Next Article in Journal
Hypoxia, but Not Normoxia, Reduces Effects of Resveratrol on Cisplatin Treatment in A2780 Ovarian Cancer Cells: A Challenge for Resveratrol Use in Anticancer Adjuvant Cisplatin Therapy
Next Article in Special Issue
Comparative Genomics and Functional Genomics Analysis in Plants
Previous Article in Journal
Effects of Semaglutide and Empagliflozin on Inflammatory Markers in Patients with Type 2 Diabetes
Previous Article in Special Issue
Insight into the Organization of the B10v3 Cucumber Genome by Integration of Biological and Bioinformatic Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analysis of Genome Structure and Its Variations in Potato Cultivars Grown in Russia

by
Dmitry I. Karetnikov
1,2,3,
Gennady V. Vasiliev
1,2,
Stepan V. Toshchakov
4,
Nikolay A. Shmakov
1,2,3,
Mikhail A. Genaev
1,2,
Mikhail A. Nesterov
1,2,
Salmaz M. Ibragimova
1,2,
Daniil A. Rybakov
5,
Tatjana A. Gavrilenko
5,
Elena A. Salina
1,2,6,
Maxim V. Patrushev
4,
Alex V. Kochetov
1,3,6 and
Dmitry A. Afonnikov
1,2,3,*
1
Federal Research Center Institute of Cytology and Genetics SB RAS, 630090 Novosibirsk, Russia
2
Kurchatov Center for Genome Research, Institute of Cytology and Genetics SB RAS, 630090 Novosibirsk, Russia
3
Faculty of Natural Sciences, Novosibirsk State University, 630090 Novosibirsk, Russia
4
Kurchatov Center for Genome Research, National Research Center Kurchatov Institute, 123182 Moscow, Russia
5
Federal Research Center the N.I. Vavilov All-Russian Institute of Plant Genetic Resources (VIR), 190000 St. Petersburg, Russia
6
Faculty of Agronomy, Novosibirsk State Agrarian University, 630039 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(6), 5713; https://doi.org/10.3390/ijms24065713
Submission received: 14 February 2023 / Revised: 10 March 2023 / Accepted: 13 March 2023 / Published: 16 March 2023
(This article belongs to the Special Issue Comparative Genomics and Functional Genomics Analysis in Plants)

Abstract

:
Solanum tuberosum L. (common potato) is one of the most important crops produced almost all over the world. Genomic sequences of potato opens the way for studying the molecular variations related to diversification. We performed a reconstruction of genomic sequences for 15 tetraploid potato cultivars grown in Russia using short reads. Protein-coding genes were identified; conserved and variable parts of pan-genome and the repertoire of the NBS-LRR genes were characterized. For comparison, we used additional genomic sequences for twelve South American potato accessions, performed analysis of genetic diversity, and identified the copy number variations (CNVs) in two these groups of potato. Genomes of Russian potato cultivars were more homogeneous by CNV characteristics and have smaller maximum deletion size in comparison with South American ones. Genes with different CNV occurrences in two these groups of potato accessions were identified. We revealed genes of immune/abiotic stress response, transport and five genes related to tuberization and photoperiod control among them. Four genes related to tuberization and photoperiod were investigated in potatoes previously (phytochrome A among them). A novel gene, homologous to the poly(ADP-ribose) glycohydrolase (PARG) of Arabidopsis, was identified that may be involved in circadian rhythm control and contribute to the acclimatization processes of Russian potato cultivars.

1. Introduction

Common potato (Solanum tuberosum L.) is one the most important crops grown worldwide. It is ranked the first highest produced non-cereal food crop and the fourth highest produced crop worldwide after wheat, corn, and rice [1]. Potato is produced as food and animal feed, and they are also grown for industrial purposes. Potatoes contribute key nutrients to the human diet including vitamin C, potassium, and dietary fiber [2].
The latest taxonomical treatment of Spooner et al. [3,4] recognized four cultivated potato species—S. tuberosum, divided into Chilean and Andean cultivar groups, and three hybrid cultivated species of “bitter potato”, S. ajanhuiri Juz. and Bukasov, S. curtilobum Juz. and Bukasov, S. juzepczukii Buk.). Thus, S. tuberosum is represented by native tetraploid cultivars (landraces) grown in the lowlands of Chile (S. tuberosum Chilotanum group) and by populations of di-, tri-, and tetraploid landraces grown in the highlands of the Andes (S. tuberosum Andigenum group). Continuing breeding has led to the development of thousands of improved cultivars; in most cases, modern improved cultivars are the product of interspecific crosses with different cultivated and wild potato species [4,5]. However, there is still a need worldwide to develop new cultivars with desirable and more effective properties.
Modern potato cultivars are autotetraploid with tetrasomic inheritance, high heterozygosity, frequent pollen sterility, and clonal method of reproduction [6]. The latter factor causes dysfunctional and deleterious alleles not to be removed during meiosis, leading to inbred depression. These factors cause complications in potato breeding [7]. In this regard, genome sequencing and analysis provide the basis for efficient research in the field of potato genetics and breeding.
The potato genome of the homozygous DM1-3 double monoploid was sequenced and assembled by the Potato Genome Sequencing Consortium [8]. Subsequently, additional DM1-3 pseudomolecules with improved sequence and annotation quality were assembled and annotated [9,10,11]. Extensive CVNs were found, affecting more than 30% of the potato genome [11]. Lately, a high-quality haplotype-resolved assembly to the chromosomes of the diploid [12,13] and autotetraploid potato cultivar (cv.) S. tuberosum “Otava” genome [14] was obtained using long reads and the chromosome conformation capture method. Recently the phased S. tuberosum genomes of several commercial cultivars from North America and Europe were sequenced and assembled [15]. Using long reads, an assembly was obtained for high-quality diploid potato genomes from more than 40 wild and cultivated representatives of the Solanum section Petota (including diploid Andean landraces) [6]. Phased potato genome assemblies allow the characterization of the potato pan-genome, the study of global rearrangements at the chromosome level between haplotypes and opens up new possibilities for directed breeding in the future.
In addition to high quality assemblies based on long reads, methods of genomic sequence reconstruction based on short reads are actively used. Recently, genomic sequences for six accessions of cultivated potato polyploid species were assembled, three of which were tetraploid landraces belonging to the Andigenum group and to the Chilotanum group and two of them were sequenced only by short reads [16]. Nucleotide and structural variations were analyzed by short-read sequencing in various potato species and landraces from South America [17], elite tetraploid cultivars [18], potato somatic hybrid, its parents and progeny [19], diploid potato clones [20]. These data made it possible to estimate such structural characteristics of the potato genomes as the set of protein-coding genes [16,19], characterize their functions [16], estimate copy number variations, single nucleotide, and small insertion–deletion polymorphisms [17,18,19,20].
Copy number variations play an important role in crop domestication and diversification [21]. In potato genomes, they provide a major contribution to the genomic diversity of clonally propagated potatoes, as well as affecting species-specific and dispensable groups of pan-genome genes [11]. Using identified CNVs, it is possible to analyze their contribution to the genetic diversity of wild and cultivated species accessions, as well as identify clusters of genes that are affected by CNVs in potato genomes (such as SAUR, gene clusters, gene clusters of metabolite biosynthesis, etc.) [17].
The aim of our work is to study the diversity and variation of genomic sequences of potato cultivars grown in Russia and to search for genes that could participate in their diversification and acclimatization through copy number changes based on CNV comparison for Russian potato cultivars and South American accessions sequenced recently. We reconstructed genome sequences using short read sequencing for 15 potato-improved cultivars, 14 developed by different Russian breeding programs, and 1 Dutch variety cultivated in Russia (further, we will call this subset Russian cultivars). We identified the conserved and variable parts of the pan-genome and estimated the functions associated with them, assessing the diversity of NBS-LRR genes. Using additional sequence data of 12 genomes of South American potato species accessions represented by 1 wild ancestor species S. bukasovii and 11 accessions of Andean and Chilean cultivated species, we performed a comparative analysis of CNVs, which allowed us to identify structural gene variations with different occurrence in these 2 subsets. Functional analysis of these genes was performed, and it was shown that they are associated with the response to abiotic and biotic stresses. A number of genes from this pool have been identified, which are associated with tuberization and control of circadian rhythms.

2. Results

2.1. Genome Assembly Statistics and Annotation for Russian Cultivars

The number of reads in our libraries varies from 60 million to 290 million, yielding the sequence coverage of the S. tuberosum group phureja DM1-3 reference potato genome from 23× for cv. Udacha to 109× for cv. Grand (Table S1, Supplementary File S1). The proportion of paired-end reads for all cultivars after preprocessing was more than 87% (Table S1, Supplementary File S1).
Table 1 shows the main statistics for the genomes of Russian potato cultivars. The number of contigs ranged from 196 kb (cv. Grand) to 551 kb (cv. Udacha), and the proportion of contigs smaller than 1000 bp without an open reading frame was from 35% (cv. Udacha) to 80% (cv. Fritella). The total lengths of assembled and filtered contigs exceeded the size of the DM1-3 reference genome, 810 Mb for all but one accession (cv. Udacha, 653 Mb) (Table 1). The GC content of the genomic sequences varies from 34.83% (cv. Nikulinsky) to 36.22% (cv. Gusar), the average for all genomes being 35.55% (34.84% for DM1-3). The contigs of largest length range from 163 Mb (cv. Krasa Meshchery) to 281 Mb (cv. Krepysh); N50 varies from 8095 bp (cv. Gusar) to 15,505 bp (cv. Grand) and shows a positive relationship with the coverage value. The average proportion of identified repeats is 61%. The statistics (Table 1) indicate the fragmentation of our assemblies is high.
The results of BUSCO analysis (Figure S1, Supplementary File S1) demonstrated that of Solanales dataset proteins more than 60% are present in all genomes of Russian potato cultivars completely and in a single copy. The fraction of duplicated proteins varies from 2.5% for cv. Udacha to 13% for cv. Fritella. The fraction of fragmented variants vary from 6% for cv. Grand to 12% for cv. Udacha. The fraction of missed proteins varies from 15% for cv. Zhukovsky to 24% for cv. Gusar.

2.2. Protein Orthogroups Analysis

The results of the identification of orthologous groups for protein-coding genes for 15 Russian potato cultivars, and 12 South American accessions from ref. [17], DM1-3 reference genome and the tomato Solanum lycopersicum genome (outgroup) are presented in Table 2. The number of open reading frames (ORFs) for Russian cultivars varies from 60,411 (cv. Udacha) to 77,417 (cv. Krasa Meshchery), this value is higher for genomes with high coverage (for DM1-3 39,021 ORFs are known). More than 90% of the proteins in all South American accessions, Russian cultivars, and reference genomes belong to common orthologous groups.
The total number of orthologous groups identified is 125,744, of which 84,450 groups have more than 1 sequence and include 2,117,217 sequences. The remaining 41,294 amino acid sequences (2% of the total number of sequences and 33% of the total number of orthogroups) represent unassigned (single-sequence) orthogroups. The average number of sequences in the orthologous group is 25.1, with a median of 9.0. The G50 metric (the number of genes in the orthogroup such that 50% of genes are in orthogroups of that size or larger) is 56 sequences, and the O50 metric (the smallest number of orthogroups such that 50% of genes are in orthogroups of that size or larger) is 9220. Only seven orthologous groups include exactly one sequence from each genome.

2.3. Analysis of ORFs in Russian Cultivars

In the group of 15 genomes of potato cultivars grown in Russia, 103,748 orthologous groups were identified. Genes in these orthogroups were classified into “core”, “softcore”, “shell” and “cloud” types by their occurrence in different accessions (see Methods, Section 4.6). The fraction of genes of these types in pan-genome and each genome is shown in Figure 1a. Of the 1,050,536 genes annotated, 465,278 (44.5% of the total number) belong to the “core” part; for individual cultivars, the proportion of this type of gene varies from 42% to 45.5%. The number of genes belonging to “shell” orthogroups is 399,045 genes (37.6%). The number of genes belonging to “softcore” orthogroups is 154,283 (15.1%) and the number of genes belonging to “cloud” orthogroups is 31,930 (2.8%). According to the ratio of genes in these categories of orthogroups, the Russian cultivars are similar, despite the difference in genomic sequence coverage by reads.
The distribution of orthogroups by the represented genome number is shown in Figure 1b. Orthogroups that contain genes from 2 to 13 genomes (“shell” genes) prevail (51,138 orthogroups). The second largest category is orthogroups represented in a single genome (“cloud”, 30,516 orthogroups). Next, are 14,992 orthogroups containing genes in all 15 genomes (“core”), and 7102 orthogroups represented in at least 14 genomes (“softcore”). The difference between fractions of “core”, “softcore”, “shell”, and “cloud” parts for genes (Figure 1a) and orthogroups (Figure 1b) can be explained by several homologs of genes in one orthogroup representing the same genome.
Using the data on the number of genes in the orthologous groups, we estimated the change in the size of the pan-genome of Russian potato cultivars and its conserved part (Figure 2). The plot demonstrates that the size of the potato pan-genome does not reach the plateau for the number of genomes we investigated.
We performed functional annotation of protein-coding genes in Russian cultivars. The number and percentage of annotated ORFs are shown in Table 3: the proportion of annotated sequences for all cultivars exceeds 53%, and the maximum fraction of annotated genes is observed for cv. Grand (59.04%), the minimum is observed for cv. Gusar (53.18%).
The proportion of annotated genes in the “core” part of the pan-genome (66.1%) is only slightly higher than in the “softcore” (61.5%), one and a half times higher than in the “shell” (41.4%), and almost three times higher than in the “cloud” (22.5%). We identified the 15 most frequent functional domains in the pan-genome and estimated their frequencies in different parts of the pan-genome. The results are shown in Figure S2 (Supplementary File S1) and indicate that domains such as cytochrome P450, protein kinases and tyrosine kinases, RNA-recognition motif, F-box, and PPR repeats have a higher frequency of occurrence in the conserved part of the pan-genome (“core”). Such domains as NB-ARCs, LRR motifs, integrases, gag domains, and reverse transcriptases are characterized by higher frequency in the variable part of the pan-genome (“cloud”, “shell”).

2.4. Analysis of NBS-LRR Genes

We identified 3404 full-length proteins of the NBS-LRR family in the amino acid sequences of Russian potato cultivars. Based on the co-occurrence in the orthogroups with known NLR-proteins of the reference genome DM1-3, NBS-LRR classes were identified for 1270 sequences. The remaining 2134 proteins fall into orthogroups with no NLR-proteins from the reference DM1-3 and thus were not attributed to particular NLR classes. Table 4 shows the number of NLR genes of each class predicted in the genome of an individual cultivar. The proportion of proteins of different classes among all NBS-LRR proteins is shown in Figure S3 (Supplementary File S1).
The highest number of NBS-LRR proteins (306) was identified for the cv. Severnoe Siyanie, the lowest (135), for the cv. Udacha (Table 4). The number of NBS-LRR correlates well with the coverage of the genome by reads (the higher the coverage, the more proteins identified). Four cultivars with the highest number of NBS-LRR genes (from 281 to 306) have coverage greater than 100× (cvs. Krasavchik, Grand, Fritella, Severnoe Siyanie). Four cultivars with the lowest number of NBS-LRR genes, from 135 to 182 (cvs. Udacha, Sudarinya, Zhukovsky, Gusar), have low coverage (23× to 48×). On average, there are 227 NLR proteins per genome. The most represented classes of the NBS-LRR proteins are CNL-1 (267 proteins in all 15 genomes), CNL-7 (251 proteins), and CNL-R (198 proteins); the least represented class is CNL-4, which contains a total of 11 genes in all 15 genomes.
Our data indicate a high diversity of NBS-LRR genes in the 15 Russian potato cultivars we studied. This is evidenced by the proportion of proteins that were assigned to orthogroups that did not include proteins from the DM1-3 genome (60% and higher) (Figure S3, Supplementary File S1). Clearly, they represent the variable part of the immune response proteins.

2.5. Genetic Differentiation among Improved Russian Cultivars and South American Landrace Accessions

To study genetic differentiation and genome diversity of cultivated potatoes (Table 2), 1012 orthologous groups were used by OrthoFinder, in which at least 51.7% of amino acid sequences represented a single copy for any of the genotypes. The resulting tree of genetic differentiation is shown in Figure 3.
One can see from Figure 3 that Russian potato cultivars form a cluster (support value 1) whose members are separated from the studied South American landraces. This cluster separates into several clusters of smaller sizes. The level of support for the cluster of South American landrace genotypes (S. tuberosum Andigenum group and S. tuberosum Chilotanum group) is also high (0.999). At the root of the cluster is the wild ancestor species S. bukasovii Juz. (=Solanum candolleanum Berthault). The remaining accessions of three highlands “bitter” cultivated potato species (S. juzepczukii, S. curtilobum, and S. ahanhuiri) form the group at the root of all other potato genotypes. These three species are of hybrid origin derived from natural crosses between wild species S. acaule, S. boliviense and cultivated species distributed in the high Andean altiplano between Southern Peru and Central Bolivia, at elevations between 3600 and 4400 m [22,23]. Thus, characteristic of the tree is the location of three “bitter” potato species from Andean highlands at the root of the tree and the presence of two sister groups of Russian cultivars and nine South American potato accessions.

2.6. CNV Characteristics for Russian Cultivars and South American Landrace Potato Accessions

Bioinformatics analysis using CNVpytor v0.4.1 allowed CNV identification and analysis in Russian potato cultivars, and South American landrace potato accessions and compare their characteristics. The distribution of CNVs for different accessions by number, maximal length, and related gene numbers is shown in Figure 4 and Table S2 (Supplementary File S1).
Figure 4a shows that the number of deletions for most genotypes exceeds the number of duplications (except for South American CUR and JUZ accessions). Russian cultivars compared to South American ones are more homogeneous in this parameter: the number of duplications varies near 10,000 except for the cv. Grand. The number of deletions varies by about 20,000 without significant deviations for any cultivars. At the same time, sharp differences in the number of duplications are observed for the South American samples: their number is ~5000 or less for STN, PHU, GON2, and GON1; high values of the number of duplications (more than 14,000) for BUK, AJH, CUR, and JUZ; moderate values, comparable to those of Russian cultivars for CHA, TBR, ADG2, and ADG1. The number of deletions approximately corresponds to that in Russian cultivars, except for the aforementioned CUR and JUZ (the lowest values among all genotypes). Interestingly, the CUR, JUZ, and AJH accessions (the number of deletions is greater than the number of duplications or close to it) belong to the group of highland “bitter” cultivated potato species located at the root of tree in Figure 3.
The distribution of the maximum size of the duplicated or deleted segment demonstrates a high homogeneity of Russian cultivars and diversity of South American accessions (Figure 4b). The maximum lengths of deletions/duplications for Russian cultivars are close to ~100,000 bp, for some cultivars the maximum length of deletions is slightly greater than the number of duplications and for others vice versa. The length of structural rearrangements is somewhat less in the cv. Grand, and cv. Nevsky is characterized by a significant excess of the maximum length of deletions (400,000 bp) over the maximum length of duplications (~100,000 bp). The Russian cultivars also demonstrate uniformity in the average size of rearrangements, close to 9000 bp except for the mentioned cv. Grand (value ~6500 bp).
A number of South American accessions are characterized by a significantly longer length of the deleted segment compared to the duplicated one. Thus, for BUK wild species accession it is over 800,000 bp, while for PHU, CUR, CHA, JUZ, TBR, and ADG2 accessions it is over 500,000 bp. At the same time, for the GON2 genotype, this value is comparable to that of Russian cultivars, while for GON1 and ADG2 it is only two times higher. The maximum length of duplications for the South American genotypes does not differ significantly from the Russian ones (not more than 2 times, maximum 175,000 bp for JUZ), but compared to the Russian cultivars, the variability of South American landraces in this parameter is higher.
While the number of deletions in the genomes we studied is in most cases higher than duplications, they affect a smaller number of protein-coding genes (ORFs) compared to duplications (Figure 4c). Interestingly, the genotypes we studied are rather homogeneous in the number of genes affected by deletions: there are no remarkable differences between the Russian and South American accessions. In terms of the number of genes affected by duplications, on the contrary, the variability is high, also among Russian cultivars. The low number of duplications (less than 5000 bp) affects genes in cvs. Grand, Symphonia, and diploid cultivated species accessions from South America—STN, PHU, GON2, GON1. A high value is observed in both Russian cultivars (Golubizna, Udacha, Krasavchik, Sudarynia, Gusar) and South American accessions (BUK, AJH, CUR, JUZ, TBR). The remaining accessions show a moderate number of gene duplications.
We performed statistical tests for a difference in means and the equality of two variances for CNV characteristics from Figure 4 in Russian cultivars and South American accessions (Table S3, Supplementary File S1). The results demonstrated that the mean values differ significantly for one CNV characteristic (maximum deletion size, Figure 4b). Variances are unequal for all characteristics except maximum duplication size (Figure 4b). These results support the hypothesis that Russian cultivars compared to South American ones are more homogeneous in most CNV parameters.
The distribution of CNV on chromosomes of both accessions from Russia and South America is shown in Figures S4 and S5 (Supplementary File S1). These figures demonstrate that CNVs are distributed unevenly both along the genome and between accessions and that CNV distribution patterns on chromosomes in Russian and South American cultivars differ. For example, for chromosome 1, all studied cultivars are characterized by a low number of duplications at the 25-30 Mb region (white pattern in Figure S4). At the same time, a number of Russian cultivars are characterized by the less frequent occurrence of CNV at 65-90 Mb (lighter shade): cv. Krasa Meshchery, cv. Grand, cv. Symphonia, cv. Nikulinsky. The same differences are observed for these cultivars on the second chromosome, and in general, they show fewer duplications compared to the remaining cultivars from Russia (in agreement with Figure 4a). Diploid Andean landraces of STN, PHU, GON2, and GON1 show similar differences relative to the rest of the South American accessions. The density of CNV duplications for them is lower on almost all chromosomes than for the rest of the South American polyploid landraces. This is also consistent with the data in Figure 4a.
A number of the chromosome segments corresponding to the differences between Russian and South American accessions are noticeable. For example, the duplications on chromosome 7: Russian cultivars are depleted in them around 25 Mb, while for South American landraces this is not observed (Figure S4, Supplementary File S1). The lower density of duplications is also characteristic of Russian cultivars in the central parts of chromosomes 11 and 12.
Figure S5 (Supplementary File S1) demonstrates no clearly distinguishable accessions that were enriched with deletions in either Russian or South American accessions (which is also consistent with Figure 4a). At the same time, a number of genome segments are noticeable in which there are differences between Russian and South American accessions. For example, the frequency of deletions in Russian cultivars is lower than in South American accessions in chromosome 10. This is also characteristic of chromosomes 3 and 9. In contrast, an increased frequency of deletions on chromosome 7 in a number of Russian cultivars is observed within the region of 0–30 Mb; a number of South American accessions are depleted of deletions in this region.
Thus, our analysis demonstrates the higher homogeneity of Russian cultivars in terms of genome structural rearrangements compared to South American cultivated potato accessions. These structural variations are characterized by the prevalence of deletions over duplications with respect to the whole genome and vice versa, duplications over deletions, in the regions containing protein-coding genes. In addition, there is great diversity in the distribution of CNVs on the chromosomes of the accessions we studied.

2.7. CNV-Based Similarity of Potato Genomes

We compared the potato genomes we studied by the similarity of CNV frequencies in the loci encoding proteins based on principal component analysis as proposed in [17]. The total number of DM1-3 v4.03 reference genome genes affected by CNV in 15 cultivars from Russia and 12 South American cultivars associated with CNV was 38,310.
The principal component analysis shows that the first component accounts for 21%, the second 8%, and the third 5% of the total variance. The bivariate PCA diagrams for the projections to the first three components are shown in Figure 5. The diagram for the two principal components (Figure 5A) shows that the first component is characterized by the division of all samples into the species of hybrid origin from the high Andes (JUZ, CUR, AJH) and all others, with strong overlapping values for the rest of the genomes of the South American accessions and Russian cultivars in this component. Interestingly, the Andean diploid landraces STN, PHU, GON1, GON2 and Russian cvs. Symphonia, Grand, Nikulinsky, Krasa Meshchery have the lowest values of the first component and the lowest number of duplications (Figure 4a). At the same time, the highest values of this component are characteristic of South American allopolyploid cultivated species (CUR, JUZ) and diploid cultivated species S. ahanhuiri (AJH) having hybrid origin with the highest number of duplications (Figure 4a). Apparently, the first component in Figure 5a is likely related to the number of duplications.
The second component, on the contrary, is associated with the separation between genomes of analyzed Russian and South American potato accessions. It is interesting that in contrast to the tree in Figure 3 in the PC1/PC2 diagram Chilean S. tuberosum (TBR) is closer to the Russian cultivars than to South American Andean landraces. On the other hand, the PC2/PC3 diagram (Figure 5b) shows that this genotype is quite far from the Russian cultivars in the third component.
It should also be noted that the samples corresponding to the Russian cultivars in these graphs form a separate cloud. At the same time, among the South American accessions, several separate groups can be distinguished (Figure 5a). The first one mentioned above, JUZ, CUR, and AJH, corresponds to the “bitter” cultivated species accessions located at the root of the tree diagram of potato genotypes (Figure 3). The second one includes of Andean diploid landraces (GON2, GON1, PHU, STN); Andean cultivated tetraploid species (ADG1, ADG2) are located closer to them, all together they belong to the S. tuberosum Andigenum group. Accessions of Andean-cultivated triploid S. chaucha (CHA) and wild ancestor species S. bukasovii (BUK) are further away, and TBR is located close to Russian cultivars. Note that the location of accessions in the PCA diagram only at the level of large clusters corresponds to their genetic differentiation (Figure 3), while in the details (at closer distances) there are differences.
We also represented the similarity of analyzed potato genomes by CNV distribution in the form of the tree shown in Figure 6. Its structure is consistent with the diagrams obtained from the principal component analysis: the tree contains a cluster of AJH, JUZ, and CUR, which is joined by South American BUK and TBR genomes. The second cluster of genomes of South American accessions in this tree corresponds to the S. tuberosum Andigenum accessions located in the PCA1/PCA2 diagram in the upper left corner (Figure 5a): ADG1, ADG2, GON1, GON2, PHU, STN with CHA joining them. The Russian cultivars on this tree belong to mostly star-like branches from the main stem of the tree.

2.8. Comparison of the CNVs Occurrence in Genomes of South American Cultivated Species and Russian Cultivars

As noted above, the tree of genetic differentiation of potato genomes (Figure 3) is characterized by the presence of two sister clusters of Russian cultivars (including reference gene DM1-3) and South American cultivated species accessions (except for JUZ, CUR, and AJH). This allows us to search for loci that have different CNV occurrences in genomes from these two sister clusters of potatoes. We additionally excluded from the South American cluster the BUK genome, which represents a wild type and forms a long branch from the root of this cluster (shown in magenta in Figure 3). Therefore, we analyzed occurrences of CNVs related to protein-coding genes in genomes of 15 Russian cultivars and 8 South American cultivated species accessions.
Each CNV locus associated with a protein-coding gene in a particular genome was characterized by three CNV types according to our analysis: significant duplications (+1), significant deletions (-1), and without any significant CNV (0) (see Materials and Methods, Section 4.8). To identify loci in which occurrence of CNV of these types unevenly distributed between Russian and South American cultivars, we performed Fisher’s exact test (2 × 3 contingency table analysis). As a result, we identified 1742 genes for which this test showed significance at the p-value < 0.01 (Supplementary File S2).
We further narrowed the above list of genes by distinguishing specific patterns of CNV occurrence between genomes of studied Russian cultivars and South American accessions of several types (see Materials and Methods, Section 4.8): RUDup (duplications in Russian cultivars); SADup (duplications in South American landraces); RUDel (deletions in Russian cultivars); SADel (deletions in South American landraces). We identified 236 genes for RUDup, 33 for SADup, 109 for RUDel, and 219 for SADel CNV patterns (Supplementary File S2).
Enrichment analysis for the entire list of significant genes using the DAVID yielded 10 significant functional clusters (FDR < 0.05; Supplementary File S2). They represent the following terms: NB-ARC/plant defense/leucine-rich repeat (cluster 1); poly(ADP-ribose) glycohydrolase activity (cluster 2); replication factor A (cluster 3); vesicle-mediated transport/vacuolar sorting protein 39 (cluster 4); leucine-rich repeat (cluster 5); aspartic-type endopeptidase activity (cluster 6); voltage-gated anion channel activity/porin domain (cluster 7); FBD domain (cluster 8); RNA methylation (cluster 9); TIR/signal transduction (cluster 10).
No significant functional clusters were found for SADup and RUDel patterns (Supplementary File S2). For the group of 236 genes with the RUDup pattern, the following functional annotation clusters were identified: vesicle-mediated transport/vacuolar sorting protein 39 (cluster 1); RNA methylation (cluster 2); voltage-gated anion channel activity/porin (cluster 3); fucose metabolic process (cluster 4); GTPase activity/guanyl nucleotide binding (cluster 5); methyltransferase activity (cluster 6); isomerase/peptidyl-prolyl cis-trans isomerase activity (cluster 7); carbohydrate-binding/isomerase/aldose 1-epimerase activity (cluster 8); KDPG/KHG aldolase (cluster 9).
Four functional gene annotation clusters were found for the SADel pattern (Supplementary File S2): polyphenol oxidase (cluster 1); ascorbate biosynthesis/inositol oxygenase activity (cluster 2); peptidase M76, ATP23 (cluster 3); translation elongation factor EF1A (cluster 4).

2.9. Comparison of the CNVs Occurrence in Genomes of South American and Russian Potato Accessions

Genes associated with tuberization and photoperiod control in potatoes are actively studied in their relationship to adaptation to the long day typical for European latitudes [4,24]. To evaluate the influence of CNV on these genes in genomes of Russian and South American accessions, we searched for genes associated with tuberization and photoperiod control (Supplementary File S3) among the genes with significant differences in CNV occurrence (Supplementary File S2). We found such differences for four genes (Table 5).
In addition to the genes identified previously as associated with tuberization and photoperiod control in potatoes (Supplementary File S3), we found another gene, Poly(ADP-ribose) glycohydrolase, which demonstrated different CNV occurrence in genomes of Russian and South American accessions analyzed in the present study (Table 5). We identified significant duplications of this gene for all 15 Russian cultivars, and only for 3 South American accessions (CHA, TBR, ADG2). The homolog of this gene in Arabidopsis, PARG, is involved in circadian rhythm control [27].
According to Table 5, deletions are common for tuberization and photoperiod control genes in South American potato accessions (four genes out of five). In Russian cultivars, a remarkable number of deletions, 4 accessions out of fifteen, is observed only for gene PGSC0003DMG400012838, while in South American cultivars the number of accessions with deletions for this gene is seven out of eight. Other genes in Table 5 are characterized by the absence of any frequent deletions in Russian genomes and by the remarkable proportion of duplications (8 of 15 for the gene PGSC0003DMG400015766 and 15 of 15 for the gene PGSC0003DMG400029361).

2.10. CNVs in SAUR Gene Clusters

Previously, CNV analysis in potatoes demonstrated the presence of a large number of structural variations in the loci associated with SAUR (small auxin-up RNA) gene clusters involved in auxin signaling [11]. These are gene clusters located on chromosomes 1 (86.97-87.17 Mb), 4 (54.17-54.37 Mb), 6 (56.29-56.49 Mb), and 11 (0.87-1.11 Mb) [17]. Our analysis showed that these loci are enriched in CNV not only in South American landraces but also in Russian cultivars (Figures S6–S9, Supplementary File S1). The cluster of SAUR genes on chromosome 1 demonstrates the greatest CNV enrichment (Figure S6, Supplementary File S1). On chromosome 4, the dominance of duplications in Russian cultivars over deletions and over CNV of both types in South American accessions (Figure S7, Supplementary File S1) was observed. However, we observed significant differences in CNV occurrence in the South American and Russian accessions only for two SAUR genes. The first gene PGSC0003DMG400016568 (auxin-induced SAUR) located on chromosome 4 at ~54.187 Mb (Supplementary File S2, see also Figure S7, Supplementary File S1) and affected by deletions in four South American potato genomes and is not affected by any CNV in Russian genomes. The second gene is PGSC0003DMG400021565 (SAUR family protein), for which 10 duplications in Russian cultivars and no significant CNVs in South American potato accessions are observed (Supplementary File S2). Interestingly, this gene does not belong to any of the mentioned clusters and is located on chromosome 9 at position ~54.81 Mb. Thus, despite the presence of a large number of CNVs in the SAUR cluster regions, their specificity with respect to Russian or South American accessions is weakly expressed.

3. Discussion

3.1. Genome Assemblies

We used short-read sequencing to analyze the genomic sequences of 15 potato cultivars grown in Russia. Analysis of the quality of their genome assemblies showed that the N50 values for our assemblies are slightly higher than that obtained in assemblies of South American potato accessions based on Illumina sequencing [16]. L50 values in our assemblies are lower than for genomes of South American accessions, also indicating the generally longer contigs that we were able to assemble. Given higher coverage for a number of genomes of Russian cultivars, the results can be considered generally comparable. In terms of the number of identified ORFs, our results are also consistent with the analysis of the South American accessions [16]. We identified about 80% of the genes from the BUSCO proteins (approximately the same proportion for all accessions), which is also comparable to the results for South American polyploid landraces [16]. Note that for the four haplotypes of the tetraploid cv. Otava, the number of identified ORFs was 153,000 [14] and the BUSCO completeness score was 97.3%. Six tetraploid genomes of European potato cultivars yielded the number of gene models ranged from 103,000 to 180,000 [15], which is comparable to the number of identified genes in the cv. Otava genome. On average, the number of identified protein-coding genes for the cv. Otava tetraploid genome assembly turns out to be about twice as high as our assemblies. This comparison demonstrates that protein-coding genes are underrepresented in our assemblies. This implies that the information about the functional characteristics and the abundance of proteins of different functional classes could be obtained only roughly from our data.

3.2. Protein Coding Genes Identification and Analysis

Short reads do not allow accurate reconstruction of the genomic sequences, which makes it difficult to estimate the conserved and variable parts of the pan-genome. For our assemblies, core and softcore genes account for approximately 60% of individual genomes, shell ~38%, and cloud ~2%. Assembling and analyzing six genomes of commercial potato cultivars with haplotype resolution resulted in twice the estimate of the proportion of core genes, ~80% [15], shell genes comprise 19%, and cloud 1–2%. Thus, our results underestimate the fraction of core genes in comparison with full tetraploid assemblies [6,15]. The difference in the two estimates is most likely due to the fact that our analysis did not allow separating similar genes belonging to different homologous chromosomes/haplotypes of potatoes.
Analysis of the pan-genome for our cultivars demonstrated its size did not reach a plateau. On the one hand, this can be explained by the small number of genomes studied: for diploid potato species, a plateau is reached when the number of genomes is close to 40 [6]. On the other hand, it can be partially explained by the high complexity of tetraploid potato genomes by gene content, expression, and function [15].
Despite the indicated inaccuracies, the general picture of the distribution of gene functions in core/shell parts of the pan-genome is similar to the results of plant pan-genome analysis from other works [6,15,28,29,30]: the variable part is more associated with genes of immune response and response to environmental stress conditions, the conserved part is primarily related to basic genome functions. In our work, the conservative part of the pan-genome of cultivars grown in Russia is enriched in genes involved in plant growth and development (PPR repeats, RNA recognition motifs, etc.), while the more variable part of the pan-genome is enriched in genes associated with plant immunity.

3.3. Diversity of the NBS-LRR Genes

R-genes mediate the plant response to various pathogens and pests. The diversity of immune response genes in potatoes is the basis for the development of new improved cultivars resistant to biotic stresses [31,32,33]. Reconstruction of potato genomes and transcriptomes reveals sequences encoding R-genes for further study [34,35]. The average number of complete NBS-LRR proteins in Russian potato cultivar genomes is ~230. Note that in the DM1-3 genome, 257 NBS-LRR genes were identified [34,36]. Estimates of the number of NBS domains alone for several dozen potato cultivars based on nucleotide marker alignment are about twice that number, 575–590 [35]. Due to the incompleteness of the genomic sequences of Russian potato cultivars, our estimates on the number of proteins belonging to different classes of NBS-LRR proteins determined for the reference DM1-3 genome are imprecise (Table 4). Some proteins belonging to specific classes are missed. For example, no proteins were detected at all for some classes in several Russian potato cultivars. However, we detected sequences dissimilar with annotated NBS-LRR proteins from DM1-3 (not falling into corresponding orthogroup clusters). Clearly, they represent variable parts of the NBS-LRR proteins in Russian potato cultivars and their fraction is quite large (~60%). This may indicate that potato accessions contain a diverse pool of R-genes, a significant proportion of which are unique to specific genomes, which is consistent with our results obtained earlier [34].

3.4. Study of Genetic Differentiation and Diversity of Cultivated Potato Genomes

The analysis of the genetic similarity between Russian and South American accessions showed that three Andean highlands “bitter” species (JUZ, CUR, AJH) separated from other accessions used in our study. This is consistent with the results of [3] demonstrated separation of these species into a separate cluster in the tree based on SSRs for landraces of all cultivated species and closely related wild species progenitors. In the tree obtained by [3], species S. juzepczuki, S. curtilobum, and S. ajanhuiri lie at the tree root, in relation to other representatives of cultivated potatoes. Specimens from these three species form separate clusters in the tree for potato germplasm collections at the Vavilov Institute of Plant Genetic Resources [37]. Thus, in agreement with [3,37], our results support the distinctive nature of hybrid species S. curtilobum, S. juzepczukii, and S. ajanhuiri within the group of cultivated potatoes [22].
The tree (Figure 3) also clearly distinguishes two sister clusters: the representatives of the S. tuberosum Andigenum group and the analyzed subset of improved potato cultivars grown in Russia. One possible reason for the separation of Russian potato cultivars and native South American accessions may be the involvement of Mexican wild species S. demissum and/or S. stoloniferum in the lineages of almost all Russian cultivars analyzed in this work (see data on molecular markers and references to the literature in the Supplementary File S4). This grouping is in agreement with the results of the phylogenetic tree of potatoes and its relatives (wild species, landraces, and improved cultivars) reconstructed using SNPs [38]. Note that in a diverse subset of Russian cultivars, we did not observe genetic differentiation (Figure 3) corresponding to their pedigrees (see references in the Supplementary File S4). Notably, in Figure 3 two clusters can be distinguished which include cultivars with contrasting characters in terms of their adaptability. One is the cluster of cvs. Zhukovsky, Nevsky, and Udacha (they hav been released in 10, 12, and 9 regions of Russian Federation, respectively, demonstrating a wide range of adaptability). The second cluster includes cvs. Symphonia, Severnoe Siyanie, and Fritella (released in 1, 1, and 2 regions, respectively, demonstrating narrow range of adaptability). This may reflect the genetic similarity between some cultivars with similar ecological diversification patterns.

3.5. CNVs Characteristics of Potato Cultivars

In our work, within subset of South American specimens there is only one of them representing a Chilean cultivar (TBR) [17]. This allowed us to evaluate the diversity and characteristics of CNV in Russian potato cultivars and to compare them with those of South American landraces genomes. The data were in agreement with previous results of potato CNV analysis [11,17,18,19,20,39]. The number of deletions is greater than the number of duplications in most of the accessions we studied (except for South American accessions of “bitter” cultivated species CUR and JUZ), including all cultivars from Russia. At the same time, the number of genes that underwent deletion is generally less than the number of duplicated genes (except South American diploid Andean landraces of STN, PHU, GON2, and GON1). This was also observed for South American landraces in [17], in the analysis of potato somatic hybrid, parents and progeny [19], diploid potato clones [20], and one of three potato cultivars of the Ural selection [39].
For the Russian potato cvs. Grand and Symphonia, the number of deleted genes is less than the number of duplicated genes slightly. In general, cultivars from Russia are more homogeneous in the characteristics of deletions and duplications, which seems reasonable, due to their genetic similarity (Figure 3). A similar pattern of relatively small variations in CNV was observed in the analysis of commercial tetraploid potato cultivars [18]. We also showed that CNV enrichment was observed in SAUR gene loci in both Russian and South American accessions, which is also consistent with earlier results [11,17].
The grouping of potato genotypes derived from the analysis of the two principal components derived from CNV (Figure 5 and Figure 6) is also consistent with the results of [17]: South American landraces of hybrid highland cultivated species AJH, CUR, JUZ are distant from other accessions that belong to S. tuberosum Andigenum group. Diploid Andean landrace accessions (diploids of S. tuberosum Andigenum group) GON1, GON2, STN, and PHU form a dense cluster, and tetraploid Andean landraces (ADG1 and ADG2) (tetraploids of S. tuberosum Andigenum group) are located next to each other. On the large scale, the PCA results are consistent with the genetic differentiation of these accessions (Figure 3): AJH, JUZ, and CUR accessions of ‘bitter’ potatoes group together; accessions of Andigenum Group (GON1, GON2, PHU, STN, CHA, ADG1, ADG2) fall into the second large cluster of South American accessions; Russian cultivars form the third large, dispersed group of accessions. It should be noted that these trees were obtained by different approaches (protein sequence similarity vs. CNV similarity) and differ by details (especially in relative positions of different accessions on the tree within groups of Russian and South American accessions). For example, in the PCA diagram, the TBR genotype is located close to the Russian potato cultivars, while they are distant on the genetic similarity tree. However, the clustering of this TBR accession (sample CIP 705053) with commercial potato cultivars having hybrid origin might be expected, because this TBR accession possesses the W-type plastome [40] which is characteristic for wild potato species and very seldom in potato landraces [41]. Whereas another cytoplasm type T (with 241 bp specific plastid deletion) is typical for Chilean landrace populations [41,42,43]. So, we can suppose that TBR accession (sample CIP 705053) is not a native Chilean landrace, it can be represented by an interspecific hybrid genotype. Interestingly, genetic analysis of the diversity and relatedness in an Andean potato collection from Argentina demonstrated, that some accessions classified as Andean landraces, consistently clustered with commercial cultivars supporting the hypothesis that they were, in fact, reintroductions of European-bred potato cultivars [44].

3.6. Patterns of CNV Occurrence in Russian and South American Cultivars

We distinguished the genes that are significantly different in CNV occurrence in genomes of Russian and South American accessions and investigated their function. In addition, we identified genes that have a predominance of CNV of the same type (duplications/deletions) in either Russian cultivars or South American landraces. This analysis allows us to evaluate the contribution of CNVs to the rate of diversification of improved cultivars grown in Russia and South American native cultivars (landraces) during their breeding process [21].
Functional analysis of the genes having different CNV frequencies in Russian and South American accessions showed several functional clusters of terms. Interestingly, some of them were also identified in the analysis of CNV patterns associated with duplications in Russian cultivars (RUDup). According to the similarity of the annotation terms, these clusters can be divided into three groups. First of all, these are plant immune response genes containing NB-ARC, LRR, and TIR domains (clusters 1,5,10). This is not surprising since immune response receptors in potatoes show a high level of diversity [19,34,35,45], which is also consistent with the results of the NBS-LRR sequence analysis (Table 4). A specific set of combinations of genes of resistance to particular pathogens could have been formed during the selection of these cultivars. Thus, this group of genes reflects the great diversity of immune response genes for South American and Russian potato accessions.
The second group represents genes involved in transport processes. These are vesicle-mediated transport/vacuolar sorting as well as voltage-gated anion channel activity, porin domain (clusters 4, 7). Vacuoles are known to serve as accumulators of secondary plant metabolites such as alkaloids, phenolic compounds, xenobiotics, etc. [46,47]. Porins are involved in the exchange of ions and small molecules across the mitochondrial outer membrane and engaged in complex interactions driving many facets of cell function [48]. Probably, the differences in CNV associated with these functions reflect the differences in the accumulation and transport of various metabolites in South American and Russian cultivars subsets.
The third group of genes (clusters 2, 3, 6, 8, 9) can be described by the general response to abiotic stresses and affecting other important processes in plants. For example, aspartic-type endopeptidases (cluster 6) have been associated with plant defense response mechanisms, hybrid sterility, reproductive development, abiotic and biotic stresses, chloroplast homeostasis, and lateral root development [48]. It has been shown that segmental and tandem duplications are characteristic of genes in this group in the potato genome, and the expression of approximately 21% of these genes changes under salt, osmotic, or temperature treatments [49]. Genes from the 3rd functional cluster (replication factor A) are also shown to change expression associated with endoplasmic reticulum stress and potential involvement in genotoxic stress responses [50]. Genes from cluster 9 are involved in RNA methylation processes and related to the development and abiotic stress processes in plants [51]. Genes from cluster 8 (FBD domain) are also involved in plant development and stress response, ubiquitinylation processes [52]. The cluster 2 gene (Poly(ADP-ribose) glycohydrolase, gene ID PGSC0003DMG400029361), whose homolog in Arabidopsis (PARG) regulates immune gene expression and defense responses [53], can also be assigned to this group. In plants, unlike animals, these genes are present in the genome in multiple copies and are involved in a wide range of important biological processes [54]. These genes are also involved in the response to a number of abiotic stresses [54,55,56,57]. In general, the variability of the third group of genes is consistent with the processes of diversification of cultivated potato accessions and their adaptation to new growing conditions.

3.7. CNV Differences in Genes Related to Tuberization/Photoperiod

The presence of CNVs in genes associated with photoperiod in crops is an important factor in their diversification during adaptation to new climatic zones [21]. CNVs of a number of genes have been shown to affect flowering in wheat [58], and barley [59], flowering and heading time in winter wheat [60,61] photoperiod sensitivity in bread wheat [62], heading time in durum wheat [63].
Our analysis showed that genomes of Russian cultivars adapted to the long-day high northern latitudes and short-day adapted Andean landraces differ in their CNV occurrence for 4 of 48 known genes related to tuber formation and response to photoperiod changes (Table 5). One of them is phytochrome A, which is involved in circadian clock control in potatoes [26] and potentially involved in resistance to adverse factors [64], three were identified as differentially expressed in two tetraploid cultivars with short and long tuberization times [25].
In addition, we detected the poly(ADP-ribose) glycohydrolase gene (PARG, gene ID PGSC0003DMG400029361) (Table 5) located on chromosome 12. We identified significant duplications of this gene for all 15 Russian tetraploid cultivars, and only for 3 South American polyploid landraces (CHA, TBR, ADG2). Experiments show that poly(ADP-ribose) glycohydrolases in Arabidopsis thaliana play an important role in the regulation of circadian oscillator [27]: an inhibitor of PARG shortened the period length of wild-type plants. Interestingly, the function of this gene in potatoes in relation to the control of circadian rhythms has not been previously reported.
Our results reflect the importance of CNV in the adaptation of European and, in particular, Russian potato cultivars to the conditions of a longer photoperiod characteristic of higher latitudes in Europe and Russia [4,65]. However, we may suppose that the CNVs are not common for the most tuberization and photoperiod control genes (Supplementary File S3), and their variability seems to be described mostly by the accumulation of nucleotide replacements or short insertions/deletions in the genes themselves [66,67] or their upstream regions [15].
Thus, the comparative analysis of CNV occurrence in the genomes South American and Russian potato cultivars shows that structural variations are closely related to the processes of diversification in potato cultivars. First of all, these are well-known processes for cultivated plants: immune response, stress response, and processes of transport of metabolites and ions in cells. In addition, there are a number of specific processes, important for potato adaptation, such as the control of photoperiod or/and tuberization.

4. Materials and Methods

4.1. Plant Material

We analyzed the genomes of 15 potato cultivars grown in Russia. They include a subset of 14 improved (contemporary) potato cultivars with diverse breeding backgrounds developed in Russia (Supplementary File S4) [37,68,69]. The plants of 10 cultivars were obtained from the in vitro collection maintained at the N.I. Vavilov All-Russian Institute of Plant Genetic Resources (VIR), Saint Petersburg, Russia, and plants of the rest 4 cultivars were obtained from the in vitro collection GeneAgro of the Institute of Cytology and Genetics SB RAS, Novosibirsk, Russia. Names, details of pedigree (parental genotypes), information on the year of release, and place of origin (Breeding Centre or Company) are listed in Supplementary File S4. This file also provides links to external resources with more information on genotyping data (as DOI for the corresponding articles). These 14 cultivars previously were genotyped using nSSR fingerprinting; in addition, they were screened with 15 DNA markers associated with 10 R-genes conferring resistance to diseases and pests and their cytoplasm types were established using the commonly used set of organelle DNA-specific primers (Supplementary File S4).
These cultivars represent breeding programs conducted in different breeding centers located in the central part of Russian Federation, Moscow region (Russian Potato Research Centre; LLC Agrocenter “Korenevo”) and in the north-west part of Russian Federation, Saint Petersburg region (Leningrad Research Agriculture Institute — Branch of Russian Potato Research Centre; LLC Selection firm “LIGA”). Fourteen Russian cultivars represent different maturity types and ranges of use (table, starch, and processing potato cultivars). They were selected to represent a wide range of genetic diversity and origin that had been identified based on previous results of molecular studies and pedigree information (Supplementary File S4).
These cultivars can be divided into three groups according to their cytoplasm types which are corresponded to their maternal parentage. Group 1: four cultivars (Grand, Gusar, Meteor, Sudarinya) have W/gamma type cytoplasm that is typical for interspecific hybrids with wild Mexican species S. stoloniferum × S. tuberosum. Group 2: seven cultivars (Fritella, Krasa Meshchery, Krepysh, Nevsky, Nikulinsky, Udacha, Zhukovsky ranny) have the D-type cytoplasm as known from the results of molecular analysis; the D-type cytoplasm is typical for wild Mexican species S. demissum. Group 3: three cultivars (Golubizna, Krasavchik, Severnoe siyanie)— typical for Chilean S. tuberosum cytoplasm type T-W wild species were involved in their pedigrees as paternal parents.
In addition to 14 Russian cultivars, 1 foreign Dutch cv. Symphonia was also involved in our study. This cultivar is characterized by high resistance to wart disease, potato cyst nematode, and moderate resistance to scab, but is susceptible to late blight. It is actively used in breeding and genetic research in Russia [70,71,72,73,74]. We use term “Russian cultivars” for all these 15 cultivars, including Symphonia for simplicity in this paper.
These cultivars demonstrate very broad adaptability (ecological plasticity) under different climatic and ecological conditions (Supplementary File S4). For example, cvs. Nevsky, Zhukovsky, and Udacha have been approved for 12, 10, and 9 regions of the Russian Federation, respectively, relating to the contrasting light zones (The State Register of Selection Achievements Approved for Use, in Russian, https://reestr.gossortrf.ru/, accessed on 2 March 2023). On the other hand, there are cultivars that have been released only in one particular region of the Russian Federation (for example, cvs. Krasavchik and Severnoe Siyanie); the Dutch cv. Symphonia is among them (approved only for one region of the Russian Federation (Supplementary File S4). Since the number of narrow adaptive cultivars in this study was small, we included Symphony in the analysis as well.
Most Russian cultivars were registered in the VIR Genebank both as living accessions and as nomenclature standards in the form of an herbarium voucher according to the International Code of Nomenclature for Cultivated Plants (ICNCP) [75]. Supplementary File S4 includes information about code of the herbarium vouchers (nomenclatural standards) which are maintained in the WIR Herbarium of cultivated plants, and their wild relatives located in Saint Petersburg. The plant material was provided to the WIR Herbarium by the authors of the cultivars.

4.2. DNA Samples Preparation and Sequencing

Green leave samples of 0.1–0.15 g were frozen in liquid nitrogen. The frozen leaves were ground in a ceramic mortar with the addition of liquid nitrogen. Using a cold spatula, the ground mass was transferred to a 1.5 mL test tube. DNA extraction was performed using a DNeasy Plant Mini Kit, QIAGEN (Germany), according to the instructions. An amount of 400 μL of AP1 solution and 4 μL of RNase A (100 mg/mL) were added to the ground mass. The mixture was then incubated in a water bath at 65 °C for 10 min. After that, 130 µL of P3 solution was added, stirred, and incubated on ice for 5 min. After standing on ice, it was centrifuged for 5 min at 14,000 rpm. The lysate was transferred to a QIAshedder Mini spin column, then centrifuged again for 2 min at 14,000 rpm. The purified lysate was carefully transferred to a 1.5 mL tube, where it was mixed with 1.5× volume of AW1 solution and pipetted. The resulting solution was filtered on DNeasy Mini spin columns by centrifugation for 1 min at 8000 rpm. The columns were then transferred to a new tube, and double purification with AW2 solution was performed. DNA elution was performed with AE solution with a total volume of 150 µL in two steps.
Two mkg DNA from potato leaves was fragmented using a Covaris M220 sonicator with parameters optimized for a maximum fragment size of approximately 400 bp for library preparation. Barcoded genome libraries were prepared using 100 ng of fragmented DNA, with Roche KAPA Hyper Prep Kit, and KAPA UDI adapters (ROCHE, Basel, Switzerland), according to the manufacturer’s protocol for dual size selection. Nine PCR cycles were used for amplification, followed by AMPure XP (Agencourt, Brea, CA, USA) purification. Final libraries quantification was performed with a Bioanalyzer 2100 and a DNA High Sensitivity Kit (Agilent, Santa Clara, CA, USA). After normalization, barcoded libraries were pooled and sequenced on a NextSeq 550 or Novaseq 6000 platform (Illimina, San Diego, CA, USA) with 2 × 150 bp paired-end reads. The *.bcl files were converted to fastq format and demultiplexed using the bcl2fastq software (https://support.illumina.com/sequencing/sequencing_software/bcl2fastq-conversion-software.html, accessed at 21 January 2022) according to the developer’s instructions.

4.3. Reference Genome and Sequences of South American Potato Landraces

We used the DM1-3 v4.04 assembly [8], which was downloaded from the SpudDB database—Potato Genomic Resources (http://spuddb.uga.edu/, accessed on 12 February 2022), as the reference genome of potato S. tuberosum. AGAT v. 0.9.2 [76] was used to obtain the amino acid sequences of the DM1-3 reference genome based on DM1-3 genome annotation v4.03. The amino acid sequences of tomato S. lycopersicum SL3.0 [77] downloaded from the Ensembl Plants database (http://plants.ensembl.org, accessed on 12 February 2022) were used as an outgroup for the analysis of genetic diversity in potato accessions studied.
We obtained a library of potato transposable elements using the EDTA package [78] and the DM1-3 reference genomic sequence. This library was used to further identify TEs in the contigs we assembled.
The sequences of South American S. tuberosum potato genomes [17] were downloaded from NCBI by BioProject identifier PRJNA556263. Raw data from SRA: SRR10244436 (BUK), SRR10244437 (AJH), SRR10244438 (STN), SRR10244439 (PHU), SRR10244440 (GON2), SRR10244441 (GON1), SRR10248510 (CUR), SRR10248511 (CHA), SRR10248512 (JUZ), SRR10248513 (TBR), SRR10248514 (ADG2), SRR10248515 (ADG1). Genome assemblies from NCBI genomic database: ASM984962v1 (CHA), ASM984964v1 (CUR), ASM984968v1 (JUZ), ASM984970v1 (ADG1), ASM984972v1 (ADG2), ASM984974v1 (TBR), ASM984975v1 (PHU), ASM984978v1 (STN), ASM984980v1 (AJH), ASM984981v1 (BUK), ASM984986v2 (GON1), ASM984990v1 (GON2). Description of the South American accessions is shown in Table S4 (Supplementary File S1).

4.4. Genome Assembly and Quality Estimation

Genomic sequences of potato varieties cultivated in Russia were processed using the following bioinformatics pipeline:
  • Genome assembly up to contig level using MaSuRCA v3.4.2 [79];
  • Assembly quality estimation using QUAST v5.2.0 [80];
  • TEs masking using obtained TE libraries and RepeatMasker (http://www.repeatmasker.org/, accessed 2 February 2020) [81];
  • ORF identification using AUGUSTUS v3.4.0 [82];
  • Filtering contigs with length below 1000 bps not containing ORFs;
  • Evaluation of the genome completeness performed using BUSCO v5.3.0 [83] with Solanales DB10 (5 August 2020).
The open reading frames for South American potato genomic sequences identified these genomes using AUGUSTUS v3.4.0 [82], see step 4 of the pipeline.

4.5. Orthologous Gene Groups Identification and Genetic Diversity Analysis

Identification and analysis of orthologous groups were performed for protein-coding sequences of 15 cultivars grown in Russia, 12 South American potato accessions, the potato reference genome DM1-3 and the outgroup S. lycopersicum tomato genome.
OrthoFinder v2.5.2 [84] was used to identify orthologous groups. This program was also used to build a tree reflecting genetic diversity between potato accessions. OrthoFinder was run with the -m MSA parameter, which allowed us to reconstruct the species tree using an algorithm based on the reconstruction of phylogenetic trees of individual orthogroups and their combinations. This method takes into account possible duplications and loss of genes within individual orthogroups, which is important in our case of genomic sequences reconstructed from short reads. The phylogenetic tree was visualized using iTOL v.6 [85].

4.6. Orthologous Gene Groups Identification and Phylogenetic Reconstruction

The orthologs for genomes of 15 Russian cultivars were classified into 4 classes: core (the orthogroup includes sequences from all 15 genomes), softcore (the orthogroup includes sequences from 14 genomes), shell (the orthogroup includes sequences from 2–13 genomes), and cloud (the orthogroup includes sequences from single genome). Pan-genome size modeling based on orthogroup data was performed via PanGP v1.0.1 [86] with a random algorithm and sample size of 500.
Functional annotation of proteins from each genome for Russian and South American accessions was performed using InterProScan v5.51.85 [87]. The frequency of occurrence of a particular protein function term was calculated based on annotation of Pfam domains [88] and InterPro database identifiers. Next, the frequency of occurrence of the protein function term in the variable and conserved parts of the potato pan-genome was estimated.

4.7. Identification and Analysis of NBS-LRR Genes

NBS-LRR domains of proteins associated with plant immunity were searched in the amino acid sequences of the genomes of Russian potato cultivars. The NLR-Parser program was used for this purpose [89]. Only those proteins in which the complete (“complete”) and true (“true”) domain structure of NLR proteins was reported by NLR-Parser were taken for further consideration. To classify these sequences into different classes (TNL, CNL-R and CNL1-8 groups), we used the partitioning of protein sequences into orthogroups (see Section 4.5 above). The orthogroup sequences in which the corresponding classes of NBS-LRR proteins of the potato reference genome DM1-3 [36] were represented were assigned to these classes. The partitioning of the reference genome proteins into the corresponding classes was taken from ref. [36] (listed also in the Supplementary file “Identical_NB_ARC.xlsx” from ref. [34]).

4.8. CNV Identification and Analysis

Illumina short reads were used to evaluate and analyze Copy Number Variation in Russian and South American potato genomes. Quality filtering of reads and trimming of adapters was performed using Trimmomatic v0.39 [90] with the following parameters: TruSeq3-PE.fa:2:30:10 SLIDINGWINDOW:5:20 LEADING:20 TRAILING:20 MINLEN:50. Trimmed reads were aligned to the DM1-3 v4.04 reference gene using BWA v0.7.17 [91]. Reads were labeled for duplicates via Picard v2.26.1 (https://broadinstitute.github.io/picard/, accessed 2 February 2020) and sorted and indexed using SAMTOOLS v1.12 [92]. Only properly paired reads were used for further analysis.
Alignment results were used to detect CNV using CNVpytor v0.4.1 [93]. CNVs were detected on all chromosomes of the DM1-3 v4.04 reference genome as well as on pseudomolecules (chr00 and ChrUn). CNVs calls were filtered as follows: length greater than 1 kb, P-value (first e-value) < 0.01, q0 < 50%, and pN < 50%. The R package intansv v1.12.0 [94] was used to find correspondence between the identified CNVs and the genes in the potato genome. For this purpose, the CNVpytor output files were converted to the format required for input by removing the last two columns (pN and dG).
Visualization of the position of CNVs in the DM1-3 genome for Russian and South American cultivars was obtained using the Circos [95].
The list of CNVs was formatted as a table with rows corresponding to the potato genotype and columns corresponding to the genes for which significant CNVs were identified for further statistical processing. The table element was +1 if the CNV corresponded to a duplication of a region, -1 if the CNV was a deletion, and 0 if no significant CNVs for the accession were found in that region.
To evaluate the similarity of Russian and South American potato accessions by their CNV characteristics, we performed principal component analysis (PCA) using column of the above table using as variables by the Scikit-learn v1.1.2 package [96]. A tree for potato genomes based on their CNV similarity was built using the PARS algorithm of PHYLIP [97].
In order to identify CNVs with significant differences in occurrences within the groups of Russian and South American potato genomes, we used the analysis of 2 × 3 contingency tables for the table of CNV +1/0/-1 types for genes described above. Genotypes were classified into 2 classes: Russian and South American. We considered eight South American accessions (CHA, ADG1, ADG2, TBR, PHU, STN, GON1, GON2) forming a sister cluster in relation to genomes of Russian cultivars analyzed here. CNVs were classified into three types: −1, +1, 0 (see above). The significance of associations between potato variety type and CNVs was assessed using Fisher’s exact test implemented in the Python rpy2 library (https://rpy2.github.io/, accessed 02.07.2021). The association between CNV and genome types was considered significant at a p-value < 0.01. Additionally, we classified genes with CNV according to specific patterns of representation in potato variety groups. RUDup: genes with significant duplications detected in 50% or more of Russian cultivars, while they are absent in South American accessions. RUDel: genes that have significant deletions in 50% or more of Russian cultivars but are absent in South American accessions. SADup: genes that have significant duplications in 50% or more of South American accessions, but do not have them in Russian cultivars. SADel: genes found to have significant deletions in 50% or more of South American accessions, but not in Russian cultivars.
Functional analysis of gene groups with specific CNV patterns was performed using the DAVID web service [98].
Adaptation to the long day period is an important diversification factor for potato varieties. Therefore, we compiled a list of genes related to tuberization and photoperiodicity processes in potatoes and searched these genes among those having specific CNV patterns in populations of genomes of Russian cultivars and South American accessions analyzed in the present study. For this purpose, the initial list of tuberization and photoperiodicity-related genes was taken from a review [24] based on references given there [67,99,100,101,102,103,104,105] and supplemented with genes from refs. [25,26]. A total of 48 potato genes were included in the list (Supplementary File S3).

5. Conclusions

In the present work, we performed short-read sequencing, assembly, and structural analysis of genomes of 15 cultivars grown in Russia (14 cultivars of Russian origin and 1 Dutch cultivar). The main characteristics of our assemblies are consistent with those obtained from the assemblies for short reads of various potato genomes (~60% of the genes belong to the conserved part of the pan-genome, 38% to the shell, and 2% to the cloud part). ORFs from the pan-genome core are related to basal gene function, the variable part is associated with immune response, and genes are responsible for environmental adaptation. The set of the NBS-LRR genes in the accessions is highly variable: on average we identified 227 complete NBS-LRR sequences per genome; the fraction of classified genes into CNL/TNL classes is about 40%.
South American potato genomes included in our comparative analysis demonstrated separation from the Russian potato genomes on the genetic differentiation tree: highland Andean (S. juzepczukii, S. curtilobum, and S. ahanhuiri), other South American and Russian accessions. We performed CNV analysis, the results of which on the distribution of their main characteristics in Russian and South American accessions are in agreement with previously published data: the number of deletions exceeds the number of duplications, with a higher number of genes with duplications than with deletions. Russian cultivars demonstrated homogeneity in CMV characteristics in comparison with South American potato landraces.
We identified genes with CNV with different occurrences in South American and Russian potato accessions. This allowed us to evaluate the functions of the genes associated with the diversification of Russian and South American potato cultivars. The functional analysis of these genes showed that a significant part of them is related to the immune response or response to abiotic stress. At the same time, a detailed analysis of genes related to tuberization, and photoperiod control revealed significant differences in CNV occurrence in four of the known genes and identified an additional gene homologous to the PARG gene of Arabidopsis, which may be involved in circadian rhythm control processes related to the acclimation processes of Russian potato cultivars.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24065713/s1. References [3,4,17,22,106,107] are cited in the supplementary materials.

Author Contributions

Conceptualization, E.A.S., A.V.K. and D.A.A.; methodology, D.I.K., G.V.V., M.A.N., S.V.T., M.A.G., N.A.S., S.M.I., D.A.R. and D.A.A.; software, D.I.K.; validation, D.I.K. and D.A.A.; formal analysis, D.I.K. and D.A.A.; investigation, D.I.K. and D.A.A.; resources, M.V.P., E.A.S. and A.V.K.; data curation, T.A.G., D.I.K. and D.A.A.; writing—original draft preparation, G.V.V., M.A.N., S.V.T., N.A.S., T.A.G., D.I.K. and D.A.A.; writing—review and editing, E.A.S., T.A.G. and D.A.A.; visualization, D.I.K. and D.A.A.; supervision, E.A.S., A.V.K. and D.A.A.; project administration, E.A.S. and D.A.A.; funding acquisition, M.V.P., E.A.S. and A.V.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Kurchatov Genome Center of the Institute of Cytology and Genetics of Siberian Branch of the Russian Academy of Sciences, in agreement with the Ministry of Education and Science of the Russian Federation, no. 075-15-2019-1662.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The assembled genomes and all raw sequencing data have been deposited under NCBI BioProject PRJNA933976.

Acknowledgments

The data analysis was performed using computational resources of the Novosibirsk State University High-Performance Computing Center and the “Bioinformatics” Joint Computational Center supported by the budget project № FWNR-2022-0020. Multiplication of the plant material was conducted in the Laboratory of Artificial Plant Growth of ICG SB RAS within the budgetary project FWNR-2022-0017.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CNVCopy number variation
cv. Cultivar
NBS-LRR Nucleotide-binding site and leucine-rich repeats
ORFOpen reading frame
PCPrincipal component
PCAPrincipal component analysis
SAUR Small auxin-up RNA

References

  1. Dahal, K.; Li, X.-Q.; Tai, H.; Creelman, A.; Bizimungu, B. Improving Potato Stress Tolerance and Tuber Yield Under a Climate Change Scenario–A Current Overview. Front. Plant Sci. 2019, 10, 563. [Google Scholar] [CrossRef] [PubMed]
  2. Beals, K.A. Potatoes, Nutrition and Health. Am. J. Potato Res. 2019, 96, 102–110. [Google Scholar] [CrossRef] [Green Version]
  3. Spooner, D.M.; Núñez, J.; Trujillo, G.; del Rosario Herrera, M.; Guzmán, F.; Ghislain, M. Extensive Simple Sequence Repeat Genotyping of Potato Landraces Supports a Major Reevaluation of Their Gene Pool Structure and Classification. Proc. Natl. Acad. Sci. USA 2007, 104, 19398–19403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Spooner, D.M.; Ghislain, M.; Simon, R.; Jansky, S.H.; Gavrilenko, T. Systematics, Diversity, Genetics, and Evolution of Wild and Cultivated Potatoes. Bot. Rev. 2014, 80, 283–383. [Google Scholar] [CrossRef]
  5. Van Berloo, R.; Hutten, R.C.B.; van Eck, H.J.; Visser, R.G.F. An Online Potato Pedigree Database Resource. Potato Res. 2007, 50, 45–57. [Google Scholar] [CrossRef] [Green Version]
  6. Tang, D.; Jia, Y.; Zhang, J.; Li, H.; Cheng, L.; Wang, P.; Bao, Z.; Liu, Z.; Feng, S.; Zhu, X.; et al. Genome Evolution and Diversity of Wild and Cultivated Potatoes. Nature 2022, 606, 535–541. [Google Scholar] [CrossRef]
  7. Watanabe, K. Potato Genetics, Genomics, and Applications. Breed. Sci. 2015, 65, 53–68. [Google Scholar] [CrossRef] [Green Version]
  8. The Potato Genome Sequencing Consortium Genome Sequence and Analysis of the Tuber Crop Potato. Nature 2011, 475, 189–195. [CrossRef] [Green Version]
  9. Sharma, S.K.; Bolser, D.; de Boer, J.; Sønderkær, M.; Amoros, W.; Carboni, M.F.; D’Ambrosio, J.M.; de la Cruz, G.; Di Genova, A.; Douches, D.S.; et al. Construction of Reference Chromosome-Scale Pseudomolecules for Potato: Integrating the Potato Genome with Genetic and Physical Maps. G3 Genes Genomes Genet. 2013, 3, 2031–2047. [Google Scholar] [CrossRef] [Green Version]
  10. Pham, G.M.; Hamilton, J.P.; Wood, J.C.; Burke, J.T.; Zhao, H.; Vaillancourt, B.; Ou, S.; Jiang, J.; Buell, C.R. Construction of a Chromosome-Scale Long-Read Reference Genome Assembly for Potato. GigaScience 2020, 9, giaa100. [Google Scholar] [CrossRef]
  11. Hardigan, M.A.; Crisovan, E.; Hamilton, J.P.; Kim, J.; Laimbeer, P.; Leisner, C.P.; Manrique-Carpintero, N.C.; Newton, L.; Pham, G.M.; Vaillancourt, B.; et al. Genome Reduction Uncovers a Large Dispensable Genome and Adaptive Role for Copy Number Variation in Asexually Propagated Solanum tuberosum. Plant Cell 2016, 28, 388–405. [Google Scholar] [CrossRef] [Green Version]
  12. Zhou, Q.; Tang, D.; Huang, W.; Yang, Z.; Zhang, Y.; Hamilton, J.P.; Visser, R.G.F.; Bachem, C.W.B.; Robin Buell, C.; Zhang, Z.; et al. Haplotype-Resolved Genome Analyses of a Heterozygous Diploid Potato. Nat. Genet. 2020, 52, 1018–1023. [Google Scholar] [CrossRef] [PubMed]
  13. Freire, R.; Weisweiler, M.; Guerreiro, R.; Baig, N.; Hüttel, B.; Obeng-Hinneh, E.; Renner, J.; Hartje, S.; Muders, K.; Truberg, B.; et al. Chromosome-Scale Reference Genome Assembly of a Diploid Potato Clone Derived from an Elite Variety. G3 Genes Genomes Genetics 2021, 11, jkab330. [Google Scholar] [CrossRef]
  14. Sun, H.; Jiao, W.-B.; Krause, K.; Campoy, J.A.; Goel, M.; Folz-Donahue, K.; Kukat, C.; Huettel, B.; Schneeberger, K. Chromosome-Scale and Haplotype-Resolved Genome Assembly of a Tetraploid Potato Cultivar. Nat. Genet. 2022, 54, 342–348. [Google Scholar] [CrossRef] [PubMed]
  15. Hoopes, G.; Meng, X.; Hamilton, J.P.; Achakkagari, S.R.; de Alves Freitas Guesdes, F.; Bolger, M.E.; Coombs, J.J.; Esselink, D.; Kaiser, N.R.; Kodde, L.; et al. Phased, Chromosome-Scale Genome Assemblies of Tetraploid Potato Reveal a Complex Genome, Transcriptome, and Predicted Proteome Landscape Underpinning Genetic Diversity. Mol. Plant 2022, 15, 520–536. [Google Scholar] [CrossRef] [PubMed]
  16. Kyriakidou, M.; Anglin, N.L.; Ellis, D.; Tai, H.H.; Strömvik, M.V. Genome Assembly of Six Polyploid Potato Genomes. Sci. Data 2020, 7, 88. [Google Scholar] [CrossRef] [Green Version]
  17. Kyriakidou, M.; Achakkagari, S.R.; Gálvez López, J.H.; Zhu, X.; Tang, C.Y.; Tai, H.H.; Anglin, N.L.; Ellis, D.; Strömvik, M.V. Structural Genome Analysis in Cultivated Potato Taxa. Theor. Appl. Genet. 2020, 133, 951–966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Pham, G.M.; Newton, L.; Wiegert-Rininger, K.; Vaillancourt, B.; Douches, D.S.; Buell, C.R. Extensive Genome Heterogeneity Leads to Preferential Allele Expression and Copy Number-dependent Expression in Cultivated Potato. Plant J. 2017, 92, 624–637. [Google Scholar] [CrossRef] [Green Version]
  19. Tiwari, J.K.; Rawat, S.; Luthra, S.K.; Zinta, R.; Sahu, S.; Varshney, S.; Kumar, V.; Dalamu, D.; Mandadi, N.; Kumar, M.; et al. Genome Sequence Analysis Provides Insights on Genomic Variation and Late Blight Resistance Genes in Potato Somatic Hybrid (Parents and Progeny). Mol. Biol. Rep. 2021, 48, 623–635. [Google Scholar] [CrossRef]
  20. Achakkagari, S.R.; Kyriakidou, M.; Gardner, K.M.; De Koeyer, D.; De Jong, H.; Strömvik, M.V.; Tai, H.H. Genome Sequencing of Adapted Diploid Potato Clones. Front. Plant Sci. 2022, 13, 954933. [Google Scholar] [CrossRef]
  21. Lye, Z.N.; Purugganan, M.D. Copy Number Variation in Domestication. Trends Plant Sci. 2019, 24, 352–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ovchinnikova, A.; Krylova, E.; Gavrilenko, T.; Smekalova, T.; Zhuk, M.; Knapp, S.; Spooner, D.M. Taxonomy of Cultivated Potatoes (Solanum Section Petota: Solanaceae): Cultivated potato taxonomy. Bot. J. Linn. Soc. 2011, 165, 107–155. [Google Scholar] [CrossRef] [Green Version]
  23. Spooner, D.M.; Gavrilenko, T.; Jansky, S.H.; Ovchinnikova, A.; Krylova, E.; Knapp, S.; Simon, R. Ecogeography of Ploidy Variation in Cultivated Potato (Solanum Sect. Petota). Am. J. Bot. 2010, 97, 2049–2060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Osnato, M.; Cota, I.; Nebhnani, P.; Cereijo, U.; Pelaz, S. Photoperiod Control of Plant Growth: Flowering Time Genes Beyond Flowering. Front. Plant Sci. 2022, 12, 805635. [Google Scholar] [CrossRef]
  25. Niu, Y.; Li, G.; Jian, Y.; Duan, S.; Liu, J.; Xu, J.; Jin, L. Genes Related to Circadian Rhythm Are Involved in Regulating Tuberization Time in Potato. Hortic. Plant J. 2022, 8, 369–380. [Google Scholar] [CrossRef]
  26. Yanovsky, M.J.; Izaguirre, M.; Wagmaister, J.A.; Gatz, C.; Jackson, S.D.; Thomas, B.; Casal, J.J. Phytochrome A Resets the Circadian Clock and Delays Tuber Formation under Long Days in Potato: Resetting of the Circadian Clock by PhyA. Plant J. 2000, 23, 223–232. [Google Scholar] [CrossRef]
  27. Panda, S.; Poirier, G.G.; Kay, S.A. Tej Defines a Role for Poly(ADP-Ribosyl)Ation in Establishing Period Length of the Arabidopsis Circadian Oscillator. Dev. Cell 2002, 3, 51–61. [Google Scholar] [CrossRef] [Green Version]
  28. Pronozin, A.Y.; Bragina, M.K.; Salina, E.A. Crop Pangenomes. Vestn. VOGiS 2021, 25, 57–63. [Google Scholar] [CrossRef]
  29. Bayer, P.E.; Golicz, A.A.; Scheben, A.; Batley, J.; Edwards, D. Plant Pan-Genomes Are the New Reference. Nat. Plants 2020, 6, 914–920. [Google Scholar] [CrossRef]
  30. Hirsch, C.N.; Foerster, J.M.; Johnson, J.M.; Sekhon, R.S.; Muttoni, G.; Vaillancourt, B.; Peñagaricano, F.; Lindquist, E.; Pedraza, M.A.; Barry, K.; et al. Insights into the Maize Pan-Genome and Pan-Transcriptome. Plant Cell 2014, 26, 121–135. [Google Scholar] [CrossRef] [Green Version]
  31. Rodewald, J.; Trognitz, B. Solanum Resistance Genes against Phytophthora Infestans and Their Corresponding Avirulence Genes: Solanum Rpi Genes and Their Avirulence Counterparts. Mol. Plant Pathol. 2013, 14, 740–757. [Google Scholar] [CrossRef]
  32. Khiutti, A.V.; Antonova, O.Y.; Mironenko, N.V.; Gavrilenko, T.A.; Afanasenko, O.S. Potato Resistance to Quarantine Diseases. Russ J Genet Appl Res 2017, 7, 833–844. [Google Scholar] [CrossRef]
  33. Kochetov, A.V.; Gavrilenko, T.A.; Afanasenko, O.S. New Genetic Tools for Plant Defense against Parasitic Nematodes. Vestn. VOGiS 2021, 25, 337–343. [Google Scholar] [CrossRef] [PubMed]
  34. Kochetov, A.V.; Afonnikov, D.A.; Shmakov, N.; Vasiliev, G.V.; Antonova, O.Y.; Shatskaya, N.V.; Glagoleva, A.Y.; Ibragimova, S.M.; Khiutti, A.; Afanasenko, O.S.; et al. NLR Genes Related Transcript Sets in Potato Cultivars Bearing Genetic Material of Wild Mexican Solanum Species. Agronomy 2021, 11, 2426. [Google Scholar] [CrossRef]
  35. Prakash, C.; Trognitz, F.C.; Venhuizen, P.; von Haeseler, A.; Trognitz, B. A Compendium of Genome-Wide Sequence Reads from NBS (Nucleotide Binding Site) Domains of Resistance Genes in the Common Potato. Sci. Rep. 2020, 10, 11392. [Google Scholar] [CrossRef]
  36. Jupe, F.; Pritchard, L.; Etherington, G.J.; MacKenzie, K.; Cock, P.J.; Wright, F.; Sharma, S.K.; Bolser, D.; Bryan, G.J.; Jones, J.D.; et al. Identification and Localisation of the NB-LRR Gene Family within the Potato Genome. BMC Genom. 2012, 13, 75. [Google Scholar] [CrossRef] [Green Version]
  37. Gavrilenko, T.; Antonova, O.; Ovchinnikova, A.; Novikova, L.; Krylova, E.; Mironenko, N.; Pendinen, G.; Islamshina, A.; Shvachko, N.; Kiru, S.; et al. A Microsatellite and Morphological Assessment of the Russian National Cultivated Potato Collection. Genet. Resour. Crop. Evol. 2010, 57, 1151–1164. [Google Scholar] [CrossRef]
  38. Hardigan, M.A.; Laimbeer, F.P.E.; Newton, L.; Crisovan, E.; Hamilton, J.P.; Vaillancourt, B.; Wiegert-Rininger, K.; Wood, J.C.; Douches, D.S.; Farré, E.M.; et al. Genome Diversity of Tuber-Bearing Solanum Uncovers Complex Evolutionary History and Targets of Domestication in the Cultivated Potato. Proc. Natl. Acad. Sci. USA 2017, 114, E9999–E10008. [Google Scholar] [CrossRef] [Green Version]
  39. Lihodeevskiy, G.A.; Shanina, E.P. Structural Variations in the Genome of Potato Varieties of the Ural Selection. Agronomy 2021, 11, 1703. [Google Scholar] [CrossRef]
  40. Achakkagari, S.R.; Kyriakidou, M.; Tai, H.H.; Anglin, N.L.; Ellis, D.; Strömvik, M.V. Complete Plastome Assemblies from a Panel of 13 Diverse Potato Taxa. PLoS ONE 2020, 15, e0240124. [Google Scholar] [CrossRef]
  41. Hosaka, K. Who Is the Mother of the Potato? Restriction Endonuclease Analysis of Chloroplast DNA of Cultivated Potatoes. Theoret. Appl. Genetics 1986, 72, 606–618. [Google Scholar] [CrossRef] [PubMed]
  42. Gavrilenko, T.; Antonova, O.; Shuvalova, A.; Krylova, E.; Alpatyeva, N.; Spooner, D.M.; Novikova, L. Genetic Diversity and Origin of Cultivated Potatoes Based on Plastid Microsatellite Polymorphism. Genet. Resour. Crop Evol. 2013, 60, 1997–2015. [Google Scholar] [CrossRef]
  43. Gavrilenko, T.; Chukhina, I.; Antonova, O.; Krylova, E.; Shipilina, L.; Oskina, N.; Kostina, L. Comparative Analysis of the Genetic Diversity of Chilean Cultivated Potato Based on a Molecular Study of Authentic Herbarium Specimens and Present-Day Gene Bank Accessions. Plants 2022, 12, 174. [Google Scholar] [CrossRef] [PubMed]
  44. Sucar, S.; Carboni, M.F.; Rey Burusco, M.F.; Castellote, M.A.; Massa, G.A.; Monte, M.N.; Feingold, S.E. Assessment of Genetic Diversity and Relatedness in an Andean Potato Collection from Argentina by High-Density Genotyping. Horticulturae 2022, 8, 54. [Google Scholar] [CrossRef]
  45. Gu, B.; Cao, X.; Zhou, X.; Chen, Z.; Wang, Q.; Liu, W.; Chen, Q.; Zhao, H. The Histological, Effectoromic, and Transcriptomic Analyses of Solanum Pinnatisectum Reveal an Upregulation of Multiple NBS-LRR Genes Suppressing Phytophthora Infestans Infection. Int. J. Mol. Sci. 2020, 21, 3211. [Google Scholar] [CrossRef] [PubMed]
  46. Poustka, F.; Irani, N.G.; Feller, A.; Lu, Y.; Pourcel, L.; Frame, K.; Grotewold, E. A Trafficking Pathway for Anthocyanins Overlaps with the Endoplasmic Reticulum-to-Vacuole Protein-Sorting Route in Arabidopsis and Contributes to the Formation of Vacuolar Inclusions. Plant Physiol. 2007, 145, 1323–1335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Shitan, N.; Yazaki, K. Dynamism of Vacuoles toward Survival Strategy in Plants. Biochim. Biophys. Acta BBA-Biomembr. 2020, 1862, 183127. [Google Scholar] [CrossRef]
  48. Young, M.J.; Bay, D.C.; Hausner, G.; Court, D.A. The Evolutionary History of Mitochondrial Porins. BMC Evol. Biol. 2007, 7, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Norero, N.; Rey Burusco, M.; D’Ippólito, S.; Décima Oneto, C.; Massa, G.; Castellote, M.; Feingold, S.; Guevara, M. Genome-Wide Analyses of Aspartic Proteases on Potato Genome (Solanum Tuberosum): Generating New Tools to Improve the Resistance of Plants to Abiotic Stress. Plants 2022, 11, 544. [Google Scholar] [CrossRef]
  50. Herath, V.; Verchot, J. Comprehensive Transcriptome Analysis Reveals Genome-Wide Changes Associated with Endoplasmic Reticulum (ER) Stress in Potato (Solanum tuberosum L.). Int. J. Mol. Sci. 2022, 23, 13795. [Google Scholar] [CrossRef]
  51. Hu, J.; Manduzio, S.; Kang, H. Epitranscriptomic RNA Methylation in Plant Development and Abiotic Stress Responses. Front. Plant Sci. 2019, 10, 500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Zhang, X.; Gonzalez-Carranza, Z.H.; Zhang, S.; Miao, Y.; Liu, C.; Roberts, J.A. F-Box Proteins in Plants. In Annual Plant Reviews online; Roberts, J.A., Ed.; John Wiley and Sons: Hoboken, NJ, USA, 2019; pp. 307–328. ISBN 978-1-119-31299-4. [Google Scholar]
  53. Feng, B.; Liu, C.; de Oliveira, M.V.V.; Intorne, A.C.; Li, B.; Babilonia, K.; de Souza Filho, G.A.; Shan, L.; He, P. Protein Poly(ADP-Ribosyl)Ation Regulates Arabidopsis Immune Gene Expression and Defense Responses. PLoS Genet. 2015, 11, e1004936. [Google Scholar] [CrossRef] [PubMed]
  54. Rissel, D.; Peiter, E. Poly(ADP-Ribose) Polymerases in Plants and Their Human Counterparts: Parallels and Peculiarities. IJMS 2019, 20, 1638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Amor, Y.; Babiychuk, E.; Inzé, D.; Levine, A. The Involvement of Poly(ADP-Ribose) Polymerase in the Oxidative Stress Responses in Plants. FEBS Lett. 1998, 440, 1–7. [Google Scholar] [CrossRef]
  56. Lamb, R.S.; Citarelli, M.; Teotia, S. Functions of the Poly(ADP-Ribose) Polymerase Superfamily in Plants. Cell. Mol. Life Sci. 2012, 69, 175–189. [Google Scholar] [CrossRef]
  57. Vainonen, J.; Shapiguzov, A.; Vaattovaara, A.; Kangasjärvi, J. Plant PARPs, PARGs and PARP-like Proteins. Curr. Protein Pept. Sci. 2016, 17, 713–723. [Google Scholar] [CrossRef]
  58. Díaz, A.; Zikhali, M.; Turner, A.S.; Isaac, P.; Laurie, D.A. Copy Number Variation Affecting the Photoperiod-B1 and Vernalization-A1 Genes Is Associated with Altered Flowering Time in Wheat (Triticum aestivum). PLoS ONE 2012, 7, e33234. [Google Scholar] [CrossRef] [Green Version]
  59. Nitcher, R.; Distelfeld, A.; Tan, C.; Yan, L.; Dubcovsky, J. Increased Copy Number at the HvFT1 Locus Is Associated with Accelerated Flowering Time in Barley. Mol Genet. Genom. 2013, 288, 261–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Würschum, T.; Boeven, P.H.G.; Langer, S.M.; Longin, C.F.H.; Leiser, W.L. Multiply to Conquer: Copy Number Variations at Ppd-B1 and Vrn-A1 Facilitate Global Adaptation in Wheat. BMC Genet. 2015, 16, 96. [Google Scholar] [CrossRef] [Green Version]
  61. Würschum, T.; Langer, S.M.; Longin, C.F.H.; Tucker, M.R.; Leiser, W.L. A Three-Component System Incorporating Ppd-D1, Copy Number Variation at Ppd-B1, and Numerous Small-Effect Quantitative Trait Loci Facilitates Adaptation of Heading Time in Winter Wheat Cultivars of Worldwide Origin: Heading Time Control in Wheat. Plant Cell Env. 2018, 41, 1407–1416. [Google Scholar] [CrossRef]
  62. Beales, J.; Turner, A.; Griffiths, S.; Snape, J.W.; Laurie, D.A. A Pseudo-Response Regulator Is Misexpressed in the Photoperiod Insensitive Ppd-D1a Mutant of Wheat (Triticum aestivum L.). Ther. Appl Genet. 2007, 115, 721–733. [Google Scholar] [CrossRef] [PubMed]
  63. Würschum, T.; Rapp, M.; Miedaner, T.; Longin, C.F.H.; Leiser, W.L. Copy Number Variation of Ppd-B1 Is the Major Determinant of Heading Time in Durum Wheat. BMC Genet. 2019, 20, 64. [Google Scholar] [CrossRef] [PubMed]
  64. Demenkov, P.S.; Saik, O.V.; Ivanisenko, T.V.; Kolchanov, N.A.; Kochetov, A.V.; Ivanisenko, V.A. Prioritization of Potato Genes Involved in the Formation of Agronomically Valuable Traits Using the Solanum tuberosum Knowledge Base. Vestn. VOGiS 2019, 23, 312–319. [Google Scholar] [CrossRef]
  65. Hawkes, J.G. History of the Potato. In The Potato Crop; Harris, P.M., Ed.; Springer: Dordrecht, The Netherlands, 1992; pp. 1–12. ISBN 978-94-010-5034-0. [Google Scholar]
  66. Gutaker, R.M.; Weiß, C.L.; Ellis, D.; Anglin, N.L.; Knapp, S.; Luis Fernández-Alonso, J.; Prat, S.; Burbano, H.A. The Origins and Adaptation of European Potatoes Reconstructed from Historical Genomes. Nat. Ecol. Evol. 2019, 3, 1093–1101. [Google Scholar] [CrossRef]
  67. Kloosterman, B.; Abelenda, J.A.; del Mar Carretero Gomez, M.; Oortwijn, M.; de Boer, J.M.; Kowitwanich, K.; Horvath, B.M.; van Eck, H.J.; Smaczniak, C.; Prat, S.; et al. Naturally Occurring Allele Diversity Allows Potato Cultivation in Northern Latitudes. Nature 2013, 495, 246–250. [Google Scholar] [CrossRef] [PubMed]
  68. Rybakov, D.А.; Аntonova, O.Y.; Chukhina, I.G.; Fomina, N.А.; Klimenko, N.S.; Zheltova, V.V.; Meleshin, А.А.; Kochieva, E.Z.; Oves, E.V.; Аpshev, K.K.; et al. Nomenclatural standards and genetic passports of potato cultivars bred in the A.G. Lorkh All-Russian Research Institute of Potato Farming. Biotehnol. Sel. Rastenij 2021, 3, 5–52. [Google Scholar] [CrossRef]
  69. Klimenko, N.S.; Gavrilenko, T.A.; Chukhina, I.G.; Gadzhiev, N.M.; Evdokimova, Z.Z.; Lebedeva, V.A. Nomenclatural standards and genetic passports of potato cultivars bred at the Leningrad Research Institute for Agriculture “Belogorka”. Biotehnol. Sel. Rastenij 2021, 3, 18–54. [Google Scholar] [CrossRef]
  70. Gavrilenko, T.A.; Antonova, O.Y.; Kostina, L.I. Study of Genetic Diversity in Potato Cultivars Using PCR Analysis of Organelle DNA. Russ. J. Genet. 2007, 43, 1301–1305. [Google Scholar] [CrossRef]
  71. Touloosh, V.P.; Kanool, S.K. Promising potato varieties for the Tuva Republic. Bull. Tuva State Univ. 2011, 2, 85–87. [Google Scholar]
  72. Makarova, T.A.; Makarov, P.N. Prospects of growing of potatoes with different maturation periods and methods of their protection from fungal diseases in the northern districts of Tyumen region. Plant Prot. News 2019, 4, 22–28. [Google Scholar] [CrossRef]
  73. Rodina, E.S.; Renev, N.O.; Nurpeisova, A.S.; Zhusupov, E.K. Productivity of nematode resistant sorts of potato in Tyumen region. IOSAU 2021, 91, 53–56. [Google Scholar]
  74. Gavrilenko, Т.A.; Klimenko, N.S.; Alpatieva, N.V.; Kostina, L.I.; Lebedeva, V.A.; Evdokimova, Z.Z.; Apalikova, O.V.; Novikova, L.Y.; Antonova, O.Y. Cytoplasmic Genetic Diversity of Potato Varieties Bred in Russia and FSU Countries. Vestn. VOGiS 2019, 23, 753–764. [Google Scholar] [CrossRef] [Green Version]
  75. Brickell, C.D.; Alexander, C.; Cubey, J.J.; David, J.C.; Hoffman, M.H.A.; Leslie, A.C.; Malécot, V.; Jin, X. International Code of Nomenclature for Cultivated Plants (ICNCP or Cultivated Plant Code), 9th ed. Scripta horticulturae. 18, 190.
  76. Dainat, J.; Hereñú, D.; Pucholt, P. NBISweden/AGAT: AGAT-v0.7.0. 2021. Available online: https://github.com/NBISweden/AGAT/releases (accessed on 12 March 2023).
  77. The Tomato Genome Consortium the Tomato Genome Sequence Provides Insights into Fleshy Fruit Evolution. Nature 2012, 485, 635–641. [CrossRef] [Green Version]
  78. Ou, S.; Su, W.; Liao, Y.; Chougule, K.; Agda, J.R.A.; Hellinga, A.J.; Lugo, C.S.B.; Elliott, T.A.; Ware, D.; Peterson, T.; et al. Benchmarking Transposable Element Annotation Methods for Creation of a Streamlined, Comprehensive Pipeline. Genome Biol. 2019, 20, 275. [Google Scholar] [CrossRef] [Green Version]
  79. Zimin, A.V.; Marçais, G.; Puiu, D.; Roberts, M.; Salzberg, S.L.; Yorke, J.A. The MaSuRCA Genome Assembler. Bioinformatics 2013, 29, 2669–2677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Gurevich, A.; Saveliev, V.; Vyahhi, N.; Tesler, G. QUAST: Quality Assessment Tool for Genome Assemblies. Bioinformatics 2013, 29, 1072–1075. [Google Scholar] [CrossRef] [Green Version]
  81. Smit, A.F.A.; Hubley, R.; Green, P. RepeatMasker Open-4.0. 2013–2015, 2015; 289–300.
  82. Stanke, M.; Morgenstern, B. AUGUSTUS: A Web Server for Gene Prediction in Eukaryotes That Allows User-Defined Constraints. Nucleic Acids Res. 2005, 33, W465–W467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Manni, M.; Berkeley, M.R.; Seppey, M.; Zdobnov, E.M. BUSCO: Assessing Genomic Data Quality and Beyond. Curr. Protoc. 2021, 1, e323. [Google Scholar] [CrossRef]
  84. Emms, D.M.; Kelly, S. OrthoFinder: Phylogenetic Orthology Inference for Comparative Genomics. Genome Biol. 2019, 20, 238. [Google Scholar] [CrossRef] [Green Version]
  85. Letunic, I.; Bork, P. Interactive Tree of Life (ITOL) v5: An Online Tool for Phylogenetic Tree Display and Annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef]
  86. Zhao, Y.; Jia, X.; Yang, J.; Ling, Y.; Zhang, Z.; Yu, J.; Wu, J.; Xiao, J. PanGP: A Tool for Quickly Analyzing Bacterial Pan-Genome Profile. Bioinformatics 2014, 30, 1297–1299. [Google Scholar] [CrossRef] [Green Version]
  87. Blum, M.; Chang, H.-Y.; Chuguransky, S.; Grego, T.; Kandasaamy, S.; Mitchell, A.; Nuka, G.; Paysan-Lafosse, T.; Qureshi, M.; Raj, S.; et al. The InterPro Protein Families and Domains Database: 20 Years On. Nucleic Acids Res. 2021, 49, D344–D354. [Google Scholar] [CrossRef] [PubMed]
  88. Mistry, J.; Chuguransky, S.; Williams, L.; Qureshi, M.; Salazar, G.A.; Sonnhammer, E.L.L.; Tosatto, S.C.E.; Paladin, L.; Raj, S.; Richardson, L.J.; et al. Pfam: The Protein Families Database in 2021. Nucleic Acids Res. 2021, 49, D412–D419. [Google Scholar] [CrossRef] [PubMed]
  89. Steuernagel, B.; Jupe, F.; Witek, K.; Jones, J.D.G.; Wulff, B.B.H. NLR-Parser: Rapid Annotation of Plant NLR Complements. Bioinformatics 2015, 31, 1665–1667. [Google Scholar] [CrossRef] [Green Version]
  90. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A Flexible Trimmer for Illumina Sequence Data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [Green Version]
  91. Li, H.; Durbin, R. Fast and Accurate Short Read Alignment with Burrows–Wheeler Transform. Bioinformatics 2009, 25, 1754–1760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Li, H.; Handsaker, B.; Wysoker, A.; Fennell, T.; Ruan, J.; Homer, N.; Marth, G.; Abecasis, G.; Durbin, R. 1000 Genome Project Data Processing Subgroup the Sequence Alignment/Map Format and SAMtools. Bioinformatics 2009, 25, 2078–2079. [Google Scholar] [CrossRef] [Green Version]
  93. Suvakov, M.; Panda, A.; Diesh, C.; Holmes, I.; Abyzov, A. CNVpytor: A Tool for Copy Number Variation Detection and Analysis from Read Depth and Allele Imbalance in Whole-Genome Sequencing. GigaScience 2021, 10, giab074. [Google Scholar] [CrossRef]
  94. Jia, L.; Liu, N.; Huang, F.; Zhou, Z.; He, X.; Li, H.; Wang, Z.; Yao, W. Intansv: An R Package for Integrative Analysis of Structural Variations. PeerJ 2020, 8, e8867. [Google Scholar] [CrossRef]
  95. Krzywinski, M.; Schein, J.; Birol, İ.; Connors, J.; Gascoyne, R.; Horsman, D.; Jones, S.J.; Marra, M.A. Circos: An Information Aesthetic for Comparative Genomics. Genome Res. 2009, 19, 1639–1645. [Google Scholar] [CrossRef] [Green Version]
  96. Pedregosa, F.; Varoquaux, G.; Gramfort, A.; Michel, V.; Thirion, B.; Grisel, O.; Blondel, M.; Prettenhofer, P.; Weiss, R.; Dubourg, V.; et al. Scikit-Learn: Machine Learning in Python. Mach. Learn. 2011, 12, 2825–2830. [Google Scholar]
  97. Felsenstein, J. PHYLIP (Phylogeny Inference Package), version 3.5c. Available online: https://evolution.genetics.washington.edu/phylip.html (accessed on 12 March 2023).
  98. Sherman, B.T.; Hao, M.; Qiu, J.; Jiao, X.; Baseler, M.W.; Lane, H.C.; Imamichi, T.; Chang, W. DAVID: A Web Server for Functional Enrichment Analysis and Functional Annotation of Gene Lists (2021 Update). Nucleic Acids Res. 2022, 50, W216–W221. [Google Scholar] [CrossRef] [PubMed]
  99. Abelenda, J.A.; Cruz-Oró, E.; Franco-Zorrilla, J.M.; Prat, S. Potato StCONSTANS-Like1 Suppresses Storage Organ Formation by Directly Activating the FT-like StSP5G Repressor. Curr. Biol. 2016, 26, 872–881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Martin, A.; Adam, H.; Díaz-Mendoza, M.; Żurczak, M.; González-Schain, N.D.; Suárez-López, P. Graft-Transmissible Induction of Potato Tuberization by the MicroRNA MiR172. Development 2009, 136, 2873–2881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Navarro, C.; Abelenda, J.A.; Cruz-Oró, E.; Cuéllar, C.A.; Tamaki, S.; Silva, J.; Shimamoto, K.; Prat, S. Control of Flowering and Storage Organ Formation in Potato by Flowering Locus T. Nature 2011, 478, 119–122. [Google Scholar] [CrossRef]
  102. Ramírez Gonzales, L.; Shi, L.; Bergonzi, S.B.; Oortwijn, M.; Franco-Zorrilla, J.M.; Solano-Tavira, R.; Visser, R.G.F.; Abelenda, J.A.; Bachem, C.W.B. Potato CYCLING DOF FACTOR 1 and Its LncRNA Counterpart StFLORE Link Tuber Development and Drought Response. Plant J. 2021, 105, 855–869. [Google Scholar] [CrossRef]
  103. Sharma, P.; Lin, T.; Hannapel, D.J. Targets of the StBEL5 Transcription Factor Include the FT Ortholog StSP6A. Plant Physiol. 2016, 170, 310–324. [Google Scholar] [CrossRef] [Green Version]
  104. Teo, C.-J.; Takahashi, K.; Shimizu, K.; Shimamoto, K.; Taoka, K. Potato Tuber Induction Is Regulated by Interactions Between Components of a Tuberigen Complex. Plant Cell Physiol. 2016, 58, pcw197. [Google Scholar] [CrossRef]
  105. Zhou, T.; Song, B.; Liu, T.; Shen, Y.; Dong, L.; Jing, S.; Xie, C.; Liu, J. Phytochrome F Plays Critical Roles in Potato Photoperiodic Tuberization. Plant J. 2019, 98, 42–54. [Google Scholar] [CrossRef]
  106. Dodds, K.S. Classification of cultivated potatoes. In The Potato and Its Wild Relatives; Botanical Studies; Correll, D.S., Ed.; Contributions from Texas Research Foundation: Renner, TX, USA, 1962; Volume 4, pp. 517–539. [Google Scholar]
  107. Hawkes, J.G. The Potato: Evolution, Biodiversity and Genetic Resources. Belhaven Press: London, UK, 1990. [Google Scholar]
Figure 1. Summary assessments for the proportion of genes (a) and orthogroups (a) present in pan-genome of 15 assemblies of Russian potato cultivars. (a) Bar chart for the proportion of core, softcore, shell, and cloud genes in the pan-genome (top bar) and each cultivar. X-axis: the fraction of genes. Y-axis: cultivars denoted by abbreviations. (b) Pie chart for the distribution of core, softcore, shell, and cloud orthogroups in the pan-genome. Color legend for pan-genome parts is shown below the pie chart.
Figure 1. Summary assessments for the proportion of genes (a) and orthogroups (a) present in pan-genome of 15 assemblies of Russian potato cultivars. (a) Bar chart for the proportion of core, softcore, shell, and cloud genes in the pan-genome (top bar) and each cultivar. X-axis: the fraction of genes. Y-axis: cultivars denoted by abbreviations. (b) Pie chart for the distribution of core, softcore, shell, and cloud orthogroups in the pan-genome. Color legend for pan-genome parts is shown below the pie chart.
Ijms 24 05713 g001
Figure 2. Simulation of pan- and core-genome sizes, in terms of number of orthologous gene clusters and pan-genome composition. X-axis is the number of genomes; the Y-axis is the number of clusters. The plot color legend is provided for each part.
Figure 2. Simulation of pan- and core-genome sizes, in terms of number of orthologous gene clusters and pan-genome composition. X-axis is the number of genomes; the Y-axis is the number of clusters. The plot color legend is provided for each part.
Ijms 24 05713 g002
Figure 3. Genetic differentiation within two subsets of South American landraces and Russian potato cultivars accessions, as well as the reference tomato and potato DM1-3 genomes, built by Orthofinder program. Branches leading to Russian cultivars are shown in red, to South American cultivated species accessions—in green. Wild species S. bukasivii (BUK) is shown by magenta branch. The branch corresponding to the reference genome (DM1-3 PGSM v4.03) is shown in blue, the S. lycopersicum outgroup is shown in black. The scale bar is shown in the upper left corner, the numbers at the nodes correspond to the Shimodaira–Hasegawa-like support values.
Figure 3. Genetic differentiation within two subsets of South American landraces and Russian potato cultivars accessions, as well as the reference tomato and potato DM1-3 genomes, built by Orthofinder program. Branches leading to Russian cultivars are shown in red, to South American cultivated species accessions—in green. Wild species S. bukasivii (BUK) is shown by magenta branch. The branch corresponding to the reference genome (DM1-3 PGSM v4.03) is shown in blue, the S. lycopersicum outgroup is shown in black. The scale bar is shown in the upper left corner, the numbers at the nodes correspond to the Shimodaira–Hasegawa-like support values.
Ijms 24 05713 g003
Figure 4. Distribution of CNV quantitative characteristics in the genomes of Russian cultivars and South American potato species accessions. (a) Distribution of the number of duplications and deletions; (b) distribution of maximum lengths of deletions and duplications; (c) distribution of the number of genes with deletions and duplications. See Table 2 for abbreviation of the names of South American accessions.
Figure 4. Distribution of CNV quantitative characteristics in the genomes of Russian cultivars and South American potato species accessions. (a) Distribution of the number of duplications and deletions; (b) distribution of maximum lengths of deletions and duplications; (c) distribution of the number of genes with deletions and duplications. See Table 2 for abbreviation of the names of South American accessions.
Ijms 24 05713 g004
Figure 5. Principal component analysis diagrams for potato genotypes based on the CNV similarity for DM1-3 v4.03 protein-coding genes. (a) PCA plot for PC1 (X-axis) and PC2 (Y-axis) components; (b) PCA plot for PC2 (X-axis) and PC3 (Y-axis) components. Explained percentage of variance shown in parentheses. Selected for this study Russian potato cultivars shown as red dots, and South American accessions shown as green dots. See Table 1 for abbreviation of the names of 15 Russian cultivars and Table 2 for abbreviation of the names of South American accessions.
Figure 5. Principal component analysis diagrams for potato genotypes based on the CNV similarity for DM1-3 v4.03 protein-coding genes. (a) PCA plot for PC1 (X-axis) and PC2 (Y-axis) components; (b) PCA plot for PC2 (X-axis) and PC3 (Y-axis) components. Explained percentage of variance shown in parentheses. Selected for this study Russian potato cultivars shown as red dots, and South American accessions shown as green dots. See Table 1 for abbreviation of the names of 15 Russian cultivars and Table 2 for abbreviation of the names of South American accessions.
Ijms 24 05713 g005
Figure 6. Tree diagram reconstructed by potato genomes using their CNV similarity. Cultivars from Russia are shown by red branch color and South American potato accessions shown by green branch color.
Figure 6. Tree diagram reconstructed by potato genomes using their CNV similarity. Cultivars from Russia are shown by red branch color and South American potato accessions shown by green branch color.
Ijms 24 05713 g006
Table 1. Genome assembly statistics of potato varieties cultivars grown in Russia after filtering of short contigs (<1000 bp) without ORFs.
Table 1. Genome assembly statistics of potato varieties cultivars grown in Russia after filtering of short contigs (<1000 bp) without ORFs.
CultivarAbbreviationContig NumberAssembly Size, bpGC, %Largest Contig Length, bpN50L50Total Repeats, %
FritellaFri163,5621,222,793,32135.10212,76915,37919,16261.15
GolubiznaGolu257,855960,248,39235.83210,050955517,35460.82
GrandGrand155,4611,170,073,53235.21246,54115,50518,18660.86
GusarGus269,494938,553,97536.22178,581809520,02061.70
Krasa MeshcheryKrMe217,5191,297,174,75735.54162,81512,47623,86962.65
KrasavchikKras202,1871,196,647,23235.61192,44912,88620,74261.68
KrepyshKrep252,102994,694,66735.58280,787914819,97861.84
MeteorMet242,362877,113,92335.59164,801997516,04160.46
NevskyNev209,314843,177,79835.63176,08211,81613,71660.45
NikulinskyNik193,0331,149,833,87334.83204,47811,64922,01360.75
Severnoe SiyanieSevS207,7811,260,128,86635.64222,66513,42321,16062.38
SudarinyaSud259,781894,173,93636.03204,234942616,51361.34
SymphoniaSymph176,3501,119,406,98634.92216,39912,27221,68561.08
UdachaUda204,824652,821,26036.09236,853815114,11459.72
ZhukovskyZhu214,788899,426,46935.47202,35112,04914,40060.63
Table 2. Orthogroup statistics for ORFs from the genomes of Russian potato cultivars, South American accessions, potato DM1-3, and S. lycopersicum.
Table 2. Orthogroup statistics for ORFs from the genomes of Russian potato cultivars, South American accessions, potato DM1-3, and S. lycopersicum.
AccessionORF NumberNumber of ORFs in OrthogroupsFraction of ORFs in Orthogroups, %Number of Unassigned ORFsFraction of Unassigned ORFs, %
Subset of Russian cultivars:
Fritella73,20972,21298.69971.4
Golubizna72,45771,31998.411381.6
Grand70,30468,98998.113151.9
Gusar71,07969,89098.311891.7
Krasa Meshchery77,41776,04198.213761.8
Krasavchik73,81572,54598.312701.7
Krepysh68,36267,43298.69301.4
Meteor68,16567,08098.410851.6
Nevsky62,91661,93498.49821.6
Nikulinsky73,66772,72798.79401.3
Severnoe Siyanie75,70174,31698.213851.8
Sudarinya68,86467,56398.113011.9
Symphonia69,78368,93698.88471.2
Udacha60,41159,41698.49951.6
Zhukovsky64,38663,35298.410341.6
Subset of South American accessions 1:
AJH75,55574,64498.89111.2
BUK101,26798,90797.723602.3
CHA76,30175,52199.07801.0
PHU79,45178,71599.17360.9
GON155,93555,39799.05381.0
GON283,06880,51896.925503.1
STN88,44887,33498.711141.3
TBR119,470114,61695.948544.1
ADG147,96947,82299.71470.3
ADG295,32293,62798.216951.8
CUR119,710116,69197.530192.5
JUZ92,02290,68698.513361.5
S.tuberosum DM1-3 v. 40339,02837,82896.912003.1
S. lycopersicum34,42931,15990.532709.5
1 Subset of 12 South American accessions includes representatives of S. tuberosum Andigenum group (STN, PHU, GON1, GON2, CHA, ADG1, ADG2), S. tuberosum Chilotanum group (TBR), S. ajanhuiri (AJH), S. curtilobum (CUR), S. juzepczukii (JUZ), S. bukasovii (BUK) [17], see also Materials and methods Section 4.3.
Table 3. The number and fraction of protein-coding genes in genomes of Russian cultivars annotated by InterproScan program.
Table 3. The number and fraction of protein-coding genes in genomes of Russian cultivars annotated by InterproScan program.
CultivarNumber of Functionally
Annotated ORFs
Fraction of Functionally
Annotated ORFs, %
Fritella42,18758.42
Golubizna38,42853.88
Grand40,73159.04
Gusar37,17053.18
Krasa Meshchery40,97853.89
Krasavchik42,14458.09
Krepysh36,79154.56
Meteor36,32354.16
Nevsky34,39155.53
Nikulinsky40,84056.16
Severnoe siyanie41,93356.42
Sudarinya36,38653.85
Symphonia39,01256.59
Udacha32,24754.27
Zhukovsky35,34455.79
Table 4. Number of full-length NBS-LRR genes of different classes identified in genomes of Russian potato cultivars and reference genome DM1-3.
Table 4. Number of full-length NBS-LRR genes of different classes identified in genomes of Russian potato cultivars and reference genome DM1-3.
CultivarCNL TypeTNLn/aTotal
12345678R
Fritella27470782132015190302
Golubizna15400371421212118187
Grand296115112311317191298
Gusar1342143180911117182
Krasa Meshchery245305101251119148242
Krasavchik159721191731512181281
Krepysh21452351331214147229
Meteor173106101911013112192
Nevsky124017111421112111185
Nikulinsky15660381751711163251
Severnoe Siyanie1582211102541720192306
Sudarinya19410351421211110181
Symphonia255315111721714152252
Udacha113114612010483135
Zhukovsky9520351511210119181
Total by class267744111801192513419819521343404
S.tuberosum DM1-316132093123232432660257
Table 5. The list of genes associated with tuberization and photoperiod control, for which CNV occurrence in genomes of Russian and South American accessions is significantly different. Gene IDs, number of significant CNV of deletion/duplication types in compared subsets, p-values (Fisher’s test), and gene description are given.
Table 5. The list of genes associated with tuberization and photoperiod control, for which CNV occurrence in genomes of Russian and South American accessions is significantly different. Gene IDs, number of significant CNV of deletion/duplication types in compared subsets, p-values (Fisher’s test), and gene description are given.
Gene IDRU del/dupSA del/dupp-ValueGene Description
PGSC0003DMG4000006780/24/10.005Metallocarboxy-peptidase inhibitor 1
PGSC0003DMG4000128384/07/00.009Non-specific lipid-transfer protein 1
PGSC0003DMG4000232720/04/00.008Elongation factor 1-alpha 1
PGSC0003DMG4000157661/84/00.010Phytochrome A 2
PGSC0003DMG4000293610/150/30.002Poly(ADP-ribose) glycohydrolase 3
1 See ref. [25], 2 See ref. [26], 3 This work.
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

Karetnikov, D.I.; Vasiliev, G.V.; Toshchakov, S.V.; Shmakov, N.A.; Genaev, M.A.; Nesterov, M.A.; Ibragimova, S.M.; Rybakov, D.A.; Gavrilenko, T.A.; Salina, E.A.; et al. Analysis of Genome Structure and Its Variations in Potato Cultivars Grown in Russia. Int. J. Mol. Sci. 2023, 24, 5713. https://doi.org/10.3390/ijms24065713

AMA Style

Karetnikov DI, Vasiliev GV, Toshchakov SV, Shmakov NA, Genaev MA, Nesterov MA, Ibragimova SM, Rybakov DA, Gavrilenko TA, Salina EA, et al. Analysis of Genome Structure and Its Variations in Potato Cultivars Grown in Russia. International Journal of Molecular Sciences. 2023; 24(6):5713. https://doi.org/10.3390/ijms24065713

Chicago/Turabian Style

Karetnikov, Dmitry I., Gennady V. Vasiliev, Stepan V. Toshchakov, Nikolay A. Shmakov, Mikhail A. Genaev, Mikhail A. Nesterov, Salmaz M. Ibragimova, Daniil A. Rybakov, Tatjana A. Gavrilenko, Elena A. Salina, and et al. 2023. "Analysis of Genome Structure and Its Variations in Potato Cultivars Grown in Russia" International Journal of Molecular Sciences 24, no. 6: 5713. https://doi.org/10.3390/ijms24065713

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