Next Article in Journal
A High-Resolution Linkage Map Construction and QTL Analysis for Morphological Traits in Anthurium (Anthurium andraeanum Linden)
Next Article in Special Issue
Development of S Haplotype-Specific Markers to Identify Genotypes of Self-Incompatibility in Radish (Raphanus sativus L.)
Previous Article in Journal
Physiochemical Responses and Ecological Adaptations of Peach to Low-Temperature Stress: Assessing the Cold Resistance of Local Peach Varieties from Gansu, China
Previous Article in Special Issue
Genetic Variability for Micronutrient Content and Tuber Yield Traits among Biofortified Potato (Solanum tuberosum L.) Clones in Ethiopia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genomic and Cytogenetic Analysis of Synthetic Polyploids between Diploid and Tetraploid Cotton (Gossypium) Species

by
Mukhammad T. Khidirov
1,
Dilrabo K. Ernazarova
1,2,
Feruza U. Rafieva
1,
Ziraatkhan A. Ernazarova
1,
Abdulqahhor Kh. Toshpulatov
1,
Ramziddin F. Umarov
1,
Madina D. Kholova
1,
Barno B. Oripova
1,
Mukhlisa K. Kudratova
1,
Bunyod M. Gapparov
1,
Maftunakhan M. Khidirova
2,
Doniyor J. Komilov
3,
Ozod S. Turaev
1,2,
Joshua A. Udall
4,
John Z. Yu
4,* and
Fakhriddin N. Kushanov
1,2,3,*
1
Institute of Genetics and Plant Experimental Biology, Academy of Sciences of the Republic of Uzbekistan, Tashkent 111226, Uzbekistan
2
Department of Genetics, National University of Uzbekistan, Tashkent 100174, Uzbekistan
3
Department of Biology, Namangan State University, Uychi Street-316, Namangan 160100, Uzbekistan
4
United States Department of Agriculture (USDA)-Agricultural Research Service (ARS), Southern Plains Agricultural Research Center, 2881 F&B Road, College Station, TX 77845, USA
*
Authors to whom correspondence should be addressed.
Submission received: 30 October 2023 / Revised: 8 December 2023 / Accepted: 12 December 2023 / Published: 17 December 2023
(This article belongs to the Collection Advances in Plant Breeding)

Abstract

:
Cotton (Gossypium spp.) is the most important natural fiber source in the world. The genetic potential of cotton can be successfully and efficiently exploited by identifying and solving the complex fundamental problems of systematics, evolution, and phylogeny, based on interspecific hybridization of cotton. This study describes the results of interspecific hybridization of G. herbaceum L. (A1-genome) and G. mustelinum Miers ex Watt (AD4-genome) species, obtaining fertile hybrids through synthetic polyploidization of otherwise sterile triploid forms with colchicine (C22H25NO6) treatment. The fertile F1C hybrids were produced from five different cross combinations: (1) G. herbaceum subsp. frutescens × G. mustelinum; (2) G. herbaceum subsp. pseudoarboreum × G. mustelinum; (3) G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum; (4) G. herbaceum subsp. africanum × G. mustelinum; (5) G. herbaceum subsp. euherbaceum (variety A-833) × G. mustelinum. Cytogenetic analysis discovered normal conjugation of bivalent chromosomes in addition to univalent, open, and closed ring-shaped quadrivalent chromosomes at the stage of metaphase I in the F1C and F2C hybrids. The setting of hybrid bolls obtained as a result of these crosses ranged from 13.8–92.2%, the fertility of seeds in hybrid bolls from 9.7–16.3%, and the pollen viability rates from 36.6–63.8%. Two transgressive plants with long fiber of 35.1–37.0 mm and one plant with extra-long fiber of 39.1–41.0 mm were identified in the F2C progeny of G. herbaceum subsp. frutescens × G. mustelinum cross. Phylogenetic analysis with 72 SSR markers that detect genomic changes showed that tetraploid hybrids derived from the G. herbaceum × G. mustelinum were closer to the species G. mustelinum. The G. herbaceum subsp. frutescens was closer to the cultivated form, and its subsp. africanum was closer to the wild form. New knowledge of the interspecific hybridization and synthetic polyploidization was developed for understanding the genetic mechanisms of the evolution of tetraploid cotton during speciation. The synthetic polyploids of cotton obtained in this study would provide beneficial genes for developing new cotton varieties of the G. hirsutum species, with high-quality cotton fiber and strong tolerance to biotic or abiotic stress. In particular, the introduction of these polyploids to conventional and molecular breeding can serve as a bridge of transferring valuable genes related to high-quality fiber and stress tolerance from different cotton species to the new cultivars.

1. Introduction

Cotton belongs to the genus Gossypium which consists of approximately 46 diploid species and seven allotetraploid species [1,2,3]. All diploid (2n = 2x = 26) cotton species belong to eight genome groups (A-G and K), and all tetraploid (2n = 4x = 52) species are classified into one genome group (AD) [4,5]. Gossypium species spread through large geographical and ecological areas and has a wide range of morphological and genetic diversity, mainly preserved in germplasm collections and as genetic materials for cotton breeding programs worldwide. These resources can be successfully used to transfer valuable traits from wild species into elite cultivars [6,7]. Polyploidization is one mechanism that introduces the genetic diversity of diploids to polyploids, which allows easy crossing with tetraploids and obtaining fertile hybrids. Unfortunately, issues arise when diploid wild species are crossed with tetraploids. It is often impossible to obtain fertile hybrids through crossing between tetraploid and wild diploid species [8,9]. Completely sterile triploid (2n = 3x = 39) hybrids are obtained from the natural hybridization of tetraploid elite cultivars with wild diploid species [4]. Nevertheless, the fertility of hybrids can be recovered by polyploidization, resulting in hexaploid (2n = 6x = 78) plants [10].
The process of ancient genome duplication is called paleopolyploidy, which occurred at least several million years ago (MYA) [11]. In paleopolyploidy, the duplication of the genome of a single species is called autopolyploid, while combining the genomes of two species is allopolyploidy. Polyploidy is common in plants, fish, and some amphibians. Polyploids have more than two complete sets of chromosomes, which are inheritable. Diploid gametes can result in increased chromosome sets because of a failure to divide during meiosis. A triploid zygote is formed when these diploid gametes fuse with a monoploid gamete. These triploids can be unstable and sterile. When these diploid gametes fuse with another diploid gamete, a stable tetraploid zygote is formed. Polyploids can be tetraploid (4x), hexaploid (6x), and other aneuploids, higher order polyploids, or multiple chromosome pairs. The chromosome doubling through polyploidization can enhance the genetic diversity of plants.
The first comprehensive study in this field was conducted 250 years ago by Carl Linnaeus [12]. Polyploidization event was first discovered in 1907 and was considered responsible for increasing the number of chromosomes [13]. The development of plants that change as a result of polyploidization increases their adaptability to adverse conditions and environmental factors [14,15]. Currently, several chemical and gaseous agents are used in research to obtain polyploids. The most used agents are colchicine and oryzalin [16]. Colchicine (C22H25NO6) is obtained from the bulb-like corms of the plant Colchicum autumnale L. [17]. This substance is an alkaloid agent used to induce polyploidy [18]. This type of chemical binds to the tubulin protein and prevents the formation of microtubule spindle fibers (achromatin threads) during the metaphase of mitotic cell division, resulting in the duplication of chromosome bundles without spreading to the poles [19]. Colchicine can effectively stop cell division at the mitotic anaphase. At this stage, the chromosomes have already been duplicated, but mitosis has not yet occurred, and the restriction of cell wall formation at this stage leads to the formation of polyploid cells. They are usually larger than diploid species and often have thickening of the tissues, which leads to the formation of large plant organs [20].
The diversity of Gossypium species is an ideal model for understanding the evolution and domestication of polyploids, as well as the pathways and mechanisms of gene and genome evolution [21]. There are several large Gossypium germplasm collections around the world, and among the largest and richest germplasm collections are those in the USA and Uzbekistan. These germplasm collections consist of cotton accessions with high genetic diversity, and their widespread use in genetics and breeding is of great importance in solving the current problems facing the cotton industry [22,23].
The historical domestication and modern breeding of Upland cotton (Gossypium hirsutum) have significantly improved its yield and fiber quality, but this has led to a dramatically reduced genetic diversity [24,25,26,27]. Upland cotton cultivars have a narrow genetic base of fiber quality and abiotic stress tolerance traits. The gene introgression from species such as G. herbaceum and G. mustelinum allows the increase of genetic diversity by introducing new alleles to improve fiber quality and stress tolerance. G. herbaceum, also known as ‘Levant cotton’, is grown in rainfed areas of Africa, and this crop is tolerant to salinity, drought, and scorching hot weather [28]. G. mustelinum is a wild species native to the northeastern region of Brazil, among existing allotetraploids [29]. The fiber of G. mustelinum is genetically different from that of allotetraploid G. hirsutum. The quantitative trait loci (QTL) for fiber quality and the interesting alleles identified from G. mustelinum can make an important contribution to the improvement of Upland cotton germplasm [25]. Wang et al. [30] dissected the molecular genetic basis of fiber strength and fineness in G. mustelinum and G. hirsutum crosses. They identified 42 QTLs, highlighting the potential of G. mustelinum alleles to improve fiber quality in Upland cotton. The study revealed new allelic variation for cotton breeding. Yang et al. [31] made a chromosome-level genome assembly of G. mustelinum, demonstrating its efficacy in identifying genes for qualitative and quantitative traits. The study laid a foundation for cotton genetics and breeding, emphasizing the rich gene pool of G. mustelinum. Chen et al. [25] used selective genotyping to validate over 75 QTLs for fiber quality traits that were introgressed from G. mustelinum into Upland cotton. The study lays the foundation for fine mapping, marker-assisted selection, and map-based gene cloning.
In the past few decades, the following studies were conducted on obtaining interspecific hybrids exploiting the genetic diversity of Gossypium germplasm: G. hirsutum × G. klotzschianum [32], G. hirsutum × G. trilobum [33], G. arboreum × G. anomalum [34], G. hirsutum × G. anomalum [35], G. herbaceum × G. australe [36], G. hirsutum × G. australe [37], G. hirsutum × G. arboreum [38], G. capitis-viridis × (G. hirsutum × G. australe)2 [39], G. hirsutum × G. darwinii [40], G. herbaceum × G. nelsonii [41]. Despite an extensive literature review, no reports were found on obtaining allotetraploids using the cotton species G. herbaceum and G. mustelinum.
In this study, we hypothesized that obtaining fertile hybrids through the polyploidization of triploids derived from the cross between G. herbaceum and G. mustelinum could enable their use in tetraploid cotton breeding. To test this hypothesis, the goals of our research were as follows: (1) to obtain F1 triploid hybrids of G. herbaceum and G. mustelinum cotton species; (2) to obtain F1C hexaploids through doubling the genome of triploid hybrids using colchicine treatment; and (3) to select tetraploids among segregating F2C aneuploid (triploid, tetraploid, and hexaploid) hybrids. The new knowledge and genetic resources would be used in the introduction of valuable traits such as fiber quality and stress tolerance into the elite Upland cotton varieties using marker-assisted selection (MAS) technology.

2. Results

2.1. The Characteristics of Interspecific Cotton Hybrids between G. herbaceum and G. mustelinum

Several crosses were made between G. herbaceum (A1 genome) and G. mustelinum (AD4 genome) species to develop interspecific hybrids with valuable traits as initial materials for genetics and breeding studies. As a result of many crossings, the interspecific triploid F1 hybrids were obtained in five different combinations such as (1) G. herbaceum subsp. frutescens × G. mustelinum, (2) G. herbaceum subsp. pseudoarboreum × G. mustelinum, (3) G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum, (4) G. herbaceum subsp. africanum × G. mustelinum, and (5) G. herbaceum subsp. euherbaceum (variety A-833) × G. mustelinum. The boll-setting rate of the hybrids in the recipient buds after crossing was 13.8–92.2%, and complete seed-setting rate in hybrid bolls was 9.7–16.3%. The highest rate of fertility was observed between variety A-833 (subsp. euherbaceum) and subsp. frutescens (cultivated tropical form), and boll-setting rate was over 50.0%. A relatively low fertility rate was recorded in subsp. pseudoarboreum × G. mustelinum (13.8%). Complete seed-setting rates showed relatively similar among all combinations (9.7–16.3%) (Table 1).
Typically, triploids are infertile, some of them are semi-fertile and can produce both aneuploid and euploid gametes. Taking this into account, the resulting F1 triploid hybrid seeds were treated with colchicine before sowing. Triploid and polyploid hybrids were obtained as following scheme (Figure 1).
Only, G. herbaceum subsp. africanum × G. mustelinum, subsp. pseudoarboreum f. harga × G. mustelinum, subsp. frutescens × G. mustelinum (Figure 2) combinations were grown after colchicine treatment to produce viable F1C hybrids.

2.2. Inheritance of Morphological and Economically Valuable Traits in Interspecific Polyploid Hybrids F1C and F2C

2.2.1. The Number of Bolls per Plant and the Fertility of Seeds

It is known that parameters such as the number of bolls in a plant, fiber yield, weight of 1000 seeds, weight of cotton in one boll and the number of fully born seeds in one boll play an important role in determining the level of productivity. Since these indicators are inextricably linked with each other and have a correlative effect on the development of the character. In addition, qualitative and quantitative indicators of seeds born in one bag are used as a basis for determining the phylogenetic relationships of species based on classical methods. Therefore, the initial sources used in the rese—arch and the yield indicators of the polyploid hybrids obtained on their basis were analyzed. In terms of the number of bolls per plant among the subspecies of G. herbaceum, the highest rate (36.0 pieces) was recorded in the form of subsp. pseudoarboreum f. harga. In the analyzed bolls, the rate of complete seed (or, fertility of seeds) was 89.2 ± 1.9%. Also, the number of bolls per plant showed positive indicators on subspecies of subsp. africanum—27.0 pieces, and the yield of complete seeds in the boll was 94.5 ± 1.8%, in the variety A-833 (subsp. euherbaceum)—28.0 pieces, and the rate of complete seeds was 86.9 ± 1.9%. In G. mustelinum—a Brazilian endemic species, the average number of bolls per plant was—16.0, and the rate of complete seed was—86.0 ± 2.1% (Table 2).
In F1C subsp. frutescens × G. mustelinum polyploid hybrids, the number of bolls per plant was—21.0, and the “fertility rate of seeds per bolls” was very low (19.4 ± 1.9%) (Table 3). An average of 14.3 seeds were set in one boll, of which 2.8 whole seeds and 11.5 empty seeds were set. In F1C subsp. pseudoarboreum f. harga × G. mustelinum hybrids, the number of bolls per plant was—1.0, the number of complete seeds was 2.0, and the number of empty seeds was 7.0. F1C subsp. africanum × G. mustelinum hybrid plants were sterile, and no yield components were observed.
In hybrid plants of subsp. frutescens × G. mustelinum F2C generation, there was a variation of 18.0–26.0 units in terms of “number of bolls per plant”. Relatively low values (between 30.6 ± 1.2% and 49.2 ± 2.1%) were found in the “seeds fertility” indicator, also. According to the results of the analysis, in polyploid hybrids of the F1C and F2C generations, there was a slight decrease in the number of pods per plant compared to the parental samples, as well as cases of sterility in the subsp. africanum × G. mustelinum combination. A sharp decrease was noted in terms of the rate of complete seed fertility, which is the main factor of productivity. This indicates that the parental lines are phylogenetically distant and isolated species.

2.2.2. Vegetative Growth Duration

According to the results of the research, wild (subsp. africanum) and ruderal (subsp. pseudoarboreum f. harga) forms of G. herbaceum of which, the vegetative growth duration was 137.4 and 127.2 days, therefore they would require breeding for short day conditions. The vegetative growth duration in the cultivated-tropical subsp. frutescens, ruderal subsp. pseudoarboreum subspecies and cultivated subsp. euherbaceum A-833 variety is 117.2, 119.6, and 117.4 days respectively, in field conditions since they do not require breeding for short days. The wild tetraploid species G. mustelinum also requires breeding for short days since the duration of its vegetative growth duration was 151.8 days.
F1C hybrid plants obtained by cross-species with representatives of the diploid G. herbaceum species and tetraploid G. mustelinum species was between 128.6–145.2 days. Including F1C subsp. frutescens × G. mustelinum the vegetative duration of polyploid hybrid plants was 128.6 days, the coefficient of variation was 2.3%. The coefficient of dominance for the sign is equal to hp = 0.34, and the state of partial dominance of the form with a positive indicator (subsp. frutescens) was noted. The vegetative growth duration of F1C subsp. pseudoarboreum f. harga × G. mustelinum polyploid hybrid combination plants was 139.6 days on average, the variation coefficient was 1.5%, and the dominance coefficient was hp = −0.01. In this combination, the character was inherited with a partial dominance of the negative indicator form (G. mustelinum). The average vegetative growth duration of polyploid hybrid plants was 139.6 days, the variation coefficient was 1.5%, and the dominance coefficient was hp = −0.01. In this combination, the character was inherited in the partial dominance of the G. mustelinum species with a negative indicator. The polyploid hybrid plants of F1C subsp. africanum × G. mustelinum combination had an average vegetative growth duration of 145.2 days, and the coefficient of variation was 2.2% (Table 4). The coefficient of dominance for the studied character is equal to hp = −0.08, and in this combination, the character was inherited in the case of partial dominance of the negative indicator form (G. mustelinum).
From the above-mentioned F1C polyploid hybrids, only F2C hybrids of subsp. frutescens × G. mustelinum combination were obtained. Also, in this generation, the degree of variability of the indicators of the period from the germination to the opening of the first bud was studied in 50% of the plants. A total of 134 plants were obtained in the F2C generation of the subsp. frutescens × G. mustelinum combination, of which 53 strongly required breeding for short days and thus affected yield components under field conditions. In the remaining 81 F2C hybrids, the growth duration showed a wide range of variability between 116.6–152.2 days, which were divided into eight classes. Among the studied hybrid combinations, the growth duration was 115–119 day and faster recombinant forms around 120–124 were isolated.

2.2.3. Fiber Length and Strength

Cotton has been cultivated for centuries mainly for its fiber and has undergone natural and artificial selection. The wild diploid and tetraploid species are the natural resources with high genetic diversity for improving the fiber quality traits of elite cultivars. Therefore, fiber length is considered one of the major indicators of valuable economic traits in the breeding process.
According to the results of fiber quality tests the fiber length was in the range of 17.2–25.2 mm in G. herbaceum subspecies (Table 5). In this case, a relatively low result was observed in subsp. frutescens (17.2 ± 0.2 mm) and a relatively high result was observed in cultivated variety subsp. euherbaceum A-833 (25.2 ± 0.3 mm). As well as 25.1 ± 0.4 mm was in ruderal subsp. pseudoarboreum, 20.1 ± 0.2 mm was in subsp. pseudoarboreum f. harga and 24.7 ± 0.3 mm was in wild-type subsp. africanum. In the case of tetraploid G. mustelinum, the fiber length was 25.6 ± 0.2 mm.
The fiber strength varied in all of the cotton samples including the highest values of 35.5 cN/tex and 32.6 cN/tex, respectively, in subsp. frutescens and subsp. africanum. The fiber strength in A-833 variety of subsp. euherbaceum was 27.2 cN/tex, in subsp. pseudoarbareum f. harga was 24.1 cN/tex, and in G. mustelinum was 17.3 cN/tex (Table 5).
Due to two F1C allohexaploid combinations were dead after the germination and the other two being male sterile, therefore all analyses were conducted only with G. herbaceum subsp. frutescens × G. mustelinum F1C allohexaploids since they were fertile. The fiber length was 28.7 ± 0.9 mm, and a positive heterosis (hp = 1.74) was observed.
Variability was observed in the plants of F2C subsp. frutescens × G. mustelinum hybrid combinations. The analyzed results were divided into eight classes. Accordingly, two transgressive plants with long fiber of 35.1–37.0 mm, and one transgressive plant with extra-long fiber of 39.1–41.0 mm were identified. Since the fiber of the Brazilian endemic species G. mustelinum differs from that of the allotetraploid species G. hirsutum in terms of strength, maturity and length, a genetically rich source for improving fiber qualities of cultivars.
The identification of major QTL loci for fiber quality and application of a set of beneficial alleles from G. mustelinum can contribute significantly to the long-term improvement of cultivated cotton germplasm. To introduce important alleles from G. mustelinum species into the genome of G. hirsutum species for fiber quality improvement, interspecific populations consisting of plants obtained by crossing G. mustelinum species with G. hirsutum species were created [43,44,45]. In the studied families, it was determined that the genes responsible for fiber strength and fineness in alleles of G. mustelinum are dominantly inherited. Based on the previous work with G. barbadense, G. tomentosum, and G. darwinii, which included introgression of G. mustelinum alleles [43,44,45]. These hybrid genotypes identified in our study would serve as a valuable source for the development of long-fiber cultivars and future breeding programs.

2.3. Cytogenetic and Genomic Studies of Interspecific Hybrids

Chromosome conjugation, spore and pollen viability analysis at metaphase I (MI) stage of meiosis is one of the effective methods for determining the ploidy level of intergenomic hybrids (allopolyploids, autopolyploids). Cytogenetic analyzes of hybrids resulting from the crosses between diploid and tetraploid cotton species were conducted in the study, and interesting data were obtained regarding their genomic structure and their partial homology.
In the three hexaploid plants studied, disturbances in the process of the first division of meiosis (MI) were detected. F1C subsp. frutescens × G. mustelinum 38.22 ± 0.25 bivalents, 0.45 ± 0.29 univalents (the number of univalents was up to six) and 0.36 ± 0.14 quadrivalents. F1C subsp. pseudoarboreum f. harga × G. mustelinum and F1C subsp. africanum × G. mustelinum combinations had a higher number of univalents than the above combination (up to 2–10) (Table 6, Figure 3). The analysis showed that due to structural differences in chromosomes and the impossibility of normal chromosome conjugation in hybrids, there was a desynaptic effect due to early and asynchronous divergence in individual bivalents (Figure 3 a,b).
In general, the number of univalents found in maternal pollen cells was recorded from two to ten. It was known that there is weak desynapsis (meeting of several univalents along with bivalents in the female pollen cell), medium desynapsis (meeting of many univalents along with bivalents in the female pollen cell), and complete desynapsis (meeting of mainly univalents and sometimes several bivalents in the female pollen cell) [46]. Such univalents were formed as a result of early divergence of chromosomes. In the cotton samples of our study, this level—from two to ten univalent encounters—corresponded to the medium desynapsis.
As a result of research, the MI phase of meiosis was observed in some plants of the F2C subsp. frutescens × G. mustelinum hexaploid hybrid (Figure 4). The cotton plant samples with different levels of ploidy (24.73II; 38.18II; 38.00II) were identified in second generation (F2C). Disruptions in the mentioned chromosomal conjugation led to partial sterility of anthers in hexaploid hybrids and abnormal spore formation at the sporulation stage. It should also be noted that the MI phase of meiosis was carried out normally (26.00II) in two hybrid plants (Table 6).
Thus, the cause of partial infertility observed in allopolyploids was likely due to the presence of disturbances in the meiosis phase (disruption of the synchronous distribution of chromosomes in the anaphase or desynapsis phenomenon), instability of the number of chromosomes. One of the unique features of hybrids derived from the crosses between cotton species far from each other was with chromosomal aberrations. In such cases, plants of with chromosomal aberrations are due to the absence of homologous chromosomes of different species or the presence of their only partial homologue. Hybrids with homeologous genomes are viable, but sterile, i.e., sterile, due to genome duplication [47]. According to the conclusions of Sanamyan [48], hybrid combinations obtained from the diploid species (G. thurberi × G. raimondii, G. arboreum × G. thurberi, and G. herbaceum × G. thurberi) are productive in generations, on the contrary, intergenomic hybrids are very a large percentage of plants without flowers to appear.
There were difficulties in spore analysis to determine the required stage due to the very low number of spikelets and pollen grains in polyploid hybrids. As a result, ten additional samples had to be analyzed. According to the analysis of the conducted tetrads, 90.3–96.8% meiotic stage in three hybrid forms (F1C subsp. frutescens × G. mustelinum, F1C subsp. pseudoarboreum f. harga × G. mustelinum, F1C subsp. africanum × G. mustelinum) index was defined. It was noted that certain disturbances of the meiotic index in the form of micronuclear tetrads and polyads were observed in all forms. In F1C subsp. frutescens × G. mustelinum polyploid hybrid, the meiotic index was 90.36%, the rate of micronuclear tetrads was 3.27% and that of polyads was 6.37% (Table 7).
Micronuclear tetrads were not found in F1C subsp. pseudoarboreum f. harga × G. mustelinum polyploid hybrid unlike other plants. F2C subsp. frutescens × G. mustelinum polyploid hybrids T 38-12, T 30-6, T 30-7, T 30-8, T 38-3 forms meiotic index 96.05–98.54%, polyads 1.42–3.12%, micronuclear tetrads were not observed. In T 1-8 and T 51-13 forms, meiotic index was 96.15–97.45%, polyads 2.11–3.07%, micronuclear tetrads 0.69–0.89%.
In tetrads with identified micronuclei, up to 1–8 micronuclei (Figure 5c), and from polyads up to pentad, hexad, heptad, octad (Figure 5d) aneuploid spores were noted.
It was known that pollen grains formed as result of meiosis are not equal in terms of genetic characters and functional capabilities. When the chromosomal complex of pollen grains was insufficient (i.e., when the frequency of meeting homologous chromosomes was low, the pollen grains were underdeveloped, and the viable pollen decreased significantly). When a complete chromosome complex was obtained, that is, when the homologous row of chromosomes was restored, the pollen grains became pink and fertile again. As a result of cytogenetic analysis, pollen graininess in hybrid plants F1C subsp. pseudoarboreum f. harga × G. mustelinum 63.89% in the combination of, F1C subsp. frutescens × G. mustelinum combination had a low yield—32.67%, F1C subsp. africanum × G. mustelinum combination, pollen sterility was determined (Table 8, Figure 6).
Since most of the hybrid plants showed short day and late tolerance in F2C generation of the combination subsp. frutescens × G. mustelinum, 24 out of 134 plants were studied for pollen graininess (Table 9). Until late autumn, reproductive organs were not formed in many hybrids, abnormal forms were noted without pollen development in hybrids that started flowering. There were a few cases where dust grains were not formed.
The pollen fertility in the studied hybrids had different indicators. The pollen fertility of 14 hybrids was low (1.49–40.43%). Among them T 33-6, T 30-8, T49-2, T 38-7 plants were very low (1.49–10.23%), T 33-2, T 37-4, plants up to 50% or more, T 38-3, T 30-6, T 30-7, T 38-12, more than 60% (Table 8, Figure 7) indicators of dustiness were found in plants.

2.4. Molecular and Phylogenetic Analysis of Allopolyploid Forms of Cotton

The PCR analysis was conducted using 72 simple sequence repeat (SSR) markers (Table S1) associated with economically important traits to determine the genetic polymorphisms between parental genotypes as well as to determine some genomic changes in allohexaploid hybrids (Figure 8). As a result of PCR analysis, 57 DNA markers were determined as polymorphic markers (Table S2), while 14 DNA markers were monomorphic, and one DNA marker was not amplified.
Based on the results of PCR analysis, the genomic changes were determined in the genomic regions of 18 (31.6%) out of 57 polymorphic DNA markers or in 40 (26.7%) PCR amplicons (alleles) of 150 alleles in allohexaploid hybrids. In particular, 31 (26.0%), 29 (24.3%), and 27 (22.6%) alleles had a change in three F1C combinations (G. herbaceum subsp. frutescens × G. mustelinum), F1C (G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum) and F1C (G. herbaceum subsp. africanum × G. mustelinum) hybrids, of 119, 120 and 116 alleles, respectively. In these 3 allohexaploid hybrids, alleles 23, 21, and 20 were missing although they were present in parental genotypes, respectively. On the other hand, non-parental alleles 8, 8, and 7 appeared in allohexaploid hybrids, respectively.
A phylogenetic tree construction indicated that the cotton accessions were clustered into two main groups (Figure 9). The first group included three Gossypium herbaceum subspecies such as subsp. frutescens, subsp. pseudoarboreum f. harga and subsp. africanum. G. herbaceum subsp. pseudoarboreum f. harga and subsp. africanum are genetically more similar to each other than subsp. frutescens. G. mustelinum and three interspecific allohexaploid hybrids F1C (G. herbaceum subsp. frutescens × G. mustelinum), F1C (G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum) and F1C (G. herbaceum subsp. africanum × G. mustelinum) were presented in the second group of the phylogenetic tree.

3. Discussion

To obtain synthetic polyploids, some chemical mutagens, especially colchicine, are most effective. Colchicine prevents spindle formation during mitosis, which prevents the separation of daughter chromosomes in anaphase and cytokinesis, leading to a doubling of the number of chromosomes in the cell [49]. In this study, to increase the number of chromosomes (autopolyploidy), representatives of the diploid cotton species G. herbaceum were treated with 0.1 and 0.2% solutions of colchicine and synthetic tetraploid plants were developed. In studies conducted with various percentage solutions of colchicine, it was found that a 0.2% solution of colchicine was more effective than a 0.1% solution. As an effective concentration, we recommend the use of a 0.2% (20 h) of a solution of colchicine to conduct this kind of research.
A different genetic variability was observed in F2C hybrid plants. Two transgressive plants with long fiber of 35.1–37.0 mm, and one transgressive plant with extra-long fiber of 39.1–41.0 mm were identified in the F2C progeny of G. herbaceum subsp. frutescens × G. mustelinum combination. The observed genetic variability, especially in cotton polyploid genotypes with long and extra-long fiber, has direct implications for cotton breeding. These variations would provide opportunities for selecting and developing cultivars with enhanced fiber qualities, potentially leading to the creation of commercially valuable cotton varieties with improved fiber length. According to the data mentioned above [43,44,45], although the F2C allopolyploid hybrid combination obtained with G. mustelinum (subsp. frutescens × G. mustelinum) is late maturing, fiber yield is low compared to other forms, it can be concluded that the release of long fiber and extra-long fiber plants is the effect of G. mustelinum genes. Another reason for this outcome can be explained by the mutation of some sections (gene alleles) in the genome of G. herbaceum due to the application of the experimental polyploidy method, and the dis-appearance or appearance of F1C allopolyploid hybrid genome regions. That is, depending on the distribution of the signs of parental forms in the hybrids of the second generation, it is estimated that the number of polymer genes and their constant state in future generations. These isolated forms will serve as a unique genetic resource for the development of long fiber varieties in future breeding programs. For this reason, we have planned to introduce specific F2C polyploids into conventional and molecular cotton breeding programs to obtain new varieties with desired traits through hybridization and backcrossing approaches. The parental genotypes possess distinct fiber properties, with G. mustelinum exhibiting soft and long fiber, while G. herbaceum has fiber with hygroscopic properties. The fiber characteristics mentioned above, valuable in the global cotton industry, are expected to be effectively incorporated into the future cultivars.
As a result of molecular and phylogenetic studies, it was shown that tetraploid hybrids derived from the G. herbaceum × G. mustelinum cross, are close to the species G. mustelinum. Also, it was found that compared with the first cluster of the phylogenetic tree, the second cluster appeared to be small. It only had one subcluster containing the forms subsp. frutescens, subsp. pseudoarboreum f. harga, and subsp. africanum belonging to the species G. herbaceum L., which indicates their phylogenetic relatedness. At the same time, it was noted that these forms are significantly distant from the type of G. mustelinum. Another important aspect of the research results was that subsp. frutescens from representatives of G. herbaceum are close to their cultivated forms, and that its subsp. africanum is a wilder form.

4. Materials and Methods

4.1. Plant Materials

In this study, wild, semi-wild and cultivated subspecies of diploid cotton G. herbaceum L. and wild tetraploid cotton G. mustelinum Miers ex Watt as well as their five interspecific hybrids (1) G. herbaceum subsp. frutescens × G. mustelinum; (2) G. herbaceum subsp. pseudoarboreum × G. mustelinum; (3) G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum; (4) G. herbaceum subsp. africanum × G. mustelinum; (5) G. herbaceum subsp. euherbaceum (variety A-833) × G. mustelinum were used. The seeds of parental genotypes were taken from the special cotton germplasm collection at the Institute of Genetics and Plant Experimental Biology (IGPEB) Academy of Sciences of the Republic of Uzbekistan (ASRUz). Experiments were conducted during 2020–2022 in the laboratory of Experimental polyploidy and phylogeny of cotton and the nursery at the IGPEB, located in Qibray district, Tashkent.

4.2. Hybridization and Polyploidization

Diploid (2n = 2x = 26) G. herbaceum L. subspecies were crossed with tetraploid (2n = 4x = 52) G. mustelinum Miers ex Watt and obtained five combinations of triploid (2n = 3x = 39) hybrid genotypes. Experiments were conducted under two conditions: (1) F1 hybrid seeds of triploids were treated with 0.1% colchicine for 24 h, and (2) 0.2% colchicine was applied for 20 h. Both conditions involved darkness at room temperature (22 °C) during seed germination, reaching approximately 1 cm in shoot length. Consequently, the duration of the colchicine treatment was reduced by four hours with an increased concentration. As a result, F1C (C for colchicine) synthetic allohexaploid (2n = 6x = 78) hybrid (G. herbaceum subsp. frutescens × G. mustelinum; G. herbaceum subsp. pseudoarboreum × G. mustelinum; G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum; G. herbaceum subsp. africanum × G. mustelinum; G. herbaceum subsp. euherbaceum (variety A-833) × G. mustelinum) genotypes have been obtained. Synthetic allohexaploid genotypes were planted aim to obtain F2C generations. Unfortunately, seed of two allohexaploid genotypes (G. herbaceum subsp. pseudoarboreum × G. mustelinum and G. herbaceum subsp. euherbaceum (variety A-833) × G. mustelinum) died at the germination stage. Two F1C allohexaploids (G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum and G. herbaceum subsp. africanum × G. mustelinum) were male-sterile. Only, F1C G. herbaceum subsp. frutescens × G. mustelinum allohexaploids were fertile and obtained F2C genotypes.

4.3. Cytological Analysis

Analysis of sporads and pollen viability were carried out according to Pausheva [50]. For the analysis of tetrads, 2–4 mm cotton buds (in the pinhead square and match-head square stages) were collected early morning and were fixed in the ethanol-acetic acid mixture (7:3 v/v). The samples were screened using a trinocular microscope (N-300M(MD101), Ningbo Yongxin Optics Co., Ltd., Ningbo, China) after staining with acetocarmine. For the analysis of tetrads, the meiotic index (Mi) was calculated as the percentage of normal tetrads over total sporads. In order to determine the pollen viability, pollen was collected from newly opened cotton flowers in the first half of the day and was stained with acetocarmine. Next cytological analysis was caried out using a Leica CM E microscope with a Leica EC3 camera (The Leica Microsystems Inc., Wetzlar, Germany).

4.4. Phenotypic Observation

During the 2021–2022 planting seasons, G. herbaceum L. subspecies, G. mustelinum Miers ex Watt and their F1-2C allohexaploids were evaluated for phenotypic traits in the field condition. To study morpho-biological traits, such as (i) vegetative growth period (VP); (ii) number of bolls per plant (NB); (iii) number of seeds per boll (NSB); (iv) boll weight (BW); (v) 1000-seed weight (SW); (vi) fiber length (FL); fiber strength (FS); and (vii) male fertility (MF) and sterility (MS) of pollens.

4.5. DNA Isolation and SSR Analysis

The genomic DNA from cotton plant leaf tissues was isolated using the cetyltrimethylammonium bromide (CTAB) method [51]. The DNA concentration was calculated by measuring the absorbance of 1 µL of the samples at 260/280 nm using the NanoDrop Eight spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA). DNA samples were diluted to a working concentration of 25 ng/µL. A total of 72 SSR (simple sequence repeat) markers were selected from CottonGen the cotton marker database (https://www.cottongen.org/data/markers) (accessed on 20 February 2022) [52]. PCR-based SSR genotyping was conducted as described previously [6,53,54]. The construction and visualization of the phylogenetic tree was performed using NCSS 12.

4.6. Statistical Analysis

Statistical analyses of the obtained experimental results were carried out according to Dospekhov [42]. The mean-square deviation from the mean was determined for the number of measurements n = 10 according to the formula:
σ = X i X ¯ 2 n 1
Mean-square deviation from the mean value is calculated by the following formula:
σ X ¯ = σ n 1
Substituting the reliability criteria (Student’s coefficient) into the formula, the boundaries of the confidence interval for the arithmetic mean were obtained:
= t c σ X ¯
The relative error of the results of a series of x measurements at a confidence level of 95% will be equal to:
S = 100 σ X ¯ t c X ¯
V—the coefficient of variation was calculated by the below-mentioned formula:
V = 100 σ X ¯ X ¯

5. Conclusions

In this study, we made several crosses between cotton species and obtained interspecific triploid F1 hybrids in five different combinations: (1) G. herbaceum subsp. frutescens × G. mustelinum; (2) G. herbaceum subsp. pseudoarboreum × G. mustelinum; (3) G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum; (4) G. herbaceum subsp. africanum × G. mustelinum; (5) G. herbaceum subsp. euherbaceum (variety A-833) × G. mustelinum. Fertile allohexaploid F1C hybrids were obtained by polyploidization of triploid forms. Cytogenetic analysis showed the existence of univalent, open, and closed ring-shaped quadrivalent chromosomes at the stage of metaphase I in the F1C and F2C hybrids. Also, various meiotic anomalies (tetrad cells with micronuclei and polyads), as well as low rates of pollen (36.6–63.8%) and complete sterility (F1C subsp. africanum × G. mustelinum) were observed in the analysis of tetrads. This indicates the structural heterozygosity of polyploid hybrids.
A total of 57 out of the 72 SSR markers identified in the study were determined polymorphic between cotton species. Genomic changes were observed in allohexaploid hybrids. Allele changes were observed in F1C (G. herbaceum subsp. frutescens × G. mustelinum), F1C (G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum) and F1C (G. herbaceum subsp. africanum × G. mustelinum) hybrids, 31 (26.0%), 29 (24.3%) and 27 (22.6%) out of 119, 120 and 116 alleles, respectively.
As a result of phylogenetic tree construction, the cotton accessions were clustered into two groups. The first group included three Gossypium herbaceum subspecies which are genetically similar to each other and subsp. frutescens. G. mustelinum and three interspecific allohexaploid hybrids F1C (G. herbaceum subsp. frutescens × G. mustelinum), F1C (G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum) and F1C (G. herbaceum subsp. africanum × G. mustelinum) were presented in the second group of the phylogenetic tree. Transgressive plants with long fiber were identified for potential improvement of fiber length in cultivated cottons. These synthetic allotetraploid cottons could also serve as valuable sources in the introgression of economically important traits including biotic and abiotic stress tolerance into the elite Upland cotton varieties.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/plants12244184/s1, Table S1: The panel of SSR markers associated with economically important traits; Table S2: Polymorphic SSR markers, and their polymorphism information content (PIC) and heterozygosity (He) values. References [55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98] are cited in the supplementary materials.

Author Contributions

M.T.K., D.K.E., J.Z.Y. and F.N.K. conceived the research. M.T.K., D.K.E., Z.A.E., F.U.R., M.D.K., B.B.O., M.K.K., B.M.G., A.K.T., D.J.K., M.M.K. and R.F.U. conducted experiments. O.S.T., D.K.E., F.U.R., J.A.U., J.Z.Y. and F.N.K. analyzed the data and wrote the manuscript. F.N.K. and D.K.E. managed the project. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Acknowledgments

The authors would like to remember Abdumavlyan A. Abdullaev and Sofiya M. Rizaeva for providing valuable advice to this study. This study was supported by Academy of Sciences of the Republic of Uzbekistan and USDA-ARS Cooperative Agreement 3091-21000-048-003N.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fryxell, P. The Natural History of the Cotton Tribe (Malvaceae, Tribe Gossypieae); Texas A & M University Press: College Station, TX, USA, 1979; p. 245. [Google Scholar]
  2. Wendel, J.F.; Grover, C.E. Taxonomy and evolution of the cotton genus, Gossypium. In Cotton; Agronomy Monograph; Fang, D.D., Percy, R.G., Eds.; American Society of Agronomy Inc.: Madison, WI, USA, 2015; p. 57. [Google Scholar]
  3. Wendel, J.; Brubaker, C.; Alvarez, I.; Cronn, R.; Stewart, J. Evolution and natural history of the cotton genus. In Genetics and Genomics of Cotton; Paterson, A.H., Ed.; Springer: New York, NY, USA, 2009; pp. 3–22. [Google Scholar]
  4. Brubaker, C.; Paterson, A.; Wendel, J. Comparative genetic mapping of allotetraploid cotton and its diploid progenitors. Genome 1999, 42, 184–203. [Google Scholar] [CrossRef]
  5. Wendel, J.; Cronn, R. Polyploidy and the evolutionary history of cotton. Adv. Agron. 2003, 78, 139–186. [Google Scholar] [CrossRef]
  6. Kushanov, F.; Komilov, D.; Turaev, O.; Ernazarova, D.; Amanboyeva, R.; Gapparov, B.; Yu, J. Genetic analysis of mutagenesis that induces the photoperiod insensitivity of wild cotton Gossypium hirsutum subsp. purpurascens. Plants 2022, 11, 3012. [Google Scholar] [CrossRef] [PubMed]
  7. Kushanov, F.; Turaev, O.; Ernazarova, D.; Gapparov, B.; Oripova, B.; Kudratova, M.; Rafieva, F.; Khalikov, K.; Erjigitov, D.; Khidirov, M.; et al. Genetic diversity, QTL mapping, and marker-assisted selection technology in cotton (Gossypium spp.). Front. Plant Sci. 2021, 12, 779386. [Google Scholar] [CrossRef] [PubMed]
  8. Altman, D. Exogenous hormone applications at pollination for in vitro and in vivo production of cotton interspecific hybrids. Plant Cell Rep. 1988, 7, 257–261. [Google Scholar] [CrossRef] [PubMed]
  9. Meyer, V. Interspecific cotton breeding. Econ. Bot. 1974, 28, 56–60. [Google Scholar] [CrossRef]
  10. Montes, E.; Coriton, O.; Eber, F.; Huteau, V.; Lacape, J.; Reinhardt, C.; Marais, D.; Hofs, J.; Chevre, A.; Pannetier, C. Assessment of gene flow between Gossypium hirsutum and G. herbaceum: Evidence of unreduced gametes in the diploid progenitor. G3 2017, 7, 2185–2193. [Google Scholar] [CrossRef] [PubMed]
  11. Huang, G.; Wu, Z.; Percy, R.G.; Bai, M.; Li, Y.; Frelichowski, J.E.; Hu, J.; Wang, K.; Yu, J.Z.; Zhu, Y. Genome sequence of Gossypium herbaceum and genome updates of Gossypium arboreum and Gossypium hirsutum provide insights into cotton A-genome evolution. Nat. Genet. 2020, 52, 516–524. [Google Scholar] [CrossRef]
  12. Pershina, L. On the role of distant hybridization and polyploidy in plant evolution. Vogis Bull. 2009, 13, 2. (In Russian) [Google Scholar]
  13. Alam, H.; Razaq, M. Induced polyploidy as a tool for increasing tea (Camellia sinensis L.) production. Northeast Agric. Univ. 2015, 22, 43–47. [Google Scholar] [CrossRef]
  14. Vavilov, N. Doubling the number of chromosomes as a method for obtaining new plant forms. Sel. Work. 1965, 5, 156–168. (In Russian) [Google Scholar]
  15. Kamburova, V.; Salakhutdinov, I.; Shermatov, S.; Buriev, Z.; Abdurakhmonov, I. Cotton as a Model for Polyploidy and Fiber Development Study. In Model Organisms in Plant Genetics; IntechOpen: London, UK, 2021; pp. 1–17. [Google Scholar] [CrossRef]
  16. Bae, S.; Islam, M.; Kim, H.; Lim, K. Induction of tetraploidy in watermelon with oryzalin treatments. Hort. Sci. Technol. 2020, 38, 385–393. [Google Scholar] [CrossRef]
  17. Dasgeb, B.; Kornreich, D.; McGuinn, K.; Okon, L.; Brownell, I.; Sackett, D. Colchicine: An ancient drug with novel applications. Br. J. Derm. 2018, 178, 350–356. [Google Scholar] [CrossRef] [PubMed]
  18. Stanys, V.; Staniene, G.; Siksnianas, T. In vitro induction of ploidy in Ribes. Acta Uni. Latv. Biol. 2004, 676, 235–237. [Google Scholar]
  19. Petersen, K.; Hagberg, P.; Kristiansen, K. Colchicine and oryzalin mediated chromosome doubling in different genotypes of Miscanthus sinensis. Plant Cell Tis. Organ. Cult. 2003, 73, 137–146. [Google Scholar] [CrossRef]
  20. Vainola, A. Polyploidization and early screening of Rhododendron hybrids. Euphytica 2000, 112, 239–244. [Google Scholar] [CrossRef]
  21. Jia, Y.H.; Sun, J.L.; Wang, X.W.; Zhou, Z.L.; Pan, Z.E.; He, S.P.; Pang, B.Y.; Wang, L.R. Molecular diversity and association analysis of drought and salt tolerance in G.hirsutum L. germplasm. J. Integ. Agr. 2013, 13, 1845–1853. [Google Scholar] [CrossRef]
  22. Abdurakhmonov, I.Y. (Ed.) World Cotton Germplasm Resources; IntechOpen: London, UK, 2014; 320p. [Google Scholar] [CrossRef]
  23. Campbell, B.; Saha, S.; Percy, R.; Frelichowski, J.; Jenkins, J.; Park, W.; Mayee, C.; Gotmare, V.; Dessauw, D.; Giband, M.; et al. Status of the global cotton germplasm resources. Crop Sci. 2010, 50, 1161–1179. [Google Scholar] [CrossRef]
  24. Abdurakhmonov, I.; Kohel, R.; Yu, J.; Pepper, A.; Abdullaev, A.; Kushanov, F.; Salakhutdinov, I.; Buriev, Z.; Saha, S.; Scheffler, B.; et al. Molecular diversity and association mapping of fiber quality traits in exotic G.hirsutum L. germoplasm. Genomics 2008, 92, 478–487. [Google Scholar] [CrossRef]
  25. Chen, Q.; Wang, W.; Wang, C.; Zhang, M.; Yu, J.; Zhang, Y.; Yuan, B.; Ding, Y.; Jones, D.C.; Paterson, A.; et al. Validation of QTLs for fiber quality introgressed from Gossypium mustelinum by selective genotyping. G3 2020, 10, 2377–2384. [Google Scholar] [CrossRef]
  26. Jareczek, J.; Grover, C.; Hu, G.; Xiong, X.; Arick, I.; Peterson, D.; Wendel, J. Domestication over speciation in allopolyploid cotton species: A stronger transcriptomic pull. Genes 2023, 14, 1301. [Google Scholar] [CrossRef] [PubMed]
  27. Grover, C.E.; Yoo, M.J.; Lin, M.; Murphy, M.D.; Harker, D.B.; Byers, R.L.; Lipka, A.E.; Hu, G.; Yuan, D.; Conover, J.L.; et al. Genetic analysis of the transition from wild to domesticated cotton (Gossypium hirsutum L.). G3 2020, 10, 731–754. [Google Scholar] [CrossRef] [PubMed]
  28. Jena, S.; Srivastava, A.; Rai, K.; Ranjan, A.; Singh, S.; Nisar, T. Development and characterization of genomic and expressed SSRs for levant cotton (Gossypium herbaceum L.). Theor. App. Genet. 2012, 124, 565–576. [Google Scholar] [CrossRef]
  29. Barroso, P.; Hoffmann, L.; Batista, C.; Freitas, R.; Alves, M.; Silva, U.; Andrade, F. In situ conservation and genetic diversity of three populations of Gossypium mustelinum Miers (exWatt). Genet. Res. Crop Evol. 2010, 57, 343–349. [Google Scholar] [CrossRef]
  30. Wang, B.; Zhuang, Z.; Zhang, Z.; Draye, X.; Shuang, L.-S.; Shehzad, T.; Lubbers, E.L.; Jones, D.; May, O.L.; Paterson, A.H.; et al. Advanced backcross QTL analysis of fiber strength and fineness in a cross between Gossypium hirsutum and G. mustelinum. Front. Plant Sci. 2017, 8, 1848. [Google Scholar] [CrossRef] [PubMed]
  31. Yang, Y.; You, C.; Wang, N.; Wu, M.; Le, Y.; Wang, M.; Zhang, X.; Yu, Y.; Lin, Z. Gossypium mustelinum genome and an introgression population enrich interspecific genetics and breeding in cotton. Theor. Appl. Genet. 2023, 136, 130. [Google Scholar] [CrossRef] [PubMed]
  32. Sun, Y.; Zhang, X.; Nie, Y.; Guo, X.; Jin, S.; Liang, S. Production and characterization of somatic hybrids between Upland cotton (Gossypium hirsutum) and wild cotton (G. klotzschianum Anderss) via electrofusion. Theor. App. Genet. 2004, 109, 472–479. [Google Scholar] [CrossRef]
  33. Yu, J.Z.; Kohel, R.J.; Fang, D.D.; Cho, J.; Van, A.D.; Ulloa, M.; Hoffman, S.M.; Pepper, A.E.; Stelly, D.M.; Jenkins, J.N.; et al. A high-density simple sequence repeat and single nucleotide polymorphism genetic map of the tetraploid cotton genome. G3 2012, 1, 43–58. [Google Scholar] [CrossRef]
  34. Newaskar, G.; Chimote, V.; Mehetre, S.; Jadhav, S. Interspecific hybridization in Gossypium L.: Characterization of progenies with different ploidy-confirmed multigenomic backgrounds. Plant Breed. 2013, 132, 211–216. [Google Scholar] [CrossRef]
  35. Zhang, X.; Zhai, C.; He, L.; Guo, Q.; Zhang, X.; Xu, P.; Su, H.; Gong, Y.; Ni, W.; Xinlian, S. Morphological, cytological and molecular analyses of a synthetic hexaploid derived from an interspecific hybrid between Gossypium hirsutum and Gossypium anomalum. Crop J. 2014, 2, 272–277. [Google Scholar] [CrossRef]
  36. Liu, Q.; Chen, Y.; Chen, Y.; Wang, Y.; Chen, J.; Zhang, T.; Zhang, T.; Zhouet, B. A new synthetic allotetraploid (A1A1G2G+) between Gossypium herbaceum and G. australe: Bridging for simultaneously transferring favorable genes from these two diploid species into Upland cotton. PLoS ONE 2015, 10, e0123209. [Google Scholar]
  37. Chen, Y.; Wang, Y.; Wang, K.; Zhu, X.; Guo, W.; Zhang, T.; Zhou, B. Construction of a complete set of alien chromosome addition lines from Gossypium australe in Gossypium hirsutum: Morphological, cytological, and genotypic characterization. Theor. App. Genet. 2014, 127, 1105–1121. [Google Scholar] [CrossRef] [PubMed]
  38. Chen, Y.; Wang, Y.; Zhao, T.; Yang, J.; Feng, S.; Nazeer, W. A new synthetic amphiploid (AADDAA) between G. hirsutum and G. arboreum lays the foundation for transferring resistances to verticillium and drought. PLoS ONE 2015, 10, e0128981. [Google Scholar] [CrossRef] [PubMed]
  39. Chen, D.; Wu, Y.; Zhang, X.; Li, F. Analysis of [Gossypium capitis-viridis × (G.hirsutum × G.australe)2] trispecific hybrid and selected characteristics. PLoS ONE 2015, 10, e0127023. [Google Scholar] [CrossRef] [PubMed]
  40. Kushanov, F.; Buriev, Z.; Shermatov, S.; Turaev, O.; Norov, T.; Pepper, A.; Saha, S.; Ulloa, M.; Yu, J.; Jenkins, J.; et al. QTL mapping for flowering-time and photoperiod insensitivity of cotton Gossypium darwinii Watt. PLoS ONE 2017, 12, e0186240. [Google Scholar] [CrossRef] [PubMed]
  41. Yin, X.; Zhan, R.; He, Y.; Song, S.; Wang, L.; Ge, Y.; Chen, D. Morphological description of a novel synthetic allotetraploid(A1A1G3G3) of Gossypium herbaceum L. and G. nelsonii Fryx. suitable for disease-resistant breeding applications. PLoS ONE 2020, 15, e0242620. [Google Scholar] [CrossRef]
  42. Dospekhov, B.A. Methodology of Field Experience (with the Basics of Statistical Processing of Research Results); Agropromizdat: Moscow, Russia, 1985; 351p. (In Russian) [Google Scholar]
  43. Wang, B.; Draye, X.; Zhuang, Z.; Zhang, Z.; Liu, M.; Lubbers, E.L.; Jones, D.; May, O.L.; Paterson, A.H.; Chee, P.W. QTL analysis of cotton fiber length in advanced backcross populations derived from a cross between Gossypium hirsutum and G. mustelinum. Theor. Appl. Genet. 2017, 130, 1297–1308. [Google Scholar] [CrossRef]
  44. Wang, B.; Liu, L.; Zhang, D.; Zhuang, Z.; Guo, H.; Qiao, X.; Wei, L.; Rong, J.; May, O.L.; Paterson, A.H.; et al. A genetic map between Gossypium hirsutum and the Brazilian endemic G. mustelinum and its application to QTL mapping. G3 2016, 6, 1673–1685. [Google Scholar] [CrossRef]
  45. Wang, B.; Draye, X.; Zhang, Z.; Zhuang, Z.; May, O.L.; Paterson, A.H.; Chee, P.W. Advanced backcross quantitative trait locus analysis of fiber elongation in a cross between Gossypium hirsutum and G. mustelinum. Crop Sci. 2016, 56, 1760–1768. [Google Scholar] [CrossRef]
  46. Shkutina, F.M. Meiosis in distant hybrids and amphidiploids. In V. Sat. Cytology and Genetics of Meiosis; Khvostova, V.V., Bogdanov, Y.F., Eds.; Publishing house “Nauka”: Moscow, Russia, 1975; pp. 292–311. (In Russian) [Google Scholar]
  47. Kozak, M.F. Mitotic rhythms in representatives of the genus Glycine L. Tsitol Genet. 2004, 6, 7–12. (In Russian) [Google Scholar]
  48. Sanamyan, M.F. Cytogenetic study of hybrids and mutants of cotton. In Abstract Candidate of Biological Sciesnces; Uzbekistan Academy of Sciences: Tashkent, Uzbekistan, 1988; 22p. (In Russian) [Google Scholar]
  49. Blasio, F.; Prieto, P.; Pradillo, M.; Naranjo, T. Genomic and meiotic changes accompanying polyploidization. Plants 2022, 11, 125. [Google Scholar] [CrossRef] [PubMed]
  50. Pausheva, Z.P. Workshop on Cytology; Kolos: Moscow, Russia, 1988; 287p. (In Russian) [Google Scholar]
  51. Paterson, A.H.; Brubaker, C.L.; Wendel, J.F. A rapid method for extraction of cotton (Gossypium spp.) genomic DNA suitable for RFLP or PCR analysis. Plant Mol. Biol. Rep. 1993, 11, 122–127. [Google Scholar] [CrossRef]
  52. Yu, J.; Jung, S.; Cheng, C.; Lee, T.; Zheng, P.; Buble, K.; Crabb, J.; Humann, J.; Hough, H.; Jones, D.; et al. CottonGen: The community database for cotton genomics, genetics, and breeding research. Plants 2021, 10, 2805. [Google Scholar] [CrossRef]
  53. Adylova, A.; Norbekov, G.; Khurshut, E.; Nikitina, E.; Kushanov, F. SSR analysis of the genomic DNA of perspective Uzbek hexaploid winter wheat varieties. Vavilov J. Genet. Breed. 2018, 22, 634–639. [Google Scholar] [CrossRef]
  54. Turaev, O.S.; Baboev, S.K.; Ziyaev, Z.M.; Norbekov, J.K.; Erjigitov, D.S.; Bakhadirov, U.S.; Tursunmurodova, B.T.; Dolimov, A.A.; Turakulov, K.S.; Ernazarova, D.K.; et al. Present status and future perspectives of wheat (Triticum aestivum L.) research in Uzbekistan. SABRAO J. Breed. Genet. 2023, 55, 1463–1475. [Google Scholar] [CrossRef]
  55. Shen, X.; Guo, W.; Zhu, X.; Yu, J.; Kohel, R.; Zhang, T. Molecular mapping of QTLs for fiber qualities in three diverse lines in Upland cotton using SSR markers. Mol. Breed. 2005, 15, 169–181. [Google Scholar] [CrossRef]
  56. Wu, J.; Gutierrez, O.; Jenkins, J.; McCarty, J.; Zhu, J. Quantitative analysis and QTL mapping for agronomic and fiber traits in an RIL population of Upland cotton. Euphytica 2008, 165, 231–245. [Google Scholar] [CrossRef]
  57. Sun, F.D.; Zhang, J.H.; Wang, S.F.; Gong, W.K.; Shi, Y.Z.; Liu, A.Y.; Yuan, Y.L. QTL mapping for fiber quality traits across multiple generations and environments in Upland cotton. Mol. Breed. 2011, 30, 569–582. [Google Scholar] [CrossRef]
  58. Liu, H.; Quampah, A.; Chen, J.; Li, J.; Huang, Z.; He, Q.; Zhu, S. QTL mapping with different genetic systems for nine non-essential amino acids of cottonseeds. Mol. Genet. Genom. 2017, 292, 671–684. [Google Scholar] [CrossRef]
  59. Ma, L.; Zhao, Y.; Wang, Y.; Shang, L.; Hua, J. QTLS analysis and validation for fiber quality traits using maternal backcross population in Upland cotton. Fron. Plant Sci. 2017, 8, 2168. [Google Scholar] [CrossRef]
  60. An, C.; Jenkins, J.N.; Wu, J.; Guo, Y.; McCarty, J.C. Use of fiber and fuzz mutants to detect QTL for yield components, seed, and fiber traits of Upland cotton. Euphytica 2009, 172, 21–34. [Google Scholar] [CrossRef]
  61. Du, L.; Cai, C.; Wu, S.; Zhang, F.; Hou, S.; Guo, W. Evaluation and exploration of favorable QTL alleles for salt stress related traits in cotton cultivars (G. hirsutum L.). PLoS ONE 2016, 11, e0151076. [Google Scholar] [CrossRef]
  62. Abdelraheem, A.; Zhu, Y.; Zhang, J. Quantitative trait locus mapping for fusarium wilt race 4 resistance in a recombinant inbred line population of pima cotton (Gossypium barbadense). Pathogens 2022, 11, 1143. [Google Scholar] [CrossRef] [PubMed]
  63. Liang, Q.; Li, P.; Hu, C.; Hua, H.; Li, Z.; Rong, Y.; Wang, K.; Hua, J. Dynamic QTL and epistasis analysis on seedling root traits in Upland cotton. J. Genet. 2014, 93, 63–78. [Google Scholar] [CrossRef] [PubMed]
  64. Liang, Q.; Hu, C.; Hua, H.; Li, Z.; Hua, J. Construction of a linkage map and QTL mapping for fiber quality traits in Upland cotton (Gossypium hirsutum L.). Chinese Sci. Bull. 2013, 58, 3233–3243. [Google Scholar] [CrossRef]
  65. Shao, Q.; Zhang, F.; Tang, S.; Liu, Y.; Fang, X.; Liu, D.; Zhang, Z. Identifying QTL for fiber quality traits with three Upland cotton (Gossypium hirsutum L.) populations. Euphytica 2014, 198, 43–58. [Google Scholar] [CrossRef]
  66. Shang, L.; Liu, F.; Wang, Y.; Abduweli, A.; Cai, S.; Wang, K.; Hua, J. Dynamic QTL mapping for plant height in Upland cotton (Gossypium hirsutum). Plant Breed. 2015, 134, 703–712. [Google Scholar] [CrossRef]
  67. Wang, F.; Zhang, C.; Liu, G.; Chen, Y.; Zhang, J.; Qiao, Q.; Zhang, J. Phenotypic variation analysis and QTL mapping for cotton (Gossypium hirsutum L.) fiber quality grown in different cotton-producing regions. Euphytica 2016, 211, 169–183. [Google Scholar] [CrossRef]
  68. Wang, B.; Guo, W.; Zhu, X.; Wu, Y.; Huang, N.; Zhang, T. QTL mapping of yiyeld and yiyeld components for elite hybrid derived-RILs in Upland cotton. Genet. Genom. 2007, 34, 35–45. [Google Scholar] [CrossRef]
  69. Shen, X.; Guo, W.; Lu, Q.; Zhu, X.; Yuan, Y.; Zhang, T. Genetic mapping of quantitative trait loci for fiber quality and yield trait by RIL approach in Upland cotton. Euphytica 2007, 155, 371–380. [Google Scholar] [CrossRef]
  70. Ulloa, M.; Wang, C.; Hutmacher, R.; Wright, S.; Davis, R.; Saski, C.; Roberts, P. Mapping Fusarium wilt race 1 resistance genes in cotton by inheritance, QTL and sequencing composition. Mol. Genet. Genet. 2011, 286, 21–36. [Google Scholar] [CrossRef] [PubMed]
  71. Jamshed, M.; Jia, F.; Gong, J.; Palanga, K.; Shi, Y.; Li, J.; Shang, H.; Liu, A.; Chen, T.; Zhang, Z.; et al. Identification of stable quantitative trait loci (QTLs) for fiber quality traits across multiple environments in Gossypium hirsutum recombinant inbred line population. BMC Genomics 2016, 17, 197. [Google Scholar] [CrossRef] [PubMed]
  72. Wang, B.; Wu, Y.; Huang, N.; Zhu, X.; Guo, W.; Zhang, T. QTL Mapping for Plant Architecture Traits in Upland Cotton Using RILs and SSR Markers. Actu Genet. Sinica 2006, 33, 161–170. [Google Scholar] [CrossRef] [PubMed]
  73. Abdelraheem, A.; Kuraparthy, V.; Zhang, J. Identification of drought and salt tolerant cotton germplasm and associated markers in the U.S. Upland germplasm. In Proceedings of the ASA, CSSA and SSSA International Annual Meetings, Phoenix, AZ, USA, 6–9 November 2016; pp. 163–1322. [Google Scholar]
  74. Deng, X.; Gong, J.; Liu, A.; Shi, Y.; Gong, W.; Ge, Q.; Li, J.; Shang, H.; Wu, Y.; Yuan, Y. QTL mapping for fiber quality and yield-related traits across multiple generations in segregating population of CCRI 70. J. Cotton Res. 2019, 2, 13. [Google Scholar] [CrossRef]
  75. Li, H.; Pan, Z.; He, S.; Jia, Y.; Geng, X.; Chen, B.; Wang, L.; Pang, B.; Du, X. QTL mapping of agronomic and economic traits for four F2 populations of Upland cotton. J. Cotton Res. 2021, 4, 3. [Google Scholar] [CrossRef]
  76. Qin, Y.-S.; Liu, R.-Z.; Mei, H.-X.; Zhang, T.-Z.; Guo, W.-Z. QTL Mapping for yield traits in Upland cotton (Gossypium hirsutum L.). Acta Agron. Sin. 2009, 35, 1812–1821. [Google Scholar] [CrossRef]
  77. Yu, J.; Ulloa, M.; Hoffman, S.; Kohel, R.; Pepper, A.; Fang, D.; Percy, R.; Burke, J. Mapping genomic loci for cotton plant architecture, yield components, and fiber properties in an interspecific (Gossypium hirsutum L. × G. barbadense L.) RIL population. Mol. Genet. Genom. 2014, 289, 1347–1367. [Google Scholar] [CrossRef]
  78. Shang, L.; Liang, Q.; Wang, Y.; Wang, X.; Wang, K.; Abduweli, A.; Ma, L.; Cai, S.; Hua, J. Identification of stable QTLs controlling fiber traits properties in multi-environment using recombinant inbred lines in Upland cotton (Gossypium hirsutum L.). Euphytica 2015, 205, 877–888. [Google Scholar] [CrossRef]
  79. Yu, X.; Chu, B.; Liu, R.; Sun, J.; Brian, J.; Wang, H.; Shuijin, Z.; Sun, Y. Characteristics of fertile somatic hybrids of G. hirsutum L. and G. trilobum generated via protoplast fusion. Theor. Appl. Genet. 2012, 125, 1503–1516. [Google Scholar] [CrossRef]
  80. Yu, Y.; Yuan, D.; Liang, S.; Li, X.; Wang, X.; Lin, Z.; Zhang, X. Genome structure of cotton revealed by a genome-wide SSR genetic map constructed from a BC1 population between Gossypium hirsutum and G. barbadense. BMC Genom. 2011, 12, 15. [Google Scholar] [CrossRef]
  81. Zhang, K.; Zhang, J.; Ma, J.; Tang, Y.; Liu, D.; Teng, Z. Genetic mapping and quantitative trait locus analysis of fiber quality traits using a three-parent composite population in Upland cotton (Gossypium hirsutum L.). Mol. Breeding 2012, 29, 335–348. [Google Scholar] [CrossRef]
  82. Zhang, S.; Feng, L.; Xing, L.; Yang, B.; Gao, X.; Zhu, X.; Zhang, T.; Zhou, B.; Jenkins, J. New QTLs for lint percentage and boll weight mined in introgression lines from two feral landraces into Gossypium hirsutum acc TM-1. Plant Breed. 2016, 135, 90–101. [Google Scholar] [CrossRef]
  83. Tan, Z.; Fang, X.; Tang, S.; Zhang, J.; Liu, D.; Teng, Z.; Li, L.; Ni, H.; Zheng, F.; Liu, D.; et al. Genetic map and QTL controlling fiber quality traits in Upland cotton (Gossypium hirsutum L.). Euphytica 2015, 203, 615–628. [Google Scholar] [CrossRef]
  84. Yu, W.; Zhang, K.; Li, S.; Yu, H.; Zhai, M.; Wu, M.; Li, X.; Fan, S.; Song, M.; Yang, D. Mapping quantitative trait loci for lint yield and fiber quality across environments in a Gossypium hirsutum × Gossypium barbadense backcross inbred line population. Theor. Appl. Genet. 2013, 126, 275–287. [Google Scholar] [CrossRef] [PubMed]
  85. Yu, J.; Kohel, R.; Smith, C. The construction of a tetraploid cotton genome wide comprehensive reference map. Genomics 2010, 95, 230–240. [Google Scholar] [CrossRef] [PubMed]
  86. Qin, H.; Guo, W.; Zhang, M.; Zhang, T. QTL mapping of yield and fiber traits based on a four-way cross population in Gossypium hirsutum L. Theor. App. Genet. 2008, 117, 883–894. [Google Scholar] [CrossRef] [PubMed]
  87. Xiao, J.; Wu, K.; Fang, D.; Stelly, D.; Yu, J.; Cantrell, R. New SSR markers for use in cotton (Gossypium spp.). Imp. J. Cotton Sci. 2009, 13, 75–157. [Google Scholar]
  88. Shi, Y.; Zhang, B.; Liu, A.; Li, W.; Li, J.; Lu, Q.; Zhang, Z.; Li, S.; Gong, W.; Shang, H.; et al. Quantitative trait loci analysis of Verticillium wilt resistance in interspecific backcross populations of Gossypium hirsutum × Gossypium barbadense. BMC Genomics 2016, 17, 877. [Google Scholar] [CrossRef]
  89. Wang, C.; Ulloa, M.; Duong, T.; Roberts, P.A. Quantitative trait loci mapping of multiple independent loci for resistance to fusarium oxysporum f.sp. vasinfectum races 1 and 4 in an interspecific cotton population. Phytopathology 2018, 108, 759–767. [Google Scholar] [CrossRef]
  90. Li, S.; Liu, A.; Kong, L.; Gong, J.; Li, J.; Gong, W.; Lu, Q.; Li, P.; Shang, H.; Xiao, X.; et al. QTL mapping and genetic effect of chromosome segment substitution lines with excellent fiber quality from Gossypium hirsutum × Gossypium barbadense. Mol. Genet. Genom. 2019, 294, 1123–1136. [Google Scholar] [CrossRef]
  91. Wang, Y.; Ning, Z.; Hu, Y.; Chen, J.; Zhao, R.; Chen, H.; Ai, N.; Guo, W.; Zhang, T. Molecular mapping of restriction-site associated DNA markers in allotetraploid Upland cotton. PLoS ONE 2015, 10, e0124781. [Google Scholar] [CrossRef] [PubMed]
  92. Nie, X.; Huang, C.; You, C.; Li, W.; Zhao, W.; Shen, C.; Zhang, B.; Wang, H.; Yan, Z.; Dai, B.; et al. Genome-wide SSR-based association mapping for fiber quality in nation-wide Upland cotton inbreed cultivars in China. BMC Genomics 2016, 17, 352. [Google Scholar] [CrossRef] [PubMed]
  93. Muhammad, S.; Anam, T.; Shagufta, P.; Amina, A.; Sumera, A.; Muhammad, R.; Xianliang, S.; Xuezhen, S.; Tianzhen, Z. Population structure, linkage disequilibrium and association mapping for cotton leaf curl disease incidence in cotton (Gossypium hirsutum L.). Research Square 2019. [Google Scholar] [CrossRef]
  94. Gutierrez, O.A.; Robinson, A.F.; Jenkins, J.N.; McCarty, J.C.; Wubben, M.J.; Callahan, F.E.; Nichols, R.L. Identification of QTL regions and SSR markers associated with resistance to reniform nematode in Gossypium barbadense L. accession GB713. Theor. App. Genet. 2010, 122, 271–280. [Google Scholar] [CrossRef] [PubMed]
  95. Huang, C.; Shen, C.; Wen, T.; Gao, B.; Zhu, D.; Li, X.; Ahmed, M.; Li, D.; Lin, Z. SSR-based association mapping of fiber quality in Upland cotton using an eight-way MAGIC population. Mol. Genet. Genom. 2018, 293, 793–805. [Google Scholar] [CrossRef]
  96. Ma, L.; Su, Y.; Wang, Y.; Nie, H.; Cui, Y.; Cheng, C.; Wang, M.; Hua, J. QTL and genetic analysis controlling fiber quality traits using paternal backcross population in Upland Cotton. J. Cotton Res. 2019, 3, 22. [Google Scholar] [CrossRef]
  97. Zhao, L.; Lv, D.; Cai, C.; Tong, C.; Chen, D.; Zhang, W.; Du, H.; Guo, H.; Guo, Z. Toward allotetraploid cotton genome assembly: Integration of a high-density molecular genetic linkage map with DNA sequence information. BMC Genomics 2012, 13, 539. [Google Scholar] [CrossRef]
  98. Han, Z.; Wang, C.; Song, X.; Guo, W.; Gou, J.; Li, C.; Chen, X.; Zhang, T. Characteristics, development and mapping of Gossypium hirsutum derived EST-SSRs in allotetraploid cotton. Theor. App. Genet. 2006, 112, 430–439. [Google Scholar] [CrossRef]
Figure 1. The scheme of obtaining cotton amphidiploids.
Figure 1. The scheme of obtaining cotton amphidiploids.
Plants 12 04184 g001
Figure 2. (♀)—diploid G. herbaceum subsp. frutescens; (♂)—tetraploid G. mustelinum; F1C—allohexaploid hybrid progeny of the first generation of G. herbaceum subsp. frutescens × G. mustelinum.
Figure 2. (♀)—diploid G. herbaceum subsp. frutescens; (♂)—tetraploid G. mustelinum; F1C—allohexaploid hybrid progeny of the first generation of G. herbaceum subsp. frutescens × G. mustelinum.
Plants 12 04184 g002
Figure 3. Conjugation of chromosomes at metaphase-I stage of meiosis: (a) F1C G. herbaceum subsp. frutescens × G. mustelinum 35II + 2IV; (b) F1C G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum 38II + 2I (arrows indicate quadrivalents and univalents).
Figure 3. Conjugation of chromosomes at metaphase-I stage of meiosis: (a) F1C G. herbaceum subsp. frutescens × G. mustelinum 35II + 2IV; (b) F1C G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum 38II + 2I (arrows indicate quadrivalents and univalents).
Plants 12 04184 g003
Figure 4. Conjugation of chromosomes at metaphase-I stage of meiosis—F2C G. herbaceum subsp. frutescens × G. mustelinum (a)—24II + 4I (arrows indicate univalents); (b)—26II.
Figure 4. Conjugation of chromosomes at metaphase-I stage of meiosis—F2C G. herbaceum subsp. frutescens × G. mustelinum (a)—24II + 4I (arrows indicate univalents); (b)—26II.
Plants 12 04184 g004
Figure 5. Microscopic images of tetrads in the example of F1C G. herbaceum subsp. frutescens × G. mustelinum hybrid: (a)—normal tetrads; (b)—micronuclear tetrads (arrows indicate micronuclei); (c)—octad; (d)—heptad (at 40× magnification).
Figure 5. Microscopic images of tetrads in the example of F1C G. herbaceum subsp. frutescens × G. mustelinum hybrid: (a)—normal tetrads; (b)—micronuclear tetrads (arrows indicate micronuclei); (c)—octad; (d)—heptad (at 40× magnification).
Plants 12 04184 g005
Figure 6. The results of pollen fertility analysis: (a)—F1C G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum 63.89 ± 3.55; (b)—F1C G. herbaceum subsp. frutescens × G. mustelinum 32.67 ± 3.55 (at 40× magnification).
Figure 6. The results of pollen fertility analysis: (a)—F1C G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum 63.89 ± 3.55; (b)—F1C G. herbaceum subsp. frutescens × G. mustelinum 32.67 ± 3.55 (at 40× magnification).
Plants 12 04184 g006
Figure 7. The results of pollen fertility analysis (F2C G. herbaceum subsp. frutescens × G. mustelinum): (a) T 1-4—40.43%; (b) T 37-4—50.2%; (c) T 49-2—1.5%; (d) T 37-14—7.9%.
Figure 7. The results of pollen fertility analysis (F2C G. herbaceum subsp. frutescens × G. mustelinum): (a) T 1-4—40.43%; (b) T 37-4—50.2%; (c) T 49-2—1.5%; (d) T 37-14—7.9%.
Plants 12 04184 g007
Figure 8. Genetic polymorphisms and genomic changes of the cotton samples using SSR markers. (a) NAU1458 marker; (b) NAU1093 marker; (c) BNL2634 marker; (d) BNL3140 marker. M—Molecular weight marker (base pairs, bp), Plants 12 04184 i001—lack alleles, Plants 12 04184 i002—additional alleles. 1—G. mustelinum; 2—G. herbaceum subsp. frutescens; 3—F1C (G. herbaceum subsp. frutescens × G. mustelinum); 4—G. herbaceum subsp. pseudoarboreum f. harga; 5—F1C (G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum); 6—G. herbaceum subsp. africanum; 7—F1C (G. herbaceum subsp. africanum × G. mustelinum).
Figure 8. Genetic polymorphisms and genomic changes of the cotton samples using SSR markers. (a) NAU1458 marker; (b) NAU1093 marker; (c) BNL2634 marker; (d) BNL3140 marker. M—Molecular weight marker (base pairs, bp), Plants 12 04184 i001—lack alleles, Plants 12 04184 i002—additional alleles. 1—G. mustelinum; 2—G. herbaceum subsp. frutescens; 3—F1C (G. herbaceum subsp. frutescens × G. mustelinum); 4—G. herbaceum subsp. pseudoarboreum f. harga; 5—F1C (G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum); 6—G. herbaceum subsp. africanum; 7—F1C (G. herbaceum subsp. africanum × G. mustelinum).
Plants 12 04184 g008
Figure 9. The phylogenetic tree of cotton species and their interspecific alloploid hybrids using 57 polymorphic SSR markers. With the phylogenetic genealogy method, the arithmetic values of the genetic regions of the cotton species and their interspecific polyploid hybrids were identified using the Neighbor-Joining method of the PAUP 4.0 (Phylogenetic Analysis Using Parsimony and other methods) program.
Figure 9. The phylogenetic tree of cotton species and their interspecific alloploid hybrids using 57 polymorphic SSR markers. With the phylogenetic genealogy method, the arithmetic values of the genetic regions of the cotton species and their interspecific polyploid hybrids were identified using the Neighbor-Joining method of the PAUP 4.0 (Phylogenetic Analysis Using Parsimony and other methods) program.
Plants 12 04184 g009
Table 1. The fertility and hybrid boll-setting rates in interspecific crosses between G. herbaceum and G. mustelinum species.
Table 1. The fertility and hybrid boll-setting rates in interspecific crosses between G. herbaceum and G. mustelinum species.
No.Hybrid CombinationNumber of CrossesNumber of Obtained Hybrid BollsFertilization Rate, % Boll Setting Rate, %
X ± SxRangeSV
1.G. herbaceum subsp. africanum × G. mustelinum621524.210.9 ± 1.07.7–18.23.0628.09
2.G. herbaceum subsp. pseudoarboreum × G. mustelinum59813.810.0 ± 1.94.3–23.15.958.7
3.G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum651524.69.7 ± 1.15.9–15.43.3934.95
4.G. herbaceum subsp. frutescens × G. mustelinum635892.213.1 ± 3.24.0–41.710.1277.25
5.G. herbaceum subsp. euherbaceum (A-833) × G. mustelinum623354.516.3 ± 1.710.0–25.05.4133.18
Note: S stands for the relative error of the results of a series of x measurements at a confidence level of 95%; and V stands for the coefficient of variation, as described by Dospekhov [42].
Table 2. The number of bolls per parental plant and the fertility of cotton seeds.
Table 2. The number of bolls per parental plant and the fertility of cotton seeds.
No.Plant SamplesNumber of BollsNumber of Seeds Per Boll Percentage of Complete Seeds Per Boll, %
Per PlantAnalyzed TotalCompleteEmpty x ¯ ± S x ¯ RangeSV %
Parental Lines
1.G. herbaceum subsp. frutescens211025.423.51.993 ± 284–1006.176.7
2.G. herbaceum subsp. pseudoarboreum201019.518.01.592 ± 285–1006.006.5
3.G. herbaceum subsp. pseudoarboreum f. harga361018.716.72.089 ± 280–1005.906.6
4.G. herbaceum subsp. africanum271018.517.51.095 ± 284–1005.706.0
5.G. herbaceum subsp. euherbaceum A-833281017.615.32.387 ± 282–1006.087.0
6.G. mustelinum Miers ex Watt161027.123.23.986 ± 2 77–96 6.628
Note: S stands for the relative error of the results of a series of x measurements at a confidence level of 95%; and V stands for the coefficient of variation, as described by Dospekhov [42].
Table 3. The number of bolls per hybrid plant of F1C and F2C polyploids.
Table 3. The number of bolls per hybrid plant of F1C and F2C polyploids.
No.Hybrid CombinationNumber of BollsNumber of Seeds Per Boll Percentage of Complete Seeds Per Boll, %
Per PlantAnalyzedTotalCompleteEmpty x ¯ ± S x ¯ RangeSV %
F1C polyploid hybrids
1.G. herbaceum subsp. frutescens × G. mustelinum211014.32.811.519 ± 28–276.1331.6
2.G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum11927----
3.G. herbaceum subsp. africanum × G. mustelinum---------
F2C polyploid hybrids
4.T 51-13 G. herbaceum subsp. frutescens × G. mustelinum261022.97.015.930 ± 125–363.7512.3
5.T 30-8 G. herbaceum subsp. frutescens × G. mustelinum241027.211.615.643 ± 138–462.505.9
6.T 38-3 G. herbaceum subsp. frutescens × G. mustelinum211026.09.516.536 ± 131–433.6910.3
7.T 30-6 G. herbaceum subsp. frutescens × G. mustelinum191018.79.29.549 ± 241–616.6013.4
8.T 38-12 G. herbaceum subsp. frutescens × G. mustelinum181022.87.815.034 ± 128–393.7811.1
Note: S stands for the relative error of the results of a series of x measurements at a confidence level of 95%; and V stands for the coefficient of variation, as described by Dospekhov [42].
Table 4. Heredity and variability of the vegetative growth duration in F1C and F2C—polyploid hybrids and parental lines.
Table 4. Heredity and variability of the vegetative growth duration in F1C and F2C—polyploid hybrids and parental lines.
No.Plant SamplesVegetative Growth Duration, Day
Range x ¯ ± S x ¯ SV %hp
Parental lines
1.G. herbaceum subsp. frutescens116–119117.2 ± 0.41.31.1-
2.G. herbaceum subsp. pseudoarboreum118–122119.6 ± 0.41.51.3-
3.G. herbaceum subsp. pseudoarboreum f. harga126–129127.2 ± 0.41.31.0-
4.G. herbaceum subsp. africanum135–139137.4 ± 0.41.81.3-
5.G. herbaceum subsp. euherbaceum A-833116–119117.4 ± 0.41.31.1-
6.G. mustelinum150–153151.8 ± 0.41.30.9-
F1C polyploid hybrids
7.G. herbaceum subsp. frutescens × G. mustelinum125–133128.6 ± 0.92.82.30.34
8.G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum137–142139.6 ± 0.72.11.5−0.01
9.G. herbaceum subsp. africanum × G. mustelinum142–150145.2 ± 1.33.12.2−0.08
F2C polyploid hybrids
10.G. herbaceum subsp. frutescens × G. mustelinum81 plants
n = 4
115–119116.6 ± 0.51.81.6-
120–124121.8 ± 0.72.11.7-
125–129127.8 ± 0.51.61.3-
130–134131.4 ± 0.62.01.5-
135–139136.8 ± 0.72.11.5-
140–144141.6 ± 0.51.51.1-
145–149147.2 ± 0.51.61.1-
150–155152.2 ± 0.82.61.7-
Note: S stands for the relative error of the results of a series of x measurements at a confidence level of 95%; and V stands for the coefficient of variation, as described by Dospekhov [42].
Table 5. Fiber length and strength in F1C and F2C—polyploid hybrids and parental lines.
Table 5. Fiber length and strength in F1C and F2C—polyploid hybrids and parental lines.
No.Plant SamplesFiber Length, mmFiber Strength, cN/tex
Range x ¯ ± S x ¯ SV %hp
Parental lines
1.G. herbaceum subsp. frutescens16.0–18.017.2 ± 0.20.633.7-35.5
2.G. herbaceum subsp. pseudoarboreum24.0–27.025.1 ± 0.41.204.8-
3.G. herbaceum subsp. pseudoarboreum f. harga19.0–21.020.1 ± 0.20.743.7-21.4
4.G. herbaceum subsp. africanum24.0–26.024.7 ± 0.30.823.3-32.6
5.G. herbaceum subsp. euherbaceum A-83324.0–26.025.2 ± 0.30.793.1-27.2
6.G. mustelinum25.0–27.025.6 ± 0.20.702.7-17.3
F1C polyploid hybrids
7.G. herbaceum subsp. frutescens × G. mustelinum25.0–32.028.7 ± 0.90.913.11.74-
F2C polyploid hybrids
8.G. herbaceum subsp. frutescens × G. mustelinum23.1–41.029.7 ± 1.23.311.057.3-
Note: S stands for the relative error of the results of a series of x measurements at a confidence level of 95%; and V stands for the coefficient of variation, as described by Dospekhov [42].
Table 6. Conjugation of cotton chromosomes at metaphase-I stage of meiosis.
Table 6. Conjugation of cotton chromosomes at metaphase-I stage of meiosis.
No.Hybrid CombinationsThe Number of Studied Maternal Anther Cells Average Number Per Cell
UnivalentsBivalentsQuadrivalents
F1C polyploid hybrids
1.G. herbaceum subsp. frutescens × G. mustelinum220.45 ± 0.2938.22 ± 0.250.36 ± 0.14
2.G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum151.73 ± 0.8537.53 ± 0.440.46 ± 0.24
3.G. herbaceum subsp. africanum × G. mustelinum161.81 ± 0.8437.56 ± 0.430.50 ± 0.24
F2C polyploid hybrids
4.T 37-14 G. herbaceum subsp. frutescens × G. mustelinum301.46 ± 0.5524.73 ± 0.290.26 ± 0.09
5.T 38-3 G. herbaceum subsp. frutescens × G. mustelinum23-26.00 ± 0.00-
6.T 37-4 G. herbaceum subsp. frutescens × G. mustelinum280.64 ± 0.2538.18 ± 0.230.25 ± 0.11
7.T 49-2 G. herbaceum subsp. frutescens × G. mustelinum220.54 ± 0.2 738.00 ± 0.2 70.27 ± 0.13
8.T 30-8 G. herbaceum subsp. frutescens × G. mustelinum18-26.00 ± 0.00-
Table 7. Cytological observation of F1C generation polyploid hybrids.
Table 7. Cytological observation of F1C generation polyploid hybrids.
No.Hybrid CombinationsTotal Number of SporesMeiotic
Index, %
Micronuclear Tetrads, %Polyads, %
F1C polyploid hybrids
1.G. herbaceum subsp. frutescens × G. mustelinum61290.36 ± 1.193.27 ± 0.726.37 ± 0.98
2.G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum21996.80 ± 1.18-3.19 ± 1.18
3.G. herbaceum subsp. africanum × G. mustelinum48196.05 ± 0.880.83 ± 0.413.12 ± 0.79
F2C polyploid hybrids
4.T 1-8 G. herbaceum subsp. frutescens × G. mustelinum41196.15 ± 0.120.89 ± 0.413.07 ± 0.72
5.T 38-12 G. herbaceum subsp. frutescens × G. mustelinum32597.05 ± 0.64-2.62 ± 0.29
6.T 30-6 G. herbaceum subsp. frutescens × G. mustelinum35497.21 ± 0.47-2.78 ± 0.71
7.T 51-13 G. herbaceum subsp. frutescens × G. mustelinum29897.45 ± 0.880.69 ± 0.232.11 ± 0.41
8.T 30-7 G. herbaceum subsp. frutescens × G. mustelinum34198.51 ± 0.48-1.48 ± 0.44
9.T 30-8 G. herbaceum subsp. frutescens × G. mustelinum40596.05 ± 0.88-3.12 ± 0.79
10.T 38-3 G. herbaceum subsp. frutescens × G. mustelinum35798.54 ± 0.81-1.42 ± 0.38
Table 8. Development of cotton pollens in F1C polyploid hybrids.
Table 8. Development of cotton pollens in F1C polyploid hybrids.
No.Hybrid CombinationsTotal Number of PollensPollen Fertility, %
1.G. herbaceum subsp. frutescens × G. mustelinum32432.7 ± 3.6
2.G. herbaceum subsp. pseudoarboreum f. harga × G. mustelinum43263.9 ± 3.6
3.G. herbaceum subsp. africanum × G. mustelinum5410
Table 9. Analysis of pollen quality in amphidiploid hybrids of F2C G. herbaceum subsp. frutescens × G. mustelinum.
Table 9. Analysis of pollen quality in amphidiploid hybrids of F2C G. herbaceum subsp. frutescens × G. mustelinum.
No.SampleTotal Number of PollensPollen Fertility, %No.SampleTotal Number of PollensPollen Fertility, %
1.T 1-13119933.9 ± 1.913.T 30-8295555.7 ± 0.8
2.T 1-470040.4 ± 3.414.T 37-448250.2 ± 5.2
3.T 1-848251.0 ± 5.215.T 37-879623.1 ± 2.2
4.T 30-2353827.1 ± 3.716.T 38-1284660.9 ± 2.8
5.T 30-666660.2 ± 3.617.T 38-3131567.8 ± 1.7
6.T 30-773865.9 ± 3.018.T 38-436738.4 ± 6.4
7.T 37-141147.9 ± 6.419.T 38-735210.2± 2.6
8.T 30-126722.9 ± 6.620.T 42-215914.5 ± 7.8
9.T 33-269046.8 ± 3.621.T 49-26061.5 ± 0.2
10.T 33-62485.6 ± 21.122.T 51-1542634.3 ± 5.3
11.T 36-17615.8 ± 17.523.T 51-1936325.1 ± 5.2
12.T 37-222817.5 ± 6.324.T 51-13153651.8 ± 1.6
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

Khidirov, M.T.; Ernazarova, D.K.; Rafieva, F.U.; Ernazarova, Z.A.; Toshpulatov, A.K.; Umarov, R.F.; Kholova, M.D.; Oripova, B.B.; Kudratova, M.K.; Gapparov, B.M.; et al. Genomic and Cytogenetic Analysis of Synthetic Polyploids between Diploid and Tetraploid Cotton (Gossypium) Species. Plants 2023, 12, 4184. https://0-doi-org.brum.beds.ac.uk/10.3390/plants12244184

AMA Style

Khidirov MT, Ernazarova DK, Rafieva FU, Ernazarova ZA, Toshpulatov AK, Umarov RF, Kholova MD, Oripova BB, Kudratova MK, Gapparov BM, et al. Genomic and Cytogenetic Analysis of Synthetic Polyploids between Diploid and Tetraploid Cotton (Gossypium) Species. Plants. 2023; 12(24):4184. https://0-doi-org.brum.beds.ac.uk/10.3390/plants12244184

Chicago/Turabian Style

Khidirov, Mukhammad T., Dilrabo K. Ernazarova, Feruza U. Rafieva, Ziraatkhan A. Ernazarova, Abdulqahhor Kh. Toshpulatov, Ramziddin F. Umarov, Madina D. Kholova, Barno B. Oripova, Mukhlisa K. Kudratova, Bunyod M. Gapparov, and et al. 2023. "Genomic and Cytogenetic Analysis of Synthetic Polyploids between Diploid and Tetraploid Cotton (Gossypium) Species" Plants 12, no. 24: 4184. https://0-doi-org.brum.beds.ac.uk/10.3390/plants12244184

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