters, and colour imaging of biomass have all been researched to determine the elevated canopy and concealed half (root system) of plants [296,297].
Examples include employing ground-filled rhizoboxes to analyse the architecture of wheat roots using RGB imaging [298] and using RGB digital imaging to phenotype plant shoots [299]. Infrared thermography to confirm the function of stomatal electrical conductance in barley and wheat saplings under salt stress is another illustration [299]. Nuclear magnetic resonance (NMR), computed tomography (CT), and X-rays have all been employed forin situ 3D imaging of root anatomy. There are not many web databases that can help people with image processing, like http://www.plant-image-analysis.org, accessed on 5 May 2023. In order to obtain real-time phenome responses to diseases, the exterior environment, and nutrients, contemporary instruments include hyperspectral imaging systems, fluorescence imaging systems, high-throughput advanced plant phenotyping platforms, laser scanners, and high-resolution IR/NIR cameras. However, to extract phenome data from the enormous databases produced by phenotyping, deep learning techniques are required. Additionally, the thorough administration of platforms and software presents significant difficulties, restricting this application to a few key commodities, like rice, maize, and wheat [196].

8. Transgenic Flax/Linseed

The initial genetic modification made to flax washerbicide tolerance. The herbicide glyphosate (Roundup) blocks the enzyme 5-enolpyruvylshikimate-3-phosphate (EPSP) polymerase in higher plants. To create transgenic flax lines resistant to 5 mM glyphosate, the glyphosate-insensitive EPSP enzyme from Petunia hybrida was transformed into flax by an Agrobacterium-mediated method of transformation [300]. Later, flax was given additional herbicide-resistant genes, including those for glufosinate ammonium [301] and chlorsulfuron [302]. Under the direction of the CaMV 35S promoter, the phosphinothricin acetyltransferase (PAT) gene from Streptomyces viridochromo was overexpressed. The PAT enzyme was introduced into flax plants to provide them with resistance to the herbicide glufosinate ammonium [301].
To tolerate abiotic stress conditions, several crop plants have undergone genetic modification. Numerous candidate genes, including functional and regulatory genes, have been modified [302]. Functional genes, such as those that encode water channel proteins, ion transporters, late embryogenesis abundant (LEA) protein, and heat shock protein (HSP), are directly engaged in the defence of cells against stressors. The regulatory genes control signal transmission and gene expression during stress reactions, and they comprise transcription factors, protein phosphatases, and protein kinases [303,304]. These studies aiming to genetically modify flax for resistance/tolerance to abiotic stress have not yet been published. When plants are subjected to abiotic stress, the activity of the transporters is first impacted, which hinders the plants’ physiological function. The transport of water and other minor solutes, including urea, ammonia, glycerol, silicic acid, boric acid, CO2, and H2O2, is significantly aided by aquaporins in plants [305]. Abiotic stress resistance in various plant species has recently been effectively demonstrated by the application of genes expressing aquaporins. Aquaporins have been discovered in a variety of agricultural plants, including flax, due to their significance [306,307]. A total of 51 AQPs from various subfamilies were found in the flax genome [210].

9. Conclusions

The production of oilseed crops is substantially affected byan array of abiotic and biotic stressors. To better understand the chemical complexes and biochemical mechanisms involved in plants’ ability to tolerate stress, large amounts of data are quickly being created and annotated on a global scale. The availability of various genomic resources in flax has dramatically grown over the past ten years. These resources include molecular markers, linkage maps, transcriptomes, and whole-genome sequencing. Through a varietal development programme, these resources may be effectively used to increase the capacity of flax for adaptation to the environment and resistance to biotic stress. Given its great economic worth, flax is used in a variety of ways in the food, bioenergy, nutrition, and nutraceutical industries. This review has covered many aspects of tolerance that are now being utilised to overcome these barriers. These strategies really work to increase a plant’s natural ability to withstand diverse biotic and abiotic stresses, possibly having long-lasting impacts and, in theory, not negatively affecting other intriguing agricultural attributes like yield and the quality of edible parts. Additionally, technological advancements like the application of high-throughput platforms to measure the tolerance traits of crop genotypes upon treatments give researchers the chance to characterise the plant responses of a subset of varieties under stringent growing circumstances or in the field, where real-world agricultural conditions occur.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kajla, P.; Sharma, A.; Sood, D.R. Flaxseed—A potential functional food source. J. Food Sci. Technol. 2015, 52, 1857–1871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Laux, M. 2011. Available online: https://www.agmrc.org/commodities-products/grains-oilseeds/flax-profile (accessed on 13 March 2012).
  3. Morris, D.H. Linseed in the Ruminant Diet—Adding Linseed to Feed Enhances the Fat Profile of Milk Winnipeg, MB, Flax Council of Canada. 2008. Available online: http://www.flaxcouncil.ca/files/web/Beef_R3_final.pdfm (accessed on 25 May 2012).
  4. Singh, K.; Mridula, D.; Rehal, J.; Barnwal, P. Flaxseed: A potential source of food, feed and fiber. Crit. Rev. Food Sci. Nutr. 2011, 51, 210–222. [Google Scholar] [CrossRef] [PubMed]
  5. Friis, I.B. Linaceae: Linum. In Flora of Ethiopia and Eritrea: Mangnoliaceae to Flacourtiaceae; Edwards, S., Tadesse, M., Demissew, S., Hedberg, I., Eds.; The National Herbarium, Addis Ababa and the Department of Systematic Botany: Uppsala, Sweden, 2000; Volume 2, pp. 352–357. [Google Scholar]
  6. Jhala, A.J.; Hall, L.M.; Hall, J.C. Potential hybridization of flax with weedy and wild relatives: An avenue for movement to engineered genes? Crop Sci. 2008, 48, 825–840. [Google Scholar] [CrossRef] [Green Version]
  7. Mansby, E.; Diaz, O.; von Bothmer, R. Preliminary study of genetic diversity in Swedish flax (Linum usitatissimum L.). Genet. Resour. Crop Evol. 2000, 47, 417–424. [Google Scholar] [CrossRef]
  8. Hoque, A.; Fiedler, J.D.; Rahman, M. Genetic diversity analysis of a flax (Linum usitatissimum L.) global collection. BMC Genom. 2020, 21, 557. [Google Scholar] [CrossRef]
  9. Kurt, O.; Evans, G.M. Genetic Basis of Variation in Linseed (Linum usitatissimum L.) Cultivars. Turk. J. Agric. For. 1998, 22, 373–379. [Google Scholar]
  10. Mohammed, A.S.; Atnaw, S.M.; Ramaya, A.V.; Alemayehu, G. A comprehensive review on the effect of ethers, antioxidants, and cetane improver additives on biodiesel-diesel blend in CI engine performance and emission characteristics. J. Energy Inst. 2023, 108, 101227. [Google Scholar] [CrossRef]
  11. Liu, N.; Wan, B.; Zhang, Z.; Fang, X.; Lin, X.; Wang, Y.; Tang, J.; Bai, X.; Li, Y.; Yao, Y.; et al. Self-healing waterborne polyurethane coatings with high transparence and haze via cellulose nanocrystal stabilized linseed oil Pickering emulsion. Int. J. Biol. Macromol. 2023, 235, 123830. [Google Scholar] [CrossRef]
  12. Goyal, A.; Sharma, V.; Upadhyay, N.; Gill, S.; Sihag, M. Flax and flaxseed oil: An ancient medicine & modern functional food. J. Food Sci. Technol. 2014, 51, 1633–1653. [Google Scholar]
  13. Parikh, M.; Maddaford, T.G.; Austria, J.A.; Aliani, M.; Netticadan, T.; Pierce, G.N. Dietary flaxseed as a strategy for improving human health. Nutrients 2019, 11, 1171. [Google Scholar] [CrossRef] [Green Version]
  14. Sethi, S.; Tyagi, S.K.; Anurag, R.K. Plant-based milk alternatives an emerging segment of functional beverages: A review. J. Food Sci. Technol. 2016, 53, 3408–3423. [Google Scholar] [CrossRef] [PubMed]
  15. Rodriguez-Leyva, D.; Weighell, W.; Edel, A.L.; LaVallee, R.; Dibrov, E.; Pinneker, R.; Maddaford, T.G.; Ramjiawan, B.; Aliani, M.; Guzman, R.; et al. Potent antihypertensive action of dietary flaxseed in hypertensive patients. Hypertension 2013, 62, 1081–1089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pohjanheimo, T.A.; Hakala, M.A.; Tahvonen, R.L.; Salminen, S.J.; Kallio, H.P. Flaxseed in bread making: Effects on sensory quality, aging, and composition of bakery products. J. Food Sci. 2006, 71, S343–S348. [Google Scholar] [CrossRef]
  17. Alam, S.; Sarker, M.; Rahman, M.; Afrin, S.; Richi, F.T.; Zhao, C.; Zhou, J.R.; Mohamed, I.N. Traditional Herbal Medicines, Bioactive Metabolites, and Plant Products Against COVID-19: Update on Clinical Trials and Mechanism of Actions. Front. Pharmacol. 2021, 12, 1248. [Google Scholar] [CrossRef]
  18. Alagirusamy, R.; Das, A. Yarns: Production, processability and properties. In Fibrous and Composite Materials for Civil Engineering Applications; Woodhead Publishing: Sawston, UK, 2011; Volume 1, pp. 29–61. [Google Scholar]
  19. Zaidi, Z.; Do Ain UzunAnaf, W.H. Katan (Linum usitatissimum Linn.) a potent unani therapeutic agent for respiratory tract diseases. J. Pharmacognocy 2021, 8, 112–118. [Google Scholar]
  20. Pramanik, A.; Tiwari, S.; Tripathi, M.K.; Tomar, R.S.; Singh, A.K. Molecular characterization of groundnut (Arachis hypogea L.) germplasm lines for yield attributed traits. Indian J. Genet. 2019, 79, 56–65. [Google Scholar] [CrossRef]
  21. Mandloi, S.; Tripathi, M.K.; Tiwari, S.; Tripathi, N. Genetic diversity analysis among late leaf spot and rust resistant and susceptible germplasm in groundnut (Arachis hypogea L.). Isr. J. Plant Sci. 2022, 69, 163–171. [Google Scholar] [CrossRef]
  22. Rajpoot, N.S.; Tripathi, M.K.; Tiwari, S.; Tomar, R.S.; Kandalkar, V.S. Characterization of Indian mustard germplasm on the basis of morphological traits and SSR markers. Curr. J. Appl. Sci. Technol. 2020, 39, 300–311. [Google Scholar] [CrossRef]
  23. Diederichsen, A.; Raney, J.P. Seed colour, seed weight and seed oil content in Linum usitaitissimum accessions held by plant gene resources of Canada. Plant Breed. 2006, 125, 372–373. [Google Scholar] [CrossRef]
  24. Saeidi, G. Genetic variation and heritability for germination, seed vigour and field emergence in brown and yellow- seeded genotypes of flax. Int. J. Plant Prod. 2012, 2, 15–22. [Google Scholar]
  25. Sharma, A.; Tripathi, M.K.; Tiwari, S.; Gupta, N.; Tripathi, N.; Mishra, N. Evaluation of soybean (Glycine max L.) genotypes on the basis of biochemical contents and anti-oxidant enzyme activities. Legume Res. 2021, 44, 1419–1429. [Google Scholar] [CrossRef]
  26. Baghel, R.; Sharma, A.K.; Tiwari, S.; Tripathi, M.K.; Tripathi, N. Genetic Diversity Analysis of Indian Mustard (Brassica spp.) Germplasm Lines using SSR Molecular Markers. Int. J. Curr. Microbiol. Appl. Sci. 2020, 9, 137–143. [Google Scholar] [CrossRef]
  27. Patzk, J. Comparison of RAPD, STS, ISSR and AFLP molecular methods used for assessment of genetic diversity in hop (Humulus lupulus L.). Euphytica 2001, 121, 9–18. [Google Scholar] [CrossRef]
  28. Krulíckova, K.; Posvec, Z.; Griga, M. Identification of flax and linseed cultivars by isozyme markers. Biol. Plant. 2002, 45, 327–336. [Google Scholar] [CrossRef]
  29. Fu, Y.B. Geographic patterns of RAPD variation in cultivated flax. Crop Sci. 2005, 45, 1084–1091. [Google Scholar] [CrossRef] [Green Version]
  30. Fu, Y.B. Redundancy and distinctness in flax germplasm as revealed by RAPD dissimilarity. Plant Genet. Resour. 2006, 4, 117–124. [Google Scholar] [CrossRef]
  31. Singh, P.K.; Akram, M.; Srivastava, R.L. Genetic diversity in linseed (Linum usitatissimum) cultivars based on RAPD-analysis. Indian J. Agric. Sci. 2009, 79, 1046–1049. [Google Scholar]
  32. Bibi, T.; Mustafa, H.S.B.; Hasan, E.U.; Rauf, S.; Mahmood, T.; Ali, Q. Analysis of genetic diversity in linseed using molecular markers. Life Sci. J. 2015, 12, 28–37. [Google Scholar]
  33. Nagabhushanam, B.; Mir, M.I.; Nagaraju, M.; Sujatha, E.; Devi, B.R.; Kumar, B.K. Genetic Diversity Analysis of Linseed (Linum usitatissimum L.) Accessions Using RAPD Markers. Emirates J. Food Agric. 2021, 33, 589. [Google Scholar] [CrossRef]
  34. Everaert, I.; De Riek, J.; De Loose, M.; Van Waes, J.; Van Bockstaele, E. Most similar variety grouping for distinctness evaluation of flax and linseed (Linum usitatissimum L.) varieties by means of AFLP and morphological data. Plant Var. Seeds 2001, 14, 69–87. [Google Scholar]
  35. Van Treuren, R.; van Soest, L.J.M.; van Hintum, T.J.L. Markerassisted rationalization of genetic resource collections: A case study in flax using AFLPs. Theor. Appl. Genet. 2001, 103, 144–152. [Google Scholar] [CrossRef]
  36. Wakjira, A.; Viljoen, C.D.; Labuschagne, M.T. Analysis of genetic diversity in linseed using AFLP markers. Ethiop. J. Sci. 2005, 28, 41–50. [Google Scholar] [CrossRef]
  37. Chandrawati Maurya, R.; Singh, P.K.; Ranade, S.A.; Yadav, H.K. Diversity analysis in Indian genotypes of linseed (Linum usitatissimum L.) using AFLP markers. Gene 2014, 549, 171–178. [Google Scholar] [CrossRef] [PubMed]
  38. Wiesnerová, D.; Wiesner, I. ISSR-based clustering of cultivated flax germplasm is statistically correlated to thousand seed mass. Mol. Biotechnol. 2004, 3, 207–214. [Google Scholar] [CrossRef] [PubMed]
  39. Rajwade, A.V.; Arora, R.S.; Kadoo, N.Y.; Harsulkar, A.M.; Ghorpade, P.B.; Gupta, V.S. Relatedness of Indian flax genotypes (Linum usitatissimum L.): An inter-simple sequence repeat (ISSR) primer assay. Mol. Biotechnol. 2010, 45, 161–170. [Google Scholar] [CrossRef]
  40. El Sayed, A.A.; Ezzat, S.M.; Mostafa, S.H.; Zedan, S.Z.; Abdel-Sattar, E.; El Tanbouly, N. Inter simple sequence repeat analysis of genetic diversity and relationship in four Egyptian flaxseed genotypes. Phcog. Res. 2018, 10, 166–172. [Google Scholar]
  41. Choudhary, S.B.; Sharma, H.K.; Kumar, A.A.; Maruthi, R.T.; Mitra, J.; Chowdhury, I.; Singh, B.K.; Karmakar, P.G. SSR and morphological trait based population structure analysis of 130 diverse flax (Linum usitatissimum L.) accessions. Comptes Rendus Biol. 2017, 340, 65–75. [Google Scholar] [CrossRef]
  42. Wu, J.; Zhao, Q.; Wu, G.; Zhang, S.; Jiang, T. Development of Novel SSR Markers for Flax (Linum usitatissimum L.) Using Reduced-Representation Genome Sequencing. Front. Plant Sci. 2017, 7, 2018. [Google Scholar] [CrossRef] [Green Version]
  43. Green, A.G.; Chen, Y.; Singh, S.P.; Dribnenki, J.C.P. Flax. In Compendium of Transgenic Crop Plants; Kole, C., Hall, T.C., Eds.; Blackwell Publishing Ltd.: Oxford, UK, 2008; pp. 199–226. [Google Scholar]
  44. Cloutier, S.; Ragupathy, R.; Niu, Z.; Duguid, S. SSR-based link-age map of flax (Linum usitatissimum L.) and mapping of QTLs underlying fatty acid composition traits. Mol. Breed. 2011, 28, 437–451. [Google Scholar] [CrossRef]
  45. Cloutier, S.; Miranda, E.; Ward, K.; Radovanovic, N.; Reimer, E.; Walichnowski, A.; Datla, R.; Rowland, G.; Duguid, S.; Ragupathy, R. Simple sequence repeat marker development from bacterial artificial chromosome end sequences and expressed sequence tags of flax (Linum usitatissimum L.). Theor. Appl. Genet. 2012, 125, 685–694. [Google Scholar] [CrossRef] [Green Version]
  46. Fu, Y.B.; Diederichsen, A.; Richards, K.W.; Peterson, G. Genetic diversity within a range of cultivars and landraces of flax (Linum usitatissimum L.) as revealed by RAPDs. Genet. Resour. Crop Evol. 2002, 49, 167–174. [Google Scholar] [CrossRef]
  47. Fu, Y.-B.; Rowland, G.G.; Duguid, S.D.; Richards, K.W. RAPD analysis of 54 north American flax cultivars. Crop Sci. 2003, 43, 1510–1515. [Google Scholar] [CrossRef]
  48. Cloutier, S.; Niu, Z.; Datla, R.; Duguid, S. Development and analysis of EST-SSRs for flax (Linum usitatissimum L.). Theor. Appl. Genet. 2009, 119, 53–63. [Google Scholar] [CrossRef] [PubMed]
  49. Asati, R.; Tripathi, M.K.; Tiwari, S.; Yadav, R.K.; Tripathi, N. Molecular approaches in the development of drought tolerance in chickpea. Life 2022, 12, 1846. [Google Scholar] [CrossRef]
  50. Yadav, P.K.; Singh, A.K.; Tripathi, M.K.; Tiwari, S.; Yadav, S.K.; Tripathi, N. Morpho-Physiological and Molecular Characterization of Maize (Zea mays L.) Genotypes for Drought Tolerance. Eur. J. Appl. Sci. 2022, 10, 65–87. [Google Scholar]
  51. Shyam, C.; Tripathi, M.K.; Tiwari, S.; Tripathi Niraj Ahuja, A. Identification of Low and High Erucic Acid Containing Genotype(S) in Indian Mustard Employing Molecular Markers. In Recent Progress in Plant and Soil Research; BP International: West Bengal, India, 2022; Volume 5, pp. 18–36. [Google Scholar] [CrossRef]
  52. Tripathi, N. Biodiversity and conservation of genetic resources: Key factor in crop improvement. In Biodiversity and Crop Improvement; Wani, S.H., Ed.; Scientific Research Publishing: Wuhan, China, 2015; pp. 104–129. ISBN 978-1-61896-068-9. [Google Scholar]
  53. Diederichsen, A. Ex-situ collections of cultivated flax (Linum usitatissimum L.) and other species of the genus Linum L. Genet. Resour. Crop Evol. 2007, 54, 661–678. [Google Scholar] [CrossRef]
  54. Diederichsen, A.; Fu, Y.B. Flax genetic diversity as the raw material for future success. In Proceedings of the International Conference on Flax and Other Bast Plants, Saskatoon, SK, Canada, 21–23 July 2008; pp. 270–280. [Google Scholar]
  55. Heslop-Harrison, J.S.; Schwarzacher, T. Genetics and genomics of crop domestication. In Plant Biotechnology and Agriculture: Prospects for the 21st Century; Altman, A., Hasegawa, P.M., Eds.; Elsevier Academic: Alpharetta, GA, USA, 2012; pp. 3–18. [Google Scholar]
  56. Brutch, N.B. The flax genetic resources collection held at the N.I. Vavilov Institute, Russian Federation. In Flax Genetic Resources in Europe; Maggioni, L.M., Pavelek, M., van Soest, L.J.M., Lipman, E., Eds.; IPGRI: Maccarese, Italy; Rome, Italy, 2002; pp. 61–65. [Google Scholar]
  57. Zhuchenko, A.A.; Rozhmina, T.A. Mobilizacija genetičeskich resursov l“na [Mobilization of flax genetic resources]. In VILAR and VNIIL; Starica: Staritsa, Russia, 2000; p. 224. [Google Scholar]
  58. Rashid, K.Y. Principal diseases of flax. In Flax the Genus Linum; Muir, A.D., Westcott, N.D., Eds.; Taylor and Francis Ltd.: London, UK, 2003; pp. 92–123. [Google Scholar]
  59. Shivaraj, S.M.; Dhakate, P.; Sonah, H.; Vuong, T.; Nguyen, H.T.; Deshmukh, R. Progress toward development of climate-smart flax: A perspective on omics-assisted breeding. In Genomic Designing Climate-Smart Oilseed Crops; Springer: Cham, Switzerland, 2019; pp. 239–274. [Google Scholar]
  60. Mishra, N.; Tripathi, M.K.; Tripathi, N.; Tiwari, S.; Gupta, N.; Sharma, A. Screening of soybean genotypes against drought on the basis of gene-linked microsatellite markers. In Book Innovations in Science and Technology; BP International: West Bengal, India, 2022; Volume 3, pp. 49–61. [Google Scholar]
  61. Tripathi, M.K.; Tripathi, N.; Tiwari, S.; Mishra, N.; Sharma, A.; Tiwari, S.; Singh, S. Identification of Indian soybean (Glycine max [L.] Merr.) genotypes for drought tolerance and genetic diversity analysis using SSR markers. Scientists 2023, 3, 31–46. [Google Scholar]
  62. Fahad, S.; Bajwa, A.A.; Nazir, U.; Anjum, S.A.; Farooq, A.; Zohaib, A.; Sadia, S.; Nasim, W.; Adkins, S.; Saud, S.; et al. Crop Production under Drought and Heat Stress: Plant Responses and Management Options. Front. Plant Sci. 2017, 8, 1147. [Google Scholar] [CrossRef] [Green Version]
  63. Nasir, M.W.; Toth, Z. Effect of Drought Stress on Potato Production: A Review. Agronomy 2022, 12, 635. [Google Scholar] [CrossRef]
  64. Molotoks, A.; Stehfest, E.; Doelman, J.; Albanito, F.; Fitton, N.; Dawson, T.P.; Smith, P. Global projections of future cropland expansion to 2050 and direct impacts on biodiversity and carbon storage. Glob. Chang. Biol. 2018, 24, 5895–5908. [Google Scholar] [CrossRef] [Green Version]
  65. Cui, Z.; Yan, B.; Gao, Y.; Wu, B.; Wang, Y.; Wang, H.; Xu, P.; Zhao, B.; Cao, Z.; Zhang, Y.; et al. Agronomic cultivation measures on productivity of oilseed flax: A review. Oil Crop Sci. 2022, 7, 53–62. [Google Scholar] [CrossRef]
  66. Zare, S.; Mirlohi, A.; Saeidi, G.; Sabzalian, M.R.; Ataii, E. Water stress intensified the relation of seed color with lignan content and seed yield components in flax (Linum usitatissimum L.). Sci. Rep. 2021, 11, 23958. [Google Scholar] [CrossRef] [PubMed]
  67. Fila, G.; Bagatta, M.; Maestrini, C.; Potenza, E.; Matteo, R. Linseed as a dual-purpose crop: Evaluation of cultivar suitability and analysis of yield determinants. J. Agric. Sci. 2018, 156, 162–176. [Google Scholar] [CrossRef]
  68. Hu, H.; Xiong, L. Genetic engineering and breeding of drought-resistant crops. Annu. Rev. Plant Biol. 2014, 65, 715–741. [Google Scholar] [CrossRef] [PubMed]
  69. Stavropoulos, P.; Mavroeidis, A.; Papadopoulos, G.; Roussis, I.; Bilalis, D.; Kakabouki, I. On the Path towards a “Greener” EU: A Mini Review on Flax (Linum usitatissimum L.) as a Case Study. Plants 2023, 12, 1102. [Google Scholar] [CrossRef]
  70. Heller, K.; Byczyńska, M. The impact of environmental factors and applied agronomy on quantitative and qualitative traits of flax fiber. J. Nat. Fibers 2015, 12, 26–38. [Google Scholar] [CrossRef]
  71. Kaur, V.; Yadav, R.; Wankhede, D.P. Linseed (Linum usitatissimum L.) genetic resources for climate change intervention and its future breeding. J. Appl. Nat. Sci. 2017, 9, 1112–1118. [Google Scholar] [CrossRef] [Green Version]
  72. Dash, P.K.; Cao, Y.; Jailani, A.K.; Gupta, P.; Venglat, P.; Xiang, D.; Rai, R.; Sharma, R.; Thirunavukkarasu, N.; Abdin, M.Z.; et al. Genome-wide analysis of drought induced gene expression changes in flax (Linum usitatissimum). GM Crops Food 2014, 5, 106–119. [Google Scholar] [CrossRef] [Green Version]
  73. Tawfik, R.S.; Badr, A.; Sammour, R.; Ibrahim, U.; Matter, M.; Sakr, M. Improvement of flax drought tolerance using gene transfer. Plant Tissue Cult. Biotechnol. 2016, 26, 197–207. [Google Scholar] [CrossRef] [Green Version]
  74. Asgarinia, P.; Mirlohi, A.; Saeidi, G.; MohamadiMirik, A.A.; Gheysari, M.; Razavi, V.S. Selection criteria for assessing drought tolerance in a segregating population of flax (Linum usitatissimum L.). Can. J. Plant Sci. 2016, 97, 424–437. [Google Scholar] [CrossRef] [Green Version]
  75. Qi, X.S.; Wang, X.R.; Xu, J.; Zhang, J.P.; Mi, J. Drought-resistance evaluation of flax germplasm at adult plant stage. Sci. Agric. Sin. 2010, 43, 3076–3087. [Google Scholar]
  76. Sharma, J.C.; Tomar, S.S.; Shivran, R.K.; Prakash, C. Water requirement water use efficiency consumptive use yield and quality parameters of linseed (Linum usitatissimum L.) varieties as influenced by fertility levels irrigation scheduling. Adv. Life Sci. 2012, 1, 180–182. [Google Scholar]
  77. Kaur, V.; Yadav, S.K.; Wankhede, D.P.; Pulivendula, P.; Kumar, A.; Chinnusamy, V. Cloning and characterization of a gene encoding MIZ1, a domain of unknown function protein and its role in salt and drought stress in rice. Protoplasma 2020, 257, 475–487. [Google Scholar] [CrossRef] [PubMed]
  78. Gowda, V.R.P.; Henry, A.; Yamauchi, A.; Shashidhar, H.E.; Serraj, R. Root biology and genetic improvement for drought avoidance in rice. Field Crop Res. 2011, 122, 1–13. [Google Scholar] [CrossRef]
  79. Tuberosa, R.; Sanguineti, M.C.; Landi, P.; Giuliani, M.M.; Salvi, S.; Conti, S. Identification of QTLs for root characteristics in maize grown in hydroponics and analysis of their overlap with QTLs for grain yield in the field at two water regimes. Plant Mol. Biol. 2002, 48, 697–712. [Google Scholar] [CrossRef]
  80. Manschadi, A.M.; Christopher, J.; deVoil, P.; Hammer, G.L. The role of root architectural traits in adaptation of wheat to water-limited environments. Funct. Plant Biol. 2006, 33, 823–837. [Google Scholar] [CrossRef]
  81. Soto-Cerda, B.J.; Cloutier, S.; Gajardo, H.A.; Aravena, G.; Quian, R. Identifying drought-resilient flax genotypes and related candidate genes based on stress indices, root traits and selective sweep. Euphytica 2019, 215, 41. [Google Scholar] [CrossRef]
  82. Soto-Cerda, B.J.; Cloutier, S.; Gajardo, H.A.; Aravena, G.; Quian, R.; You, F.M. Drought response of flax accessions and identification of quantitative trait nucleotides (QTNs) governing agronomic and root traits by genome-wide association analysis. Mol. Breed. 2020, 40, 15. [Google Scholar] [CrossRef]
  83. Mishra, N.; Tripathi, M.K.; Tiwari, S.; Tripathi, N.; Gupta, N.; Sharma, A. Morphological and physiological performance of Indian soybean [Glycine max (L.) Merrill.] genotypes in respect to drought. Legume Res. 2021, 1–9. [Google Scholar] [CrossRef]
  84. Choudhary, M.L.; Tripathi, M.K.; Tiwari, S.; Pandya, R.K.; Gupta, N.; Tripathi, N.; Parihar, P. Screening of Pearl Millet [Pennisetum glaucum (L.) R. Br.] Germplam Lines for Drought Tolerance Based on Morpho-physiological Traits and SSR Markers. Curr. J. Appl. Sci. Technol. 2021, 40, 46–63. [Google Scholar] [CrossRef]
  85. Kachare, S.; Tiwari, S.; Tripathi, N. Expression of DREB1, RBCL, PIP, SGR genes and morpho-physiological changes under water stress in soybean. J. Plant Biochem. Biotechnol. 2022, 32, 338–355. [Google Scholar] [CrossRef]
  86. Kariuki, L.W.; Masinde, P.; Githiri, S.; Onyango, A.N. Effect of water stress on growth of three linseed (Linum usitatissimum L.) varieties. SpringerPlus 2016, 5, 759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Ansari, A.; Razmjoo, J.; Karimmojeni, H. Mycorrhizal colonization and seed treatment with salicylic acid to improve physiological traits and tolerance of flaxseed (Linum usitatissimum L.) plants grown under drought stress. Acta Physiol. Plant 2016, 38, 34. [Google Scholar] [CrossRef]
  88. Dubey, S.; Bhargava, A.; Fuentes, F.; Shukla, S.; Srivastava, S. Effect of salinity stress on yield and quality parameters in flax (Linum usitatissimum L.). Not. Bot. Horti Agrobot. 2020, 48, 954–966. [Google Scholar] [CrossRef]
  89. Singh, M.; Nara, U.; Kumar, A.; Choudhary, A.; Singh, H.; Thapa, S. Salinity tolerance mechanisms and their breeding implications. J. Genet. Eng. Biotechnol. 2021, 19, 173. [Google Scholar] [CrossRef]
  90. Tao, R.; Ding, J.; Li, C.; Zhu, X.; Guo, W.; Zhu, M. Evaluating and Screening of Agro-Physiological Indices for Salinity Stress Tolerance in Wheat at the Seedling Stage. Front. Plant Sci. 2021, 12, 646175. [Google Scholar] [CrossRef]
  91. Shrivastava, P.; Kumar, R. Soil salinity: A serious environmental issue and plant growth promoting bacteria as one of the tools for its alleviation. Saudi J. Biol. Sci. 2015, 22, 123–131. [Google Scholar] [CrossRef] [Green Version]
  92. Kocak, M.Z.; Göre, M.; Kurt, O. The effect of different salinity levels on germination development of some flax (Linum usitatissimum L.) varieties. J. Food Sci. Technol. 2022, 10, 657–662. [Google Scholar] [CrossRef]
  93. El-Afry, M.M.; El-Okkiah, S.A.; El-Kady, E.S.A.; El-Yamanee, G.S.A. Exogenous application of ascorbic acid for alleviation the adverse effects of salinity stress in flax (Linum usitatissimum L.). Middle East J. 2018, 7, 716–739. [Google Scholar]
  94. Nasri, N.; Maatallah, S.; Saidi, I.; Lachaal, M. Influence of salinity on germination, seedling growth, ion content and acid phosphatase activities of Linum usitatissimum L. J. Anim. Plant Sci. 2017, 27, 517–521. [Google Scholar]
  95. Patil, N.M.; Datir, S.S.; Shah, P.V. Salt-induced physiological and biochemical changes in two varieties of Linum usitatissimum L. Int. J. Curr. Microbiol. Appl. Sci. 2015, 4, 296–304. [Google Scholar]
  96. Kaya, M.D.; Day, S.; Cikili, Y.; Arslan, N. Classification of some linseed (Linum usitatissimum L.) genotypes for salinity tolerance using germination, seedling growth, and ion content. Chil. J. Agric. Res. 2012, 72, 27–32. [Google Scholar] [CrossRef] [Green Version]
  97. Khan, N. Designing Genomic Solutions for Abiotic Traits in Flax (Linum usitatissimum L.). Ph.D. Thesis, University of Ottawa, Ottawa, ON, Canada, 2022. Submitted. [Google Scholar]
  98. Guo, R.; Zhou, J.; Ren, G.; Hao, W. Physiological responses of linseed seedlings to iso osmotic polyethylene glycol, salt, and alkali stresses. Agron. J. 2013, 105, 764–772. [Google Scholar] [CrossRef]
  99. Miart, F.; Fontaine, J.X.; Mongelard, G.; Wattier, C.; Lequart, M.; Bouton, S.; Molinie, R.; Dubrulle, N.; Fournet, F.; Demailly, H. Integument-Specific Transcriptional Regulation in the Mid-Stage of Flax Seed Development Influences the Release of Mucilage and the Seed Oil Content. Cells 2021, 10, 2677. [Google Scholar] [CrossRef]
  100. Yu, Y.; Chen, H.; Yang, Y.; Lou, D.; Liang, C.; Yuan, H.; Wu, G.; Xu, C. Identification and Characterization of Differentially Expressed microRNAs and Target Gene Related to Flax Stem Development. J. Nat. Fibers 2022, 19, 5974–5990. [Google Scholar] [CrossRef]
  101. Wei, W.; Qian, D.; Qiaoling, J. Effects of NaCl stress on the biochemical characteristics in six species fiber flax seeding. Chin. Agric. Sci. Bull. 2013, 18, 017. [Google Scholar]
  102. Belkadhi, A.; De Haro, A.; Soengas, P.; Obregon, S.; Cartea, M.E.; Djebali, W.; Chaïbi, W. Salicylic acid improves root antioxidant defense system and total antioxidant capacities of flax subjected to cadmium. OMICS 2013, 17, 398–406. [Google Scholar] [CrossRef] [Green Version]
  103. Belkadhi, A.; De Haro, A.; Obregon, S.; Chaïbi, W.; Djebali, W. Positive effects of salicylic acid pretreatment on the composition of flax plastidial membrane lipids under cadmium stress. Environ. Sci. Pollut. Res. 2015, 22, 1457–1467. [Google Scholar] [CrossRef]
  104. Kaplan, M.E.; Simmons, E.R.; Hawkins, J.C.; Ruane, L.G.; Carney, J.M. Influence of cadmium and mycorrhizal fungi on the fatty acid profile of flax (Linum usitatissimum) seeds. J. Sci. Food Agric. 2015, 95, 2528–2532. [Google Scholar] [CrossRef]
  105. Grant, C.A.; Dribnenki, J.C.P.; Bailey, L.D. Cadmium and zinc concentrations and ratios in seed and tissue of solin (cv LinolaTM 947) and flax (cvs McGregor and Vimy) as affected by nitrogen and phosphorus fertiliser and Provide (Penicillium bilaji). J. Sci. Food Agric. 2000, 80, 1735–1743. [Google Scholar] [CrossRef]
  106. Smykalova, I.; Vrbova, M.; Tejklova, E.; Vetrovcova, M.; Griga, M. Large scale screening of heavy metal tolerance in flax/linseed (Linum usitatissimum L.) tested in vitro. Ind. Crops Prod. 2010, 32, 527–533. [Google Scholar] [CrossRef]
  107. Soudek, P.; Katrušáková, A.; Sedláček, L.; Petrová, Š.; Kočí, V.; Maršík, P.; Griga, M.; Vaněk, T. Effect of heavy metals on inhibition of root elongation in 23 cultivars of flax (Linum usitatissimum L.). Arch. Environ. Contamin. Toxicol. 2010, 59, 194–203. [Google Scholar] [CrossRef]
  108. Darapuneni, M.K.; Morgan, G.D.; Ibrahim, A.M.; Duncan, R.W. Effect of vernalization and photoperiod on flax flowering time. Euphytica 2014, 195, 279–285. [Google Scholar] [CrossRef]
  109. Ramirez-Villegas, J.; Molero Milan, A.; Alexandrov, N.; Asseng, S.; Challinor, A.J.; Crossa, J.; Eeuwijk, F.V.; Ghanem, M.E.; Grenier, C.; Heinemann, A.B.; et al. CGIAR modeling approaches for resource-constrained scenarios: I. Accelerating crop breeding for a changing climate. Crop Sci. 2020, 60, 547–567. [Google Scholar] [CrossRef] [Green Version]
  110. Cross, R.H. Heat Stress Effects on Flowering and Reproduction in Linum usitatissimum (Flax). Master’s Thesis, University of Saskatchewan, Saskatoon, SK, Canada, 2002. [Google Scholar]
  111. Saha, D.; Mukherjee, P.; Dutta, S.; Meena, K.; Sarkar, S.K.; Mandal, A.B.; Dasgupta, T.; Mitra, J. Genomic insights into HSFs as candidate genes for high-temperature stress adaptation and gene editing with minimal off-target effects in flax. Sci. Rep. 2019, 9, 5581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Saha, D.; Shaw, A.K.; Datta, S.; Mitra, J. Evolution and functional diversity of abiotic stress-responsive NAC transcription factor genes in Linum usitatissimum L. Environ. Exp. Bot. 2021, 188, 104512. [Google Scholar] [CrossRef]
  113. Cross, R.H.; McKay, S.A.B.; McHughen, A.G.; Bonham-Smith, P.C. Heat stress effects on reproduction and seed set in Linum usitatissimum L. (flax). Plant Cell Environ. 2003, 26, 1013–1020. [Google Scholar] [CrossRef]
  114. Pokhrel, S.; Meyers, B.C. Heat-responsive microRNAs and phased small interfering RNAs in reproductive development of flax. Plant Direct. 2022, 6, 385. [Google Scholar] [CrossRef]
  115. Tripathi, N.; Tripathi, M.K.; Tiwari, S.; Payasi, D.K. Molecular Breeding to overcome biotic stresses in Soybean: Update. Plants 2022, 11, 1967. [Google Scholar] [CrossRef]
  116. Kenaschuk, E.O.; Rowland, G.G. Flax. In Harvest of Gold: The History of Field Crop Breeding in Canada; Slinkard, A.E., Knott, D.R., Eds.; University of Saskatchewan: Saskatoon, SK, Canada, 1993; pp. 173–176. [Google Scholar]
  117. Russell, G.E. Plant Breeding for Pest and Disease Resistance. Studies in Agriculture Sciences; Butterworth-Heinemann eBook: London, UK, 2013; p. 496. [Google Scholar]
  118. Lawrence, G.J.; Dodds, P.N.; Ellis, J.G. Transformation of the flax rust fungus, Melampsoralini: Selection via silencing of an avirulence gene. Plant J. 2010, 61, 364–369. [Google Scholar] [CrossRef] [Green Version]
  119. Flor, H. Tests for allelism of rust-resistance genes in flax. Crop Sci. 1965, 5, 415–418. [Google Scholar] [CrossRef] [Green Version]
  120. Khan, I.A. Flax, the Genus Linum Edited by Alostar D. Muir and Neil D. Westcott (Agriculture and Agri-Food Canada, Saskatoon); Taylor & Francis: London, UK, 2003; ISBN 0-415-30807-0. [Google Scholar]
  121. Kanapin, A.; Samsonova, A.; Rozhmina, T.; Bankin, M.; Logachev, A.; Samsonova, M. The Genome Sequence of Five Highly Pathogenic Isolates of Fusarium oxysporum f. sp. lini. Mol. Plant Microbe Interact. 2020, 33, 1112–1115. [Google Scholar] [CrossRef]
  122. Agrawal, S.; Narware, H.; Sahu, R. Inheritance of Fusarium wilt resistance in linseed. J. Oilseeds Res. 1991, 8, 231. [Google Scholar]
  123. Spielmeyer, W.; Green, A.; Bittisnich, D.; Mendham, N.; Lagudah, E. Identification of quantitative trait loci contributing to Fusarium wilt resistance on an AFLP linkage map of flax (Linum usitatissimum). Theor. Appl. Genet. 1998, 97, 633–641. [Google Scholar] [CrossRef]
  124. Spielmeyer, W.; Lagudah, E.; Mendham, N.; Green, A. Inheritance of resistance to flax wilt (Fusarium oxysporum f. sp. lini Schlecht) in a doubled haploid population of Linum usitatissimum L. Euphytica 1998, 101, 287–291. [Google Scholar] [CrossRef]
  125. Portyankin, D.; Karachan, V. Resistance to Fusarium wilt of perspective fiber flax varieties. Zashchita Rastenij 1999, 79, 308–314. [Google Scholar]
  126. Diederichsen, A.; Rozhmina, T.A.; Kudrjavceva, L.P. Variation patterns within 153 flax (Linum usitatissimum L.) genebank accessions based on evaluation for resistance to fusarium wilt, anthracnose and pasmo. Plant Genet. Resour. 2008, 6, 22–32. [Google Scholar] [CrossRef]
  127. Elwakil, M.A.; Arafat, N.S.; Shabana, Y.M. Identification, Variability and Phylogeny of Alternaria linicola Associated with Linseed in Egypt. Plant Pathol. J. 2022, 21, 33–40. [Google Scholar] [CrossRef]
  128. Singh, R.B.; Singh, R.N. Occurrence Status and Management of Alternaria Blight (Alternaria species) of Linseed (Linum usitatissimum). Indian J. Agric. Sci. 2005, 75, 277–280. [Google Scholar]
  129. Holi, S.K.; Meena, S. Management of Alternaria Blight of Linseed (Linum usitatissimum) Caused by Alternaria lini. J. Plant Sci. Res. 2015, 31, 47–50. [Google Scholar]
  130. Fitt, B.D.L.; Coskun, H. Occurrence and pathogenicity of Alternaria spp. on linseed in the UK. In Flax as a Fibre and Oil Bearing Crop, Proceedings of the FAO European Regional Workshop on Flax, Brno, Czech Republic, 18–20 June 1991; University of Hertfordshire Research Archieves; University of Hertfordshire: Hatfield, UK, 1991; pp. 338–346. [Google Scholar]
  131. Vloutoglou, I.; Fitt, B.; Lucas, J. Infection of Linseed by Alternaria linicola; Effects of Inoculum Density, Temperature, Leaf Wetness and Light Regime. Eur. J. Plant Pathol. 1999, 105, 585–595. [Google Scholar] [CrossRef]
  132. Singh, R.B.; Singh, H.K.; Parmar, A. Yield loss assessment due to Alternaria blight and its management in linseed. Pak. J. Biol. Sci. 2014, 17, 511–516. [Google Scholar] [CrossRef] [PubMed]
  133. Speck, A.; Trouvé, J.P.; Enjalbert, J.; Geffroy, V.; Joets, J.; Moreau, L. Genetic Architecture of Powdery Mildew Resistance Revealed by a Genome-Wide Association Study of a Worldwide Collection of Flax (Linum usitatissimum L.). Front. Plant Sci. 2022, 13, 871633. [Google Scholar] [CrossRef] [PubMed]
  134. Ashry, N.; Mansour, M.; Aly, A.; Zayed, S. Genetic studies on powdery mildew resistance of flax, yield and some yield components. Egypt J. Agric. Res. 2002, 80, 1525–1537. [Google Scholar]
  135. Rashid, K.; Duguid, S. Inheritance of resistance to powdery mildew in flax. Can. J. Plant Pathol. 2005, 27, 404–409. [Google Scholar] [CrossRef]
  136. Rashid, K.Y.; Desjardins, M.L.; Duguid, S. Diseases of flax in Manitoba and Saskatchewan in 2009. Can. Plant Dis. Surv. 2010, 90, 136–137. [Google Scholar]
  137. Cullis, C.A. Mechanisms and control of rapid genomic changes in flax. Ann. Bot. 2005, 95, 201–206. [Google Scholar] [CrossRef]
  138. Maggioni, L.; Pavelek, M.; van Soest, L.J.M.; Lipman, E. Flax genetic resources in Europe. In Proceedings of the Ad Hoc Meeting of European Cooperative Programme for Crop Genetic Resources Network ECP/GR of the International Plant Genetic Resources Institute, Prague, Czech Republic, 7–8 December 2001. [Google Scholar]
  139. Shamov, D. Status of the Bulgarian national flax collection. In Proceedings of the Ad Hoc Meeting, Prague, Czech Republic, 7–8 December 2001. [Google Scholar]
  140. UPOV. Working paper on the revision of test guidelines for flax, linseed (Linum usitatissimum L.). In Proceedings of the International Union for the Protection of New Varieties of Plants, Geneva, Technical Working Party for Agricultural Crops, Twentieth Session, Beltsville, MD, USA, 13–17 May 1991. [Google Scholar]
  141. Diederichsen, A.; Hammer, K. Variation of cultivated flax (Linum usitatissimum L. subsp. usitatissimum) and its wild progenitor pale flax (subsp. angustifolium (Huds.) Thell.). Genet. Resour. Crop Evol. 1995, 42, 263–272. [Google Scholar] [CrossRef]
  142. Pavelek, M. Recent state of international flax database and future development’, European cooperative network on flax, breeding research group. In Proceedings of the Report of Flax Genetic Resources Workshop, Second Meeting, Brno, Czech Republic, 8–10 November 1994; pp. 57–63. [Google Scholar]
  143. Pavelek, M. Further development of International Flax Data Base and special descriptors for more detail evaluation of agronomic and processing characters. In Breeding for Fibre and Oil Quality in Flax’, Proceedings of the Third Meeting of the International Flax Breeding Group, St Valery enCaux, France, 7–8 November 1995; Centre Technique Pour l’étude et l’amélioration du lin (CETEAL): Paris, France, 1995; pp. 1–13. [Google Scholar]
  144. Rakouský, S.; Tejklova, E.; Wiesner, I.; Wiesnerova, D.; Kocabek, T.; Ondrej, M. TDNA induced mutations and somaclonal variants of flax. Nat. Fibres 1998, 2, 244–246. [Google Scholar]
  145. Rakouský, S.; Tejklov_a, E.; Wiesner, I. Recent advances in flax biotechnology in the Czech Republic. Nat. Fibres 2001, 1, 151–155. [Google Scholar]
  146. Jia, G.; You, F.M.; Cloutier, S.; Deyholos, M.K.; Booker, H. The flax breeding database (FlaxDB). In Proceedings of the XXVII Plant and Animal Genome Conference, San Diego, CA, USA, 12–16 January 2019; p. PE1082. [Google Scholar]
  147. Biradar, S.A.; Ajithkumar, K.; Rajanna, B.; Savitha, A.S.; Shubha, G.V.; Shankergoud, I.; Singh, P.K. Prospects and challenges in linseed (Linum usitatissimum L.) production: A review. J. Oilseeds Res. 2016, 33, 1–13. [Google Scholar]
  148. Tadesse, N.; Lay, C.; Dybing, C.D. Comparative seed yield performance of high-by-high and low-by-high crosses in flax. Plant Breed. 1997, 116, 561–566. [Google Scholar] [CrossRef]
  149. Pavelek, M. International flax database. In Proceedings of the Flax Genetic Resources in Europe, Ad Hoc Meeting, Prague, Czech Republic, 7–8 December 2001; Maggioni, L., Pavelek, M., van Soest, L.J.M., Lipman, E., Eds.; International Plant Genetic Resources Institute: Rome, Italy, 2001. [Google Scholar]
  150. Tejklova, E. Some factors affecting anther cultures in Linum usitatissimum L. Rostl. Vyrob. 1996, 42, 249–260. [Google Scholar]
  151. Tejklova, E. Study of anther culture in flax (Linum usitatissimum L.). Nat. Fibres 1998, 2, 202–209. [Google Scholar]
  152. Yurenkova, S.I.; Kubrak, S.V.; Titok, V.V.; Khotyljova, L.V. Flax species polymorphism for isozyme and metabolic markers. Genetika 2005, 41, 334–340. [Google Scholar] [CrossRef]
  153. Ijaz, A.; Shahbaz, A.; Ullah, I.; Ali, S. Molecualar Characterization of Linseed Germplasm Using RAPD DNA Fingerprinting Markers. Am.-Eurasian J. Agric. Environ. Sci. 2013, 13, 1266–1274. [Google Scholar]
  154. Rowland, G.G. Growing Flax: Production, Management and Diagnostic Guide; Flax Council of Canada and Saskatchewan Flax Development Commission: Saskatoon, SK, Canada, 1998. [Google Scholar]
  155. Gill, K.S. Evolutionary Relationships Among Linum Species. Ph.D. Thesis, University of California, Riverside, CA, USA, 1966. [Google Scholar]
  156. Carnahan, H.L. Combining Ability in Flax (Linum usitatissimum). Master’s Thesis, University of Minnesota, Minneapolis, MN, USA, 1947. [Google Scholar]
  157. Dubey, D.K.; Singh, S.P. Extent of heterosis and problems of hybrid seed production in linseed (Linum usitatissimum L.). Balwant Vidyapeeth J. Agric. Sci. Res. 1969, 6, 83–87. [Google Scholar]
  158. Kumar, S.; Singh, S.P. Inheritance of male sterility in some introduced varieties of linseed (Linum usitatissimum L.). Indian J. Agric. Sci. 1970, 40, 184–191. [Google Scholar]
  159. Kumar, S.; Singh, S.P. Inheritance of partial male fertility in linseed (Linum usitatissimum L.). Indian J. Agric. Sci. 1972, 42, 34–38. [Google Scholar]
  160. Dubey, D.K.; Singh, S.P. Mechanism of pollen abortion in three male sterile lines of flax (Linum usitatissimum L.). Crop Sci. 1965, 5, 121–124. [Google Scholar] [CrossRef] [Green Version]
  161. Dubey, D.K.; Singh, S.P. Use of cytoplasmatic male sterility for the production of hybrid seeds in flax Linum usitatissimum L. Crop Sci. 1966, 6, 125–126. [Google Scholar] [CrossRef]
  162. Kumar, S. Haploid linseed. Agra Univ. J. Res. 1971, 40, 23–25. [Google Scholar]
  163. Pant, S.C.; Mishra, V.K. Heterosis over superior parents under diallel cross in linseed X flax (Linum usitatissimum L.). Indian J. Plant Genet. Resour. 2008, 21, 93–97. [Google Scholar]
  164. Sood, S.; Kalia, N.R.; Bhateria, S.; Kumar, S. Detection of genetic components of variation for some biometrical traits in Linum usitatissimum L. in sub-mountain Himalayan region. Euphytica 2007, 155, 107–115. [Google Scholar] [CrossRef]
  165. Bhateria, S.; Sood, S.P.; Pathania, A. Genetic analysis of quantitative traits across environments in linseed (Linum usitatissimum L.). Euphytica 2006, 150, 185–194. [Google Scholar] [CrossRef]
  166. George, K.P.; Nayar, K.K. Early dwarf mutants in linseed induced by gamma rays. Curr. Sci. 1973, 42, 137–138. [Google Scholar]
  167. Rai, M.; Das, K. Potentiality and genetic variability in irradiated population of linseed. Indian J. Genet. Plant Breed. 1976, 36, 20–25. [Google Scholar]
  168. Pospisil, B. The significance of chemical muatgens in breeding flax. Genet. Slechteni 1974, 10, 187–196. [Google Scholar]
  169. Bianu, M.; Marki, A.; Ochesanu, C. The flexuous stem-a new gene in Linum usitatissimum induced by the alkylating agents. Botanique 1972, 25, 171–175. [Google Scholar]
  170. Deshpande, R.B. A note on the occurrence of chlorophyll deficiency in linseed. Curr. Sci. 1939, 8, 168–169. [Google Scholar]
  171. Levan, A. Experimentally induced chlorophyll mutants in flax. Hereditas 1944, 30, 225–230. [Google Scholar] [CrossRef]
  172. Sharov, I. The mutation spectrum of fibre flax induced by chemical mutagens. Refrat. Zh 1971, 8, 55–119. [Google Scholar]
  173. Green, A.G. A mutant genotype of flax (Linum usitatissimum) containing very low levels of linolenic acid in its seed oil. Can. J. Plant Sci. 1986, 66, 499–503. [Google Scholar] [CrossRef]
  174. Gill, K.S. Linseed; Indian Council of Agricultural Research: New Delhi, India, 1987. [Google Scholar]
  175. Cloutier, S.; Miranda, E.; Ragupathy, R.; Radovanovic, N.; Reimer, E.; Walichnowski, A. Integrated consensus genetic and physical maps of flax (Linum usitatissimum L.). Theor. Appl. Genet. 2012, 125, 1783–1795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Li, C.; Huang, Y.; Huang, R.; Wu, Y.; Wang, W. The genetic architecture of amylose biosynthesis in maize kernel. Plant Biotechnol. J. 2018, 16, 688–695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Remington, D.L.; Thornsberry, J.M.; Matsuoka, Y.; Wilson, L.M.; Whitt, S.R.; Doebley, J.; Kresovich, S.; Goodman, M.M.; Buckler, E.S. Structure of linkage disequilibrium and phenotypic associations in the maize genome. Proc. Natl. Acad. Sci. USA 2001, 98, 11479–11484. [Google Scholar] [CrossRef]
  178. Sonah, H.; O’Donoughue, L.; Cober, E.; Rajcan, I.; Belzile, F. Identification of loci governing eight agronomic traits using a GBS-GWAS approach and validation by QTL mapping in soya bean. Plant Biotechnol. J. 2015, 3, 10. [Google Scholar] [CrossRef] [PubMed]
  179. Soto-Cerda, B.J.; Diederichsen, A.; Ragupathy, R.; Cloutier, S. Genetic characterization of a core collection of flax (Linum usitatissimum L.) suitable for association mapping studies and evidence of divergent selection between fiber and linseed types. BMC Plant Biol. 2013, 13, 78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Soto-Cerda, B.J.; Diederichsen, A.; Duguid, S.; Booker, H.; Rowland, G.; Cloutier, S. The potential of pale flax as a source of useful genetic variation for cultivated flax revealed through molecular diversity and association analyses. Mol. Breed. 2014, 34, 2091–2107. [Google Scholar] [CrossRef] [Green Version]
  181. Soto-Cerda, B.J.; Duguid, S.; Booker, H.; Rowland, G.; Diederichsen, A.; Cloutier, S. Association mapping of seed quality traits using the Canadian flax (Linum usitatissimum L.) core collection. Theor. Appl. Genet. 2014, 127, 881–896. [Google Scholar] [CrossRef] [Green Version]
  182. Cullis, C.A. Genetics and Genomics of Linum; Springer International Publishing: Cham, Switzerland, 2019. [Google Scholar]
  183. Kenaschuk, E.O. High linolenic acid flax. Google Patents Cullis CA (2007) Flax. In Genome Mapping and Molecular Breeding in Plants; Kole, C., Ed.; Springer: Berlin/Heidelberg, Germany, 2005; Volume 2, pp. 275–295. [Google Scholar]
  184. Zhang, J.; Xie, Y.; Dang, Z.; Wang, L.; Li, W.; Zhao, W.; Zhao, L.; Dang, Z. Oil content and fatty acid components of oilseed flax under different environments in China. Agron. J. 2016, 108, 365–372. [Google Scholar] [CrossRef]
  185. Botstein, D.; White, R.L.; Skolnick, M.; Davis, R.W. Construction of a genetic linkage map in man using restriction fragment length polymorphisms. Amer. J. Hum. Genet. 1980, 32, 314. [Google Scholar]
  186. Paterson, A.H.; Lander, E.S.; Hewitt, J.D.; Peterson, S.; Lincoln, S.E.; Tanksley, S.D. Resolution of quantitative traits into Mendelian factors by using a complete linkage map of restriction fragment length polymorphisms. Nature 1988, 335, 721–726. [Google Scholar] [CrossRef]
  187. Oh, T.J.; Gorman, M.; Cullis, C.A. RFLP and RAPD mapping in flax (Linumusitatisimum). Theor. Appl. Genet. 2000, 101, 590–593. [Google Scholar] [CrossRef]
  188. Kumar, S.; You, F.M.; Cloutier, S. Genome wide SNP discovery in flax through next generation sequencing of reduced representation libraries. BMC Genom. 2012, 13, 684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Fukushima, A.; Kusano, M.; Redestig, H.; Arita, M.; Saito, K. Integrated omics approaches in plant systems biology. Curr. Opin. Chem. Biol. 2009, 13, 532–538. [Google Scholar] [CrossRef]
  190. Weckwerth, W.; Ghatak, A.; Bellaire, A.; Chaturvedi, P.; Varshney, R.K. PANOMICS meets germplasm. Plant Biotechnol. J. 2020, 18, 1507–1525. [Google Scholar] [CrossRef] [Green Version]
  191. Pazhamala, L.T.; Kudapa, H.; Weckwerth, W.; Millar, A.H.; Varshney, R.K. Systems biology for crop improvement. Plant Genome 2021, 14, e20098. [Google Scholar] [CrossRef]
  192. Kumari, J.; Mahatman, K.K.; Sharma, S.; Singh, A.K.; Adhikari, S.; Bansal, R.; Kaur, V.; Kumar, S.; Yadav, M.C. Recent advances in different omics mechanism for drought stress tolerance in rice. Russ. J. Plant Physiol. 2022, 69, 18. [Google Scholar] [CrossRef]
  193. Chaudhary, J.; Patil, G.B.; Sonah, H.; Deshmukh, R.K.; Vuong, T.D.; Valliyodan, B.; Nguyen, H.T. Expanding omics resources for improvement of soybean seed composition traits. Front. Plant Sci. 2015, 6, 1021. [Google Scholar] [CrossRef] [Green Version]
  194. Chaudhary, J.; Khatri, P.; Singla, P.; Kumawat, S.; Kumari, A.; Vikram, A.; Jindal, S.K.; Kardile, H.; Kumar, R.; Sonah, H.; et al. Advances in omics approaches for abiotic stress tolerance in tomato. Biology 2019, 8, 90. [Google Scholar] [CrossRef] [Green Version]
  195. Akhmetshina, A.O.; Strygina, K.V.; Khlestkina, E.K.; Porokhovinova, E.A.; Brutch, N.B. High-throughput sequencing techniques to flax genetics and breeding. Ecol. Genet. 2020, 18, 103–124. [Google Scholar] [CrossRef] [Green Version]
  196. Yadav, B.; Kaur, V.; Narayan, O.P.; Yadav, S.K.; Kumar, A.; Wankhede, D.P. Integrated omics approaches for flax improvement under abiotic and biotic stress: Current status and future prospects. Front. Plant Sci. 2022, 13, 931275. [Google Scholar] [CrossRef]
  197. Saroha, A.; Pal, D.; Kaur, V.; Kumar, S.; Bartwal, A.; Aravind, J.; Radhamani, J.; Rajkumar, S.; Kumar, R.; Gomashe, S.S.; et al. Agro-morphological variability and genetic diversity in linseed (Linum usitatissimum L.) germplasm accessions with emphasis on flowering and maturity time. Genet. Resour. Crop Evol. 2022, 69, 315–333. [Google Scholar] [CrossRef]
  198. Kaur, V.; Kumar, S.; Yadav, R.; Wankhede, D.P.; Aravind, J.; Radhamani, J.; Rana, J.C.; Kumar, A. Analysis of genetic diversity in Indian and exotic linseed germplasm and identification of trait specific superior accessions. J. Environ. Biol. 2018, 39, 702–709. [Google Scholar] [CrossRef]
  199. Uysal, H.; Fu, Y.-B.; Kurt, O.; Peterson, G.W.; Diederichsen, A.; Kusters, P. Genetic diversity of cultivated flax (Linum usitatissimum L.) and its wild progenitor pale flax (Linum bienne mill.) as revealed by ISSR markers. Genet. Resour. Crop Evol. 2010, 57, 1109–1119. [Google Scholar] [CrossRef]
  200. Fu, Y.-B.; Peterson, G.; Diederichsen, A.; Richards, K. RAPD analysis of genetic relationships of seven flax species in the genus Linum L. Genet. Resour. Crop Evol. 2002, 49, 253–259. [Google Scholar] [CrossRef]
  201. Wang, Z.; Hobson, N.; Galindo, L.; Zhu, S.; Shi, D.; McDill, J.; Yang, L.; Hawkins, S.; Neutelings, G.; Datla, R.; et al. The genome of flax (Linum usitatissimum) assembled de novo from short shotgun sequence reads. Plant J. 2012, 72, 461–473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Spindel, J.E.; Dahlberg, J.; Colgan, M.; Hollingsworth, J.; Sievert, J.; Staggenborg, S.H.; Hutmacher, R.; Jansson, C.; Vogel, J.P. Association mapping by aerial drone reveals 213 genetic associations for Sorghum bicolor biomass traits under drought. BMC Genom. 2018, 19, 679. [Google Scholar] [CrossRef] [Green Version]
  203. Lasky, J.R.; Upadhyaya, H.D.; Ramu, P.; Deshpande, S.; Hash, C.T.; Bonnette, J.; Juenger, T.E.; Hyma, K.; Acharya, C.; Mitchell, S.E.; et al. Genome-environment associations in sorghum landraces predict adaptive traits. Sci. Adv. 2015, 1, e1400218. [Google Scholar] [CrossRef] [Green Version]
  204. Badigannavar, A.; Teme, N.; de Oliveira, A.C.; Li, G.; Vaksmann, M.; Viana, V.E.; Ganapathi, T.R.; Sarsu, F. Physiological, genetic and molecular basis of drought resilience in sorghum (Sorghum bicolor (L.)). Indian J. Plant Physiol. 2018, 23, 670–688. [Google Scholar] [CrossRef]
  205. Guo, Z.; Yang, W.; Chang, Y.; Ma, X.; Tu, H.; Xiong, F.; Jiang, N.; Feng, H.; Huang, C.; Yang, P.; et al. Genome-wide association studies of image traits reveal genetic architecture of drought resistance in rice. Mol. Plant 2018, 11, 789–805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Mangin, B.; Bonnafous, F.; Blanchet, N.; Boniface, M.C.; Bret-Mestries, E.; Carrère, S.; Cottret, L.; Legrand, L.; Marage, G.; Pegot-Espagnet, P.; et al. Genomic prediction of sunflower hybrids oil content. Front. Plant Sci. 2017, 8, 1633. [Google Scholar] [CrossRef] [Green Version]
  207. Dossa, K.; Li, D.; Zhou, R.; Yu, J.; Wang, L.; Zhang, Y.; You, J.; Liu, A.; Mmadi, M.A.; Fonceka, D.; et al. The genetic basis of drought tolerance in the high oil crop Sesamum indicum. J. Plant Biotechnol. 2019, 17, 1788–1803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Li, S.; Wang, L.; Zhang, Y.; Zhu, G.; Zhu, X.; Xia, Y.; Li, J.; Gao, X.; Wang, S.; Zhang, J.; et al. Genome-Wide Identification and Function of Aquaporin Genes During Dormancy and Sprouting Periods of Kernel-Using Apricot (Prunus armeniaca L.). Front. Plant Sci. 2021, 12, 690040. [Google Scholar] [CrossRef] [PubMed]
  209. Shivaraj, S.M.; Deshmukh, R.; Bhat, J.A.; Sonah, H.; Bélanger, R.R. Understanding aquaporin transport system in eelgrass (Zostera marina L.), an aquatic plant species. Front. Plant Sci. 2017, 8, 1334. [Google Scholar] [CrossRef] [Green Version]
  210. Thambugala, D.P. Analysis of Genetic Diversity and Expression of Genes Involved in Fatty Acid Composition in Flax (Linum usitatissimum L.) and Comparative Genomic Analysis of their Loci. Ph.D. Thesis, Department of Plant Science, University of Manitoba, Winnipeg, MB, Canada, 2015. [Google Scholar]
  211. He, L.; Xiao, J.; Rashid, K.Y.; Yao, Z.; Li, P.; Jia, G.; Wang, X.; Cloutier, S.; You, F.M. Genome-wide association studies for pasmo resistance in flax (Linum usitatissimum L.). Front. Plant Sci. 2019, 9, 1982. [Google Scholar] [CrossRef]
  212. You, F.M.; Rashid, K.Y.; Zheng, C.; Khan, N.; Li, P.; Xiao, J.; He, L.; Yao, Z.; Cloutier, S. Insights into the Genetic Architecture and Genomic Prediction of Powdery Mildew Resistance in Flax (Linum usitatissimum L.). Int. J. Mol. Sci. 2022, 23, 4960. [Google Scholar] [CrossRef]
  213. Crossa, J.; Pérez-Rodríguez, P.; Cuevas, J.; Montesinos-López, O.; Jarquín, D.; Campos, G.D.L.; Burgueño, J.; González-Camacho, J.M.; Pérez-Elizalde, S.; Beyene, Y.; et al. Genomic selection in plant breeding: Methods, models, and perspectives. Trends Plant Sci. 2017, 22, 961–975. [Google Scholar] [CrossRef]
  214. Abed, A.; Pérez-Rodríguez, P.; Crossa, J.; Belzile, F. When less can be better: How can we make genomic selection more cost-effective and accurate in barley? Theor. Appl. Genet. 2018, 131, 1873–1890. [Google Scholar] [CrossRef]
  215. Deshmukh, R.; Sonah, H.; Patil, G.; Chen, W.; Prince, S.; Mutava, R.; Vooung, T.; Valliyodan, D.; Nguyen, H.T. Integrating omic approaches for abiotic stress tolerance in soybean. Front. Plant Sci. 2014, 5, 244. [Google Scholar] [CrossRef]
  216. Budhlakoti, N.; Kushwaha, A.K.; Rai, A.; Chaturvedi, K.K.; Kumar, A.; Pradhan, A.K.; Kumar, U.; Kumar, R.R.; Juliana, P.; Mishra, D.C.; et al. Genomic selection: A tool for accelerating the efficiency of molecular breeding for development of climate-resilient crops. Front. Genet. 2022, 13, 832153. [Google Scholar] [CrossRef] [PubMed]
  217. Sertse, D.; You, F.M.; Ravichandran, S.; Soto-Cerda, B.J.; Duguid, S.; Cloutier, S. Loci harboring genes with important role in drought and related abiotic stress responses in flax revealed by multiple GWAS models. Theoret. Appl. Genet. 2021, 134, 191–212. [Google Scholar] [CrossRef] [PubMed]
  218. You, F.M.; Cloutier, S. Mapping quantitative trait loci onto chromosome-scale pseudomolecules in flax. Methods Protoc. 2020, 3, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Asgarinia, P.; Cloutier, S.; Duguid, S.R.; Rashid, K.; Mirlohi, A.; Banik, M.; Saeidi, G. Mapping quantitative trait loci for powdery mildew resistance in flax. Crop. Sci. 2013, 53, 2462–2472. [Google Scholar] [CrossRef] [Green Version]
  220. Wu, J.; Zhao, Q.; Sun, D.; Wu, G.; Zhang, L.; Yuan, H.; Yu, Y.; Zhang, S.; Yang, X.; Li, Z.; et al. Transcriptome analysis of flax (Linum usitatissimum L.) undergoing osmotic stress. Ind. Crop Prod. 2018, 116, 215–223. [Google Scholar] [CrossRef]
  221. Cartagena, J.A.; Marquez, G.P.B. Transcriptome analysis in Jatropha during abiotic stress response. In Oil Crop Genomics; Tombuloglu, H., Unver, T., Tombuloglu, G., Hakeem, K.R., Eds.; Springer: Cham, Switzerland, 2021; pp. 317–337. [Google Scholar]
  222. Moschen, S.; Di Rienzo, J.A.; Higgins, J.; Tohge, T.; Watanabe, M.; González, S.; Rivarola, M.; García-García, F.; Dopazo, J.; Hopp, H.E.; et al. Integration of transcriptomic and metabolic data reveals hub transcription factors involved in drought stress response in sunflower (Helianthus annuus L.). Plant Mol. Biol. 2017, 94, 549–564. [Google Scholar] [CrossRef]
  223. Leisner, C.P.; Yendrek, C.R.; Ainsworth, E.A. Physiological and transcriptomic responses in the seed coat of field-grown soybean (Glycine max L. Merr.) to abiotic stress. BMC Plant Biol. 2017, 17, 242. [Google Scholar] [CrossRef] [Green Version]
  224. Dmitriev, A.A.; Krasnov, G.S.; Rozhmina, T.A. Flax (Linum usitatissimum L.) response to non-optimal soil acidity and zinc deficiency. BMC Plant Biol. 2019, 19, 54. [Google Scholar] [CrossRef] [Green Version]
  225. Aghaee, P.; Rahmani, F. Biochemical and molecular responses of flax to 24-epibrassinosteroide seed priming under drought stress. J. Plant Interact. 2019, 14, 242–253. [Google Scholar] [CrossRef] [Green Version]
  226. Yu, Y.; Wu, G.; Yuan, H.; Cheng, L.; Zhao, D.; Huang, W.; Zhang, S.; Zhang, L.; Chen, H.; Zhang, J.; et al. Identification and characterization of miRNAs and targets in flax (Linum usitatissimum) under saline, alkaline, and saline-alkaline stresses. BMC Plant Biol. 2016, 16, 124. [Google Scholar] [CrossRef] [Green Version]
  227. Dash, P.K.; Rai, R.; Mahato, A.K.; Gaikwad, K.; Singh, N.K. Transcriptome landscape at different developmental stages of a drought tolerant cultivar of flax (Linum usitatissimum). Front. Chem. 2017, 5, 82. [Google Scholar] [CrossRef] [Green Version]
  228. Melnikova, N.V.; Dmitriev, A.A.; Belenikin, M.S.; Koroban, N.V.; Speranskaya, A.S.; Krinitsina, A.A.; Krasnov, G.S.; Lakunina, V.A.; Snezhkina, A.V.; Sadritdinova, A.F.; et al. Identification, expression analysis, and target prediction of flax genotroph MicroRNAs under normal and nutrient stress conditions. Front. Plant Sci. 2016, 7, 399. [Google Scholar] [CrossRef] [Green Version]
  229. Melnikova, N.V.; Dmitriev, A.A.; Belenikin, M.S.; Speranskaya, A.S.; Krinitsina, A.A.; Rachinskaia, O.A. Excess fertilizer responsive miRNAs revealed in Linum usitatissimum L. Biochimie 2015, 109, 36–41. [Google Scholar] [CrossRef] [PubMed]
  230. Zyablitsin, A.V.; Dmitriev, A.A.; Krasnov, G.S.; Bolsheva, N.L.; Rozhmina, T.A.; Muravenko, O.V.; Fedorova, M.S.; Snezhkina, A.V.; Kudryavtseva, A.V.; Melnikova, N.V. CAX3 gene is involved in flax response to high soil acidity and aluminium exposure. Mol. Biol. 2018, 52, 514–519. [Google Scholar] [CrossRef]
  231. Shivaraj, S.M.; Deshmukh, R.K.; Rai, R.; Bélanger, R.; Agrawal, P.K.; Dash, P.K. Genome-wide identification, characterization, and expression profile of aquaporin gene family in flax (Linum usitatissimum). Sci. Rep. 2017, 7, 46137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Wu, J.; Zhao, Q.; Wu, G.; Yuan, H.; Ma, Y.; Lin, H.; Pan, L.; Li, S.; Sun, D. Comprehensive analysis of differentially expressed unigenes under NaCl stress in flax (Linum usitatissimum L.) using RNA-Seq. Int. J. Mol. Sci. 2019, 20, 369. [Google Scholar] [CrossRef] [Green Version]
  233. Kostyn, K.; Czemplik, M.; Kulma, A.; Bortniczuk, M.; Skała, J.; Szopa, J. Genes of phenylpropanoid pathway are activated in early response to fusarium attack in flax plants. Plant Sci. 2012, 190, 103–115. [Google Scholar] [CrossRef]
  234. Dmitriev, A.A.; Krasnov, G.S.; Rozhmina, T.A.; Novakovskiy, R.O.; Snezhkina, A.V.; Fedorova, M.S.; Yurkevich, O.Y.; Muravenko, O.V.; Bolsheva, N.L.; Kudryavtseva, A.V.; et al. Differential gene expression in response to Fusarium oxysporum infection in resistant and susceptible genotypes of flax (Linum usitatissimum L.). BMC Plant Biol. 2017, 17, 253. [Google Scholar] [CrossRef] [Green Version]
  235. Boba, A.; Kostyn, K.; Kozak, B.; Wojtasik, W.; Preisner, M.; Prescha, A.; Gola, E.M.; Lysh, D.; Dudek, B.; Szopa, J.; et al. Fusarium oxysporum infection activates the plastidial branch of the terpenoid biosynthesis pathway in flax, leading to increased ABA synthesis. Planta 2020, 251, 50. [Google Scholar] [CrossRef] [Green Version]
  236. Wróbel-Kwiatkowska, M.; Lorenc-Kukula, K.; Starzycki, M.; Oszmiański, J.; Kepczyńska, E.; Szopa, J. Expression of β-1, 3-glucanase in flax causes increased resistance to fungi. Physiol. Mol. Plant Pathol. 2004, 65, 245–256. [Google Scholar] [CrossRef]
  237. Boba, A.; Kostyn, K.; Kozak, B.; Zalewski, I.; Szopa, J.; Kulma, A. Transcriptomic profiling of susceptible and resistant flax seedlings after Fusarium oxysporum lini infection. PLoS ONE 2021, 16, e0246052. [Google Scholar] [CrossRef]
  238. Yu, Y.; Huang, W.; Chen, H.; Wu, G.; Yuan, H.; Song, X.; Kang, Q.; Zhao, D.; Jiang, W.; Liu, Y.; et al. Identification of differentially expressed genes in flax (Linum usitatissimum L.) under saline–alkaline stress by digital gene expression. Gene 2014, 549, 113–122. [Google Scholar] [CrossRef] [PubMed]
  239. Wang, W.; Wang, L.; Wang, L.; Tan, M.; Ogutu, C.O.; Yin, Z.; Zhou, J.; Wang, J.; Wang, L.; Yan, X. Transcriptome analysis and molecular mechanism of linseed (Linum usitatissimum L.) drought tolerance under repeated drought using single-molecule long-read sequencing. BMC Genom. 2021, 22, 109. [Google Scholar] [CrossRef] [PubMed]
  240. Wu, W.; Nemri, A.; Blackman, L.M.; Catanzariti, A.-M.; Sperschneider, J.; Lawrence, G.J.; Dodds, P.N.; Jones, D.A.; Hardham, A.R. Flax rust infection transcriptomics reveals a transcriptional profile that may be indicative for rust Avr genes. PLoS ONE 2019, 14, e0226106. [Google Scholar] [CrossRef] [PubMed]
  241. Nemri, A.; Saunders, D.G.; Anderson, C.; Upadhyaya, N.M.; Win, J.; Lawrence, G.J.; Jones, D.A.; Kamoun, S.; Ellis, J.G.; Dodds, P.N. The genome sequence and effector complement of the flax rust pathogen Melampsoralini. Front. Plant Sci. 2014, 5, 98. [Google Scholar] [CrossRef] [PubMed]
  242. Galindo-Gonzalez, L.; Deyholos, M.K. RNA-seq transcriptome response of flax (Linum usitatissimum L.) to the pathogenic fungus Fusarium oxysporum f. sp. lini. Front. Plant Sci. 2016, 7, 1766. [Google Scholar] [CrossRef] [Green Version]
  243. Pontarin, N.; Molinié, R.; Mathiron, D.; Tchoumtchoua, J.; Bassard, S.; Gagneul, D.; Thiombiano, B.; Demailly, H.; Fontaine, J.X.; Guillot, X.; et al. Age-dependent metabolic profiles unravel the metabolic relationships within and between flax leaves (Linum usitatissimum). Metabolites 2020, 10, 218. [Google Scholar] [CrossRef]
  244. Weckwerth, W. Metabolomics in systems biology. Annu. Rev. Plant Biol. 2003, 54, 669–689. [Google Scholar] [CrossRef]
  245. Ghatak, A.; Chaturvedi, P.; Weckwerth, W. Metabolomics in plant stress physiology. Adv. Biochem. Eng./Biotechnol. 2018, 164, 187–236. [Google Scholar] [CrossRef]
  246. Hall, R.D. Plant metabolomics: From holistic hope, to hype, to hot topic. New Phytol. 2006, 169, 453–468. [Google Scholar] [CrossRef]
  247. Ibáñez, C.; García-Cañas, V.; Valdés, A.; Simó, C. Novel MS-based approaches and applications in food metabolomics. Trends Anal. Chem. 2013, 52, 100–111. [Google Scholar] [CrossRef]
  248. Kouba, M.; Mourot, J. A review of nutritional effects on fat composition of animal products with special emphasis on n-3 polyunsaturated fatty acids. Biochimie 2011, 93, 13–17. [Google Scholar] [CrossRef]
  249. McCann, M.J.; Gill, C.I.R.; McGlynn, H.; Rowland, I.R. Role of mammalian lignans in the prevention and treatment of prostate cancer. Nutr. Cancer 2005, 52, 1–14. [Google Scholar] [CrossRef] [PubMed]
  250. Ramsay, A.; Fliniaux, O.; Fang, J.; Molinie, R.; Roscher, A.; Grand, E.; Guillot, X.; Kovensky, J.; Fliniaux, M.A.; Schneider, S.; et al. Development of an NMR metabolomics-based tool for selection of flaxseed varieties. Metabolomics 2014, 10, 1258–1267. [Google Scholar] [CrossRef]
  251. El Sabagh, A.; Hossain, A.; Barutcular, C.; Gormus, O.; Ahmad, Z.; Hussain, S.; Islam, M.S.; Alharby, H.; Bamagoos, A.; Kumar, N.; et al. Effects of drought stress on the quality of major oilseed crops: Implications and possible mitigation strategies—A review. Appl. Ecol. Environ. Res. 2019, 17, 4019–4043. [Google Scholar] [CrossRef]
  252. Jakab, G.; Ton, J.; Flors, V.; Zimmerli, L.; Metraux, J.P.; Mauch-Mani, B. Enhancing Arabidopsis salt and drought stress synthase gene in transgenic tobacco plants: Pleiotropic phenotypes include drought tolerance. Planta 2005, 201, 293–297. [Google Scholar] [CrossRef] [Green Version]
  253. Macarisin, D.; Wisniewski, M.E.; Bassett, C.; Thannhauser, T.W. Proteomic analysis of β-aminobutyric acid priming and abscisic acid—Induction of drought resistance in crabapple (Malus pumila): Effect on general metabolism, the phenyl-propanoid pathway and cell wall enzymes. Plant Cell Environ. 2009, 32, 1612–1631. [Google Scholar] [CrossRef]
  254. Garg, A.K.; Kim, J.K.; Owens, T.G.; Ranwala, A.P.; Do Choi, Y.; Kochian, L.V.; Wu, R.J. Trehalose accumulation in rice plants confers high tolerance levels to different abiotic stresses. Proc. Natl. Acad. Sci. USA 2002, 99, 15898–15903. [Google Scholar] [CrossRef]
  255. Du, Y.L.; Wang, Z.Y.; Fan, J.W.; Turner, C.; Wang, T.; Li, F.M. B-aminobutyric acid increases abscisic acid accumulation and desiccation tolerance and decreases water use but fails to improve grain yield in two spring wheat cultivars under soil drying. J. Exp. Bot. 2012, 63, 4849–4860. [Google Scholar] [CrossRef]
  256. Bengtsson, T.; Weighill, D.; Proux-Wéra, E.; Levander, F.; Resjö, S.; Burra, D.D.; Moushib, L.I.; Hedley, P.E.; Liljeroth, E.; Jacobson, D.; et al. Proteomics and transcriptomics of the BABA-induced resistance response in potato using a novel functional annotation approach. BMC Genom. 2014, 15, 315. [Google Scholar] [CrossRef] [Green Version]
  257. Cortina, C.; Culianez-Macia, F.A. Tomato abiotic stress enhanced tolerance by trehalose biosynthesis. Plant Sci. 2005, 169, 75–82. [Google Scholar] [CrossRef]
  258. Jakab, G.; Cottier, V.; Toquin, V.; Rigoli, G.; Zimmerli, L.; Metraux, J.P.; Mauch-Mani, B. β-Aminobutyric acid-induced resistance in plants. Eur. J. Plant Pathol. 2001, 107, 29–37. [Google Scholar] [CrossRef]
  259. Singh, P.; Wu, C.C.; Zimmerli, L. β-Aminobutyric acid priming by stress imprinting. Plant Signal. Behav. 2010, 5, 878–880. [Google Scholar] [CrossRef] [Green Version]
  260. Quéro, A.; Fliniaux, O.; Elboutachfaiti, R.; Petit, E.; Guillot, X.; Hawkins, S.; Courtois, J.; Mesnard, F. Β-aminobutyric acid increases drought tolerance and reorganizes solute content and water homeostasis in flax (Linum usitatissimum). Metabolomics 2015, 11, 1363–1375. [Google Scholar] [CrossRef]
  261. Qayyum, M.A.; Bashir, F.; Maqbool, M.M.; Ali, A.; Bashir, S.; Abbas, Q. Implications of saline water irrigation for linseed on seed germination, seedling survival and growth potential. Sarhad J. Agric. 2019, 35, 1289–1297. [Google Scholar] [CrossRef]
  262. Cha-um, S.; Supaibulwatana, K.; Kirdmanee, C. Water relation, photosynthetic ability and growth of Thai jasmine rice (Oryza sativa L. ssp. indica cv. KDML105) to salt stress by application of exogenous glycine betaine and choline. J. Agron. Crop Sci. 2006, 192, 25–36. [Google Scholar] [CrossRef]
  263. Naz, T.; Akhtar, J.; Iqbal, M.M.; Haq, M.A.U.; Saqib, M. Boron toxicity in salt-affected soils and effects on plants. In Soil Science: Agricultural and Environmental Prospectives; Hakeem, K.R., Akhtar, J., Sabir, M., Eds.; Springer: Cham, Switzerland, 2016; pp. 259–286. [Google Scholar]
  264. Hamade, K.; Fliniaux, O.; Fontaine, J.X.; Molinié, R.; OtogoNnang, E.; Bassard, S.; Guénin, S.; Gutierrez, L.; Lainé, E.; Hano, C.; et al. NMR and LC-MS-based metabolomics to study osmotic stress in lignan-deficient flax. Molecules 2021, 26, 767. [Google Scholar] [CrossRef]
  265. Vo KT, X.; Rahman, M.M.; Rahman, M.M.; Trinh KT, T.; Kim, S.T.; Jeon, J.S. Proteomics and metabolomics studies on the biotic stress responses of rice: An update. Rice 2021, 14, 30. [Google Scholar] [CrossRef]
  266. Gevaert, K.; Vandekerckhove, J. Gel-Free Proteomics: Methods and Protocols; Springer: Berlin, Germany, 2011. [Google Scholar]
  267. Baggerman, G.; Vierstraete, E.; De Loof, A.; Schoofs, L. Gel-based versus gel-free proteomics: A review. Comb. Chem. High Throughput Screen 2005, 8, 669–677. [Google Scholar] [CrossRef]
  268. Klubicová, K.; Danchenko, M.; Skultéty, L.; Berezhna, V.V.; Hricová, A.; Rashydov, N.M.; Hajduch, M. Agricultural recovery of a formerly radioactive area: II. Systematic proteomic characterization of flax seed development in the remediated Chernobyl area. J. Proteom. 2011, 74, 1378–1384. [Google Scholar] [CrossRef]
  269. Hradilová, J.; Řehulka, P.; Řehulková, H.; Vrbová, M.; Griga, M.; Brzobohatý, B. Comparative analysis of proteomic changes in contrasting flax cultivars upon cadmium exposure. Electrophoresis 2010, 31, 421–431. [Google Scholar] [CrossRef] [PubMed]
  270. Balbuena, T.S.; Salas, J.J.; Martinez-Force, E.; Garces, R.; Thelen, J.J. Proteome analysis of cold acclimation in sunflower. J. Proteome Res. 2011, 10, 2330–2346. [Google Scholar] [CrossRef] [PubMed]
  271. Alvarez, S.; Berla, B.M.; Sheffield, J.; Cahoon, R.E.; Jez, J.M.; Hicks, L.M. Comprehensive analysis of the Brassica juncea root proteome in response to cadmium exposure by complementary proteomic approaches. Proteomics 2009, 9, 2419–2431. [Google Scholar] [CrossRef] [PubMed]
  272. Wendelboe-Nelson, C.; Morris, P.C. Proteins linked to drought tolerance revealed by DIGE analysis of drought resistant and susceptible barley varieties. Proteomics 2012, 12, 3374–3385. [Google Scholar] [CrossRef]
  273. Pang, Q.; Chen, S.; Dai, S.; Chen, Y.; Wang, Y.; Yan, X. Comparative proteomics of salt tolerance in Arabidopsis thaliana and Thellungiella halophila. J. Proteome Res. 2010, 9, 2584–2599. [Google Scholar] [CrossRef]
  274. Kosová, K.; Vítámvás, P.; Prášil, I.T. Proteomics of stress responses in wheat and barley—Search for potential protein markers of stress tolerance. Front. Plant Sci. 2014, 5, 711. [Google Scholar] [CrossRef] [Green Version]
  275. Caruso, G.; Cavaliere, C.; Foglia, P.; Gubbiotti, R.; Samperi, R.; Laganà, A. Analysis of drought responsive proteins in wheat (Triticum durum) by 2D-PAGE and MALDI-TOF mass spectrometry. Plant Sci. 2009, 177, 570–576. [Google Scholar] [CrossRef]
  276. Caruso, G.; Cavaliere, C.; Guarino, C.; Gubbiotti, R.; Foglia, P.; Laganà, A. Identification of changes in Triticum durum L. leaf proteome in response to salt stress by two-dimensional electrophoresis and MALDI-TOF mass spectrometry. Anal. Bioanal. Chem. 2008, 391, 381–390. [Google Scholar] [CrossRef]
  277. Rinalducci, S.; Egidi, M.G.; Karimzadeh, G.; Jazii, F.R.; Zolla, L. Proteomic analysis of a spring wheat cultivar in response to prolonged cold stress. Electrophoresis 2011, 32, 1807–1818. [Google Scholar] [CrossRef]
  278. Majoul, T.; Bancel, E.; Triboi, E.; Ben Hamida, J.; Branlard, G. Proteomic analysis of the effect of heat stress on hexaploid wheat grain: Characterization of heat-responsive proteins from non-prolamins fraction. Proteomics 2004, 4, 505–513. [Google Scholar] [CrossRef]
  279. Li, H.; Yang, M.; Zhao, C.; Wang, Y.; Zhang, R. Physiological and proteomic analyses revealed the response mechanisms of two different drought-resistant maize varieties. BMC Plant Biol. 2021, 21, 513. [Google Scholar] [CrossRef]
  280. Halder, T.; Choudhary, M.; Liu, H.; Chen, Y.; Yan, G.; Siddique, K. Wheat proteomics for abiotic stress tolerance and root system architecture: Current status and future prospects. Proteomes 2022, 10, 17. [Google Scholar] [CrossRef] [PubMed]
  281. Melnikova, N.V.; Belenikin, M.S.; Bolsheva, N.L.; Dmitriev, A.A.; Speranskaya, A.S.; Krinitsina, A.A.; Nadezda, V.K.; Samatadze, T.E.; Amosova, A.V.; Muravenko, O.V.; et al. Flax inorganic phosphate deficiency responsive miRNAs. J. Agric. Sci. 2014, 6, 156–160. [Google Scholar] [CrossRef] [Green Version]
  282. Barvkar, V.T.; Pardeshi, V.C.; Kale, S.M.; Qiu, S.Q.; Rollins, M.; Datla, R.; Gupta, V.S.; Kadoo, N.Y. Genome-wide identification and characterization of microRNA genes and their targets in flax (Linum usitatissimum). Planta 2013, 237, 1149–1161. [Google Scholar] [CrossRef]
  283. Young, L.W. High Temperature Stress and Flowering in Brassica napus L. Ph.D. Thesis, University of Saskatchewan, Saskatoon, SK, Canada, 2003. [Google Scholar]
  284. Dmitriev, A.A.; Kudryavtseva, A.V.; Krasnov, G.S.; Koroban, N.V.; Speranskaya, A.S.; Krinitsina, A.A.; Belenikin, M.S.; Snezhkina, A.V.; Sadritdinova, A.F.; Kishlyan, N.V.; et al. Gene expression profiling of flax (Linum usitatissimum L.) under edaphic stress. BMC Plant Biol. 2016, 16, 237. [Google Scholar] [CrossRef] [Green Version]
  285. Hano, C.; Addi, M.; Fliniaux, O.; Bensaddek, L.; Duverger, E.; Mesnard, F.; Lamblin, F.; Laine, E. Molecular characterization of cell death induced by a compatible interaction between Fusarium oxysporum f. sp. linii and flax (Linum usitatissimum) cells. Plant Physiol. Biochem. 2008, 46, 590–600. [Google Scholar] [CrossRef]
  286. Fujita, Y.; Venterink, H.O.; Van Bodegom, P.M.; Douma, J.C.; Heil, G.W.; Hölzel, N.; Jablonska, E.; Kotowski, W.; Okruszko, T.; Pawlikowski, P.; et al. Low investment in sexual reproduction threatens plants adapted to phosphorus limitation. Nature 2013, 505, 82–86. [Google Scholar] [CrossRef] [PubMed]
  287. Wu, J.W.; Shi, Y.; Zhu, Y.X.; Wang, Y.C.; Gong, H.J. Mechanisms of enhanced heavy metal tolerance in plants by silicon, a review. Pedosphere 2013, 23, 815–825. [Google Scholar] [CrossRef]
  288. Guo, P.; Baum, M.; Grando, S.; Ceccarelli, S.; Bai, G.; Li, R.; von Korff, M.; Varshney, R.K.; Graner, A.; Valkoun, J. Differentially expressed genes between drought-tolerant and drought-sensitive barley genotypes in response to drought stress during the reproductive stage. J. Exp. Bot. 2009, 60, 3531–3544. [Google Scholar] [CrossRef]
  289. Shi, D.C.; Wang, D. Effects of various salt-alkaline mixed stresses on Aneurolepidium chinense (Trin.) Kitag. Plant Soil 2005, 271, 15–26. [Google Scholar] [CrossRef]
  290. Guo, B.; Fedorova, N.D.; Chen, X.; Wan, C.-H.; Wang, W.; Nierman, W.C.; Bhatnagar, D.; Yu, J. Gene Expression Profiling and Identification of Resistance Genes to Aspergillus flavus Infection in Peanut through EST and Microarray Strategies. Toxins 2011, 3, 737–753. [Google Scholar] [CrossRef] [PubMed]
  291. Guo, R.; Hao, W.; Gong, D.Z. Effects of water stress on germination and growth of linseed seedlings (Linum usitatissimum L.), photosynthetic efficiency and accumulation of metabolites. J. Agric. Sci. 2012, 4, 253–265. [Google Scholar] [CrossRef] [Green Version]
  292. Guo, R.; Zhou, J.; Hao, W.; Gu, F.; Liu, Q.; Li, H.; Xia, X.; Mao, L. Germination, growth, chlorophyll fluorescence and ionic balance in linseed seedlings subjected to saline and alkaline stresses. Plant Prod. Sci. 2014, 17, 20–31. [Google Scholar] [CrossRef] [Green Version]
  293. White, J.W.; Andrade-Sanchez, P.; Gore, M.A.; Bronson, K.F.; Coffelt, T.A.; Conley, M.M.; Feldmann, K.A.; French, A.N.; Heun, J.T.; Hunsaker, D.J.; et al. Field-based phenomics for plant genetics research. Field Crops Res. 2012, 133, 101–112. [Google Scholar] [CrossRef]
  294. Pratap, A.; Gupta, S.; Nair, R.M.; Gupta, S.K.; Schafleitner, R.; Basu, P.S.; Singh, C.M.; Prajapati, U.; Gupta, A.K.; Nayyar, H.; et al. Using Plant Phenomics to Exploit the Gains of Genomics. Agronomy 2019, 9, 126. [Google Scholar] [CrossRef] [Green Version]
  295. Walter, A.; Liebisch, F.; Hund, A. Plant phenotyping: From bean weighing to image analysis. Plant Methods 2015, 11, 14. [Google Scholar] [CrossRef] [Green Version]
  296. Yang, W.; Doonan, J.H.; Hawkesford, M.J.; Pridmore, T.; Zhou, J. Editorial: State-of-the-Art Technology and Applications in Crop Phenomics. Front. Plant Sci. 2021, 12, 767324. [Google Scholar] [CrossRef]
  297. Tao, H.; Xu, S.; Tian, Y.; Li, Z.; Ge, Y.; Zhang, J.; Wang, Y.; Zhou, G.; Deng, X.; Zhang, Z.; et al. Proximal and remote sensing in plant phenomics: 20 years of progress, challenges, and perspectives. Plant Commun. 2022, 3, 100344. [Google Scholar] [CrossRef]
  298. Bekkering, C.S.; Huang, J.; Tian, L. Image-Based, Organ-Level Plant Phenotyping for Wheat Improvement. Agronomy 2020, 10, 1287. [Google Scholar] [CrossRef]
  299. Sirault, X.R.R.; James, R.A.; Furbank, R.T. A new screening method for osmotic component of salinity tolerance in cereals using infrared thermography. Funct. Plant Biol. 2009, 36, 970–977. [Google Scholar] [CrossRef]
  300. Jordan, M.C.; McHughen, A. Glyphosate tolerant flax plants from mediated gene transfer. Plant Cell Rep. 1988, 7, 281–284. [Google Scholar] [CrossRef] [PubMed]
  301. McHughen, A.; Holm, F. Development and preliminary field testing of a glufosinate-ammonium tolerant transgenic flax. Can. J. Plant Sci. 1995, 75, 117–120. [Google Scholar] [CrossRef]
  302. Haughn, G.W.; Smith, J.; Mazur, B.; Somerville, C. Transformation with a mutant Arabidopsis acetolactate synthase gene renders tobacco resistant to sulfonylurea herbicides. Mol. Gen. Genet. 1988, 211, 266–271. [Google Scholar] [CrossRef]
  303. Shinozaki, K.; Yamaguchi-Shinozaki, K.; Seki, M. Regulatory network of gene expression in the drought and cold stress responses. Curr. Opin. Plant Biol. 2003, 6, 410–417. [Google Scholar] [CrossRef]
  304. Wang, H.; Wang, H.; Shao, H.; Tang, X. Recent advances in utilizing transcription factors to improve plant abiotic stress tolerance by transgenic technology. Front. Plant Sci. 2016, 7, 67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Deshmukh, R.K.; Sonah, H.; Bélanger, R.R. Plant Aquaporins: Genome-Wide Identification, Transcriptomics, Proteomics, and Advanced Analytical Tools. Front. Plant Sci. 2016, 7, 1896. [Google Scholar] [CrossRef] [Green Version]
  306. Singh, S.; Bhatt, V.; Kumar, V.; Kumawat, S.; Khatri, P.; Singla, P.; Shivaraj, S.M.; Nadaf, A.; Deshmukh, R.; Sharma, T.R.; et al. Evolutionary Understanding of Aquaporin Transport System in the Basal Eudicot Model Species Aquilegia coerulea. Plants 2020, 9, 799. [Google Scholar] [CrossRef] [PubMed]
  307. Patel, J.; Mishra, A. Plant aquaporins alleviate drought tolerance in plants by modulating cellular biochemistry, root-architecture, and photosynthesis. Physiol. Plant. 2021, 172, 1030–1044. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of abiotic stress regulation mechanisms by plants. The orange boxes represent coping mechanisms, while grey boxes represent disturbances in plant physiological response.
Figure 1. Schematic representation of abiotic stress regulation mechanisms by plants. The orange boxes represent coping mechanisms, while grey boxes represent disturbances in plant physiological response.
Genes 14 01461 g001
Figure 2. Disease cycle of rust in linseed. In this infectious stage, urediniospores are present on the main plant (host plant) and later transferred through teliospores, avoiding harsh, unfavourable conditions by remaining dormant on alternate hosts as basidiospores.
Figure 2. Disease cycle of rust in linseed. In this infectious stage, urediniospores are present on the main plant (host plant) and later transferred through teliospores, avoiding harsh, unfavourable conditions by remaining dormant on alternate hosts as basidiospores.
Genes 14 01461 g002
Figure 3. An overview of the use of omics approach, which depicts how the information is passed on from genomics to transcriptomics, then to proteomics, and finally to metabolomics, which bring about the final change in response to a stress or signal.
Figure 3. An overview of the use of omics approach, which depicts how the information is passed on from genomics to transcriptomics, then to proteomics, and finally to metabolomics, which bring about the final change in response to a stress or signal.
Genes 14 01461 g003
Figure 4. Diagrammatic illustration of the importance of genomics in the production of high-yielding linseed cultivars.
Figure 4. Diagrammatic illustration of the importance of genomics in the production of high-yielding linseed cultivars.
Genes 14 01461 g004
Figure 5. Schematic representation of pathways that are up- and downregulated on encounter with various types of stress and respond by changing metabolite concentrations.
Figure 5. Schematic representation of pathways that are up- and downregulated on encounter with various types of stress and respond by changing metabolite concentrations.
Genes 14 01461 g005
Table 1. Taxa of Linum conserved in Canada’s plant genetic resources.
Table 1. Taxa of Linum conserved in Canada’s plant genetic resources.
SpeciesNo. of AccessionsLife Form2nOrigin
L. campanulatum L.1Perennial28West Mediterranean
L. austriacum L.3Perennial18Asia, Siberia, Central and East Europe
L. bienne Mill.11Winter-annual30Mediterranean, West Europe, West Asia
L. decumbens Desf.1Annual30South Europe
L. capitatum1Perennial34Balkan
L. altaicum Ledeb.2Perennial18West Siberia
L. leonii1Perennial18France, Germany
L. flavum L.4Perennial28 (30)Caucasus, South and Central Europe
L. trigynum L.2Annual20Mediterranean, South and Central Europe, West Asia
L. hirsutum L.1Perennial16Central and East Europe, West Asia
L. narbonense L.3Perennial18 (20)Mediterranean
L. rigidum Pursh1Perennial20North America
L. lewisii Pursh11Perennial18North America
L. pallescens Bunge1Perennial30West Siberia
L. tenuifolium L.1Perennial16Europe, West Asia, Mediterranean,
L. perenne L.10Perennial18Siberia, West Asia, Central and East Europe
L. strictum L.3Annual18 (30, 22)Mediterranean, West Asia, South Europe,
L. grandiflorum Desf.7Annual16Algeria
Table 2. List of varieties grown under different climate regimes in India.
Table 2. List of varieties grown under different climate regimes in India.
Varietal CharactersName of Variety
Seed typeJLS95, JLS66, LSL93, RLC 148, Pratap 2, T 397, JLS27, PKV-NL 260, Kota Barani Alsi 4, JL41, JLS67, JLS73, Arpita, Padmini, Divya, Priyam, Indu (LCK 1108), Rajan, Surya, TL 99, Kota Barani Alsi 5, Kota Barani Alsi 6, Jawahar Linseed 165, RLC 164, RLC 167, RLC 171, Sabour Tisi-2, Sabour Tisi-3, Birsa Tisi-1, LCK 1611, JLS122, SHA5
Irrigated conditionsPratap 2, JLS27, Kota Alsi 6, Priyam, Indu (LCK 1108), Rajan, Surya, TL 99, SHA5, LCK 1611, JLS95, JLS66
Rainfed conditionJLS95, JLS66, LSL93, RLC 148, T 397, Sharda, Padmini, KV-NL 260, Kota Barani Alsi 4, JL 41, JLS67, JLS73, Arpita, Padmini, Divya, JLS122
Utera situationRLC 143, RLC 153, Sabour Tisi-2, Sabour Tisi-3, R 552
Cultivars with dual purpose (seed + fibre)Him Alsi-2, Nagarkot, Jeevan, RLU-6, Rashmi, Parvati, Meera, Shikha, Tiara (JRF 2)
Table 3. List of linseed varieties that are tolerant to various abiotic and biotic stresses.
Table 3. List of linseed varieties that are tolerant to various abiotic and biotic stresses.
S. No.Name of VarietyYear of Release and Notification No.PedigreeOriginating CentreDurationAvg. Yield kg/haOil Content (%)Recommended StatesSpecial Features
1K-21975
44 E, 21 August 1975
Rust-resistant strain from UP x Kangra localLudhiana
(Punjab)
170–1751110(I)40.04Punjab, Haryana, and Uttar PradeshWhite flower, dark brown, resistant to wilt and rust
2LC-1851975
44 E, 21 August 1975
NP(RR)-37x Kangra localLudhiana
(Punjab)
150–160500(U)38.89PunjabBlue flower, resistant to powdery mildew and rust
3HIRA1978
429 E, 19 December 1978
H 342 x NP(RR)-9Kanpur
(U.P.)
130–1351200(R)36.36Bundelkhand area and U.P.White flower, resistant to wilt and rust
4MUKTA1978
429 E, 19 December 1978
H 626 x NP(RR)-9Kanpur
(U.P.)
127–1321200(I)41.40Eastern U.P.Large white flower, medium- and brown-seeded, resistant to wilt and rust
5CHAMBAL1978
13 E, 19 December 1978
Local x RR45Kota
Rajasthan
125–130900(I)40.11RajasthanBlue flower, brown- and bold-seeded, moderately resistant to rust, wilt, and powdery mildew
6NEELUM1978
13 E, 19 December 1978
T-1 x NP(RR)-9Kanpur
(U.P.)
140–1451500(I)43.00Mid Central and Western Uttar PradeshBlue flower, brown- and bold-seeded, tolerant to wilt and rust
7JAWAHAR-11982
19 E, 14 January 1982
Selection of KP 29Jabalpur
(M.P.)
116–120900(R)38.34Madhya PradeshRed-violet flower, dark-brown- and bold-seeded, resistant to rust
8JAWAHAR-7 (JLS-1)1982
19 E, 14 January 1982
Selection of No. 55Raipur
(C.G.)
116–120700(R), 300(U)37.79Madhya PradeshRed-violet flower, dark-brown and bold-seeded, resistant to rust
9JAWAHAR-17 (JLS-7)1982
19 E, 14 January 1982
Selection of No. 55Raipur
(C.G.)
117–1201300(I), 800(R)37.61Madhya PradeshRed-violet flower, dark-brown- and bold-seeded, resistant to rust
10NEELA1982
19 E, 14 January 1982
SPS from indigenous collectionBehrampore (W.B.)127–135850(R)34.93West BengalBlue flower, brown- and medium-seeded, resistant to bud fly
11LC-541982
19 E, 14 January 1982
K2 x Kangra localLudhiana
(Punjab)
155–1651330(I)42.00Punjab, H.P. and HaryanaWhite flower, brown- and medium-seeded resistant to wilt and rust
12C-4291983
429 E, 3 January 1983
No.3 x IP-135Nagpur
(M.S.)
125–1301000(R)39.07MaharashtraRed-violet flower, dark-brown- andsmall-seeded, resistant to rust
13T-3971984
596 E, 13 August 1984
T-491 x T-1103-1Kanpur
(U.P.)
120–1251100(I)44.00Bundelkhand of U.P., Bihar, Assam, M.P. and RajasthanViolet flower, dark-brown-seeded, tolerant to rust, wilt, and drought
14JAWAHAR-552 (R-552)1984
596 E, 13 August 1984
Sel. Number 55x B-67Raipur
(C.G.)
118–125900(R)44.00Madhya PradeshViolet flower, light-brown-seeded, medium toleranceto rust, wilt, and powdery mildew
15PUSA-2 1985
295 E, 9 April 1985
Selection of BS 12New Delhi125–150730(R)38.31Punjab, H.P., Haryana and RajasthanWhite flower, brown-seeded, medium height, resistant to rust
16PUSA-31985
295 E, 9 April 1985
K2 x T-603New Delhi125–150800(I)37.65Punjab, H.P., Haryana and RajasthanWhite flower, brown-seeded, medium height, resistant to rust
17S-36-Selection of Local SolapurPune (MH)130–135400(R)34.80KarnatakaBlue flower, light-brown- and small-seeded
18HIMALINI1985 295 E, 9 April 1985K2 x Kangra localPalampur
(H.P.)
150–1751310(I)42.00Punjab, H.P., Haryana and adjoining of RajasthanWhite flower, brown-seeded, medium height, resistant to rust and wilt
19JAWAHAR-231985
540 E, 24 July 1985
EC 9832 x HiraJabalpur
(MP)
120–1301000(I)43.00M.P., Odisha, Raj.
Bundelkhand of U.P., M.S., Karnataka
White flower, brown- andmedium-seeded, resistant to wilt and rust and powdery mildew
20GARIMA1985
540 E, 24 July 1985
T-126 x NeelumKanpur
(UP)
120–1301490(I)42.00U.P. (Excl. Bundelkhand), Bihar, W.B. and AssamBlue flower, medium height, brown-seeded, tolerant to PM, altenararia blight, and wilt, resistant to rust
21SWETA1985
540 E, 24 July 1985
Mukta x T-1206Kanpur (UP)130–135880(R)44.00U.P. (Excl. Bundelkhand), Bihar, W.B. and AssamBlue flower, mediumheight, brown-seeded, tolerant to PM, altenararia blight, and wilt, resistant to rust
22SHUBHRA1985
540 E, 24 July 1985
Mukta x K 2Kanpur
(UP)
130–1351390(I), 870 (R)45.00U.P. (Excl. Bundelkhand), Bihar, W.B and AssamDark-brown-seeded, medium, white flower, resistant to rust
23LAXMI-271987
165 E,
6 March 1987
Neelum/R1//Neelum x NPRR-9//
Neelum x R1//
Neelum x Afg-8
Mauranipur (UP)110–1201260(I),1020(R)45.00Bundelkhand of U.P.Red-violet flower, dark-brown-seeded medium height, resistant to rust, wilt, and PM
24GAURAV1987
834 E, 18 September 1987
Sel.3 x EC-1552Kanpur
(UP)
137–1401050 (S)
950(F)
43.00Assam, U.P. (Excl. Bundelkhand), Bihar, West BengalViolet-red flower, brown- and medium-seeded, resistant to wilt and rust
25KIRAN1988
10 E, 1 January 1988
(R1x Afg-8) x R1Raipur
(C.G.)
120–126750(R)43.00M.P., Odisha, Bundelkhand of U.P., M.S., Karnataka, and RajasthanViolet-red flower, brown- and medium-seeded, resistant to wilt, powdery mildew, and rust
26JANKI1988New River x LC 216Palampur
(H.P.)
165–1701200(I)42.00Himachal PradeshViolet-red flower, brown- and medium-seeded, resistant to wilt, powdery mildew, and rust
27JEEVAN1988
10 E, 1 January 1988
Summit x LC-216Palampur
(H.P.)
175–1801090(S 1100(F)45.00Punjab and Himachal PradeshViolet-red flower, brown- and medium-seeded, resistant to wilt, powdery mildew, and rust
28SURABHI1995
408 E, 4 May 1995
LC 216 x LC 185Palampur
(H.P.)
165–1701000(U)44.00Himachal PradeshWhite flower, yellow- and small-seeded, resistant to wilt, powdery mildew, and rust
29NAGARKOT1995
400 E, 4 May 1995
New river x LC 216Palampur
(H.P.)
165–1701150 (S)
950 (F)
43.00Panjab, U.P., H.P., W.B., Assam, Haryana and Raj.Blue flower, brown- and medium-seeded, resistant to wilt, powdery mildew, and rust
30SHIKHA1997
401 E, 15 May 1998
Hira xCristaKanpur
(U.P.)
135–1401233 (S)
1033 (F)
42.00U.P. (Excl. Bundelkhand), Bihar, W.B. and Assam Blue flower, brown- and medium-seeded, resistant to wilt and rust
31LC 20231998Flake x LC 54Ludhiana
(Punjab)
158–1631294 (I)38.8Punjab-
32PADMINI1998
1050 E, 26 October 1999
(EC 41,628 x EC 77959) x (DPL 20 x Neelum)Manranipur (U.P.)120–125943 (R)43.00M.P., Rajasthan, Bundelkhand of U.P., Maharashtra, Odissa and KarnatakaBlue flower, brown-seeded, resistant to powdery mildew
33JAWAHAR LINSEED SAGAR-9
(JLS 9)
1998
425 E, 8 June 1999
(RL-102 x R7) x J 23Sagar
(M.P.)
115–1251250 (I), 1000 (R)42.00Madhya PradeshWhite flower, mediumheight, brown- and bold-seeded, resistant to wilt, powdery mildew, and rust
34RASHMI1999
1050 E, 26 October 1999
Gaurav x JankiKanpur
(U.P.)
135–1401003 (S), 719 (F)41.00U.P., Bihar, W.B. and Rajasthan Very light blue flower, brown-seeded, resistant to wilt, powdery mildew, and rust
35MEERA 2000
340 E, 3 April 2000
(RL75-6-2 x RL 29-8) x LCK 8528Kota
(Raj.)
135–1401439 (S), 1011 (F)42.00U.P., Bihar, W.B. and RajasthanViolet-blue flower, brown-seeded, resistant to wilt, powdery mildew, and rust
36RL-9142000
973 E, 4 September 2002
(RR 9 x R-93) x (Flake 1 x LC 54)Kota
(Raj.)
135–1401617(I)41.10Kota command area of RajasthanViolet-blue flower, medium height, brown-seeded, resistant to wilt, powdery mildew, and rust
37PARVATI2001
92 E, 2 February 2001
EC 41628/
EC77959//DPL20 x Neelam///EC216 x Hira//BR 1x NP 440
Manranipur (U.P.)140–1461600 (S), 1020 (F)42.00U.P., Bihar, W.B. and Rajasthan Blue flower, medium height, brown-seeded, resistant to powdery mildew and rust
38SHEELA2001
92 E, 2 February 2001
Gaurav x JankiKanpur
(U.P.)
155–1601379 (R)41.00Himachal Pradesh, Panjab, Haryana, and Jammu and KashmirBlue flower, shining-brown-seeded, resistant to wilt, powdery mildew, and rust
39SHEKHAR2001
92 E, 2 February 2001
Laxmi 27 x EC 1387Kanpur
(U.P.)
135–1401555(I), 920 (R)43.00U.P. (Excl. Bundelkhand), Bihar, W.B. and AssamViolet-blue flower, shining-brown-seeded, moderately resistant to bud fly, alternaria blight, powdery mildew, and rust
40NL-972001
92 E, 2 February 2001
R-7 x RLC 4Nagpur
(M.H.)
115–120641 (R)42.00Bidarbha RegionBlue flower, brown- and medium-seeded
41JAWAHAR LINSEED SAGAR-27
(SUYOG)
2002
161 E, 4 February 2004
(Kran x KL-168) x KiranSagar
(M.P.)
118–1251509 (I)41.43Raj. Bundelkhand of U.P., M.P., C.G., M.H., Odissa, A.P., and KarnatakaWhite flower, light-brown-seeded, resistant to rust, powdery mildew, and bud fly
42BINWA (KL-210)2004
122 E, 2 February 2005
Flake 1 x SPS-47/7-10-3Palampur
(H.P.)
179–186858 (I)40.00Haryana, Panjab, Himachal Pradesh, and J&KBlue flower, yellow-seeded, resistant to wilt and rust
43BANER (KL-224)2005
1177 E, 25 August 2005
EC 21741 x LC 216Palampur
(H.P.)
171–203511 (U)39.70 Haryana, Panjab, Himachal Pradesh, and J&KPurple flower, brown-coloured seeds, resistant to rust
44INDIRA ALSI-32 (RLC-76)2005
1566 E, 5 November 2005
Kiran x RLC 29Raipur
(C.G.)
110–115780(R)39.18CG, Maharashtra, Karnataka, and OdishaBlue flower, dark-brown-seeded, powdery mildew
45KARTIKA (RLC-76)2005
1566 E,
5 November 2005
Kiran x LCK-88062Raipur
(C.G.)
103–1081078(R)42.93Rainfed areas of ChhattisgarhBlue flower, light-brown and medium-seeded, resistant to bud fly and powdery mildew
46DEEPIKA (RLC 78)2006
1178 E, 20 July 2007
Kiran x AyogiRaipur
(C.G.)
110–1151272 (SI&U)41.39Partially irrigated as well as Utera Situation of CGBlue flower, brown-seeded, resistant to powdery mildew
47SHARDA (LMS-4-27)2006
1572 E, 20 September 2006
(Shubhra x J1) x (J1 x Kiran)Manranipur (C.G.)100–105762 (R)41.32CG, Maharashtra, Karnataka, AP and OdissaWhite flower, brown-seeded, MR to wilt, powdery mildew, and bud fly
48PRATAP ALSI-1 (RLU-6)2007
1703 E, 5October 2007
Acc. 750 x RL 29-8Kota
(Raj.)
129–1351997 (S) 834 (F)41.08Rajasthan Kota command areaWhite flower, brown-seeded, MR to rust, wilt, and bud fly
49LC-20632007
1108 E, 8 May 2008
1509 x LC 54Ludhiana
(Punjab)
115–1251200 (I)37.40Irrigated area of Punjab StateBlue flower, dark-brown-seeded, MR to bud fly
50HIMANI KL-2142008
2458 E, 16 October 2008
DPL 20 x KLS-1Palampur
(H.P.)
177–200583 (U)36.40HP, PB, Haryana and J&KBlue flower, small- andbrown-seeded, MR to powdery mildew and rust
51AZAD ALSI-1 (LMS 9-2K)2008
2458 E, 16 October 2008
RL 904 x KiranManranipur (U.P.)125–1301610 (I)39.92BKD area of UP, MP and Rajasthan Red violet, MR to wilt, PM blue flower, dark-brown-seeded, resistant to bud fly and rust
52RLC 922008
2458 E, 16 October 2008
Jeevan x LCK 9209Raipur
(C.G.)
1111196 (I)37.70CG, Maharashtra, Karnataka, AP, and OdissaTinyblue flower, light-violet flower, brown- and medium-seeded, moderately resistant to powdery mildew, rust, and bud fly
53SHIVAL (SLS-67)2010
733 E, 1 April 2010
LCK 9610 x LMS 127Sagar
(M.P.)
108–1101252 (R)40.16BKD area of UP, MP, and RajasthanWhite flower, light-brown-seeded, MR to powdery mildew and rust
54JAWAHAR LINSEED-41 2011
2326 E, 10 October2011
Kiran x Acc. No.443Hoshangabad (M.P.)115–1201600 (I)40.00Area of MP state with limited facilityWhite flower, brown- and bold-seeded, resistant to rust, powdery mildew, and rust
55 BHAGSU (KL-215)2010
2137 E, 31 August 2010
RL-50-3 x SurbhiPalampur
(H.P.)
175–201428(U)36.38Himachal Pradesh, J&K, Uttaranchal, Punjab, and HaryanaBlue flower, brown- and small-seeded, MR to rust
56RUCHI (LCK 5021)2011
283 E, 7 February 2011
Garima x LCK 88062Kanpur
(U.P.)
1341366 (S), 1055 (F)39.84UP. (Excl. Bundelkhand), Bihar, WB and AssamBlue flower, shining-brown-seeded, MR to rust, powdery mildew, and bud fly
57JAWAHAR LINSEED SAGAR-73
(JLS 73)
2011
632 E,
25 March 2011
Padmini x Laxmi 27Sagar
(M.P.)
111–1141090 (R)38.82Bundelkh and region of UP, MP, RajasthanLarge blue flower, bold-seeded, high omega 3(58.20), high oil content (42%), light-brown-seeded, resistant to rust, powdery mildew, and bud fly
58MAU AZAD
ALSI-2
(LMS 149-4)
2011
2326 E,
10 October 2011
KL 178 x HiraManranipur (U.P.)105–110815 (R)40.17CG, Maharashtra,
Karnataka, AP, Odisha
Medium white flower, medium- and brown-seeded, resistant to rust
59NDL 2004-052011Garima x RL 993Faizabaad
(U.P.)
125–1301800(I)40.0Eastern Uttar PradeshLight flower, medium- and dark-brown-seeded, resistant to Alternaria blight, rust, and powdery mildew
60NDL 20022011Garima x EC 44Faizabaad
(U.P.)
130–1351600 (I & R)39.3Eastern Uttar PradeshLight-blue flower, medium- and dark-brown-seeded, resistant to rust and powdery mildew and tolerant to bud fly
61Pratap Alsi 2
(RL 26016)
2012
268 E,
28 January 2015
RL 914 x NL 93Kota
(Raj.)
129–1351957 (I)41.81Rajasthan Blue flower, shining-brown- and bold-seeded, powdery mildew, wilt, and bud fly. Moderately resistant to Alternaria blight, powdery mildew, wilt, and bud fly
62PKVNL-2602015
112 E,
12 January 2015
R 552 x RLC6Nagpur
(M.S.)
102–106963 (R)37.67MaharashtraLight flower, brown-seeded, moderately resistant to powdery mildew and bud fly
63Kota Barani Alsi-3
(RL-292002)
2015RL 903 x AyogiKota
(Raj.)
119–1241370 (R)38.73RajasthanBlue flower, shining dark-brown-seeded, resistant to rust and moderately resistant to altarnaria blight, powdery mildew, and bud fly
64Kota Barani Alsi-4
(RL-10193)
2015Triveni x RL-1011Kota
(Raj.)
120–1261100 (R)40.37UP, MP, RajasthanWhite flower, shining dark-brown-seeded, moderately resistant to altarnaria blight and powdery mildew
65ChattisgadhAlsi-1
(RLC-133)
2015NL-14 x ACC-26Raipur
(C.G.)
110–113844 (R)36.4CGWhite flower, early, moderately resistant to bud fly, altarnaria blight, powdery mildew, and rust
66Tiara
(JRF-2)
2015
1228 E
FT-889 x FT-895CRIJEF (W.B.)135–1501290 H.P., Uttarakhand. Sikkim, A.P., U.P, Kashmir, and W.B.White flower, tall, suitable for fibre yield
67Arpita (OL 98-13-1)2016
3540 E,
22 November 2016
RLC 29 x R 1871Keonjhar
(Odisha)
102–106849 (R)35.67OdishaBlue flower, light-brown-seeded, resistant to wilt and powdery mildew
68UteraAlsi
(RLC-143)
2016LCK-88062 x T-397Raipur
(C.G.)
115–118570 (U)34.1C.G., M.P., U.P., Jharkhand, Maharashtra, and OdishaSuitable for Utera/para, blue flower, resistant to powdery mildew and rust, moderately resistant to budfly and altarnaria blight
69Jawahar Linseed Sagar-79 (JLS-79)2016
3540 E, 22 November 2016
Padmini x Laxmi-27Sagar
(M.P.)
1762(I)40.35Madhya PradeshBlue flower, suitable for irrigated conditions, moderately tolerant to powdery mildew, wilt, Alternaria blight, and bud fly
70PKV-NL-260
(NL-260)
2016
3540 E, 2 November 2016
R552 x RLC-6Nagpur
(M.S.)
102–106940(R)35.56MaharashtraBlue flower, suitable for rainfed conditions, moderately tolerant to powdery mildew, wilt, Alternaria blight, and bud fly
71DIVYA(BAU-06-03)2016
3540 E,
22 November 2016
BAU-1008 x KiranKanke
(J.K.)
1801538(I)39.80Punjab, Haryana, Himachal Pradesh, and UttarakhandWhite flower, resistant to rust and wilt
72INDU
(LCK 1108)
2016
1007 E, 30 March 2017
LMS1-2Kx
LMS4-27
Kanpur
(U.P.)
133–137955 (I)40.00Whole of U.P.Blue flower, bold- and brown-seeded, highly resistant to rust, resistant to alternaria blight and powdery mildew.
73UMA
(LCK-1101)
2016
1007 E,
30 March 2017
Polf 34 x
Padmini
Kanpur
(U.P.)
120–123868 (R)37.60Whole of U.P.Blue flower, medium- and light-brown-seeded, moderately resistant to rust, wilt, and Alternaria blight
74VARSHA ALSI
(RLC-148)
2017SIKO 10 x KiranRaipur
(C.G.)
110–1141033 (R)35.70CG, MP, UP, JH, M.S., and OdishaBlue flower, suitable for rainfed situation, moderately resistant to budfly, moderately susceptible to wilt and powdery mildew
75.PRIYAM
(BAU-2012-1)
2017
2805 E,
25 August 2017
BAU-2k-14 x GarimaRanchi (Jharkhand)1731151(R)37.15Punjab, Haryana,
H.P., and
Uttarakhand.
White flower, dark brown seed
76.JAWAHAR
LINSEED
SAGAR-95
(JLS-95)
2018
1379 E,
27 March 2018
JLS-27x GS-281Sagar (M.P.)1331240(R)39.40Bundelkhand part of Uttar Pradesh, Raj.
Madhya Pradesh, Karnataka, Odisha,
Maharashtra, Chhattisgarh
White flower, bold-seeded, high omega 3 (53.50), high oil content (40%), suitable for rainfed and mechanical harvesting, resistant to rust, moderately resistant to wilt and powdery mildew, alternaria blight, and budfly
77JAWAHAR LINSEED SAGAR-66 (JLS 66) 2018
399 E,
24 January 2018
Selection from Accn. 2511 Sagar
(MP)
1141200 (R)40.50Madhya Pradesh Blue flower, suitable for rainfed conditions, bold-seeded, high omega 3 (55.96), high oil content (42%), resistant to rust, moderately resistant to wilt, powdery mildew, Alternaria blight, and budfly
78SABOUR TISI-1
(BAUP-101)
2018
1379 E,
27 March 2018
Selection from Kishanganj localPatna
(Bihar)
124453(U)32.68U.P., excluding Bundelkhand,
Bihar, Jharkhand, W.B, and Assam
Blue flower, suitable for Utera,
moderately resistant to rust and powdery mildew
80HIM PALAM ALSI-2
(KL-263)
2018
1379 E,
27 March 2018
KL-223 x KL-224Palampur 1301633(I)35.58H.P., Haryana, and Jammu
and Kashmir
Blue flower, suitable for rainfed conditions,
moderately resistant to rust and powdery mildew
81RAJAN2018
1498 E,
1 April 2019
Gaurav x GS 234Kanpur130–1351528(S)
1093(F)
38.01Whole of U.P.Blue flower, brown- and medium-seeded, highly resistant to rust, moderately resistant to alternaria blight
82SURYA
(LCK 1404)
2018
3220 E,
5 September 2019
Shubhra x H-25Kanpur155–1601431(I)35.49H.P., Punjab, Haryana, and J&KBlue flower, brown- and medium-seeded, highly resistant to rust, moderately resistant to wilt
83UTERA ALSI-2 (RLC-153)2019
1326 E, 2 April 2019
LCK88062x EC-1424Raipur118–120519(U)34.8Assam, Bihar, Chhattisgarh, Jharkhand, Orrisa, M.P., M.H. and U.P.Suitable for Utera conditions, moderately resistant to budfly
84LSL-932019
3220 E,
5 September 2019
Selection from SLS-93Latur 106846(R)37.07MaharashtraWhite flower, suitable for rainfed conditions, resistant to rust, moderately resistant to wilt, powdery mildew, alternaria blight, and bud fly
85TL 992019
3220 E,
5 September 2019
RLC 6 X SolinBARC
Trombey
1201274(I)36.56U.P., Bihar, Jharkhand, W.B., Assam, NagalandLow linolenic acid content (2–3%), edible oil purpose, resistant to wilt and rust, moderately resistant to budfly
86JAWAHAR LINSEED 1652019
3220 E,
5 September 2019
Meera x Polf 17Hoshangabad (M.P.)1251392(I)33.84H.P, Punjab, and Jammu and KashmirBlue flower, suitable for irrigated conditions, resistant to rust and moderately resistant to powdery mildew, wilt, and bud fly
87KOTA BARANI ALSI 5
(RL 29005)
2020
99 E, 9 November 2020
PKDL 41 x Meera Kota
(Raj.)
115–1201150(R)36.00RajasthanModerately resistant to wilt, alternaria blight, and bud fly
88KOTABARANI ALSI 6
(RL 15584)
2020
3482 E, 7 October 2020
RL 101 x RLU 6Kota
(Raj.)
115–1201200(R)32.00H.P, Punjab, and Jammu and KashmirModerately resistant to wilt, powdery mildew, rust, alternaria blight, and bud fly
89KOTA ALSI 6
(RL 13165)
2020
3482 E, 7 October 2020
RLC 25127 x RL 26016Kota
(Raj.)
1200(R)32.00UP, Jharkhand, Bihar, W.B., and Assam Moderately resistant to wilt and powdery mildew, rust, alternaria blight, and bud fly
90RLC 161
(SUVEE)
2020
3482 E, 7 October 2020
Ayogi x GS 234Raipur (C.G.) 1262(R)32.10H.P, Punjab, and JammuModerately resistant to alternaria blight and budfly
92RLC 1642020
500 E, 29 January 2021
Polf 22 x JRF 5Raipur (C.G.) 1161(R)32.60H.P, Punjab and JammuResistant to rust, moderately resistant to alternaria blight and bud fly
93RLC 1672020
500 E, 29 January 2021
T 397 x Polf 22 Raipur (C.G.) 1297(R)34.20H.P, Punjab and JammuModerately resistant to wilt, powdery mildew, and budfly
93APARNA (LCK1611)2020
500 E, 29 January 2021
BAU610 x ParvatiKanpur (U.P.) 1342(I)33.00H.P, Punjab, and JammuResistant to Alternaria blight, moderately resistant to wilt and powdery mildew
95
BUAT ALSI-4
(LMS 15-31)
2020
500 E, 29 January 2021
P 18 x Accn. No. 2299Mauranipur (U.P.)
1271(I)31.80Bundelkhand part of U.P., M.P., MH, K.N., Orisha, C.G.Moderately resistant to powdery mildew and budfly
96
BIRSA TISI 1
(BAU 15-03)
2021
8 E, 24 December 2021
(Shekhar x RLC 80) x GarimaRanchi, Jharkhand128–1301141(R)41.80Jharkhand Resistant to wilt and powdery mildew, moderately resistant to alternaria blight and bud fly
97PRIYAM
(BAU 12-1)
2021
8 E, 24 December 2021
BAU 2K-1 x GarimaRanchi, Jharkhand128–1301253(R)40.70H.P., Punjab, and Jammu and Jharkhand Resistant to rust, wilt, alternaria blight, and powdery mildew, moderately resistant to bud fly
98DIVYA
(BAU 06-3)
2021
8 E, 24 December 2021
BAU 1008 x Kiran Ranchi, Jharkhand128–1301538(I)39.80H.P., Punjab, Jammu, and JharkhandResistant to powdery mildew and rust, moderately resistant to alternaria blight and wilt
99SHUATS ALSI 2
(SHA-2)
2021
8 E, 24 December 2021
LCK 8879 x CI 1399Naini (U.P.)123–1251110(R)37.40Uttar PradeshResistant to powdery mildew and wilt, moderately resistant to rust
100SABOUR TISI 2
(BRLS-101)
2021
8 E, 24 December 2021
SLS 72 x ShekharSabour, Bihar118547(U)38.20U.P., Bihar, Jharkhand, W.B., Assam, M.P., M.H.,
C.G., Nagaland, Raj., Odisha, Karnataka
Resistant to rust and wilt, moderately resistant to Alternaria blight
101SABOUR TISI 3
(BRLS-107-1)
2021
8 E, 24 December 2021
LCK 7035 x ShekharSabour, Bihar1221883(I)37.80U.P., Bihar, Jharkhand, W.B., Assam, M.P., Nagaland, Raj., M.H., C.G., Odisha, KarnatakaResistant to powdery mildew and wilt, moderately resistant to rust
102(RLC 171)2022
3254 E, 27 July 2022
Polf 22 x JRF 22Raipur
(C.G.)
1321073(R)34.53Assam, Bihar, C.G., H.P., Jammu, JK, K.N., M.P., M.H., Nagaland, Orrisa, Punjab, Raj., and U.P.Blue flower, moderately resistant to rust and bud fly
103SABOUR TISI 4
(BRLS 121)
2023
1056 E, 6 March 2023
(NL 260 x Shekhar) x PKDL 71Sabour, Bihar120100034.00UP., Bihar, Jharkhand, W.B., Assam and NagalandBlue flower, moderately resistant to rust and bud fly
104JAWAHAR LINSEED SAGAR 122
(JLS 122)
2023
1056 E, 6 March 2023
JLS 73 x JLS 66Sagar
M.P.
116964(R)41.00Madhya PradeshBlue flower, semi-dehiscent capsule suitable for mechanical harvesting, bold-seeded (8 g), high omega 3 (54.20), resistant to rust, moderately resistant to powdery mildew, Alternaria blight, and bud fly
105AZAD PRAGYA
(LCK 1516)
1056 E, 6 March 2023Shubhra x Shikha Kanpur (U.P.)1281345(I)35.00Uttar PradeshBlue flower, resistant to rust and powdery mildew, moderately resistant to wilt and bud fly
106SHUATS ALSI 5
(SHA-5)
2023
1056 E, 6 March 2023
RL 10205 x IC 564620Naini (U.P.)125–1281250(I)35.80Uttar PradeshBlue flower, resistant to rust and powdery mildew, moderately resistant to wilt and bud fly
Table 4. Genome-wide association studies discovered quantitative trait nucleotides/loci for key abiotic stressors [196].
Table 4. Genome-wide association studies discovered quantitative trait nucleotides/loci for key abiotic stressors [196].
TraitCandidate GeneTypeQTN/QTLFunctionReference
Root surface
area stability
Lus10034840Calcium transporting ATPase9, Plasma membrane typeLu5-4,774,423Pollen development[82]
Lus10016017Catalase isozyme 3Lu6-15,939,492Response to water deprivation, promotion of drought stress tolerance[81]
Total root
length stability
Lus10039723IAA amidosynthetaseGH3.6Lus-20,209,630Response to stress and root development[82]
Lus10039747Diacylglycerol kinase 5 Cold and drought stress tolerance[81]
Lus10021019Allene oxide synthase Lu6-19,733,117Stomatal closure and drought tolerance[180]
Lus10020997S/Tproteinkinase SRK2E Response to water deprivation and regulation of stomatal closure[81]
Stress
tolerance index
Lus10019811Probable cinnamyl alcohol
dehydrogenase1
Lu6-17,376,408Drought tolerance[82]
Lus10014978aquaporinPIP2-2Lu14-23,517,150Drought tolerance[180]
Lus10019781L-Ascorbate peroxidase Enhanced salt tolerance, drought tolerance, and cold tolerance[181]
Total root volume
stability
Lus10016017Catalase Isozyme CLu6-15,961,789Promotion of drought stress tolerance and response to water deprivation[81]
Bundle weight under drought stressLus100403333-Ketoacyl-CoA synthase 19Chr9:4203006Drought tolerance and biomass-related traits[217]
Canopy temperature under drought stressLus10013240Xyloglucan endotransglucosylase/hydrolase protein 27Chr2:23123754Leaf size and veins and drought susceptibility index[82]
Lus10019365Stromal cell-derived factor 2Chr3:9279281Heat stress and better stress tolerance indices[217]
Lus10024816Cytochrome Chr9:18937269Moisture stress tolerance[81]
Seeds per bollLus10021766Mitogen-activated protein kinase kinasekinase 5Chr9:15446958Drought susceptibility index[81]
Grain yieldLus10042229CBL-interacting protein kinaseChr11:3972867Drought [180]
Lus10042231Translocon chloroplast 110 Heat shock and drought susceptibility index[181]
Thousand-seed weight under
drought stress
Lus10029127Kelchrepeat F-boxChr1:7029139Ovule development and stress tolerance index[82]
Lus10029115Ribosomal pentatricopeptide repeat protein 4 Seed development and stress tolerance[82]
Lus10030137Nuclear factor Y subunit A1Chr12:10910146Seed development and drought stress tolerance[81]
Lus10030142Translocated promoter region, nuclear pore anchor Flowering, auxin signalling[81]
Plant height under drought stressLus10029690/1Cellulose synthase interactive Chr5:1375386Flax fibre and stress tolerance index[82]
YieldLus10031398Inositol Monophosphatase 1Chr12:20557728Drought tolerance[81]
Table 5. Genome-wide association studies that discovered quantitative trait nucleotides/loci for key biotic stressors [196].
Table 5. Genome-wide association studies that discovered quantitative trait nucleotides/loci for key biotic stressors [196].
TraitQTN/QTLCandidate
Gene
TypeFunctionReferences
Pasmo resistanceQTL 45Lus10031043LRR receptor kinaseBacterial pathogen-associated molecular pattern[211]
(PAMP) receptor
Lus10031058Elongation FactorEffector-trigger immunity[135]
Fusarium wilt resistanceafB13-- --[218]
Powdery mildew resistanceQPM-crc-LG1-- --[219]
-Pm1 [135]
Lu4-12,432,479Lus10036891 RGA (WRKY transcription factor)[212]
Table 6. Global transcriptome study of gene expression patterns in flax in response to significant abiotic and biotic stressors [196].
Table 6. Global transcriptome study of gene expression patterns in flax in response to significant abiotic and biotic stressors [196].
TraitTissueDEGs/DEUsKey PointsPlatform/ToolReferences
Response to salt stressFlaxseeds77,361,566 in neutral-salt stress, alkaline-salt stress, and alkaline stress Photosynthesis-, pathogen-, and wax-related genesIllumina HiSeq 2000[238]
Flaxseeds3374 upregulated and 18,040 downregulated Provide high-impact gene expression profileIllumina high-throughput sequencing[232]
Response to drought stressSeeds, roots, and shoots183:72 upregulated and 111 downregulatedMaintain homeostasis and growthCombi matrix 90 K array[72]
Response to drought-sensitive and tolerant varieties Flax leavesIn cv.Z1413,245 upregulated and 4167 downregulatedDNA repair from ROS damage and proline biosynthesisPacBIo ISO-seq [239]
Response to osmotic stress under PEG-induced condition Flaxseeds3922:1487 upregulated
and 2432 downregulated
Signal transduction and biochemical pathwayIllumina platform [240]
Response to aluminium stress and high soil acidity Flax seedlings-Compartmentalisation of Ca2+ in vacuoles Illumina Platform[230]
Response against flax rust (Melanosporalini)Leaf tissues-Gene encoding PR protein and hydrolysis and uptake of nutrientsIllumina genome analyzer II[241]
Response to Fusarium wilt Roots, stem Transduction and reception of pathogen signalsIllumina HiSeq 2000[242]
Table 7. Important genes in flax that are activated and deactivated in response to biotic and abiotic stressors.
Table 7. Important genes in flax that are activated and deactivated in response to biotic and abiotic stressors.
CharacterUpregulated/Downregulated GenesReferences
Salinity and alkalinityHSP70 NAC family members, ABA, WRKY, PrxR[238]
LUS-miRNAs and miRNA-targeted genes [281]
UBE2 gene, mitochondrial termination factor family protein and auxin signalling F-box, Myb domain protein [282]
DroughtLipid transferase protein, cell wall synthesis genes, RUBISCO, PS1, r2r3-MYB transcription factor, EF-tu, LEA5, cytochrome P450 family proteins, brassinosteroid insensitive I-associated receptor kinase1, and AP2/ERF domain-containing transcription factor[72]
NAC domain protein[112]
HeatHSF[283]
Heat shock factors (HSFs)[111]
miRNAs and phasiRNAs[114]
GUS activity in sepals, petals, and pistils[110]
Nutrient stressWRKY, ING1 family JAZ and HARB11[284]
Aluminium stressmiR319, UDP-glycosyl-transferasemiR390, glutathione-S-transferase and miR393[234]
RustAvrs and CWDEs[240]
Fusarium wiltPAL, PCBER, SRG1, UGT73C3, AAA-ATPase ASD, mitochondrial (AATPA), glucan endo-1,3-β-glucosidase, MYB transcription factors, ERD dehydrins, and auxin-responsive protein SAUR, WKY3, WRKY70, WRKY75, MYB113, and MYB108[234,285]
Fusarium culmorumPAL, CCR, CAD, UGT, and TD[233]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Paliwal, S.; Tripathi, M.K.; Tiwari, S.; Tripathi, N.; Payasi, D.K.; Tiwari, P.N.; Singh, K.; Yadav, R.K.; Asati, R.; Chauhan, S. Molecular Advances to Combat Different Biotic and Abiotic Stresses in Linseed (Linum usitatissimum L.): A Comprehensive Review. Genes 2023, 14, 1461. https://doi.org/10.3390/genes14071461

AMA Style

Paliwal S, Tripathi MK, Tiwari S, Tripathi N, Payasi DK, Tiwari PN, Singh K, Yadav RK, Asati R, Chauhan S. Molecular Advances to Combat Different Biotic and Abiotic Stresses in Linseed (Linum usitatissimum L.): A Comprehensive Review. Genes. 2023; 14(7):1461. https://doi.org/10.3390/genes14071461

Chicago/Turabian Style

Paliwal, Shruti, Manoj Kumar Tripathi, Sushma Tiwari, Niraj Tripathi, Devendra K. Payasi, Prakash N. Tiwari, Kirti Singh, Rakesh Kumar Yadav, Ruchi Asati, and Shailja Chauhan. 2023. "Molecular Advances to Combat Different Biotic and Abiotic Stresses in Linseed (Linum usitatissimum L.): A Comprehensive Review" Genes 14, no. 7: 1461. https://doi.org/10.3390/genes14071461

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