Next Article in Journal
Advanced Research on Biological Properties—A Study on the Activity of the Apis mellifera Antioxidant System and the Crystallographic and Spectroscopic Properties of 7-Diethylamino-4-hydroxycoumarin
Previous Article in Journal
Selective MicroRNA Packaging Reveals Distinct Core Signatures in Human Mesenchymal-Stromal-Cell-Derived Extracellular Vesicles
Previous Article in Special Issue
LRR Receptor-like Protein in Rapeseed Confers Resistance to Sclerotinia sclerotiorum Infection via a Conserved SsNEP2 Peptide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Thriving or Withering? Plant Molecular Cytogenetics in the First Quarter of the 21st Century

by
Elzbieta Wolny
1,*,
Luis A. J. Mur
2,
Nobuko Ohmido
3,
Zujun Yin
4,5,
Kai Wang
6 and
Robert Hasterok
1,*
1
Plant Cytogenetics and Molecular Biology Group, Faculty of Natural Sciences, Institute of Biology, Biotechnology and Environmental Protection, University of Silesia in Katowice, 40-032 Katowice, Poland
2
Department of Life Sciences, Aberystwyth University, Edward Llwyd Building, Aberystwyth SY23 3DA, UK
3
Graduate School of Human Development and Environment, Kobe University, Kobe 657-8501, Japan
4
State Key Laboratory of Cotton Bio-Breeding and Integrated Utilization, Institute of Cotton Research, Chinese Academy of Agricultural Sciences, Anyang 455000, China
5
Western Agricultural Research Center, Chinese Academy of Agricultural Sciences, Changji 831100, China
6
School of Life Sciences, Nantong University, Nantong 226019, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(14), 7013; https://doi.org/10.3390/ijms26147013
Submission received: 7 June 2025 / Revised: 15 July 2025 / Accepted: 16 July 2025 / Published: 21 July 2025
(This article belongs to the Collection Feature Papers in Molecular Plant Sciences)

Abstract

Nearly four decades have passed since fluorescence in situ hybridisation was first applied in plants to support molecular cytogenetic analyses across a wide range of species. Subsequent advances in DNA sequencing, bioinformatic analysis, and microscopy, together with the immunolocalisation of various nuclear components, have provided unprecedented insights into the cytomolecular organisation of the nuclear genome in both model and non-model plants, with crop species being perhaps the most significant. The ready availability of sequenced genomes is now facilitating the application of state-of-the-art cytomolecular techniques across diverse plant species. However, these same advances in genomics also pose a challenge to the future of plant molecular cytogenetics, as DNA sequence analysis is increasingly perceived as offering comparable insights into genome organisation. This perception persists despite the continued relevance of FISH-based approaches for the physical anchoring of genome assemblies to chromosomes. Furthermore, cytogenetic approaches cannot currently rival purely genomic methods in terms of throughput, standardisation, and automation. This review highlights the latest key topics in plant cytomolecular research, with particular emphasis on chromosome identification and karyotype evolution, chromatin and interphase nuclear organisation, chromosome structure, hybridisation and polyploidy, and cytogenetics-assisted crop improvement. In doing so, it underscores the distinctive contributions that cytogenetic techniques continue to offer in genomic research. Additionally, we critically assess future directions and emerging opportunities in the field, including those related to CRISPR/Cas-based live-cell imaging and chromosome engineering, as well as AI-assisted image analysis and karyotyping.

1. Introduction

Modern molecular cytogenetics integrates genetic, genomic, molecular, and cell biology approaches with advanced microscopic imaging techniques. Its flagship technique, DNA–DNA in situ hybridisation (ISH), enables the precise mapping of target sequences across a range of cytogenetic substrates, including mitotic and meiotic chromosomes at various stages of division, interphase nuclei, and extended chromatin fibres. Each of these substrates offers distinct advantages in terms of mapping resolution, individual chromosome identification, and suitability for addressing specific research questions [1,2]. ISH was first developed in the late 1960s for Xenopus laevis [3] (Figure 1) and was later applied to plants in the mid-1980s. However, its full potential was initially constrained by the technical limitations of non-fluorescent detection systems, such as autoradiographic [4] (Figure 1) and enzymatic [5]. In 1989, Maluszynska and Schweizer demonstrated the chromosomal localisation of ribosomal RNA (rRNA) genes on B chromosomes of Crepis capillaris, using biotinylated 18S and 25S ribosomal DNA (rDNA) probes, detected through the biotin–avidin–fluorescein isothiocyanate (FITC) system [6]. This work represents one of the earliest published applications of DNA–DNA fluorescence in situ hybridisation (FISH) in plants (Figure 1). The first attempts at in situ localisation of single- or low-copy chromosome-specific sequences followed relatively soon thereafter [7,8,9]. Nevertheless, early applications of FISH in plants primarily focused on mapping repetitive DNA sequences, including evolutionarily conserved elements (e.g., rDNA and T3AG3 telomeric motifs), as well as species- and genus-specific sequences such as total genomic DNA (gDNA), selected tandem repeats, and transgenes. These efforts initially used single-colour (e.g., [10,11,12]), then dual-colour (e.g., [13,14,15]), and occasionally multi-colour (e.g., [16,17]) visualisation. The breakthrough came in 2000 with the sequencing of the first plant nuclear genome—of the model angiosperm Arabidopsis thaliana (arabidopsis) [18]. Building on this, Lysak et al. employed collections of ordered bacterial artificial chromosome (BAC) clones containing large inserts of arabidopsis gDNA to specifically paint chromosomes—first in this species [19] (Figure 1), and later in its relatives [20]. This approach significantly advanced plant genome analysis by enabling, for example, comparative studies of karyotype structure and evolution, albeit at individual chromosome segment resolution [21]. It also allowed the spatial positioning of chromosomes within the interphase nucleus to be determined [22]. However, effective chromosome painting (CP) using this method was limited to a handful of small-genome plant species, such as tomato, potato [23], and representatives of the model grass genus Brachypodium [24,25] (Figure 1). This problem reflected the limited availability of sequenced plant genomes at the time, the technical complexity of BAC library construction and handling [26], and the difficulty in obtaining contiguous, chromosome-specific signals in plants with large and/or repeat-rich genomes. Many of these limitations were overcome with the advent and widespread application of probes based on single- or low-copy oligonucleotides in the so-called oligo-FISH approach, a technique originally developed for animal genomes (e.g., [27,28,29]) and later adapted for use in plants [30] (Figure 1). This approach relied on the development of bioinformatic tools for genome mining to identify suitable probe targets (for recent reviews, see [31,32]), advances in the commercial large-scale synthesis of customised oligonucleotides, and, most importantly, the rapid increase in the number of sequenced plant genomes [33,34]. These advances not only paved the way for CP in species that were previously considered challenging or even intractable (e.g., [35,36,37]), but also facilitated the relatively straightforward generation of FISH probes from virtually any sequence of interest.
FISH analyses can be supplemented by, or combined with, immunofluorescent approaches targeting chromatin modifications (for a comprehensive review, see [38]) or other nuclear components, thereby offering deeper insights into the developmental [39,40] and functional dynamics of the nucleus [41,42,43], as well as its epigenetic responses to experimentally induced stresses [44,45,46,47]. These techniques, supported by state-of-the-art microscopy platforms (for reviews, see, e.g., [48,49,50]), along with advanced image processing and analysis platforms (e.g., [51,52]), have yielded unprecedented insights into cytomolecular organisation of nuclear genomes across a wide spectrum of model, non-model, and crop plants. Notably, in a broader genomic context, FISH-based approaches using chromosome- or genome-specific probes have played a vital role in physically anchoring genome sequence assemblies to chromosomes, as demonstrated in foundational work [18] and subsequent earlier studies (e.g., [53,54]), and continue to do so in the most recent work (e.g., [55]).
Meanwhile, rapid advances in sequencing technologies and bioinformatics have made large-scale genomic analyses faster and more affordable than ever before. According to published data [33,34], by the end of 2024, the total number of plant species with sequenced genomes had exceeded 1800, including 370 species sequenced and published for the first time that year. This milestone demonstrated the feasibility of the Ten Thousand Plant Genome Project’s objectives, which had been set only a few years earlier [56]. While these developments create opportunities to apply advanced cytomolecular approaches to a wide range of plant species, they also raise critical questions about the future of plant molecular cytogenetics, a field where techniques are often inherently low-throughput, difficult to standardise, labour-intensive, time-consuming, and largely resistant to automation (e.g., [2,57]). In this review, we summarise recent developments (2020–2025) in major areas of plant molecular cytogenetics, with particular focus on various aspects of chromosome identification and karyotype evolution, chromatin and interphase nuclear organisation, chromosome structure, hybridisation and polyploidy, and cytogenetics-assisted crop development (Figure 2). We also explore future prospects, including CRISPR (clustered regularly interspaced short palindromic repeats)/Cas (CRISPR-associated protein)-based live cell imaging and CRISPR/Cas chromosome engineering, together with their applications and limitations in plant cytomolecular research, as well as artificial intelligence (AI)-assisted image analysis and karyotyping.

2. Chromosome Identification and Karyotype Evolution

The identification of individual chromosomes and their segments within a given species has been, and continues to be, central to most cytomolecular analyses, regardless of their specific objectives. Unsurprisingly, this focus is consistent across a wide range of species, as evidenced in many recent publications (Table 1). To achieve this, robust chromosome markers are essential. As comprehensively reviewed for Brachypodium [70] and Saccharum [71], various FISH probes can serve this purpose. However, their effectiveness depends not only on the sequences employed and the availability of well-developed genomic resources, but also on the specific organisation of the target genome, including its size and the proportion of single- or low-copy sequences relative to repetitive DNA content. The most advanced studies, particularly those involving comparative analyses of the structure and evolution of individual chromosomes and entire karyotypes within related species, require markers capable of specifically painting entire chromosomes or their substantial segments, or, at the very least, providing a chromosome-unique barcode. In recent years, such chromosomal landmarks have become the gold standard in plant karyotyping (Table 1; section I of Figure 2). This progress can be partly attributed to the use of long-established BAC libraries, which have enabled, for example, the reconstruction of karyotype organisation within previously understudied groups of the Brassicaceae [59,72] (Figure 2B), annual [73] and perennial [74]. Brachypodium polyploids, as well as various representatives of the Fabaceae [75,76,77] and Passiflora [78].
Most current studies routinely rely on the application of oligo-FISH probe libraries, which either complement existing BAC-based resources (e.g., [58,79,80,81,82]) or have been developed for a range of both monocot (e.g., [83,84,85,86,87,88,89,90,91]) and eudicot (e.g., [92,93,94,95,96,97,98,99]) species, for which effective CP or comparative chromosome painting (CCP) would otherwise be limited or unavailable. Some studies primarily focus on chromosome identification within individual species, such as Tripidium arundinaceum [100] and Gossypium hirsutum [96]. However, currently available marker systems enable precise tracking of the structure and evolution of individual chromosomes and entire karyotypes across various taxonomic levels. These range from intraspecific analyses within Cicer arietinum [58] (Figure 2A), to intrageneric comparisons in Brachypodium [73,74], Phaseolus [76,101], Vigna [81,82], Lupinus [80], Passiflora [78], Avena [86], Rhynchospora [87], Musa [60,89] (Figure 2C), Citrus [92], Fragaria [93], Ipomoea [94,99], Cucumis [97,98], Glycyrrhiza [95], and Silene [102]. They also extend to complex intergeneric comparisons involving evolutionarily more or less distant groups, such as the Saccharum complex [90], and the Thlaspideae [72], Phaseoleae [103], and Triticeae [83,84,88] tribes. These comparative analyses have greatly advanced our understanding of nuclear genome evolution at the chromosomal level across diverse plant groups. For example, they have uncovered extensive, and in some cases complete, chromosomal synteny across a range of divergence times, such as between Tripidium arundinaceum and Zea mays (maize; ~18 million years; MYs), Tripidium arundinaceum and sorghum (~9 MYs) [100], within the genera Citrus (~9 MYs) [92] and Glycyrrhiza (~3–12 MYs) [95], and between Brachypodium hybridum and its putative diploid progenitors, B. distachyon (Brachypodium) and B. stacei (~0.14–1.4 MYs) [73]. On the other hand, numerous studies across various genera highlight the prevalence of species- and genus-specific chromosome rearrangements (CRs), such as duplications, inversions, and translocations, which contribute to karyotype diversification and speciation [24,58,72,75,76,79,80,87,88,90,97,98,102,104]. Particularly noteworthy are recent analyses employing CP in large-genome grasses [83,88], including allopolyploids such as Campeiostachys (Elymus) nutans [84], which has a genome size exceeding 10 Gb [55]. Thus, the evolutionary histories of entire genomes and individual chromosomes in species that were previously highly challenging, or even intractable, to study can now be elucidated.
Given the current availability and versatility of oligo-based probes, it is not surprising that the use of other types of FISH probes, long employed in chromosome identification and karyotyping, is gradually declining. For instance, 5S and 35S rDNA-targeting sequences, which are among the methodologically simplest, are now typically used only as auxiliary probes. ([58,72,75,76,77,78,79,81,85,90,91,92,93,94,95,96,100,101,105,106]; see Table 1 for details). Although such sequences may still be useful for preliminary comparative studies and in supporting molecular phylogenetic analyses in plant species where more specific chromosome landmarks are not yet available, as illustrated in the genus Onobrychis [107]. Other simple probes, for example, those based on gDNA, which is essential for genomic in situ hybridisation (GISH), are effectively obsolete as an approach (e.g., [72,85]). However, gDNA can be useful in the study of allopolyploids and in tracking individual chromosomes or chromosome fragments within alien genomic backgrounds. These applications are particularly important for the rapid integration of cytogenetic approaches into plant breeding, as discussed further in this review.
Table 1. A selection of recent research articles on chromosome identification and karyotype evolution.
Table 1. A selection of recent research articles on chromosome identification and karyotype evolution.
Research ObjectResearch ApproachAims and Main FindingsReferences
Thlaspideae (Brassicaceae):
Alliaria petiolata (2x, 6x), Didymophysa fenestrata, Graellsia saxifragifolia, G. stylosa, Parlatoria cakiloidea, Peltaria turkmena, Peltariopsis grossheimii,
P. planisiliqua, Pseudocamelina glaucophylla,
P. szowitsii, Thlaspi arvensa
Multiple FISH approaches: CCP with
Arabidopsis thaliana BAC contigs; GISH with gDNA of Pa. cakiloidea and A. petiolata; oligo-FISH with probes localising satDNAs, BAC-FISH with clones targeting 5S and 35S rDNA
Reconstruction of chromosomal organisation in the ancestral genome and analysis of karyotype structure and evolution in 12 Thlaspideae representatives; detection of genus- and species-specific CRs (e.g., pericentric inversions); evidence for allohexaploid origin of A. petiolata (6x) from diploid A. petiolata and Pa. cakiloidea[72]
Catolobus pendulusCCP with Arabidopsis thaliana BAC contigsKaryotype organisation of an understudied species; the hypotetraploid C. pendulus genome originated from
a whole-genome duplication in a genome resembling the ACK, followed by chromosomal rediploidisation
[59]
B. distachyon, B. stacei,
B. hybridum
BAC-FISH with chromosome-specific
B. distachyon clones
Reconstruction of chromosome evolution between genome D and S in annual Brachypodium diploids and their derived allopolyploid; complete chromosomal synteny observed between B. hybridum and its
progenitors
[73]
Brachypodium boissieri, B. mexicanum,
B. phoenicoides, B. retusum, B. rupestre
BAC-FISH with chromosome-specific
B. distachyon clones
Chromosome identification in all species; reconstruction of chromosome evolution among the A1.1, A1.2, A2, E1, E2, and G genome in perennial Brachypodium polyploids; demonstration of ‘orphan’ genomes in the model grass genus[74]
Phaseolus vulgaris, Vigna aconitifolia,
V. unguiculata
BAC-FISH with chromosome-specific
P. vulgaris and V. unguiculata clones; FISH with 5S and 35S rDNA-targeting probes
Intergeneric analyses of karyotype organisation; detection of macrosynteny breaks between Vigna and Phaseolus; CRs (duplications, inversions, and translocations) contribute to karyotype divergence in Vigna[75]
Phaseolus leptostachyus, P. macvaughiiBAC-FISH with chromosome-specific Phaseolus vulgaris clones; FISH with 5S and 35S rDNA-targeting probesInterspecific analyses of karyotype organisation; CRs observed in P. leptostachyus are not shared with
P. macvaughii; only one nested chromosome fusion (chromosomes 10 and 11), is common to both; pericentric inversions detected in chromosomes 3 and 4 exclusively in P. macvaughii
[76]
Macroptilium atropurpureum, M. bracteatum,
M. erythroloma, M. gracile, M. lathyroides,
M. martii
BAC-FISH with chromosome-specific Phaseolus vulgaris clones; FISH with 5S and 35S rDNA-targeting probesInterspecific analyses of karyotype organisation; BAC markers show synteny on orthologous chromosomes, while karyotype differentiation primarily driven by the number and distribution of rDNA loci[77]
Passiflora alata, P. watsonianaBAC-FISH with Passiflora edulis clones; Ty1-Copia and Ty3-Gypsy elements from P. edulis; FISH with 5S and 35S rDNA-targeting probesInterspecific analyses of karyotype organisation; despite karyotype variability, no synteny breaks were observed in the chromosomal distribution of BACs and rDNA sites, except for an additional 35S rDNA locus on chromosome 3 of P. watsoniana; LTRs were uniformly dispersed, with occasional slight accumulation in proximal chromosome regions[78]
Phaseolus vulgaris, Vigna angularis,
V. unguiculata
BAC-FISH with V. unguiculata clones; oligo-FISH with CCP probes designed from the P. vulgaris genome sequence;
FISH with a probe targeting 35S rDNA
Intergeneric analyses of karyotype organisation; first
oligo-FISH CP in legumes; combination of BAC- and
oligo-FISH resources, establishing a cytogenetic map of
V. angularis; detection of CRs (translocations, inversions); chromosomes 2 and 3 identified as hotspots for CRs and de novo centromere formation
[79]
Phaseolus vulgaris, Vigna unguiculataOligo-FISH with barcode probes designed from the Vigna unguiculata genome sequenceIntergeneric analyses of karyotype organisation; in silico integration of previously established BAC-based chromosome-specific and rDNA markers with a newly developed oligo-FISH-based chromosome identification system; alignment of cytogenetic data with genome sequence data for representatives of two distinct genera; confirmation of known and detection of novel CRs (translocations, peri- and paracentric inversions)[104]
Phaseoleae (Fabaceae):
Phaseolus vulgaris, Vigna unguiculata,
Lablab purpureus, Macroptilium atropurpureum
Oligo-FISH with painting probes specific to Pv2 and Pv3 chromosomes of P. vulgarisIntergeneric analyses of karyotype organisation; inference of basic chromosome number and number of genomic blocks in the APK; the translocation between APK2 and APK3 is exclusive to Phaseolus, as chromosomes 2 and 3 of L. purpureus and M. atropurpureum resemble the orthologous chromosomes of V. unguiculata and are more closely related to the APK[103]
Phaseolus acutifolius, P. coccineus, P. dumosus,
P. filiformis, P. leptostachyus, P. lunatus,
P. macvaughii, P. microcarpus, P. vulgaris,
P. vulgaris
CCP with oligo probes designed from the Vigna unguiculata genome sequence and oligo probes specific for chromosomes 2 and 3 of P. vulgaris; FISH with 5S and 35S rDNA-targeting probes and the CentPv1 repeat probeKaryotype evolution in the genus Phaseolus; detection of chromosomal rearrangements (translocations, inversions, duplications, deletions), primarily in species of the Leptostachyus group; P. leptostachyus experienced rapid genome reshuffling without whole-genome duplication, resulting in a reduction of its chromosome number from 11 to 10 pairs[101]
Vigna
subgenera: Vigna; Plectrotropis, Haydonia,
Lasiospron, Ceratotropis
CCP with oligo probes specific for chromosomes Pv2 and Pv3 of Phaseolus vulgaris; barcode oligo probes designed from the Vigna unguiculata genome sequence; BAC clones derived from P. vulgaris and V. unguiculata; FISH with 5S and 35S rDNA-targeting probesKaryotype evolution in the genus Vigna; chromosome identification across eight taxa; macrosynteny observed for chromosomes 2, 3, 4, 6, 7, 8, 9 and 10 in all taxa except V. vexillata, which possesses the most divergent karyotype; only minor differences in painting patterns observed among the subgenera[81]
Vigna lasicarpa, V. unguiculataCCP with oligo probes specific for chromosomes Pv1, Pv2, Pv3, and Pv5 of Phaseolus vulgaris; barcode oligo probes designed from the Vigna unguiculata reference genome; BAC clones derived from V. unguiculata and P. vulgaris; FISH with 5S and 35S rDNA-targeting probes and the T3AG3 telomeric repeatKaryotype evolution in the genus Vigna; demonstration of conserved oligo-FISH patterns on chromosomes 2, 6, 8, 10 and 11 between V. unguiculata and V. lasicarpa; paracentric inversions in Vla3 and Vla9; descending dysploidy in V. lasicarpa driven by end-to-end fusion of homoeologous chromosomes 5 and 7[82]
Lupinus angustifolius, L. cryptanthus,
L. micranthus, L. cosentinii, L. pilosus
BAC-FISH with L. angustifolius clones;
oligo-FISH with probes specific for
chromosome Lang06 of L. angustifolius
Karyotype evolution among five wild Lupinus species; demonstration of putative CRs within the Lang06 region, altering synteny and associated with speciation[80]
Cicer arietinumOligo-FISH with chromosome-specific probes designed from the C. arietinum kabuli morphotype genome sequence; BAC-FISH with single-copy clones from the desi morphotype of C. arietinum; FISH with 5S and 35S rDNA-targeting probes and the T3AG3 telomeric repeatComparative analysis of closely related chickpea genotypes; individual chromosome identification and karyotype development; identification of CRs contributing to genome diversification among chickpea cultivars[58]
Tripidium arundinaceumOligo-FISH with maize-derived painting probes; FISH with 5S and 35S rDNA-targeting probesChromosome identification and karyotyping in T. arundinaceum (sugarcane relative), effective despite 18 MY divergence from maize; conserved synteny with sorghum over 9 MYs[100]
Saccharum complex (Erianthus fulvus,
E. rockii, Miscanthus sinensis, Narenga porphyrocoma, Saccharum officinarum, S. spontaneum, S. robustum)
Oligo-FISH with painting probes designed from the S. officinarum genome sequence; FISH with 5S and 35S rDNA-targeting probesDevelopment of a set of 10 chromosome-specific landmarks effective for comparative karyotype analysis within the Saccharum complex; detection of CRs and novel cytotypes; chromosome fusions are common in various polyploids of the complex and alter the basic chromosome numbers[90]
Aegilops markgrafii, Ae. tauschii,
Ae. umbellulata, Ae. uniaristata, Ae. speltoides,
Triticum aestivum
Oligo-FISH with D genome chromosome-specific painting probes designed from the Triticum aestivum genome sequenceIdentification of the D genome in diploid (Aegilops) and polyploid (T. aestivum) species; detection of translocations involving D chromosomes, including two novel translocations (3D–7D and 4D–5D–7D) in three Ae. tauschii accessions; determination of the precise positions of chromosomal breakpoints in Ae. tauschii accessions; painting probes produce signals in four different genomes (U, C, M, N); CRs were identified in Ae. umbellulata, Ae. markgrafii, and Ae. uniaristata[88]
Triticeae (Poaceae)
genera: Aegilops, Campeiostachys, Dasypyrum,
Elymus, Hordeum, Roegneria, Thinopyrum
Oligo-FISH with painting probes specific for chromosomes 1St to 7St, based on the Pseudoroegneria libanotica and Triticum
aestivum reference genomes
Analysis of Triticeae karyotype organisation; identification of individual chromosomes of the St genome; conservation of St chromosomes in St-containing Triticeae representatives; weak St hybridisation signals observed in Y-genome chromosomes suggest an origin from the St genome[84]
Thinopyrum elongatum, Th. bessarabicum
wheat–tetraploid Th. elongatum substitution lines, Triticum durum–Th. elongatum amphidiploid
Oligo-FISH with painting probes designed from the Th. elongatum and Triticum aestivum genome sequences; tandem-repeat oligo probes pSc119.2 and pTa535Development of a complete set of the E-genome painting probes to facilitate the detection of alien material in wheat breeding; chromosome identification; Th. bessarabicum (2x) shows a close genetic relationship with diploid Th. elongatum; five of the seven E-genome chromosomes exhibit complete synteny in both diploids, except for a reciprocal translocation between 4E and 5Eb; a reciprocal translocation between 5E and 7E is present in one of the diploid Th. elongatum accessions[83]
Avena eriantha, A. fatua, A. nuda, A. sativa,
A. ventricosa, A. wiestii
Oligo-FISH with painting probes based on syntenic regions between wheat and barley; tandem-repeat oligo probesComparative karyotyping of eleven hexaploid and diploid Avena accessions; a high-resolution standard karyotype of A. sativa was established based on distinct FISH signals from multiple oligo probes[86]
Elymus dahuricus, Hordeum vulgareOligo-FISH with painting probes based on syntenic regions between wheat and barley; oligo-FISH with repeat-based probes: pSc119.2, pTa535, Po5, 7E-716, 7E-599, 5S rDNA, 18S rDNA, 3A1, 13-J1011, 7E-744, d01-135, Ae334, and (GAA)7; GISH using Pseudoroegneria spicata gDNA; CENH3
immunolocalisation
Establishment of a universal karyotyping nomenclature system for E. dahuricus; precise determination of the linkage groups and sub-genomes of individual chromosomes; detection of a novel intergenomic rearrangement between the 2H and 5Y chromosomes in this allopolyploid[85]
Rhynchospora (Cyperaceae):
representative species from sections: Albae, Dichromena, Cephalotae, Pauciflorae, Polycephalae, Pseudocapitatae, Tenues
Oligo-FISH with two barcode probes
(Rbv-I and Rbv-II) designed from the
R. breviuscula genome sequence
Investigation of karyotype evolution and chromosomal variations in highly dynamic holocentric karyotypes; identification of all chromosomes in R. breviuscula; probes mapped in 13 other Rhynchospora representatives reveal CRs, including fusions, fissions, inversions, and translocations, as well as whole-genome duplication in R. pubera[87]
Musa acuminata, M. balbisiana: wild subspecies, cultivars, and hybridsOligo-FISH with chromosome-specific probes designed using
the M. acuminata genome sequence
Comparative karyotype analysis; detection of numerous accession-specific chromosome translocations correlated with banana speciation; demonstration of the complexity of banana genome evolution; identification of putative progenitors of banana cultivars[60,89]
Elaeis guineensis, E. oleifera, Cocos nucifera, Phoenix dactyliferaOligo-FISH with chromosome-specific probes designed using the E. guineensis genome sequence; FISH with a 5S rDNA-targeting probeEstablishment of a reference karyotype for E. guineensis and E. oleifera, identification of homoeologous regions in related species[91]
Citrus maxima, C. medica, C. mangshanensis,
C. reticulata, Microcitrus australasica,
Poncirus trifoliata
Oligo-FISH with painting probes designed using the Citrus maxima genome; FISH with the 180 bp satellite repeat and with probes targeting 5S and 35S rDNAIdentification of all chromosomes in the studied species; complete chromosomal synteny was observed among six Citrus species over approximately 9 MYs of divergence, with no interchromosomal rearrangements identified in any species[92]
FragariaOligo-FISH with painting probes designed using the Fragaria vesca genome; FISH with 5S and 45S rDNA-targeting probesIdentification of individual chromosomes in 11 Fragaria representatives with different ploidy levels; comparative karyotyping; Fragaria species exhibit conserved karyotypes, with no interchromosomal rearrangements observed; differences in rDNA loci organisation patterns were found among polyploids; variations in signal intensity of oligo probes among homologous chromosomes in Fragaria polyploids provides new insights into their origins[93]
IpomoeaOligo-FISH with barcode probes designed using I. nil; FISH with 5S and 35S rDNA-targeting probesComparative chromosome analysis in I. batatas and its wild relatives with different ploidy levels; I. trifida is the most closely related diploid to I. batatas; providing
cytogenetic evidence for the segmental allopolyploid hypothesis of sweet potato origin
[94]
Sixteen diploid Ipomoea species representing all seven minor cladesOligo-FISH with painting probes designed using I. nil chromosomes 7 and 15; FISH with 5S and 35S rDNA-targeting probesComparative chromosome analysis across the genus; significant cytogenetic divergence between 2n = 28 and 2n = 30 species, questioning molecular phylogeny-based classifications that group them into the same clade; significant interspecific variation in rDNA loci distribution complements CCP-based analyses[99]
Cucumis sativus var. sativus, C. sativus var. hardwickii, C. hystrix, C. melo, C. metuliferus, C. subsericeus, C. dipsaceus, C. zeyheri,
C. anguria
Oligo-FISH with painting probes designed using the C. sativus genome sequenceDesigning chromosome painting oligo probe libraries; reconstruction of the ancestral karyotype for the genus; comparative analysis reveals the genome structure of all studied species and complex CRs that occurred during Cucumis karyotype evolution; compared to African species, Asian-origin species possess genomes that are highly reshuffled due to large-scale inversions, centromere repositioning, and chromothripsis-like events[97,98]
Glycyrrhiza eglandulosa, G. glanndulosa,
G. eurycarpa, G. glabra, G. inflata,
G. prostrata, G. uralensis
Oligo-FISH with painting probes designed using the Glycyrrhiza uralensis genome sequence; FISH with 5S and 45S rDNA-targeting probesChromosome identification in G. uralensis and its relatives; exceptionally conserved chromosomal synteny was observed after 3–12 MYs of divergence, with no cytologically visible interchromosomal rearrangements detected by CP[95]
Gossypium hirsutumOligo-FISH with probes designed using the Gossypium hirsutum genome sequence and targeting single and multiple chromosomes; FISH with oligo probes targeting
telomeric sites and 5S and 45S rDNA loci
Developing robust markers for chromosome
identification in previously intractable species
[96]
Pulmonaria officinalis group (P. obscura,
P. officinalis s. str.)
Multicolour FISH with tandem-repeat probes (five newly identified satellite DNAs, 5S and 45S rDNA)Designing a new set of chromosome-specific landmarks; comparative karyotyping; chromosome structure in P. officinalis s. str. is more variable than in P. obscura; confirmation of the hybrid status of 2n = 15 putative hybrids collected from mixed populations of P. obscura and P. officinalis s. str.[105]
Silene latifolia, S. dioica, S. vulgaris,
S. maritima
Comparative oligo-FISH with an
X-chromosome-scaffold-originated probe designed from the S. latifolia genome sequence; Silene STAR-C centromeric and X43.1 repeat DNA probes
Development of a more robust probe for visualising Silene sex chromosomes than any previous markers, and investigation of their evolution; the hybridisation of this probe to the short arms of several autosomes in S. vulgaris and S. maritima suggests that extensive CRs played
a role in the evolution of Silene sex chromosomes
[102]
Twelve wild representatives
of the Hordeum genus
Comparative BAC-FISH with the Hbog_46L9 clone from H. bogdanii, which contains a Panicum-derived chromosomal segment; FISH with a 45S rDNA-targeting probeDemonstration of horizontal gene transfer by identifying and characterising a foreign chromosomal segment from a Panicum-like donor in wild barley species, highlighting its evolution and dynamics within host genomes[106]
Nearly 30 species of the Onobrychis genusFISH with 5S and 35S rDNA-targeting probesDetermination of the number and chromosomal distribution of rDNA loci, along with the identification of selected chromosomes; demonstration of polymorphism in rDNA chromosomal patterns in diploids, contrasted with its absence in polyploids; inference of the ancestral basic chromosome number, rDNA loci counts, and mechanisms such as polyploidisation and descending dysploidy that have shaped chromosome number
evolution in the genus
[107]

3. Chromatin and Interphase Nuclear Organisation

Plant nuclear DNA is arranged into chromatin, which not only carries genetic information but also plays crucial roles in processes such as DNA replication, repair, and transcriptional regulation. The three-dimensional (3D) organisation, epigenetic status, and cis- and trans-interactions of chromatin within the nucleus are essential for its function as well as dynamic responses to developmental signals and environmental cues (4D genomics) [108]. Significant progress has been made in deciphering spatial nuclear architecture in plants, largely due to the development of chromosome conformation capture (3C) techniques, such as Hi-C (high-throughput chromatin conformation capture), ChIA-PET (chromatin interaction analysis by paired-end tag sequencing), and HiChIP (in situ Hi-C followed by chromatin immunoprecipitation) (for recent reviews, see [109,110,111]). Furthermore, a broad array of FISH-based approaches, including 3D-FISH and oligo-FISH with painting probes, combined with super-resolution microscopy, have provided valuable insights into how the spatial arrangement of chromosome territories (CTs) influences gene expression [66,112] (section II of Figure 2). Although the existence of CTs in plants was demonstrated some time ago, first in arabidopsis [19] and subsequently in other species (e.g., [24,113]), the mechanisms underlying their establishment and maintenance remain poorly characterised. Recent discoveries related to these processes are summarised in Table 2 and outlined below. Attachment of chromatin to the nuclear envelope, mediated by lamin-like nuclear matrix constituent proteins such as CRWN1 and CRWN4, appears to be a key factor [63,114] (Figure 2G). In addition to CRWN proteins, condensin complexes also contribute to interphase chromatin architecture. Specifically, the role of arabidopsis CAP-D3 in centromere and telomere positioning has been reported [115]. Studies using cap-d3 mutants demonstrate that arabidopsis forms distinct condensin I and II complexes, and that CAP-D3 facilitates the spatial separation of chromocenters without affecting global DNA or histone methylation patterns [62] (Figure 2F).
The interphase nuclei of arabidopsis show a rosette-like CC-loop organisation with telomeres clustering near the nucleolus, whereas related crucifer species with larger genomes display Rabl-like or dispersed configurations. This suggests that genomic properties, such as genome size and degree of longitudinal chromosome compartmentalisation, rather than phylogenetic position, determine interphase nuclear organisation in crucifer genomes [116]. Similarly, stable chromosome positioning independent of genome size, as well as conserved DNA replication dynamics, have been observed across seven Poaceae species [117]. A Rabl-like chromosome configuration in interphase nuclei, revealed through 3D-FISH, was also detected in Limnanthaceae (Brassicales), a family closely related to arabidopsis [118]. In the monocot Oryza sativa (rice) most somatic cell nuclei lack Rabl configuration [119], contrasting with another monocot model plant, Brachypodium, which exhibits Rabl organisation in certain cell types [120]. A recent study on O. sativa interphase nuclei revealed tissue-specific chromatin architecture, with differences in condensation levels and CT arrangements between leaf and root cells. Chromosome positioning, CT volumes, and spatial associations were shown to depend on nucleolus size and the activity of the 45S rDNA locus, underscoring the connection with nucleolar architecture [61] (Figure 2D,E). The dynamic organisation of chromatin in relation to the cell cycle and developmental cues has also been investigated in Hordeum vulgare (barley). Distinct centromere and telomere arrangements observed between cycling and endoreduplicated nuclei in embryo and endosperm tissues indicate that the Rabl configuration is established and maintained through mitotic divisions, and that it is further influenced by tissue identity and seed developmental stage [121].
Table 2. A selection of recent research articles on molecular cytogenetic analyses of chromatin and interphase nuclear organisation.
Table 2. A selection of recent research articles on molecular cytogenetic analyses of chromatin and interphase nuclear organisation.
Research ObjectResearch ApproachAims and Main FindingsReferences
Arabidopsis thaliana, Hordeum vulgare3D FISH using centromeric, 35S rDNA-targeting, telomeric, subtelomeric, and H5L-specific painting oligo probes; immunodetection of RNA polymerase II, ASY1, ZYP1, DMC1, HEI10, SSSU, and H3K27me3; visualisation of
a GFP-tagged protein associated with the nuclear envelope; imaging performed using diffraction-limited confocal microscopy and super-resolution microscopy
A compendium of strategies to analyse the spatial distribution of nuclear and chromosomal signals from 3D image stacks[66]
Arabidopsis thaliana (Col-0 and ddm1-2)Image analysis using the semi-automatic ImageJ plug-in iCRAQ (https://github.com/gschivre/iCRAQ, accessed on 15 July 2025) and the DL-based tool Nucl.Eye.D (https://zenodo.org/records/7075507, accessed on 15 July 2025)Detection and quantification of A. thaliana nuclear features using two segmentation methods: iCRAQ (semi-automated) and Nucl.Eye.D (deep learning), enabling precise analysis[112]
Populus trichocarpa3D-FISH on frozen root tip sections; FISH with a 45S rDNA-targeting probe and oligopainting probes for chromosomes 17 and 19Improved signal quality compared to paraffin sections; chromosome-specific oligo probes enabled 3D analysis of chromosome territories; autosome pair 17 associated more frequently than sex chromosome 19[113]
Arabidopsis thaliana WT and crwn1-1, crwn4-1,
and kaku4-2 mutants
BAC-FISH using clones specific to A. thaliana chromosomes 1 and 3Plant chromatin organisation is flexible, adapting to developmental and environmental cues; under heat stress, the nuclear lamina disassembles, and chromatin domains relocate from the nuclear envelope to the inner nucleus while remaining associated with CRWN1; CRWN1 plays a key role in genome folding dynamics during stress[63]
Arabidopsis thaliana Columbia-0
and cap-d3 T-DNA insertion mutants
FISH using 180 bp centromeric repeat, 5S and 45S rDNA-targeting probes; immunostaining with antibodies against histone modifications H3K27me3, H3K9me1, H3K9me2, H3K4me3, H3K9ac, H3K14ac, H3K18ac or H3K9 + 14 + 18 + 23 + 27ac, and 5-methylcytosineEvaluation of the role of CAP-D3 in interphase chromatin organisation and function; in cap-d3 mutants, heterochromatic sequences show increased association, while nuclear size and the general histone and DNA methylation patterns remain unchanged[62]
Arabidopsis thaliana, Arabis cypria, Bunias orientalis,
Cardamine amara, Descurainia preauxiana,
Euclidium syriacum, Hesperis sylvestris
2D and 3D FISH using centromere-specific oligo probes: pAL, ArCy1, CARCEN, HeSy1, and de novo identified 156-bp repeat of D. preauxiana; telomeric repeat and A. thaliana BAC clone T15P10 (AF167571) containing 35S rRNA genesThe CC-loop model in Arabidopsis thaliana links telomeres to the nucleolus; in crucifers, small genomes exhibit nucleolus-associated telomere clustering, whereas large genomes display a Rabl-like configuration or a dispersed chromosomal distribution[116]
Avena sativa, Brachypodium distachyon, Hordeum vulgare, Oryza sativa, Secale cereale, Triticum aestivum, Zea maysNuclei sorting by flow cytometry;
5-ethynyl-2′-deoxyuridine labelling; FISH using a telomere oligo probe; centromere immunovisualisation with an antibody against OsCenH3 (rice centromeric histone H3 variant)
Conserved DNA replication dynamics and chromosome positioning across seven Poaceae species with varying genome sizes[117]
Limnanthaceae, Brassicales2D and 3D FISH using centromere-specific oligo probes, telomeric repeat, Arabidopsis thaliana BAC clone T15P10 (AF167571) containing 35S rRNA genes, and clone pCT4.2 (M65137) corresponding to the 5S rDNA repeatFive chromosome pairs in the interphase nuclei of Limnanthes species adopt a Rabl-like configuration[118]
Oryza sativaOligo-FISH using painting probes specific for chromosome 9 and the S and L arms of chromosome 2; FISH with probes targeting centromeric, telomeric, and 45S rDNA sitesSix chromosome territory (CT) configurations were identified in O. sativa root meristematic nuclei and four in leaf nuclei, showing variations in CT volume and association frequency; the association of chromosome 9 CTs was influenced by 45S rDNA activity, linking nuclear organisation to the position and size of the nucleolus[61]
Hordeum vulgareNuclei sorting by flow cytometry; oligo-FISH using a barley centromere-specific probe; FISH with telomeric repeats, 5S and 45S rDNA-targeting probes; immunostaining with antibodies against H. vulgare CENH3Analysis of nuclear morphology and chromosome organisation in cycling and endoreduplicated nuclei isolated from embryo and endosperm tissues of developing barley seeds; endoreduplicated nuclei exhibit irregular shapes, show reduced sister chromatid cohesion at 5S rDNA loci, and decreased CENH3 levels; progressive endoreduplication leads to intermingling centromeres and telomeres[121]

4. Chromosome Structure

The metaphase chromosome represents the structural state in which chromatin reaches its highest level of compaction. Recent studies examining plant chromatin organisation throughout the cell cycle, as well as the architecture of metaphase chromosomes, are presented in Table 3, section III of Figure 2, and briefly outlined below. Two models have been proposed to describe the higher-order organisation of metaphase chromosomes: helical and non-helical. Recent studies combining Hi-C analysis, biopolymer modelling, and structured illumination microscopy have confirmed a helical organisation of barley metaphase chromosomes, consistent with observations in chicken and HeLa cells [122]. Imaging techniques such as transmission electron microscopy [123], scanning electron microscopy [124], and super-resolution microscopy [125] (Figure 2J) have been employed to validate the presence of this helical chromatin structure. Among these techniques, super-resolution microscopy enables the localisation of topoisomerase (Topo) proteins along chromosomes [126]. The localisation and distribution of Topo II have also been investigated using high-voltage transmission electron microscopy and ultra-high-voltage transmission electron microscopy combined with immunogold labelling [127]. Advances in the analysis of plant chromosome structure using various electron microscopy techniques have recently been reviewed by Ohmido et al. [128].
Our understanding of mitotic dynamics has traditionally relied on fixed-sample techniques, which limit insights into the kinetics of chromatin, nucleoli, microtubules, and the duration of individual mitotic stages. However, studies employing fluorescent protein translational fusion lines in barley, combined with confocal microscopy, have yielded valuable insights into nuclear organisation and mitotic dynamics in living root meristematic cells [129]. During eukaryotic cell division, each daughter cell must inherit a balanced chromosome complement, a process critically dependent on properly functioning centromeres. Centromeres assemble kinetochore complexes essential for spindle microtubule attachment. Comprehensive reviews on the structure, function, and evolution of plant centromeres have recently been published [130,131,132]. In plants, kinetochore assembly sites are marked by the presence of the centromeric histone H3 variant (CENH3), which can be localised using specific antibodies. Recent studies have successfully identified and localised this protein in many species, including Agave tequilana relatives [133], Prionium serratum [65] (Figure 2I), Chionographis japonica [134], Cuscuta spp. [135], and Gossypium species [136]. Unlike canonical histones, CENH3 evolves rapidly, and its N-terminal tail exhibits high variability even among closely related species. Due to this variability, antibodies targeting conserved domains of outer kinetochore proteins, such as KNL1 and NDC80, offer greater versatility for centromere immunolabeling across diverse plant taxa [137].
Most plants possess monocentric chromosomes, characterised by centromeres localised to a single chromosomal region. However, numerous species have independently evolved holocentric chromosomes, where spindle microtubules attach along the entire chromosome length [138]. A comprehensive analysis of holocentric chromosome architecture, incorporating oligo-FISH with satellite repeats, immunostaining of CENH3 and kinetochore proteins, and histone modification profiling, has recently been conducted in Chionographis japonica [134] and Luzula sylvatica [139]. Another form of centromere organisation, known as metapolycentricity, has been observed in Pisum [64] (Figure 2H). These chromosomes are cytologically characterised by extended primary constrictions and multiple discrete domains of CENH3 chromatin. A transition from monocentricity to holocentricity has been reported within the genus Cuscuta, where the kinetochore protein gene KNL2 is absent, CENH3 is enriched in heterochromatic regions, and microtubule attachment occurs along the full chromosomal length [135].
Table 3. A selection of recent research articles addressing various aspects of chromosome structure from a cytomolecular perspective.
Table 3. A selection of recent research articles addressing various aspects of chromosome structure from a cytomolecular perspective.
Research ObjectResearch ApproachAims and Main FindingsReferences
Hordeum vulgareImmunostaining with antibodies against Topo IIα and grass CENH3 (centromeric histone H3 variant); structured illumination microscopy (SIM) and photoactivated localisation microscopyTopo IIα is dispersed along chromosome arms but accumulates at centromeres, telomeres, and NORs; at centromeres, Topo IIα intermingles with CENH3-containing chromatin [126]
Hordeum vulgareOligo-FISH with probes specific to the 5HL chromosome; 5-ethynyl-2′-deoxyuridine (EdU) labelling; analysis of purified metaphase chromosomes; biopolymer modelling; spatial SIM of large fluorescently labelled chromosome segments Direct differential visualisation of a condensed chromatin fibre confirms the helical model; revealing chromonemas—helically wound, 400-nm-thick chromatin threads that form the chromatids of mitotic chromosomes [122]
Agave tequilana, Hesperaloe funifera,
H. parviflora, Hesperoyucca whipplei, Yucca carnerosana,
Y. constricta, Y. elata
Immunostaining with antibodies against agavoid CENH3; 3D super-resolution microscopy; scaling relationship of kinetochore size to chromosome size in the karyotypeA positive intra-karyotype relationship between kinetochore and chromosome size, similar to that observed in other eucaryotes; the scaling of total kinetochore size to genome size may originate from the mechanics of cell division[133]
Prionium serratumImmunostaining with antibodies against P. serratum CENH3, α-tubulin, histone H3S28ph, and histone H2A120phP. serratum exhibits a monocentric chromosome organisation, in contrast to the holocentricity observed in other species of the Cyperid clade (Thurniceae-Juncaceae-Cyperaceae)[65]
Chionographis japonicaOligo-FISH with probes for C. japonica satellite repeats (centromeric Chio1 and Chio2), LTR transposable elements, telomeric repeat, and 45S rDNA FISH; immunostaining with antibodies against C. japonica CENH3, MIS12, NDC80, and α-tubulin, as well as histone modifications including H3K4me2, H3K9me2, H3S10ph, H3S28ph, H3T3ph, and H2AT120ph; EdU labellingHolocentric chromatids of C. japonica consist of 7–11 evenly spaced, megabase-sized centromere-specific histone H3-positive units, which contain satellite arrays of 23- and 28-bp-long monomers; the large-scale eu- and heterochromatin arrangement differs between C. japonica and other known holocentric species[134]
Luzula sylvaticaOligo-FISH with probes for satellite repeats Lusy1 and Lusy2; immunostaining with antibodies against L. elegans CENH3, KNL1, NDC80, and α-tubulinL. sylvatica holocentromeres are predominantly associated with two satellite DNA repeats, Lusy1 and Lusy2, while CENH3 also binds to satellite-free gene-poor regions; Lusy1 plays a crucial role in centromere function across most Luzula species; holocentric chromosomes in Luzula may have originated from chromosome fusions of ancestral monocentric chromosomes and the expansion of CENH3-associated satDNA[139]
Pisum sativum and related representatives
of the Fabeae tribe (Pisum, Lathyrus, Vicia)
Oligo-FISH with painting probes PS6, for
P. sativium chromosome 6, satDNA-based FabTR probes; immunostaining with anti-CENH3 antibodies
Assembly and analysis of a 177.6 Mb region of P. sativum chromosome 6 which includes 81.6 Mb centromere region (CEN6) and adjacent segments of both chromosome arms; three satellite repeats were associated with CENH3-enriched chromatin, while five others were not; comparative analysis revealed that the evolution of metapolycentromeres is driven by the expansion of centromeric chromatin into neighbouring chromosomal regions, accompanied by the accumulation of novel satellite repeats, which are complemented by CRs in some species[64]
Cuscuta europaea, C. epithymum, C. australis, C. campestris, C. reflexaImmunostaining with antibodies against Cuscuta CENH3, KNL1 and KLN2, CENP-C, MIS12, NDC80, BUB3;1/2, borealin, and α-tubulin The transition from monocentricity to holocentricity in the genus Cuscuta was accompanied by dramatic changes in the kinetochore, including the loss of centromeric localisation of CENH3, CENP-C, KNL1, MIS12, and NDC80 proteins, as well as and the degeneration of the spindle assembly checkpoint (SAC); these changes indicate that holocentric Cuscuta species have lost the ability to form a standard kinetochore and no longer utilise the SAC to regulate microtubule attachment to chromosomes[135]
Arabidopsis thaliana, Chionographis japonica, Cuscuta reflexa, Dionaea muscipula, Drosera capensis, Juncus effusus, Luzula nivea, Nelumbo nucifera, Nymphaea alba, Ocimum basilicum, Picea abies, Pisum sativum, Raphanus sativus, Rhynchospora pubera, Triticum aestivumImmunostaining with antibodies against KNL1, NDC80, and α-tubulinThe KNL1 and NDC80 antibodies effectively labelled centromeres in condensed chromosomes during cell division, as well as the interphase nuclei of most species tested; KNL1 and NDC80 antibodies are better suited for immunolabeling centromeres than CENH3 antibodies, providing greater versatility across different plant species and enabling the study of centromere organisation in non-model species[137]
Gossypium anomalum, G. arboreum,
G. hirsutum, G. raimondii
Immunostaining with antibodies against cotton CENH3; FISH with probes for centromeric repeats of G. anomalumCharacterisation of G. anomalum centromeric sequences using chromatin immunoprecipitation against CENH3 antibodies; G. anomalum centromeres contained only retrotransposon-like repeats and lacked long arrays of satellite DNA[136]
Petunia axillaris subsp. axillaris, P. axillaris subsp. parodi, P. integrifolia subsp. inflata,
P. × hybrida
FISH with PSAT1, PSAT3, PSAT4, PSAT5, PSAT6, and PSAT7 satellite repeat probesSeven repeat families (PSAT1, PSAT3, PSAT4, PSAT5, PSAT6, PSAT7, PSAT8) exhibited high sequence similarity and organisation across the four Petunia genomes; these repeat families occupy distinct chromosomal niches, differing in copy number and organisation[140]
Rosa arvensis, R. multiflora, R. rugosa,
R. majalis, R. nitida, R. persica
FISH with 18S rDNA and probes derived from the 5S rDNA genic region, as well as 5S_B and 5S_A IGS subregionsLocus-specific probes determined the number and chromosomal position of 5S rDNA families; two major 5S rDNA families (5S_A and 5S_B) were identified in Rosa diploids and polyploids; the 5S_B family often co-localised with 35S rDNA at NORs, while such co-localisation of the 5S_A family was rare[141]
41 woody plants representing 37 species and 27 genera, and 18 families, Zea maysOligo-FISH with the (AG3T3)3 probeThe AG3T3 sequence was observed at chromosome termini in 38 plants; its non-telomeric signals were detected in 23 plants, being particularly abundant in Chimonanthus campanulatus[142]
Hordeum vulgare fluorescent marker linesTime-lapse confocal microscopy imagingDevelopment of unique materials enabling detailed live-cell imaging of mitosis and cytokinesis; determination of the duration of mitosis and its stages in barley; demonstration that chromosome condensation in barley often precedes the mitotic preprophase[129]
Hordeum vulgareElectron tomographyDissecting the 3D higher-order structure of metaphase chromosomes using a thin carbon film in electron tomography revealed periodic structures with a 300–400 nm pitch along the barley chromosome axis; their periodicity was twice that of the corresponding structures found in human chromosomes[123]
Hordeum vulgareImmunogold labelling; immunostaining with barley-specific antibody against Topo II; high-voltage transmission electron microscopy (HVTEM) and ultra-high-voltage transmission electron microscopy (UHVTEM)HVTEM and UHVTEM combined with immunogold labelling are effective for detecting structural proteins such as Topo II; Topo II molecules are distributed along barley chromosomes in a non-specific pattern, with distinct accumulation at the chromosome termini, nucleolus organiser, and centromeric regions[127]
Hordeum vulgareImmunostaining with barley-specific antibodies against Topo IIα applied to flow-sorted chromosomes; sub-diffraction variants of fluorescence super-resolution microscopy, such as structured illumination, stimulated emission depletion, and single-molecule localisation microscopyProtein imaging in barley metaphase chromosomes: comparing selected super-resolution approaches with conventional wide-field and confocal microscopy in terms of mapping resolution and accuracy[125]
Hordeum vulgareScanning electron microscopy (SEM)Investigating the role of calcium ions (Ca2+) in the chromosome structure of barley; BAPTA treatment led to a less condensed, dispersed chromosome structure due to Ca2+ chelation; high-resolution SEM provided detailed visualisation of chromosome ultrastructure under different calcium ion conditions[124]
A human–Arabidopsis thaliana hybrid cell line containing a neo-chromosomeFISH using fifteen probes targeting A. thaliana single-copy genome regions, A. thaliana centromere repeat (Atcen, 180 bp), and telomeric short repeats of human (T2AG3) and A. thaliana (T3AG3)The structure and function of plant and animal chromosomes are largely conserved, enabling the creation of a human–A. thaliana hybrid cell line; a neo-chromosome was formed by inserting segments of A. thaliana chromosomes 2–5 into human chromosome 15; the neo-chromosome contained A. thaliana centromeric repeats and human telomeres; however the A. thaliana centromere was not functional; most A. thaliana DNA was eliminated during culture[143]
Secale cereale, Triticum aestivum, Aegilops speltoidesFISH with 5S rDNA, S. cereale genome-specific repeat Revolver, B-chromosome-specific repeats (D1100, E3900, Sc9c130, Sc26c38), and the DCR28 gene familyBased on a newly assembled ~430 Mb rye B chromosome pseudomolecule, five candidate genes were identified as trans-acting moderators influencing targeted B chromosome nondisjunction during the first pollen mitosis; among them is DCR28, a microtubule-associated gene; the DCR28 gene family appears to be neo-functionalised and is uniquely highly expressed during the first pollen mitosis in rye[144]
Sorghum purpureosericeumFISH on embryo sections, pollen grains, and meiocytes using the B-specific repeat CL135 and the centromeric probe CL29B chromosome occurrence is tissue- and organ-specific, primarily due to extensive elimination during embryo development, which continues throughout plant growth; accumulation of B chromosomes results either from nondisjunction during the first pollen mitosis or from additional nuclear divisions during pollen development[145]
Aegilops speltoidesFISH with repetitive DNA probes Spelt1, Spelt52, pSc119.2, pTa71 (45S rDNA), As5SDNAE (5S rDNA), CCS1 (centromeric), and T3AG3 (telomeric)Ectopic associations between B and A chromosomes were observed, along with cell-specific rearrangements of B chromosomes in both mitosis and microgametogenesis; the copy numbers of selected transposable elements and tandem repeats varied with genotype and tissue type, but were unaffected by the presence or absence of B chromosomes[146]
Supernumerary B chromosomes are often distinguishable from the standard A chromosomes within the karyotype. Their enigmatic role, non-Mendelian behaviour during meiosis, and frequent irregular segregation in postmeiotic mitoses continue to make them compelling subjects of study (reviewed by [147]). For example, Chen et al. [144] employed a newly assembled Secale cereale (rye) B chromosome pseudomolecule to identify five candidate genes acting as trans-acting regulators of B chromosome drive in developing pollen. These included the DCR28 gene family, which encodes a protein associated with cell division and was also identified on Aegilops speltoides B chromosomes. In contrast, studies in Sorghum purpureosericeum demonstrate B chromosome instability during plant ontogenesis, with their elimination occurring primarily during embryo development. This leads to distribution patterns restricted to specific tissues or organs, which are largely preserved post-embryonically [145]. Interestingly, in Ae. speltoides, the presence or absence of B chromosomes did not appear to influence the copy number dynamics of mobile elements and tandem repeats, despite observed ectopic associations between supernumerary and standard chromosomes. This complicates efforts to disentangle the specific roles of B chromosomes from other factors affecting nuclear genome integrity and dynamics [146]. However, recent findings in maize demonstrate that such distinctive roles do exist., For example, they influence the distribution and characteristics of R-loops on A chromosomes in a tissue-specific manner, thereby affecting various cellular processes, including gene expression [148].

5. Natural and Induced Hybridisation and Polyploidy

Polyploidisation, the multiplication of complete chromosome sets, is a widespread phenomenon and plays a crucial role in angiosperm speciation, adaptation and evolution [149,150,151]. Cytomolecular approaches using FISH with various sequences as probes, particularly gDNA, have long proved effective in studying plant polyploids by enabling the visualisation of chromosome sets contributed by their evolutionary parents (e.g., [152,153]). In some cases, they have provided the first, yet robust, indication of the polyploid nature of newly discovered species (e.g., [154]). Recent cytomolecular studies have further yielded valuable insights into taxonomic delineation and genome evolution across diverse taxa (see Table 4 and section IV of Figure 2). For example, they have clarified the complex origins of Hieracium allopolyploids [155], revealed interspecific relationships and non-homologous chromosomal rearrangements in Triticeae species [36], and contributed to tracing the evolution of the Camelina genus by identifying a C. neglecta-like genome (C. intermedia) as a potential ancestor of C. sativa [156]. Furthermore, GISH-based analyses have shed light on hybridisation events between the native Opuntia rioplatensis and the introduced North American species O. ficus-indica, which likely led to the formation of the taxon described as O. cristalensis [157]. The same approach was employed by Shimomai et al. [158] to demonstrate how polyploidisation in Commelina contributes to enhanced survival in urban environments.
In addition to analyses of naturally evolved polyploids, molecular cytogenetic methods have proven invaluable for studying synthetic auto- and allopolyploids. These, apart from their usefulness in breeding, serve as informative model systems for investigating early genome evolution following polyploidisation [159,160]. The formation of synthetic polyploids can trigger various genomic responses, including non-homologous chromosome pairing, translocations, deletions and other structural rearrangements [152,159]. Variation in chromosome number, accompanied by chromosome translocations, has been observed in synthetic Triticum turgidum × Aegilops umbellulata hybrids [161].
Table 4. A selection of recent research articles on cytomolecular analyses related to natural hybridisation and polyploidy.
Table 4. A selection of recent research articles on cytomolecular analyses related to natural hybridisation and polyploidy.
Research ObjectResearch ApproachAims and Main FindingsReferences
Saccharum spontaneumOligo-FISH with haplotypic probes of
S. spontaneum specific to chromosomes 8A, 8B, 8C, and 8D
Whole genome duplications in autopolyploid sugarcane AP85–441; no chromosomal aberrations found between autotetraploid AP85–441 and its spontaneously doubled version indicate strict regulation of chromosome duplication[162]
Hieracium intybaceum, H pallidiflorum.
H. picroides, H. prenanthoides
GISH with gDNA of H. intybaceum and H. prenanthoides; FISH with 5S and 35S rDNA-targeting probesOne of the first multiapproach studies of apomictic Hieracium allopolyploids; multiple origins of hybridogenous H. pallidiflorum and H. picroides from the same diploid–polyploid parental species H. intybaceum and H. prenanthoides; new insight into the taxonomic delineation of the species[155]
OpuntiaFISH with 5S and 35S rDNA-targeting probesO. × cristalensis appears to be a hybrid between the native
Argentine species O. rioplatensis and the North American introduced species O. ficus-indica; the number of 5S rDNA sites in O. × cristalensis reflects its ploidy level, whereas the number of 35S rDNA sites does not; probably the first documented case of hybridisation between North and South American Opuntia species
[157]
TriticeaeFISH with (1) single copy oligos associated with each of the A-, B-, and D-genome chromosomes of Triticum aestivum; (2) oligos from 1H to 7H chromosomes designed using Hordeum vulgare genome; (3) the conserved oligos based on a wheat reference genome; (4) Synt1 to Synt7 oligos from the syntenic region with > 96% homology in wheat-barley linkage groups; (5) Synt7SL barcoding oligosDevelopment of a chromosome-specific painting using oligo pools for large-genome Triticeae species; high-throughput karyotyping of Triticeae and some wheat-alien derivatives; tracking interspecific chromosome homologous relationships and non-homologous CRs[36]
Commelina benghalensis, C. communis (Cc),
C. communis f. ciliata (Ccfc)
GISH with gDNA of CcfcInvestigation of the role of polyploidisation in the distribution and survival of Cc and its subspecies Ccfc across urban-rural gradients; urban areas were dominated by Cc, whereas both Cc and Ccfc coexisted in rural areas; polyploidy and an additional genome provide Cc with enhanced survival in urban environments[158]
Synthetic
Triticum turgidum–Aegilops umbellulata
hybrids
GISH with gDNA of Ae. umbellulata; FISH with following probes: oligo-pTa-535 (pTa535), oligo-pSc119.2 (pSC119.2), oligo-pTa71 (pTa71—35S rDNA-targeting probe), and (AAC)5Investigation of unreduced gamete formation mechanisms in T. turgidum–Ae. umbellulata triploid F1 hybrid crosses and the chromosome compositions in their F2 generations; chromosome numbers in F2 plants ranged from 35 to 43, with variations in chromosome loss/gain among genomes, chromosome loss was highest in the U genome; three types of chromosome translocations and polymorphic FISH karyotypes were identified[161]
Camelina intermedia, C. hispida, C. laxa,
C. neglecta, C. sativa
GISH with gDNA of C. hispida, C. laxa, C. neglecta, and C. intermedia; CCP with Arabidopsis thaliana BAC contigs as painting probesThe identification of the maternal genome of the allohexaploid C. sativa; a tetraploid C. neglecta-like genome (C. intermedia) is hypothesised to be the likely maternal ancestor of the C. sativa based on its high collinearity with two maternally inherited subgenomes; the study contributes to completing the image of the evolution of the Camelina genus[156]
Synthetic and natural allotetraploid wheat hybridsFISH with centromere-specific retrotransposon of wheat and the telomere repeat; multicolour GISH with gDNA of Triticum urartu, Aegilops longissima and Ae. tauschiiA series of nascent allotetraploid wheats from three diploid genomes (A, S*, and D) was synthesised; most progeny had consistent chromosome numbers, with each genome containing 14 chromosomes, suggesting stable chromosome number inheritance due to diploidisation; detected aneuploids have affected centromere pairing and clustering in early meiosis[163]
Autopolyploid (4x, 8x, 10x) clones of Saccharum
spontaneum
Oligo-FISH with painting probes specific for chromosomes 1 (Chr. 1), 7 (Chr. 7), and 8 (Chr. 8) of
S. spontaneum
All clones showed stable, diploid-like chromosome behaviour during meiosis; in the 4x clone, two and two copies of Chr. 8 are of different size, and the pairing likely occurs between the homologs of similar size; considering high sequence similarity among Chr. 8 homologues, some unknown mechanisms are responsible for their peculiar pairing behaviour in the 4x clone[67]
Brachypodium hybridumFISH with 5S and 35S rDNA-targeting probesInvestigation of nucleolar dominance (ND) stability in
B. hybridum genotype 3-7-2 compared to the reference genotype ABR113 revealed differences in tissue-specific expression; in ABR113, ND remained stable across all tissues, including primary and adventitious roots, leaves, and spikes; genotype 3-7-2 exhibited a strong upregulation of S-subgenome units in adventitious roots, but not in other tissues
[164]
Brachypodium hybridum, B. distachyon,
B. stacei
FISH with 5S and 35S rDNA-targeting probesAnalysis of the structure, expression, and epigenetic landscape of 35S rDNA in allopolyploid B. hybridum and its diploid progenitors, B. distachyon and B. stacei; in B. hybridum, the copy number of B. stacei 35S rDNA homoeologues was reduced, accompanied by their transcriptional inactivation; DNA methylation played a role in the silencing of 35S rDNA loci in the S-subgenome[165]
Brachypodium hybridum, B. distachyon,
B. stacei
FISH with 5S and 35S rDNA-targeting probesComparative analysis of repetitive DNA, focusing on rDNA, in two B. hybridum genotypes of significantly different evolutionary ages; in the younger genotype, ABR113, partial elimination of 35S rDNA units was detected; the older genotype, Bhyb26, exhibited a tendency toward diploidisation, with a reduction in the number of both 35S and 5S rDNA loci[166]
Multiple genotypes of Festuca × Lolium hybrids, Festuca × Festuca interspecific hybridsFISH with a 45S rDNA-targeting probe; GISH with gDNA of F. pratensis and F. glaucescensInvestigating ND in Festulolium and fescue hybrids; providing new evidence that this phenomenon is maternity-independent, aligns with genome dominance, and occurs early after hybrid genome merging, being completed in the F2 generation[167]
Two Tragopogon porrifolius lines, por1 and por2, which significantly differ in their 35S rDNA copy numberFISH with 5S and 35S rDNA-targeting probesA positive correlation between the lower 35S rDNA copy number in por1 and the size of NORs on chromosomes D; both L- and S-variants of 35S rDNA were detected in por2,whereas only the S-rDNA variant was found in por1; in por1, the expression of S-rDNA was linked to secondary constrictions (SCs) of NORs located on both chromosomes A; in por2, silencing of S-rDNA was accompanied by NOR condensation on chromosomes A, the presence of SCs on D-NORs, and the expression of L-rDNA, suggesting bidirectional ND[168]
In contrast, autotetraploids of Saccharum spontaneum exhibit strict control over chromosome duplication [162]. Meiotic stability was observed in a series of autopolyploid clones of S. spontaneum, which displayed diploid-like chromosome behaviour [67] (Figure 2K,L). Similarly, stable inheritance patterns resulting from diploidisation were detected in the majority of progeny from both synthetic and natural allotetraploid wheat hybrids [163].
The 35S rDNA sequence has been the focus of numerous studies on natural plant and resynthesised allopolyploids and hybrids, particularly regarding nucleolar dominance (ND), as recently reviewed [169]. In ND, 35S rDNA loci inherited from one progenitor are transcriptionally dominant over those from the other. For example, B. hybridum, a natural allopolyploid and an annual representative of the model grass genus Brachypodium, has long been [170] and continues to be extensively studied regarding ND, including its epigenetic landscape [41], and more recently, its tissue specificity [164], and intraspecific variation [165,166]. The contribution of parental genomes to ND has also been investigated in Festulolium and Festuca hybrids [167]. In Tragopogon porrifolius lines differing in their 35S rDNA composition, a positive correlation has been observed between a lower 35S rDNA copy number and the size of the nucleolus organiser region (NOR) [168]. While uniparental silencing of 35S rDNA in interspecific hybrids and allopolyploids is well-documented, evidence for similar silencing of 5S rRNA genes has long been lacking. This issue was addressed in 2024 when Mandáková et al. [171] reported the first instance of uniparental silencing of 5S rDNA in Cardamine polyploids, opening avenues for further research into the regulatory roles of these genes within complex polyploid genomes.

6. Cytogenetics-Assisted Crop Improvement

The rising demand for plant-based foods and products, set against the backdrop of climate change, presents a significant challenge for breeders, who must rapidly develop crop varieties that are more productive, resilient, and resource-efficient. Crop wild relatives (CWRs) represent a valuable reservoir of genetic diversity, particularly for improving tolerance to both biotic and abiotic stresses. In addition, they offer potential for improving yield, quality, and adaptability to harsh environmental conditions, whilst also broadening the genetic base of cultivated crops [172]. In this context, chromosome manipulation remains one of the most important tools available to plant breeders for introducing novel variation into crop varieties (for reviews, see [173,174,175]).
Genomic resources, such as genetic markers, reference genomes, transcriptomes, gene expression profiles, and protein databases, are essential tools in plant breeding. They support the identification of key traits, the analysis of genetic diversity, genomic mapping, and marker-assisted selection, and help to accelerate the development of improved cultivars. Modern genomic approaches are increasingly displacing traditional cytogenetic analyses in current research [176]. However, molecular cytogenetic studies of chromosomes in both crops and CWRs remain vital for understanding their evolution, genetic recombination patterns, and karyotypic stability [175,177]. To maintain the desirable traits bred into elite cultivars, the amount of alien genetic material introduced must be carefully regulated. For this purpose, a combination of cytomolecular tools, including FISH and GISH, is widely employed to examine and select desirable genotypes (section V of Figure 2). Examples of recent achievements in the study of breeding lines using molecular cytogenetic methods are presented in Table 5.
Cereals are the primary staple crops in most regions of the world and represent one of the crop groups in which considerable efforts have been directed towards introgression breeding. Interest in chromosome organisation in the diploid progenitors of common wheat, as well as in wild wheat species, arises largely from their value as sources of novel genes that were lost during domestication. Recent research includes the development and cytogenetic characterisation of introgression lines of bread wheat (Triticum aestivum) with rye [178,179,180,181], and numerous CWRs within the Triticeae tribe that exhibit various ploidy levels, such as Agropyron cristatum [68,182] (Figure 2M), Aegilops biuncialis [183], Ae. geniculata [69,184,185] (Figure 2N), Elymus sibiricus [186], Leymus mollis [187], Psathyrostachys huashanica [188,189], Thinopyrum intermedium [190], and Th. ponticum [191]. The primary traits of interest have included resistance to leaf rust, stripe rust, Fusarium head blight, and powdery mildew, together with desirable agronomic characteristics such as plant height, spike length, elongated glumes, and increased grain size. Similarly, chromosomal translocations carrying leaf rust resistance genes have been characterised in triticale–Aegilops kotschyi, and triticale–Ae. tauschii translocation lines [178].
Table 5. A selection of recent research articles on cytogenetics-assisted crop improvement.
Table 5. A selection of recent research articles on cytogenetics-assisted crop improvement.
Research ObjectResearch ApproachTrait(s) of Interest, Aims, and Key FindingsReferences
×Triticosecale introgression linesGISH with gDNA of Aegilops sharonensis and Ae. taushiiLeaf rust caused by Puccinia triticina: identification of Ae. kotschyi and
Ae. tauschii chromosome segments in triticale translocation lines carrying resistance genes
[178]
Triticum aestivum–Secale cereale introgression linesGISH with gDNA of S. cereale; FISH with the repetitive sequence pAs1 and pSc119.2 probes; the 6c6 wheat-specific centromeric probe; the pMD-CEN3 S. cereale-specific centromeric probe; and an Arabidopsis thaliana-type (T3AG3) telomeric probeStripe rust caused by Puccinia striiformis: cytomolecular characterisation of the wheat–rye T1RS.1BL translocation line, including the presence of complex chromosome translocations[179,180]
×Triticosecale × wheat derivativesGISH with gDNA of S. cereale; FISH with pSc119.2, pTa71 (35S rDNA), and pAs1 probesYellow rust resistance: identification of the 1RS.1BL translocation in triticale × wheat progenies[181]
Triticum aestivum–Agropyron cristatum
introgression lines
GISH with gDNA of A. cristatum; oligo-FISH with pSc119.2-1, pTa531-1, pAcCR1, and CCS1 probesPlant height and leaf size: identification of spontaneous T1AL.1PS and T1AS.1PL Robertsonian translocations in the wheat–A. cristatum translocation lines[68]
Triticum aestivum–Agropyron cristatum introgression lineGISH with gDNA of A. cristatumMultiple elite agronomic traits, including high resistance to powdery mildew and leaf rust: characterisation of the wheat–A. cristatum disomic 6P addition line[182]
Triticum aestivum–Aegilops biuncialis introgression lineGISH with gDNA of Ae. umbelulata, Ae. comosa, and T. turgidum; oligo-FISH with pAs1 and pSc119.2 probesGlume properties: characterisation of the wheat–Ae. biuncialis 5Mb disomic addition line[183]
Triticum aestivum–Aegilops geniculata introgression linesGISH with gDNA of Ae. geniculata; oligo-FISH with pAs1 and pSc119.2 probesFusarium head blight, powdery mildew, and stripe rust resistance: characterisation of substitution lines with high resistance to these diseases, derived from hybrid progeny between Ae. geniculata and hexaploid wheat[69]
Triticum aestivum–Aegilops geniculata introgression linesGISH with gDNA of Ae. geniculata; oligo-FISH with pSc119.2 and pTa535 probesStripe rust (3Mg DAL) and powdery mildew (7Mg DAL) resistance: characterisation of wheat–Ae. geniculata disomic addition lines[184,185]
Triticum aestivum–Elymus sibiricus introgression lineGISH with gDNA of E. sibiricus; FISH with 35S rDNA-targeting probeLeaf rust resistance: characterisation of the novel wheat–E. sibiricus 3St addition line[186]
Triticum aestivum–Leymus mollis introgression lineGISH with gDNA of L. mollis; oligo-FISH with pSc119.2 and pTa535 probesStripe rust resistance, spike length: characterisation of a novel wheat–L. mollis 2Ns (2D) disomic substitution line[187]
Interspecific derivatives between Triticum aestivum and Psathyrostachys huashanicaGISH with gDNA of P. huashanica; oligo-FISH with pSc119.2 and oligo-pTa535 probesFusarium head blight resistance: identifying and characterising two pathogen-resistant interspecific derivatives: wheat–P. huashanica 1Ns long arm ditelosomic addition line and 2Ns substitution line[188]
Triticum aestivum–Psathyrostachys huashanica introgression lineGISH with gDNA of P. huashanica; oligo-FISH with pSc119.2 and pTa535 probesSeveral elite agronomic traits, including elongated glumes, longer spikes, larger grains, and resistance to Fusarium head blight: characterisation of the wheat–P. huashanica 3Ns disomic 6P addition line[189]
Triticum aestivum × Thinopyrum intermedium derivativesGISH with gDNA of Th. bessarabicum; oligo-FISH with probes pAs1-1, pAs1-3, AFA-4, (GAA)10, and pSc119.2-1Fusarium head blight resistance: examining the chromosome composition of five wheat–Th. intermedium partial amphiploids with J-genome chromosomes[190]
Triticum aestivum–Thinopyrum intermedium and Triticum aestivum–Th. ponticum introgression linesGISH with gDNA of Th. bessarabicum, Th. intermedium, and Th. ponticum; oligo-FISH with pSc119.2 and pTa535 probesStripe rust resistance: characterisation of wheat–Thinopyrum disomic
substitution lines
[191]
Brassica juncea–B. fruticulosa introgression linesGISH with gDNA of B. fruticulosa and B. nigra; oligo-FISH with probes designed using the
B. rapa genome and the repetitive sequence CentBr2 probe
Mustard aphid resistance: identification of introgressions from wild species into the crop and tracking the stability of introgressed fragments of interest across generations[192]
Hibiscus cannabinusFISH with 18S-1 (35S rDNA), pXV1 (5S rDNA), and pLT11 (telomeric) probesInitial comparative cytogenetic characterisation of kenaf landrace and breeding lines[193]
Phaseolus vulgarisFISH with 5S and 35S rDNA-targeting probesCytogenetic characterisation of 154 common bean accessions: high polymorphism in the number of 45S rDNA sites among the five accessions studied by FISH[194]
Camelina sativaFISH with 5S and 35S rDNA-targeting probesCytogenetic characterisation of nine C. sativa genotypes: high polymorphism in the number of 5S and 45S rDNA sites[195]
Veronica species/cultivars and their progeniesFISH with 5S and 35S rDNA-targeting probesImproving Veronica breeding programmes: pre-screening of hybrids, identification of true hybrids, self-pollinated progenies, and false hybrids[196]
Gentiana cruciata and G. tibetica somatic hybridsGISH with gDNA of G. tibetica; FISH with 5S and 35S rDNA-targeting probesCytogenetic characterisation of interspecific somatic hybrids: relatively high chromosomal stability with a predominance of G. cruciata
chromosomes
[197]
Lilium davidii var. unicolor, L. regale and Lilium intersectional hybridsGISH with gDNA of L. longiflorum
and L. speciosum ‘gloriosoides’; oligo-FISH with pTA794 (5S rDNA), telomeric, and pITS probes
Improving lily breeding: characterising the genomic composition of hybrid progeny and determining the parental origin of specific chromosomes; non-denaturing FISH provides an advantage when reprobing slides[198]
Gossypium hirsutum–G. anomalum chromosome segment substitution linesOligo-FISH with painting probes specific for chromosomes 6 (Chr. 06), 9 (Chr. 9), and 11 (Chr. 11) of G. anomalumChromosome-specific identification of G. anomalum introgressions in a G. hirsutum background, supporting the SSR and resequencing data[199]
FISH, using various probes with particular attention to gDNA and sequences targeting 5S and 35S rDNA, has also proven useful in the characterisation of breeding lines of Brassica junceaB. fruticulosa [192], Hibiscus cannabinus [193], Phaseolus vulgaris [194], and Camelina sativa [195], in supporting breeding programmes for Veronica species [196], and in the analysis of somatic hybrids of Gentiana cruciata and G. tibetica [197], as well as inter-sectional hybrids of Lilium [198]. The potential of oligo-FISH using CP probes to detect chromosome segments introgressed into Gossypium hirsutum from its stress-tolerant wild diploid relative G. anomalum, has recently been demonstrated by Xu et al. [199]. This highlighted the value of this approach as a tool to facilitate interspecific introgression breeding in this key fibre crop.

7. Further Current Fields of Plant Cytomolecular Research

In the previous sections, we described the latest achievements in the main research areas related to plant molecular cytogenetics. Here, we briefly highlight several additional current fields in which modern molecular cytogenetic approaches play an important role. Among these is the analysis of the repeatome, excluding studies focused on rDNA, which were discussed earlier. Repetitive DNA sequences constitute a major component of nuclear genomes, acting as a structural backbone in centromeres and telomeres, driving genome evolution, and contributing to the regulation of gene expression [200]. In some angiosperms, they may account for up to 90% of the genome [201]. Their variable abundance, high sequence diversity, and distinct chromosomal distributions contribute significantly to interspecific genome divergence and, together with polyploidy, to the remarkable range of genome sizes observed among seed plants [202]. The advent of high-throughput sequencing technologies has enabled detailed exploration of the repetitive fraction of plant genomes and facilitated comparative repeatome analyses [203]. These advances have been further supported by the development of novel computational tools, such as RepeatExplorer [204] (and its updated version, RepeatExplorer2 [205]), as well as TAREAN [206], which enable de novo identification of diverse classes of repetitive DNA elements. These tools employ sequence clustering algorithms to generate graphical representations (graph layouts) of repeat clusters based on sequence similarity and graph connectivity. Each cluster is then analysed for similarity to known repeats in existing databases, such as the database of retrotransposon protein domains (REXdb). The resulting cluster shapes are often characteristic of specific repeat types: for example, circular or ring-like graph structures typically indicate tandemly arranged satellite DNA, whereas more complex, branching patterns are commonly associated with dispersed elements, such as long terminal repeat (LTR) retrotransposons. This graphical approach thus serves as a valuable visual aid in interpreting repeatome composition and dynamics. However, the precise chromosomal localisation of predicted repeats often requires validation through FISH analysis. In recent years, the combined application of RepeatExplorer and FISH has been employed to identify and characterise DNA repeats in a range of species, including Ensete glaucum [207], Juncus effusus [208], Cenchrus ciliaris [209], members of the Phaseoleae tribe [210], Cuscuta [211], Hydrangea [212], Erythrostemon hughesii [213], and representatives of the Leptostachyus group within the genus Phaseolus [214]. Repetitive DNA often exhibits species-specific genomic profiles, which can aid in understanding interspecific relationships and support taxonomic classification [215]. Genome-specific repetitive sequences, when used as FISH probes, have enabled the discrimination of different genomes in polyploid Urochloa [216] and the tracking of intergenomic translocation patterns across various Avena species [217]. In Petunia × hybrida, the application of a set of satellite repeat families demonstrated that recent hybridisation during breeding preserved the chromosomal positions of repeats but altered their copy numbers [140]. Recent studies have also focused on the characterisation of repetitive DNA in terms of its sequence composition and its contribution to various chromosomal structures, such as centromeres. This has been extensively investigated across a range of plant species, including Juncus effusus [208], Sorghum bicolor [215], representatives of Saccharum [218], Gossypium anomalum [136], and Populus trichocarpa [219]. Notably, in Triticum aestivum, centromeric localisation of sequences other than satellite DNA has been reported, while satellite DNAs were predominantly localised to subtelomeric regions. This pattern reveals asymmetry in subtelomere organisation among the bread wheat subgenomes and suggests its potential significance in facilitating homologous chromosome recognition and pairing during meiosis [220]. Similarly, four satellite DNAs and several LTR retrotransposons have been identified in most subtelomeric regions of Erythrostemon hughesii chromosomes [213].
The application of molecular cytogenetic techniques is also crucial for the reliable assessment of early genetic effects on nuclear genome instability following exposure to chemical and physical mutagens, with particular emphasis on the investigation of micronuclei, as reviewed in [221]. In this context, FISH with various repetitive DNA probes has proven effective in barley [222], Brachypodium [223], and maize [224], with the model grass providing particularly comprehensive insight into micronuclei composition through the use of CP probes [225,226]. Furthermore, recent studies in Brachypodium on DNA methylation and histone modifications, using fluorescently labelled antibodies, have shed new light on the role of epigenetic regulation in micronuclei induction under mutagenic conditions [44,45]. In the future, such approaches may contribute to more accurate assessments of the impact of environmental stress on plant genome integrity.
Molecular cytogenetics provides valuable tools for dissecting various aspects of meiosis. For example, Zhang et al. [67] employed FISH with CP probes to examine chromosome behaviour during meiosis in a series of autopolyploid clones of Saccharum spontaneum. Despite their broad ploidy range, all clones exhibited stable, diploid-like chromosome behaviour, with homologues predominantly forming bivalents (Figure 2K). Similarly, homologous chromosome pairing has been studied in rice, where native chromosome ends have been shown to play a critical role in initiating the process [227]. Corresponding analyses in tetraploid maize highlighted the importance of DNA sequence similarity in promoting preferential homologous pairing [228]. These factors, along with other mechanisms such as the well-characterised Ph1 locus in polyploid wheats [229] and the recently identified BnaPh1 QTL in Brassica napus [230], contribute to the regulation of chromosome pairing in allopolyploids.
Recent studies have also highlighted the role of molecular cytogenetics, with particular emphasis on fluorescent immunolocalisation, in improving our understanding of various aspects of meiotic recombination in arabidopsis. For example, Blackwell et al. [231] demonstrated that the mismatch repair protein MSH2 promotes crossovers in genomic regions with higher sequence diversity. Similarly, Zhu et al. [232] reported the importance of the SMC5/6 (STRUCTURAL MAINTENANCE OF CHROMOSOMES) complex in maintaining the progression of meiotic recombination. Natural variation in the SNI1 (SUPPRESSOR OF NPR1-1 INDUCIBLE 1) gene, which encodes a component of this complex, was shown to affect its function and may modulate the crossover landscape under varying environmental conditions. In their most recent work, these authors also shed light on the role of the ATR (ATAXIA TELANGIECTASIA AND RAD3-RELATED) kinase, whose inactivation leads to a marked redistribution of crossovers, with a decrease in pericentromeric regions and an increase within the chromosome arms [233].

8. Concluding Remarks and Future Perspectives

Over the past decade, plant molecular cytogenetics has undoubtedly undergone significant actual or potential breakthroughs, driven by advances in genomic and molecular methodologies. One such breakthrough was the introduction and widespread adoption of oligo-FISH-based CP, which enabled the expansion of complex nuclear genome analyses at the microscopic level beyond a limited number of small-genome models, such as arabidopsis and Brachypodium and their relatives, to dozens, if not hundreds, of species, including both non-model and crop species representing a wide range of angiosperm families and nuclear genome sizes. On the other hand, some initially promising approaches, such as CRISPR/Cas [234,235]-based live-cell imaging, have not fulfilled all expectations. This technique utilises a fluorescence-tagged, nuclease-deficient Cas (dCas) protein to track specific DNA sequences in vivo in a programmable manner, and it holds considerable potential for visualising various aspects of nuclear dynamics (for review, see [236]). Although this approach has proven successful in human [237] and murine [238] living cells, its application in plants has so far been limited to the demonstration of telomere dynamics in Nicotiana benthamiana [239]. A similar attempt in arabidopsis [240] and brachypodium resulted in ectopic distribution of GFP signals, exposing the limitations of this methodology in small-genome plants, although it may retain some potential when applied to fixed material [241].
The CRISPR/Cas system has also recently been applied to chromosome-scale engineering through the generation of targeted CRs [242]. As demonstrated in arabidopsis, inversions induced using this approach can influence local recombination patterns, either restoring meiotic crossovers in chromosomal regions that were previously inert to genetic exchange [243], or, conversely, massively suppressing meiotic recombination [244]. This new method of manipulating chromosomes holds considerable potential in plant biology and biotechnology. Its applications include breaking genetic linkages between specific genes, reversing natural inversions that suppress recombination in breeding programmes, artificially establishing genetic isolation, and constructing mini-cargo chromosomes, as reviewed by Puchta and Houben [245]. The most recent studies highlight the usefulness of targeted chromosome engineering for investigating telomere dynamics, chromatin structure, gene expression, and phenotypic stability in arabidopsis [246], and are already providing valuable materials and inspiration for advanced cytomolecular analyses [247].
Another potentially significant breakthrough may lie in the application of AI to cytogenetics, particularly in the context of advanced image analysis [248]. One especially promising development is the integration of deep learning-based AI algorithms into karyotyping software. As recently reviewed by Rosenblum et al. [249], four commercially available AI-assisted karyotyping platforms are already in use for the analysis of human chromosomes in both clinical and research contexts. These systems address various limitations inherent in traditional automated karyotyping, particularly those related to image acquisition, segmentation, chromosome classification, and analytical accuracy, and they hold the potential to redefine the current paradigm of chromosome analysis. Similar approaches could also hold transformative potential in plant systems, although their implementation is likely to be considerably more challenging due to the vast number of plant species that may serve as targets for analysis, as well as the extensive diversity of their karyotypes. Nevertheless, an emerging AI-based image analysis pipeline is currently under development as part of the fifth generation of the long-established CHIAS chromosome image analysis system [250,251,252,253], and it has already yielded encouraging preliminary results for potential application in plant cytogenetics. Furthermore, Oxford Instruments reports the integration of AI into the latest release of its Imaris imaging package [51]. Thus, in light of the current body of evidence, the future of cytomolecular analyses in plants appears prospective, particularly as an effective means of bridging genomic and transcriptomic data with advanced microscopy-based observations and interpretations.

Author Contributions

Conceptualisation, R.H.; writing—original draft preparation, E.W. and R.H.; writing—review and editing, E.W., L.A.J.M., N.O., Z.Y., K.W. and R.H. All authors have read and agreed to the published version of the manuscript.

Funding

E.W. and R.H. acknowledge support from the Research Excellence Initiative of the University of Silesia in Katowice. This work was also supported by the Japan Society for the Promotion of Science (JSPS) KAKENHI Programme (project no. JP25K01986, awarded to N.O.).

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
2DTwo-dimensional
3DThree-dimensional
3CChromosome conformation capture
ACKAncestral Crucifer Karyotype
AIArtificial intelligence
APKAncestral Phaseoleae Karyotype
arabidopsisArabidopsis thaliana
BACBacterial artificial chromosome
BrachypodiumBrachypodium distachyon
CasCRISPR-associated protein
CCPComparative chromosome painting
CENH3Centromeric histone H3 variant
ChIA-PETChromatin interaction analysis by paired-end tag sequencing
CPChromosome painting
CRChromosome rearrangement
CRISPRClustered regularly interspaced short palindromic repeats
CTChromosome territory
CWRCrop wild relative
dCasNuclease-deficient (‘dead’) CRISPR-associated protein
EdU5-ethynyl-2′-deoxyuridine
FISHFluorescence in situ hybridisation
FITCFluorescein isothiocyanate
gDNATotal genomic DNA
GISHGenomic in situ hybridisation
Hi-CHigh-throughput chromatin conformation capture
HiChIPIn situ Hi-C followed by chromatin immunoprecipitation
HVTEMHigh-voltage transmission electron microscopy
ISHIn situ hybridisation
LTRLong terminal repeat
MYMillion years
NDNucleolar dominance
Oligo-FISHOligonucleotide fluorescence in situ hybridisation
rDNARibosomal DNA
REXdbDatabase of retrotransposon protein domains
rRNARibosomal RNA
SEMScanning electron microscopy
SIMStructured illumination microscopy
TEMTransmission electron microscopy
TopoTopoisomerase
UHVTEMUltra-high-voltage transmission electron microscopy
WTWild type

References

  1. Fransz, P.; van de Belt, J.; de Jong, H. Extended DNA fibers for high-resolution mapping. Methods Mol. Biol. 2023, 2672, 351–363. [Google Scholar] [CrossRef]
  2. Schwarzacher, T.; Heslop-Harrison, J.S. Practical In Situ Hybridization; BIOS Scientific Publishers: Oxford, UK, 2000. [Google Scholar]
  3. Pardue, M.L.; Gall, J.G. Molecular hybridization of radioactive DNA to the DNA of cytological preparations. Proc. Natl. Acad. Sci. USA 1969, 64, 600–604. [Google Scholar] [CrossRef] [PubMed]
  4. Schubert, I.; Wobus, U. In situ hybridization confirms jumping nucleolus organizing regions in Allium. Chromosoma 1985, 92, 143–148. [Google Scholar] [CrossRef]
  5. Weiss, H.; Pasierbek, P.; Maluszynska, J. An improved nonfluorescent detection system for in situ hybridization in plants. Biotech. Histochem. 2000, 75, 49–53. [Google Scholar] [CrossRef] [PubMed]
  6. Maluszynska, J.; Schweizer, D. Ribosomal RNA genes in B chromosomes of Crepis capillaris detected by non-radioactive in situ hybridization. Heredity 1989, 62 Pt 1, 59–65. [Google Scholar] [CrossRef] [PubMed]
  7. Jiang, J.; Gill, B.S.; Wang, G.L.; Ronald, P.C.; Ward, D.C. Metaphase and interphase fluorescence in situ hybridization mapping of the rice genome with bacterial artificial chromosomes. Proc. Natl. Acad. Sci. USA 1995, 92, 4487–4491. [Google Scholar] [CrossRef] [PubMed]
  8. Woo, S.S.; Jiang, J.; Gill, B.S.; Paterson, A.H.; Wing, R.A. Construction and characterization of a bacterial artificial chromosome library of Sorghum bicolor. Nucleic Acids Res. 1994, 22, 4922–4931. [Google Scholar] [CrossRef]
  9. Fuchs, J.; Kloos, D.U.; Ganal, M.W.; Schubert, I. In situ localization of yeast artificial chromosome sequences on tomato and potato metaphase chromosomes. Chromosome Res. 1996, 4, 277–281. [Google Scholar] [CrossRef]
  10. Hagemann, S.; Scheer, B.; Schweizer, D. Repetitive sequences in the genome of Anemone blanda: Identification of tandem arrays and of dispersed repeats. Chromosoma 1993, 102, 312–324. [Google Scholar] [CrossRef]
  11. Moscone, E.A.; Matzke, M.A.; Matzke, A.J. The use of combined FISH/GISH in conjunction with DAPI counterstaining to identify chromosomes containing transgene inserts in amphidiploid tobacco. Chromosoma 1996, 105, 231–236. [Google Scholar] [CrossRef]
  12. Fuchs, J.; Strehl, S.; Brandes, A.; Schweizer, D.; Schubert, I. Molecular-cytogenetic characterization of the Vicia faba genome--heterochromatin differentiation, replication patterns and sequence localization. Chromosome Res. 1998, 6, 219–230. [Google Scholar] [CrossRef]
  13. Hasterok, R.; Jenkins, G.; Langdon, T.; Jones, R.N.; Maluszynska, J. Ribosomal DNA is an effective marker of Brassica chromosomes. Theor. Appl. Genet. 2001, 103, 486–490. [Google Scholar] [CrossRef]
  14. Moscone, E.A.; Klein, F.; Lambrou, M.; Fuchs, J.; Schweizer, D. Quantitative karyotyping and dual-color FISH mapping of 5S and 18S-25S rDNA probes in the cultivated Phaseolus species (Leguminosae). Genome 1999, 42, 1224–1233. [Google Scholar] [CrossRef]
  15. Snowdon, R.J.; Friedt, W.; Köhler, A.; Köhler, W. Molecular cytogenetic localization and characterization of 5S and 25S rDNA loci for chromosome identification in oilseed rape (Brassica napus L.). Ann. Bot. 2000, 86, 201–204. [Google Scholar] [CrossRef]
  16. Hasterok, R.; Dulawa, J.; Jenkins, G.; Leggett, M.; Langdon, T. Multi-substrate chromosome preparations for high throughput comparative FISH. BMC Biotechnol. 2006, 6, 20. [Google Scholar] [CrossRef] [PubMed]
  17. Hasterok, R.; Langdon, T.; Taylor, S.; Jenkins, G. Combinatorial labelling of DNA probes enables multicolour fluorescence in situ hybridisation in plants. Folia Histochem. Cytobiol. 2002, 40, 319–323. [Google Scholar] [PubMed]
  18. The Arabidopsis Genome Initiative. Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 2000, 408, 796–815. [Google Scholar] [CrossRef] [PubMed]
  19. Lysak, M.A.; Fransz, P.F.; Ali, H.B.; Schubert, I. Chromosome painting in Arabidopsis thaliana. Plant J. 2001, 28, 689–697. [Google Scholar] [CrossRef] [PubMed]
  20. Lysak, M.A.; Pecinka, A.; Schubert, I. Recent progress in chromosome painting of Arabidopsis and related species. Chromosome Res. 2003, 11, 195–204. [Google Scholar] [CrossRef]
  21. Lysak, M.A.; Berr, A.; Pecinka, A.; Schmidt, R.; McBreen, K.; Schubert, I. Mechanisms of chromosome number reduction in Arabidopsis thaliana and related Brassicaceae species. Proc. Natl. Acad. Sci. USA 2006, 103, 5224–5229. [Google Scholar] [CrossRef]
  22. Pecinka, A.; Schubert, V.; Meister, A.; Kreth, G.; Klatte, M.; Lysak, M.A.; Fuchs, J.; Schubert, I. Chromosome territory arrangement and homologous pairing in nuclei of Arabidopsis thaliana are predominantly random except for NOR-bearing chromosomes. Chromosoma 2004, 113, 258–269. [Google Scholar] [CrossRef] [PubMed]
  23. Tang, X.; Szinay, D.; Lang, C.; Ramanna, M.S.; van der Vossen, E.A.; Datema, E.; Lankhorst, R.K.; de Boer, J.; Peters, S.A.; Bachem, C.; et al. Cross-species bacterial artificial chromosome-fluorescence in situ hybridization painting of the tomato and potato chromosome 6 reveals undescribed chromosomal rearrangements. Genetics 2008, 180, 1319–1328. [Google Scholar] [CrossRef] [PubMed]
  24. Idziak, D.; Betekhtin, A.; Wolny, E.; Lesniewska, K.; Wright, J.; Febrer, M.; Bevan, M.W.; Jenkins, G.; Hasterok, R. Painting the chromosomes of Brachypodium: Current status and future prospects. Chromosoma 2011, 120, 469–479. [Google Scholar] [CrossRef]
  25. Betekhtin, A.; Jenkins, G.; Hasterok, R. Reconstructing the evolution of Brachypodium genomes using comparative chromosome painting. PLoS ONE 2014, 9, e115108. [Google Scholar] [CrossRef]
  26. Peterson, D.G.; Tomkins, J.P.; Frisch, D.A.; Wing, R.A.; Paterson, A.H. Construction of Plant Bacterial Artificial Chromosome (BAC) Libraries: An illustrated Guide. Available online: https://www.mgel.msstate.edu/pubs/bacman2.pdf (accessed on 15 July 2025).
  27. Beliveau, B.J.; Joyce, E.F.; Apostolopoulos, N.; Yilmaz, F.; Fonseka, C.Y.; McCole, R.B.; Chang, Y.; Li, J.B.; Senaratne, T.N.; Williams, B.R.; et al. Versatile design and synthesis platform for visualizing genomes with Oligopaint FISH probes. Proc. Natl. Acad. Sci. USA 2012, 109, 21301–21306. [Google Scholar] [CrossRef]
  28. Boyle, S.; Rodesch, M.J.; Halvensleben, H.A.; Jeddeloh, J.A.; Bickmore, W.A. Fluorescence in situ hybridization with high-complexity repeat-free oligonucleotide probes generated by massively parallel synthesis. Chromosome Res. 2011, 19, 901–909. [Google Scholar] [CrossRef] [PubMed]
  29. Yamada, N.A.; Rector, L.S.; Tsang, P.; Carr, E.; Scheffer, A.; Sederberg, M.C.; Aston, M.E.; Ach, R.A.; Tsalenko, A.; Sampas, N.; et al. Visualization of fine-scale genomic structure by oligonucleotide-based high-resolution FISH. Cytogenet. Genome Res. 2011, 132, 248–254. [Google Scholar] [CrossRef] [PubMed]
  30. Han, Y.; Zhang, T.; Thammapichai, P.; Weng, Y.; Jiang, J. Chromosome-specific painting in Cucumis species using bulked oligonucleotides. Genetics 2015, 200, 771–779. [Google Scholar] [CrossRef]
  31. Liu, G.; Zhang, T. Single copy oligonucleotide fluorescence in situ hybridization probe design platforms: Development, application and evaluation. Int. J. Mol. Sci. 2021, 22, 7124. [Google Scholar] [CrossRef]
  32. Harun, A.; Liu, H.; Song, S.; Asghar, S.; Wen, X.; Fang, Z.; Chen, C. Oligonucleotide fluorescence in situ hybridization: An efficient chromosome painting method in plants. Plants 2023, 12, 2816. [Google Scholar] [CrossRef]
  33. Schwacke, R.; Bolger, M.; Usadel, B. PubPlant—A continuously updated online resource for sequenced and published plant genomes. Front. Plant Sci. 2025, 16, 1603547. [Google Scholar] [CrossRef]
  34. Schwacke, R.; Bolger, M.; Usadel, B. PubPlant—A Continuously Updated Resource for Sequenced and Published Plant Genomes. Available online: https://www.plabipd.de/pubplant_timeline.html (accessed on 15 July 2025).
  35. Albert, P.S.; Zhang, T.; Semrau, K.; Rouillard, J.M.; Kao, Y.H.; Wang, C.R.; Danilova, T.V.; Jiang, J.; Birchler, J.A. Whole-chromosome paints in maize reveal rearrangements, nuclear domains, and chromosomal relationships. Proc. Natl. Acad. Sci. USA 2019, 116, 1679–1685. [Google Scholar] [CrossRef]
  36. Li, G.R.; Zhang, T.; Yu, Z.H.; Wang, H.J.; Yang, E.N.; Yang, Z.J. An efficient Oligo-FISH painting system for revealing chromosome rearrangements and polyploidization in Triticeae. Plant J. 2021, 105, 978–993. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, X.; Sun, S.; Wu, Y.; Zhou, Y.; Gu, S.; Yu, H.; Yi, C.; Gu, M.; Jiang, J.; Liu, B.; et al. Dual-color oligo-FISH can reveal chromosomal variations and evolution in Oryza species. Plant J. 2020, 101, 112–121. [Google Scholar] [CrossRef] [PubMed]
  38. Fuchs, J.; Demidov, D.; Houben, A.; Schubert, I. Chromosomal histone modification patterns—from conservation to diversity. Trends Plant Sci. 2006, 11, 199–208. [Google Scholar] [CrossRef] [PubMed]
  39. Wolny, E.; Braszewska-Zalewska, A.; Hasterok, R. Spatial distribution of epigenetic modifications in Brachypodium distachyon embryos during seed maturation and germination. PLoS ONE 2014, 9, e101246. [Google Scholar] [CrossRef] [PubMed]
  40. Wolny, E.; Braszewska-Zalewska, A.; Kroczek, D.; Hasterok, R. Histone H3 and H4 acetylation patterns are more dynamic than those of DNA methylation in Brachypodium distachyon embryos during seed maturation and germination. Protoplasma 2017, 254, 2045–2052. [Google Scholar] [CrossRef]
  41. Borowska-Zuchowska, N.; Hasterok, R. Epigenetics of the preferential silencing of Brachypodium stacei-originated 35S rDNA loci in the allotetraploid grass Brachypodium hybridum. Sci. Rep. 2017, 7, 5260. [Google Scholar] [CrossRef] [PubMed]
  42. Zheng, T.; Nibau, C.; Phillips, D.W.; Jenkins, G.; Armstrong, S.J.; Doonan, J.H. CDKG1 protein kinase is essential for synapsis and male meiosis at high ambient temperature in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2014, 111, 2182–2187. [Google Scholar] [CrossRef]
  43. Choi, K.; Zhao, X.; Tock, A.J.; Lambing, C.; Underwood, C.J.; Hardcastle, T.J.; Serra, H.; Kim, J.; Cho, H.S.; Kim, J.; et al. Nucleosomes and DNA methylation shape meiotic DSB frequency in Arabidopsis thaliana transposons and gene regulatory regions. Genome Res. 2018, 28, 532–546. [Google Scholar] [CrossRef]
  44. Bara, A.W.; Braszewska, A.; Kwasniewska, J. DNA nethylation-an epigenetic mark in mutagen-treated Brachypodium distachyon cells. Plants 2021, 10, 1408. [Google Scholar] [CrossRef] [PubMed]
  45. Bara-Halama, A.W.; Idziak-Helmcke, D.; Kwasniewska, J. Unraveling the DNA methylation in the rDNA foci in mutagen-induced Brachypodium distachyon micronuclei. Int. J. Mol. Sci. 2022, 23, 6797. [Google Scholar] [CrossRef] [PubMed]
  46. Wolny, E.; Skalska, A.; Braszewska, A.; Mur, L.A.J.; Hasterok, R. Defining the cell wall, cell cycle and chromatin landmarks in the responses of Brachypodium distachyon to salinity. Int. J. Mol. Sci. 2021, 22, 949. [Google Scholar] [CrossRef] [PubMed]
  47. Ohmido, N.; Polosoro, A. Chromatin immunostaining of plant nuclei. In Plant Cytogenetics and Cytogenomics. Methods in Molecular Biology; Heitkam, T., Garcia, S., Eds.; Humana: New York, NY, USA, 2023; Volume 2672, pp. 233–244. [Google Scholar]
  48. Heintzmann, R.; Huser, T. Super-resolution structured illumination microscopy. Chem. Rev. 2017, 117, 13890–13908. [Google Scholar] [CrossRef] [PubMed]
  49. Hickey, S.M.; Ung, B.; Bader, C.; Brooks, R.; Lazniewska, J.; Johnson, I.R.D.; Sorvina, A.; Logan, J.; Martini, C.; Moore, C.R.; et al. Fluorescence microscopy-an outline of hardware, biological handling, and fluorophore considerations. Cells 2021, 11, 35. [Google Scholar] [CrossRef] [PubMed]
  50. Chen, X.; Zhong, S.; Hou, Y.; Cao, R.; Wang, W.; Li, D.; Dai, Q.; Kim, D.; Xi, P. Superresolution structured illumination microscopy reconstruction algorithms: A review. Light Sci. Appl. 2023, 12, 172. [Google Scholar] [CrossRef]
  51. Imaris AI Microscopy Image Analysis Software. Available online: https://imaris.oxinst.com/ (accessed on 15 July 2025).
  52. ImageJ Image Processing and Analysis in Java. Available online: https://imagej.net/ij/ (accessed on 15 July 2025).
  53. Febrer, M.; Goicoechea, J.L.; Wright, J.; McKenzie, N.; Song, X.; Lin, J.; Collura, K.; Wissotski, M.; Yu, Y.; Ammiraju, J.S.; et al. An integrated physical, genetic and cytogenetic map of Brachypodium distachyon, a model system for grass research. PLoS ONE 2010, 5, e13461. [Google Scholar] [CrossRef]
  54. The International Brachypodium Initiative. Genome sequencing and analysis of the model grass Brachypodium distachyon. Nature 2010, 463, 763–768. [Google Scholar] [CrossRef]
  55. Xiong, Y.; Yuan, S.; Xiong, Y.; Li, L.; Peng, J.; Zhang, J.; Fan, X.; Jiang, C.; Sha, L.N.; Wang, Z.; et al. Analysis of allohexaploid wheatgrass genome reveals its Y haplome origin in Triticeae and high-altitude adaptation. Nat. Commun. 2025, 16, 3104. [Google Scholar] [CrossRef]
  56. Ten Thousand Plant Genome Project. Available online: https://en.genomics.cn/en-project-dzwyjs-6184.html (accessed on 15 July 2025).
  57. Jenkins, G.; Hasterok, R. BAC ‘landing’ on chromosomes of Brachypodium distachyon for comparative genome alignment. Nat. Protoc. 2007, 2, 88–98. [Google Scholar] [CrossRef]
  58. Dolezalova, A.; Sladekova, L.; Simonikova, D.; Holusova, K.; Karafiatova, M.; Varshney, R.K.; Dolezel, J.; Hribova, E. Karyotype differentiation in cultivated chickpea revealed by oligopainting fluorescence in situ hybridization. Front. Plant Sci. 2021, 12, 791303. [Google Scholar] [CrossRef]
  59. Farhat, P.; Mandakova, T.; Divisek, J.; Kudoh, H.; German, D.A.; Lysak, M.A. The evolution of the hypotetraploid Catolobus pendulus genome—The poorly known sister species of Capsella. Front. Plant Sci. 2023, 14, 1165140. [Google Scholar] [CrossRef] [PubMed]
  60. Berankova, D.; Cizkova, J.; Majzlikova, G.; Dolezalova, A.; Mduma, H.; Brown, A.; Swennen, R.; Hribova, E. Striking variation in chromosome structure within Musa acuminata subspecies, diploid cultivars, and F1 diploid hybrids. Front. Plant Sci. 2024, 15, 1387055. [Google Scholar] [CrossRef] [PubMed]
  61. Dolezalova, A.; Berankova, D.; Kolackova, V.; Hribova, E. Insight into chromatin compaction and spatial organization in rice interphase nuclei. Front. Plant Sci. 2024, 15, 1358760. [Google Scholar] [CrossRef] [PubMed]
  62. Municio, C.; Antosz, W.; Grasser, K.D.; Kornobis, E.; Van Bel, M.; Eguinoa, I.; Coppens, F.; Brautigam, A.; Lermontova, I.; Bruckmann, A.; et al. The Arabidopsis condensin CAP-D subunits arrange interphase chromatin. New Phytol. 2021, 230, 972–987. [Google Scholar] [CrossRef]
  63. Wang, N.; Wang, Z.; Tzourtzou, S.; Wang, X.; Bi, X.; Leimeister, J.; Xu, L.; Sakamoto, T.; Matsunaga, S.; Schaller, A.; et al. The plant nuclear lamina disassembles to regulate genome folding in stress conditions. Nat. Plants 2023, 9, 1081–1093. [Google Scholar] [CrossRef]
  64. Macas, J.; Avila Robledillo, L.; Kreplak, J.; Novak, P.; Koblizkova, A.; Vrbova, I.; Burstin, J.; Neumann, P. Assembly of the 81.6 Mb centromere of pea chromosome 6 elucidates the structure and evolution of metapolycentric chromosomes. PLoS Genet. 2023, 19, e1010633. [Google Scholar] [CrossRef]
  65. Baez, M.; Kuo, Y.T.; Dias, Y.; Souza, T.; Boudichevskaia, A.; Fuchs, J.; Schubert, V.; Vanzela, A.L.L.; Pedrosa-Harand, A.; Houben, A. Analysis of the small chromosomal Prionium serratum (Cyperid) demonstrates the importance of reliable methods to differentiate between mono- and holocentricity. Chromosoma 2020, 129, 285–297. [Google Scholar] [CrossRef]
  66. Randall, R.S.; Jourdain, C.; Nowicka, A.; Kaduchova, K.; Kubova, M.; Ayoub, M.A.; Schubert, V.; Tatout, C.; Colas, I.; Kalyanikrishna; et al. Image analysis workflows to reveal the spatial organization of cell nuclei and chromosomes. Nucleus 2022, 13, 277–299. [Google Scholar] [CrossRef]
  67. Zhang, X.; Meng, Z.; Han, J.L.; Khurshid, H.; Esh, A.; Hasterok, R.; Wang, K. Characterization of meiotic chromosome behavior in the autopolyploid Saccharum spontaneum reveals preferential chromosome pairing without distinct DNA sequence variation. Crop J. 2023, 11, 1550–1558. [Google Scholar] [CrossRef]
  68. Han, B.H.; Wang, X.; Sun, Y.Y.; Kang, X.L.; Zhang, M.; Luo, J.W.; Han, H.M.; Zhou, S.H.; Lu, Y.Q.; Liu, W.H.; et al. Pre-breeding of spontaneous Robertsonian translocations for density planting architecture by transferring Agropyron cristatum chromosome 1P into wheat. Theor. Appl. Genet. 2024, 137, 110. [Google Scholar] [CrossRef] [PubMed]
  69. Yang, X.; Xu, M.; Wang, Y.; Cheng, X.; Huang, C.; Zhang, H.; Li, T.; Wang, C.; Chen, C.; Wang, Y.; et al. Development and molecular cytogenetic identification of two wheat-Aegilops geniculata Roth 7Mg chromosome substitution lines with resistance to Fusarium head blight, powdery mildew and stripe rust. Int. J. Mol. Sci. 2022, 23, 7056. [Google Scholar] [CrossRef] [PubMed]
  70. Hasterok, R.; Wang, K.; Jenkins, G. Progressive refinement of the karyotyping of Brachypodium genomes. New Phytol. 2020, 227, 1668–1675. [Google Scholar] [CrossRef]
  71. Wang, K.; Zhang, H.; Khurshid, H.; Esh, A.; Wu, C.W.; Wang, Q.N.; Piperidis, N. Past and recent advances in sugarcane cytogenetics. Crop J. 2023, 11, 1–8. [Google Scholar] [CrossRef]
  72. Bayat, S.; Lysak, M.A.; Mandakova, T. Genome structure and evolution in the cruciferous tribe Thlaspideae (Brassicaceae). Plant J. 2021, 108, 1768–1785. [Google Scholar] [CrossRef]
  73. Gordon, S.P.; Contreras-Moreira, B.; Levy, J.J.; Djamei, A.; Czedik-Eysenberg, A.; Tartaglio, V.S.; Session, A.; Martin, J.; Cartwright, A.; Katz, A.; et al. Gradual polyploid genome evolution revealed by pan-genomic analysis of Brachypodium hybridum and its diploid progenitors. Nat. Commun. 2020, 11, 3670. [Google Scholar] [CrossRef] [PubMed]
  74. Sancho, R.; Inda, L.A.; Diaz-Perez, A.; Des Marais, D.L.; Gordon, S.; Vogel, J.P.; Lusinska, J.; Hasterok, R.; Contreras-Moreira, B.; Catalan, P. Tracking the ancestry of known and ‘ghost’ homeologous subgenomes in model grass Brachypodium polyploids. Plant J. 2022, 109, 1535–1558. [Google Scholar] [CrossRef] [PubMed]
  75. Oliveira, A.; Martins, L.D.V.; Bustamante, F.O.; Munoz-Amatriain, M.; Close, T.; da Costa, A.F.; Benko-Iseppon, A.M.; Pedrosa-Harand, A.; Brasileiro-Vidal, A.C. Breaks of macrosynteny and collinearity among moth bean (Vigna aconitifolia), cowpea (V. unguiculata), and common bean (Phaseolus vulgaris). Chromosome Res. 2020, 28, 293–306. [Google Scholar] [CrossRef]
  76. Ferraz, M.E.; Fonseca, A.; Pedrosa-Harand, A. Multiple and independent rearrangements revealed by comparative cytogenetic mapping in the dysploid Leptostachyus group (Phaseolus L., Leguminosae). Chromosome Res. 2020, 28, 395–405. [Google Scholar] [CrossRef]
  77. de Barros, D.; Montenegro, C.; Gomes, M.; Ferraz, M.E.; Miotto, S.T.S.; Pedrosa-Harand, A. Cytogenetic characterization and karyotype evolution in six Macroptilium species (Leguminosae). Genome 2023, 66, 165–174. [Google Scholar] [CrossRef]
  78. Dias, Y.; Sader, M.A.; Vieira, M.L.C.; Pedrosa-Harand, A. Comparative cytogenetic maps of Passiflora alata and P. watsoniana (Passifloraceae) using BAC-FISH. Plant Syst. Evol. 2020, 306, 51. [Google Scholar] [CrossRef]
  79. Do Vale Martins, L.; de Oliveira Bustamante, F.; da Silva Oliveira, A.R.; da Costa, A.F.; de Lima Feitoza, L.; Liang, Q.; Zhao, H.; Benko-Iseppon, A.M.; Munoz-Amatriain, M.; Pedrosa-Harand, A.; et al. BAC- and oligo-FISH mapping reveals chromosome evolution among Vigna angularis, V. unguiculata, and Phaseolus vulgaris. Chromosoma 2021, 130, 133–147. [Google Scholar] [CrossRef] [PubMed]
  80. Bielski, W.; Ksiazkiewicz, M.; Simonikova, D.; Hribova, E.; Susek, K.; Naganowska, B. The puzzling fate of a lupin chromosome revealed by reciprocal oligo-FISH and BAC-FISH mapping. Genes 2020, 11, 1489. [Google Scholar] [CrossRef] [PubMed]
  81. Dias, S.; de Oliveira Bustamante, F.; do Vale Martins, L.; da Costa, V.A.; Montenegro, C.; Oliveira, A.; de Lima, G.S.; Braz, G.T.; Jiang, J.; da Costa, A.F.; et al. Translocations and inversions: Major chromosomal rearrangements during Vigna (Leguminosae) evolution. Theor. Appl. Genet. 2024, 137, 29. [Google Scholar] [CrossRef]
  82. Serafim, L.; Silva, J.H.; Dias, S.; Oliveira, A.; Nunes, M.C.; da Costa, A.F.; Benko-Iseppon, A.M.; Jiang, J.; do Vale Martins, L.; Brasileiro-Vidal, A.C. “End-to-end chromosome fusion” as the main driver of descending dysploidy in Vigna lasiocarpa (Mart. ex Benth.) Verdc. (Leguminosae Juss.). Plants 2025, 14, 1872. [Google Scholar] [CrossRef]
  83. Chen, C.; Han, Y.; Xiao, H.; Zou, B.; Wu, D.; Sha, L.; Yang, C.; Liu, S.; Cheng, Y.; Wang, Y.; et al. Chromosome-specific painting in Thinopyrum species using bulked oligonucleotides. Theor. Appl. Genet. 2023, 136, 177. [Google Scholar] [CrossRef] [PubMed]
  84. Chen, C.; Zhang, X.; Li, Y.; Zou, B.; Xiao, H.; Han, Y.; Yang, X.; Wu, D.; Sha, L.; Yang, C.; et al. Chromosome-specific painting reveals the Y genome origin and chromosome rearrangements of the St genome in Triticeae. Plant Physiol. 2024, 196, 870–882. [Google Scholar] [CrossRef] [PubMed]
  85. Jiang, C.; Liu, X.; Yang, Z.; Li, G. Chromosome rearrangement in Elymus dahuricus revealed by ND-FISH and oligo-FISH painting. Plants 2023, 12, 3268. [Google Scholar] [CrossRef] [PubMed]
  86. Jiang, W.; Jiang, C.; Yuan, W.; Zhang, M.; Fang, Z.; Li, Y.; Li, G.; Jia, J.; Yang, Z. A universal karyotypic system for hexaploid and diploid Avena species brings oat cytogenetics into the genomics era. BMC Plant Biol. 2021, 21, 213. [Google Scholar] [CrossRef]
  87. Mata-Sucre, Y.; Parteka, L.M.; Ritz, C.M.; Gatica-Arias, A.; Felix, L.P.; Thomas, W.W.; Souza, G.; Vanzela, A.L.L.; Pedrosa-Harand, A.; Marques, A. Oligo-barcode illuminates holocentric karyotype evolution in Rhynchospora (Cyperaceae). Front. Plant Sci. 2024, 15, 1330927. [Google Scholar] [CrossRef]
  88. Shi, P.; Sun, H.; Liu, G.; Zhang, X.; Zhou, J.; Song, R.; Xiao, J.; Yuan, C.; Sun, L.; Wang, Z.; et al. Chromosome painting reveals inter-chromosomal rearrangements and evolution of subgenome D of wheat. Plant J. 2022, 112, 55–67. [Google Scholar] [CrossRef]
  89. Simonikova, D.; Nemeckova, A.; Cizkova, J.; Brown, A.; Swennen, R.; Dolezel, J.; Hribova, E. Chromosome painting in cultivated bananas and their wild relatives (Musa spp.) reveals differences in chromosome structure. Int. J. Mol. Sci. 2020, 21, 7915. [Google Scholar] [CrossRef]
  90. Yu, F.; Zhao, X.; Chai, J.; Ding, X.; Li, X.; Huang, Y.; Wang, X.; Wu, J.; Zhang, M.; Yang, Q.; et al. Chromosome-specific painting unveils chromosomal fusions and distinct allopolyploid species in the Saccharum complex. New Phytol. 2022, 233, 1953–1965. [Google Scholar] [CrossRef]
  91. Zaki, N.M.; Schwarzacher, T.; Singh, R.; Madon, M.; Wischmeyer, C.; Hanim Mohd Nor, N.; Zulkifli, M.A.; Heslop-Harrison, J.S.P. Chromosome identification in oil palm (Elaeis guineensis) using in situ hybridization with massive pools of single copy oligonucleotides and transferability across Arecaceae species. Chromosome Res. 2021, 29, 373–390. [Google Scholar] [CrossRef] [PubMed]
  92. He, L.; Zhao, H.; He, J.; Yang, Z.; Guan, B.; Chen, K.; Hong, Q.; Wang, J.; Liu, J.; Jiang, J. Extraordinarily conserved chromosomal synteny of Citrus species revealed by chromosome-specific painting. Plant J. 2020, 103, 2225–2235. [Google Scholar] [CrossRef] [PubMed]
  93. Qu, M.; Zhang, L.; Li, K.; Sun, J.; Li, Z.; Han, Y. Karyotypic stability of Fragaria (strawberry) species revealed by cross-species chromosome painting. Chromosome Res. 2021, 29, 285–300. [Google Scholar] [CrossRef] [PubMed]
  94. Sun, J.; Zhang, Q.; Xu, M.; Yan, M.; Liu, X.; Sun, J.; Cao, Q.; Wang, H.; Yang, J.; Li, Z.; et al. Comparative karyotype analysis provides cytogenetic evidence for the origin of sweetpotato. Chromosome Res. 2024, 32, 14. [Google Scholar] [CrossRef] [PubMed]
  95. Meng, Z.; Zheng, Q.; Shi, S.; Wang, W.; Wang, F.; Xie, Q.; Chen, X.; Shen, H.; Xiao, G.; Li, H. Whole-chromosome oligo-painting in licorice unveils interspecific chromosomal evolutionary relationships and possible origin of triploid genome species. Plant J. 2024, 120, 2089–2100. [Google Scholar] [CrossRef] [PubMed]
  96. Xu, M.; Guo, H.; Wang, Y.; Zhou, B. Identification of chromosomes by fluorescence in situ hybridization in Gossypium hirsutum via developing oligonucleotide probes. Genome 2024, 67, 64–77. [Google Scholar] [CrossRef]
  97. Bi, Y.; Zhao, Q.; Yan, W.; Li, M.; Liu, Y.; Cheng, C.; Zhang, L.; Yu, X.; Li, J.; Qian, C.; et al. Flexible chromosome painting based on multiplex PCR of oligonucleotides and its application for comparative chromosome analyses in Cucumis. Plant J. 2020, 102, 178–186. [Google Scholar] [CrossRef]
  98. Zhao, Q.; Meng, Y.; Wang, P.; Qin, X.; Cheng, C.; Zhou, J.; Yu, X.; Li, J.; Lou, Q.; Jahn, M.; et al. Reconstruction of ancestral karyotype illuminates chromosome evolution in the genus Cucumis. Plant J. 2021, 107, 1243–1259. [Google Scholar] [CrossRef] [PubMed]
  99. Sun, J.; Chen, L.; Sun, J.; Liu, Z.; Han, Y. Cross-species chromosome painting offers new insights into the phylogenetic relationships among 16 representative species of Ipomoeeae. Front. Plant Sci. 2025, 16, 1610698. [Google Scholar] [CrossRef] [PubMed]
  100. Yu, F.; Chai, J.; Li, X.; Yu, Z.; Yang, R.; Ding, X.; Wang, Q.; Wu, J.; Yang, X.; Deng, Z. Chromosomal characterization of Tripidium arundinaceum revealed by oligo-FISH. Int. J. Mol. Sci. 2021, 22, 8539. [Google Scholar] [CrossRef] [PubMed]
  101. Nascimento, T.; Pedrosa-Harand, A. High rates of structural rearrangements have shaped the chromosome evolution in dysploid Phaseolus beans. Theor. Appl. Genet. 2023, 136, 215. [Google Scholar] [CrossRef]
  102. Bacovsky, V.; Cegan, R.; Simonikova, D.; Hribova, E.; Hobza, R. The formation of sex chromosomes in Silene latifolia and S. dioica was accompanied by multiple chromosomal rearrangements. Front. Plant Sci. 2020, 11, 205. [Google Scholar] [CrossRef]
  103. Montenegro, C.; do Vale Martins, L.; Bustamante, F.O.; Brasileiro-Vidal, A.C.; Pedrosa-Harand, A. Comparative cytogenomics reveals genome reshuffling and centromere repositioning in the legume tribe Phaseoleae. Chromosome Res. 2022, 30, 477–492. [Google Scholar] [CrossRef] [PubMed]
  104. De Oliveira Bustamante, F.; do Nascimento, T.H.; Montenegro, C.; Dias, S.; do Vale Martins, L.; Braz, G.T.; Benko-Iseppon, A.M.; Jiang, J.; Pedrosa-Harand, A.; Brasileiro-Vidal, A.C. Oligo-FISH barcode in beans: A new chromosome identification system. Theor. Appl. Genet. 2021, 134, 3675–3686. [Google Scholar] [CrossRef] [PubMed]
  105. Kobrlova, L.; Cizkova, J.; Zoulova, V.; Vejvodova, K.; Hribova, E. First insight into the genomes of the Pulmonaria officinalis group (Boraginaceae) provided by repeatome analysis and comparative karyotyping. BMC Plant Biol. 2024, 24, 859. [Google Scholar] [CrossRef] [PubMed]
  106. Mahelka, V.; Krak, K.; Fehrer, J.; Caklova, P.; Nagy Nejedla, M.; Cegan, R.; Kopecky, D.; Safar, J. A Panicum-derived chromosomal segment captured by Hordeum a few million years ago preserves a set of stress-related genes. Plant J. 2021, 105, 1141–1164. [Google Scholar] [CrossRef]
  107. Yucel, G.; Betekhtin, A.; Cabi, E.; Tuna, M.; Hasterok, R.; Kolano, B. The chromosome number and rDNA loci evolution in Onobrychis (Fabaceae). Int. J. Mol. Sci. 2022, 23, 11033. [Google Scholar] [CrossRef]
  108. Kumar, S.; Kaur, S.; Seem, K.; Kumar, S.; Mohapatra, T. Understanding 3D genome organization and its effect on transcriptional gene regulation under environmental stress in plant: A chromatin perspective. Front. Cell Dev. Biol. 2021, 9, 774719. [Google Scholar] [CrossRef] [PubMed]
  109. Dogan, E.S.; Liu, C. Three-dimensional chromatin packing and positioning of plant genomes. Nat. Plants 2018, 4, 521–529. [Google Scholar] [CrossRef] [PubMed]
  110. Pei, L.; Li, G.; Lindsey, K.; Zhang, X.; Wang, M. Plant 3D genomics: The exploration and application of chromatin organization. New Phytol. 2021, 230, 1772–1786. [Google Scholar] [CrossRef] [PubMed]
  111. Ouyang, W.; Xiong, D.; Li, G.; Li, X. Unraveling the 3D genome architecture in plants: Present and future. Mol. Plant 2020, 13, 1676–1693. [Google Scholar] [CrossRef] [PubMed]
  112. Johann To Berens, P.; Schivre, G.; Theune, M.; Peter, J.; Sall, S.O.; Mutterer, J.; Barneche, F.; Bourbousse, C.; Molinier, J. Advanced image analysis methods for automated segmentation of subnuclear chromatin domains. Epigenomes 2022, 6, 34. [Google Scholar] [CrossRef]
  113. Ning, Y.; Shang, D.; Xin, H.; Ni, R.; Wang, Z.; Zhen, Y.; Liu, G.; Xi, M. Establishing of 3D-FISH on frozen section and its applying in chromosome territories analysis in Populus trichocarpa. Plant Cell Rep. 2024, 43, 255. [Google Scholar] [CrossRef]
  114. Hu, B.; Wang, N.; Bi, X.; Karaaslan, E.S.; Weber, A.L.; Zhu, W.; Berendzen, K.W.; Liu, C. Plant lamin-like proteins mediate chromatin tethering at the nuclear periphery. Genome Biol. 2019, 20, 87. [Google Scholar] [CrossRef]
  115. Schubert, V.; Lermontova, I.; Schubert, I. The Arabidopsis CAP-D proteins are required for correct chromatin organisation, growth and fertility. Chromosoma 2013, 122, 517–533. [Google Scholar] [CrossRef]
  116. Shan, W.; Kubova, M.; Mandakova, T.; Lysak, M.A. Nuclear organization in crucifer genomes: Nucleolus-associated telomere clustering is not a universal interphase configuration in Brassicaceae. Plant J. 2021, 108, 528–540. [Google Scholar] [CrossRef]
  117. Nemeckova, A.; Kolackova, V.; Vrana, J.; Dolezel, J.; Hribova, E. DNA replication and chromosome positioning throughout the interphase in three-dimensional space of plant nuclei. J. Exp. Bot. 2020, 71, 6262–6272. [Google Scholar] [CrossRef]
  118. Zuo, S.; Mandakova, T.; Kubova, M.; Lysak, M.A. Genomes, repeatomes and interphase chromosome organization in the meadowfoam family (Limnanthaceae, Brassicales). Plant J. 2022, 110, 1462–1475. [Google Scholar] [CrossRef] [PubMed]
  119. Dong, F.; Jiang, J. Non-Rabl patterns of centromere and telomere distribution in the interphase nuclei of plant cells. Chromosome Res. 1998, 6, 551–558. [Google Scholar] [CrossRef] [PubMed]
  120. Robaszkiewicz, E.; Idziak-Helmcke, D.; Tkacz, M.A.; Chrominski, K.; Hasterok, R. The arrangement of Brachypodium distachyon chromosomes in interphase nuclei. J. Exp. Bot. 2016, 67, 5571–5583. [Google Scholar] [CrossRef]
  121. Nowicka, A.; Ferkova, L.; Said, M.; Kovacik, M.; Zwyrtkova, J.; Baroux, C.; Pecinka, A. Non-Rabl chromosome organization in endoreduplicated nuclei of barley embryo and endosperm tissues. J. Exp. Bot. 2023, 74, 2527–2541. [Google Scholar] [CrossRef]
  122. Kubalova, I.; Camara, A.S.; Capal, P.; Beseda, T.; Rouillard, J.M.; Krause, G.M.; Holusova, K.; Toegelova, H.; Himmelbach, A.; Stein, N.; et al. Helical coiling of metaphase chromatids. Nucleic Acids Res. 2023, 51, 2641–2654. [Google Scholar] [CrossRef]
  123. Hayashida, M.; Sartsanga, C.; Phengchat, R.; Malac, M.; Harada, K.; Akashi, T.; Fukui, K.; Ohmido, N. Higher-order structure of barley chromosomes observed by electron tomography. Micron 2022, 160, 103328. [Google Scholar] [CrossRef]
  124. Siregar, A.Y.; Sartsanga, C.; Arifudin, F.S.; Phengchat, R.; Salamah, A.; Ohmido, N.; Fukui, K.; Dwiranti, A. Calcium ion significance on the maintenance of barley (Hordeum vulgare) chromosome compaction. Micron 2021, 145, 103046. [Google Scholar] [CrossRef]
  125. Kubalova, I.; Nemeckova, A.; Weisshart, K.; Hribova, E.; Schubert, V. Comparing super-resolution microscopy techniques to analyze chromosomes. Int. J. Mol. Sci. 2021, 22, 1903. [Google Scholar] [CrossRef]
  126. Kubalova, I.; Weisshart, K.; Houben, A.; Schubert, V. Super-resolution microscopy reveals the number and distribution of topoisomerase IIalpha and CENH3 molecules within barley metaphase chromosomes. Chromosoma 2023, 132, 19–29. [Google Scholar] [CrossRef] [PubMed]
  127. Sartsanga, C.; Phengchat, R.; Wako, T.; Fukui, K.; Ohmido, N. Localization and quantitative distribution of a chromatin structural protein Topoisomerase II on plant chromosome using HVTEM and UHVTEM. Micron 2024, 179, 103596. [Google Scholar] [CrossRef] [PubMed]
  128. Ohmido, N.; Sartsanga, C.; Dwiranti, A. Structure and compaction of plant chromosomes: Studies using advanced electron microscopy. Micron 2025, 196–197, 103860. [Google Scholar] [CrossRef]
  129. Kaduchova, K.; Marchetti, C.; Ovecka, M.; Galuszka, P.; Bergougnoux, V.; Samaj, J.; Pecinka, A. Spatial organization and dynamics of chromosomes and microtubules during barley mitosis. Plant J. 2023, 115, 602–613. [Google Scholar] [CrossRef]
  130. Kuo, Y.T.; Schubert, V.; Marques, A.; Schubert, I.; Houben, A. Centromere diversity: How different repeat-based holocentromeres may have evolved. Bioessays 2024, 46, e2400013. [Google Scholar] [CrossRef] [PubMed]
  131. Naish, M.; Henderson, I.R. The structure, function, and evolution of plant centromeres. Genome Res. 2024, 34, 161–178. [Google Scholar] [CrossRef] [PubMed]
  132. Schubert, V.; Neumann, P.; Marques, A.; Heckmann, S.; Macas, J.; Pedrosa-Harand, A.; Schubert, I.; Jang, T.S.; Houben, A. Super-resolution microscopy reveals diversity of plant centromere architecture. Int. J. Mol. Sci. 2020, 21, 3488. [Google Scholar] [CrossRef] [PubMed]
  133. Plackova, K.; Zedek, F.; Schubert, V.; Houben, A.; Bures, P. Kinetochore size scales with chromosome size in bimodal karyotypes of Agavoideae. Ann. Bot. 2022, 130, 77–84. [Google Scholar] [CrossRef] [PubMed]
  134. Kuo, Y.T.; Camara, A.S.; Schubert, V.; Neumann, P.; Macas, J.; Melzer, M.; Chen, J.; Fuchs, J.; Abel, S.; Klocke, E.; et al. Holocentromeres can consist of merely a few megabase-sized satellite arrays. Nat. Commun. 2023, 14, 3502. [Google Scholar] [CrossRef] [PubMed]
  135. Neumann, P.; Oliveira, L.; Jang, T.S.; Novak, P.; Koblizkova, A.; Schubert, V.; Houben, A.; Macas, J. Disruption of the standard kinetochore in holocentric Cuscuta species. Proc. Natl. Acad. Sci. USA 2023, 120, e2300877120. [Google Scholar] [CrossRef]
  136. Ding, W.; Zhu, Y.; Han, J.; Zhang, H.; Xu, Z.; Khurshid, H.; Liu, F.; Hasterok, R.; Shen, X.; Wang, K. Characterization of centromeric DNA of Gossypium anomalum reveals sequence-independent enrichment dynamics of centromeric repeats. Chromosome Res. 2023, 31, 12. [Google Scholar] [CrossRef]
  137. Oliveira, L.; Neumann, P.; Mata-Sucre, Y.; Kuo, Y.T.; Marques, A.; Schubert, V.; Macas, J. KNL1 and NDC80 represent new universal markers for the detection of functional centromeres in plants. Chromosome Res. 2024, 32, 3. [Google Scholar] [CrossRef]
  138. Melters, D.P.; Paliulis, L.V.; Korf, I.F.; Chan, S.W. Holocentric chromosomes: Convergent evolution, meiotic adaptations, and genomic analysis. Chromosome Res. 2012, 20, 579–593. [Google Scholar] [CrossRef]
  139. Mata-Sucre, Y.; Kratka, M.; Oliveira, L.; Neumann, P.; Macas, J.; Schubert, V.; Huettel, B.; Kejnovsky, E.; Houben, A.; Pedrosa-Harand, A.; et al. Repeat-based holocentromeres of the woodrush Luzula sylvatica reveal insights into the evolutionary transition to holocentricity. Nat. Commun. 2024, 15, 9565. [Google Scholar] [CrossRef] [PubMed]
  140. Alisawi, O.; Richert-Poggeler, K.R.; Heslop-Harrison, J.S.P.; Schwarzacher, T. The nature and organization of satellite DNAs in Petunia hybrida, related, and ancestral genomes. Front. Plant Sci. 2023, 14, 1232588. [Google Scholar] [CrossRef] [PubMed]
  141. Vozarova, R.; Herklotz, V.; Kovarik, A.; Tynkevich, Y.O.; Volkov, R.A.; Ritz, C.M.; Lunerova, J. Ancient origin of two 5S rDNA families dominating in the genus Rosa and their behavior in the Canina-type meiosis. Front. Plant Sci. 2021, 12, 643548. [Google Scholar] [CrossRef] [PubMed]
  142. Luo, X.; He, Z.; Liu, J.; Wu, H.; Gong, X. FISH mapping of telomeric and non-telomeric (AG3T3)3 reveal the chromosome numbers and chromosome rearrangements of 41 woody plants. Genes 2022, 13, 1239. [Google Scholar] [CrossRef]
  143. Liu, Y.; Liaw, Y.M.; Teo, C.H.; Capal, P.; Wada, N.; Fukui, K.; Dolezel, J.; Ohmido, N. Molecular organization of recombinant human-Arabidopsis chromosomes in hybrid cell lines. Sci. Rep. 2021, 11, 7160. [Google Scholar] [CrossRef] [PubMed]
  144. Chen, J.; Bartos, J.; Boudichevskaia, A.; Voigt, A.; Rabanus-Wallace, M.T.; Dreissig, S.; Tulpova, Z.; Simkova, H.; Macas, J.; Kim, G.; et al. The genetic mechanism of B chromosome drive in rye illuminated by chromosome-scale assembly. Nat. Commun. 2024, 15, 9686. [Google Scholar] [CrossRef]
  145. Karafiatova, M.; Bojdova, T.; Stejskalova, M.; Harnadkova, N.; Kumar, V.; Houben, A.; Chen, J.; Dolezalova, A.; Honys, D.; Bartos, J. Unravelling the unusual: Chromosome elimination, nondisjunction and extra pollen mitosis characterize the B chromosome in wild sorghum. New Phytol. 2024, 243, 1840–1854. [Google Scholar] [CrossRef]
  146. Shams, I.; Raskina, O. Supernumerary B chromosomes and plant genome changes: A snapshot of wild populations of Aegilops speltoides Tausch (Poaceae, Triticeae). Int. J. Mol. Sci. 2020, 21, 3768. [Google Scholar] [CrossRef]
  147. Chen, J.; Birchler, J.A.; Houben, A. The non-Mendelian behavior of plant B chromosomes. Chromosome Res. 2022, 30, 229–239. [Google Scholar] [CrossRef]
  148. Liu, Q.; Liu, Y.; Yi, C.; Gao, Z.; Zhang, Z.; Zhu, C.; Birchler, J.A.; Han, F. Genome assembly of the maize B chromosome provides insight into its epigenetic characteristics and effects on the host genome. Genome Biol. 2025, 26, 47. [Google Scholar] [CrossRef]
  149. Soltis, P.S.; Marchant, D.B.; Van de Peer, Y.; Soltis, D.E. Polyploidy and genome evolution in plants. Curr. Opin. Genet. Dev. 2015, 35, 119–125. [Google Scholar] [CrossRef] [PubMed]
  150. Salse, J. Deciphering the evolutionary interplay between subgenomes following polyploidy: A paleogenomics approach in grasses. Am. J. Bot. 2016, 103, 1167–1174. [Google Scholar] [CrossRef] [PubMed]
  151. Heslop-Harrison, J.S.P.; Schwarzacher, T.; Liu, Q. Polyploidy: Its consequences and enabling role in plant diversification and evolution. Ann. Bot. 2023, 131, 1–10. [Google Scholar] [CrossRef] [PubMed]
  152. Chester, M.; Leitch, A.R.; Soltis, P.S.; Soltis, D.E. Review of the application of modern cytogenetic methods (FISH/GISH) to the study of reticulation (polyploidy/hybridisation). Genes 2010, 1, 166–192. [Google Scholar] [CrossRef]
  153. Han, J.; Zhou, B.; Shan, W.; Yu, L.; Wu, W.; Wang, K. A and D genomes spatial separation at somatic metaphase in tetraploid cotton: Evidence for genomic disposition in a polyploid plant. Plant J. 2015, 84, 1167–1177. [Google Scholar] [CrossRef] [PubMed]
  154. Hasterok, R.; Draper, J.; Jenkins, G. Laying the cytotaxonomic foundations of a new model grass, Brachypodium distachyon (L.) Beauv. Chromosome Res. 2004, 12, 397–403. [Google Scholar] [CrossRef] [PubMed]
  155. Chrtek, J.; Mraz, P.; Belyayev, A.; Pastova, L.; Mrazova, V.; Caklova, P.; Josefiova, J.; Zagorski, D.; Hartmann, M.; Jandova, M.; et al. Evolutionary history and genetic diversity of apomictic allopolyploids in Hieracium s.str.: Morphological versus genomic features. Am. J. Bot. 2020, 107, 66–90. [Google Scholar] [CrossRef] [PubMed]
  156. Mandáková, T.; Lysak, M.A. The identification of the missing maternal genome of the allohexaploid camelina (Camelina sativa). Plant J. 2022, 112, 622–629. [Google Scholar] [CrossRef]
  157. Köhler, M.; Oakley, L.J.; Font, F.; Peñas, M.L.L.; Majure, L.C. On the continuum of evolution: A putative new hybrid speciation event in Opuntia (Cactaceae) between a native and an introduced species in southern South America. Syst. Biodivers. 2021, 19, 1026–1039. [Google Scholar] [CrossRef]
  158. Shimomai, H.; Taichi, N.; Katsuhara, K.R.; Kato, S.; Ushimaru, A.; Ohmido, N. Allopolyploidy enhances survival advantages for urban environments in the native plant genus Commelina. Ann. Bot. 2024, 134, 1055–1066. [Google Scholar] [CrossRef]
  159. Amosova, A.V.; Zemtsova, L.V.; Yurkevich, O.Y.; Zhidkova, E.N.; Ksiazczyk, T.; Shostak, N.G.; Muravlev, A.A.; Artemyeva, A.M.; Samatadze, T.E.; Zoshchuk, S.A.; et al. Genomic changes in generations of synthetic rapeseed-like allopolyploid grown under selection. Euphytica 2017, 213, 217. [Google Scholar] [CrossRef]
  160. Kruppa, K.; Turkosi, E.; Mayer, M.; Toth, V.; Vida, G.; Szakacs, E.; Molnar-Lang, M. McGISH identification and phenotypic description of leaf rust and yellow rust resistant partial amphiploids originating from a wheat x Thinopyrum synthetic hybrid cross. J. Appl. Genet. 2016, 57, 427–437. [Google Scholar] [CrossRef]
  161. Song, Z.P.; Zuo, Y.Y.; Li, W.J.; Dai, S.F.; Liu, G.; Pu, Z.J.; Yan, Z.H. Chromosome stability of synthetic Triticum turgidum–Aegilops umbellulata hybrids. BMC Plant Biol. 2024, 24, 391. [Google Scholar] [CrossRef] [PubMed]
  162. Meng, Z.; Shi, S.D.; Shen, H.T.; Xie, Q.L.; Li, H.B. Haplotype-specific chromosome painting provides insights into the chromosomal characteristics in self-duplicating autotetraploid sugarcane. Ind. Crops Prod. 2023, 202, 117085. [Google Scholar] [CrossRef]
  163. Zhang, J.; Fan, C.; Liu, Y.; Shi, Q.; Sun, Y.; Huang, Y.; Yuan, J.; Han, F. Cytological analysis of the diploid-like inheritance of newly synthesized allotetraploid wheat. Chromosome Res. 2023, 32, 1. [Google Scholar] [CrossRef] [PubMed]
  164. Borowska-Zuchowska, N.; Robaszkiewicz, E.; Mykhailyk, S.; Wartini, J.; Pinski, A.; Kovarik, A.; Hasterok, R. To be or not to be expressed: The first evidence of a nucleolar dominance tissue-specificity in Brachypodium hybridum. Front. Plant Sci. 2021, 12, 768347. [Google Scholar] [CrossRef]
  165. Borowska-Zuchowska, N.; Kovarik, A.; Robaszkiewicz, E.; Tuna, M.; Tuna, G.S.; Gordon, S.; Vogel, J.P.; Hasterok, R. The fate of 35S rRNA genes in the allotetraploid grass Brachypodium hybridum. Plant J. 2020, 103, 1810–1825. [Google Scholar] [CrossRef]
  166. Trunova, D.; Borowska-Zuchowska, N.; Mykhailyk, S.; Xia, K.; Zhu, Y.; Sancho, R.; Rojek-Jelonek, M.; Garcia, S.; Wang, K.; Catalan, P.; et al. Does time matter? Intraspecific diversity of ribosomal RNA genes in lineages of the allopolyploid model grass Brachypodium hybridum with different evolutionary ages. BMC Plant Biol. 2024, 24, 981. [Google Scholar] [CrossRef]
  167. Mahelka, V.; Kopecky, D.; Majka, J.; Krak, K. Uniparental expression of ribosomal RNA in xFestulolium grasses: A link between the genome and nucleolar dominance. Front. Plant Sci. 2023, 14, 1276252. [Google Scholar] [CrossRef]
  168. Matyasek, R.; Kalfusova, R.; Kuderova, A.; Rehurkova, K.; Sochorova, J.; Kovarik, A. Transcriptional silencing of 35S rDNA in Tragopogon porrifolius correlates with cytosine methylation in sequence-specific manner. Int. J. Mol. Sci. 2024, 25, 7540. [Google Scholar] [CrossRef]
  169. Borowska-Zuchowska, N.; Mykhailyk, S.; Robaszkiewicz, E.; Matysiak, N.; Mielanczyk, L.; Wojnicz, R.; Kovarik, A.; Hasterok, R. Switch them off or not: Selective rRNA gene repression in grasses. Trends Plant Sci. 2023, 28, 661–672. [Google Scholar] [CrossRef] [PubMed]
  170. Idziak, D.; Hasterok, R. Cytogenetic evidence of nucleolar dominance in allotetraploid species of Brachypodium. Genome 2008, 51, 387–391. [Google Scholar] [CrossRef] [PubMed]
  171. Mandakova, T.; Krumpolcova, A.; Matyasek, R.; Volkov, R.; Lysak, M.A.; Kovarik, A. Uniparental silencing of 5S rRNA genes in plant allopolyploids—Insights from Cardamine (Brassicaceae). Plant J. 2024, 119, 1313–1326. [Google Scholar] [CrossRef] [PubMed]
  172. Gramazio, P.; Prohens, J.; Toppino, L.; Plazas, M. Editorial: Introgression breeding in cultivated plants. Front. Plant Sci. 2021, 12, 764533. [Google Scholar] [CrossRef]
  173. Kashyap, A.; Garg, P.; Tanwar, K.; Sharma, J.; Gupta, N.C.; Ha, P.T.T.; Bhattacharya, R.C.; Mason, A.S.; Rao, M. Strategies for utilization of crop wild relatives in plant breeding programs. Theor. Appl. Genet. 2022, 135, 4151–4167. [Google Scholar] [CrossRef] [PubMed]
  174. Kwiatek, M.T.; Nawracala, J. Chromosome manipulations for progress of triticale (×Triticosecale) breeding. Plant Breed. 2018, 137, 823–831. [Google Scholar] [CrossRef]
  175. Rustgi, S. Plant cytogenetics blurring disciplinary boundaries to sustain global food security. Nucleus 2023, 66, 239–243. [Google Scholar] [CrossRef]
  176. Kumar, R.; Das, S.P.; Choudhury, B.U.; Kumar, A.; Prakash, N.R.; Verma, R.; Chakraborti, M.; Devi, A.G.; Bhattacharjee, B.; Das, R.; et al. Advances in genomic tools for plant breeding: Harnessing DNA molecular markers, genomic selection, and genome editing. Biol. Res. 2024, 57, 80. [Google Scholar] [CrossRef]
  177. Salina, E.A.; Adonina, I.G. Cytogenetics in the dtudy of chromosomal rearrangement during wheat evolution and breeding. In Cytogenetics—Past, Present and Further Perspectives; Larramendy, M.L., Soloneski, S., Eds.; IntechOpen: Rijeka, Croatia, 2018. [Google Scholar]
  178. Kwiatek, M.T.; Belter, J.; Ulaszewski, W.; Skowronska, R.; Noweiska, A.; Wisniewska, H. Molecular identification of triticale introgression lines carrying leaf rust resistance genes transferred from Aegilops kotschyi Boiss. and Ae. tauschii Coss. J. Appl. Genet. 2021, 62, 431–439. [Google Scholar] [CrossRef]
  179. Li, Z.; Jiang, Q.; Fan, T.; Zhao, L.; Ren, Z.; Tan, F.; Luo, P.; Ren, T. Molecular cytogenetic and physiological characterization of a novel wheat-rye T1RS.1BL translocation line from Secale cereal L. Weining with resistance to stripe rust and functional “Stay Green” trait. Int. J. Mol. Sci. 2022, 23, 4626. [Google Scholar] [CrossRef] [PubMed]
  180. Li, Z.; Ren, Z.; Tan, F.; Luo, P.; Ren, T. Molecular cytogenetic characterization of novel 1RS.1BL translocation and complex chromosome translocation lines with stripe rust resistance. Int. J. Mol. Sci. 2022, 23, 2731. [Google Scholar] [CrossRef] [PubMed]
  181. Jamwal, N.S.; Badiyal, A.; Chaudhary, H.K.; Schwarzacher, T.; Heslop-Harrison, J.S. Molecular cytogenetic analysis of newly developed progenies from triticale x wheat crosses for yield and stress tolerance. Cereal Res. Commun. 2024, 52, 859–865. [Google Scholar] [CrossRef]
  182. Li, Q.F.; Lu, Y.Q.; Pan, C.L.; Wang, Z.J.; Liu, F.L.; Zhang, J.P.; Yang, X.M.; Li, X.Q.; Liu, W.H.; Li, L.H. Characterization, identification and evaluation of a novel wheat-Agropyron cristatum (L.) Gaertn. disomic addition line II-30-5. Genet. Resour. Crop Evol. 2020, 67, 2213–2223. [Google Scholar] [CrossRef]
  183. Song, L.; Zhao, H.; Zhang, Z.; Zhang, S.; Liu, J.; Zhang, W.; Zhang, N.; Ji, J.; Li, L.; Li, J. Molecular cytogenetic identification of wheat-Aegilops biuncialis 5Mb disomic addition line with tenacious and black glumes. Int. J. Mol. Sci. 2020, 21, 4053. [Google Scholar] [CrossRef] [PubMed]
  184. Wang, Y.; Cheng, X.; Yang, X.; Wang, C.; Zhang, H.; Deng, P.; Liu, X.; Chen, C.; Ji, W.; Wang, Y. Molecular cytogenetics for a wheat-Aegilops geniculata 3Mg alien addition line with resistance to stripe rust and powdery mildew. BMC Plant Biol. 2021, 21, 575. [Google Scholar] [CrossRef]
  185. Wang, Y.; Fan, J.; Xiao, Y.; Feng, X.; Zhang, H.; Chen, C.; Ji, W.; Wang, Y. Genetic analysis of resistance to powdery mildew on 7Mg chromosome of wheat-Aegilops geniculata, development and utilization of specific molecular markers. BMC Plant Biol. 2022, 22, 564. [Google Scholar] [CrossRef]
  186. Motsnyi, I.I.; Halaiev, O.V.; Alieksieieva, T.G.; Chebotar, G.O.; Chebotar, S.V.; Betekhtin, A.; Hasterok, R.; Armoniene, R.; Rahmatov, M. Cytogenetic and molecular identification of novel wheat-Elymus sibiricus addition lines with resistance to leaf rust and the presence of leaf pubescence trait. Front. Plant Sci. 2024, 15, 1482211. [Google Scholar] [CrossRef]
  187. Feng, X.; Du, X.; Wang, S.; Deng, P.; Wang, Y.; Shang, L.; Tian, Z.; Wang, C.; Chen, C.; Zhao, J.; et al. Identification and DNA marker development for a wheat-Leymus mollis 2Ns (2D) disomic chromosome substitution. Int. J. Mol. Sci. 2022, 23, 2676. [Google Scholar] [CrossRef]
  188. Hou, C.; Han, J.; Zhang, L.; Geng, Q.; Zhao, L.; Liu, S.; Yang, Q.; Chen, X.; Wu, J. Identification of resistance to Fusarium head blight and molecular cytogenetics of interspecific derivatives between wheat and Psathyrostachys huashanica. Breed. Sci. 2022, 72, 213–221. [Google Scholar] [CrossRef]
  189. Pang, J.; Huang, C.; Wang, Y.; Wen, X.; Deng, P.; Li, T.; Wang, C.; Liu, X.; Chen, C.; Zhao, J.; et al. Molecular cytological analysis and specific marker development in wheat-Psathyrostachys huashanica Keng 3Ns additional line with elongated glume. Int. J. Mol. Sci. 2023, 24, 6726. [Google Scholar] [CrossRef]
  190. Wang, H.; Cheng, S.; Shi, Y.; Zhang, S.; Yan, W.; Song, W.; Yang, X.; Song, Q.; Jang, B.; Qi, X.; et al. Molecular cytogenetic characterization and fusarium head blight resistance of five wheat-Thinopyrum intermedium partial amphiploids. Mol. Cytogenet. 2021, 14, 15. [Google Scholar] [CrossRef] [PubMed]
  191. Wang, S.; Wang, C.; Feng, X.; Zhao, J.; Deng, P.; Wang, Y.; Zhang, H.; Liu, X.; Li, T.; Chen, C.; et al. Molecular cytogenetics and development of St-chromosome-specific molecular markers of novel stripe rust resistant wheat-Thinopyrum intermedium and wheat-Thinopyrum ponticum substitution lines. BMC Plant Biol. 2022, 22, 111. [Google Scholar] [CrossRef] [PubMed]
  192. Agrawal, N.; Gupta, M.; Atri, C.; Akhatar, J.; Kumar, S.; Heslop-Harrison, P.; Banga, S.S. Anchoring alien chromosome segment substitutions bearing gene(s) for resistance to mustard aphid in Brassica juncea-B. fruticulosa introgression lines and their possible disruption through gamma irradiation. Theor. Appl. Genet. 2021, 134, 3209–3224. [Google Scholar] [CrossRef] [PubMed]
  193. Ankrah, N.A.; El-nagish, A.; Breitenbach, S.; Tetteh, A.Y.; Heitkam, T. Comparative cytogenetics of kenaf (Hibiscus cannabinus L.) breeding lines reveal chromosomal variability and instability. Genet. Resour. Crop Evol. 2024, 72, 3423–3436. [Google Scholar] [CrossRef]
  194. Savas-Tuna, G.; Yücel, G.; Asçiogul, T.K.; Ates, D.; Esiyok, D.; Tanyolaç, M.B.; Tuna, M. Molecular cytogenetic characterization of common bean (Phaseolus vulgaris L.) accessions. Turk. J. Agric. For. 2020, 44, 612–630. [Google Scholar] [CrossRef]
  195. Kwiatek, M.T.; Drozdowska, Z.; Kurasiak-Popowska, D.; Noweiska, A.; Nawracala, J. Cytomolecular analysis of mutants, breeding lines, and varieties of camelina (Camelina sativa L. Crantz). J. Appl. Genet. 2021, 62, 199–205. [Google Scholar] [CrossRef]
  196. Park, H.W.; Sevilleno, S.S.; Ha, M.; Cabahug-Braza, R.A.; Yi, J.H.; Lim, K.B.; Cho, W.; Hwang, Y.J. The application of fluorescence in situ hybridization in the prescreening of Veronica hybrids. Plants 2024, 13, 1264. [Google Scholar] [CrossRef]
  197. Tomiczak, K. Molecular and cytogenetic description of somatic hybrids between Gentiana cruciata L. and G. tibetica King. J. Appl. Genet. 2020, 61, 13–24. [Google Scholar] [CrossRef]
  198. Zhou, M.; Yong, X.; Zhu, J.; Xu, Q.; Liu, X.; Zhang, L.; Mou, L.; Zeng, L.; Wu, M.; Jiang, B.; et al. Chromosomal analysis of progenies between Lilium intersectional hybrids and wild species using ND-FISH and GISH. Front. Plant Sci. 2024, 15, 1461798. [Google Scholar] [CrossRef]
  199. Xu, Z.; Chen, J.; Meng, S.; Xu, P.; Zhai, C.; Huang, F.; Guo, Q.; Zhao, L.; Quan, Y.; Shangguan, Y.; et al. Genome sequence of Gossypium anomalum facilitates interspecific introgression breeding. Plant Commun. 2022, 3, 100350. [Google Scholar] [CrossRef] [PubMed]
  200. Liao, X.; Zhu, W.; Zhou, J.; Li, H.; Xu, X.; Zhang, B.; Gao, X. Repetitive DNA sequence detection and its role in the human genome. Commun. Biol. 2023, 6, 954. [Google Scholar] [CrossRef] [PubMed]
  201. Mehrotra, S.; Goyal, V. Repetitive sequences in plant nuclear DNA: Types, distribution, evolution and function. Genom. Proteom. Bioinform. 2014, 12, 164–171. [Google Scholar] [CrossRef]
  202. Leitch, I.J.; Johnston, E.; Pellicer, J.; Hidalgo, O.; Bennett, M.D. Plant DNA C-Values Database (Release 7.1, Apr 2019). Available online: https://cvalues.science.kew.org/ (accessed on 15 July 2025).
  203. Belyayev, A.; Josefiova, J.; Jandova, M.; Kalendar, R.; Krak, K.; Mandak, B. Natural history of a aatellite DNA family: From the ancestral genome component to species-specific sequences, concerted and non-concerted evolution. Int. J. Mol. Sci. 2019, 20, 1201. [Google Scholar] [CrossRef]
  204. Novak, P.; Neumann, P.; Pech, J.; Steinhaisl, J.; Macas, J. RepeatExplorer: A Galaxy-based web server for genome-wide characterization of eukaryotic repetitive elements from next-generation sequence reads. Bioinformatics 2013, 29, 792–793. [Google Scholar] [CrossRef]
  205. Novak, P.; Neumann, P.; Macas, J. Global analysis of repetitive DNA from unassembled sequence reads using RepeatExplorer2. Nat. Protoc. 2020, 15, 3745–3776. [Google Scholar] [CrossRef] [PubMed]
  206. Novak, P.; Avila Robledillo, L.; Koblizkova, A.; Vrbova, I.; Neumann, P.; Macas, J. TAREAN: A computational tool for identification and characterization of satellite DNA from unassembled short reads. Nucleic Acids Res. 2017, 45, e111. [Google Scholar] [CrossRef] [PubMed]
  207. Wang, Z.; Rouard, M.; Biswas, M.K.; Droc, G.; Cui, D.; Roux, N.; Baurens, F.C.; Ge, X.J.; Schwarzacher, T.; Heslop-Harrison, P.J.S.; et al. A chromosome-level reference genome of Ensete glaucum gives insight into diversity and chromosomal and repetitive sequence evolution in the Musaceae. Gigascience 2022, 11, giac027. [Google Scholar] [CrossRef] [PubMed]
  208. Dias, Y.; Mata-Sucre, Y.; Thangavel, G.; Costa, L.; Baez, M.; Houben, A.; Marques, A.; Pedrosa-Harand, A. How diverse a monocentric chromosome can be? Repeatome and centromeric organization of Juncus effusus (Juncaceae). Plant J. 2024, 118, 1832–1847. [Google Scholar] [CrossRef]
  209. Rathore, P.; Schwarzacher, T.; Heslop-Harrison, J.S.; Bhat, V.; Tomaszewska, P. The repetitive DNA sequence landscape and DNA methylation in chromosomes of an apomictic tropical forage grass, Cenchrus ciliaris. Front. Plant Sci. 2022, 13, 952968. [Google Scholar] [CrossRef]
  210. Ribeiro, T.; Vasconcelos, E.; Dos Santos, K.G.B.; Vaio, M.; Brasileiro-Vidal, A.C.; Pedrosa-Harand, A. Diversity of repetitive sequences within compact genomes of Phaseolus L. beans and allied genera Cajanus L. and Vigna Savi. Chromosome Res. 2020, 28, 139–153. [Google Scholar] [CrossRef]
  211. Ibiapino, A.; Baez, M.; Garcia, M.A.; Costea, M.; Stefanovic, S.; Pedrosa-Harand, A. Karyotype asymmetry in Cuscuta L. subgenus Pachystigma reflects its repeat DNA composition. Chromosome Res. 2022, 30, 91–107. [Google Scholar] [CrossRef] [PubMed]
  212. Ishiguro, S.; Taniguchi, S.; Schmidt, N.; Jost, M.; Wanke, S.; Heitkam, T.; Ohmido, N. Repeatome landscapes and cytogenetics of hortensias provide a framework to trace Hydrangea evolution and domestication. Ann. Bot. 2025, 135, 549–564. [Google Scholar] [CrossRef] [PubMed]
  213. Mata-Sucre, Y.; Sader, M.; Van-Lume, B.; Gagnon, E.; Pedrosa-Harand, A.; Leitch, I.J.; Lewis, G.P.; Souza, G. How diverse is heterochromatin in the Caesalpinia group? Cytogenomic characterization of Erythrostemon hughesii Gagnon & G.P. Lewis (Leguminosae: Caesalpinioideae). Planta 2020, 252, 49. [Google Scholar] [CrossRef] [PubMed]
  214. Ferraz, M.E.; Ribeiro, T.; Sader, M.; Nascimento, T.; Pedrosa-Harand, A. Comparative analysis of repetitive DNA in dysploid and non-dysploid Phaseolus beans. Chromosome Res. 2023, 31, 30. [Google Scholar] [CrossRef] [PubMed]
  215. Kuo, Y.T.; Ishii, T.; Fuchs, J.; Hsieh, W.H.; Houben, A.; Lin, Y.R. The evolutionary dynamics of repetitive DNA and its impact on the genome diversification in the genus Sorghum. Front. Plant Sci. 2021, 12, 729734. [Google Scholar] [CrossRef]
  216. Tomaszewska, P.; Vorontsova, M.S.; Renvoize, S.A.; Ficinski, S.Z.; Tohme, J.; Schwarzacher, T.; Castiblanco, V.; de Vega, J.J.; Mitchell, R.A.C.; Heslop-Harrison, J.S.P. Complex polyploid and hybrid species in an apomictic and sexual tropical forage grass group: Genomic composition and evolution in Urochloa (Brachiaria) species. Ann. Bot. 2023, 131, 87–108. [Google Scholar] [CrossRef]
  217. Tomaszewska, P.; Schwarzacher, T.; Heslop-Harrison, J.S.P. Oat chromosome and genome evolution defined by widespread terminal intergenomic translocations in polyploids. Front. Plant Sci. 2022, 13, 1026364. [Google Scholar] [CrossRef]
  218. Huang, Y.; Ding, W.; Zhang, M.; Han, J.; Jing, Y.; Yao, W.; Hasterok, R.; Wang, Z.; Wang, K. The formation and evolution of centromeric satellite repeats in Saccharum species. Plant J. 2021, 106, 616–629. [Google Scholar] [CrossRef]
  219. Xin, H.; Wang, Y.; Zhang, W.; Bao, Y.; Neumann, P.; Ning, Y.; Zhang, T.; Wu, Y.; Jiang, N.; Jiang, J.; et al. Celine, a long interspersed nuclear element retrotransposon, colonizes in the centromeres of poplar chromosomes. Plant Physiol. 2024, 195, 2787–2798. [Google Scholar] [CrossRef]
  220. Galvez-Galvan, A.; Garrido-Ramos, M.A.; Prieto, P. Bread wheat satellitome: A complex scenario in a huge genome. Plant Mol. Biol. 2024, 114, 8. [Google Scholar] [CrossRef] [PubMed]
  221. Kwasniewska, J.; Bara, A.W. Plant cytogenetics in the micronuclei investigation-the past, current status, and perspectives. Int. J. Mol. Sci. 2022, 23, 1306. [Google Scholar] [CrossRef] [PubMed]
  222. Juchimiuk-Kwasniewska, J.; Brodziak, L.; Maluszynska, J. FISH in analysis of gamma ray-induced micronuclei formation in barley. J. Appl. Genet. 2011, 52, 23–29. [Google Scholar] [CrossRef] [PubMed]
  223. Kus, A.; Kwasniewska, J.; Hasterok, R. Brachypodium distachyon—A useful model in the qualification of mutagen-induced micronuclei using multicolor FISH. PLoS ONE 2017, 12, e0170618. [Google Scholar] [CrossRef]
  224. Tostes, N.V.; Ferreira, M.V.R.; Soares, F.A.F.; Silva, J.C.; Bhering, L.L.; Clarindo, W.R. DNA content, repeatome composition and origin of the Zea mays micronuclei. Sci. Rep. 2025, 15, 14997. [Google Scholar] [CrossRef]
  225. Kus, A.; Kwasniewska, J.; Szymanowska-Pulka, J.; Hasterok, R. Dissecting the chromosomal composition of mutagen-induced micronuclei in Brachypodium distachyon using multicolour FISH. Ann. Bot. 2018, 122, 1161–1171. [Google Scholar] [CrossRef]
  226. Kus, A.; Szymanowska-Pulka, J.; Kwasniewska, J.; Hasterok, R. Detecting Brachypodium distachyon chromosomes Bd4 and Bd5 in MH- and X-ray-induced micronuclei using mcFISH. Int. J. Mol. Sci. 2019, 20, 2848. [Google Scholar] [CrossRef]
  227. You, H.; Tang, D.; Liu, H.; Zhou, Y.; Li, Y.; Shen, Y.; Gong, Z.; Yu, H.; Gu, M.; Jiang, J.; et al. Chromosome ends initiate homologous chromosome pairing during rice meiosis. Plant Physiol. 2024, 195, 2617–2634. [Google Scholar] [CrossRef]
  228. Braz, G.T.; Yu, F.; Zhao, H.; Deng, Z.; Birchler, J.A.; Jiang, J. Preferential meiotic chromosome pairing among homologous chromosomes with cryptic sequence variation in tetraploid maize. New Phytol. 2021, 229, 3294–3302. [Google Scholar] [CrossRef]
  229. Riley, R.; Chapman, V.; Kimber, G. Genetic control of chromosome pairing in intergeneric hybrids with wheat. Nature 1959, 183, 1244–1246. [Google Scholar] [CrossRef]
  230. Higgins, E.E.; Howell, E.C.; Armstrong, S.J.; Parkin, I.A.P. A major quantitative trait locus on chromosome A9, BnaPh1, controls homoeologous recombination in Brassica napus. New Phytol. 2021, 229, 3281–3293. [Google Scholar] [CrossRef] [PubMed]
  231. Blackwell, A.R.; Dluzewska, J.; Szymanska-Lejman, M.; Desjardins, S.; Tock, A.J.; Kbiri, N.; Lambing, C.; Lawrence, E.J.; Bieluszewski, T.; Rowan, B.; et al. MSH2 shapes the meiotic crossover landscape in relation to interhomolog polymorphism in Arabidopsis. EMBO J. 2020, 39, e104858. [Google Scholar] [CrossRef] [PubMed]
  232. Zhu, L.; Fernandez-Jimenez, N.; Szymanska-Lejman, M.; Pele, A.; Underwood, C.J.; Serra, H.; Lambing, C.; Dluzewska, J.; Bieluszewski, T.; Pradillo, M.; et al. Natural variation identifies SNI1, the SMC5/6 component, as a modifier of meiotic crossover in Arabidopsis. Proc. Natl. Acad. Sci. USA 2021, 118, e2021970118. [Google Scholar] [CrossRef] [PubMed]
  233. Zhu, L.; Dluzewska, J.; Fernandez-Jimenez, N.; Ranjan, R.; Pele, A.; Dziegielewski, W.; Szymanska-Lejman, M.; Hus, K.; Gorna, J.; Pradillo, M.; et al. The kinase ATR controls meiotic crossover distribution at the genome scale in Arabidopsis. Plant Cell 2024, 37, koae292. [Google Scholar] [CrossRef] [PubMed]
  234. Charpentier, E.; Doudna, J.A. Biotechnology: Rewriting a genome. Nature 2013, 495, 50–51. [Google Scholar] [CrossRef]
  235. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef]
  236. Khosravi, S.; Ishii, T.; Dreissig, S.; Houben, A. Application and prospects of CRISPR/Cas9-based methods to trace defined genomic sequences in living and fixed plant cells. Chromosome Res. 2020, 28, 7–17. [Google Scholar] [CrossRef]
  237. Chen, B.; Gilbert, L.A.; Cimini, B.A.; Schnitzbauer, J.; Zhang, W.; Li, G.W.; Park, J.; Blackburn, E.H.; Weissman, J.S.; Qi, L.S.; et al. Dynamic imaging of genomic loci in living human cells by an optimized CRISPR/Cas system. Cell 2013, 155, 1479–1491. [Google Scholar] [CrossRef]
  238. Anton, T.; Bultmann, S.; Leonhardt, H.; Markaki, Y. Visualization of specific DNA sequences in living mouse embryonic stem cells with a programmable fluorescent CRISPR/Cas system. Nucleus 2014, 5, 163–172. [Google Scholar] [CrossRef]
  239. Dreissig, S.; Schiml, S.; Schindele, P.; Weiss, O.; Rutten, T.; Schubert, V.; Gladilin, E.; Mette, M.F.; Puchta, H.; Houben, A. Live-cell CRISPR imaging in plants reveals dynamic telomere movements. Plant J. 2017, 91, 565–573. [Google Scholar] [CrossRef]
  240. Fujimoto, S.; Matsunaga, S. Visualization of chromatin loci with transiently expressed CRISPR/Cas9 in plants. Cytologia 2017, 82, 559–562. [Google Scholar] [CrossRef]
  241. Ishii, T.; Schubert, V.; Khosravi, S.; Dreissig, S.; Metje-Sprink, J.; Sprink, T.; Fuchs, J.; Meister, A.; Houben, A. RNA-guided endonuclease—In situ labelling (RGEN-ISL): A fast CRISPR/Cas9-based method to label genomic sequences in various species. New Phytol. 2019, 222, 1652–1661. [Google Scholar] [CrossRef] [PubMed]
  242. Beying, N.; Schmidt, C.; Pacher, M.; Houben, A.; Puchta, H. CRISPR-Cas9-mediated induction of heritable chromosomal translocations in Arabidopsis. Nat. Plants 2020, 6, 638–645. [Google Scholar] [CrossRef] [PubMed]
  243. Schmidt, C.; Fransz, P.; Ronspies, M.; Dreissig, S.; Fuchs, J.; Heckmann, S.; Houben, A.; Puchta, H. Changing local recombination patterns in Arabidopsis by CRISPR/Cas mediated chromosome engineering. Nat. Commun. 2020, 11, 4418. [Google Scholar] [CrossRef] [PubMed]
  244. Ronspies, M.; Schmidt, C.; Schindele, P.; Lieberman-Lazarovich, M.; Houben, A.; Puchta, H. Massive crossover suppression by CRISPR-Cas-mediated plant chromosome engineering. Nat. Plants 2022, 8, 1153–1159. [Google Scholar] [CrossRef]
  245. Puchta, H.; Houben, A. Plant chromosome engineering—Past, present and future. New Phytol. 2024, 241, 541–552. [Google Scholar] [CrossRef]
  246. Helia, O.; Matusova, B.; Havlova, K.; Hyskova, A.; Lycka, M.; Beying, N.; Puchta, H.; Fajkus, J.; Fojtova, M. Chromosome engineering points to the cis-acting mechanism of chromosome arm-specific telomere length setting and robustness of plant phenotype, chromatin structure and gene expression. Plant J. 2025, 121, e70024. [Google Scholar] [CrossRef]
  247. Khosravi, S.; Hinrichs, R.; Ronspies, M.; Haghi, R.; Puchta, H.; Houben, A. Epigenetic state and gene expression remain stable after CRISPR/Cas-mediated chromosomal inversions. New Phytol. 2025, 245, 2527–2539. [Google Scholar] [CrossRef]
  248. Revolutionizing Cytogenetics with AI. Available online: https://www.hnl.com/resources/blog/revolutionizing-cytogenetics-with-ai (accessed on 15 July 2025).
  249. Rosenblum, L.S.; Holmes, J.; Taghiyev, A.F. The emergence of artificial intelligence-guided karyotyping: A review and reflection. Genes 2025, 16, 685. [Google Scholar] [CrossRef]
  250. Nakayama, S.; Fukui, K. Quantitative chromosome mapping of small plant chromosomes by improved imaging on CHIAS II. Genes Genet. Syst. 1997, 72, 35–40. [Google Scholar] [CrossRef]
  251. Kato, S.; Fukui, K. Condensation pattern (CP) analysis of plant chromosomes by an improved chromosome image analysing system, CHIAS III. Chromosome Res. 1998, 6, 473–479. [Google Scholar] [CrossRef]
  252. Kato, S.; Ohmido, N.; Hara, M.; Kataoka, R.; Fukui, K. Image analysis of small plant chromosomes by using an improved system, CHIAS IV. Chromosome Sci. 2009, 12, 43–50. [Google Scholar]
  253. Fukui, K. Standardization of karyotyping plant chromosomes by a newly developed chromosome image analyzing system (CHIAS). Theor. Appl. Genet. 1986, 72, 27–32. [Google Scholar] [CrossRef]
Figure 1. Milestones in the development and application of DNA–DNA in situ hybridisation, with a specific focus on plant research. The six milestones correspond to the following publications: [3,4,6,19,24,30]. Figure created using BioRender (https://BioRender.com, accessed on 15 July 2025).
Figure 1. Milestones in the development and application of DNA–DNA in situ hybridisation, with a specific focus on plant research. The six milestones correspond to the following publications: [3,4,6,19,24,30]. Figure created using BioRender (https://BioRender.com, accessed on 15 July 2025).
Ijms 26 07013 g001
Figure 2. Current major areas in plant cytomolecular research (IV). (A) Karyogram of Cicer arietinum CDC Frontiers (kabuli type) chromosomes, discriminated using FISH on mitotic metaphase chromosomes with a combination of painting probes: CAF-OP1 (green), CAF-OP2 (red), and 5S rDNA (yellow) (adapted from [58]). (B) Ideogram showing the chromosome-level structure of the Catalobus pendulus genome, based on CCP analysis, indicating the positions of 44 genomic blocks on its 15 chromosomes (Cp1–Cp15). The photomicrograph shows an example of CCP with mitotic and pachytene chromosomes Cp9 and Cp10 painted using arabidopsis BAC contigs representing ancestral genomic blocks V, W, and X, respectively (adapted from [59]). (C) Exemplary chromosome translocations identified by oligo-FISH on mitotic metaphase chromosomes of Musa acuminata ITC0660 ‘Khae (Phrae)’, using probes specific for the long arm of chromosome 10 and the short and long arms of chromosome 7 (adapted from [60]). (D,E) FISH with centromeric, telomeric, and chromosome 2-specific probes applied to ultra-thin root sections of Oryza sativa, prepared using a cryomicrotome (adapted from [61]). The images show evidence of Rabl configuration in xylem (D) and cortex (E) cells. (F) 3D-SIM maximum intensity projections of FISH with the pAL centromeric repeat on structurally preserved, acrylamide-embedded 4C leaf nuclei of arabidopsis WT and cap-d3 mutants. The cap-d3 mutation affects centromere association but does not alter their overall spatial arrangement within the nuclei (adapted from [62]). (G) Changes in the localisation of the arabidopsis nuclear lamina protein CRWN1 under salt stress (adapted from [63]). (H) Multicolour FISH labelling of the Lathyrus sativus homoeologue of pea chromosome 6 using PS6 painting probes, along with probes for the satellite repeats FabTR-54, which fills the gap in the PS6-C signal, and FabTR-2, which is associated with CENH3 chromatin in L. sativus (adapted from [64]). (I) Interaction of the centromeric protein CENH3 (red) with α-tubulin (green) in metaphase chromosomes of Prionium serratum, visualised using spatial SIM (adapted from [65]). (J) Overview of the image analysis workflow for examining the ultrastructure of the Hordeum vulgare 5H metaphase chromosome after FISH using centromeric, 35S rDNA-targeting, telomeric, subtelomeric, and 5HL-specific oligo probes (adapted from [66]). (K,L) Meiotic chromosome behaviour in the autodecaploid Saccharum spontaneum clone Yunnan 82–106 revealed using CP oligo probes (adapted from [67]). (K) Pairing configurations at pachytene showing ten copies of chromosome 8 forming two bivalents and one hexavalent. (L) Five bivalents of chromosome 8 with their detailed structure visualised using dual-colour barcoded painting FISH. (M) GISH using gDNA of Agropyron cristatum (red) reveals the presence and stable meiotic behaviour of the translocated A. cristatum 1P chromosome segment in the wheat background of the T1AS.1PL translocation line (adapted from [68]). (N) Visualisation of Aegilops geniculata chromosomes (indicated by arrows) at mitotic metaphase in a Triticum aestivum–A. geniculata substitution line, using GISH with A. geniculata genomic DNA (green) and FISH with the pTa535 D-genome-specific probe (red) (adapted from [69]). Figure created using BioRender (https://BioRender.com, accessed on 15 July 2025). All materials presented in this figure, except for the photomicrograph in panel (M), were published under the terms of the Creative Commons Attribution (CC-BY 4.0) licence (https://creativecommons.org/licenses/by/4.0/). The image in panel (M) has been published under an exclusive licence to Springer-Verlag GmbH Germany, with the authors of the original publication retaining copyright.
Figure 2. Current major areas in plant cytomolecular research (IV). (A) Karyogram of Cicer arietinum CDC Frontiers (kabuli type) chromosomes, discriminated using FISH on mitotic metaphase chromosomes with a combination of painting probes: CAF-OP1 (green), CAF-OP2 (red), and 5S rDNA (yellow) (adapted from [58]). (B) Ideogram showing the chromosome-level structure of the Catalobus pendulus genome, based on CCP analysis, indicating the positions of 44 genomic blocks on its 15 chromosomes (Cp1–Cp15). The photomicrograph shows an example of CCP with mitotic and pachytene chromosomes Cp9 and Cp10 painted using arabidopsis BAC contigs representing ancestral genomic blocks V, W, and X, respectively (adapted from [59]). (C) Exemplary chromosome translocations identified by oligo-FISH on mitotic metaphase chromosomes of Musa acuminata ITC0660 ‘Khae (Phrae)’, using probes specific for the long arm of chromosome 10 and the short and long arms of chromosome 7 (adapted from [60]). (D,E) FISH with centromeric, telomeric, and chromosome 2-specific probes applied to ultra-thin root sections of Oryza sativa, prepared using a cryomicrotome (adapted from [61]). The images show evidence of Rabl configuration in xylem (D) and cortex (E) cells. (F) 3D-SIM maximum intensity projections of FISH with the pAL centromeric repeat on structurally preserved, acrylamide-embedded 4C leaf nuclei of arabidopsis WT and cap-d3 mutants. The cap-d3 mutation affects centromere association but does not alter their overall spatial arrangement within the nuclei (adapted from [62]). (G) Changes in the localisation of the arabidopsis nuclear lamina protein CRWN1 under salt stress (adapted from [63]). (H) Multicolour FISH labelling of the Lathyrus sativus homoeologue of pea chromosome 6 using PS6 painting probes, along with probes for the satellite repeats FabTR-54, which fills the gap in the PS6-C signal, and FabTR-2, which is associated with CENH3 chromatin in L. sativus (adapted from [64]). (I) Interaction of the centromeric protein CENH3 (red) with α-tubulin (green) in metaphase chromosomes of Prionium serratum, visualised using spatial SIM (adapted from [65]). (J) Overview of the image analysis workflow for examining the ultrastructure of the Hordeum vulgare 5H metaphase chromosome after FISH using centromeric, 35S rDNA-targeting, telomeric, subtelomeric, and 5HL-specific oligo probes (adapted from [66]). (K,L) Meiotic chromosome behaviour in the autodecaploid Saccharum spontaneum clone Yunnan 82–106 revealed using CP oligo probes (adapted from [67]). (K) Pairing configurations at pachytene showing ten copies of chromosome 8 forming two bivalents and one hexavalent. (L) Five bivalents of chromosome 8 with their detailed structure visualised using dual-colour barcoded painting FISH. (M) GISH using gDNA of Agropyron cristatum (red) reveals the presence and stable meiotic behaviour of the translocated A. cristatum 1P chromosome segment in the wheat background of the T1AS.1PL translocation line (adapted from [68]). (N) Visualisation of Aegilops geniculata chromosomes (indicated by arrows) at mitotic metaphase in a Triticum aestivum–A. geniculata substitution line, using GISH with A. geniculata genomic DNA (green) and FISH with the pTa535 D-genome-specific probe (red) (adapted from [69]). Figure created using BioRender (https://BioRender.com, accessed on 15 July 2025). All materials presented in this figure, except for the photomicrograph in panel (M), were published under the terms of the Creative Commons Attribution (CC-BY 4.0) licence (https://creativecommons.org/licenses/by/4.0/). The image in panel (M) has been published under an exclusive licence to Springer-Verlag GmbH Germany, with the authors of the original publication retaining copyright.
Ijms 26 07013 g002
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

Wolny, E.; Mur, L.A.J.; Ohmido, N.; Yin, Z.; Wang, K.; Hasterok, R. Thriving or Withering? Plant Molecular Cytogenetics in the First Quarter of the 21st Century. Int. J. Mol. Sci. 2025, 26, 7013. https://doi.org/10.3390/ijms26147013

AMA Style

Wolny E, Mur LAJ, Ohmido N, Yin Z, Wang K, Hasterok R. Thriving or Withering? Plant Molecular Cytogenetics in the First Quarter of the 21st Century. International Journal of Molecular Sciences. 2025; 26(14):7013. https://doi.org/10.3390/ijms26147013

Chicago/Turabian Style

Wolny, Elzbieta, Luis A. J. Mur, Nobuko Ohmido, Zujun Yin, Kai Wang, and Robert Hasterok. 2025. "Thriving or Withering? Plant Molecular Cytogenetics in the First Quarter of the 21st Century" International Journal of Molecular Sciences 26, no. 14: 7013. https://doi.org/10.3390/ijms26147013

APA Style

Wolny, E., Mur, L. A. J., Ohmido, N., Yin, Z., Wang, K., & Hasterok, R. (2025). Thriving or Withering? Plant Molecular Cytogenetics in the First Quarter of the 21st Century. International Journal of Molecular Sciences, 26(14), 7013. https://doi.org/10.3390/ijms26147013

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop