Next Article in Journal
Agrobacterium tumefaciens-Mediated Genetic Transformation of Eclipta alba
Previous Article in Journal
Phytotoxicity of Two Bauhinia Species on Four Triticum aestivum Varieties in Laboratory Bioassay
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Genetic Diversity Studies: Insights from DNA Marker Analyses

by
Nongthombam Bidyananda
1,
Imlitoshi Jamir
2,
Karolina Nowakowska
3,
Vanlalrinchhani Varte
4,
Wagner A. Vendrame
5,
Rajkumari Sanayaima Devi
6 and
Potshangbam Nongdam
1,*
1
Department of Biotechnology, Manipur University, Canchipur 795003, Manipur, India
2
Department of Biotechnology, Nagaland University, Dimapur 797112, Nagaland, India
3
Section of Ornamental Plants, Institute of Horticultural Sciences, Warsaw University of Life Sciences (SGGW), Nowoursynowska 159, 02-776 Warsaw, Poland
4
Department of Neurology, McKnight Brain Institute, Norman Fixel Institute for Neurological Diseases, University of Florida, Gainesville, FL 32611, USA
5
Environmental Horticulture Department, University of Florida, Gainesville, FL 32611, USA
6
Department of Botany, Deen Dayal Upadhyaya College, University of Delhi, New Delhi 110078, India
*
Author to whom correspondence should be addressed.
Int. J. Plant Biol. 2024, 15(3), 607-640; https://doi.org/10.3390/ijpb15030046
Submission received: 1 June 2024 / Revised: 27 June 2024 / Accepted: 4 July 2024 / Published: 8 July 2024
(This article belongs to the Section Plant Ecology and Biodiversity)

Abstract

:
The plant adaptation response to a changing environment depends on the genetic diversity level it possesses. Genetic diversity and a thorough understanding of population indices are pivotal for decoding plant adaptation to dynamic environmental stressors. The development of polymerase chain reaction (PCR)-based molecular markers enables comprehensive population analyses and the precise detection of individuals and groups with unique genetic variations. Various molecular markers have been employed to assess genetic diversity, examine population structure, and delineate cluster patterns within and among populations. DNA markers revolutionize plant diversity studies by allowing detailed analyses of genetic variations, including economically significant trait-influencing genes. Despite their simplicity, they offer high reproducibility, ensuring accurate estimations of plant variation. Integrating multiple marker systems with advanced high-throughput sequencing techniques is poised to enhance the understanding and management of depleting plant genetic resources by providing a comprehensive picture of diversity at the genome-wide level. This review explores diverse molecular markers, elucidating their advantages and limitations, and highlights their impact on evaluating the genetic diversity and population structure of plants.

1. Introduction

Molecular markers have been widely used for plant genetic diversity and population genetics studies, essential for breeding and crop improvement, conservation, protection, introduction, and reintroduction of endangered and valuable plants [1,2,3]. They also enable the identification of new plant varieties and detect genetic changes from the known ones, providing valuable insights into the existing genetic variations within and between plant populations [4]. Genetic diversity is vital for adapting and adjusting plants to environmental changes [5]. The responses of plants and their adaptive abilities to climatic change depend on their genetic diversity levels [6]. Molecular markers offer essential insights into the variation in plant genetic composition and population structures, thereby playing vital roles in optimizing plant utilization and ensuring effective management [7]. They also assist in quickly identifying the wild relatives of cultivated plants, thereby providing the possibility of improving crop varieties with increased resistance to diseases and environmental stress [8]. With the recent increase in the understanding of plant genome sequences and the varied molecular roles of plant genes, the field of plant molecular genetics has been revolutionized, increasing its efficacy in the application of plant genetic variety study and breeding programs [9].
The morphological characters and cytological and ethnological parameters traditionally employed to estimate genetic diversity levels are not reliable and effective, as only a small part of the plant genome is represented by these traits, and they are easily affected by environmental factors [10]. They provide an incomplete picture of the complex genetic structure variation within and between species, genera, and plant groups [11]. Enhancements in biochemistry and molecular biology have led to the development of more powerful biochemical and molecular markers, overcoming the inherent limitations of traditional weak markers [12]. The novel invention of polymerase chain reaction (PCR) by Kary Mullis in 1983 has dramatically simplified plant genetic variation analyses by developing sophisticated DNA markers with high reproducibility [13]. Biochemical isozyme and DNA markers have been employed to determine genetic diversity, population structures, and phylogenetic relationships, identify cultivars, and construct genetic linkage maps in several plant species [14,15,16]. However, DNA markers are more effective and favorable than isozymes due to their abundance, dominant and co-dominant inheritances, and involvement of both express and non-express sequences specific to their gene locus [17]. They are also less dependent on environmental factors and can provide more detailed information about the underlying genomic variation in an organism. Recently, plant diversity studies with single marker types have not been considered complete without applying other markers [18]. The cumulative application of multiple markers imparts more accurate and reliable genetic diversity information, population structure, and cluster determination than using a single marker type [19,20]. The genetic variability within various plants has been effectively assessed by employing a combination of diverse markers [4,21,22]. The recent advancements in DNA sequencing techniques have further bolstered the efficacy of DNA markers in diversity studies, genome mapping, and crop improvement [23]. The present review aims to provide a comprehensive overview of different DNA markers, their advantages, limitations, and applications in genetic diversity, cluster resolution, and population structure studies of plants.

2. DNA Markers

DNA markers are fragments of DNA that reveal variation among organisms [24,25]. They may even be used to detect polymorphisms between different genotypes within a population based on the divergence in a particular DNA sequence [26,27]. DNA marker development may be based on either the coding or non-coding regions of the genes, and this may or may not necessitate prior sequence information [9]. Researchers continue to strive to establish more reliable and accurate markers such as AFLP, ISSR, SCoT, iPBS, DArT, ITS, SNP, and others, which are more efficient for estimating diversity than morphological and biochemical markers [28]. However, an ideal DNA marker should possess properties such as high polymorphism, dominant or co-dominant inheritance, frequent occurrence, independence from environmental and developmental conditions, high reproducibility, easy accessibility, and fast and cost-effective assays [10,17,29]. DNA markers may broadly be classified as hybridization- and PCR-based markers (Figure 1).

3. Application of Markers for Genetic Diversity and Population Structure Studies

The enormous utility of molecular markers in plant research is excellent. After their development, these markers can be used for DNA fingerprinting, genotype characterization, the construction of genetic linkage maps, quantitative trait locus (QTL) mapping, genetic homogeneity testing, and population and genetic diversity studies [30,31]. Earlier, a single marker system was employed to evaluate plant genetic variation. However, multiple markers have been recommended to obtain more reliable, accurate, and comprehensive information about the diversity level, genetic population structure, and cluster resolution of the investigated plant species [32,33,34,35]. The combined utilization of different marker systems ensures the validation of the results from plant genetic investigations derived from the analysis of each marker system [4]. The following flowchart depicts the basic steps of gel-based molecular markers explicitly designed for plant genetic diversity studies (Figure 2).

3.1. Utilisation of Single Molecular Marker Systems

Exploring plant genetic diversity is crucial for understanding their evolutionary history, adaptation, and potential for future breeding programs. Among the various approaches to ascertain genetic diversity and population structure is using a single marker system, which involves the utilization of only specific molecular markers.

3.1.1. Random Amplification of Polymorphic DNA (RAPD)

RAPDs are popular first-generation DNA markers based on the amplification of random DNA fragments in different loci of a DNA template with short, arbitrary primers and detect polymorphisms through amplified DNA fragments analyses after electrophoresis [36]. They are simple, quick, and cost-effective and have proven effective for evaluating genetic variations in both wild and domesticated plants [37]. Bibi [38] used seven selected RAPD primers out of fifteen to detect the best linseed variety, generating precise polymorphic bands. Dhakshanamoorthy [39] also utilized 20 polymorphic RAPD primers to identify mutants in Jatropha curcas treated with gamma rays and ethyl methane sulphonate (EMS). Fu [40] demonstrated their usefulness by resolving the genetic relationships among populations of the rare, endemic species of Changium smyrnioides through a genetic diversity evaluation. Zhang [41] applied 14 RAPD primers, generating 98 polymorphic bands to genetically characterize Coptis omeiensis. RAPD markers have been widely employed in elucidating plant variability and interconnections, deciphering population genetic structures, advancing crop development, and supporting conservation initiatives [42,43]. However, the markers have limitations arising from a short primer length, lower polymorphism levels and non-reproducible PCR outcomes, and the dominant mode of inheritance, rendering them less reliable and unsuitable for marker-assisted selection (MAS) and accurate diversity studies [44].

3.1.2. Restriction Fragment Length Polymorphism (RFLP)

The RFLP is a hybridization-based marker that employs restriction enzymes for DNA fragment generation and subsequent hybridization to target fragments using specific radio-labeled probes. The marker detects differences between individual genomes due to point mutation, insertion/deletion, translocation, inversion, and duplication by generating different-sized DNA fragments through restriction digestion [45,46]. RFLPs are locus-specific markers showing high reproducibility and reliability with co-dominant inheritance and offering easy data exchange between laboratories [47]. However, the main shortcomings of the markers include the necessity of a substantial amount of high-quality DNA, costly radioactive probes with prior sequence knowledge for their development, and the requirement for long and tedious Southern blotting techniques [10]. RFLPs have been used widely in genome mapping to select desirable genes, analyze polygenic characters, and determine genetic relationships and diversity between crop plants [48,49]. Chang [50] applied RFLP markers to design a linkage map for the nuclear genome of Arabidopsis thaliana, associating it with clones of the unknown gene function exhibiting the mutant phenotype and vice versa. Benchimol [51] demonstrated the effectiveness of RFLPs in assessing genetic diversity and allocating the genotypes from tropical maize populations into heterotic groups. However, they could not determine the line crosses from genetically diverse heterotic groups. The markers also proved to be efficient in revealing the differences among wild wheat relatives from regions separated by natural barriers [52]. The RFLP method has also been employed to construct genome maps for tomatoes, potatoes, and chili peppers [53,54].

3.1.3. Amplified Fragment Length Polymorphism (AFLP)

The AFLP markers developed by Vos [55] have been used widely for genetic diversity studies of closely related plant species. Applying the markers does not require an introductory sequence, making them a powerful tool for detecting polymorphisms at the DNA level [56]. AFLPs have shown higher reproducibility and sensitivity than RAPD and RFLP markers [24]. They have emerged as the markers of preference for several investigations due to the non-necessity of previous sequence information, greater abundance, and genome-wide, high polymorphism flexibility when used for molecular assays, with the potential for automation [57]. Al-Nadabi [58] identified six distinct citrus cultivars out of thirty-three by employing an AFLP fingerprint, providing phylogenetic information on citrus cultivars and their diversity in Oman. AFLPs revealed high genetic diversity between 26 different populations of Eupatorium adenophorum from China, suggesting the plant diversification rate was dependent on the invasion time and the geographical location [59]. Murariu [60] employed AFLP markers to assess wide genetic diversity in 60 maize landrace accessions from Romania, offering the potential for incorporation into the breeding programs. Bhattacharrya [61] highlighted the utility of AFLP markers in estimating diversity for effective conservation strategies, as demonstrated in Dendrobium thyrsiflorum, which exhibited high antioxidant activity without prior genetic information on linkage maps and QTL. AFLP markers provided efficient information on genome coverage, genetic variation, and phylogenetic relationships in the Origanum and Thymus species [62].
Furthermore, their application yielded insights into eco-geographic factors and genetic differentiation patterns influencing the variation and population structure of Elymus tangutorum in Western China [63]. AFLP markers were also instrumental in the genetic restoration and conservation efforts for Panicum turgidum in Saudi Arabia [64], identifying the optimal niche for Rhodiola rosea at an altitude of 3150–3250 m [65] and facilitating effective germplasm conservation of Rhododendron concinnum [66]. Despite numerous advantages, AFLP markers have limitations, including challenges in interpreting banding profiles, their dominant nature, and the potential occurrence of co-migrating bands during homology identification [67].

3.1.4. CAAT Box-Derived Polymorphism (CBDP)

CBDPs utilize the nucleotide sequence of the CAAT box of plant promoters, characterized by a consensus GGCCAATCT sequence, and play a vital role in transcription [68]. The marker is valuable for cultivar identification, distinguishing indigenous from introduced varieties, marker-assisted selection, and construction of linkage maps [69]. Moreover, CBDPs are easy to develop and provide reproducible profiles. The marker can be applied singly or combined with other markers to accurately determine population structure and specify the genetic relationships among plant species [70,71]. The efficiency of the markers was demonstrated in analyzing the genetic diversity of wheat and lentil germplasms [72,73]. CBDP markers also exhibited tremendous potential for the effective genetic diversity evaluation of the Salvia species, which is necessary for crop improvement programs involving hybridization and QTL mapping [74]. They unveiled a high level of genetic variation in Iranian Aegilops, providing vital information for potential valuable genes for wheat breeding programs [75]. However, these markers are also associated with limitations, as the reproducibility of the assay depends on the DNA sample quality and the PCR conditions used [68]. Additionally, the transferability of CBDP markers across species can be limited [74]. Furthermore, the CBDP markers have relatively low discriminatory power, making them less suitable for complex genetic relationship studies in some plant species [72].

3.1.5. Sequence-Related Amplified Polymorphism (SRAP)

SRAP markers are a valuable tool in molecular genetics research, amplifying coding regions of DNA using primers that target open reading frames [76]. These markers have demonstrated high variability and robustness, comparable to the commonly used AFLP markers, with the added benefit of being less technically demanding [77]. A significant advantage of using an SRAP marker is that one SRAP primer can combine with an unlimited number of other primers [78]. SRAP markers have primarily been used in agronomic and horticultural research, including developing quantitative trait loci in advanced hybrids and assessing genetic diversity in extensive germplasm collections [79]. Additionally, they were utilized to better understand the genetic diversity, molecular characteristics, and population structure of prairie grass (Bromus catharticus) and Bermuda grass (Cynodon dactylon) [80,81]. The investigation of gene flow between populations using SRAP markers provided valuable insights into the conservation and breeding efforts of Simao pine and Coloneaster shantungensis [82,83].
Furthermore, the high discriminatory power and ability to amplify gene-rich regions of the genome have made SRAP markers a promising tool for genetic mapping studies aimed at marker-assisted introgression in sugarcane [84]. They have also been applied in Morus macroura genetic differentiation studies, where primer E was found to distinguish between male and female PCR bands [85]. Although SRAP markers have been predominantly used in the investigation of plant genetic diversity, their applications in taxonomic studies could be influenced by anthropogenic or different types of selection pressures [77].

3.1.6. Start Codon-Targeted Polymorphism (SCoT)

SCoT markers are based on short, conserved regions flanking the ATG start codon in plant genes [86]. They are more reproducible and effective than arbitrary RAPD and ISSR markers in assessing the diversity and population structures of many plant varieties [19]. In the bottle gourd germplasm, 20 SCoT markers produced 161 amplicons and generated 82.61% polymorphisms, demonstrating the efficacy of SCoT markers in identifying and characterizing genetic diversity [87]. They have been applied to identify and analyze the genetic diversity of mango cultivars, enhancing breeding strategies and facilitating germplasm management [88]. Furthermore, they were utilized to identify inter-varietal distinctions among eight cereal grass varieties [89]. SCoT markers have also been instrumental in investigating the genetic foundation of Elymus sibiricus and Boehmeria nivea, which are cross-pollinated heterozygous species [90,91]. The markers were also deployed to differentiate selected Glycine max cultivars and to study rose genotypes’ genetic backgrounds [92,93]. The detection of induced allelic variations within Fusarium yellow tolerant/resistant lines of ginger was also performed using SCoT markers [94]. SCoT markers are rarely used directly for species identification, though they have been prominently employed for genetic diversity and population structure studies of several plants [95].

3.1.7. Cleaved Amplified Polymorphism Sequence (CAPS)

The CAPS, also known as PCR-restriction fragment length polymorphism (PCR-RFLP) markers, are designed by first performing the PCR amplification of DNA fragments generated via restriction enzyme digestion using specific primers, and the resulting products are separated using an agarose gel [96]. They present several advantages, including the co-dominance of alleles, the ability to utilize a small amount of template DNA, locus specificity, and high reproducibility. Unlike RFLP markers, CAPS markers eliminate the need for technically demanding procedures such as Southern blot hybridization and radioactive detection [97]. However, the size of amplified fragments and the need for sequence data for designing the marker are significant shortcomings of the marker [98].
CAPS markers are primarily applied in gene mapping investigations [97] and serve as a valuable tool for detecting SNPs within watermelon genomes [99] and for the identification of distinct Glycyrrhiza glabra genotypes [100]. These markers have been harnessed to differentiate fiber quality in two cultivated cotton varieties, Gossypium hirsutum and G. barbadense. This distinction facilitates marker-assisted selection initiatives, aiding in incorporating phytochrome and/or HY5 genes between these species [99]. CAPS markers have also been used for genetic diversity and population structure analyses and to manage emerging button mushroom genetic resources [101].

3.1.8. Inter-Primer Binding Site (iPBS)

The iPBS marker system has emerged as a robust DNA fingerprinting technique that operates effectively without requiring sequence information. This system has become the favored universal marker for discerning genetic distinctions within and between various eukaryotic organisms, spanning both intra-specific and inter-specific levels [102,103]. iPBS was described by Kalendar [104] as a universal DNA labeling method used in plants based on the primer binding site for the reverse transcription enzyme of the LTR retrotransposon. This marker system has been successfully applied for genetic diversity investigations related to breeding and conserving rare and economic plants like Allium ledebourianum, Juglans regia, and cotton germplasm [105,106,107]. Similarly, iPBS markers have identified an extensive diversity of scarlet eggplant, forming the genetic base resource of a specific breeding program in African indigenous vegetables [108]. The first-ever safflower (Carthamus tinctorius) plant-breeding parents were successfully determined using iPBS markers [109]. These markers exhibit genome-inherent plasticity, enabling them to uncover variations within Turkish okra germplasm [110] effectively. Furthermore, iPBS markers have been employed to distinguish well-defined genetic relationships among various tree peony germplasm, accurately classifying the varieties [111]. The markers have also been utilized for the DNA fingerprinting of natural hybrids that share morphological similarities with their parents, determination of phylogenies, and taxonomic discrimination in species belonging to Fagaceae [112].

3.1.9. Simple Sequence Repeats (SSRs)

SSRs, also known as microsatellites, are DNA markers based on short tandem repeated motifs, the number of which varies at a specific locus [113]. SSR markers offer several advantages over other marker systems. Firstly, they are highly reproducible, making them valuable for genetic analysis. Additionally, ultra-pure DNA templates are not required for their use. Secondly, SSR markers are highly polymorphic, allowing for the detection of allelic variations even among closely related varieties [10]. Thirdly, SSR polymorphisms are co-dominant, allowing for the easy interpretation of results. Finally, SSR markers are abundant and well-distributed throughout eukaryotic genomes [114]. They have been successfully used in various genetic applications, such as evaluating the genetic diversity and relationships among Chrysanthemum morifolium cultivars [115], mutation detection in pineapple [116], cultivar and germplasm selection and differentiation, genetic improvements, population structure determination and marker-assisted selection breeding of new varieties [117,118]. SSR markers have also been applied for the genotype identification of Tunisian citrus species [119], linkage mapping for Curculigo latifolia, Chrysanthemum indicum, and Capsicum frutescens [120,121,122,123], and genetic diversity evaluation of Lathyrus species and Lactuca indica [124,125]. Bhardwaj et al. [126] analyzed the genetic diversity of 353 Solanum tuberosum accessions using 25 SSR markers and found a high level of allelic polymorphisms among the accessions. However, SSR marker development is complex and costly, requiring sequence information to design primer sets for a specific species. Alternative approaches, such as next-generation sequencing, EST libraries, and enriched genome library searches, have been designed to substantially increase the production of SSR loci [127].

3.1.10. Inter-Retrotransposon Amplified Polymorphism (IRAP)

IRAP markers are retrotransposon-based markers that identify the insertion sequence between two retrotransposons with outward-facing long terminal repeats (LTRs) [128]. Retrotransposons are abundant in plant genomes and occur irreversibly in high copy numbers, making them particularly useful for phylogenetic and gene mapping studies [129]. IRAP markers were applied to determine the genetic diversity of 29 citrus genotypes and confirm Pommelo and Mandarin as true citrus species [130]. Additionally, they have been widely used to differentiate among hybrids and assess inter-species variations in different plants [131]. The first species-specific LTR retrotransposons cloned from five rare relic species of drugs plants, namely Adenophora lilifolia, Adonis sibirica, Adonis vernalis, Digitalis grandiflora, and Paeonia anomala, revealed the genetic diversity among the six populations of A. vernalis [132].
Many variations of retrotransposon-based markers are in existence, which include the sequence-specific amplified polymorphism (S-SAP), retrotransposon-microsatellite amplified polymorphism (REMAP), and retrotransposon-based insertional polymorphism (RBIP). S-SAP constitutes a multiplex marker system explicitly designed to detect variations in DNA flanking retrotransposon insertion sites [133]. In contrast, REMAP employs a single primer system grounded in the LTR target sequence and a simple sequence repeat motif to amplify regions exhibiting polymorphisms [128]. RBIP, on the other hand, operates as a co-dominant marker system utilizing PCR primers designed from the retrotransposon and flanking DNA to examine the insertional polymorphisms related to individual retrotransposons [134]. The primary historical limitation associated with retrotransposon-based marker systems has been the necessity for sequence data [129].

3.1.11. Conserved DNA-Derived Polymorphism (CDDP)

CDDP markers have been derived from extensively characterized plant genes responsive to biotic and abiotic stresses and developmental processes [135]. This approach results in the creation of functional markers that are directly associated with particular plant phenotypes. The markers produced detectable length polymorphisms with conserved DNA regions that share the same priming site but differ in genomic distribution [136]. Investigations of plants employing CDDP markers have revealed that genetic diversity is influenced by geographical distances and the level of gene exchange among groups and sexes of plants under different environmental conditions [137]. These markers have been successfully applied in evaluating genetic diversity in various plant species, aiding in genetic rescue, in situ and ex situ conservation efforts, and cultivar identification [67,138]. CDDP markers have many advantages, including convenience, lower cost, and rich polymorphisms, which can be utilized to produce targeted trait markers [24].

3.1.12. Diversity Array Technique (DArT)

The DArT was developed as a sequence-independent, hybridization-based marker performed on a microarray platform to capture the allelic diversity inherent to the target organism [139]. They offer the advantage of concurrently assessing numerous polymorphisms based on restriction sites across genotypes without necessitating the availability of DNA sequence information or site-specific oligonucleotides [140]. The polymorphism observed through DArT markers primarily arises from validated restriction site variations, a characteristic substantiated through investigations involving the model plant Arabidopsis thaliana [140].
They have become promising markers in genomic discovery, comparative mapping, directed breeding of superior oat varieties, and generating consensus maps of rye and oat [141]. Sánchez-Sevilla [142] showed that a comprehensive set of 603 DArT markers was highly efficient for classifying 62 strawberry cultivars based on historical, geographical, and pedigree-based cues. DArT markers have also demonstrated robustness among different mapping plant populations, allowing for map alignment [143]. In addition, they have proven highly efficient and cost-effective for genetic mapping in apples, providing moderate genome coverage [144]. DArT markers also have the potential to provide comprehensive genetic diversity analysis, whole-genome profiling, and high-density mapping of complex traits essential for marker-assisted breeding. Consequently, molecular analyses using DArT markers have significantly improved various crop species [145].
However, applying DArT markers in plant research poses significant technical challenges, characterized by intricate and time-consuming procedures requiring trained personnel. Moreover, this microarray-based technique demands substantial investments in physical energy, costs, and access to advanced laboratory facilities. These factors collectively constrain the widespread utilization of DArT markers in plant diversity studies [10].

3.1.13. Internal Transcribed Spacer (ITS)

ITS markers are indispensable for accurate species identification and play a pivotal role in DNA barcoding, focusing on the spacer DNA located within repetitive ribosomal RNA gene sequences [146]. The nuclear ribosomal DNA ITS region exhibits significant divergence across species while maintaining a higher level of conservation within a specific species, making them the preferred genetic markers for species-level identification [147]. Eukaryotic ITS regions possess several advantageous attributes, including their compact size, highly conserved flanking sequences, and ease of detection, even when working with limited DNA quantities due to the abundance of rRNA gene clusters [148].
Furthermore, they exhibit substantial variation, even among closely related species, and undergo rapid concerted evolution driven by unequal crossing over and gene conversion [149]. ITS1 has been successfully used to distinguish medicinal herbs, such as Amomum villosum (Zingiberaceae), and to differentiate Boerhavia diffusa (Punarnava) from Boerhavia erecta [150,151]. ITS1 has also been exploited for the reliable authentication of medicinal plants, detecting adulterants and substitutes of Gmelina arborea (Gambhari) [152]. Combining the ITS rDNA marker and cytogenetic analysis has helped resolve many taxonomic riddles, including citrus species and subspecies [153]. The main concern with this marker is its weak discrimination power for lower taxa and lack of species resolution, so carefully selecting the ITS marker is of utmost importance to avoid negative results [154].

3.1.14. Directed Amplification of Minisatellite DNA (DAMD)

DAMD markers were elucidated by Heath [155] and have found varied applications in plant diversity research in Cucumis sativus [156], Capsicum [157], and more recently, in citrus [158] and Musa cavendishii [159]. Numerous DAMD primers have been successfully harnessed to assess genetic diversity among distinct plant genotypes, as evidenced by the works of Ince and Karaca [160], Jain [161], and Saleh [162]. The DAMD markers have the advantages of not requiring prior sequence information, inheriting through Mendelian fashion, and behaving as dominant markers, although they may have limitations in reproducibility and the non-homology of generated fragments [155].
A concise overview, including examples, of applying a single marker system in genetic diversity studies of some plant species is illustrated in Table 1.

3.2. Utilization of Combined Molecular Markers

Using a single marker system in plant genetic diversity studies has limitations, necessitating the adoption of a multi-marker approach [19]. These arise because the conclusions drawn from analyzing a specific marker type can be effectively supported and validated by the findings derived from another marker. The simultaneous use of multiple markers can yield more precise outcomes with enhanced reliability and accuracy, thereby aiding in developing effective plant conservation strategies and improving crop varieties [6].

3.2.1. Cumulative Applications of Dominant Markers

Goswami [163] utilized RAPD and SCoT markers to investigate genetic variation in Lasiurus sindicus, revealing high percentages of polymorphic bands from both markers. The study emphasized substantial within-population genetic variation (90%) compared to among-population variation (10%) and provided valuable insights into the plant’s genetic diversity and potential for conservation and enhancement. Hromadova [164] assessed the effectiveness of 10 RAPD and 10 SCoT markers in detecting genetic diversity among 33 common bean genotypes. The SCoT markers (MI = 7.474, DI = 2.265) proved more effective than the RAPD ones (MI = 5.323, DDI = 1.612) with higher diversity detection index (DDI) and marker index (MI) values. Dendrograms and PCoA plots from both markers confirmed the distinct separation of bean genotypes, with the SCoT markers outperforming the RAPD markers in detecting genetic diversity.
Mansoory [165] employed ISSR and SCoT markers to evaluate the genetic diversity of 57 Diospyros genotypes from Iran. A cluster analysis grouped the genotypes into four clusters, with D. kaki in groups one and two, D. lotus in group three, and D. virginiana in group four. Combining both markers enhanced genotype separation, indicating Iran’s rich Diospyros germplasm and the utility of multiple markers for accurate diversity assessment. Alzahrani [166] examined genetic divergence in sixteen Medicago sativa (Alfalfa) cultivars (twelve from Saudi Arabia and four from Egypt) using ISSR and SCoT markers. ISSR generated 163 bands (60% polymorphism), while SCoT produced 150 bands (77% polymorphism), with the cultivars clustering into two populations in the dendrogram, aiding alfalfa breeding for drought tolerance and high yield. SCoT and IRAP markers have also proven helpful for assessing inter- and intra-specific genetic diversity in Diospyros germplasms for genotype identification and the adoption of effective conservation efforts [167]. Bhattacharyya [168] deployed ISSR and DAMD markers cumulatively to accurately assess the genetic relationships among distinct Dendrobium nobile germplasms. Khodaee [169] studied the genetic diversity of 48 Aegilops triuncialis accessions in Iran using SCoT, CBDP, and ISSR markers. A total of 359 DNA fragments were generated, with ISSR showing the highest diversity (PIC = 0.3, MI, Rp) among the multiple markers applied. Genetic diversity was greater in the Alborz population, with the UPGMA classification of accessions aligned with the geographical distribution. Combined markers offer more significant insights for future wheat breeding programs.
Arya [170] explored the genetic diversity of 20 Morinda tomentosa genotypes using 131 SCoT, 97 RAPD, and 70 ISSR markers. The SCoT markers exhibited the highest polymorphism (70.23%), Nei’s gene diversity (0.20), and geographic clustering, followed by ISSR and RAPD. The efficacy of SCoT and ISSR markers for genetic diversity analysis and geographic patterning of M. tomentosa was emphasized, enabling potential strategies for conservation and collection across regions and globally. EL-Mansy [171] also investigated the divergence aspects of six tomato lines (G1, G2, G3, G4, G5, and G6) using RAPD, ISSR, and SCoT markers. The marker analysis highlighted ISSR primers 49A, HB-14, 49A, 49B, and 89B as the most informative, with ISSR providing the highest unique specific markers (six), followed by RAPD (four) and SCoT (three). The cluster analysis showed the grouping of G1, G2, and G3 together and separated other lines. Apana [6] conducted an investigation into the genetic diversity and population structure of Clerodendrum serratum utilizing diverse molecular markers, including CBDP, iPBS, ISSR, and SCoT. The findings revealed that SCoT markers exhibited more efficacy in detecting polymorphisms and distinguishing genotypes than the remaining markers in the study. The multiple marker analysis revealed moderate gene flow and low genetic differentiation among the populations, with no significant correlation between the geographic and genetic distances. Three genetic clusters were determined using individual markers, while five genetic groups with high admixture were observed using pooled marker data, aiding in the conservation and management of C. serratum. The combined use of 15 ISSR, 11 SCoT, and 10 iPBS primers by Amom [4] in the genetic diversity and population structure study of endemic Dendrocalamus manipureanus showed significant genetic differentiation among the populations due to low gene flow, as indicated by the GST (0.684) and Nm (0.230) values. The BARRIER analysis identified nine genetic barriers, suggesting hindrances to gene flow. The findings from the cumulative analyses hold significant implications for effectively managing and enhancing the genetic characteristics of indigenous bamboo.
Tahir [20] applied ISSR, CDDP, and SCoT markers to explore genetic diversity in 59 barley accessions, generating 391 bands (255 ISSR, 35 CDDP, and 101 SCoT). SCoT displayed superior allelic diversity assessment with gene diversity averages at 0.77 (ISSR), 0.67 (CDDP), and 0.81 (SCoT), and the PIC recorded at 0.74 (ISSR), 0.63 (CDDP), and 0.80 (SCoT). Barley was grouped into two main clusters, with 15%, 9%, and 14% variability among the populations. Tiwari [172] utilized two sets of markers for the analysis of genetic diversity in 39 Andrographis paniculata specimens. They employed gene-targeted markers, specifically twenty-two SCoT and nineteen CBDP primers, alongside arbitrary amplified markers comprising eighteen RAPD and five ISSR primers. Notably, gene-targeted markers yielded more amplified amplicons, with 132 and 97 for SCoT and CBDP, respectively, compared to 124 and 32 for the RAPD and ISSR markers. Furthermore, it was observed that the polymorphic information content (PIC) values ranged from 0.09 to 0.48, with an average value of 0.34 and 0.41 per primer for the SCoT and CBDP markers, respectively. The resolving power ranged from 2.36 to 10.54, averaging 1.39 to 13.15 per primer for the SCoT and CBDP markers, respectively. Both the PIC and resolving power exhibited high values in the gene-targeted markers, whereas for the RAPD and ISSR markers, the PIC values extended from 0.32 to 0.45, and the resolving power ranged from 2.13 to 10.03. These findings underscore the clear applicability and reliability of gene-targeted markers for assessing genetic diversity in A. paniculata.
Table 1. Plant genetic diversity studies using a single marker system.
Table 1. Plant genetic diversity studies using a single marker system.
Molecular MarkersApplicationsPlants
Investigated
References
AFLPAmplified fragment length polymorphism: uses restriction enzymes and primers specific to genomic DNA to amplify DNA fragments of different sizes.Detects genetic variation within and among populations, linkage mapping, discrimination of cultivars, and association analyses.Tectona grandis; Brassica oleracea; Glehnia littoralis; Solanum tuberosum; Daucus carota[173,174,175,176,177]
ISSRInter-simple sequence repeat: uses primers specific to inter-microsatellite regions to amplify DNA fragments of different sizes.Evaluates genetic variation within and among populations, linkage mapping, and association analyses.Lepidium sativum; Balanites aegyptiaca; Prunus armeniaca; Vigra unguiculata; Camellia yuhsienensis; Clitaria ternatea[178,179,180,181,182,183]
RAPDRandom amplified polymorphic DNA: uses arbitrary primers to amplify DNA fragments of different sizes.Detects genetic variation within and among populations and genetic similarity.Carica papaya; Coffee canephora; Allium sativum; Dendrobium species; Nigella sativa[184,185,186,187,188]
SSRSimple sequence repeat: uses primers specific to microsatellite regions to amplify DNA fragments of different sizes.Ascertains genetic variation within and among populations, linkage mapping, association analyses, and plant breeding.Solanum tuberosum; Cajanus cajan; Vicia amoena; Allium sativum; Curcuma longa[119,126,189,190,191]
RFLPRestriction fragment length polymorphism: uses restriction enzymes to cut DNA at specific sites, and the resulting fragments are separated via gel electrophoresis.Detects genetic variation within and among populations and DNA fingerprinting. Oryza sativa; Fragaria x Ananassa; Brassica juncea[192,193,194]
DArTDiversity array technology: a high-throughput marker technology that uses a combination of restriction enzymes and a microarray platform. Determines genetic variation within and among populations and marker-assisted selection.Lesquerella species; Glycine max; Vigna unguiculata; Camellia sinensis[195,196,197,198]
SCARSequence-characterized amplified region: uses primers specific to a known DNA sequence to amplify a fragment of a specific size.Detects specific genes or alleles in a population and marker-assisted selection.Calanthe species;
Poa pratensis; Dendrobium officinale; Musa species; Moringa oleifera
[199,200,201,202,203]
CAPSCleaved amplified polymorphic sequence: uses restriction enzymes and primers specific to a known DNA sequence to amplify a fragment of a specific size.Detects specific genes or alleles in a population, identification of cultivars, and marker-assisted selection. Glycyrrhiza species; Lathyrus sativum; Citrullus lanatus; Zingiber officinale; Capsicum annum[100,204,205,206,207]
IRAPInter retrotransposon amplified polymorphism: uses primers specific to transposable elements to amplify DNA fragments of different sizes.Evaluates genetic variation within and among populations.Sorghum bicolor; Piper nigrum; Hordeum vulgare; Pinus sylvestris; Sakura species[208,209,210,211,212]
CDDPConserved DNA-derived polymorphism: uses a single primer constructed with a conserved area of functional genes.Ascertains genetic variation within and among populations.Salix taishanensis; Pistacia vera; Musa species; Arachis hypogaea; Amomum tsao-kosaleh[138,213,214,215,216]
DAMDDirected amplification of minisatellite-region DNA: uses a single primer specific to inter-microsatellite regions. Assesses genetic variation within and among populations.Capsicum; Origanum syriacum; Salvia officinalis; Ficus sycomorus[157,217,218,219]
SRAPSequence-related amplified polymorphism: uses arbitrary forward and reverse primer combinations targeting ORFs to amplify a coding region.Detects genetic variation within and among populations, mapping and tagging genes, germplasm identification, and sex determination.Cuminum cyminum; Pinus yunnanensis; Lavandula angustifolia; Aspergillus flavus; Zea mays[220,221,222,223,224]
SCoTStart codon-targeted polymorphism: uses a short-conserved region flanking the start codon, producing highly reproducible amplification of targeted DNA fragments of different sizes.Detects genetic variation within and among populations, determines population structures, identifies cultivars, QTL mapping, and DNA fingerprinting.Ardisia crenata; Avena nuda; Scutellaria baicalensis; Trigonella species; Triticum aestivum; Crataegus monogyna[225,226,227,228,229,230]
ITS2Internal transcribed spacer 2: a segment of the internal transcribed spacer (ITS) region, utilized as an alternative for species differentiation, involves the spacer DNA located within the tandem repeats separating the small and large subunits of ribosomal RNA (rRNA). ITS primers are designed to amplify the gene sequence containing the fastest-evolving region of the rRNA gene, resulting in fragments of varying sizes for differentiating species.Evaluates genetic variation within and among populations, intraspecific variation, species identification, authentication of plant variety, and detection of adulterants.Dendrobium species; Physalis species; Astragalus species;[231,232,233]
iPBSInter-primer binding site: uses the primer binding site for the reverse transcription enzyme of the LTR retrotransposon. No prior sequence information to amplify DNA fragments of different sizes, a preferred universal marker system.Detects genetic differentiation at both the intra-specific and inter-specific levels, marker-assisted selection, and breeding.Abelmoschus esculentus; Alfalfa; Phaseolus vulgaris; Triticum species; Brassica species; Castanea sativa [234,235,236,237,238,239]
CBDPCAAT-box derived polymorphism: Uses the CAAT box consensus sequence of the plant promoter upstream of the start codon to amplify DNA fragments of different sizes.Detects genetic diversity among and within species/populations, cultivar identification, linkage map construction, and marker-assisted selection.Triticum durum; Salvia species; Lens culinaris[73,74,240]
STSSequence-tagged site: Short DNA sequences of known locations that are easily detectable using PCR and serve as landmarks in the physical map of the genome.Variation analysis, gene expression, genome mapping, and gene silencing.Cenchrus species; Triticum aestivum; Oryza sativa; Agropyron cristatum; Secale cereale; Thinopyrum intermedium[241,242,243,244,245,246]
Amom [19] applied four markers (RAPD, ISSR, iPBS, and SCoT) to analyze 50 genotypes of native bamboo in North-East India. Forty primers of four marker systems generated varying polymorphic bands, with SCoT being the most informative and discriminatory. The Mantel test revealed a highly positive correlation between the markers, ranging from 0.60 (SCoT and RAPD) to 0.83 (iPBS and ISSR), indicating their effectiveness. The genetic clustering of bamboo genotypes is based on DNA markers aligned with their geographical origins. The multiple markers analysis produced precise genetic relationship determination among the native bamboo. Gene-targeted markers like SCoT and CBDP, used in conjunction with other molecular markers such as DAMD, CDDP, and IRAPs, have significantly contributed to the analysis of genetic relationships, gene mapping, conservation, breeding, and conservation of many medicinal and food crops [247,248].

3.2.2. Cumulative Application of Dominant and Co-Dominant Markers

Using both dominant and co-dominant markers in plant genetic diversity studies offers several advantages, as these markers provide good genome coverage and more accurate genetic data [249]. Additionally, they compensate for each other’s limitations. The studies conducted on Maize [250], Shorea curtisii [251], and Stenotaphrum secundatum [252] using AFLP and SSR markers facilitated the identification of quantitative trait loci (QTL) on specific chromosome regions. Applying AFLP and SSR markers has also been instrumental in explaining the high phenotypic variance observed between rice varieties [253]. Basu [254] assessed genetic diversity in jute cultivars (Corchorus olitorius and C. capsularis) using SSR and AFLP markers. The study revealed high variation between the two jute species, indicating distant maternal and possible different origins. However, some prominent Indian cultivars were closely related to wild accessions with unique genotypes from India and Kenyan accessions.
The genetic relatedness of 82 walnut genotypes from the Himalayan region was examined using 13 SSR and 20 RAPD primers [255]. High genetic diversity was evident within populations, with SSR primers displaying one to five alleles per locus and RAPD primers showing two to six alleles. Polymorphic loci were at 100%, and the average similarity was 49% (12% to 79%). The dendrogram analysis using these two markers revealed four subclusters, significantly affecting walnut breeding and conservation strategies. Zargar [256] employed 15 RAPD and 23 SSR markers to assess diversity among 51 common bean genotypes exhibiting high polymorphism, generating 171 polymorphic RAPD and 268 SSR bands. SSRs demonstrated a greater PIC value (0.300) and resolving power (5.241) than RAPDs, while RAPDs had a higher marker index (2.69). Hierarchical clustering accurately grouped genotypes based on cultivation area, and STRUCTURE analysis revealed three subpopulations aligned with distance-based groupings, indicating significant genetic diversity. Dar [257] explored genetic diversity among 47 sesame accessions using 22 RAPD and 18 SSR primers, with RAPD primers yielding 191 polymorphic bands while SSR primers produced only polymorphic fragments. SSRs exhibited higher polymorphic information content (0.194), while RAPDs showed a greater marker index (1.426) and resolving power (4.012). The genetic information derived from the cumulative application of markers emphasized their potential applications in DNA fingerprinting, germplasm conservation, and crop enhancement for Sesamum indicum.
Nascimento [258] observed high polymorphisms (95% for SSR and 75.8% for ISSR) while assessing the genetic diversity of 53 Dioscorea trifida accessions using eight SSR and sixteen ISSR markers. The dendrogram analysis of both markers showed the accessions clustering into three main groups, with the Bayesian and principal coordinate analyses supporting the grouping. While high variation was observed within groups (66.5% for SSR and 60.6% for SSR), the genetic and geographic distances showed slight correlations (r = 0.08, p = 0.0007 for SSR; r = 0.16, p = 0.0002 for ISSR). Hammami [259] demonstrated significant differences among populations (67%) but lower variation within populations (24%) in wild Brachypodium using SSR and ISSR markers. SSR and ISSR analyses revealed higher polymorphic fragments in B. hybridum than in B. distachyon with species-specific clustering. The principal component analysis linked genetic traits, climate, and geography, separating the two species. Ramzan [260] investigated the genetic diversity in twenty-one Tamarix samples using ten ISSR and six SSR primers, with significant polymorphisms (88.5% for ISSR and 80.28% for SSR) and high mean PIC values of 0.34 (ISSR) and 0.35 (SSR). The genetic variability among ecotypes was high, with dissimilarity indexes ranging from 0.00 to 0.77, and the Kalurkot and Bhakkar specimens showed the highest dissimilarity. Nazir [261] showed that using SSR markers in a study on 63 buckwheat genotypes, including local variants from India’s northwestern Himalayas, produced effective polymorphisms. ISSRs exhibited higher resolving power (4.38) than SSRs (1.42), while SSRs showed a greater average PIC value (0.43) than ISSRs (0.36). Geographical clustering using the two marker systems was accurate, and the STRUCTURE analysis unveiled substantial genetic diversity within the population, benefiting buckwheat breeding and conservation endeavors. Papaioannou [262] examined the genetic diversity of 27 garlic accessions using SSR and ISSR markers, revealing 26 distinct alleles for SSR and 84 for ISSR. SSR markers exhibited a higher redundancy level than ISSRs, potentially indicating duplicated accessions. An AMOVA highlighted that most molecular diversity originated from within-accession differences while clustering analyses using UPGMA, STRUCTURE, and PCoA based on SSRs showed consistent results.
The comparative analysis of gene diversity using dominant DArT and co-dominant SSR and SNP markers in Lolium perenne revealed that the DArT marker exhibited the highest consistency and reproducibility [263]. Additionally, genome SSR and CAPS markers were proven effective in identifying suitable candidates for breeding salt-tolerant rice (Oryza sativa L.) and locating high sodium transport-associated genes for mapping [264]. Shahnazari [265] employed SSR and CAPS-SSR markers to genetically fingerprint 13 sweet orange cultivars using SSR markers, which enabled hybrid prediction in orange cultivars, showing high diversity among sweet orange trees. The cultivars exhibited high genetic variability (with an average polymorphism of 98.46%), with Behshahr and Jadeh Ghadim 2 genotypes showing the highest and lowest genetic diversity values. Additionally, K-means clustering divided the cultivars into two main groups, while genetic similarity suggested potential cases of homonymy or synonymy. The applications of combined markers in the genetic diversity studies of different plants are briefly described in Table 2.

4. Drawbacks and Recent Developments in DNA Marker Technology

The applicability of DNA markers is enormous, but they are endowed with many drawbacks that limit their uses in plant research. One of the main disadvantages is the high cost of marker technology, which may require expensive equipment like PCR and DNA sequencers, commercial kits, reagents, etc. [285,286]. The high expenses may restrict access to this technology for a smaller group of researchers with limited funds [23]. Other weaknesses of molecular markers include longer time consumption, especially for large-scale investigations, and potential environmental effects on the results of molecular marker studies due to factors such as temperature, light, moisture, etc. [24]. Specific markers must be selected for certain investigations, as the choice of marker may influence the outcome of the study [287]. Hussain and Nisar [2] also emphasized the significance of selecting suitable markers for plant genetic diversity studies, as some may not work for other species. However, markers specifically designed for a particular species are not readily available and are difficult to establish, as the available reference genomes or markers influence the development of new markers [288]. Many markers cannot represent the entire genome, as they cannot provide complete coverage of the whole plant genome, resulting in an inaccurate evaluation of genetic diversity and relationships. Guo [289] highlighted the importance of combining molecular markers and other techniques to acquire a more comprehensive and accurate picture of genetic variation. Jagtap [290] also stressed the validation of the results of molecular marker analysis with other methods, as they are associated with a high false positive rate. Another drawback of molecular markers is the difficulty in interpreting and analyzing complex data. The interpretation of extensive molecular data requires experts with good knowledge of biology and population genetics [13]. Furthermore, managing and analyzing large datasets generated from high-throughput genotyping by sequencing platforms is a daunting task that involves advanced software and expertise. Using these markers requires technical personnel with specialized skills and knowledge in genetics and molecular biology [9].
Despite several limitations, molecular markers remain vital tools for understanding genetic diversity and the evolution of plant species for the last three decades. However, the necessity to develop more efficient and novel markers for assessing genetic variation, species identification, and molecular systematic studies is becoming increasingly apparent. The advancement of marker technology in recent times has partly addressed some limitations and challenges in using markers in plant research. One significant advancement was the emergence of SNPs, representing a third-generation molecular marker technology succeeding RFLPs and SSRs, among others. SNPs, which determine variation within a single nucleotide of DNA, can be easily detected with recent advancements in genomics [291]. Coined by Eric S. Lander in 1996, SNPs originated from sequence polymorphisms resulting from single nucleotide mutations at specific loci within DNA sequences [292] (Figure 3).
SNPs have yielded profound insights into genetic diversity, facilitating the elucidation of relationships between distinct varieties and empowering cultivators to enhance crop yields and safeguard germplasm integrity [293]. Notably, SNP markers have effectively delineated Gossypium hirsutum from other Gossypium species and have further demarcated wild from cultivated G. hirsutum [294]. Similarly, they have facilitated discrimination within the notably diverse Ethiopian sorghum population [295]. Employing SNP sequences from nuclear and chloroplast gene regions has proven advantageous in diverse applications, including phylogenetic analysis, evolutionary studies, and inheritance determination [296]. Remarkable genetic diversity within maize inbred lines and heterotic groups has been revealed through SNP genotyping [296]. Furthermore, genome-wide SNPs within various Camellia sinensis varieties have been identified through genotyping-by-sequencing [297]. SNPs are preferred over conventional SSR markers as they are economical, reliable, effective, stable, and amenable to automation [298,299]. However, developing SNPs in plants is tough with the unavailability of many sequenced model plants and the possibility of duplication of complex genomes with rich repeat sequences [300].
The rapid progression of next-generation sequencing (NGS) technology, cost reduction, and the development of new bioinformatics pipelines have enabled the discovery of SNPs on a large scale in several plants [300]. Genotyping-by-sequencing (GBS) is a rapid, high-throughput, and affordable NGS-based approach for SNP identification in a combined one-step marker detection and genotyping process without requiring the reference genome [301,302]. A simplified representation of experimental steps involved in GBS technology is shown in Figure 4. SNPs generated through GBS application are extremely helpful for genetic diversity analysis, genome-wide association studies (GWAS), QTL mapping, genomic selection, and breeding improvement without known markers in several non-model plant species prevailing across the globe [303,304]. Tomar [305] used a total of 14,563 high-quality SNPs identified using GBS to genotype and illustrate the population structure and genetic variation within and between subgroups of 141 elite advanced breeding lines of spring wheat from CIMMYT (Mexico). The determination of low heterozygosity between advanced wheat breeding lines within subgroups and the moderate variation among subgroups revealed the possibility of applying the elite wheat breeding lines for further GWAS studies. Diaz [306] utilized the SNPs from GBS to analyze the genetic diversity and population structure of the Acrocomia genus, consisting of 172 samples from seven species. The study affirmed the classical taxonomy of the genus, showing specific groups and the genetic differentiation of A. aculeata, A. totai, A. intumescens, and A. crispa. Dang [307] also employed 92,719 high-quality SNPs originating from restriction-site-associated DNA sequencing (RADseq/GBS) technology to determine low genetic diversity (HO = 0.249 and HE = 0.208) and population differentiation in Reaumuria trigyna.
They also observed the positioning of 353 outlier SNPs in 243 gene coding sequences in the R. trigyna transcriptome with potential sites of diversifying selection in the genes related to secondary metabolite synthesis and hormone regulation. The SNPs identified through GBS proved to be highly efficient markers that have been applied for genetic diversity studies of several plants, such as Cenchrus americanus [308], maize [309], and Ipomea batata [310]. A systematic investigation of 128 maize inbred lines by Dube [311] using 11,450 SNP markers revealed significant genetic diversity (p < 0.001) in key phenotypic traits. The mean gene diversity (GD) and polymorphic information content (PIC) were 0.40 and 0.31, indicating substantial variation. The population structure analysis identified three subpopulations consistent with the phylogenetic analysis. These findings from SNP marker analysis highlighted considerable genetic diversity in maize inbred lines, providing a foundation for selecting lines with favorable alleles and suggesting potential applications of marker-assisted selection for key agronomic traits. Haung [312] proposed constructing a broccoli fingerprint using SNPs for cultivar identification, understanding global broccoli diversity, and providing insights for advancing breeding programs. They analyzed 161 broccoli cultivars using 10 selected pairs of SNP primers, generating 78 alleles. The polymorphic information content (PIC) ranged from 0.64 to 0.90, revealing genetic similarities and distinctions between domestic and foreign cultivars.
SNP microarrays are another high-throughput genotyping platform that relies on hybridization and fluorescence principles and can genotype thousands to millions of SNPs across the genome in a single experiment [313]. These microarrays are extensively used in plant genetic research due to their efficiency, accuracy, and ability to provide comprehensive insights into the genetic makeup of plant populations [313]. The SNP array is a specialized type of DNA microarray that includes carefully designed probes, each carrying information about specific SNP positions. In the hybridization process, these probes interact with fragmented DNA to determine the distinct alleles of all SNPs present on the array for a specific DNA sample [314]. The exhaustive scrutiny of SNP data yields insights into genetic variations and structural modifications within the genome, allowing for a meticulous characterization of genomic abnormalities [315]. Several SNP arrays have demonstrated successful applications in genotyping diploid plants. Notable examples include the Apple 480 K SNP array, the Maize 600 K SNP array, and the Rice 700 K SNP array [316]. Each array provided a comprehensive platform for genotyping, offering insights into the genetic variations within their respective plant species [317]. Creating and fine-tuning SNP arrays involves a significant investment of time and effort and a notable challenge in this process is the occurrence of ascertainment bias [318]. This bias often arises from non-random polymorphism sampling or limited SNP discovery panels [319]. Various strategies are employed to address and minimize such bias. High-coverage whole-genome sequencing is one approach that aims to provide a more comprehensive and unbiased representation of genetic variation [320]. Additionally, updates to SNP array markers are implemented to incorporate discoveries and enhance the accuracy of genotyping information. Another tactic involves the integration of markers from multiple arrays, offering a more inclusive perspective on genetic diversity [321].
The development of DNA barcode techniques also facilitates plant diversity research by correctly identifying plant samples in a repeatable and reliable manner and determining the consistency of species definition across plant lineages with a measure of genetic variability based on the DNA barcode sequence data [321]. The efficiency of DNA barcodes relies on combining the strengths of molecular genetics, sequencing technologies, and bioinformatics [321]. Ribulose bisphosphate carboxylase large chain (rbcL) and maturase K (matK) genes are used as core DNA barcodes for seed plants, while the psbA-trnH intergenic spacer (psbA-trnH) and internal transcribed spacer (ITS) sequences are employed as supplementary DNA barcodes [322]. DNA barcode technology is utilized for accurate identification, genetic differentiation, and phylogenetic and species discrimination studies on several plants [323,324]. Enhancements in the field of epigenetics may also play a significant role in diversity study as epigenetic modification, like DNA methylation, can alter gene expression, influencing different adaptation responses of plants to environmental changes [17]. Massicote [325] mentioned the dependence of epigenetic processes on genetic variation. Wang [326] stated that epigenetic variation is the absolute downstream effect of genetic changes, while some considered it an independent phenomenon [327]. The assimilation of marker data with epigenetic information may potentially provide new insights into plant genetic diversity studies. The recent development of many improved bioinformatics tools also enabled the integration of molecular marker data with other sources of genetic information like genomic sequences, making it more efficient in plant genetic diversity analysis [328]. Advancements in molecular genetics, next-generation sequencing technologies, and bioinformatics have accelerated the development of more efficient and advanced molecular markers, which help address the challenges and limitations of using molecular markers in plant research.

5. Conclusions

Various molecular markers have been employed over the past three decades to study varied aspects of genetic diversity, including assessing the gene flow, population structure, and cluster analysis of several plants. The development of more advanced gene-targeted markers through rapid progression in molecular genetics has enabled the generation of high-resolution genetic data to make accurate decisions about appropriate conservation strategies for many important plants. Despite their many useful characteristics, the markers are also associated with several limitations that must be resolved. Developing efficient and cost-effective markers that can offer more precise and complete information on plant diversity level, population genetic structure, and cluster assignment is highly essential. The combination of multiple markers and genomic data generated through high throughput sequencing technologies will immensely help accelerate the understanding of plant genetic structure by providing a more comprehensive picture of diversity at the genome-wide level. With the continuously evolving technology, the prospect of molecular markers in the genetic analysis of plants is promising and bright, offering great potential in expanding our knowledge in properly preserving and utilizing increasingly depleting plant resources.

Author Contributions

Conceptualization, N.B. and P.N.; original draft preparation and funding acquisition, P.N., N.B. and I.J.; review and editing, K.N., V.V., W.A.V. and R.S.D.; visualization K.N., V.V., W.A.V. and R.S.D. All authors have read and agreed to the published version of the manuscript.

Funding

The Financial assistance provided to Potshangbam Nongdam under Core Research Grant (Plant sciences), Grant number-EMR/2017/000706 by SERB (Science Engineering and Research Board), New Delhi, India, is highly acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tikendra, L.; Koijam, A.S.; Nongdam, P. Molecular Markers based Genetic Fidelity Assessment of Micropropagated Dendrobium chrysotoxum Lindl. Meta Gene 2019, 20, 100562. [Google Scholar] [CrossRef]
  2. Hussain, H.; Nisar, M. Assessment of Plant Genetic Variations using Molecular Markers: A Review. J. Appl. Biol. Biotechnol. 2020, 8, 099–109. [Google Scholar] [CrossRef]
  3. Tuvesson, S.D.; Larsson, C.T.; Ordon, F. Use of Molecular Markers for Doubled Haploid Technology: From Academia to Plant Breeding Companies. In Double Haploid Technology; Segui-Simarro, J.M., Ed.; Humana: New York, NY, USA, 2021; Volume 2288, pp. 49–72. ISBN 9781071613351. [Google Scholar] [CrossRef]
  4. Amom, T.; Tikendra, L.; Potshangbam, A.M.; Bidyananda, N.; Devi, R.S.; Dey, A.; Sahoo, M.R.; Vendrame, W.A.; Jamir, I.; Nongdam, P. Conservation strategies for endemic Dendrocalamus manipureanus: A study on genetic diversity and population structure based on molecular and phytochemical markers. S. Afr. J. Bot. 2023, 152, 106–123. [Google Scholar] [CrossRef]
  5. Pereira-Dias, L.; Vilanova, S.; Fita, A.; Prohens, J.; Rodríguez-Burruezo, A. Genetic diversity, population structure, and relationships in a collection of pepper (Capsicum spp.) landraces from the Spanish centre of diversity revealed by genotyping-by-sequencing (GBS). Hort. Res. 2019, 6, 54. [Google Scholar] [CrossRef]
  6. Apana, N.; Amom, T.; Tikendra, L.; Angamba, P.; Dey, A.; Nongdam, P. Genetic diversity and population structure of Clerodendrum serratum (L.) Moon using CBDP, iPBS and SCoT markers. J. Appl. Res. Med. 2021, 25, 100349. [Google Scholar] [CrossRef]
  7. Gyani, P.C.; Bollinedi, H.; Gopala Krishnan, S.; Vinod, K.K.; Sachdeva, A.; Bhowmick, P.K.; Ellur, R.K.; Nagarajan, M.; Singh, A.K. Genetic Analysis and Molecular Mapping of the Quantitative Trait Loci Governing Low Phytic Acid Content in a Novel LPA Rice Mutant, PLM11. Plants 2020, 9, 1728. [Google Scholar] [CrossRef] [PubMed]
  8. Garrido-Cardenas, J.A.; Mesa-Valle, C.; Manzano-Agugliaro, F. Trends in plant research using molecular markers. Planta 2018, 247, 543–557. [Google Scholar] [CrossRef] [PubMed]
  9. Nadeem, M.A.; Nawaz, M.A.; Shahid, M.Q.; Doğan, Y.; Comertpay, G.; Yıldız, M.; Hatipoğlu, R.; Ahmad, F.; Alsaleh, A.; Labhane, N.; et al. DNA molecular markers in plant breeding: Current status and recent advancements in genomic selection and genome editing. Biotechnol. Biotechnol. Equip. 2018, 32, 261–285. [Google Scholar] [CrossRef]
  10. Amom, T.; Nongdam, P. The use of molecular marker methods in plants: A review. Int. J. Curr. Res. Rev. 2017, 9, 1–7. [Google Scholar]
  11. Avise, J.C. Molecular Markers, Natural History and Evolution, 2nd ed.; Sinauer Associates Inc.: Sunderland, MA, USA, 2004; ISBN 9780412037719. [Google Scholar]
  12. Ramesh, P.; Mallikarjuna, G.; Sameena, S.; Kumar, A.; Gurulakshmi, K.; Reddy, B.V.; Reddy, P.C.O.; Sekhar, A.C. Advancements in molecular marker technologies and their applications in diversity studies. J. Biosci. 2020, 45, 1–15. [Google Scholar] [CrossRef]
  13. Miah, G.; Rafii, M.Y.; Ismail, M.R.; Puteh, A.B.; Rahim, H.A.; Islam, K.N.; Latif, M.A. A Review of Microsatellite Markers and Their Applications in Rice Breeding Programs to Improve Blast Disease Resistance. Int. J. Mol. Sci. 2013, 14, 22499–22528. [Google Scholar] [CrossRef] [PubMed]
  14. Isshiki, S.; Okubo, H.; Fujieda, K. Isozyme variation in Cucumber (Cucumis sativus L.). J. Jpn. Soc. Hortic. Sci. 1992, 61, 595–601. [Google Scholar] [CrossRef]
  15. Nongdam, P.; Nirmala, C. Genetic variability in four species of Cymbidium based on isozyme markers. Physiol. Mol. Biol. Plants 2007, 13, 65–68. [Google Scholar]
  16. Zarei, A.; Erfani-Moghadam, J. SCoT markers provide insight into the genetic diversity, population structure and phylogenetic relationships among three Pistacia species of Iran. Genet. Resour. Crop Evol. 2021, 68, 1625–1643. [Google Scholar] [CrossRef]
  17. Kumar, A.; Shahina, P.S.; Singh, R.S.; Singh, D.N. Prospect of molecular markers in precision plant breeding. In Recent Advances in Chemical Sciences and Biotechnology; New Delhi Publishers: New Delhi, India, 2020; pp. 131–142. [Google Scholar]
  18. Tikendra, L.; Rahaman, H.; Dey, A.; Sahoo, M.R.; Nongdam, P. Applicability of Molecular Markers in Ascertaining Genetic Diversity and Relationship between Five Edible Bamboos of North-East India. In Molecular Marker Techniques; : Kumar, N., Ed.; Springer: Singapore, 2023; pp. 141–160. [Google Scholar] [CrossRef]
  19. Amom, T.; Tikendra, L.; Apana, N.; Goutam, M.; Sonia, P.; Koijam, A.S.; Potshangbam, A.M.; Rahaman, H.; Nongdam, P. Efficiency of RAPD, ISSR, iPBS, SCoT and phytochemical markers in the genetic relationship study of five native and economical important bamboos of North-East India. Phytochemistry 2020, 174, 112330. [Google Scholar] [CrossRef]
  20. Tahir, N.; Lateef, D.; Rasul, K.; Rahim, D.; Mustafa, K.; Sleman, S.; Mirza, A.; Aziz, R. Assessment of genetic variation and population structure in Iraqi barley accessions using ISSR, CDDP, and SCoT markers. Czech J. Genet. Plant Breed. 2023, 59, 148–159. [Google Scholar] [CrossRef]
  21. Aly, A.A.; Eliwa, N.E.; Borik, Z.M.; Safwat, G. Physiological variation of irradiated red raddish plants and their phylogenic relationship using SCoT and CDDP marker. Not. Bot. Hortic. Agrobot. Cluj Napoca 2021, 49, 12396. [Google Scholar] [CrossRef]
  22. Abaza, N.O.; Yousief, S.S.; Moghaieb, R.E.A. The efficiency of SCoT, ISSR, and SRAP markers for detecting genetic polymorphism among Egyptian barley genotypes. J. Pharm. Negat. Results 2022, 13, 1851–1863. [Google Scholar] [CrossRef]
  23. Pudake, R.N.; Kumari, M. Assessment of Genetic Diversity in Indigenous Plants from Northeast India Using Molecular Marker Technology. In Bioprospecting of Indigenous Bioresources of North-East India; Purkayastha, J., Ed.; Springer: Singapore, 2016; pp. 181–192. [Google Scholar] [CrossRef]
  24. Amiteye, S. Basic concepts and methodologies of DNA marker systems in plant molecular breeding. Heliyon 2021, 7, e08093. [Google Scholar] [CrossRef] [PubMed]
  25. Nilkanta, H.; Amom, T.; Tikendra, L.; Rahaman, H.; Nongdam, P. ISSR marker-based population genetic study of Melocanna baccifera (Roxb.) Kurz: A commercially important bamboo of Manipur, North-East India. Scientifica 2017, 2017, 3757238. [Google Scholar] [CrossRef] [PubMed]
  26. Amom, T.; Tikendra, L.; Rahaman, H.; Potshangbam, A.; Nongdam, P. Evaluation of genetic relationship between 15 bamboo species of North-East India based on ISSR marker analysis. Mol. Biol. Res. Commun. 2018, 7, 7–15. [Google Scholar] [CrossRef] [PubMed]
  27. Kordrostami, M.; Rahimi, M. Molecular markers in plants: Concepts and applications. Genet. Millenn. 2015, 13, 4024–4031. [Google Scholar]
  28. Chen, W.; Hou, L.; Zhang, Z.; Pang, X.; Li, Y. Genetic diversity, population structure, and linkage disequilibrium of a core collection of Ziziphusjujuba assessed with genome-wide SNPs developed by genotyping-by-sequencing and SSR markers. Front. Plant Sci. 2017, 8, 575. [Google Scholar]
  29. Hasan, N.; Choudhary, S.; Naaz, N.; Sharma, N.; Laskar, R.A. Recent advancements in molecular marker-assisted selection and applications in plant breeding programmes. J. Genet. Eng. Biotechnol. 2021, 19, 128. [Google Scholar] [CrossRef] [PubMed]
  30. Tikendra, L.; Potshangbam, A.M.; Dey, A.; Devi, T.R.; Sahoo, M.R.; Nongdam, P. RAPD, ISSR, and SCoT markers based genetic stability assessment of micropropagated Dendrobium fimbriatum Lindl. var. oculatum Hk. f.- an important endangered orchid. Physiol. Mol. Biol. Plants 2021, 27, 341–357. [Google Scholar] [CrossRef] [PubMed]
  31. Rahimi, M.; Ahmadi Afzadi, M.; Kordrostami, M. Genetic diversity in Sickleweed (Falcaria vulgaris) and using stepwise regression to identify marker associated with traits. Sci. Rep. 2023, 13, 12142. [Google Scholar] [CrossRef] [PubMed]
  32. Tikendra, L.; Amom, T.; Nongdam, P. Molecular genetic homogeneity assessment of micropropagated Dendrobium moschatum Sw.—A rare medicinal orchid, using RAPD and ISSR markers. Plant Gene 2019, 19, 100196. [Google Scholar] [CrossRef]
  33. Tikendra, L.; Potshangbam, A.M.; Amom, T.; Dey, A.; Nongdam, P. Understanding the genetic diversity and population structure of Dendrobium chrysotoxum Lindl. -An endangered medicinal orchid and implication for its conservation. S. Afr. J. Bot. 2021, 138, 364–376. [Google Scholar] [CrossRef]
  34. Chen, M.Y.; He, X.H.; Zhang, Y.L.; Lu, T.T.; He, W.Q.C.; Yang, J.H.; Huang, X.; Zhu, J.W.; Yu, H.X.; Luo, C. Genetic diversity and relationship analyses of mango (Mangifera indica L.) germplasm resources with ISSR, SRAP, CBDP and CEAP markers. Sci. Hortic. 2022, 301, 111146. [Google Scholar] [CrossRef]
  35. Pradhan, S.; Paudel, Y.P.; Qin, W.; Pant, B. Genetic fidelity assessment of wild and tissue cultured regenerants of a threatened orchid, Cymbidium aloifolium using molecular markers. Plant Gene 2023, 34, 100418. [Google Scholar] [CrossRef]
  36. Williams, J.G.; Kubelik, A.R.; Livak, K.J.; Rafalski, J.A.; Tingey, S.V. DNA polymorphisms amplified by arbitrary primers are useful as genetic markers. Nucleic Acids Res. 1990, 18, 6531–6535. [Google Scholar] [CrossRef] [PubMed]
  37. Mendonça, E.G.; de Souza, A.M.; de Almeida Vieira, F.; Estopa, R.A.; Reis, C.A.F.; de Carvalho, D. Using Random Amplified Polymorphic DNA to Assess Genetic Diversity and Structure of Natural Calophyllumbrasiliense (Clusiaceae). Populations in Riparian Forests. Int. J. For. Res. 2014, 2014, 1–8. [Google Scholar] [CrossRef]
  38. 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]
  39. Dhakshanamoorthy, D.; Selvaraj, R.; Chidambaram, A. Utility of RAPD marker for genetic diversity analysis in gamma rays and ethyl methane sulphonate (EMS)-treated Jatropha curcas plants. C. R. Biol. 2015, 338, 75–82. [Google Scholar] [CrossRef] [PubMed]
  40. Fu, C.; Qiu, Y.; Kong, H. RAPD analysis for genetic diversity in Changium smyrnioides (Apiaceae), an endangered plant1. Bot. Bull. Acad. Sin. 2003, 44, 13–18. [Google Scholar]
  41. Zhang, C.; He, P.; He, J.; Zhang, Y.; Qiao, Y.; Zhang, M.; Shi, Z.; Hu, S. RAPD analysis for genetic diversity of medicinal plant Coptis omeiensis. Zhongguo Zhong Yao Za Zhi = Zhongguo Zhongyao Zazhi = China J. Chin. Mater. Medica 2010, 35, 138–141. [Google Scholar] [CrossRef]
  42. Ruiz-Chután, J.A.; Salava, J.; Janovská, D.; Žiarovská, J.; Kalousová, M.; Fernández, E. Assessment of genetic diversity in Sorghum bicolor using RAPD markers. Genetika 2019, 51, 789–803. [Google Scholar] [CrossRef]
  43. Mortazavi Moghadam, F.A.; Qaderi, A.; Sharifi-Sirchi, G.R. Evaluation of Genetic Diversity of 17 Populations (Lepidium sativum L.) Plant Collected from Different Regions of Iran by RAPD Marker. ACS Agric. Sci. Technol. 2021, 1, 684–690. [Google Scholar] [CrossRef]
  44. Ben-Ari, G.; Lavi, U. Marker-assisted selection in plant breeding. In Plant Biotechnology and Agriculture; Academic Press: Cambridge, MA, USA, 2012; pp. 163–184. [Google Scholar] [CrossRef]
  45. Miller, J.C.; Tanksley, S.D. RFLP analysis of phylogenetic relationships and genetic variation in the genus Lycopersicon. Theor. Appl. Genet. 1990, 80, 437–448. [Google Scholar] [CrossRef]
  46. Mondini, L.; Arshiya, N.; Pagnotta, M.A. Assessing Plant Genetic Diversity by Molecular Tools. Diversity 2009, 1, 19–35. [Google Scholar] [CrossRef]
  47. Mir, R.R.; Hiremath, P.J.; Riera-Lizarazu, O.; Varshney, R.K. Evolving molecular marker technologies in plants: From RFLPs to GBS. In Diagnostics in Plant Breeding; Springer: Dordrecht, The Netherlands, 2013; pp. 229–247. [Google Scholar]
  48. Cui, Y.X.; Xu, G.W.; Magill, C.W.; Schertz, K.F.; Hart, G.E. RFLP-based assay of Sorghum bicolor (L.) Moench genetic diversity. Theor. Appl. Genet. 1995, 90, 787–796. [Google Scholar] [CrossRef] [PubMed]
  49. Maizura, I.; Rajanaidu, N.; Zakri, A.H.; Cheah, S.C. Assessment of Genetic Diversity in Oil Palm (Elaeis guineensis Jacq.) using Restriction Fragment Length Polymorphism (RFLP). Genet. Resour. Crop Evol. 2006, 53, 187–195. [Google Scholar] [CrossRef]
  50. Chang, C.; Bowman, J.L.; De John, A.W.; Lander, E.S.; Meyerowitz, E.M. Restriction fragment length polymorphism linkage map for Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 1998, 85, 6856–6860. [Google Scholar] [CrossRef] [PubMed]
  51. Benchimol, L.L.; de Souza, C.L., Jr.; Garcia, A.A.F.; Kono, P.M.S.; Mangolin, C.A.; Barbosa, A.M.M.; Coelho, A.S.G.; de Souza, A.P. Genetic diversity in tropical maize inbred lines: Heterotic group assignment and hybrid performance determined by RFLP markers. Plant Breed. 2000, 119, 491–496. [Google Scholar] [CrossRef]
  52. Zaharleva, M.; Santoni, S.; David, J. Use of RFLP markers to study genetic diversity and to build a core-collection of the wild wheat relative Ae. geniculata Roth (=Ae. ovata L.). Genet. Sel. Evol, 2001; 33, (Suppl. 1), S269–S288. [Google Scholar] [CrossRef]
  53. Bonierbale, M.W.; Plaisted, R.L.; Tanksley, S. RFLP maps based on a common set of clones reveal modes of chromosomal evolution in potato and tomato. Genetics 1998, 120, 1095–1103. [Google Scholar] [CrossRef] [PubMed]
  54. Tanksley, S.D.; Young, N.D.; Paterson, A.H.; Bonierbale, M.W. RFLP mapping in plant breeding: New tools for an old science. Biotechnology 1989, 7, 257–264. [Google Scholar] [CrossRef]
  55. Vos, P.; Hogers, R.; Bleeker, M.; Reijans, M.; Lee, T.V.D.; Hornes, M.; Friters, A.; Pot, J.; Paleman, J.; Kuiper, M.; et al. AFLP: A new technique for DNA fingerprinting. Nucleic Acids Res. 1995, 23, 4407–4414. [Google Scholar] [CrossRef] [PubMed]
  56. Brugmans, B.; van der Hulst, R.G.; Visser, R.G.; Lindhout, P.; van Eck, H.J. A new and versatile method for the successful conversion of AFLP™ markers into simple single locus markers. Nucleic Acids Res. 2003, 31, e55. [Google Scholar] [CrossRef] [PubMed]
  57. Ganapathy, K.N.; Gnanesh, B.N.; Byre Gowda, M.; Venkatesha, S.C.; Gomashe, S.S.; Channamallikarjuna, V. AFLP analysis in pigeonpea (Cajanus cajan (L.) Millsp.) Revealed close relationship of cultivated genotypes with some of its wild relatives. Genet. Resour. Crop Evol. 2011, 58, 837–847. [Google Scholar] [CrossRef]
  58. Al-Nadabi, H.; Khan, M.; Al-Yahyai, R.A.; Al-Sadi, A.M. AFLP fingerprinting analysis of citrus cultivars and wild accessions from Oman suggests the presence of six distinct cultivars. Agriculture/Pol’nohospodárstvo 2018, 64, 173–182. [Google Scholar] [CrossRef]
  59. Huang, W.K.; Wan, F.H.; Guo, J.Y.; Gao, B.D.; Xie, B.Y.; Peng, D.L. AFLP analyses of genetic variation of Eupatorium adenophorum (Asteraceae) populations in China. Can. J. Plant Sci. 2009, 89, 119–126. [Google Scholar] [CrossRef]
  60. Murariu, D.; Plăcintă, D.D.; Simioniuc, V. Assessing genetic diversity in Romanian maize landraces, using molecular markers. Rom. Agric. Res. 2019, 36, 3–9. [Google Scholar] [CrossRef]
  61. Bhattacharyya, P.; Ghosh, S.; Mandi, S.S.; Kumaria, S.; Tandon, P. Genetic variability and association of AFLP markers with some important biochemical traits in Dendrobium thyrsiflorum, a threatened medicinal orchid. S. Afr. J. Bot. 2017, 109, 214–222. [Google Scholar] [CrossRef]
  62. El-Demerdash, E.S.S.; Elsherbeny, E.A.; Salama, Y.A.M.; Ahmed, M.Z. Genetic diversity analysis of some Egyptian Origanum and Thymus species using AFLP markers. J. Genet. Eng. Biotechnol. 2019, 17, 13. [Google Scholar] [CrossRef] [PubMed]
  63. Wu, W.D.; Liu, W.H.; Sun, M.; Zhou, J.Q.; Liu, W.; Zhang, C.L.; Zhang, X.Q.; Peng, Y.; Huang, L.K.; Ma, X. Genetic diversity and structure of Elymus tangutorum accessions from western China as unraveled by AFLP markers. Hereditas 2019, 156, 8. [Google Scholar] [CrossRef] [PubMed]
  64. Assaeed, A.M.; Al-Faifi, S.A.; Migdadi, H.M.; El-Bana, M.I.; Al Qarawi, A.A.; Khan, M.A. Evaluation of genetic diversity of Panicumturgidum Forssk from Saudi Arabia. Saudi J. Biol. Sci. 2018, 25, 123–129. [Google Scholar] [CrossRef]
  65. Wang, Q.; Ruan, X.; Jiang, H.; Meng, Q.; Wang, L. Genetic diversity of different geographical populations of Rhodiolarosea based on AFLP markers. Zhongguo Zhong Yao Za Zhi = Zhongguo Zhongyao Zazhi = China J. Chin. Mater. Medica 2009, 34, 2279–2284. [Google Scholar]
  66. Zhao, B.; Yin, Z.F.; Xu, M.; Wang, Q.C. AFLP analysis of genetic variation in wild populations of five Rhododendron species in Qinling Mountain in China. Biochem. Syst. Ecol. 2012, 45, 198–205. [Google Scholar] [CrossRef]
  67. Poczai, P.; Varga, I.; Laos, M.; Cseh, A.; Bell, N.; Valkonen, J.P.; Hyvönen, J. Advances in plant gene-targeted and functional markers: A review. Plant Methods 2013, 9, 6. [Google Scholar] [CrossRef] [PubMed]
  68. Singh, A.K.; Rana, M.K.; Singh, S.; Kumar, S.; Kumar, R.; Singh, R. CAAT box-derived polymorphism (CBDP): A novel promoter-targeted molecular marker for plants. J. Plant Biochem. Biotechnol. 2014, 23, 175–183. [Google Scholar] [CrossRef]
  69. Imani, A.; Ahmadi, J.; Heydari, M. Molecular variation and genetic relationships among Iranian and foreign pistachio cultivars using gene-targeted CAAT box-derived markers. Agric. Biotechnol. J. 2022, 14, 41–62. [Google Scholar]
  70. Tomar, P.; Malik, C.P. Genetic diversity assessment in Trachyspermum ammi L. Sprague using CDDP and CBDP markers. J. Plant Sci. Res. 2016, 32, 27. [Google Scholar]
  71. Ahmed, D.A.; Tahir, N.A.; Salih, S.H.; Talebi, R. Genome diversity and population structure analysis of Iranian landrace and improved barley (Hordeum vulgare L.) genotypes using arbitrary functional gene-based molecular markers. Genet. Resour. Crop Evol. 2021, 68, 1045–1060. [Google Scholar] [CrossRef]
  72. Etminan, A.; Pour-Aboughadareh, A.; Mohammadi, R.; Noori, A.; Ahmadi-Rad, A. Applicability of CAAT Box-derived Polymorphism (CBDP) Markers for Analysis of Genetic Diversity in Durum Wheat. Cereal Res. Commun. 2018, 46, 1–9. [Google Scholar] [CrossRef]
  73. Sarvmeili, J.; Saidi, A.; Farrokhi, N.; Pouresmael, M.; Talebi, R. Genetic diversity and population structure analysis of landrace and wild relatives of lentil germplasm using CBDP marker. Cytol. Genet. 2020, 54, 566–573. [Google Scholar] [CrossRef]
  74. Fabriki-Ourang, S.; Karimi, H. Assessment of genetic diversity and relationships among Salvia species using gene targeted CAAT box-derived polymorphism markers. J. Genet. 2019, 98, 75. [Google Scholar] [CrossRef] [PubMed]
  75. Eslamzadeh, M.; Omidi, M.; Rashidi, V.; Etminan, A. Evaluation of genetic diversity and population structure analysis in some Aegilops species using CBDP markers. IGS 2021, 16, 1–8. [Google Scholar]
  76. Zhou, Y.; Wang, X.; Zhang, X. Development and application of a SRAP marker for the identification of sex in Buchloe dactyloides. Euphytica 2011, 181, 261–266. [Google Scholar] [CrossRef]
  77. Robarts, D.W.; Wolfe, A.D. Sequence-related amplified polymorphism (SRAP) markers: A potential resource for studies in plant molecular biology1. Appl. Plant Sci. 2014, 2, 1400017. [Google Scholar] [CrossRef] [PubMed]
  78. Li, G.; McVetty, P.B.E.; Quiros, C.F. SRAP molecular marker technology in plant science. In Plant Breeding from Laboratories to Fields; Andersen, S.B., Ed.; InTech: Rijeka, Croatia, 2013. [Google Scholar] [CrossRef]
  79. Zhou, L.; Yarra, R.; Cao, H.; Zhao, Z. Sequence-Related Amplified Polymorphism (SRAP) Markers Based Genetic Diversity and Population Structure Analysis of Oil Palm (Elaeis guineensis Jacq.). Trop. Plant Biol. 2021, 14, 63–71. [Google Scholar] [CrossRef]
  80. Zheng, Y.; Xu, S.; Liu, J.; Zhao, Y.; Liu, J. Genetic diversity and population structure of Chinese natural bermudagrass [Cynodon dactylon (L.) Pers.] germplasm based on SRAP markers. PLoS ONE 2017, 12, e0177508. [Google Scholar] [CrossRef] [PubMed]
  81. Yi, L.; Dong, Z.; Lei, Y.; Zhao, J.; Xiong, Y.; Yang, J.; Xiong, Y.; Gou, W.; Ma, X. Genetic diversity and molecular characterization of worldwide prairie grass (Bromus catharticus Vahl) accessions using SRAP markers. Agronomy 2021, 11, 2054. [Google Scholar] [CrossRef]
  82. Wang, D.; Shen, B.; Gong, H. Genetic diversity of Simao pine in China revealed by SRAP markers. PeerJ 2019, 7, e6529. [Google Scholar] [CrossRef]
  83. Agyenim-Boateng, K.G.; Lu, J.; Shi, Y.; Zhang, D.; Yin, X. SRAP analysis of the genetic diversity of wild castor (Ricinus communis L.) in South China. PLoS ONE 2019, 14, e0219667. [Google Scholar] [CrossRef] [PubMed]
  84. Suman, A.; Kimbeng, C.; Edmé, S.; Veremis, J. Sequence-related amplified polymorphism (SRAP) markers for assessing genetic relationships and diversity in sugarcane germplasm collections. Plant Genet. Resour. 2008, 6, 222–231. [Google Scholar] [CrossRef]
  85. Aseny, N.; Syamsuardi, S.; Nurainas, N. Molecular characterization of andalas tree dioecious plant [Morus macroura Miq.] using SRAP marker. IOP Conf. Ser. Earth Environ. Sci. 2021, 741, 012050. [Google Scholar] [CrossRef]
  86. Collard, B.C.; Mackill, D.J. Start codon targeted (SCoT) polymorphism: A simple, novel DNA marker technique for generating gene-targeted markers in plants. Plant Mol. Biol. Rep. 2009, 27, 86–93. [Google Scholar] [CrossRef]
  87. Abdin, M.Z.; Arya, L.; Verma, M. Use of SCoT markers to assess the gene flow and population structure among two different populations of bottle gourd. Plant Gene 2017, 9, 80–86. [Google Scholar] [CrossRef]
  88. Luo, C.; He, X.H.; Chen, H.; Ou, S.J.; Gao, M.P. Analysis of diversity and relationships among mango cultivars using Start Codon Targeted (SCoT) markers. Biochem. Syst. Ecol. 2010, 38, 1176–1184. [Google Scholar] [CrossRef]
  89. Mavlyutov, Y.M.; Shamustakimova, A.O.; Klimenko, I.A. Application of SCoT markers for accessing of genetic diversity of gramineous forage grass species. IOP Conf. Ser. Earth Environ. Sci. 2021, 901, 012038. [Google Scholar] [CrossRef]
  90. Zhang, J.; Xie, W.; Wang, Y.; Zhao, X. Potential of start codon targeted (SCoT) markers to estimate genetic diversity and relationships among Chinese Elymus sibiricus accessions. Molecules 2015, 20, 5987–6001. [Google Scholar] [CrossRef] [PubMed]
  91. Satya, P.; Karan, M.; Jana, S.; Mitra, S.; Sharma, A.; Karmakar, P.G.; Ray, D.P. Start codon targeted (SCoT) polymorphism reveals genetic diversity in wild and domesticated populations of ramie (Boehmeria nivea L. Gaudich.), a premium textile fiber producing species. Meta Gene 2015, 3, 62–70. [Google Scholar] [CrossRef] [PubMed]
  92. Rayan, W.A.; Osman, S.A. Phylogenetic relationships of some Egyptian soybean cultivars (Glycine max L.) using SCoT marker and protein pattern. Bull. Natl. Res. Cent. 2019, 43, 161. [Google Scholar] [CrossRef]
  93. Agarwal, A.; Gupta, V.; Haq, S.U.; Jatav, P.K.; Kothari, S.L.; Kachhwaha, S. Assessment of genetic diversity in 29 rose germplasms using SCoT marker. J. King Saud Univ. Sci. 2019, 31, 780–788. [Google Scholar] [CrossRef]
  94. Sharma, V.; Thakur, M. Applicability of SCoT markers for detection of variations in Fusarium yellows resistant lines of ginger (Zingiber officinale Rosc.) induced through gamma irradiations. S. Afr. J. Bot. 2021, 140, 454–460. [Google Scholar] [CrossRef]
  95. Feng, S.; Zhu, Y.; Yu, C.; Jiao, K.; Jiang, M.; Lu, J.; Shen, C.; Ying, Q.; Wang, H. Development of species-specific SCAR markers, based on a SCoT analysis, to authenticate Physalis (Solanaceae) species. Front. Genet. 2018, 9, 192. [Google Scholar] [CrossRef] [PubMed]
  96. Shavrukov, Y.N. CAPS markers in plant biology. Russ. J. Genet. Appl. Res. 2016, 6, 279–287. [Google Scholar] [CrossRef]
  97. Konieczny, A.; Ausubel, F.M. A procedure for mapping Arabidopsis mutations using co-dominant ecotype-specific PCR-based markers. Plant J. 1993, 4, 403–410. [Google Scholar] [CrossRef] [PubMed]
  98. Shavrukov, Y. Cleaved Amplified Polymorphic Sequences (CAPS) Markers in Plant Biology; Nova Science Publishers, Inc.: Hauppauge, NY, USA, 2014; 251p, ISBN 9781631175534. [Google Scholar]
  99. Liu, S.; Gao, P.; Zhu, Q.; Luan, F.; Davis, A.R.; Wang, X. Development of cleaved amplified polymorphic sequence markers and a CAPS-based genetic linkage map in watermelon (Citrullus lanatus [Thunb.] Matsum. And Nakai) constructed using whole-genome re-sequencing data. Breed. Sci. 2016, 66, 244–259. [Google Scholar] [CrossRef] [PubMed]
  100. Jo, I.H.; Sung, J.; Hong, C.E.; Raveendar, S.; Bang, K.H.; Chung, J.W. Development of cleaved amplified polymorphic sequence (CAPS) and high-resolution melting (HRM) markers from the chloroplast genome of Glycyrrhiza species. 3 Biotech 2018, 8, 220. [Google Scholar] [CrossRef] [PubMed]
  101. Kushanov, F.N.; Pepper, A.E.; Yu, J.Z.; Buriev, Z.T.; Shermatov, S.E.; Saha, S.; Ulloa, M.; Jenkins, J.N.; Abdukarimov, A.; Abdurakhmonov, I.Y. Development, genetic mapping and QTL association of cotton PHYA, PHYB, and HY5-specific CAPS and dCAPS markers. BMC Genet. 2016, 17, 1–11. [Google Scholar] [CrossRef] [PubMed]
  102. An, H.; Lee, H.Y.; Shim, D.; Choi, S.H.; Cho, H.; Hyun, T.K.; Jo, I.H.; Chung, J.W. Development of CAPS markers for evaluation of genetic diversity and population structure in the germplasm of button mushroom (Agaricus bisporus). J. Fungi 2021, 7, 375. [Google Scholar] [CrossRef] [PubMed]
  103. Erper, I.; Ozer, G.; Kalendar, R.; Avci, S.; Yildirim, E.; Alkan, M.; Turkkan, M. Genetic diversity and pathogenicity of Rhizoctonia spp. isolates associated with red cabbage in Samsun (Turkey). J. Fungus 2021, 7, 234. [Google Scholar] [CrossRef] [PubMed]
  104. Ouyang, Z.; Wang, Y.; Ma, T.; Kanzana, G.; Wu, F.; Zhang, J. Genome-wide identification and development of LTR retrotransposon-based molecular markers for the Melilotus Genus. Plants 2021, 10, 890. [Google Scholar] [CrossRef] [PubMed]
  105. Kalendar, R.; Antonius, K.; Smýkal, P.; Schulman, A.H. iPBS: A universal method for DNA fingerprinting and retrotransposon isolation. Theor. Appl. Genet. 2010, 121, 1419–1430. [Google Scholar] [CrossRef] [PubMed]
  106. Khapilina, O.; Turzhanova, A.; Danilova, A.; Tumenbayeva, A.; Shevtsov, V.; Kotukhov, Y.; Kalendar, R. Primer Binding Site (PBS) Profiling of Genetic Diversity of Natural Populations of Endemic Species Allium ledebourianum Schult. BioTech 2021, 10, 23. [Google Scholar] [CrossRef] [PubMed]
  107. Başak, İ.; Özer, G.; Muradoğlu, F. Morphometric traits and iPBS based molecular characterizations of walnut (Juglans regia L.) genotypes. Genet. Resour. Crop Evol. 2022, 69, 2731–2743. [Google Scholar] [CrossRef]
  108. Baran, N.; Shimira, F.; Nadeem, M.A.; Altaf, M.T.; Andirman, M.; Baloch, F.S.; Temiz, M.G. Exploring the genetic diversity and population structure of upland cotton germplasm by iPBS-retrotransposons markers. Mol. Biol. Rep. 2023, 50, 4799–4811. [Google Scholar] [CrossRef]
  109. Shimira, F.; Boyaci, H.F.; Çilesiz, Y.; Nadeem, M.A.; Baloch, F.S.; Taşkin, H. Exploring the genetic diversity and population structure of scarlet eggplant germplasm from Rwanda through iPBS-retrotransposon markers. Mol. Biol. Rep. 2021, 48, 6323–6333. [Google Scholar] [CrossRef] [PubMed]
  110. Ali, F.; Yılmaz, A.; Nadeem, M.A.; Habyarimana, E.; Subaşı, I.; Nawaz, M.A.; Chaudhary, H.J.; Shahid, M.Q.; Ercişli, S.; Zia, M.A.B.; et al. Mobile genomic element diversity in world collection of safflowers (Carthamus tinctorius L.) panel using iPBS-retrotransposon markers. PLoS ONE 2019, 14, e0211985. [Google Scholar] [CrossRef] [PubMed]
  111. Yıldız, M.; Koçak, M.; Baloch, F.S. Genetic bottlenecks in Turkish okra germplasm and utility of iPBS retrotransposon markers for genetic diversity assessment. Genet. Mol. Res. 2015, 14, 10588–10602. [Google Scholar] [CrossRef] [PubMed]
  112. Duan, Y.B.; Guo, D.L.; Guo, L.L.; Wei, D.F.; Hou, X.G. Genetic diversity analysis of tree peony germplasm using iPBS markers. Genet. Mol. Res. 2015, 14, 7556–7566. [Google Scholar] [CrossRef] [PubMed]
  113. Coutinho, J.P.; Carvalho, A.; Martín, A.; Lima-Brito, J. Molecular characterization of Fagaceae species using inter-primer binding site (iPBS) markers. Mol. Biol. Rep. 2018, 45, 133–142. [Google Scholar] [CrossRef] [PubMed]
  114. Tautz, D. Hypervariability of simple sequences as a general source for polymorphic DNA markers. Nucleic Acids Res. 1989, 17, 6463–6471. [Google Scholar] [CrossRef] [PubMed]
  115. Singh, A.K.; Singh, N.K.; Singh, V.K.; Singh, D.P.; Singh, N.P. Tools for simple sequence repeat (SSR) markers. Agric. Update 2016, 11, 163–172. [Google Scholar] [CrossRef]
  116. Feng, S.; He, R.; Lu, J.; Jiang, M.; Shen, X.; Jiang, Y.; Wang, Z.; Wang, H. Development of SSR markers and assessment of genetic diversity in medicinal Chrysanthemum morifolium cultivars. Front. Genet. 2016, 7, 113. [Google Scholar] [CrossRef] [PubMed]
  117. Feng, S.; Tong, H.; Chen, Y.; Wang, J.; Chen, Y.; Sun, G.; He, J.; Wu, Y. Development of pineapple microsatellite markers and germplasm genetic diversity analysis. BioMed Res. Int. 2013, 2013, 317912. [Google Scholar] [CrossRef] [PubMed]
  118. Yang, S.; Zhong, Q.; Tian, J.; Wang, L.; Zhao, M.; Li, L.; Sun, X. Characterization and development of EST-SSR markers to study the genetic diversity and populations’ analysis of Jerusalem artichoke (Helianthus tuberosus L.). Genes Genom. 2018, 40, 1023–1032. [Google Scholar] [CrossRef] [PubMed]
  119. Kimaro, D.; Melis, R.; Sibiya, J.; Shimelis, H.; Shayanowako, A. Analysis of genetic diversity and population structure of pigeonpea [Cajanus cajan (L.) Millsp] accessions using SSR markers. Plants 2020, 9, 1643. [Google Scholar] [CrossRef] [PubMed]
  120. Romdhane, M.B.; Riahi, L.; Selmi, A.; Zoghlami, N. Patterns of genetic structure and evidence of gene flow among Tunisian Citrus species based on informative nSSR markers. C. R. Biol. 2016, 339, 371–377. [Google Scholar] [CrossRef] [PubMed]
  121. Babaei, N.; Abdullah, N.A.P.; Saleh, G.; Abdullah, T.L. Isolation and characterization of microsatellite markers and analysis of genetic variability in Curculigo latifolia Dry and. Mol. Biol. Rep. 2012, 39, 9869–9877. [Google Scholar] [CrossRef] [PubMed]
  122. Han, Z.; Ma, X.; Wei, M.; Zhao, T.; Zhan, R.; Chen, W. SSR marker development and intraspecific genetic divergence exploration of Chrysanthemum indicum based on transcriptome analysis. BMC Genom. 2018, 19, 291. [Google Scholar] [CrossRef] [PubMed]
  123. Zhong, Y.; Cheng, Y.; Ruan, M.; Ye, Q.; Wang, R.; Yao, Z.; Zhou, G.; Liu, J.; Yu, J.; Wan, H. High-throughput SSR marker development and the analysis of genetic diversity in Capsicum frutescens. Horticulturae 2021, 7, 187. [Google Scholar] [CrossRef]
  124. Rahman, M.M.; Quddus, M.R.; Ali, M.O.; Liu, R.; Li, M.; Yan, X.; Li, G.; Ji, Y.; Hossain, M.M.; Wang, C.; et al. Genetic diversity of Lathyrus sp collected from different geographical regions. Mol. Biol. Rep. 2022, 49, 519–529. [Google Scholar] [CrossRef] [PubMed]
  125. Oliya, B.K.; Kim, M.Y.; Lee, S.H. Development of genic-SSR markers and genetic diversity of Indian lettuce (Lactuca indica L.) in South Korea. Genes Genom. 2018, 40, 615–623. [Google Scholar] [CrossRef] [PubMed]
  126. Bhardwaj, V.; Kumar, A.; Sharma, S.; Singh, B.; Sood, S.; Dipta, B.; Singh, R.; Siddappa, S.; Thakur, A.K.; Dalamu, D.; et al. Analysis of Genetic Diversity, Population Structure, and Association Mapping for Late Blight Resistance in Potato (Solanum tuberosum L.) Accessions Using SSR Markers. Agronomy 2023, 13, 294. [Google Scholar] [CrossRef]
  127. Ahmad, A.; Wang, J.D.; Pan, Y.B.; Sharif, R.; Gao, S.J. Development and use of simple sequence repeats (SSRs) markers for sugarcane breeding and genetic studies. Agronomy 2018, 8, 260. [Google Scholar] [CrossRef]
  128. Kalendar, R.; Grob, T.; Regina, M.; Suoniemi, A.; Schulman, A. IRAP and REMAP: Two new retrotransposon-based DNA fingerprinting techniques. Theor. Appl. Genet. 1999, 98, 704–711. [Google Scholar] [CrossRef]
  129. Alzohairy, A.M.; Gyulai, G.; Ramadan, M.F.; Edris, S.; Sabir, J.S.; Jansen, R.K.; Eissa, H.F.; Bahieldin, A. Retrotransposon-based molecular markers for assessment of genomic diversity. Funct. Plant Biol. 2014, 41, 781–789. [Google Scholar] [CrossRef] [PubMed]
  130. Abedinpour, H.; Ranjbar, G.A.; Jelodar, N.B.; Golein, B. Evaluation of genetic diversity in Citrus genotypes by IRAP molecular marker. Int. J. Farm. Allied Sci. 2014, 3, 230–234. [Google Scholar]
  131. Widyawan, M.H.; Wulandary, S. Genetic diversity analysis of yardlong bean genotypes (Vigna unguiculata subsp. sesquipedalis) based on IRAP marker. Biodiversitas 2020, 21, d210333. [Google Scholar] [CrossRef]
  132. Boronnikova, S.V.; Kalendar, R.N. Using IRAP markers for analysis of genetic variability in populations of resource and rare species of plants. Russ. J. Genet. 2010, 46, 36–42. [Google Scholar] [CrossRef]
  133. Du, X.; Zhang, Q.; Luo, Z. Development of retrotransposon primers and their utilization for germplasm identification in Diospyros spp. (Ebenaceae). Tree Genet. Genomes 2009, 5, 235–245. [Google Scholar] [CrossRef]
  134. Flavell, A.J.; Knox, M.R.; Pearce, S.R.; Ellis, T.N. Retrotransposon-based insertion polymorphisms (RBIP) for high throughput marker analysis. Plant J. 1998, 16, 643–650. [Google Scholar] [CrossRef] [PubMed]
  135. Collard, B.C.Y.; Mackill, D.J. Conserved DNA-derived polymorphism (CDDP): A simple and novel method for generating DNA markers in plants. Plant Mol. Biol. Rep. 2009, 27, 558–562. [Google Scholar] [CrossRef]
  136. Bilčíková, J.; Farkasová, S.; Žiarovská, J. Genetic variability of commercially important apple varieties (Malus x domestica Borkh.) assessed by CDDP markers. Acta Fytotech. Zootech. 2021, 24, 21–26. [Google Scholar]
  137. Haffar, S.; Baraket, G.; Usai, G.; Aounallah, A.; Ben Mustapha, S.; Ben Abdelkrim, A.; Salhi Hannachi, A. Conserved DNA-derived polymorphism as a useful molecular marker to explore genetic diversity and relationships of wild and cultivated Tunisian figs (Ficus carica L.). Trees 2022, 36, 723–735. [Google Scholar] [CrossRef]
  138. Liu, H.; Zang, F.; Wu, Q.; Ma, Y.; Zheng, Y.; Zang, D. Genetic diversity and population structure of the endangered plant Salix taishanensis based on CDDP markers. Glob. Ecol. Conserv. 2020, 24, e01242. [Google Scholar] [CrossRef]
  139. Novoselović, D.; Bentley, A.R.; Šimek, R.; Dvojković, K.; Sorrells, M.E.; Gosman, N.; Horsenell, R.; Drezner, G.; Šatović, Z. Characterizing Croatian wheat germplasm diversity and structure in a European context by DArT markers. Front. Plant Sci. 2016, 7, 184. [Google Scholar] [CrossRef] [PubMed]
  140. Alexander, H.; Wittenberg, J.; Van, T.; Lee, D.; Cayla, C.; Kilian, A.; Visser, R.G.F.; Schouten, H.J. Validation of the high-throughput marker technology DArT using the model plant Arabidopsis thaliana. Mol. Genet. Genom. 2005, 274, 30. [Google Scholar] [CrossRef]
  141. Gawroński, P.; Pawełkowicz, M.; Tofil, K.; Uszyński, G.; Sharifova, S.; Ahluwalia, S.; Tyrka, M.; Wędzomy, M.; Kilian, A.; Bolibok-Brągozewska, H. DArT markers effectively target gene space in the rye genome. Front. Plant Sci. 2016, 7, 1600. [Google Scholar] [CrossRef] [PubMed]
  142. Sánchez-Sevilla, J.F.; Horvath, A.; Botella, M.A.; Gaston, A.; Folta, K.; Kilian, A.; Denoyes, B.; Amaya, I. Diversity arrays technology (DArT) marker platforms for diversity analysis and linkage mapping in a complex crop, the octoploid cultivated strawberry (Fragaria x ananassa). PLoS ONE 2015, 10, e0144960. [Google Scholar] [CrossRef] [PubMed]
  143. Iorizzo, M.; Gao, L.; Mann, H.; Traini, A.; Chiusano, M.L.; Kilian, A.; Aversano, R.; Carputo, D.; Bradeen, J.M. A DArT marker-based linkage map for wild potato Solanum bulbocastanum facilitates structural comparisons between Solanum A and B genomes. BMC Genet. 2014, 15, 123. [Google Scholar] [CrossRef] [PubMed]
  144. Schouten, H.J.; Van De Weg, W.E.; Carling, J.; Khan, S.A.; McKay, S.J.; van Kaauwen, M.P.; Wittenberg, A.H.J.; Koehorst-van Putten, H.J.J.; Noordijk, Y.; Gao, Z.; et al. Diversity arrays technology (DArT) markers in apple for genetic linkage maps. Mol. Breed. 2012, 29, 645–660. [Google Scholar] [CrossRef] [PubMed]
  145. Deres, D.; Feyissa, T. Concepts and applications of diversity array technology (DArT) markers for crop improvement. J. Crop Improv. 2023, 37, 913–933. [Google Scholar] [CrossRef]
  146. Chen, S.; Yao, H.; Han, J.; Liu, C.; Song, J.; Shi, L.; Zhu, Y.; Ma, X.; Gao, T.; Pang, X.; et al. Validation of the ITS2 region as a novel DNA barcode for identifying medicinal plant species. PLoS ONE 2010, 5, e8613. [Google Scholar] [CrossRef]
  147. Cheng, T.; Xu, C.; Lei, L.; Li, C.; Zhang, Y.; Zhou, S. Barcoding the kingdom Plantae: New PCR primers for ITS regions of plants with improved universality and specificity. Mol. Ecol. Resour. 2016, 16, 138–149. [Google Scholar] [CrossRef]
  148. Zhu, S.; Liu, Q.; Qiu, S.; Dai, J.; Gao, X. DNA barcoding: An efficient technology to authenticate plant species of traditional Chinese medicine and recent advances. Chin. Med. 2022, 17, 112. [Google Scholar] [CrossRef]
  149. Banchi, E.; Ametrano, C.G.; Greco, S.; Stanković, D.; Muggia, L.; Pallavicini, A. PLANiTS: A curated sequence reference dataset for plant ITS DNA metabarcoding. Database 2020, 2020, baz155. [Google Scholar] [CrossRef] [PubMed]
  150. Qiao, C.F.; Han, Q.B.; Zhao, Z.L.; Wang, Z.T.; Xu, L.S.; Xu, H.X. Sequence analysis based on ITS1 region of nuclear ribosomal DNA of Amomum villosum and ten species of Alpinia. J. Food Drug Anal. 2009, 17, 1. [Google Scholar] [CrossRef]
  151. Selvaraj, D.; Shanmughanandhan, D.; Sarma, R.K.; Joseph, J.C.; Srinivasan, R.V.; Ramalingam, S. DNA barcode ITS effectively distinguishes the medicinal plant Boerhavia diffusa from its adulterants. Genomics Proteomics Bioinform. 2012, 10, 364–367. [Google Scholar] [CrossRef] [PubMed]
  152. Manokar, J.; Balasubramani, S.P.; Venkatasubramanian, P. Nuclear ribosomal DNA–ITS region based molecular marker to distinguish Gmelina arborea Roxb. Ex Sm. from its substitutes and adulterants. J. Ayurveda Integr. Med. 2018, 9, 290–293. [Google Scholar] [CrossRef] [PubMed]
  153. Hynniewta, M.; Malik, S.K.; Rao, S.R. Genetic diversity and phylogenetic analysis of Citrus (L.) from north-east India as revealed by meiosis, and molecular analysis of internal transcribed spacer region of rDNA. Meta Gene 2014, 2, 237–251. [Google Scholar] [CrossRef] [PubMed]
  154. Li, Q.J.; Wang, X.; Wang, J.R.; Su, N.; Zhang, L.; Ma, Y.P.; Chang, Z.Y.; Zhao, L.; Potter, D. Efficient identification of Pulsatilla (Ranunculaceae) using DNA barcodes and micro-morphological characters. Front. Plant Sci. 2019, 10, 1196. [Google Scholar] [CrossRef] [PubMed]
  155. Heath, D.D.; Lwama, G.K.; Devlin, R.H. PCR primed with VNTR core sequences yields species specific patterns and hypervariable probes. Nucleic Acids Res. 1993, 21, 5782–5785. [Google Scholar] [CrossRef] [PubMed]
  156. Hu, J.B.; Li, J.W.; Wang, L.J.; Liu, L.J.; Si, S.W. Utilization of a set of high-polymorphism DAMD markers for genetic analysis of a cucumber germplasm collection. Acta Physiol. Plant 2011, 33, 227–231. [Google Scholar] [CrossRef]
  157. Ince, A.G.; Karaca, M.; Onus, A.N. Development and utilization of diagnostic DAMD-PCR markers for Capsicum accessions. Genet Resour. Crop Evol. 2009, 56, 211–221. [Google Scholar] [CrossRef]
  158. Pınar, H.; Bulut, M.; Altunoz, D.; Uzun, A.; Seday, U.; Yılmaz, K.U. Determination of genetic diversity and relationships within citrus and related genera using DAMD markers. Bangladesh J. Bot. 2017, 46, 163–170. [Google Scholar]
  159. Pinar, H.; Uzun, A.; Unlu, M.; Yaman, M. Genetic diversity in Turkish banana (Musa cavendishii) genotypes with DAMD markers. Fresenius Environ. Bull. 2019, 28, 459–463. [Google Scholar]
  160. İnce, A.G.; Karaca, M. Td-DAMD-PCR assays for fingerprinting of commercial carnations. Turk. J. Biol. 2015, 39, 290–298. [Google Scholar] [CrossRef]
  161. Jain, J.R.; Timsina, B.; Satyan, K.B.; Manohar, S.H. A comparative assessment of morphological and molecular diversity among Sechium edule (Jacq.) Sw. accessions in India. 3 Biotech 2017, 7, 106. [Google Scholar] [CrossRef] [PubMed]
  162. Saleh, B. Genetic diversity of Salvia tomentosa Miller (Lamiaceae) species using touch-down directed amplification of minisatellite DNA (Td-DAMD) molecular markers. Acta Biol. Szeged. 2019, 63, 135–141. [Google Scholar] [CrossRef]
  163. Goswami, B.; Gadi, B.R.; Rao, S.R. Morphological and molecular markers-based assessment of genetic diversity of a valuable endemic plant Lasiurus sindicus Henr. in the arid region of Rajasthan, India. Arid Land Res. Manag. 2022, 36, 298–313. [Google Scholar] [CrossRef]
  164. Hromadová, Z.; Gálová, Z.; Mikolášová, L.; Balážová, Ž.; Vivodík, M.; Chňapek, M. Efficiency of RAPD and SCoT markers in the Genetic Diversity Assessment of the Common Bean. Plants 2023, 12, 2763. [Google Scholar] [CrossRef] [PubMed]
  165. Mansoory, A.; Khademi, O.; Naji, A.M.; Rohollahi, I.; Sepahvand, E. Evaluation of genetic diversity in three Diospyros species, collected from different regions in Iran, using ISSR and SCoT molecular markers. Int. J. Fruit Sci. 2022, 22, 235–248. [Google Scholar] [CrossRef]
  166. Alzahrani, O.R.; Alshehri, M.A.; Alasmari, A.; Ibrahim, S.D.; Oyouni, A.A.; Siddiqui, Z.H. Evaluation of genetic diversity among Saudi Arabian and Egyptian cultivars of alfalfa (Medicago sativa L.) using ISSR and SCoT markers. J. Taibah Univ. Sci. 2023, 17, 2194187. [Google Scholar] [CrossRef]
  167. Guan, C.; Chachar, S.; Zhang, P.; Hu, C.; Wang, R.; Yang, Y. Inter-and intra-specific genetic diversity in Diospyros using SCoT and IRAP markers. Hortic. Plant J. 2020, 6, 71–80. [Google Scholar] [CrossRef]
  168. Bhattacharyya, P.; Kumaria, S.; Tandon, P. Applicability of ISSR and DAMD markers for phyto-molecular characterization and association with some important biochemical traits of Dendrobium nobile, an endangered medicinal orchid. Phytochem. 2015, 117, 306–316. [Google Scholar] [CrossRef] [PubMed]
  169. Khodaee, L.; Azizinezhad, R.; Etminan, A.R.; Khosroshahi, M. Assessment of genetic diversity among Iranian Aegilops triuncialis accessions using ISSR, SCoT, and CBDP markers. J. Genet. Eng. Biotechnol. 2021, 19, 1–9. [Google Scholar] [CrossRef] [PubMed]
  170. Arya, L.; Narayanan, R.K.; Verma, M.; Singh, A.K.; Gupta, V. Genetic diversity and population structure analyses of Morinda tomentosa Heyne, with neutral and gene-based markers. Genet. Resour. Crop Evol. 2014, 61, 1469–1479. [Google Scholar] [CrossRef]
  171. El-Mansy, A.B.; Abd El-Moneim, D.; ALshamrani, S.M.; Safhi, F.A.; Abdein, M.A.; Ibrahim, A.A. Genetic diversity analysis of tomato (Solanum lycopersicum L.) with morphological, cytological, and molecular markers under heat stress. Horticulturae 2021, 7, 65. [Google Scholar] [CrossRef]
  172. Tiwari, G.; Singh, R.; Singh, N.; Choudhury, D.R.; Paliwal, R.; Kumar, A.; Gupta, V. Study of arbitrarily amplified (RAPD and ISSR) and gene targeted (SCoT and CBDP) markers for genetic diversity and population structure in Kalmegh [Andrographis paniculata (Burm. f.) Nees]. Ind. Crops Prod. 2016, 86, 1–11. [Google Scholar] [CrossRef]
  173. Vaishnaw, V.; Mohammad, N.; Wali, S.A.; Kumar, R.; Tripathi, S.B.; Negi, M.S.; Ansari, S.A. AFLP markers for analysis of genetic diversity and structure of teak (Tectona grandis) in India. Can. J. For. Res. 2015, 45, 297–306. [Google Scholar] [CrossRef]
  174. El-Esawi, M.A.; Germaine, K.; Bourke, P.; Malone, R. AFLP analysis of genetic diversity and phylogenetic relationships of Brassica oleracea in Ireland. C. R. Biol. 2016, 339, 163–170. [Google Scholar] [CrossRef]
  175. Li, B.; Wang, A.; Zhang, P.; Li, W. Genetic diversity and population structure of endangered Glehnia littoralis (Apiaceae) in China based on AFLP analysis. Biotechnol. Biotechnol. Equip. 2019, 33, 331–337. [Google Scholar] [CrossRef]
  176. del Rio, A.; Bamberg, J. An AFLP Marker Core Subset for the Cultivated Potato Species Solanum phureja ( Solanum tuberosum L. subsp. andigenum). Am. J. Potato Res. 2021, 98, 374–383. [Google Scholar] [CrossRef]
  177. Domblides, A.; Domblides, E. Rapid Genetic Assessment of Carrot Varieties Based on AFLP Analysis. Horticulturae 2023, 9, 298. [Google Scholar] [CrossRef]
  178. Kumar, V.; Yadav, H.K. Assessment of genetic diversity in Lepidium sativum L. using inter simple sequence repeat (ISSR) marker. Physiol. Mol. Biol. Plants 2019, 25, 399–406. [Google Scholar] [CrossRef] [PubMed]
  179. Mint Abdelaziz, S.; Medraoui, L.; Alami, M.; Pakhrou, O.; Makkaoui, M.; Ould Mohamed Salem Boukhary, A.; Filali-Maltouf, A. Inter simple sequence repeat markers to assess genetic diversity of the desert date (Balanites aegyptiaca Del.) for Sahelian ecosystem restoration. Sci. Rep. 2020, 10, 14948. [Google Scholar] [CrossRef] [PubMed]
  180. Sheikh, Z.N.; Sharma, V.; Shah, R.A.; Sharma, N.; Summuna, B.; Al-Misned, F.A.; El-Serehy, H.A.; Mir, J.I. Genetic diversity analysis and population structure in apricot (Prunus armeniaca L.) grown under the north-western Himalayas using ISSR markers. Saudi J. Biol. Sci. 2021, 28, 5986–5992. [Google Scholar] [CrossRef]
  181. Saxena, A.; Rukam, T.S. Assessment of genetic diversity in cowpea (Vigna unguiculata L. Walp.) through ISSR marker. Res. J. Biotechnol. 2020, 15, 66–71. [Google Scholar]
  182. Zou, Y.; Ge, X.; Yan, C.; Zhong, Q.; Chen, D.; Chen, Z.; Yuan, Y.; Guo, H.; Zhou, Y.; Wang, J.; et al. Assessment of genetic diversity of Camellia yuhsienensis based on leaf structure and inter simple sequence repeat (ISSR) markers. Genet. Resour. Crop Evol. 2024, 19, 1–4. [Google Scholar] [CrossRef]
  183. Nurhasanah; Hindersah, R.; Suganda, T.; Concibido, V.; Sundari; Karuniawan, A. The First Report on the Application of ISSR Markers in Genetic Variance Detection among Butterfly Pea (Clitoria ternatea L.) Accession in North Maluku Province, Indonesia. Horticulturae 2023, 9, 1059. [Google Scholar] [CrossRef]
  184. Kumar, D.; Yadav, P.; Priyanka, Y.A.; Dwivedi, U.N.; Yadav, K. Genetic diversity analysis among papaya (Carica papaya L.) varieties using RAPD markers. Int. J. Tradit. Med. App. 2019, 1, 22–27. [Google Scholar] [CrossRef]
  185. Ramadiana, S.; Hapsoro, D.; Evizal, R.; Setiawan, K.; Karyanto, A.; Yusnita, Y. Genetic diversity among 24 clones of Robusta coffee in Lampung based on RAPD markers. Biodiversitas 2021, 22, d220614. [Google Scholar] [CrossRef]
  186. Sheuly, K.N.; Hoque, M.E.; Syfullah, K.; Bashar, M.A.; Rahman, M.H.; Siddique, A.B. Genetic Diversity Analysis of Garlic (Allium sativum L.) Genotypes Using Rapd Markers. Plant Cell Biotechnol. Mol. Biol. 2022, 13, 61–68. [Google Scholar] [CrossRef]
  187. Hartati, S.; Samanhudi, S.; Cahyono, O. Genetic similarity among Dendrobium species from Indonesia using RAPD markers. Biodiversitas 2023, 24, d240945. [Google Scholar] [CrossRef]
  188. Aydın, A. Determining the genetic diversity of some black cumin genotypes collected in different regions of Türkiye using RAPD markers. Int. J. Agric. Environ. Food Sci. 2024, 8, 294–300. [Google Scholar] [CrossRef]
  189. Wu, F.; Zhang, S.; Gao, Q.; Liu, F.; Wang, J.; Wang, X. Genetic diversity and population structure analysis in a large collection of Vicia amoena in China with newly developed SSR markers. BMC Plant Biol. 2021, 21, 1–12. [Google Scholar]
  190. Li, X.; Qiao, L.; Chen, B.; Zheng, Y.; Zhi, C.; Zhang, S.; Pan, Y.; Cheng, Z. SSR markers development and their application in genetic diversity evaluation of garlic (Allium sativum) germplasm. Plant Divers. 2022, 44, 481–491. [Google Scholar] [CrossRef] [PubMed]
  191. Gogoi, A.; Munda, S.; Paw, M.; Begum, T.; Siddiqui, M.H.; Gaafar, A.-R.Z.; Kesawat, M.S.; Lal, M. Molecular genetic divergence analysis amongst high curcumin lines of Golden Crop (Curcuma longa L.) using SSR marker and use in trait-specific breeding. Sci. Rep. 2023, 13, 19690. [Google Scholar] [CrossRef] [PubMed]
  192. Kojima, Y.; Ebana, K.; Fukuoka, S.; Nagamine, T.; Kawase, M. Development of an RFLP-based rice diversity research set of germplasm. Breed. Sci. 2005, 55, 431–440. [Google Scholar] [CrossRef]
  193. Kunihisa, M.; Fukino, N.; Matsumoto, S. Development of PCR-RFLP marker on strawberry and the identification of Cultivars and their progeny. Acta Hortic. 2006, 708, 517–522. [Google Scholar] [CrossRef]
  194. Mir, J.I.; Shahidul, I.; Rajdeep, K. Evaluation of genetic diversity in Brassica juncea (L.) using protein profiling and molecular marker (RFLP). Int. J. Plant Breed. Genet. 2015, 9, 77–85. [Google Scholar]
  195. Cruz, V.M.V.; Kilian, A.; Dierig, D.A. Development of DArT marker platforms and genetic diversity assessment of the US collection of the new oilseed crop Lesquerella and related species. PLoS ONE 2013, 8, e64062. [Google Scholar] [CrossRef] [PubMed]
  196. Lukanda, M.M.; Dramadri, I.O.; Adjei, E.A.; Edema, R.; Ssemakula, M.O.; Tukamuhabwa, P.; Tusiime, G. Genetic Diversity and Population Structure of Ugandan Soybean (Glycine max L.) Germplasm Based on DArTseq. Plant Mol. Biol. Rep. 2023, 41, 417–426. [Google Scholar] [CrossRef]
  197. Gbedevi, K.M.; Boukar, O.; Ishikawa, H.; Abe, A.; Ongom, P.O.; Unachukwu, N.; Rabbi, I.; Fatokun, C. Genetic diversity and population structure of cowpea [Vigna unguiculata (L.) Walp.] germplasm collected from Togo based on DArT markers. Genes 2021, 12, 1451. [Google Scholar] [CrossRef] [PubMed]
  198. Malebe, M.; Mphangwe, N.; Myburg, A.; Apostolides, Z. Assessment of genome-wide DArT-seq markers for tea Camellia sinensis (L.) O. Kuntze germplasm analysis. Tree Genet. Genomes 2019, 15, 48. [Google Scholar] [CrossRef]
  199. Nutthapornnitchakul, S.; Peyachoknagul, S.; Sangin, P.; Kongbungkerd, A.; Punjansing, T.; Nakkuntod, M. Genetic relationship of orchids in the Calanthe group based on sequence-related amplified polymorphism markers and development of sequence-characterized amplified regions markers for some genus/species identification. Agric. Nat. Resour. 2019, 53, 340–347. [Google Scholar]
  200. Mingyue, T.U.; Yali, H.E.; Xiaoli, L.I.; Ying, Z.O.U.; Xiaojun, Y.U.A.N. Development of SCAR markers related to heat tolerance in Kentucky bluegrass. Not. Bot. Horti Agrobot. Cluj-Napoca 2020, 48, 509–522. [Google Scholar]
  201. Zheng, K.; Cai, Y.; Chen, W.; Gao, Y.; Jin, J.; Wang, H.; Feng, S.; Lu, J. Development, identification, and application of a germplasm specific SCAR Marker for Dendrobium officinale Kimura et Migo. Front. Plant Sci. 2021, 12, 669458. [Google Scholar] [CrossRef] [PubMed]
  202. Qv, M.; Feng, G.; Chen, S.; Chen, H.; Chen, C.; Wang, F.; Lv, S.; Dai, L.; Liu, H.; Huang, B.; et al. The development and utilization of two SCAR markers linked to the resistance of banana (Musa spp. AAA) to Fusarium oxysporum f. sp. cubense race 4. Euphytica 2024, 220, 69. [Google Scholar] [CrossRef]
  203. Ravi, D.; Siril, E.A.; Nair, B.R. SCAR Marker Development for the Identification of Elite Germplasm of Moringa Oleifera Lam.-A Never Die Plant. Plant Mol. Biol. Rep. 2021, 39, 850–861. [Google Scholar] [CrossRef]
  204. Shiferaw, E.; Porceddu, E.; Pé, E.; Ponnaiah, M. Application of CAPS markers for diversity assessment in grass pea (L.). Biodivers. Res. Conserv. 2017, 48, 11–18. [Google Scholar] [CrossRef]
  205. Osae, B.A.; Amanullah, S.; Liu, H.; Liu, S.; Saroj, A.; Zhang, C.; Liu, T.; Gao, P.; Luan, F. CAPS marker-base genetic linkage mapping and QTL analysis for watermelon ovary, fruit and seed-related traits. Euphytica 2022, 218, 39. [Google Scholar] [CrossRef]
  206. Kang, J.-N.; Lee, G.-H.; Yu, J.; Choi, M.-H.; Lee, S. Development of Cleaved Amplified Polymorphic Sequence Markers for Classifying Ginger (Zingiber officinale) Cultivars Using Reference Sequencing. Plant Breed. Biotechnol. 2023, 11, 130–140. [Google Scholar] [CrossRef]
  207. Bongiorno, G.; Di Noia, A.; Ciancaleoni, S.; Marconi, G.; Cassibba, V.; Albertini, E. Development and Application of a Cleaved Amplified Polymorphic Sequence Marker (Phyto) Linked to the Pc5.1 Locus Conferring Resistance to Phytophthora capsici in Pepper (Capsicum annuum L.). Plants 2023, 12, 2757. [Google Scholar] [CrossRef]
  208. Martiwi, I.N.A.; Nugroho, L.H.; Daryono, B.S.; Susandarini, R. Genotypic variability and relationships of Sorghum bicolor accessions from Java Island, Indonesia based on IRAP markers. Biodiversitas 2020, 21, d211220. [Google Scholar] [CrossRef]
  209. Dongare, M.D.; Alex, S.; Soni, K.B.; Sindura, K.P.; Nair, D.S.; Stephen, R.; Jose, E. Cross-species transferability of IRAP retrotransposon markers and polymorphism in black pepper (Piper nigrum L.). Genet. Resour. Crop Evol. 2023, 70, 2593–2605. [Google Scholar] [CrossRef]
  210. Zayed, E.M.; Ghonaim, M.M.; Attya, A.M.; Morsi, N.A.; Hussein, K.A. IRAP-PCR technique for determining the biodiversity between Egyptian barley cultivars. Egyptian J. Bot. 2022, 62, 359–370. [Google Scholar] [CrossRef]
  211. Voronova, A.; Ruņģis, D. Development and Characterisation of Irap Markers From Expressed Retrotransposon-like sequences in Pinus sylvestris L. Proc. Latv. Acad. Sci. Sect. B Nat. Exact Appl. Sci. 2014, 67, 485–492. [Google Scholar]
  212. Stepanov, I.; Balapanov, I.; Drygina, A. Search of effective IRAP markers for sakura genotyping. BIO Web Conf. 2020, 25, 03006. [Google Scholar] [CrossRef]
  213. Aouadi, M.; Guenni, K.; Abdallah, D.; Louati, M.; Chatti, K.; Baraket, G.; Salhi Hannachi, A. Conserved DNA-derived polymorphism, new markers for genetic diversity analysis of Tunisian Pistacia vera L. Physiol. Mol. Biol. Plants 2019, 25, 1211–1223. [Google Scholar] [CrossRef] [PubMed]
  214. Igwe, D.O.; Ihearahu, O.C.; Osano, A.A.; Acquaah, G.; Ude, G.N. Genetic diversity and population assessment of Musa L. (Musaceae) employing CDDP markers. Plant Mol. Biol. Rep. 2021, 39, 801–820. [Google Scholar] [CrossRef]
  215. Klongová, L.; Kyseľ, M.; Fialková, V.; Fernández-Cusimamani, E.; Kovacik, A.; Ziarovska, J. Utilization of CDDP Markers in analysis of genetic variability of Arachis hypogaea L. J. Microbiol. Biotechnol. Food Sci. 2023, 13, E9884. [Google Scholar] [CrossRef]
  216. Ma, M.; Yan, Z.; Lu, B. Assessment of genetic diversity of the medicinal and aromatic crop, Amomum Tsao-Ko, using paap and CDDP markers. Agriculture 2022, 12, 1536. [Google Scholar] [CrossRef]
  217. Saleh, B. Genetic diversity of Origanum syriacum L. (Lamiaceae) species through Touch-Up Direct Amplification of Minisatellite-region DNA (TU-DAMD) marker. Not. Sci. Biol. 2022, 14, 11174. [Google Scholar] [CrossRef]
  218. Saleh, B. Genetic Diversity of Salvia officinalis L. (Lamiaceae) and its Related Species using TU-DAMD Analysis. Open Agric. J. 2023, 17, e187433152305080. [Google Scholar] [CrossRef]
  219. Saleh, B. Molecular characterization using Directed Amplification of Minisatellite-region DNA (DAMD) Marker in Ficus sycomorus L.(Moraceae). Open Agric. J. 2019, 13, 74–81. [Google Scholar] [CrossRef]
  220. Bhatt, J.; Kumar, S.; Patel, S.; Solanki, R. Sequence-related amplified polymorphism (SRAP) markers based genetic diversity analysis of cumin genotypes. Ann. Agrarian Sci. 2017, 15, 434–438. [Google Scholar] [CrossRef]
  221. Wang, X.; Gao, C.; Li, K. Strategy for constructing Pinus yunnanensis germplasm bank for timber based on the SRAP molecular marker. Plant Sci. J. 2019, 37, 211–220. [Google Scholar]
  222. Zagorcheva, T.; Stanev, S.; Rusanov, K.; Atanassov, I. SRAP markers for genetic diversity assessment of lavender (Lavandula angustifolia mill.) varieties and breeding lines. Biotechnol. Biotechnol. Equip. 2020, 34, 303–308. [Google Scholar] [CrossRef]
  223. Zhang, C.; Zhu, L.; Wang, M.; Tang, Y.; Zhou, H.; Sun, Q.; Yu, Q.; Zhang, J. Evaluation of SRAP markers efficiency in genetic diversity of Aspergillus flavus from peanut-cropped soils in China. OCS 2022, 7, 135–141. [Google Scholar] [CrossRef]
  224. Fareghi, S.; Mirlohi, A.F.; Saeidi, G.; Khamisabadi, H. Evaluation of SRAP marker efficiency in identifying the relationship between genetic diversities of corn inbred lines with seed quantity and quality in derived hybrids. CMB 2019, 65, 6–14. [Google Scholar] [CrossRef]
  225. Suparman, S.; Pornpongrungrueng, P.; Chantaranothai, P. Molecular studies of coralberry (Ardisia crenata Sims; Primulaceae) from Thailand based on SCoT markers. Biodiversitas 2023, 24, d240611. [Google Scholar] [CrossRef]
  226. Chňapek, M.; Mikolášova, L.; Vivodík, M.; Gálová, Z.; Hromadová, Z.; Ražná, K.; Balážová, Ž. Genetic diversity of oat genotypes using SCoT markers. Biol. Life Sci. Forum 2021, 11, 29. [Google Scholar] [CrossRef]
  227. Gawroński, J.; Dyduch-Siemińska, M. Potential of In Vitro Culture of Scutellaria baicalensis in the Formation of Genetic Variation Confirmed by ScoT Markers. Genes 2022, 13, 2114. [Google Scholar] [CrossRef] [PubMed]
  228. Mirzahosein-Tabrizi, M.; Ghanavati, F.; Azizinezhad, R.; Etminan, A. Genetic diversity revealed by phytochemical and molec-ular analyses among and within eight Trigonella sp. J. Crop. Sci. Biotechnol. 2022, 26, 345–357. [Google Scholar] [CrossRef]
  229. Emam, M.A.; Abd El-Mageed, A.M.; Niedbała, G.; Sabrey, S.A.; Fouad, A.S.; Kapiel, T.; Piekutowska, M.; Mahmoud, S.A. Genetic characterization and agronomic evaluation of drought tolerance in an Egyptian wheat (Triticum aestivum L.) Cultivars. Agronomy 2022, 12, 1217. [Google Scholar] [CrossRef]
  230. Zhang, X.; Sun, X.; Wang, J.; Xue, M.; Sun, C.; Dong, W. Evaluation of molecular and phenotypic diversity of Crataegus bretschneideri CK Schneid. and related species in China. Genet. Resour. Crop. Evol. 2023, 70, 221–234. [Google Scholar] [CrossRef]
  231. Singh, H.K.; Parveen, I.; Raghuvanshi, S.; Babbar, S.B. The loci recommended as universal barcodes for plants on the basis of floristic studies may not work with congeneric species as exemplified by DNA barcoding of Dendrobium species. BMC Res. Notes 2012, 5, 42. [Google Scholar] [CrossRef]
  232. Feng, S.; Jiang, M.; Shi, Y.; Jiao, K.; Shen, C.; Lu, J.; Ying, Q.; Wang, H. Application of the ribosomal DNA ITS2 region of Physalis (Solanaceae): DNA barcoding and phylogenetic study. Front. Plant Sci. 2016, 7, 1047. [Google Scholar] [CrossRef] [PubMed]
  233. Zhang, D.; Jiang, B. Species identification in complex groups of medicinal plants based on DNA barcoding: A case study on Astragalus spp. (Fabaceae) from southwest China. Conserv. Genet. Resour. 2020, 12, 469–478. [Google Scholar] [CrossRef]
  234. Kantar, F.; Yemşen, S.N.; Bülbül, C.; Yilmaz, N.; Mutlu, N. Phenotypic and iPBS-retrotransposon marker diversity in okra (Abelmoschus esculentus (L.) Moench) germplasm. Biotech Stud. 2021, 30, 7–15. [Google Scholar] [CrossRef]
  235. Eren, B.; Keskin, B.; Demirel, F.; Demirel, S.; Türkoğlu, A.; Yilmaz, A.; Haliloğlu, K. Assessment of genetic diversity and population structure in local alfalfa genotypes using iPBS molecular markers. Genet. Resour. Crop Evol. 2023, 70, 617–628. [Google Scholar] [CrossRef]
  236. Haliloğlu, K.; Türkoğlu, A.; Öztürk, H.I.; Özkan, G.; Elkoca, E.; Poczai, P. iPBS-Retrotransposon Markers in the Analysis of Genetic Diversity among Common Bean (Phaseolus vulgaris L.) Germplasm from Türkiye. Genes 2022, 13, 1147. [Google Scholar] [CrossRef] [PubMed]
  237. Demirel, F.; Yıldırım, B.; Eren, B.; Demirel, S.; Türkoğlu, A.; Haliloğlu, K.; Nowosad, K.; Bujak, H.; Bocianowski, J. Revealing Genetic Diversity and Population Structure in Türkiye’s Wheat Germplasm Using iPBS-Retrotransposon Markers. Agronomy. 2024, 14, 300. [Google Scholar] [CrossRef]
  238. Sameeullah, M.; Kayaçetin, F.; Khavar, K.M.; Perkasa, A.Y.; Maesaroh, S.; Waheed, M.T.; Çiftçi, V. Decoding genetic diversity and population structure of Brassica species by inter primer binding site (iPBS) retrotransposon markers. Genet. Resour. Crop Evol. 2024. [Google Scholar] [CrossRef]
  239. Orhan, E.; Kara, D. Use of retrotransposon based iPBS markers for determination of genetic relationship among some Chestnut Cultivars (Castanea sativa Mill.) in Türkiye. Mol. Biol. Rep. 2023, 50, 8397–8405. [Google Scholar] [CrossRef] [PubMed]
  240. Etminan, A.; Pour-Aboughadareh, A.; Mehrabi, A.A.; Shooshtari, L.; Ahmadi-Rad, A.; Moradkhani, H. Molecular characterization of the wild relatives of wheat using CAAT-box derived polymorphism. Plant Biosyst. 2019, 153, 398–405. [Google Scholar] [CrossRef]
  241. Wang, L.; Li, G.; Peña, R.J.; Xia, X.; He, Z. Development of STS markers and establishment of multiplex PCR for Glu-A3 alleles in common wheat (Triticum aestivum L.). J. Cereal Sci. 2010, 51, 305–312. [Google Scholar] [CrossRef]
  242. Lee, S.; Heo, H.; Kwon, Y.; Lee, B. Development of gene-based STS markers in wheat. Korean J. Crop Sci. 2012, 57, 71–77. [Google Scholar] [CrossRef]
  243. Dewi, A.K.; Rahayu, S.; Dwimahyani, I.; Reflinur, R. Analysis of yield and genetic diversity among Kewal local rice mutant lines based on STS markers. AIP Conf. Proc. 2021, 2381, 020012. [Google Scholar]
  244. Zhang, J.; Liu, W.; Lu, Y.; Liu, Q.; Yang, X.; Li, X.; Li, L. A resource of large-scale molecular markers for monitoring Agropyron cristatum chromatin introgression in wheat background based on transcriptome sequences. Sci. Rep. 2017, 7, 11942. [Google Scholar] [CrossRef]
  245. Wang, C.M.; Li, L.H.; Zhang, X.T.; Gao, Q.; Wang, R.F.; An, D.G. Development and application of EST-STS markers specific to chromosome 1RS of Secale cereale. Cereal Res. Commun. 2009, 37, 13–21. [Google Scholar] [CrossRef]
  246. Qiao, L.; Liu, S.; Li, J.; Li, S.; Yu, Z.; Liu, C.; Li, X.; Liu, J.; Ren, Y.; Zhang, P.; et al. Development of sequence-tagged site marker set for identification of J, JS, and St sub-genomes of Thinopyrum intermedium in wheat background. Front. Plant Sci. 2021, 12, 685216. [Google Scholar] [CrossRef]
  247. Srivastava, N.; Bajpai, A.; Chandra, R.; Rajan, S.; Muthukumar, M.; Srivastava, M.K. Comparison of PCR based marker systems for genetic analysis in different cultivars of mango. J. Environ. Biol. 2012, 33, 159. [Google Scholar]
  248. Talebi, R.; Nosrati, S.; Etminan, A.; Naji, A.M. Genetic diversity and population structure analysis of landrace and improved safflower (Cartamus tinctorious L.) germplasm using arbitrary functional gene-based molecular markers. Biotechnol. Biotechnol. Equip. 2018, 32, 1183–1194. [Google Scholar] [CrossRef]
  249. Aravanopoulos, F.A. Clonal identification based on quantitative, codominant, and dominant marker data: A comparative analysis of selected willow (Salix L.) clones. Int. J. For. Res. 2010, 2010, 906310. [Google Scholar] [CrossRef]
  250. Garcia, A.A.; Benchimol, L.L.; Barbosa, A.M.; Geraldi, I.O.; Souza Jr, C.L.; Souza, A.P.D. Comparison of RAPD, RFLP, AFLP and SSR markers for diversity studies in tropical maize inbred lines. Genet. Mol. Biol. 2004, 27, 579–588. [Google Scholar] [CrossRef]
  251. Ho, W.S.; Wickneswari, R.; Mahani, M.C.; Shukor, M.N. Comparative genetic diversity studies of Shorea curtisii (Dipterocarpaceae): An assessment using SSR and DAMD markers. J. Trop. For. Sci. 2006, 22–35. [Google Scholar]
  252. Luo, Y.; Zhang, X.; Xu, J.; Zheng, Y.; Pu, S.; Duan, Z.; Li, Z.; Liu, G.; Chen, J.; Wang, Z. Phenotypic and molecular marker analysis uncovers the genetic diversity of the grass Stenotaphrum secundatum. BMC Genet. 2020, 21, 86. [Google Scholar] [CrossRef] [PubMed]
  253. Kumar, A.; Kumar, S.; Kumar, R.; Kumar, V.; Prasad, L.; Kumar, N.; Singh, D. Identification of blast resistance expression in rice genotypes using molecular markers (RAPD & SCAR). Afr. J. Biotechnol. 2010, 9, 3501–3509. [Google Scholar]
  254. Basu, A.; Ghosh, M.; Meyer, R.; Powell, W.; Basak, S.L.; Sen, S.K. Analysis of genetic diversity in cultivated jute determined by means of SSR markers and AFLP profiling. Crop Sci. 2004, 44, 678–685. [Google Scholar] [CrossRef]
  255. Ahmed, N.; Mir, J.I.; Mir, R.R.; Rather, N.A.; Rashid, R.; Wani, S.H.; Shafi, W.; Mir, H.; Sheikh, M.A. SSR and RAPD analysis of genetic diversity in walnut (Juglans regia L.) genotypes from Jammu and Kashmir, India. Physiol. Mol. Biol. Plants 2012, 18, 149–160. [Google Scholar] [CrossRef] [PubMed]
  256. Zargar, S.M.; Farhat, S.; Mahajan, R.; Bhakhri, A.; Sharma, A. Unraveling the efficiency of RAPD and SSR markers in diversity analysis and population structure estimation in common bean. Saudi J. Biol. Sci. 2016, 23, 139–149. [Google Scholar] [CrossRef] [PubMed]
  257. Dar, A.A.; Mudigunda, S.; Mittal, P.K.; Arumugam, N. Comparative assessment of genetic diversity in Sesamum indicum L. using RAPD and SSR markers. 3 Biotech 2017, 7, 10. [Google Scholar] [CrossRef] [PubMed]
  258. Nascimento, W.F.; Rodrigues, J.F.; Koehler, S.; Gepts, P.; Veasey, E.A. Spatially structured genetic diversity of the Amerindian yam (Dioscorea trifida L.) assessed by SSR and ISSR markers in Southern Brazil. Genet. Resour. Crop Evol. 2013, 60, 2405–2420. [Google Scholar] [CrossRef]
  259. Hammami, R.; Jouve, N.; Soler, C.; Frieiro, E.; González, J.M. Genetic diversity of SSR and ISSR markers in wild populations of Brachypodium distachyon and its close relatives B. stacei and B.hybridum (Poaceae). Plant Syst. Evol. 2014, 300, 2029–2040. [Google Scholar] [CrossRef]
  260. Ramzan, M.; Sarwar, S.; Kauser, N.; Saba, R.; Hussain, I.; Shah, A.A.; Aslam, M.N.; Alkahtani, J.; Alwahibi, M.S. Assessment of Inter simple sequence repeat (ISSR) and simple sequence repeat (SSR) markers to reveal genetic diversity among Tamarix ecotypes. J. King Saud Univ. Sci. 2020, 32, 3437–3446. [Google Scholar] [CrossRef]
  261. Nazir, M.; Mahajan, R.; Hashim, M.J.; Iqbal, J.; Alyemeni, M.N.; Ganai, B.A.; Zargar, S.M. Deciphering allelic variability and population structure in buckwheat: An analogy between the efficiency of ISSR and SSR markers. Saudi J. Biol. Sci. 2021, 28, 6050–6056. [Google Scholar]
  262. Papaioannou, C.; Fassou, G.; Petropoulos, S.A.; Lamari, F.N.; Bebeli, P.J.; Papasotiropoulos, V. Evaluation of the Genetic Diversity of Greek Garlic (Allium sativum L.) Accessions Using DNA Markers and Association with Phenotypic and Chemical Variation. Agriculture 2023, 13, 1408. [Google Scholar] [CrossRef]
  263. Liu, S.; Feuerstein, U.; Luesink, W.; Schulze, S.; Asp, T.; Studer, B.; Becker, H.C.; Dehmer, K.J. DArT, SNP, and SSR analyses of genetic diversity in Lolium perenne L. using bulk sampling. BMC Genet. 2018, 19, 10. [Google Scholar] [CrossRef] [PubMed]
  264. Jayabalan, S.; Pulipati, S.; Ramasamy, K.; Jaganathan, D.; Venkatesan, S.D.; Vijay, G.; Kumari, K.; Raju, K.; Hariharan, G.N.; Venkataraman, G. Analysis of genetic diversity and population structure using SSR markers and validation of a Cleavage Amplified Polymorphic Sequences (CAPS) marker involving the sodium transporter OsHKT1;5 in saline tolerant rice (Oryza sativa L.) landraces. Gene 2019, 713, 143976. [Google Scholar] [CrossRef]
  265. Shahnazari, N.; Noormohammadi, Z.; Sheidai, M.; Koohdar, F. A new insight on genetic diversity of sweet oranges: CAPs-SSR and SSR markers. J. Genet. Eng. Biotechnol. 2022, 20, 105. [Google Scholar] [CrossRef] [PubMed]
  266. Xanthopoulou, A.; Ganopoulos, I.; Kalivas, A.; Nianiou-Obeidat, I.; Ralli, P.; Moysiadis, T.; Tsaftris, A.; Madesis, P. Comparative analysis of genetic diversity in Greek Genebank collection of summer squash (‘Cucurbita pepo’) landraces using start codon targeted (SCoT) polymorphism and ISSR markers. Aust. J. Crop Sci. 2015, 9, 14–21. [Google Scholar]
  267. Sharma, U.; Rai, M.K.; Shekhawat, N.S.; Kataria, V. Genetic homogeneity revealed in micropropagated Bauhinia racemosa Lam. using gene-targeted markers CBDP and SCoT. Physiol. Mol. Biol. Plants 2019, 25, 581–588. [Google Scholar] [CrossRef]
  268. Ghobadi, G.; Etminan, A.; Mehrabi, A.M.; Shooshtari, L. Molecular diversity analysis in hexaploid wheat (Triticum aestivum L.) and two Aegilops species (Aegilops crassa and Aegilops cylindrica) using CBDP and SCoT markers. J. Genet. Eng. Biotechnol. 2021, 19, 56. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  269. Al-Khayri, J.M.; Mahdy, E.M.B.; Taha, H.S.A.; Eldomiaty, A.S.; Abd-Elfattah, M.A.; Abdel Latef, A.A.H.; Rezk, A.A.; Shehata, W.F.; Almaghasla, M.I.; Shalaby, T.A.; et al. Genetic and Morphological Diversity Assessment of Five Kalanchoe Genotypes by SCoT, ISSR and RAPD-PCR Markers. Plants 2022, 11, 1722. [Google Scholar] [CrossRef] [PubMed]
  270. Osman, S.A.; Ali, H.B. Research Article Genetic Diversity of Five Lathyrus Species using RAPD, ISSR and SCoT Markers. Asian J. Plant Sci. 2020, 19, 152–165. [Google Scholar] [CrossRef]
  271. Sun, Q.B.; Li, L.F.; Li, Y.; Wu, G.J.; Ge, X.J. SSR and AFLP markers reveal low genetic diversity in the biofuel plant Jatropha curcas in China. Crop Sci. 2008, 48, 1865–1871. [Google Scholar] [CrossRef]
  272. Cao, J.; Zhou, Z.; Tu, J.; Cheng, S.; Yao, J.; Xu, F.; Wang, G.; Zhang, J.; Ye, J.; Liao, Y.; et al. Genetic diversity and population structure analysis of sand pear (Pyrus pyrifolia) ‘Nakai’varieties using SSR and AFLP markers. Not. Bot. Horti Agrobot. Cluj-Napoca 2019, 47, 970–979. [Google Scholar] [CrossRef]
  273. Saghir, K.; Abdelwahd, R.; Iraqi, D.; Lebkiri, N.; Gaboun, F.; El Goumi, Y.; Ibrahimi, M.; Abbas, Y.; Diria, G. Assessment of genetic diversity among wild rose in Morocco using ISSR and DAMD markers. J. Genet. Eng. Biotechnol. 2022, 20, 150. [Google Scholar] [CrossRef] [PubMed]
  274. Adu, B.G.; Akromah, R.; Amoah, S.; Nyadanu, D.; Yeboah, A.; Aboagye, L.M.; Amoah, R.A.; Owasu, E.G. High-density DArT-based Silico DArT, and SNP markers for genetic diversity and population structure studies in cassava (Manihot esculenta Crantz). PLoS ONE 2021, 16, e0255290. [Google Scholar] [CrossRef] [PubMed]
  275. Shaibu, A.S.; Ibrahim, H.; Miko, Z.L.; Mohammed, I.B.; Mohammed, S.G.; Yusuf, H.L.; Kamara, A.Y.; Omoigui, L.O.; Karikari, B. Assessment of the genetic structure and diversity of soybean (Glycine max L.) germplasm using diversity array technology and single nucleotide polymorphism markers. Plants 2021, 11, 68. [Google Scholar] [CrossRef] [PubMed]
  276. Sun, J.; Wang, J.; Su, D.; Yang, J.; Wang, E.; Wu, S.; Li, M.; Ma, L. Discrimination of tobacco cultivars using SCAR and RAPD markers. Czech J. Genet. Plant Breed. 2020, 56, 170–173. [Google Scholar] [CrossRef]
  277. Dou, J.; Lu, X.; Ali, A.; Zhao, S.; Zhang, L.; He, N.; Liu, W. Genetic mapping reveals a marker for yellow skin in watermelon (Citrullus lanatus L.). PLoS ONE 2018, 13, e0200617. [Google Scholar] [CrossRef] [PubMed]
  278. Katzir, N.; Tadmor, Y.; Tzuri, G.; Leshzeshen, E.; Mozes-Daube, N.; Danin-Poleg, Y.; Paris, H.S. Further ISSR and preliminary SSR analysis of relationships among accessions of Cucurbita pepo. In Proceedings of the VII Eucarpia Meeting on Cucurbit Genetics and Breeding, Ma’ale Ha Hamisha, Israel, 19–23 March 2000; Volume 510, pp. 433–440. [Google Scholar]
  279. Sagar, M.R.; Kumar, S.; Patidar, D.; Sakure, A.A. Morphological, physico-biochemical and marker-based diversity of desi cotton (Gossypium herbaceum L.) germplasm. J. King Saud Univ. Sci. 2022, 34, 102336. [Google Scholar] [CrossRef]
  280. Ghasemzadeh Baraki, S.; Nikzat Siahkolaee, S. Assessment of SCoT and DAMD molecular markers in genetic diversity and species delimitation of three moss species grown in Iran. Iran. J. Genet. Plant Breed. 2018, 7, 33–41. [Google Scholar]
  281. Alaaddin Ahmed, A.; Anwar Qadir, S.; Tahir, N.A.R. CDDP and ISSR markers-assisted diversity and structure analysis in Iraqi Mazu (Quercus infectoria Oliv.) accessions. All Life 2022, 15, 247–261. [Google Scholar] [CrossRef]
  282. Premjet, D.; Boonsrangsom, T.; Sujipuli, K.; Rattanasut, K.; Kongbungkerd, A.; Premjet, S. Morphological and Molecular Characterizations of Musa (ABB)’Mali-Ong’in Thailand. Biology 2022, 11, 1429. [Google Scholar] [CrossRef] [PubMed]
  283. Hu, C.Y.; Tsai, Y.Z.; Lin, S.F. Development of STS and CAPS markers for variety identification and genetic diversity analysis of tea germplasm in Taiwan. Bot. Stud. 2014, 55, 12. [Google Scholar] [CrossRef] [PubMed]
  284. Guo, Y.; Zhai, L.; Long, H.; Chen, N.; Gao, C.; Ding, Z.; Jin, B. Genetic diversity of Bletilla striata assessed by SCoT and IRAP markers. Hereditas 2018, 155, 35. [Google Scholar] [CrossRef] [PubMed]
  285. Kumar, P.; Gupta, V.K.; Misra, A.K.; Modi, D.R.; Pandey, B.K. Potential of molecular markers in plant biotechnology. Plant Omics 2009, 2, 141–162. [Google Scholar]
  286. Jonah, P.M.; Bello, L.L.; Lucky, O.; Midau, A.; Moruppa, S.M. The importance of molecular markers in plant breeding programmes. Glob. J. Sci. Front. Res. 2011, 11, 5–12. [Google Scholar]
  287. Varshney, R.K.; Mahendar, T.; Aggarwal, R.K.; Börner, A. Genic Molecular Markers in Plants: Development and Applications. In Genomics-Assisted Crop Improvement; Varhney, R.K., Tuberosa, R., Eds.; Springer: Dordrecht, The Netherland, 2007; Volume 1, pp. 13–29. ISBN 9781402062957. [Google Scholar]
  288. Jin, J.; Lu, P.; Xu, Y.; Tao, J.; Li, Z.; Wang, S.; Yu, S.; Wang, C.; Xie, X.; Gao, J.; et al. PCMDB: A curated and comprehensive resource of plant cell markers. Nucleic Acids Res. 2022, 50, D1448–D1455. [Google Scholar] [CrossRef] [PubMed]
  289. Guo, Z.; Wang, H.; Tao, J.; Ren, Y.; Xu, C.; Wu, K.; Zou, C.; Zhang, J.; Xu, Y. Development of multiple SNP marker panels affordable to breeders through genotyping by target sequencing (GBTS) in maize. Mol. Breed. 2019, 39, 37. [Google Scholar] [CrossRef]
  290. Jagtap, A.B.; Vikal, Y.; Johal, G.S. Genome-wide development and validation of cost-effective KASP marker assays for genetic dissection of heat stress tolerance in maize. Int. J Mol. Sci. 2020, 21, 7386. [Google Scholar] [CrossRef] [PubMed]
  291. Mammadov, J.; Aggarwal, R.; Buyyarapu, R.; Kumpatla, S. SNP markers and their impact on plant breeding. Int. J. Plant Genom. 2012, 2012, 728398. [Google Scholar] [CrossRef]
  292. Wakeley, J.; Nielsen, R.; Liu-Cordero, S.N.; Ardlie, K. The discovery of single-nucleotide polymorphisms—And inferences about human demographic history. Am. J. Hum. Genet. 2001, 69, 1332–1347. [Google Scholar] [CrossRef] [PubMed]
  293. Morgil, H.; Can Gercek, Y.; Tulum, I. Single nucleotide polymorphisms (SNPs) in plant genetics and breeding. In The Recent Topics in Genetic Polymorphisms; IntechOpen: London, UK, 2020; pp. 400–825. [Google Scholar] [CrossRef]
  294. Hinze, L.L.; Hulse-Kemp, A.M.; Wilson, I.W.; Zhu, Q.H.; Llewellyn, D.J.; Taylor, J.M.; Spriggs, A.; Fang, D.D.; Ulleoa, M.; Burke, J.J.; et al. Diversity analysis of cotton (Gossypium hirsutum L.) germplasm using the CottonSNP63K Array. BMC Plant Biol. 2017, 17, 37. [Google Scholar] [CrossRef] [PubMed]
  295. Mengistu, G.; Shimelis, H.; Laing, M.; Lule, D.; Assefa, E.; Mathew, I. Genetic diversity assessment of sorghum (Sorghum bicolor (L.) Moench) landraces using SNP markers. S. Afr. J. Plant Soil 2020, 37, 220–226. [Google Scholar] [CrossRef]
  296. Oliveira, L.S.D.; Schuster, I.; Novaes, E.; Pereira, W.A. SNP genotyping for fast and consistent clustering of maize inbred lines into heterotic groups. Crop Breed. Appl. Biotechnol. 2021, 21, e367121110. [Google Scholar] [CrossRef]
  297. Niu, S.; Song, Q.; Koiwa, H.; Qiao, D.; Zhao, D.; Chen, Z.; Liu, X.; Wen, X. Genetic diversity, linkage disequilibrium, and population structure analysis of the tea plant (Camellia sinensis) from an origin center, Guizhou plateau, using genome-wide SNPs developed by genotyping-by-sequencing. BMC Plant Biol. 2019, 19, 328. [Google Scholar] [CrossRef] [PubMed]
  298. McCouch, S.R.; Zhao, K.; Wright, M.; Tung, C.W.; Ebana, K.; Thomson, M.; Reynolds, A.; Wang, D.; DeClerck, G.; Ali, M.L.; et al. Development of genome-wide SNP assays for rice. Breed. Sci. 2010, 60, 524–535. [Google Scholar] [CrossRef]
  299. Huang, X.; Feng, Q.; Qian, Q.; Zhao, Q.; Wang, L.; Wang, A.; Guan, J.; Fan, D.; Weng, Q.; Huang, T.; et al. High-throughput genotyping by whole-genome resequencing. Genome Res. 2009, 19, 1068–1076. [Google Scholar] [CrossRef] [PubMed]
  300. Tang, W.; Wu, T.; Ye, J.; Sun, J.; Jiang, Y.; Yu, J.; Tang, J.; Chen, G.; Wang, C.; Wan, J. SNP-based analysis of genetic diversity reveals important alleles associated with seed size in rice. BMC Plant Biol. 2016, 16, 93. [Google Scholar]
  301. Poland, J.A.; Rife, T.W. Genotyping-by-sequencing for plant breeding and genetics. Plant Genome 2012, 5, 92–102. [Google Scholar] [CrossRef]
  302. Fu, Y.B. Genetic diversity analysis of highly incomplete SNP genotype data with imputations: An empirical assessment. G3-Genes Genom Genet. 2014, 4, 891–900. [Google Scholar] [CrossRef] [PubMed]
  303. Montero-Pau, J.; Blanca, J.; Esteras, C.; Martínez-Pérez, E.M.; Gómez, P.; Monforte, A.J.; Cañizares, J.; Picó, B. An SNP-based saturated genetic map and QTL analysis of fruit-related traits in Zucchini using Genotyping-by-sequencing. BMC Genom. 2017, 18, 94. [Google Scholar] [CrossRef] [PubMed]
  304. Imai, A.; Nonaka, K.; Kuniga, T.; Yoshioka, T.; Hayashi, T. Genome-wide association mapping of fruit-quality traits using genotyping-by-sequencing approach in citrus landraces, modern cultivars, and breeding lines in Japan. Tree Genet. Genomes 2018, 14, 24. [Google Scholar] [CrossRef]
  305. Tomar, V.; Dhillon, G.S.; Singh, D.; Singh, R.P.; Poland, J.; Joshi, A.K.; Tiwari, B.S.; Kumar, U. Elucidating SNP-based genetic diversity and population structure of advanced breeding lines of bread wheat (Triticum aestivum L.). PeerJ 2021, 9, e11593. [Google Scholar] [CrossRef] [PubMed]
  306. Díaz, B.G.; Zucchi, M.I.; Alves-Pereira, A.; de Almeida, C.P.; Moraes, A.C.L.; Vianna, S.A.; Azevedo-Filho, J.; Colombo, C.A. Genome-wide SNP analysis to assess the genetic population structure and diversity of Acrocomia species. PLoS ONE 2021, 16, e0241025. [Google Scholar] [CrossRef] [PubMed]
  307. Dang, Z.; Li, J.; Liu, Y.; Song, M.; Lockhart, P.J.; Tian, Y.; Niu, M.; Wang, Q. RADseq-based population genomic analysis and environmental adaptation of rare and endangered recretohalophyte Reaumuria trigyna. Plant Genom. 2023, 17, e20303. [Google Scholar] [CrossRef]
  308. Serba, D.D.; Muleta, K.T.; St. Amand, P.; Bernardo, A.; Bai, G.; Perumal, R.; Bashir, E. Genetic diversity, population structure, and linkage disequilibrium of pearl millet. Plant Genome 2019, 12, 180091. [Google Scholar] [CrossRef] [PubMed]
  309. Wang, B.; Lin, Z.; Li, X.; Zhao, Y.; Zhao, B.; Wu, G.; Ma, X.; Wang, H.; Xie, Y.; Li, Q.; et al. Genome-wide selection and genetic improvement during modern maize breeding. Nat. Genet. 2020, 52, 565–571. [Google Scholar] [CrossRef] [PubMed]
  310. Slonecki, T.J.; Rutter, W.B.; Olukolu, B.A.; Yencho, G.C.; Jackson, D.M.; Wadl, P.A. Genetic diversity, population structure, and selection of breeder germplasm subsets from the USDA sweet potato (Ipomoea batatas) collection. Front. Plant Sci. 2023, 13, 1022555. [Google Scholar] [CrossRef] [PubMed]
  311. Dube, S.P.; Sibiya, J.; Kutu, F. Genetic diversity and population structure of maize inbred lines using phenotypic traits and single nucleotide polymorphism (SNP) markers. Sci. Rep. 2023, 13, 17851. [Google Scholar] [CrossRef] [PubMed]
  312. Huang, J.; Liu, Y.; Han, F.; Fang, Z.; Yang, L.; Zhuang, M.; Zhang, Y.; Lv, H.; Wong, Y.; Ji, J.; et al. Genetic diversity and population structure analysis of 161 broccoli cultivars based on SNP markers. Hortic. Plant J. 2021, 7, 423–433. [Google Scholar] [CrossRef]
  313. Zhang, L.; Huang, Y.W.; Huang, J.L.; Ya, J.D.; Zhe, M.Q.; Zeng, C.X.; Zhang, Z.R.; Zhang, S.B.; Li, D.Z.; Li, H.T.; et al. DNA barcoding of Cymbidium by genome skimming: Call for next-generation nuclear barcodes. Mol. Ecol. Resour. 2023, 23, 424–439. [Google Scholar] [CrossRef] [PubMed]
  314. ŞapcıSelamoğlu, H. DNA barcoding of two narrow endemic plants; Astragalus argaeus and Astragalus stenosemioides from Mount Erciyes, Turkey. Conserv. Genet. Resour. 2022, 14, 81–84. [Google Scholar] [CrossRef]
  315. You, Q.; Yang, X.; Peng, Z.; Xu, L.; Wang, J. Development and applications of a high throughput genotyping tool for polyploid crops: Single nucleotide polymorphism (SNP) array. Front. Plant Sci. 2018, 9, 104. [Google Scholar] [CrossRef] [PubMed]
  316. Thomson, M.J. High-throughput SNP genotyping to accelerate crop improvement. Plant Breed. Biotechnol. 2014, 2, 195–212. [Google Scholar] [CrossRef]
  317. LaFramboise, T. Single nucleotide polymorphism arrays: A decade of biological, computational and technological advances. Nucleic Acids Res. 2009, 37, 4181–4193. [Google Scholar] [CrossRef]
  318. Heslot, N.; Rutkoski, J.; Poland, J.; Jannink, J.L.; Sorrells, M.E. Impact of marker ascertainment bias on genomic selection accuracy and estimates of genetic diversity. PLoS ONE 2013, 8, e74612. [Google Scholar] [CrossRef] [PubMed]
  319. Albrechtsen, A.; Nielsen, F.C.; Nielsen, R. Ascertainment biases in SNP chips affect measures of population divergence. Mol. Biol. Evol. 2010, 27, 2534–2547. [Google Scholar] [CrossRef] [PubMed]
  320. Chat, V.; Ferguson, R.; Morales, L.; Kirchhoff, T. Ultra-low-coverage whole-genome sequencing as an alternative to genotyping arrays in genome-wide association studies. Front. Genet. 2022, 12, 790445. [Google Scholar] [CrossRef] [PubMed]
  321. Clevenger, J.; Chu, Y.; Chavarro, C.; Agarwal, G.; Bertioli, D.J.; Leal-Bertioli, S.C.M.; Pandey, M.K.; Vaughn, J.; Abernathy, B.; Barkley, N.A.; et al. Genome-wide SNP Genotyping Resolves Signatures of Selection and Tetrasomic Recombination in Peanut. Mol. Plant 2017, 10, 309–322. [Google Scholar] [CrossRef] [PubMed]
  322. Kress, W.J. Plant DNA barcodes: Applications today and in the future. J. Syst. Evol. 2017, 55, 291–307. [Google Scholar] [CrossRef]
  323. Cowan, R.S.; Chase, M.W.; Kress, W.J.; Savolainen, V. 300,000 species to identify: Problems, progress, and prospects in DNA barcoding of land plants. Taxon 2006, 55, 611–616. [Google Scholar] [CrossRef]
  324. Mursyidin, D.H.; Setiawan, A. Assessing diversity and phylogeny of Indonesian breadfruit (Artocarpus spp.) using internal transcribed spacer (ITS) region and leaf morphology. J. Genet. Eng. Biotechnol. 2023, 21, 15. [Google Scholar]
  325. Massicotte, R.; Whitelaw, E.; Angers, B. DNA methylation: A source of random variation in natural populations. Epigenetics 2011, 6, 421–427. [Google Scholar] [CrossRef] [PubMed]
  326. Wang, M.Z.; Li, H.L.; Li, J.M.; Yu, F.H. Correlations between genetic, epigenetic and phenotypic variation of an introduced clonal herb. Heredity 2020, 124, 146–155. [Google Scholar] [CrossRef] [PubMed]
  327. Verhoeven, K.J.; Jansen, J.J.; Van Dijk, P.J.; Biere, A. Stress-induced DNA methylation changes and their heritability in asexual dandelions. New Phytol. 2010, 185, 1108–1118. [Google Scholar] [CrossRef] [PubMed]
  328. Chung, Y.S.; Choi, S.C.; Jun, T.H.; Kim, C. Genotyping-by-sequencing: A promising tool for plant genetics research and breeding. Hortic. Environ. Biotechnol. 2017, 58, 425–431. [Google Scholar] [CrossRef]
Figure 1. The broad classification of DNA markers.
Figure 1. The broad classification of DNA markers.
Ijpb 15 00046 g001
Figure 2. Schematic representation of the utilization of gel-based DNA markers for genetic diversity studies in plants.
Figure 2. Schematic representation of the utilization of gel-based DNA markers for genetic diversity studies in plants.
Ijpb 15 00046 g002
Figure 3. SNPs generated due to single nucleotide mutation (transition mutation).
Figure 3. SNPs generated due to single nucleotide mutation (transition mutation).
Ijpb 15 00046 g003
Figure 4. Schematic representation of the steps involved in GBS (genotype-by-sequencing) technology for plant research.
Figure 4. Schematic representation of the steps involved in GBS (genotype-by-sequencing) technology for plant research.
Ijpb 15 00046 g004
Table 2. Genetic diversity studies in plants using multiple marker systems.
Table 2. Genetic diversity studies in plants using multiple marker systems.
Marker Combination Marker TypesNumber
of Marker Systems
Plant SpeciesReferences
SCoT + ISSRDominantTwoDendrobium crysotoxum (SCoT = nine primers; ISSR = twenty primers; genetic diversity within population: ISSR = 86%; SCoT = 74% and between population: ISSR = 14%, SCoT = 26%)
Diospyros species (ISSR = seven primers; SCoT = ten primers; average PIC: ISSR = 0.30, SCoT = 0.36; average marker index (MI): ISSR = 1.81; SCoT = 1.79)
Cucurbita pepo (seven SCoT primers produced forty-nine polymorphic bands and six ISSR primers generated forty-two bands)
[33,165,266]
CBDP + SCoTDominantTwoBauhinia racemose (out of 25 CBDP primers, 21 produced 97 scorable bands, and for SCoT, 18 out of 36 primers produced 88 scorable bands)
Triticum aestivum, Aegilops cylindrical, and A. crassa (CBDP = 15 primers; SCoT = 15 primers; PIC for SCoT: 0.31–0.39, CBDP: 0.28–0.36; cluster analysis: all samples were grouped based on their genomic constitution)
[267,268]
SCoT + ISSR + RAPDDominantThreeKalanchoe genotype (ScoT, ISSR, and RAPD = 10 primers each; polymorphism percentage: SCoT = 57%; ISSR = 15%, RAPD = 60.25%)
Lathyrus species (SCoT = eight primers; ISSR = eight primers; RAPD = six primers; polymorphism: SCoT = 96%; ISSR = 96.81%; RAPD = 94.2%)
[269,270]
SSR + AFLPCo-dominant and dominantTwoJatropha curcas (seven AFLP primer combinations produced seventy amplified polymorphic loci; thirty SSR primers were used, out of which seventeen were amplified in an appropriate size range)
Pyrus pyrifolia (SSR; AFLP = 10 primers each; average PIC for SSR = 0.7585; polymorphism percentage for AFLP = 86.46%; genetic diversity: rich and highly representative)
[271,272]
ISSR + DAMD DominantTwoRosa species (ISSR = ten primers; DAMD = eight primers; genetic variation within population = 86%, between populations = 14%)[273]
DArT + SNPDominant and
co-dominant
TwoManihot esculenta (DArT = 10,521 markers; SNP = 10,808 markers; average PIC for DArT = 0.36; SNP = 0.28)
Glycine max (DArT = 16,116 markers; SNP = 19,505 markers; genetic variance: DArT = 98%; SNP = 97%)
[274,275]
DArT + SNP + SSRDominant and co-ominantThreeLolium perenne (DArT = 1384 markers; SNP = 182 markers; SSR = 48 markers; Genetic diversity: DArT = 0.26; SNP = 0.32; SSR = 0.45)[263]
SCAR + RAPD Co-dominant and dominantTwoNicotiana tabacum (two out of eight SCAR markers; seven out of two hundred RAPD markers efficiently discriminated a large number of Tobacco cultivars)[276]
CAPS + SSRCo-dominantTwoOryza sativa (a set of twenty-eight genome-wide SSR markers; eleven salt-responsive genic SSR markers; eight salt QT-linked SSR markers; CAPS markers: OsHKT1; 5v395)
Citrus sinensis (a total of five markers; average genetic polymorphism = 98.46%; CAPs-SSR indicated more genetic variability)
[264,265]
CAPS + SSR + SNPCo-dominantThreeCitrullus lanatus (CAPS = fifteen markers; SSR = six markers; SNP = two markers; mapping confirmation of BSA-seq: yellow skin)[277]
SSR + ISSRCo-dominant and dominantTwoA total of 28 accessions of Curcubita pepo were compared utilizing ISSR markers, detecting 90 polymorphic bands. Additionally, SSR markers were proposed to further elucidate infra-specific relationships within C. pepo.
Gossypium herbaceum (SSR = thirteen markers; ISSR = five markers; average coefficient similarity = 0.32; low correlation and high variation)
[278,279]
SCoT + DAMDDominant and co-dominantTwoMosses (the inaugural genetic diversity study of three moss species incorporated the utilization of SCoT and DAMD markers to enhance the discriminatory power and precision within the species)[280]
CDDP+ ISSR DominantTwoQuercus infectoria (ISSR = twelve primers; CDDP = nine primers; population variance within: ISSR = 92.97%; CDDP = 94.17%)[281]
ISSR + SRAPDominantTwoMusa species (ISSR = eight primers; SRAP = seven primers; polymorphic bands: ISSR = 81.6%; SRAP = 87.7%)[282]
STS + CAPSCo-dominantTwoCamelia sinensis (STS = two primers; CAPS = thirty-seven primers; high genetic diversity between the two varieties: C. sinensis var. sinensis and C. sinensis var. assamica)[283]
SCoT + IRAPDominantThreeBletilla striata (SCoT = twenty primers; IRAP = eight primers; polymorphic bands: SCoT = 96.17%; IRAP = 94%)[284]
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

Bidyananda, N.; Jamir, I.; Nowakowska, K.; Varte, V.; Vendrame, W.A.; Devi, R.S.; Nongdam, P. Plant Genetic Diversity Studies: Insights from DNA Marker Analyses. Int. J. Plant Biol. 2024, 15, 607-640. https://doi.org/10.3390/ijpb15030046

AMA Style

Bidyananda N, Jamir I, Nowakowska K, Varte V, Vendrame WA, Devi RS, Nongdam P. Plant Genetic Diversity Studies: Insights from DNA Marker Analyses. International Journal of Plant Biology. 2024; 15(3):607-640. https://doi.org/10.3390/ijpb15030046

Chicago/Turabian Style

Bidyananda, Nongthombam, Imlitoshi Jamir, Karolina Nowakowska, Vanlalrinchhani Varte, Wagner A. Vendrame, Rajkumari Sanayaima Devi, and Potshangbam Nongdam. 2024. "Plant Genetic Diversity Studies: Insights from DNA Marker Analyses" International Journal of Plant Biology 15, no. 3: 607-640. https://doi.org/10.3390/ijpb15030046

Article Metrics

Back to TopTop