Next Article in Journal
Abiotic Stress Effect on Agastache mexicana subsp. mexicana Yield: Cultivated in Two Contrasting Environments with Organic Nutrition and Artificial Shading
Previous Article in Journal
Characterization of Plant-Growth-Promoting Rhizobacteria for Tea Plant (Camellia sinensis) Development and Soil Nutrient Enrichment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sesame, an Underutilized Oil Seed Crop: Breeding Achievements and Future Challenges

1
Department of Plant Breeding and Genetics, College of Agriculture, University of Sargodha, Sargodha 40100, Pakistan
2
Department of Plant Breeding, Swedish University of Agricultural Sciences, P.O. Box 101, 23053 Lomma, Sweden
3
Department of Genetic and Bioengineering, Engineering Faculty, Trakya University, Edirne 22030, Turkey
*
Author to whom correspondence should be addressed.
Plants 2024, 13(18), 2662; https://doi.org/10.3390/plants13182662
Submission received: 26 July 2024 / Revised: 18 September 2024 / Accepted: 20 September 2024 / Published: 23 September 2024
(This article belongs to the Section Plant Genetic Resources)

Abstract

:
Sesame seeds and their edible oil are highly nutritious and rich in mono- and polyunsaturated fatty acids. Bioactive compounds such as sterols, tocopherols, and sesamol provide significant medicinal benefits. The high oil content (50%) and favorable mono- and polyunsaturated fatty acid balance, as well as resilience to water stress, make sesame a promising candidate crop for global agricultural expansion. However, sesame production faces challenges such as low yields, poor response to agricultural inputs, and losses due to capsule dehiscence. To enhance yield, traits like determinate growth, dwarfism, a high harvest index, non-shattering capsules, disease resistance, and photoperiod sensitivity are needed. These traits can be achieved through variation or induced mutation breeding. Crossbreeding methods often result in unwanted genetic changes. The gene editing CRISPR/Cas9 technology has the potential to suppress detrimental alleles and improve the fatty acid profile by inhibiting polyunsaturated fatty acid biosynthesis. Even though sesame is an orphan crop, it has entered the genomic era, with available sequences assisting molecular breeding efforts. This progress aids in associating single-nucleotide polymorphisms (SNPs) and simple sequence repeats (SSR) with key economic traits, as well as identifying genes related to adaptability, oil production, fatty acid synthesis, and photosynthesis. Additionally, transcriptomic research can reveal genes involved in abiotic stress responses and adaptation to diverse climates. The mapping of quantitative trait loci (QTL) can identify loci linked to key traits such as capsule size, seed count per capsule, and capsule number per plant. This article reviews recent advances in sesame breeding, discusses ongoing challenges, and explores potential strategies for future improvement. Hence, integrating advanced genomic tools and breeding strategies provides promising ways to enhance sesame production to meet global demands.

1. Introduction

Sesame (Sesamum indicum L.) is among the oldest oilseed crops known to humanity and is often referred to as the “queen of oilseed crops”. It is a diploid species (2n = 2x = 26 chromosomes), although variability in chromosome numbers has been identified in other species within the genus Sesamum [1]. Species such as S. calycinum, S. latifolium, and S. angolense are also diploids (2n = 2x) but with 32 chromosomes [1].
Sesame is an oilseed crop containing more than 50% edible oil and a balanced ratio of unsaturated and saturated fatty acids. Seed is also a rich source of functional molecules such as tocopherols, sterols, and sesamol. These molecules are not only beneficial for human health, but they also improve the shelf life of the oil due to their antioxidative properties. However, sesame has low yield potential when compared with other oilseed crops such as canola or sunflower, and little breeding efforts were made to improve yield potential, the introgression of host plant resistance to pathogens and pests, and shattering resistance [2]. The incorporation of genes for these characteristics may help to improve the harvestable yield of the sesame crop.
Sesame is grown in an area of 12.81 million ha, and 6.55 million metric tons were produced worldwide in 2022 [3,4]. Evidence suggests that the crop originated in Africa, where significant genetic diversity includes many wild relatives that are endemic to the various regions of this continent. Hence, Africa is considered its center of diversity. Sudan is the largest producer of sesame seed in the world followed by Myanmar and India [3,4].
Sesame is a self-pollinated, semi-erect plant that grows to a height of 0.6 to 1.2 m. Its ovate or lanceolate leaves vary, with basal leaves being tri-lobed and upper leaves irregular and serrated. As a drought-tolerant crop, sesame prefers well-drained soils with a pH of 5.5 to 6.8 and is suited to temperate tropical climates. The bell-shaped flowers have petal colors ranging from pale yellow to purple. However, sesame has a low yield potential and typically responds poorly to farm inputs, thus making it an unattractive option for growers compared to other oilseed crops. Continuous cultivation of sesame in the same field leads to declining yield, protein, and mineral content over time [5]. Additionally, the crop faces challenges such as shattering at maturity, which makes harvest difficult, and insect and disease infestations causing significant yield loss globally. Concrete efforts in breeding and genetics are needed to develop elite, high-yielding sesame cultivars with host plant resistance to insects and other pathogens. Current developments in breeding, genomics, and marker-assisted selection, therefore, highlight the importance of this crop.

2. Germplasm Resource Collection and Diversity Study in Sesame

2.1. Germplasm Resource Collection and Conservation

The primary breeding objectives for sesame are to increase seed yield, improve plant morphology, enhance tolerance to biotic and abiotic stresses, develop indehiscent capsules, and improve oil quality. These varied breeding goals highlight the critical importance of germplasm collection and conservation. A robust and diverse genetic repository is essential for breeders, providing access to a wide range of traits and alleles needed to develop improved cultivars that meet evolving agricultural challenges and consumer demands. Maintaining a comprehensive germplasm collection ensures the preservation of genetic diversity, which is crucial for the long-term sustainability and advancement of sesame breeding programs. This genetic resource is the foundation for future crop improvement innovations, climate change adaptation, and food security. Hence, germplasm collection and conservation is a continuous process to protect wild resources for sustainable agriculture. It has been thought that sesame originated from the wild populations native to Pakistan and Afghanistan. Alternative to this hypothesis, several wild species that are native to Africa could be also ancestors of cultivated sesame. Genetic research may be required to establish a relationship between modern cultivars and their parental population.
Sesamum indicum germplasm is conserved, inter alia, at the Brazilian Gene Bank (BGB, Brazil), International Crops Research Institute for the Semi-Arid Tropics (ICRISAT, India), the National Bureau of Plant Genetic Resources (NBPGR, India), the National Institute of Agricultural Science and Technology (NIAST, Republic of Korea), the Nigerian National Gene Bank (NNGB), the United States Department of Agriculture—Agricultural Research Service (USDA-ARS, USA), and Vavilov Institute of Plant Industry (VIR, Russia). ICRISAT conserves 1302 accessions of sesame germplasm, including wild accessions and those collected from international germplasm exchange. Similarly, NBPGR conserves 5288 accessions of sesame germplasm [6], while 11,116 accessions are held at the USDA-ARS repository, 9221 accessions at VIR, 3317 accessions at NIAST, 1455 accessions at BGB, and 1039 accessions at NNGB [6]. Evaluation and characterization of these accessions for morphological, agronomical, and molecular traits have been carried out, resulting in the identification of unique genotypes with high yield, disease resistance, and quality traits. For example, the FGsesame (http://ncgr.ac.cn/SesameFG/, accessed on 9 May 2024) database includes core collections of 705 sesame accessions from 29 countries, each identified based on origin, ecotype, sequencing coverage, and group information [7].
Germplasm resources are also crucial for ecosystem sustainability and nutritional security. The utilization of these resources in breeding programs, therefore, can lead to the development of improved sesame cultivars. The wild relatives of sesame contain unique genes that are not found in the cultigen pool [8]. It is, therefore, important to protect them from over-exploitation by local communities for culinary, medicinal, and cosmetic uses to prevent their extinction and maintain ecosystem balance.

2.2. Diversity Studies in Sesame

Research has demonstrated high genetic diversity among wild sesame relatives and cultivated germplasm from various regions. Wild relatives of sesame offer valuable traits for breeding programs, such as cytoplasmic male sterility, heat tolerance, disease resistance, and fatty acid modification. For instance, S. radiatum has more linolenic acid than other species [9]. The wild population of S. orientale closely resembles cultivated types (S. indicum) in both morphology and cytogenetics. Both species are diploid with 26 chromosomes and produce fertile hybrids under spontaneous or artificial reciprocal hybridization. CpDNA-based fingerprinting also supports a close relationship between S. orientale and S. indicum [10]. Its origin is still uncertain but some authors suggest it originates from the African population of S. latifolium as it is widely cultivated and found growing wild in tropical regions [11].
Crosses between wild and cultivated species have successfully introgressed desirable traits, such as charcoal rot resistance [12]. However, crosses between S. indicum and S. latifolium were largely unsuccessful, producing only a few seeds due to chromosomal number and structural differences (S. latifolium has 2n = 32 chromosomes). Both species also show qualitative differences, such as the absence of sesamolin in S. latifolium, thereby indicating both cytological and morphological differentiation between the species.
Molecular, biochemical, and morphological markers have been used to assess diversity among sesame accessions. Isozyme analysis and morphological trait comparisons have differentiated wild and cultivated species [1,13]. The presence of polymorphic co-dominant cathodic and anodic peroxidase isozymes, Co-Px-2 and An-Px-2, is unique to wild sesame germplasm. Additionally, the Sh-1 marker distinguishes S. indicum from wild S. mulayanum. Similarly, the Sh-1 marker can distinguish S. orientale from other wild accessions, confirming genetic diversity in wild sesame.
The development of molecular markers, such as simple sequence repeats (SSR) and single-nucleotide polymorphisms (SNP), has also been instrumental in the genetic diversity analysis and mapping of quantitative trait loci (QTL) in sesame [14]. SSR-based research has shown genetic diversity indices and divergence among Indian accessions [15]. Expressed sequence tag (EST) databases were exploited for the development and characterization of gene-derived SSR markers using the computer-based software MISA version 1.1.2, while SNPs were detected and visualized using the QualitySNPng software tool (http://www.bioinformatics.nl/QualitySNPng, accessed on 9 May 2024). SNP variants were identified through a genome comparison of sesame. The sesame genome survey identified SSR markers that could be exploited to determine the genetic diversity and QTL mapping of economic traits. There were 23,438 SSR (five repeats) markers, with dinucleotides being the most prevalent repeat pattern (84.24%), followed by trinucleotides (13.53%), tetranucleotides (1.65%), pentanucleotides (0.3%), and hexanucleotides (0.28%) [14]. Identifying microsatellites in the sesame genome led to the development of 218 polymorphic SSR markers for screening sesame germplasm for genetic diversity analysis [14]. The sesame cultivar “Sweeta” genome sequences were obtained through Illumina paired-end sequencing and 454 shotgun sequence technologies. Reports on the genetic diversity within diverse global accessions or native landraces are given in Table 1. Improved assembly and annotation of the genome were performed by high-quality sesame genome assembly using PacBio and Hi-C technology, improving continuity and chromosome anchoring. The new assembly identifies 24,345 protein-coding genes and reveals ancient whole-genome duplication events. These advancements will support future genetic research and breeding efforts in sesame [16]. The GINMicrosatDb database of microsatellites was developed, containing five sets of primer pairs for each microsatellite locus [17]. The database also included GC content, melting point, and flanking sequences [17].
Transcriptomic profiling of sesame genes revealed 57 MADS-box genes across 14 linkage groups that are essential for sesame development and growth [7]. Structural analysis and functional differences among MADS-box genes showed similarities and differences with model species Arabidopsis thaliana and Solanum lycopersicum [7]. The MIKCC-type MADS-box genes are also shown related to sesame flower initiation and seed development [7]. Transcriptomic analysis of interspecific hybrids has also identified genes related to disease resistance [18]. Studies of 4.4 Mbps related to fatty acid metabolism show genetic diversity loss during the domestication of sesame [19]. Expression profiles of traits such as capsule, seed leaf, and root development in cultivar “Zhongzhi13”, as well as the transcriptomic profile of genes related to waterlogging tolerance in “ZZM2541” and oil content in high-oil-yielding cultivar “Zhongfengzhi1”, were investigated [7]. Candidate genes related to seed yield, lipid metabolism, phenological traits, or disease resistance are available in the database [7]. Hence, wild and cultivated accessions should be further evaluated and characterized for morphological, agronomical, and molecular traits and consequently used as parental material to develop pre-breeding lines or improved sesame genotypes with high yield, disease resistance, and quality traits. A comprehensive summary illustrating the concepts discussed in the text on sesame breeding and conservation is illustrated in Figure 1 below.
Table 1. Genetic diversity estimates in diverse sesame germplasm across global and local accessions.
Table 1. Genetic diversity estimates in diverse sesame germplasm across global and local accessions.
GermplasmOriginMarkersDiversityReference
183 accessions: 101 Chinese accessions, 62 landraces, and 20 exotic accessionsChina25 simple sequence repeat (SSR) markers0.27–0.65 Shanon information index
Moderate
[20]
2496 sesame accessions: 17 countries and 26 states of IndiaIndia43 SSR markers1.14–2.07 Shanon information index[21]
35 sesame landracesGreece, Bulgaria, Turkey, Iraq, Yeman, Tajikistan, Korea7 expressed sequence tag (EST)-SSR markers0.27–0.35 Shanon information index
0.15–0.21 Diversity index
[22]
6 sesame cultivarsSaudi Arabia20 inter simple sequence repeat (ISSR) and 25 SCoT primersInter simple sequence repeat (ISSR) vs. start codon targeted (SCoT) polymorphisms according to polymorphism information content (PIC)
65% vs. 78%
0.26–0.31 Shanon information index
[22]
100 sesame accessionsEthiopia27 SSR markersMean gene diversity 0.30 [23]
300 sesame accessionsAfrica and Asia5065 silico diversity array technology (DArT) and 5821 single-nucleotide polymorphism (SNP) markersAverage diversity 0.14 Asia (0.17) vs. Africa (0.14)
Africa (0.21) vs. North Africa (0.23)
[24]
501 accessionsAfrica, Europe, America, Asia24,735 SNP markersHigh degree of differentiation (0.27) [25]
2496 sesame accessionsNational Genebank of ICAR-National Bureau of Plant Genetic Resources64,910 SNP markers384 accession (15%) represent maximum diversity. Captured 60% of the SNP diversity[26]

3. Oil Content and Fatty Acid Composition Analysis in Sesame Germplasm

Oil content in sesame accessions varies depending on genotype and environmental interaction, ranging from 45% to 53% among accessions from Turkey and 11 other countries [27]. The predominant fatty acids in sesame oil are oleic acid (36–43%), linoleic acid (39–46%), and palmitic acid (8–10%) [27]. Sesame oil is rich in unsaturated fatty acids, making it suitable for culinary and industrial applications (Table 2). A recent study utilized genome-wide association analysis (GWAS) and transcriptome analysis to identify novel loci and regulatory genes involved in fatty acid biosynthesis [28]. The study discovered that specific gene clusters, including FAD2, LOC10515945, LOC105161564, and LOC105162196, play crucial roles in regulating unsaturated fatty acid biosynthesis. Additionally, a regulatory co-expression network was constructed based on differentially expressed genes, aiding in understanding the transcriptional and translational regulation of oil biosynthesis in sesame seeds. These findings provide a molecular platform for improving the quality and yield of sesame oil, potentially benefiting breeding programs aimed at developing sesame cultivars with higher levels of unsaturated fatty acids.
Major lignans in sesame seeds include sesamin and sesamolin, with higher concentrations found in white-seeded cultivars [29]. Sesame seeds are also a rich source of minerals, with calcium being the predominant cation, followed by potassium and magnesium [30]. The seeds contain various phenolic compounds that impart antioxidant properties, which may help prevent disease [31].
Table 2. Natural variation for oil content (%) and fatty acid composition (%) in sesame germplasm.
Table 2. Natural variation for oil content (%) and fatty acid composition (%) in sesame germplasm.
AccessionOil
Content
ProteinFatty AcidsReference
16:018:018:118:218.322:1
22 cultivars, 4 landraces of S. mulayanum, and
7 accessions of 4 wild species
9–1537–5230–520–8[19]
4 Pakistani sesame varieties50–5419–233–195–22105–1316[32]
Market sample48 5134336[33]
Brown non-roasted vs. roasted sesame49 vs. 515.6 vs. 7.15.3 vs. 6.840.3 vs. 58.746.1 vs. 250.4 vs. 0.2[34]
103 sesame landraces41–638–1029–4141–49[35]
Collection of 12 countries45–535–68–1036–4439–460.3–04 [36]
6 cultivated (C)
3 wild spp. (W)
53–55 (C)
54–59 (W)
4–9 (C)
4–8 (W)
4.5–10 (C)
4–5 (W)
33–38 (C)
35–37 (W)
42–52 (C)
43–46 (W)
4–9.5 (C)
5–10 (W)
[9]
Indian sesame cultivar42–5416–278–124–841.5–5032–430.2–0.35[36]

4. Fatty Acid Biosynthesis and Metabolism in Sesame

Fatty acids in sesame are synthesized de novo within the chloroplasts and can be esterified with glycerol to form triglycerides, which are stored as oil bodies in plant tissues. The fatty acid composition dictates the end use of the oil, with higher polyunsaturated fatty acids being unsuitable for deep frying due to oxidation [37]. Efforts to modify fatty acids in crops like Brassica, sunflower, and soybean through genetic techniques have aimed to improve oxidative stability and nutritional value [38,39,40]. Fatty acid accumulation starts 21–33 days after flowering. There are 38 different types of fatty acids in sesame, which were detected through gas and liquid chromatography [41]. Candidate genes for the oil and fatty acids can be identified through RNA profiling. RNA profiling in sesame showed the role of various candidate genes, e.g., CCase, FAD2, DGAT, G3PDH, PEPCase, WRI1, and WRI1-like genes [42]. Efforts to enhance ω-3 fatty acid content could address deficiencies in human diets, although such oils may be unsuitable for cooking due to rapid oxidation [38].
To increase stearic acid concentration, it is crucial to enhance the activity of acyl-ACP thioesterase and suppress the SAD1 gene [43,44]. Blocking ω-6 fatty acid desaturase activity can improve oleic acid concentration in sesame oil [45]. The sequence homology of FAD2 transcripts obtained from wild and cultivated species showed divergence and SNP among both groups of germplasm [46]. Polymorphism within nucleotides led to the change of amino acids in the active site of the enzymes, leading to the differential ability of enzymes to catalyze linoleic acid contents within wild and cultivated germplasm [46]. It has been shown that abscisic acid induces seed-specific expression of the SeFAD2 gene, which is controlled by introns [47]. FAD2 gene expression may be blocked through anti-sense technology that helps to raise the concentration of oleic acid. FAD2 complete transcripts in various crop species such as soybean, sunflower, and sesame were obtained through an online gene bank database. Sequences were downloaded, and their relative sequence homology was compared on MEGA X software version 11.0.10 (Figure 2). It showed that species could be characterized based on the sequence homology for FAD2 genes showing their independent orthologous evolution. High genetic differentiation within the FAD2 gene between species may lead to various versions of the oleate desaturase enzyme. Soybean accessions showed the highest sequence homology, showing low genetic differentiation within species while sunflower had the highest genetic differentiation between their accessions (Figure 2).

5. Sesame Wilting Disease

Sudden wilting in sesame is a common production constraint in hot and dry regions of the world. Initially, it was known that the disease was caused by Fusarium oxysporum [48]. However, it has been observed that sudden wilting syndrome was caused by cumulative biotic and abiotic factors. In Pakistan, sudden wilting syndrome was one of the major production constraints due to the hot climate and caused 50–100% yield loss across the belt of central Punjab, Pakistan. Sudden wilting is observed in hot climates (40 °C) after delayed irrigation, and plants experience wilting within 24 h. The infected plants had Fusarium oxysporium and Macrophomina phaseolina [49]. However, many of the samples did not carry any of the two pathogens, but abiotic factors such as heat and drought often lead to the sesame wilt. The incidence of wilting increased with delayed irrigation [49].
A significant increase in the browning of the stem in the wilted plants of sesame was observed after 5–6 days of rain (20 mm) in heavy soils of Shahkot, Pakistan, indicative of charcoal rot infestation after wilting (Figure 3). Resistance breeding may be required to cope with the sudden wilting in the sesame. It may require the screening of breeding material and the development of cultivars that may survive hypoxia conditions within roots after heavy rainfall or irrigation [50]. Moreover, the introgression of charcoal or root rot disease resistance may further reduce the chances of fungal infestation after the wilting. Combining ability analyses of the sesame breeding lines were performed to determine the breeding value for charcoal rot resistance. Selectable variation for charcoal rot resistance was low to high, and there were significant environmental and dominance effects associated with the charcoal rot resistance which indicates that recurrent selection and genotypic selection will be practiced to improve the charcoal rot resistance [51]. The sesame genetic map was developed from 424 new polymorphic markers and 548 recombinant inbred lines [52]. These markers detected 14 QTLs which explained 3–14% of the total genetic variation contributing to the charcoal rot resistance [20]. The qCRR12.2 contributed the highest to charcoal rot resistance in all four environments, and qCRR8.2 and qCRR8.3 were environment specific. These QTLs were used to predict the disease-resistance genes. Genome-wide association research may be required to tag genes for resistance breeding.

6. Abiotic Stress Tolerance in Sesame

Sesame is a drought-tolerant crop with nitrogen recovery ability, making it suitable for arid climates with both low rainfall and low soil fertility [53]. Climate change, characterized by increasing CO2 levels and temperature fluctuations, poses significant challenges for sesame production. While sesame thrives in high temperatures, it experiences reproductive failures when temperatures exceed 33 °C [53]. Developing new cultivars with enhanced drought and heat tolerance, combined with high yield potential, is essential for sesame cultivation under changing climate conditions [54].
Selection for high yield potential and stability has been effectively carried out through the QTL mapping of economically important traits linked with yield, resistance against stresses, and stability (Table 3). Molecular markers and QTL mapping have been used to select high-yielding and stress-resistant cultivars [55]. Transcriptomic analysis of drought-tolerant sesame accessions has identified genes related to osmotic stress [56]. The root profile of drought-tolerant sesame accessions showed differential expression of genes related to catalytic activity, ion binding, and transferase activity compared to susceptible accessions [56]. This study, therefore, helps to identify the genes related to abiotic stresses such as osmotic stress.
Sesame is a short-day plant, and flowering is initiated in response to day length and temperature, with day length having a significant effect on sesame flowering time [65]. A gene named CONSTANS (CO) promotes flowering time in sesame [65]. It shows similarities to the zinc finger transcription factors of Arabidopsis and dictates the vegetative growth period of sesame. Understanding the genetic regulation of flowering and maturity can help develop cultivars suited to specific environments [57]. Flowering time may be a multigenic trait and numerous loci may show genetic interactions. In this regard, generation means analysis of flowering time in sesame indicated that two genes have additive, dominant, and epistatic effects [65]. It also showed a negative relationship with other morphological, growth, and yield-related traits [66]. Accessions with later flowering had a small reproductive period.
Gene regulation of flowering initiation in sesame could help to understand the mechanism underlying flowering and maturity under a specific set of environments. The gene CO has two homologs (SiCOL1 and SiCOL2. SICOL2) lacking the B-box motif. It delays flowering compared to genotypes with the SiCOL1 sequence. SiCOL1 is similar to the CO gene, showing high expression in leaves before flowering and diurnal rhythmic expression in response to day length. It has 16 haplotypes within Asian genotypes, and the mutant type does not respond to day length, indicating it is non-functional or photoperiod-insensitive. Genotypes having a mutant allele of SiCOL1 initiated early flowering under long days [57]. Hence, genome sequencing and annotation may help identify differentially expressed genes and understand the adaptability of sesame genotypes to a particular environment.

7. Inducing New Variability within Sesame Germplasm

7.1. Mutation Breeding

Sesame crop has defective phenology, with pod shattering, susceptibility to diseases, poor response to farm inputs, and insect infestation. To overcome these problems and to obtain new genetic variability within elite germplasm, a mutation breeding program was initiated in various parts of the world. New genetic variation was induced through the use of various physical and chemical mutagens (Table 4). At least 25 mutants have been released for general cultivation with improved agronomic and stress resistance under various mutation breeding programs in eight countries [67]. A dose of chemical mutagen of 0.5% or 1% ethyl methane sulfonate (EMS) has been recommended to induce novel genetic variability, which generated novel capsule-bearing traits such as tetra- or tri-capsule per leaf axil, determinate growth habit, and seed color [68]. The optimum lethal dose for gamma radiation was 389–417 Gy [69]. Selection within the mutant population and progeny evaluation of advanced mutant lines (M5) led to the selection of candidate lines SM-06, SM-07, and SM-04. These candidate lines were superior in terms of early maturity, and higher yield components such as capsule plant-1 when compared with standard checks [70]. Mutation breeding had significant effects on traits related to earliness and days to 50% flowering. These traits also showed significant variability within mutant lines [71]. In another study, advanced mutants (M3) were screened for high oleic acid content (70%). Six mutants from M3 were advanced to M6–M7 which showed oleic acid contents of about 70% when compared with control lines [42]. The genetic basis of high oleic acid content (%) was also investigated. A nucleotide mutation within the FAD2 gene led to the substitution of amino acid in oleate desaturase, thus inducing an increase in the oleic acid contents [46].
Mutation breeding programs generally had random effects on the genome, and generated mutants may contain several mutations within non-targeted traits. Traditional mutants are used as pre-breeding material to transfer genetic variation into elite germplasm. Contrastingly, CRISPR/Cas9 could be used for targeted genome modifications in crops. The mutants released through CRISPR/Cas9 may have the least detrimental effects on yield and related traits. In this era of genomics, new genes for photoperiod sensitivity, fatty acids, shattering, growth habit, pathogen, and insect resistance will be constantly identified. CRISPR/Cas9 may be used as a tool to downregulate undesirable genes. The downregulation of genes related to fatty acids may help to develop high-steric- or –oleic-acid mutants, while genes related to the shattering problem may be knocked out to develop shattering-resistant mutants which helps to minimize yield losses and improve seed quality as well [72]. Sesame-specific adaptability may be improved by downregulating genes related to photoperiod. Photoperiod insensitivity helps to initiate flowering under long day lengths [73]. Two sgRNA were designed in an attempt to downregulate the two cytochrome genes (CYP81Q1 and CYP92B14) that are involved in the biosynthesis of sesamin and sesamolin [74]. Sequence data showed a mutation frequency of 90.63 and 93.33% in CYP81Q1 and CYP92B14, respectively [75]. However, there were no off-target mutations identified in the study. The biosynthesis of sesamin and sesamolin was significantly disrupted in the transformed roots, which also confirmed the key functions of the genes in lignan biosynthesis [75].
Table 4. Exposure of sesame for the establishment of a mutant population.
Table 4. Exposure of sesame for the establishment of a mutant population.
MutagenEstablished PopulationNovel VariationReference
Sodium azideM2Glabrous stem[76]
Sodium azide
(15 mM)
M4 and M5Early maturity, yield advantage of 21% during the Kharif season[77]
600 Gy CoM4Salinity tolerance, number of capsules, 1000 seed mass[78]
300 Gy gamma rays416 M3 mutantP 97-1 Phytophthora nicotianae resistance[79]
40 KR gamma rays
1.5 mM
M2, M3Early maturity, no. of capsules, seed yield plant−1[80]
-M2Tetra-carpellate per leaf, higher number of seed capsules−1[81]
40 mL 0.8% ethyl methane sulfonate (EMS) in 100 mM phosphate bufferM3M5 generation was subjected to screening, and plants with high oleic acid contents (75%) were selected[82]
0.5% EMS Sorenson phosphate for 24 h at 4 °CM2A significant change in both qualitative and quantitative traits when 14 Ethiopian sesame genotypes were exposed to mutagen[83]
250 Gy gamma radiationM0Germination was reduced by 50%[84]

7.2. Development of Semi-Dwarf Plants and Various Mutant Types

Mutation breeding led to the evolution of a semi-dwarf sesame genotype (dw607). Semi-dwarf genotypes were characterized by reduced intermodal length, higher stem lodging resistance, and higher thousand seeds weight (TSW) due to larger seeds [85]. Molecular analysis of 824 accessions showed that Sidwf1 affects plant height. These 824 accessions had 58 different genomic variants for Sidwf1, which has an SNP mutation of nucleotide substitution from C to T, thereby causing an amino acid change from proline to serine [85].
An EMS-induced sesame mutant (sc1) with few seed capsules and short capsules was due to a mutation of the SICS1 allele [7]. Sequencing of loci SICS1 showed a point mutation at intron 5 and exon 6 junctions which induced a nonsense mutation due to a frameshift of the reading frame. The overall effect was a defective protein [7]. The indeterminate growth habit in sesame is dominant over the determinate type [86]. Multiple allelism for determinate growth habit dt1 and dt2 was noticed in crosses between two growth types [70]. Long and dense capsules were dominant over short and sparse capsules when two types were crossed, and the F1 hybrid was phenotypically characterized. These two types showed a segregating ratio of 12:3:1, which indicated that the two genes controlling the long and dense capsules had epistasis interactions [87]. In another study, the yl1 mutant was phenotypically characterized. The yellow-leaf dwarf mutant had lower carotenoid and chlorophyll contents than wild types. The yl1 mutant also had a longer growth duration with a slower photosynthesis rate, later flower initiation, and shorter plant height than the wild type [88].

7.3. Indehiscence Capsule

Capsule shattering is a major production constraint in sesame and leads to 50% yield loss [72]. To reduce yield loss due to capsule shattering, farmers tend to harvest manually with partial maturity. Capsule dehiscence is a dominant trait in sesame, and crossing the indehiscent parent with the dehiscent parent results in an indehiscent F1 hybrid [89]. Moreover, a Mendelian ratio (3:1) was observed in the F2 generation where a single recessive allele was controlling the indehiscent capsule [90]. The identified allele for shattering resistance was id Sindi0765000. However, it has several pleiotropic effects and induces undesirable effects on plant phenotype. Generation means analysis obtained by crossing capsule-shattering-resistant and -susceptible parents also confirmed this finding [90]. Anatomical analysis of shattering-resistant mutant 12M07 showed a parenchyma cell was loosely arranged but adhered to the seed coat [91]. Marker-assisted selection has been proposed, and gene-controlling shattering or non-shattering types were mapped onto SNP marker S8_5062843 (78.9 cM) near the distal end of LG8 (chromosome 8) [91]. Sequence analysis indicated a frameshift mutation in the non-shattering alleles, which caused a shift in the splicing and introduced SNP variation compared with the shattering alleles [91]. In another study, association mapping showed that SNP SV12002 in the 5′ upstream region of gene Sindi0765000 in chromosome 3 was closely associated with shattering resistance [92]. To facilitate the selection of non-shattering types, cleaved amplified polymorphic sequence (CAPS) markers were developed [65]. The shattering trait was mapped by 782 SNP markers, and shattering resistance was mapped on the distal part of LG8 in the population generated from Muganli-57 and PI 599446 [93]. A candidate gene for the shattering trait was identified that encodes a protein of 440 amino acids [93]. CRISPR/Cas9-based knockout within the candidate caused shattering resistance and leaf curling [93].

7.4. Male Sterility and Hybrid Vigor

Hybrid breeding provides another possibility for improving yield. Breeding lines with superior specific combining ability tend to show dominance and may also have shown superiority in yield and related traits. Economical heterosis ranged between 9.5 and 32.7%, as estimated in 1636 hybrid offspring [94]. Reliable genetic male sterility is a prerequisite for hybrid seed production. Genetic male sterility, including nuclear or cytoplasmic male sterility, may be induced through mutations. A list of various nuclear alleles identified or induced in sesame is given in Table 5.
Cytoplasmic male sterility sources were also discovered after interspecific hybridization. S. malabaricum was identified as a donor species for the sources of cytoplasmic male sterility [99]. Nuclear male sterility is controlled by two homozygous recessive alleles, which may be maintained in heterozygous conditions and upon self-pollination. The obtained progeny was subjected to the selection of male sterile plants through visual markers at the seedling stage [84]. Wrinkled leaves were used as a marker to differentiate male sterile plants [84]. Nuclear male sterility could also be induced spontaneously. For instance, recessive male sterility was identified in the Chinese cultivar “Zhuzhi”. This male sterility was transferred and maintained through sib mating and crossing with male fertile plants [100]. This source of nuclear male sterility was “D248A” [97]. Male sterility source “D248 A” had rudimentary small green anthers that were sterile and did not contain pollen. Abnormality in the pollen mother cells led to the anucleated microspores with less cytoplasm [97]. These microspore mother cells were not able to undergo normal meiotic division [97]. The segregation ratio of nuclear male sterility was 1:3 in the F2 generation. A dominant source of the genetic male sterility allele was also found in sesame [96]. This type of male sterility was easier to maintain than recessive genetic male sterility [97].
Male sterility may be restored through chemical agents which provide ease for the maintenance of genetic male sterility in sesame [101]. Genotype CRA-02 had 33% of male fertility when sprayed at bud initiation stages [101]. Male sterility and leaf curling were induced by the putative region at chromosome 4, which was identified through bulk segregant analyses followed by next-generation tools [93]. Region association with male sterility was further narrowed down through MAS by SSR and InDel markers in the backcross population BC1. Gene expression analyses showed suppression of four genes associated with sterile pollen and one gene during leaf development [92]. The gene cell wall invertase, CWINV1, was suppressed in male sterile plants, which were supposed to supply the carbohydrates to the developing pollen [102]. Cloning research showed that gene Sicwinv1 associated with male sterility in sesame accumulates during the tetrad stage in the tapetum cells of the anther [102].

8. Genetics of Economic Traits

Yield is a complex trait, which is significantly affected by numerous loci that may interact with each other and with the environment or crop management [103]. QTL for yield components has been mapped on various linkage groups. In a study, major QTLs for seed size and seed coat color on LG04 and LG11 were mapped. There were about 155 candidate genes associated with seed coat and size in these linkage groups [103]. The genome-wide association study (GWAS) approach enabled the mapping of QTL in sesame with precision due to high-density markers within the genome. This GWAS approach has been instrumental in uncovering the genetic architecture of oil content and fatty acid composition in sesame seeds [55].
Whole-genome sequencing allowed high-density markers such as SNP or SSR to map quantitative traits of sesame [103,104]. For instance, a study in sesame used 2159 SNP markers on 13 linkage groups with a marker distance of about 0.99 cM [103]. About 83,135 non-redundant SSR marker positions and motif sequences have been published and obtained from 151 published genomic sequences [104]. It enabled the mapping of “hotspots” within the sesame genome that were related to various economically important traits. In another study, association mapping with a high-density linkage map of 19,309 markers revealed 84 QTLs associated with yield components and seed mineral components [92]. Thirteen QTLs on 7 LG and 17 QTLs on 10 LG were linked with yield components which were mapped through 1230 markers in the recombinant inbred line (RIL) population [105]. High-throughput phenotyping technologies have been combined with SNP mapping to characterize labor-intensive traits with more precision and greater efficiency. For example, near-infrared spectroscopy (NIRS) has been recommended to determine several seed-related traits such as fatty acid and oil contents, which may provide a rapid alternative to other techniques such as gas chromatography or soxhlet apparatus, and it may facilitate the screening of large breeding populations [106].
A study comprising 3528 high-density specific locus amplified fragment (SLAF) markers having an average distance of 0.37 cM was used to map quantitative traits. SLAF makers mapped 46 significant QTLs for seven yield components across four environments [107]. Association mapping revealed 23 stable QTLs, which were detected in all environments. Selectable QTLs were mostly in the Chinese line “Yuzhi4” [108]. Furthermore, integrated multi-omics data from genomics, transcriptomics, and metabolomics, have provided a more comprehensive understanding of sesame’s molecular mechanisms underlying complex traits. This systems biology approach has revealed intricate gene regulatory networks involved in oil biosynthesis and stress responses [28].
Association genetics using a whole-genome re-sequencing strategy with 1354 markers over 12 chromosomes identified QTL related with water-stress-related traits such as root length and relative root length [108]. There was significant genetic variation for yield components such as capsule size, capsule number, and seed-size-related traits. A QTL qLS15-1 affecting the leaf size and development was mapped on LG15 [109]. Furthermore, this QTL (qLS15-1) was subjected to the sequencing of QTLs and their transcriptome analyses led to the identification of SIN_1004875, SIN_1004882, and SIN_1004883 I as candidate genes [109]. These QTLs were stable across environments, were in LG15, and affected leaf size and development [109]. Their sequencing and transcriptome analyses also confirmed these candidate genes [109].
Capsule number and length are important yield components in sesame and may dictate the yield potential of an accession. A study revealed two novel candidate genes (SiLPT3 and SiACS8c) affecting capsule number and capsule length. In another study, a high-density linkage map of 20,294 SNP markers led to identifying major QTL on LG2 for flowering date and yield components [110]. A significant QTL was related to the candidate gene SIN_1019016 on chromosome 10, gave Phytophthora blight resistance, and explained up to 13.34% of its phenotypic variation [111]. The advent of genome editing, particularly CRISPR-Cas9, has opened new avenues for targeted modification of sesame traits. While still in its early stages for sesame, this technology holds promise for precisely manipulating genes controlling key agronomic traits, thereby accelerating the breeding process [75].
Genomic selection (GS) is a transformative approach in plant breeding. It enhances significantly the efficiency of breeding programs through the incorporation of vast genomic data. Unlike traditional marker-assisted selection, which focuses narrowly on specific genes or QTL, GS leverages comprehensive whole-genome information. This methodology allows breeders to predict the breeding values of plants by analyzing numerous genetic markers that are distributed throughout the genome. As a result, GS is particularly advantageous for improving complex multigenic traits, such as yield, oil content, and host plant resistance to pathogens.
Recent research conducted by Sabag et al. [112] highlights the potential of GS in sesame breeding. Their study demonstrated that single-environment analyses yielded moderate to high genomic prediction accuracy for various agronomic traits. Importantly, the authors concluded that employing multi-environmental trial data could further elevate genomic prediction accuracy, thus enhancing the development of sesame cultivars better suited to semi-arid climates. This finding underscores the importance of environmental context in genomic selection, suggesting that integrating diverse growing conditions can lead to more robust predictive models. Despite promising developments, it is noteworthy that comprehensive research on genomic selection and prediction specifically within sesame has been limited. As demand for resilient and high-yielding sesame cultivars grows, the application of GS offers a valuable avenue for breeders. By increasing the accuracy of trait predictions and accelerating the breeding process, genomic selection holds the potential to significantly boost genetic gains in sesame, ensuring a more sustainable agricultural future. As research in this area continues to expand, the incorporation of genomic methods and tools will undoubtedly play a pivotal role in advancing global crop improvement efforts. Hence, whole-genome sequencing, genomic selection, genomic prediction, and GWAS pave the way for breeding sesame in the 21st century.

9. Prospects for Sesame Improvement

Prospects for sesame improvement lie in integrating crossbreeding with advanced methods such as marker-assisted selection, genomic selection, and gene editing. Recent advances in high-throughput phenotyping and genotyping have significantly accelerated sesame improvement. For example, unmanned aerial vehicles (UAVs) equipped with multispectral sensors now allow for the rapid assessment of large-scale field trials, thus enabling more efficient selection of superior genotypes that will expand the area of this crop for new germplasm with adaptability to diverse environmental conditions. On the other hand, sesame has low yield potential due to poor response to farm inputs. Sesame accessions having characteristics such as determinate growth, semi-dwarf stature, and lodging resistance are required to improve response to farm inputs and the same ideotype for mechanized harvesting. Moreover, accessions with indehiscence capsules should be bred to reduce yield loss during harvesting. Additionally, developing sesame cultivars with synchronized maturity is another critical breeding objective, as it would significantly improve the efficiency of mechanical harvesting and reduce post-harvest losses. Some of the work has already been conducted, and germplasm accessions with novel agronomic traits were developed through mutation breeding, selection, or hybridization. The application of CRISPR-Cas9 technology holds promise for targeted gene editing to develop high-yielding, pest-resistant, and climate-resilient sesame cultivars. For example, CRISPR-Cas9 could modify genes involved in lignin biosynthesis, potentially developing cultivars with improved content of oil and easier extraction. Furthermore, expanding transcriptomic and proteomic research will enhance our understanding of the regulatory mechanisms controlling important agronomic traits. Hence, integrating omics approaches with metabolomics data can offer a comprehensive view of the molecular pathways underlying sesame’s nutritional and medicinal properties, potentially opening new avenues for value-added product development. However, sesame cultivation in developing or resource-poor countries may have seldom access to advanced techniques. Hence, international collaborative efforts among institutes, researchers, breeders, and policymakers are essential to address sesame cultivation’s challenges and exploit its full potential as a sustainable oilseed crop. Climate-smart sesame cultivars must be developed to withstand extreme weather events and utilize resources more efficiently to ensure the crop’s sustainability in the face of climate change. Thus, this requires a multidisciplinary approach that combines crossbreeding techniques, genomic tools, and advanced agronomic practices.

10. Conclusions

In summary, sesame provides valuable opportunities for oil production and medicinal uses. However, its cultivation is hindered by several factors such as low yields, poor input response, and susceptibility to pests and diseases. Recent advances in molecular breeding provide new ways to overcome these challenges and improve the productivity and profitability of sesame farming. Researchers and breeders can develop superior sesame cultivars by utilizing the extensive genetic diversity found in sesame germplasm and employing advanced breeding techniques. These cultivars are designed to meet growers’ and consumers’ demands worldwide, thus enhancing the crop’s utility and market potential. This comprehensive approach will promote the sustainable growth of the sesame industry, thereby making sesame a more resilient and valuable crop for future generations.

Author Contributions

Conceptualization, S.R. and T.B.; Software, Y.K., S.R. and T.B.; Data analysis, all coauthors; Writing—original draft, Y.K., S.R., T.B., M.E., A.G., M.R. and R.O.; Writing—review and editing, all coauthors; Funding acquisition, R.O. All authors have read and agreed to the published version of the manuscript.

Funding

A.G. and R.O. acknowledge grant funding CTS 23:2654 from the Carl Trygger Foundation for the project “Fine mapping of plant height and seed coat color quantitative trait loci of sesame (Sesamum indicum L.) using single nucleotide polymorphisms”.

Acknowledgments

We would like to thank the Department of Plant Breeding and Genetics, University of Agriculture, Faisalabad, Pakistan, and the Swedish University of Agricultural Sciences for their support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Onkware, A.; Were, B.A.; Gudu, S.; Nyongesa, B.O.; Dangasuk, O.G. Genetic Relationship between sesame (Sesamum indicum L.) and related wild species based on chromosome counts and isozyme markers. Afr. J. Agric. Res. 2015, 9, 1052–1060. [Google Scholar]
  2. Elsafy, M. Status of Sesame Breeding. In Faculty of Landscape Architecture, Horticulture and Crop Production Science; Swedish University of Agricultural Sciences: Alnarp, Sweden, 2023; p. 2023:2. [Google Scholar]
  3. Canton, H. Food and Agriculture Organization of the United Nations—FAO. In The Europa Directory of International Organizations 2021; Routledge: London, UK, 2021; pp. 297–305. [Google Scholar]
  4. FAO. Statistical Yearbook 2022; Food and Agricultural Organization: Rome, Italy, 2022. [Google Scholar]
  5. Wacal, C.; Ogata, N.; Sasagawa, D.; Handa, T.; Basalirwa, D.; Acidri, R.; Ishigaki, T.; Yamamoto, S.; Nishihara, E. Seed yield, crude protein and mineral nutrient contents of sesame during a two-year continuous cropping on upland field converted from a paddy. Field Crops Res. 2019, 240, 125–133. [Google Scholar] [CrossRef]
  6. Singh, A.K.; Srinivasan, K.; Dhillon, B.; Mehra, K. Plant Genetic Resources Management in India: A Historical Perspective. In Hundred Years of Plant Genetic Resources Management in India; Singh, A.K., Srinivasan, K., Saxena, S., Dhillon, B.S., Eds.; NBPGR (ICAR): New Delhi, India, 2007; pp. 29–48. [Google Scholar]
  7. Wei, X.; Gong, H.; Yu, J.; Liu, P.; Wang, L.; Zhang, Y.; Zhang, X. SesameFG: An integrated database for the functional genomics of sesame. Sci. Rep. 2017, 7, 2342. [Google Scholar] [CrossRef] [PubMed]
  8. Bedigian, D. Slimy leaves and oily seeds: Distribution and use of wild relatives of sesame in Africa. Econ. Bot. 2004, 58, S3–S33. [Google Scholar] [CrossRef]
  9. Azeez, M.A.; Morakinyo, J.A. Genetic diversity of fatty acids in sesame and its relatives in Nigeria. Eur. J. Lipid Sci. Technol. 2011, 113, 238–244. [Google Scholar] [CrossRef]
  10. Bedigian, D. Characterization of sesame (Sesamum indicum L.) germplasm: A critique. Genet. Resour. Crop Evol. 2010, 57, 641–647. [Google Scholar]
  11. Kole, C. Wild Crop Relatives: Genomic and Breeding Resources; Springer: Cham, Switzerland, 2011. [Google Scholar]
  12. Yang, M.M.; Liu, H.Y.; Zhou, T.; Qu, H.H.; Yang, Y.X.; Wei, X.; Zuo, Y.; Zhao, Y.Z. Production and identification of F1 interspecific hybrid between Sesamum indicum and wild relative S. indicatum. Sci. Agric. Sinica 2017, 50, 1763–1771. [Google Scholar]
  13. Stavridou, E.; Lagiotis, G.; Kalaitzidou, P.; Grigoriadis, I.; Bosmali, I.; Tsaliki, E.; Tsiotsiou, S.; Kalivas, A.; Ganopoulos, I.; Madesis, P. Characterization of the genetic diversity present in a diverse sesame landrace collection based on phenotypic traits and EST-SSR markers coupled with an HRM analysis. Plants 2021, 10, 656. [Google Scholar] [CrossRef]
  14. Wei, X.; Wang, L.; Zhang, Y.; Qi, X.; Wang, X.; Ding, X.; Zhang, J.; Zhang, X. Development of simple sequence repeat (SSR) markers of sesame (Sesamum indicum) from a genome survey. Molecules 2014, 19, 5150–5162. [Google Scholar] [CrossRef]
  15. Nyongesa, B.O.; Were, B.A.; Gudu, S.; Dangasuk, O.G.; Onkware, A.O. Genetic diversity in cultivated sesame (Sesamum indicum L.) and related wild species in East Africa. J. Crop. Sci. Biotechnol. 2013, 16, 9–15. [Google Scholar] [CrossRef]
  16. Wang, M.; Huang, J.; Liu, S.; Liu, X.; Li, R.; Luo, J.; Fu, Z. Improved assembly and annotation of the sesame genome. DNA Res. 2022, 29, dsac041. [Google Scholar] [CrossRef] [PubMed]
  17. Purru, S.; Sahu, S.; Rai, S.; Rao, A.R.; Bhat, K.V. GinMicrosatDb: A genome-wide microsatellite markers database for sesame (Sesamum indicum L.). Physiol. Mol. Biol. Plants 2018, 24, 929–937. [Google Scholar] [CrossRef] [PubMed]
  18. Dutta, D.; Awon, V.K.; Gangopadhyay, G. Transcriptomic dataset of cultivated (Sesamum indicum), wild (S. mulayanum), and interspecific hybrid sesame in response to induced Macrophomina phaseolina infection. Data Brief 2020, 33, 106448. [Google Scholar] [CrossRef] [PubMed]
  19. Mondal, N.; Bhat, K.V.; Srivastava, P.S.; Sen, S.K. Effects of domestication bottleneck and selection on fatty acid desaturases in Indian sesame germplasm. Plant Genet. Resour. 2016, 14, 81–90. [Google Scholar] [CrossRef]
  20. Wang, Z.; Zhou, F.; Tang, X.; Yang, Y.; Zhou, T.; Liu, H. Morphology and SSR markers-based genetic diversity analysis of sesame (Sesamum indicum L.) cultivars released in China. Agriculture 2023, 13, 1885. [Google Scholar] [CrossRef]
  21. Maurya, R.; Singh, S.; Babu, Y.S.; Khan, F.N.; Nawade, B.; Vishwakarma, H.; Kumar, A.; Yadav, R.; Jalli, R.; Angamuthu, M.; et al. Molecular diversity studies and core development in sesame germplasm (Sesamum indicum L.) using SSR markers. Plant Mol. Biol. Rep. 2024. [Google Scholar] [CrossRef]
  22. Mesfer Alshamrani, S.; Safhi, F.A.; Alshaya, D.S.; Ibrahim, A.A.; Mansour, H.; El Moneim, D.A. Genetic diversity using biochemical, physiological, karyological and molecular markers of Sesamum indicum L. Front. Genet. 2022, 13, 1035977. [Google Scholar] [CrossRef]
  23. Teklu, D.H.; Shimelis, H.; Tesfaye, A.; Shayanowako, A.I.T. Analyses of genetic diversity and population structure of sesame (Sesamum indicum L.) germplasm collections through seed oil and fatty acid compositions and SSR markers. J. Food Composition Anal. 2022, 110, 104545. [Google Scholar] [CrossRef]
  24. Tesfaye, T.; Tesfaye, K.; Keneni, G.; Ziyomo, C.; Alemu, T. Genetic diversity of sesame (Sesamum indicum L) using high throughput diversity array technology. J. Crop. Sci. Biotechnol. 2022, 25, 359–371. [Google Scholar] [CrossRef]
  25. Seay, D.; Szczepanek, A.; De La Fuente, G.N.; Votava, E.; Abdel-Haleem, H. Genetic diversity and population structure of a large USDA sesame collection. Plants 2024, 13, 1765. [Google Scholar] [CrossRef]
  26. Ruperao, P.; Bajaj, P.; Yadav, R.; Angamuthu, M.; Subramani, R.; Rai, V.; Tiwari, K.; Rathore, A.; Singh, K.; Singh, G.P.; et al. Double-digest restriction-associated DNA sequencing-based genotyping and its applications in sesame germplasm management. Plant Genome 2024, 17, e20447. [Google Scholar] [CrossRef] [PubMed]
  27. Kurt, C. Variation in oil content and fatty acid composition of sesame accessions from different origins. Grasas Y Aceites 2018, 69, e241. [Google Scholar] [CrossRef]
  28. Zhang, Y.-P.; Zhang, Y.-Y.; Thakur, K.; Zhang, F.; Hu, F.; Zhang, J.-G.; Wei, P.-C.; Wei, Z.-J. Integration of miRNAs, degradome, and transcriptome omics uncovers a complex regulatory network and provides insights into lipid and fatty acid synthesis during sesame seed development. Front. Plant Sci. 2021, 12, 709197. [Google Scholar] [CrossRef] [PubMed]
  29. Kancharla, P.K.; Arumugam, N. Variation of oil, sesamin, and sesamolin content in the germplasm of the ancient oilseed crop Sesamum indicum L. J. Amer. Oil Chem. Soc. 2020, 97, 475–483. [Google Scholar] [CrossRef]
  30. Mohammed, F.; Abdulwali, N.; Guillaume, D.; Tenyang, N.; Ponka, R.; Al-Gadabi, K.; Bchitou, R.; Abdullah, A.H.; Naji, K.M. Chemical composition and mineralogical residence of sesame oil from plants grown in different Yemeni environments. Microchem. J. 2018, 140, 269–277. [Google Scholar] [CrossRef]
  31. Zeb, A.; Muhammad, B.; Ullah, F. Characterization of sesame (Sesamum indicum L.) seed oil from Pakistan for phenolic composition, quality characteristics and potential beneficial properties. J. Food Meas. Charact. 2017, 11, 1362–1369. [Google Scholar] [CrossRef]
  32. Asghar, A.; Majeed, M.N. Chemical characterization and fatty acid profile of different sesame verities in Pakistan. Am. J. Sci. Ind. Res. 2013, 4, 540–545. [Google Scholar]
  33. Carvalho, R.H.R.; Galvão, E.L.; Barros, J.Â.C.; Conceição, M.M.; Sousa, E.M.B.D. Extraction, fatty acid profile and antioxidant activity of sesame extract (Sesamum Indicum L.). Brazil. J. Chem. Eng. 2012, 29, 409–420. [Google Scholar] [CrossRef]
  34. Hama, J.R. Comparison of fatty acid profile changes between unroasted and roasted brown sesame (Sesamum indicum L.) seeds oil. Intl. J. Food Prop. 2017, 20, 957–967. [Google Scholar] [CrossRef]
  35. Uzun, B.; Arslan, Ç.; Furat, Ş. Variation in fatty acid compositions, oil content and oil yield in a germplasm collection of sesame (Sesamum indicum L.). J. Am. Oil Chem. Soc. 2008, 85, 1135–1142. [Google Scholar] [CrossRef]
  36. Awasthi, C.P.; Sood, S.; Acharya, M.K.; Sharma, N. Biochemical composition and fatty acid profile of some promising sesame (Sesamum indicum L.) genotypes. Indian J. Agric. Biochem. 2006, 9, 67–70. [Google Scholar]
  37. Santos, H.O.; Earnest, C.P.; Tinsley, G.M.; Izidoro, L.F.; Macedo, R.C. Small dense low-density lipoprotein-cholesterol (sdLDL-C): Analysis, effects on cardiovascular endpoints and dietary strategies. Prog. Cardiovasc. Dis. 2020, 63, 503–509. [Google Scholar] [CrossRef] [PubMed]
  38. Rauf, S.; Jamil, N.; Tariq, S.A.; Khan, M.; Kausar, M.; Kaya, Y. Progress in modification of sunflower oil to expand its industrial value. J. Sci. Food Agric. 2017, 97, 1997–2006. [Google Scholar] [CrossRef] [PubMed]
  39. Huang, H.; Cui, T.; Zhang, L.; Yang, Q.; Yang, Y.; Xie, K.; Fan, C.; Zhou, Y. Modifications of fatty acid profile through targeted mutation at BnaFAD2 gene with CRISPR/Cas9-mediated gene editing in Brassica napus. Theor. Appl. Genet. 2020, 133, 2401–2411. [Google Scholar] [CrossRef]
  40. Lee, K.-R.; Jeon, I.; Yu, H.; Kim, S.-G.; Kim, H.-S.; Ahn, S.-J.; Lee, J.; Lee, S.-K. Increasing monounsaturated fatty acid contents in hexaploid Camelina sativa seed oil by FAD2 gene knockout using CRISPR-Cas9. Front. Plant Sci. 2021, 12, 702930. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Gong, H.; Cui, X.; Gao, C.; Li, N.; Pu, Y.; Zhang, X.; Zhao, J. Integrated lipidomic and transcriptomic analyses reveal the mechanism of lipid biosynthesis and accumulation during seed development in sesame. Front. Plant Sci. 2024, 14, 1211040. [Google Scholar] [CrossRef]
  42. Zhang, P.; Li, F.; Tian, Y.; Wang, D.; Fu, J.; Rong, Y.; Wu, Y.; Gao, T.; Zhang, H. Transcriptome analysis of sesame (Sesamum indicum L.) reveals the lncrna and mrna regulatory network responding to low nitrogen stress. Intl. J. Mol. Sci. 2024, 25, 5501. [Google Scholar]
  43. Salas, J.J.; Bootello, M.A.; Martínez-Force, E.; Calerón, M.V.; Garcés, R. High stearic sunflower oil: Latest advances and applications. OCL 2021, 28, 35. [Google Scholar] [CrossRef]
  44. Dar, A.A.; Choudhury, A.R.; Kancharla, P.K.; Arumugam, N. The FAD2 gene in plants: Occurrence, regulation, and role. Front. Plant Sci. 2017, 8, 1789. [Google Scholar] [CrossRef]
  45. Jin, U.-H.; Lee, J.-W.; Chung, Y.-S.; Lee, J.-H.; Yi, Y.-B.; Kim, Y.-K.; Hyung, N.-I.; Pyee, J.-H.; Chung, C.-H. Characterization and temporal expression of a ω-6 fatty acid desaturase cDNA from sesame (Sesamum indicum L.) seeds. Plant Sci. 2001, 161, 935–941. [Google Scholar] [CrossRef]
  46. Chen, Z.; Tonnis, B.; Morris, B.; Wang, R.B.; Zhang, A.L.; Pinnow, D.; Wang, M.L. Variation in seed fatty acid composition and sequence divergence in the FAD2 gene coding region between wild and cultivated sesame. J. Agric. Food Chem. 2014, 62, 11706–11710. [Google Scholar] [CrossRef]
  47. Kim, M.J.; Kim, H.; Shin, J.S.; Chung, C.-H.; Ohlrogge, J.B.; Suh, M.C. Seed-specific expression of sesame microsomal oleic acid desaturase is controlled by combinatorial properties between negative cis-regulatory elements in the SeFAD2 promoter and enhancers in the 5′-UTR intron. Mol. Genet Genom. 2006, 276, 351–368. [Google Scholar] [CrossRef] [PubMed]
  48. Kavak, H.; Boydak, E. Trends of sudden wilt syndrome in sesame plots irrigated with delayed intervals. Afr. J. Microbiol. Res. 2011, 5, 1837–1841. [Google Scholar]
  49. Kavak, H.; Boydak, E. Screening of the resistance levels of 26 sesame breeding lines to Fusarium wilt disease. Plant Path. J. 2006, 5, 157–160. [Google Scholar] [CrossRef]
  50. Wang, L.; Zhang, Y.; Qi, X.; Li, D.; Wei, W.; Zhang, X. Global gene expression responses to waterlogging in roots of sesame (Sesamum indicum L.). Acta Physiol. Plant. 2012, 34, 2241–2249. [Google Scholar] [CrossRef]
  51. Mahmoud, S.A.; El-Sharkawy, E.S.; Emam, S. Breeding sesame for resistance to charcoal rot caused by Macrophomina phaseolina. Intl. J. Agric. Sci. 2024, 6, 18–35. [Google Scholar] [CrossRef]
  52. Wang, L.; Zhang, Y.; Zhu, X.; Zhu, X.; Li, D.; Zhang, X.; Gao, Y.; Xiao, G.; Wei, X.; Zhang, X. Development of an SSR-based genetic map in sesame and identification of quantitative trait loci associated with charcoal rot resistance. Sci. Rep. 2017, 7, 8349. [Google Scholar] [CrossRef]
  53. Baath, G.S.; Kakani, V.G.; Northup, B.K.; Gowda, P.H.; Rocateli, A.C.; Singh, H. Quantifying and modeling the influence of temperature on growth and reproductive development of sesame. J. Plant Growth Regul. 2022, 41, 143–152. [Google Scholar] [CrossRef]
  54. Kim, M.-S.; Shim, K.-B.; Park, S.-H.; Kim, K.-S. Changes in cuticular waxes of developing leaves in sesame (Sesamum indicum L.). J. Crop. Sci. Biotechnol. 2009, 12, 161–167. [Google Scholar] [CrossRef]
  55. Zhou, R.; Dossa, K.; Li, D.; Yu, J.; You, J.; Wei, X.; Zhang, X. Genome-wide association studies of 39 seed yield-related traits in sesame (Sesamum indicum L.). Intl. J. Molec. Sciences 2018, 19, 2794. [Google Scholar] [CrossRef]
  56. Song, Q.; Joshi, M.; Wang, S.; Johnson, C.D.; Joshi, V. Comparative analysis of root transcriptome profiles of sesame (Sesamum indicum L.) in response to osmotic stress. J. Plant Growth Regul. 2021, 40, 1787–1801. [Google Scholar] [CrossRef]
  57. Zhou, R.; Liu, P.; Li, D.; Zhang, X.; Wei, X. Photoperiod response-related gene SiCOL1 contributes to flowering in sesame. BMC Plant Biol. 2018, 18, 343. [Google Scholar] [CrossRef] [PubMed]
  58. Li, D.; Liu, P.; Yu, J.; Wang, L.; Dossa, K.; Zhang, Y.; Zhou, R.; Wei, X.; Zhang, X. Genome-wide analysis of WRKY gene family in the sesame genome and identification of the WRKY genes involved in responses to abiotic stresses. BMC Plant Biol. 2017, 17, 152. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, L.; Dossa, K.; You, J.; Zhang, Y.; Li, D.; Zhou, R.; Yu, J.; Wei, X.; Zhu, X.; Jiang, S.; et al. High-resolution temporal transcriptome sequencing unravels ERF and WRKY as the master players in the regulatory networks underlying sesame responses to waterlogging and recovery. Genomics 2021, 113, 276–290. [Google Scholar] [CrossRef]
  60. Wang, Y.; Zhang, Y.; Zhou, R.; Dossa, K.; Yu, J.; Li, D.; Liu, A.; Mmadi, M.A.; Zhang, X.; You, J. Identification and characterization of the bZIP transcription factor family and its expression in response to abiotic stresses in sesame. PLoS ONE 2018, 13, e0200850. [Google Scholar] [CrossRef] [PubMed]
  61. Chowdhury, S.; Basu, A.; Kundu, S. Overexpression of a new osmotin-like protein gene (SindOLP) confers tolerance against biotic and abiotic stresses in sesame. Front. Plant Sci. 2017, 8, 410. [Google Scholar] [CrossRef]
  62. Dossa, K.; Li, D.; Wang, L.; Zheng, X.; Liu, A.; Yu, J.; Wei, X.; Zhou, R.; Fonceka, D.; Diouf, D.; et al. Transcriptomic, biochemical and physio-anatomical investigations shed more light on responses to drought stress in two contrasting sesame genotypes. Sci. Rep. 2017, 7, 8755. [Google Scholar] [CrossRef]
  63. Baghery, M.A.; Kazemitabar, S.K.; Dehestani, A.; Mehrabanjoubani, P. Sesame (Sesamum indicum L.) response to drought stress: Susceptible and tolerant genotypes exhibit different physiological, biochemical, and molecular response patterns. Physiol. Mol. Biol. Plants 2023, 29, 1353–1369. [Google Scholar]
  64. Su, R.; Dossou, S.S.; Dossa, K.; Zhou, R.; Liu, A.; Zhong, Y.; Fang, S.; Zhang, X.; Wu, Z.; You, J. Genome-wide characterization and identification of candidate ERF genes involved in various abiotic stress responses in sesame (Sesamum indicum L.). BMC Plant Biol. 2022, 22, 256. [Google Scholar]
  65. Zhang, T.D.; Du, Z.W.; Mei, H.X.; Liu, Y.Y.; Cui, C.Q.; Hu, C.D.; Ma, Q.R.; Zheng, Y.Z. Genetic analysis of flowering time with the mixed major gene plus polygene inheritance model in sesame. J. Henan Agric. Sci. 2019, 48, 67–72. [Google Scholar]
  66. Shim, K.B.; Goo, B.I.; Shin, M.N.; Jeon, W.T. Effect of temperature and daylength on flowering and growth characteristics. Korean J. Crop Sci. 2020, 65, 241–247. [Google Scholar]
  67. Yu, M.; Zhou, Q.F.; Zhao, J.G.; Zhang, G.G. Research advances on induced mutation breeding of sesame. Biotechnol. Bull. 2017, 33, 8. [Google Scholar]
  68. Sandhiya, V.; Kumar, M.; Parameswari, C.; Vanniarajan, C.; Kumaravadivel, N.; Sakthivel, N.; Badigannavar, A.M. Determination of optimum dose of chemical mutagen for large scale seed treatment of white seeded sesame (Sesamum indicum L.) varieties. Electron. J. Plant Breed. 2020, 11, 238–242. [Google Scholar]
  69. Parthasarathi, G.; Pillai, M.A.; Kannan, R.; Kumari, S.M.P.; Binodh, A.K. Optimal lethal dose determination for gamma rays and EMS induced mutagenesis in TMV7 and SVPR1 sesame (Sesamum indicum L.) varieties. Curr. J. Appl. Sci. Technol. 2020, 39, 136–144. [Google Scholar] [CrossRef]
  70. Bhuiyan, M.S.H.; Malek, M.A.; Sarkar, M.A.; Islam, M.; Akram, M.W. Genetic variance and performance of sesame mutants for yield contributing characters. Malays. J. Sustain. Agric. 2019, 3, 27–30. [Google Scholar] [CrossRef]
  71. Patil, M.K.; Lokesha, R.; Diwan, J. Genetic divergence of advanced mutant breeding lines in sesame (Sesamum indicum L.) assessed through D2 statistics. Intl. J. Curr. Microbiol. Appl. Sci. 2018, 6, 3133–3139. [Google Scholar] [CrossRef]
  72. Qureshi, M.; Langham, D.R.; Lucas, S.J.; Uzun, B.; Yol, E. Breeding history for shattering trait in sesame: Classic to genomic approach. Mol. Biol. Rep. 2022, 49, 7185–7194. [Google Scholar] [CrossRef]
  73. Ashri, A. Sesamum indicum. In Handbook of Flowering; CRC Press: Boca Raton, FL, USA, 2019; pp. 309–312. [Google Scholar]
  74. Yol, E.; Basak, M.; Kızıl, S.; Lucas, S.J.; Uzun, B. A high-density SNP genetic map construction using ddRAD-seq and mapping of capsule shattering trait in sesame. Front. Plant Sci. 2021, 12, 679659. [Google Scholar] [CrossRef]
  75. You, J.; Li, D.; Yang, L.; Dossou, S.S.K.; Zhou, R.; Zhang, Y.; Wang, L. CRISPR/Cas9-mediated efficient targeted mutagenesis in sesame (Sesamum indicum L.). Front. Plant Sci. 2022, 13, 935825. [Google Scholar] [CrossRef]
  76. Weldemichael, M.Y.; Baryatsion, Y.T.; Sbhatu, D.B.; Abraha, G.G.; Juhar, H.M.; Kassa, A.B.; Sibhatu, F.B.; Gebremedhn, H.M.; Gebrelibanos, T.S.; Mossa, M.M.; et al. Effect of sodium azide on quantitative and qualitative stem traits in the M2 generation of Ethiopian sesame (Sesamum indicum L.) genotypes. Sci. World J. 2021, 2021, 6660711. [Google Scholar] [CrossRef]
  77. Jayaramachandran, M.; Saravanan, S.; Motilal, A.; Prabhu, P.C.; Hepziba, S.J.; Swain, H.; Boopathi, N.M. Genetic improvement of a neglected and underutilised oilseed crop: Sesame (Sesamum indicum L.) through mutation breeding. Nucleus 2020, 63, 293–302. [Google Scholar] [CrossRef]
  78. Aristya, V.E.; Taryono, T.; Wulandari, R.A. Yield components of some sesame mutant populations induced by gamma irradiation. Bul. Tanam. Tembakau Serat Miny. Ind. 2019, 10, 64–71. [Google Scholar] [CrossRef]
  79. Kumari, V.; Singh, A.; Chaudhary, H.K.; Kumar, A.; Prasad, R.; Jambhulkar, S.; Sanju, S. Identification of Phytophthora blight resistant mutants through induced mutagenesis in sesame (Sesamum indicum L.). Indian Phytopath. 2019, 72, 71–77. [Google Scholar] [CrossRef]
  80. Ravichandran, V.; Jayakumar, S. Effect of mutagens on quantitative characters in M2 and M3 generation of sesame (Sesamum indicum L.). J. Pharmacogn. Phytochem. 2018, 42, 2833–2836. [Google Scholar]
  81. Gadri, Y.; Williams, L.E.; Peleg, Z. Tradeoffs between yield components promote crop stability in sesame. Plant Sci. 2020, 295, 110105. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, M.L.; Tonnis, B.; Li, X.; Morris, J.B. Generation of sesame mutant population by mutagenesis and identification of high oleate mutants by GC analysis. Plants 2023, 12, 1294. [Google Scholar] [CrossRef]
  83. Weldemichael, M.Y.; Bitima, T.D.; Abrha, G.T.; Tesfu, K.; Gebremedhn, H.M.; Kassa, A.B.; Kindeya, Y.B.; Mossa, M.M. Improving desirable agronomic traits of M2 lines on fourteen Ethiopian sesame (Sesamum indicum L.) genotypes using ethyl methyl sulphonate (EMS). PLoS ONE 2023, 18, e0287246. [Google Scholar] [CrossRef]
  84. Jhar, A.; Sabbagh, S.; Alkhattab, E.; Abrak, S.; Hatipoğlu, H.; Brennan, A.C. Optimising gamma irradiation seed treatment of sesame (Sesamum indicum L.) varieties for potential future application in mutation breeding. Crop Breed. Genet. Genom. 2024, 6, e240005. [Google Scholar]
  85. Miao, H.; Li, C.; Duan, Y.; Wei, L.; Ju, M.; Zhang, H. Identification of a Sidwf1 gene controlling short internode length trait in the sesame dwarf mutant dw607. Theor. Appl. Genet. 2020, 133, 73–86. [Google Scholar] [CrossRef]
  86. Zhang, H.; Miao, H.; Li, C.; Wei, L.; Duan, Y.; Ma, Q.; Kong, J.; Xu, F.; Chang, S. Ultra-dense SNP genetic map construction and identification of SiDt gene controlling the determinate growth habit in Sesamum indicum L. Sci. Rep. 2016, 6, 31556. [Google Scholar] [CrossRef]
  87. Yol, E. Inheritance of long and dense capsule characteristics in sesame. Turkish J. Field Crops 2017, 22, 8–13. [Google Scholar] [CrossRef]
  88. Liu, H.Y.; Zhou, F.; Li, J.; Yang, M.M.; Zhou, T.; Hao, G.C.; Zhao, Y.Z. Anatomical structure and photosynthetic characteristics of a yellow leaf mutant YL1 in sesame (Sesamum indicum L.). Acta Agron. Sin. 2017, 43, 1856–1863. [Google Scholar] [CrossRef]
  89. Yol, E.; Uzun, B. Inheritance of indehiscent capsule character, heritability and genetic advance analyses in the segregation generations of dehiscent x indehiscent capsules in sesame. J. Agric Sci. 2019, 25, 79–85. [Google Scholar] [CrossRef]
  90. Ju, M.; Li, G.; Tian, Q.; Cao, H.; Ma, Q.; Duan, Y.; Guo, H.; Zhang, Z.; Huang, Y.; Wang, H.; et al. Deletion of a 1,049 bp sequence from the 5’ UTR upstream of the SiHEC3 gene induces a seed non-shattering mutation in sesame. J. Integr. Agric. 2024, 2, 2589–2604. [Google Scholar] [CrossRef]
  91. Zhang, H.; Miao, H.; Wei, L.; Li, C.; Duan, Y.; Xu, F.; Qu, W.; Zhao, R.; Ju, M.; Chang, S. Identification of a SiCL1 gene controlling leaf curling and capsule indehiscence in sesame via cross-population association mapping and genomic variants screening. BMC Plant Biol. 2018, 18, 296. [Google Scholar] [CrossRef]
  92. Teboul, N.; Gadri, Y.; Berkovich, Z.; Reifen, R.; Peleg, Z. Genetic architecture underpinning yield components and seed mineral–nutrients in sesame. Genes 2020, 11, 1221. [Google Scholar] [CrossRef]
  93. Liu, H.; Zhou, F.; Zhou, T.; Yang, Y.; Zhao, Y. Fine mapping of a novel male-sterile mutant showing wrinkled-leaf in sesame by BSA-Seq technology. Ind. Crops Prod. 2020, 156, 112862. [Google Scholar] [CrossRef]
  94. Duhoon, S. Exploitation of Heterosis for Raising Productivity in Sesame. In 4th International Crop Science Congress, Brisbane, Australia. 2004. Available online: http://www.agronomyaustraliaproceedings.org/images/sampledata/2004/poster/3/4/4/1899_duhoons.pdf (accessed on 14 September 2024).
  95. Zhao, Y.; Yang, M.; Wu, K.; Liu, H.; Wu, J.; Liu, K. Characterization and genetic mapping of a novel recessive genic male sterile gene in sesame (Sesamum indicum L.). Mol. Breed. 2013, 32, 901–908. [Google Scholar] [CrossRef]
  96. Liu, H.; Zhou, X.; Wu, K.; Yang, M.; Zhao, Y. Inheritance and molecular mapping of a novel dominant genic male-sterile gene in Sesamum indicum L. Mol. Breed. 2015, 35, 1–14. [Google Scholar] [CrossRef]
  97. Liu, H.; Zhou, F.; Zhou, T.; Yang, Y.; Zhao, Y. Cytological characterization and molecular mapping of a novel recessive genic male sterility in sesame (Sesamum indicum L.). PLoS ONE 2018, 13, e0204034. [Google Scholar] [CrossRef]
  98. Wu, K.; Liu, H.; Zuo, Y.; Yang, M.; Zhao, Y. Histological and transcriptional characterization of a novel recessive genic male sterility mutant in sesame (Sesamum indicum L.). Acta Physiol. Plant. 2014, 36, 421–431. [Google Scholar] [CrossRef]
  99. Prabakaran, A.; Rangasamy, S.S.; Ramalingam, R. Identification of cytoplasm-induced male sterility in sesame through wide hybridization. Curr. Sci. 1995, 68, 1044–1047. [Google Scholar]
  100. Liu, H.; Yang, M.; Wu, K.; Zhou, X.; Zhao, Y. Development, inheritance and breeding potential of a recessive genic male sterile line D248A in Sesame (Sesamum indicum L.). SpringerPlus 2013, 2, 268. [Google Scholar] [CrossRef] [PubMed]
  101. Zheng, Y.Z.; Wang, W.Q.; Mei, H.X.; Zhang, T.D. An analysis of chemical maintaining effect on genic male sterility in sesame (Sesamum indicum L.) I. The effect of CRA in restoring fertility for genic male sterility in sesame. Chin. J. Oil Crop Sci. 2000, 22, 13–15. [Google Scholar]
  102. Zhou, T.; Hao, G.; Yang, Y.; Liu, H.; Yang, M.; Zhao, Y. Sicwinv1, a cell wall invertase from sesame, is involved in anther development. J. Plant Growth Regulation 2019, 38, 1274–1286. [Google Scholar] [CrossRef]
  103. Du, H.; Zhang, H.; Wei, L.; Li, C.; Duan, Y.; Wang, H. A high-density genetic map constructed using specific length amplified fragment (SLAF) sequencing and QTL mapping of seed-related traits in sesame (Sesamum indicum L.). BMC Plant Biol. 2019, 19, 588. [Google Scholar] [CrossRef]
  104. Dossa, K. A physical map of important QTLs, functional markers and genes available for sesame breeding programs. Physiol. Mol. Biol. Plants 2016, 22, 613–619. [Google Scholar] [CrossRef]
  105. Wu, K.; Liu, H.; Yang, M.; Tao, Y.; Ma, H.; Wu, W.; Zuo, Y.; Zhao, Y. High-density genetic map construction and QTLs analysis of grain yield-related traits in sesame (Sesamum indicum L.) based on RAD-Seq technology. BMC Plant Biol. 2014, 14, 274. [Google Scholar] [CrossRef]
  106. Rife, T.W.; Courtney, C.; Hershberger, J.; Gore, M.A.; Neilsen, M.; Poland, J. Prospector: A mobile application for portable, high-throughput near-infrared spectroscopy phenotyping. Plant Phenome J. 2021, 4, e20024. [Google Scholar] [CrossRef]
  107. Mei, H.; Liu, Y.; Cui, C.; Hu, C.; Xie, F.; Zheng, L.; Du, Z.; Wu, K.; Jiang, X.; Zheng, Y.; et al. QTL mapping of yield-related traits in sesame. Mol. Breed. 2021, 41, 43. [Google Scholar] [CrossRef]
  108. Liang, J.; Sun, J.; Ye, Y.; Yan, X.; Yan, T.; Rao, Y.; Zhou, H.; Le, M. QTL mapping of PEG-induced drought tolerance at the early seedling stage in sesame usingwhole genome re-sequencing. PLoS ONE 2021, 16, e0247681. [Google Scholar] [CrossRef] [PubMed]
  109. Sheng, C.; Song, S.; Zhou, R.; Li, D.; Gao, Y.; Cui, X.; Tang, X.; Zhang, Y.; Tu, J.; Zhang, X.; et al. QTL-seq and transcriptome analysis disclose major QTL and candidate genes controlling leaf size in sesame (Sesamum indicum L.). Front. Plant Sci. 2021, 12, 580846. [Google Scholar] [CrossRef] [PubMed]
  110. Sabag, I.; Morota, G.; Peleg, Z. Genome-wide association uncovers the genetic architecture of tradeoff between flowering date and yield components in sesame. BMC Plant Biol. 2021, 21, 549. [Google Scholar] [CrossRef] [PubMed]
  111. Asekova, S.; Oh, E.; Kulkarni, K.P.; Siddique, M.I.; Lee, M.H.; Kim, J.I.; Lee, J.-D.; Kim, M.; Oh, K.-W.; Ha, T.-J.; et al. An integrated approach of QTL mapping and genome-wide association analysis identifies candidate genes for phytophthora blight resistance in sesame (Sesamum indicum L.). Front. Plant Sci. 2021, 12, 604709. [Google Scholar] [CrossRef]
  112. Sabag, I.; Bi, Y.; Peleg, Z.; Morota, G. Multi-environment analysis enhances genomic prediction accuracy of agronomic traits in sesame. Front. Genet. 2023, 14, 1108416. [Google Scholar] [CrossRef]
Figure 1. Flow chart of activities related to sesame breeding and germplasm conservation.
Figure 1. Flow chart of activities related to sesame breeding and germplasm conservation.
Plants 13 02662 g001
Figure 2. Cluster and phylogenetic analysis among various accessions of Glycine max, Sesamum indicum, and Helianthus annuus based on the FAD2 gene. Phylogenetic analysis was based on the sequence homology of FAD2 transcripts downloaded from gene bank https://www.ncbi.nlm.nih.gov/genbank/, accessed on 9 May 2024 and analyzed through MEGA X software.
Figure 2. Cluster and phylogenetic analysis among various accessions of Glycine max, Sesamum indicum, and Helianthus annuus based on the FAD2 gene. Phylogenetic analysis was based on the sequence homology of FAD2 transcripts downloaded from gene bank https://www.ncbi.nlm.nih.gov/genbank/, accessed on 9 May 2024 and analyzed through MEGA X software.
Plants 13 02662 g002
Figure 3. Sesame wilting syndrome after delayed irrigation (A), under high temperature (B), high-temperature effects on sesame causing stunted growth, flower shedding, and mal capsule formation (C), and sesame pod rotting in humid conditions (D). Phyllody disease causing the conversion of the capsule into a flower-like structure (E) and the sesame crop affected by the weed infestation (Cucumis callosus) causing the suppression of pod formation during the reproductive phase (F).
Figure 3. Sesame wilting syndrome after delayed irrigation (A), under high temperature (B), high-temperature effects on sesame causing stunted growth, flower shedding, and mal capsule formation (C), and sesame pod rotting in humid conditions (D). Phyllody disease causing the conversion of the capsule into a flower-like structure (E) and the sesame crop affected by the weed infestation (Cucumis callosus) causing the suppression of pod formation during the reproductive phase (F).
Plants 13 02662 g003aPlants 13 02662 g003b
Table 3. Transcriptomic profiling of sesame accessions for various stresses.
Table 3. Transcriptomic profiling of sesame accessions for various stresses.
Transcription FactorsResultsStress ToleranceReference
HD-Zip I-IV75% of SiHDz genes were differentially expressed in response to drought and salinity stressDrought and salinity[57]
SiWRKY65 genes were mapped on 15-linkage groupGrowth and development
Waterlogging and drought stress
[58]
ERF and WRKY47 genes related to waterlogging resistanceWaterlogging[59]
SibZIPsDifferential expression in response to abiotic stress Drought, waterlogging, osmotic, salt, and cold stress[60]
SindOLPExpression under various biotic and abiotic stressesHigh expression of ROS scavengers, chlorophyll contents, proline, and low lipid peroxidation[61]
fIRAK1/4Expression under drought stressPeroxidase, heat shock protein, interleukin protein, APETALA2/ethylene-responsive-element-binding protein, and mitogen-activated protein kinase[62]
GS1-2, PAL, CHS, CAB21,MYB54, MYB88 and NAC75Three nitrogen levelsLow nitrogen levels affect nitrogen stress tolerance genes through regulator lncRNA MSTRG.13854.1 which affects the MYB54 [42]
MnSOD1, MnSOD2, and PDHA-MDrought stressHigher expression of drought-related genes induced through manganese [63]
SiERFsWater stress and waterloggingSiERF23 and SiERF54 were induced by both stress conditions[64]
Table 5. Sources of male sterility in sesame.
Table 5. Sources of male sterility in sesame.
Male SterilitySourceEffectsReference
Genetic male sterility
95ms-5AB
Mutation in Yuzhi 4
by gamma rays Co60
Defective pollen, shriveled anthers, recessive gene Sms1[95]
GMS line (W1098A)Wild accession Yezhi2 (Sesamum mulayanum Nair)Dominant genetic male sterility pollen abortion due to abnormal tapetum[96]
RGMS (D248A)Spontaneous male sterility in cultivar ZuzhiRecessive male sterility. Ms gene may be selected by SB2993 and LG1-170[97]
RGMS95ms-5A-induced mutant vs. 95ms-5A.A total of 27 differentially expressed transcripts were identified in sterile vs. fertile buds. Eleven transcripts were involved in energy metabolism, signal transduction, and cell development[98]
CMSSesamum malabaricumInterspecific hybridization showed male sterility. Reciprocal effects were identified indicating cytoplasmic inheritance of organelles[99]
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

Rauf, S.; Basharat, T.; Gebeyehu, A.; Elsafy, M.; Rahmatov, M.; Ortiz, R.; Kaya, Y. Sesame, an Underutilized Oil Seed Crop: Breeding Achievements and Future Challenges. Plants 2024, 13, 2662. https://doi.org/10.3390/plants13182662

AMA Style

Rauf S, Basharat T, Gebeyehu A, Elsafy M, Rahmatov M, Ortiz R, Kaya Y. Sesame, an Underutilized Oil Seed Crop: Breeding Achievements and Future Challenges. Plants. 2024; 13(18):2662. https://doi.org/10.3390/plants13182662

Chicago/Turabian Style

Rauf, Saeed, Taiyyibah Basharat, Adane Gebeyehu, Mohammed Elsafy, Mahbubjon Rahmatov, Rodomiro Ortiz, and Yalcin Kaya. 2024. "Sesame, an Underutilized Oil Seed Crop: Breeding Achievements and Future Challenges" Plants 13, no. 18: 2662. https://doi.org/10.3390/plants13182662

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