Next Article in Journal
Micronutrient Status and Selected Physiological Parameters of Roots in Nickel-Exposed Sinapis alba L. Affected by Different Sulphur Levels
Next Article in Special Issue
Structure–Function Analysis Reveals the Singularity of Plant Mitochondrial DNA Replication Components: A Mosaic and Redundant System
Previous Article in Journal
Developing Chloroplast Genomic Resources from 25 Avena Species for the Characterization of Oat Wild Relative Germplasm
Previous Article in Special Issue
De Novo Assembly Discovered Novel Structures in Genome of Plastids and Revealed Divergent Inverted Repeats in Mammillaria (Cactaceae, Caryophyllales)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Organization Features of the Mitochondrial Genome of Sunflower (Helianthus annuus L.) with ANN2-Type Male-Sterile Cytoplasm

by
Maksim S. Makarenko
1,2,*,
Alexander V. Usatov
1,
Tatiana V. Tatarinova
2,3,4,5,
Kirill V. Azarin
1,
Maria D. Logacheva
2,6,
Vera A. Gavrilova
7,
Igor V. Kornienko
1,8 and
Renate Horn
9
1
Department of Genetics, Southern Federal University, Rostov-on-Don 344006, Russia
2
The Institute for Information Transmission Problems, Moscow 127051, Russia
3
Department of Biology, University of La Verne, La Verne, CA 91750, USA
4
Vavilov Institute of General Genetics, Moscow 119333, Russia
5
School of Fundamental Biology and Biotechnology, Siberian Federal University, Krasnoyarsk 660041, Russia
6
Skolkovo Institute of Science and Technology, Moscow 121205, Russia
7
The N.I. Vavilov All-Russian Institute of Plant Genetic Resources, Saint Petersburg 190121, Russia
8
Southern Scientific Center of the Russian Academy of Sciences, Rostov-on-Don 344006, Russia
9
Institute of Biological Sciences, Plant Genetics, University of Rostock, 18059 Rostock, Germany
*
Author to whom correspondence should be addressed.
Plants 2019, 8(11), 439; https://doi.org/10.3390/plants8110439
Submission received: 18 August 2019 / Revised: 18 October 2019 / Accepted: 19 October 2019 / Published: 23 October 2019
(This article belongs to the Special Issue Plant Organelle DNA Maintenance)

Abstract

:
This study provides insights into the flexibility of the mitochondrial genome in sunflower (Helianthus annuus L.) as well as into the causes of ANN2-type cytoplasmic male sterility (CMS). De novo assembly of the mitochondrial genome of male-sterile HA89(ANN2) sunflower line was performed using high-throughput sequencing technologies. Analysis of CMS ANN2 mitochondrial DNA sequence revealed the following reorganization events: twelve rearrangements, seven insertions, and nine deletions. Comparisons of coding sequences from the male-sterile line with the male-fertile line identified a deletion of orf777 and seven new transcriptionally active open reading frames (ORFs): orf324, orf327, orf345, orf558, orf891, orf933, orf1197. Three of these ORFs represent chimeric genes involving atp6 (orf1197), cox2 (orf558), and nad6 (orf891). In addition, orf558, orf891, orf1197, as well as orf933, encode proteins containing membrane domain(s), making them the most likely candidate genes for CMS development in ANN2. Although the investigated CMS phenotype may be caused by simultaneous action of several candidate genes, we assume that orf1197 plays a major role in developing male sterility in ANN2. Comparative analysis of mitogenome organization in sunflower lines representing different CMS sources also allowed identification of reorganization hot spots in the mitochondrial genome of sunflower.

1. Introduction

Low substitution rate in genes along with considerable variability in size and structure are distinct features of plant mitochondrial genomes (mitogenome) [1,2]. Reorganization events in mitochondrial DNA (mtDNA) are primarily caused by disruption of a fragile equilibrium of intramolecular recombinations, maintained by nuclear-mitochondrial genetic interactions [3,4]. Runaway recombination of mtDNA can lead to changes in gene content and expression patterns of mitochondria [5,6]. The mitogenomes of flowering plants carry genes for rRNAs, tRNA, subunits of the respiratory chain complexes, as well as genes for the ribosomal proteins (rps and rpl). Maturase-related protein gene (matR) and genes responsible for the biogenesis of cytochrome c (ccmB, ccmC, ccmFC, and ccmFN) are also part of the plant mitochondrial gene set [7]. Alterations in transcription activity of the mitochondrial genes can have profound effects on the functionality of mitochondria and, thus, on different plant traits. Among the phenotypic traits caused by mitochondrial impairments, special attention is devoted to cytoplasmic male sterility (CMS) [8]. CMS is an inability to produce or shed functional pollen, which has been described in more than 140 higher plant species [9]. As a result of mtDNA rearrangements, new open reading frames are created, leading to male sterility [10]. In turn, dominant nuclear-encoded restorer-of-fertility genes (Rf genes) can restore normal development of pollen. Hence CMS/Rf systems are important both for studying pollen development in plants and for commercial applications [11,12]. The existence of the CMS phenotype in plants eliminates the need for laborious, manual emasculations for a directional crossing of plants, thus promoting its utilization in hybrid breeding [13].
Comparing mtDNA configurations in sunflower is especially interesting, as more than 70 CMS sources have been described within the Food and Agriculture Organization (F.A.O.) program for sunflower [14], even though modern sunflower hybrid breeding predominantly relies on a single male-sterile cytoplasm, the so-called CMS PET1 [15]. This CMS source originated from an interspecific cross of H. petiolaris with H. annuus [16]. The molecular characterization of the CMS mechanism helps to introduce new CMS sources into breeding programs. So far, the mitogenomes of only a few CMS sources have been sufficiently characterized to be used in sunflower hybrid production. The CMS phenotype can arise spontaneously in wild populations, while in breeding lines—after wide crosses, interspecific exchange of nuclear and/or cytoplasmic genomes, and mutagenesis [11]. It has been demonstrated that some CMS sources obtained from different inter- or intraspecific crosses showed the same mechanism of male sterility formation as the CMS PET1 type [17]. Even though these CMS sources had different origins, they have the same mitochondrial genome organization indicating that some configurations may be preferentially maintained in sunflowers [17].
Less is known about the spontaneously occurring CMS sources in Helianthus annuus L. [14]. ANN2 was derived from wild sunflower population N517 in Texas [18]. In this population, 40% of the plants were male-sterile [19]. However, sunflower cultivars with ANN2 CMS type, developed from the N517 population by maintaining with lines like HA89 or RHA265, showed 100% male-sterile progenies, which indicates a stable mitochondrial DNA configuration and absence of heteroplasmy. ANN2 and other spontaneously occurring CMS sources like ANN1, ANN3, and ANN4 maintained by RHA265 were hardly restored [20]. None of the tested maintainer and restorer lines of CMS PET1 were able to restore pollen production in CMS ANN1 and CMS ANN3. Only 12.5% and 15.8% of all investigated lines showed restorer capacity towards CMS ANN2 and CMS ANN4, respectively, indicating very different CMS mechanisms compared to CMS PET1. Three restorer lines, Rf ANN2-PI 413178, Rf ANN2-P21, and Rf ANN2-RMAX1, carry a restorer-of-fertility gene for ANN2 [21,22]. A suppressor gene S1 overpowering the restorer gene action has been recently described by Liu et al. [23], thus making the CMS-Rf interactions in the ANN2 source even more complicated.
Previous mtDNA investigations of some spontaneously occurring CMS sources were based on Southern blot hybridizations with mitochondrial genes [24]. Hybridizations of the CMS sources ANN1, ANN2, ANN3 with atp6, atp9, cob, cox1, cox2, cox3, rrn18/rrn5/nad5, orfH522, orfH708, or orfH873 as a probe revealed unique banding patterns for 4 out of the 10 probes [24]. Besides, the analyses showed that CMS ANN4 and ANN5 are very similar to each other and form a distinct group from CMS ANN1, ANN2, and ANN3 [24]. It was also shown that ANN1/ANN2/ANN3 mtDNA-type significantly differs from both the male-fertile sunflower cytoplasm and the CMS PET1 source [24].
We describe the first assembly of the CMS source ANN2, which occurred spontaneously in Helianthus annuus L. The current study also provides insights into the flexibility of sunflower mitochondrial genome by comparing different isonuclear male-sterile lines HA89 (ANN2, MAX1, PET1, PET2) and the male-fertile line HA89, allowing identification of hot spots for rearrangements in the sunflower mitochondrial genome. For the CMS mechanism in ANN2, new open reading frames were identified, which were transcriptionally active. The ANN2 CMS source may be interesting not only for oilseed hybrid breeding, but also for horticultural purposes, as it is difficult to restore. It is a highly desirable trait in ornamental sunflowers since pollen production is usually not required nor looked-for, except if the pollen color enhances the contrast with the florets.

2. Results

2.1. Rearrangements in the Mitochondrial Genome of the Male-Sterile Line HA89(ANN2)

We assembled the complete mitochondrial genome of the HA89 sunflower line with ANN2 cytoplasmic sterility type (NCBI accession MN175741.1). The master chromosome of HA89 (ANN2) consists of 306,018 bp (Figure 1), and it is 5071 bp longer than the mitogenome of the male-fertile isonuclear line HA89 (NCBI accession MN171345.1).
The HA89 (ANN2) mitochondrial genome has a wide range of rearrangements as compared to the male-fertile HA89 mitogenome. The summary of whole mitogenome alignment of male-sterile and male-fertile lines is presented in Table 1.
The mitochondrial genomes of male-sterile ANN2 and male-fertile HA89 share 14 complementary regions, but their localizations and orientation differ. We showed the localization of complementary regions in a scheme with both genomes shown in linear forms in Figure 2. Since regions #1 and #14 in the case of the circular molecule represent the same region, we classified the other twelve regions (#2–#13) as rearrangements.
In most cases, rearrangements only resulted in a reversal of a gene’s direction or a change in gene order. However, the 8584 bp (#7) and 21,433 bp (#8) rearrangements influenced the coding sequence of nad6, and the 6029 bp rearrangement (#12) impaired atp6. The largest part of the nad6 gene sequence (~88%) is in the rearrangement #8, while the 3′ terminal part of nad6 lies in the rearrangement #7. As a result of the convergence of #8 and #10 rearrangements in mitochondrial DNA of HA89(ANN2), the new nad6-chimeric open reading frame—orf891—was created. Analyses of orf891 transcription pattern gave ambiguous results. Transcripts were detected for both nad6 (HA89) and orf891 (HA89(ANN2)) when using primers derived from the 5′ identical sequence of their mRNA (Supplementary Table S1). Nevertheless, using the same forward primer, but different reverse primers (Supplementary Table S1), complementary to the 3′ sequence of nad6 and orf891, transcription was detected only for nad6 (the fertile line), but not for orf891 (CMS line). It is important to note that almost all the rearrangements found in mtDNA of HA89(ANN2) are accompanied by other types of genome reorganizations—deletions and insertions.

2.2. Deletions and Insertions in the Mitochondrial Genome of the Male-Sterile Line HA89(ANN2)

In comparison to the male-fertile analog, we identified nine longer than 100 bp deletions in the mtDNA of HA89(ANN2), which are shown in Table 2. Most deletions did not affect the protein-coding sequences, except for two deletions of 316 bp and 1165 bp. The 1165 bp deletion resulted in the total elimination of orf777, while the 316 bp deletion affected the part of the atp6 gene. Interestingly, in previous studies, we also discovered the removal of orf777 from the mitochondrial genomes of two other CMS lines—HA89(PET2) [25] and HA89(MAX1) [26].
Seven longer than 100 bp insertions were detected in mtDNA of the HA89(ANN2) CMS line (Table 3). As a result of these insertions in the mitochondrial DNA of HA89(ANN2), five new open reading frames, namely orf324, orf327, orf345, orf558, and orf933, have appeared. All five ORFs are transcribed in the case of ANN2, contrary to the HA89 line.
A search for transmembrane domains (TDs) revealed that the protein encoded by orf558 contained a single TD. In the case of orf933, two TDs were detected. The orf933 encoded protein did not show homology to other sunflower proteins in GeneBank, and had only limited similarity (40–60 amino acids) to hypothetical mitochondrial proteins with unknown functions in Lactuca sativa (accession PLY70338.1), Salvia miltiorrhiza (accession YP_008992338.1), Beta vulgaris (accession CBJ23356.1), etc. Forty-six amino acids of the N-terminus of the protein encoded by orf558 matched the N-terminus of cytochrome c oxidase subunit 2 (cox2 gene). Moreover, most of the amino acids that form the transmembrane domain in orf558 protein are identical to those in COX2. However, the sunflower cytochrome c oxidase subunit 2 has two TD and the protein encoded by orf558—only a single one (Figure 3). So the orf558 represents a chimeric cox2 gene and could potentially play a role in the ANN2 CMS phenotype development.
The most complex among the discovered ORFs in the HA89(ANN2) mitogenome was orf1197, which has appeared from three simultaneous reorganization events involving the 316 bp deletion, the 430 bp insertion, and the 6029 bp rearrangement. The orf1197 represents a chimeric atp6 gene, with transcription activity specific for the CMS line HA89(ANN2). In sunflower, the atp6 gene typically encodes a protein consisting of 351 aa, whereas the predicted size of the translation product of the orf1197 is 399 aa. Both proteins share 251 identical amino acids in the C-terminus. Thus, the protein encoded by the orf1197 carries all seven TDs present in the C-terminus of the ATP synthase Fo subunit 6 from mitochondria of male-fertile sunflower (Figure 4).

3. Discussion

Recently, we had investigated complete mitochondrial DNA sequences for three CMS sources in sunflower, coming from interspecific crosses: PET1, PET2, MAX1 [25,26]. Comparison of the HA89(ANN2) mitogenome with mitochondrial genome assemblies of male-fertile lines [25,27] and the other HA89 male-sterile analogs provides insights into reorganizations of mtDNA associated with CMS phenotypes. The male-fertile lines (HA89, HA412) have only slight variations in the mtDNA sequence [25]. Whereas the mitogenomes of the CMS sources (HA89(ANN2), HA89(PET1), (HA89(PET2), HA89(MAX1)) showed significant differences as compared to their alloplasmic male-fertile analog. The complete mitochondrial genome of the male-fertile line HA89 adds up to 300,947 bp (NCBI accession MN171345.1), while HA89(PET1) has a size of 305,217 bp (NCBI accession MG735191.1), HA89(PET2) of 316,582 bp (NCBI accession MG770607.2), HA89(MAX1) of 295,586 bp (NCBI accession MH704580.1) and HA89(ANN2) of 306,018 bp (MN175741.1). The difference in genomes sizes is due to several deletions and insertions. For instance, in the mtDNA of all the investigated CMS sources, except HA89(PET1), a similar deletion in the nad4L-orf777-atp8 region was observed. In the case of HA89(PET2), this is due to a 711 bp deletion, in HA89(MAX1) has a 978 bp deletion, and in HA89(ANN2) there is an 1195 bp deletion. All these deletions resulted in removal of orf777 from the mtDNA in the CMS lines. Another region enriched by deletions is the area between cob-ccmFC, here three overlapping deletions were detected: a 451 bp deletion in HA89(PET1), one of 3780 bp in HA89(PET2) and another one of 4204 bp in HA89(ANN2).
The coincidence between locations of these deletions is not accidental. There are three 265 bp repeats present in the sunflower mitochondrial genome, with the following positions in the mtDNA of the male-fertile HA89 line: 36537-36801 (adjacent to atp8), 190074-190338 (next to cob), and 202902-202638 (between orf873-atp1). These repeat regions are shown by red stars in Figure 5. Repeats represent common recombination points in mtDNA molecules [4,28]. The identification of small repeats involved in recombination is important because they influence the maintenance and evolution of mitochondrial genomes [28]. An imbalance in the nuclear-mitochondrion relationship that may occur in distant hybridizations impairs the recombination of mtDNA sub-genomic molecules, therefore, leading to reorganizations in the mitochondrial master chromosome. For instance, in HA89(PET1) the deletion, insertion, and inversion were mentioned in the cob-atp1 region, directly between two repeats (Figure 5). In HA89(PET2), there were also several rearrangements in hot spots, resulting in the formation of a new gene cluster cob-atp8-cox3, as well as in the translocation of the ccmFC-orf873-atp1 gene cluster into the nad4L-orf777 region, combined with a deletion and huge insertion (Figure 5). In the mtDNA of both lines, HA89(MAX1) and HA89(ANN2), the specific atp1-atp8-cox3 gene order was created, while in the MAX1 CMS source ccmFC-orf873 translocated into the nad4L-orf777 region (with deletion and huge insertion in this region). In the case of ANN2, the ccmFC-orf873 region is located next to the cob gene (Figure 5). Thus, we have established that these three 265 bp repeats represent a reorganization hot spots in the sunflower mitochondrial genome.
Considering the insertions discovered in the HA89(ANN2) mitogenome, we also observed a similarity between the insertions of different CMS sources in sunflower. For instance, about 85% of the 1027 bp insertion sequence (ANN2) is complementary to the part of the 15,885 bp insertion (PET2). The 3757 bp insertion (ANN2) contains 1959 nucleotides identical to the 15,885 bp insertion (PET2) and 1215—to the 5272 bp insertion (MAX1). Also, 2343 bp of the 5338 bp insertion in ANN2 are similar to another region of sunflower mtDNA proximal to the cob gene (position 185,987–188,330 in the mtDNA of HA89 fertile line). So, this sequence is duplicated in the mitochondrial genome of the HA89(ANN2) CMS line. About 10% of 6452 bp insertion (ANN2) is complementary to both the 5050 bp (PET2) and the 5272 bp (MAX1) insertions. As well as 1158 bp of the 9044 bp insertion (ANN2) are identical to the 5050 bp and 15,885 bp insertions (PET2) and 5272 bp insertion (MAX1).
Although there are similarities in deletions, insertions, and rearrangements between mitochondrial genomes of the HA89(ANN2) and other CMS lines, the discovered ORFs were different. We summarized the data of all identified ORFs in the male-sterile cytoplasms in Table 4. The ORFs encoding proteins with similarity to other mitochondrial proteins and especially having transmembrane domains are of particular interest since the chimeric proteins with TD most often cause CMS phenotypes [10,29]. In the mtDNA of HA89(ANN2), we detected three new transcriptionally active ORFs, encoding proteins with TD—orf558 (one TD), orf933 (two TD), orf1197 (seven TD). The orf933 shows no homology to other sunflower genes, while orf558 represents a chimeric cox2 gene, and orf1197 a chimeric atp6 gene. It is difficult to estimate the exact contribution of orf558, orf933, orf1197 to the development of the male-sterile phenotype in ANN2. However, the possibility of involvement of more than one open reading frame might be one explanation that ANN2 is so difficult to restore. On the other hand, the presence of a suppressor gene S1 discovered by Liu et al. [23] might be the reason for low rates of fertility restoration of the ANN2 CMS source. Previous studies indicate that the chimeric atp6 genes or new ORFs that are co-transcribed with atp6 most often cause CMS phenotypes in flowering plants [10,30,31,32,33]. Therefore, we suggest that orf1197 is the major CMS candidate gene for ANN2 CMS source. Moreover, chimeric atp6 genes were also identified in MAX1 [26] and CMS3/ANT1 [34] CMS types of sunflower. In CMS lines, chimeric atp6 genes encode N-terminal extended proteins compared to the normal ATP synthase subunit 6 (351 aa): ANN2—399 aa, MAX1—429 aa, (AYV91168.1), CMS3/ANT1—437 aa (CAA57790.1). Moreover, 397 of 399 amino acids in the orf1197 protein are identical to the chimeric ATP6 of CMS3/ANT1 line, and therefore this protein represents a shorter version of this 437 aa-long protein. Such similarities support our hypothesis about the importance of orf1197 in shaping the CMS phenotype in ANN2.
The orf558 (as the chimeric cox2 gene) might also cause cytoplasmic male sterility in ANN2. In other plants species, modified cox2 sequences seem to be involved in the male sterility. For instance, the CMS specific pcf gene in petunia is composed from sequences of the 5′ portion of atp9, segments of cox2, and a large region of unknown origin—urfS [35]. In wild beets, the CMS-associated orf129 shows homology to the 5′ flanking and the coding sequence of cox2 [36]. In the mitogenome of the inap CMS source of Brassica napus, which was created by a somatic hybridization with Isatis indigotica, a novel cox2-2 gene was detected, which represents recombination of the cox2 of woad and cox2-2 of rapeseed [37].
Another unique feature of HA89(ANN2) mitogenome is the formation of orf891. According to the ORFs predictions, a 3′ elongation of the nad6 gene (orf891) may occur. However, the cDNA analyses did not agree with the genomic data. Perhaps due to nad6 mRNA editing instead of a 3′ elongated transcript, the shorter one is formed. Heteromorphism in nad6 transcript length was also observed in Mimulus guttatus x M. nasutus hybrids with the CMS phenotype [38]. Both male-fertile and male-sterile hybrids have a single copy of the nad6 gene, and the divergence in mRNA length was only observed in male-sterile plants [38].

4. Materials and Methods

4.1. Plant Material

The CMS line HA89(ANN2) of sunflower was obtained from the genetic collection of the N. I. Vavilov All-Russian Institute of Plant Genetic Resources (Saint Petersburg, Russia). The original source of the ANN2 male-sterile cytoplasm was obtained by Serieys in 1984 [18]. All sunflower lines were grown in regularly irrigated pots in the growth chamber KBWF 720 (Binder, Tuttlingen, Germany) with the following growth conditions: temperature—26 °C, humidity—70%, photoperiod—10/14 h (dark/light).

4.2. Mitochondrial DNA Extraction, NGS Library Preparation, and Sequencing

First, the organelle fraction from leaves of 14-days-old sunflower seedlings was isolated, as described by Makarenko et al. [39]. Such preparations significantly reduced the amount of nuclear DNA. Then DNA extraction was performed with PhytoSorb kit (Syntol, Moscow, Russia), according to the manufacturer’s protocol. Equal amounts of DNA from seven plants were mixed, and we used 1 ng of DNA pull for the NGS library preparation step. The library was made with Nextera XT DNA Library Prep Kit (Illumina, Mountain View, CA, USA), following the guidelines of Illumina. The quality of the library was evaluated using Bioanalyzer 2100 (Agilent, Santa Clara, CA, USA). The library was quantified at the Qubit fluorimeter (Invitrogen, Carlsbad, CA, USA) and by qPCR, then diluted up to the concentration of 8 pM. Sequencing was performed on two different Illumina sequencing platforms: HiSeq 2000 using TruSeq SBS Kit v3-HS 200-cycles and MiSeq using MiSeq Reagent Kit v2 500-cycles (Illumina, Mountain View, CA, USA). A total number of 3,063,836 reads (100-bp paired) and 3,305,268 reads (250-bp paired) were generated.

4.3. Mitochondrial Genome Assembly and Annotation

Quality control of reads was done using FastQC (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/). Trimming of adapter-derived and low quality (Q-score below 25) reads was performed with Trimmomatic software [40]. For contig generation, we used SPAdes Genome Assembler v.3.10.1 [41]. The whole mitochondrial genome was manually assembled using scaffolds based on high coverage (depths > 70) contigs (length > 1000 kbp) and available bridge contigs (length = 0.3–1 kbp). The genome assembly was validated by remapping the initially obtained reads using Bowtie 2 v.2.3.3 [42]. All observed rearrangements were verified by PCR analysis. For variant calling, we used samtools/bcftools software [43] and manually revised polymorphic sites using the IGV tool [44].
The mitochondrial genome was annotated with MITOFY [45], BLAST tool [46], and ORFfinder (https://www.ncbi.nlm.nih.gov/orffinder). Using GeSeq [47], we provided comparisons of our annotation with the current reference annotations (NCBI accessions NC_023337.1, CM007908.1) of the sunflower mitochondrial genome. Graphical genome maps were generated using the OGDRAW tool v.1.3.1 [48]. The prediction of transmembrane domains was made with TMHMM Server v.2.0 (http://www.cbs.dtu.dk/services/TMHMM-2.0/). The scheme of the bioinformatic pipeline is presented in Supplementary Figure S1.

4.4. Expression Analyses

RNA was extracted from leaves of seven 28-days-old sunflower plants using a guanidinium thiocyanate-phenol-chloroform based method with the ExtractRNA reagent kit (Evrogen, Moscow, Russia). The quality and concentration of the RNA were evaluated with the Qubit fluorimeter (Invitrogen, Carlsbad, CA, USA) and the NanoDrop 2000 spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA). Total RNA (0.5 μg) was treated with DNAse I (Thermo Fisher Scientific, Waltham, MA, USA), and then cDNA was synthesized using the MMLV RT kit (Evrogen, Moscow, Russia) with random primers. As a negative control for each DNAse treated mRNA sample, the same reverse transcription protocol was performed, but without the MMLV enzyme. The quantitative PCR was performed with EvaGreen based RT-PCR kit (Syntol, Moscow, Russia) on Rotor-Gene 6000 (Corbett Research, Mortlake, NSW, Australia). A summary of all primer sequences is given in Supplementary Table S1. The primers annealing temperature (Tm) was 60 °C.

5. Conclusions

Assembly of CMS ANN2 mitochondrial genome (HA89(ANN2) line) revealed several rearrangements, insertions, and deletions, as well as seven new open reading frames: orf324, orf327, orf345, orf558, orf891, orf933, and orf1197. Transcripts were detected for all new open reading frames in CMS ANN2, but not in the fertile cytoplasm. Only orf558, orf891, orf933, and orf1197 encoded proteins that contained membrane domains, making them the most likely CMS candidate genes for the ANN2 source. Notably, orf1197 represents a chimeric atp6 gene and presumably plays a major role in the CMS phenotype development of ANN2. However, CMS ANN2 may be caused by simultaneous action of several candidate genes. Hot spots for rearrangements (265 bp repeats) were identified, and we propose that they influence the maintenance and evolution of the mitochondrial genome in sunflower.

Supplementary Materials

The following are available online at https://www.mdpi.com/2223-7747/8/11/439/s1, Suppl Table S1. The primers sets used for transcription activity analysis, Suppl Table S2. The protein coding genes annotated in HA89(ANN2) mitochondrial genome, Suppl Figure S1. Bioinformatic pipeline for mitochondrial genome analysis. Trimmed reads are available at https://doi.org/10.6084/m9.figshare.8945084.v1.

Author Contributions

Conceptualization, K.V.A. and V.A.G.; methodology, M.S.M. and I.V.K.; software, T.V.T. and M.D.L.; validation, M.S.M., K.V.A. and I.V.K.; formal analysis, M.S.M. and T.V.T.; investigation, M.S.M., K.V.A. and M.D.L.; resources, V.A.G.; data curation, M.S.M. and M.D.L.; writing—original draft preparation, M.S.M., T.V.T. and R.H.; writing—review and editing, M.S.M., T.V.T. and R.H.; visualization, R.H.; supervision, A.V.U. and R.H.; project administration, A.V.U. and V.A.G.; funding acquisition, A.V.U. and M.D.L.

Funding

The study was supported by the Ministry of Education and Science of Russia project no. 6.929.2017/4.6. The NGS sequencing was provided with the support of budgetary subsidy to IITP RAS (Laboratory of Plant Genomics 0053-2019-0005). Analytical work was carried out on the equipment of centers for collective use of Southern Federal University “High Technology”.

Acknowledgments

We are grateful to the cooperators of Northern Crop Science Laboratory (Fargo, USA) for providing seeds of HA89 alloplasmic lines of sunflower to the genetic collection of the N. I. Vavilov Institute of Plant Genetic Resources. We thank reviewers and the editor for their suggestions and comments, which has helped to improve the manuscript, and Todd Lorenz (tlorenz@laverne.edu), Duane R. Chartier (drc@authentica.org) for editing assistance.

Conflicts of Interest

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

References

  1. Yang, J.; Liu, G.; Zhao, N.; Chen, S.; Liu, D.; Ma, W.; Hu, Z.; Zhan, M. Comparative mitochondrial genome analysis reveals the evolutionary rearrangement mechanism in Brassica. Plant Biol. 2015, 18, 527–536. [Google Scholar] [CrossRef]
  2. Wang, S.; Song, Q.; Li, S.; Hu, Z.; Dong, G.; Song, C.; Huang, H.; Liu, Y. Assembly of a Complete Mitogenome of Chrysanthemum nankingense Using Oxford Nanopore Long Reads and the Diversity and Evolution of Asteraceae Mitogenomes. Genes 2018, 9, 547. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, H.; Cui, P.; Zhan, K.; Lin, Q.; Zhuo, G.; Guo, X.; Ding, F.; Yang, W.; Liu, D.; Hu, S.; et al. Comparative analysis of mitochondrial genomes between a wheat K-type cytoplasmic male sterility (CMS) line and its maintainer line. BMC Genom. 2011, 12, 163. [Google Scholar] [CrossRef] [PubMed]
  4. Storchova, H.; Stone, J.D.; Sloan, D.B.; Abeyawardana, O.A.J.; Muller, K.; Walterova, J.; Pazoutova, M. Homologous recombination changes the context of Cytochrome b transcription in the mitochondrial genome of Silene vulgaris KRA. BMC Genom. 2018, 19, 874. [Google Scholar] [CrossRef] [PubMed]
  5. Marechal, A.; Brisson, N. Recombination and the maintenance of plant organelle genome stability. New Phytol. 2010, 186, 299–317. [Google Scholar] [CrossRef]
  6. Gualberto, J.M.; Mileshina, D.; Wallet, C.; Niazi, A.K.; Weber-Lotfi, F.; Dietrich, A. The plant mitochondrial genome: Dynamics and maintenance. Biochimie 2013, 100, 107–120. [Google Scholar] [CrossRef]
  7. Zervas, A.; Petersen, G.; Seberg, O. Mitochondrial genome evolution in parasitic plants. BMC Evol. Biol. 2019, 19, 87. [Google Scholar] [CrossRef]
  8. Ding, B.; Hao, M.; Mei, D.; Zaman, Q.U.; Sang, S.; Wang, H.; Wang, W.; Fu, L.; Cheng, H.; Hu, Q. Transcriptome and Hormone Comparison of Three Cytoplasmic Male-sterile Systems in Brassica napus. Int. J. Mol. Sci. 2018, 19, 4022. [Google Scholar] [CrossRef]
  9. Laser, K.D.; Lersten, N.R. Anatomy and cytology of microsporogenesis in cytoplasmic male sterile angiosperms. Bot. Rev. 1972, 33, 337–346. [Google Scholar] [CrossRef]
  10. Horn, R.; Gupta, K.J.; Colombo, N. Mitochondrion role in molecular basis of cytoplasmic male sterility. Mitochondrion 2014, 19, 198–205. [Google Scholar] [CrossRef]
  11. Chen, L.; Liu, Y.-G. Male sterility and fertility restoration in crops. Annu. Rev. Plant Biol. 2014, 65, 579–606. [Google Scholar] [CrossRef] [PubMed]
  12. Wu, Z.; Hu, K.; Yan, M.; Song, L.; Wen, J.; Ma, C.; Shen, J.; Fu, T.; Yi, B.; Tu, J. Mitochondrial genome and transcriptome analysis of five alloplasmic male-sterile lines in Brassica juncea. BMC Genom. 2019, 20, 348. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, Z.; Zhao, N.; Li, S.; Grover, C.E.; Nie, H.; Wendel, J.F.; Hua, J. Plant Mitochondrial Genome Evolution and Cytoplasmic Male Sterility. Crit. Rev. Plant Sci. 2017, 36, 55–69. [Google Scholar] [CrossRef]
  14. Serieys, H. Identification, study and utilisation in breeding programs of new CMS sources in the FAO subnetwork. In Proceedings of the 2005 Sunflower Subnetwork Progress Report, Novi Sad, Serbia and Montenegro, 17–20 July 2005; FAO: Rome, Italy, 2005; pp. 47–53. [Google Scholar]
  15. Vear, F. Changes in sunflower breeding over the last fifty years. OCL 2016, 23, D202. [Google Scholar] [CrossRef] [Green Version]
  16. Leclercq, P. Une sterilite male chez le tournesol. Ann. Amelior. Plant. 1969, 19, 99–106. [Google Scholar]
  17. Horn, R.; Friedt, W. CMS sources in sunflower: Different origin but same mechanism? Theor. Appl. Genet. 1999, 98, 195–201. [Google Scholar] [CrossRef]
  18. Serieys, H. Report on the activities of the F.A.O. working group: “Identification, study and utilization in breeding programs of new CMS sources”, for the period 1991–1993. HELIA 1994, 17, 93–102. [Google Scholar]
  19. Serieys, H.; Vincourt, P. Caracterisation de nouvelles sources de sterility male chez le tournesol. Les Colloq. lNRA Paris 1987, 45, 53–64. [Google Scholar]
  20. Horn, R.; Friedt, W. Fertility restoration of new CMS sources in sunflower (Helianthus annum L.). Plant Breed. 2006, 116, 317–322. [Google Scholar] [CrossRef]
  21. Jan, C.C. Cytoplasmic male sterility in two wild Helianthus annuus L. accessions and their fertility restoration. Crop Sci. 2000, 40, 1535–1538. [Google Scholar] [CrossRef]
  22. Jan, C.C. Silencing of fertility restoration genes in sunflower. HELIA 2003, 26, 1–6. [Google Scholar] [CrossRef]
  23. Liu, Z.; Long, Y.; Xu, S.S.; Seiler, G.; Jan, C.C. Unique fertility restoration suppressor genes for male-sterile CMS ANN2 and CMS ANN3 cytoplasms in sunflower (Helianthus annuus L.). Mol. Breed. 2019, 39, 22. [Google Scholar] [CrossRef]
  24. Horn, R. Molecular diversity of male sterility inducing and male-fertile cytoplasms in the genus Helianthus. Theor. Appl. Genet. 2002, 104, 562–570. [Google Scholar] [CrossRef] [PubMed]
  25. Makarenko, M.; Kornienko, I.; Azarin, K.; Usatov, A.; Logacheva, M.; Markin, N.; Gavrilova, V. Mitochondrial genomes organization in alloplasmic lines of sunflower (Helianthus annuus) with various types of cytoplasmic male sterility. PeerJ 2018, 6, e5266. [Google Scholar] [CrossRef] [PubMed]
  26. Makarenko, M.S.; Usatov, A.V.; Tatarinova, T.V.; Azarin, K.V.; Logacheva, M.D.; Gavrilova, V.A.; Horn, R. Characterization of the mitochondrial genome of the MAX1 type of cytoplasmic male-sterile sunflower. BMC Plant Biol. 2019, 19, 51. [Google Scholar] [CrossRef]
  27. Grassa, C.J.; Ebert, D.P.; Kane, N.C.; Rieseberg, L.H. Complete Mitochondrial Genome Sequence of Sunflower (Helianthus annuus L.). Genome Announc. 2016, 4, e00981-16. [Google Scholar] [CrossRef]
  28. Wynn, E.L.; Christensen, A.C. Repeats of Unusual Size in Plant Mitochondrial Genomes: Identification, Incidence and Evolution. G3 Genes Genomes Genet. 2019, 9, 549–559. [Google Scholar] [CrossRef] [Green Version]
  29. Mower, J.P.; Case, A.L.; Floro, E.R.; Willis, J.H. Evidence against Equimolarity of Large Repeat Arrangements and a Predominant Master Circle Structure of the Mitochondrial Genome from a Monkeyflower (Mimulus guttatus) Lineage with Cryptic CMS. Genome Biol. Evol. 2012, 4, 670–686. [Google Scholar] [CrossRef]
  30. Yamamoto, M.P.; Kubo, T.; Mikami, T. The 5’-leader sequence of sugar beet mitochondrial atp6 encodes a novel polypeptide that is characteristic of Owen cytoplasmic male sterility. MGG 2005, 273, 342–349. [Google Scholar] [CrossRef]
  31. Kim, D.H.; Kim, B.D. The organization of mitochondrial atp6 gene region in male-fertile and CMS lines of pepper (Capsicum annuum L.). Curr. Genet. 2006, 49, 59–67. [Google Scholar] [CrossRef]
  32. Jing, B.; Heng, S.; Tong, D.; Wan, Z.; Fu, T.; Tu, J.; Ma, C.; Yi, B.; Wen, J.; Shen, J. A male sterility-associated cytotoxic protein ORF288 in Brassica juncea causes aborted pollen development. J. Exp. Bot. 2012, 63, 1285–1295. [Google Scholar] [CrossRef] [PubMed]
  33. Tan, G.F.; Wang, F.; Zhang, X.Y.; Xiong, A.I. Different lengths, copies and expression levels of the mitochondrial atp6 gene in male-sterile and fertile lines of carrot (Daucus carota L.). Mitochondrial DNA Part A 2017, 29, 446–454. [Google Scholar] [CrossRef] [PubMed]
  34. Spassova, M.; Moneger, F.; Leaver, C.J.; Petrov, P.; Atanassov, A.; Nijkamp, H.J.; Hille, J. Characterisation and expression of the mitochondrial genome of a new type of cytoplasmic male-sterile sunflower. Plant Mol. Biol. 1994, 26, 1819–1831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Young, E.G.; Hanson, M.R. A fused mitochondrial gene associated with cytoplasmic male sterility is developmentally regulated. Cell 1987, 50, 41–49. [Google Scholar] [CrossRef]
  36. Yamamoto, M.P.; Shinada, H.; Onodera, Y.; Komaki, C.; Mikami, T.; Kubo, T. A male sterility-associated mitochondrial protein in wild beets causes pollen distribution in transgenic plants. Plant J. 2008, 54, 1027–1036. [Google Scholar] [CrossRef]
  37. Kang, L.; Li, P.; Wang, A.; Ge, X.; Li, Z. A Novel Cytoplasmic Male Sterility in Brassica napus (inap CMS) with Carpelloid Stamens via Protoplast Fusion with Chinese Woad. Front. Plant Sci. 2017, 8, 529. [Google Scholar] [CrossRef]
  38. Case, A.L.; Willis, J.H. Hybrid male sterility in Mimulus (Phrymaceae) is associated with a geographically restricted mitochondrial rearrangement. Evolution 2008, 62, 1026–1039. [Google Scholar] [CrossRef]
  39. Makarenko, M.S.; Usatov, A.V.; Markin, N.V.; Azarin, K.V.; Gorbachenko, O.F.; Usatov, N.A. Comparative Genomics of Domesticated and Wild Sunflower: Complete Chloroplast and Mitochondrial Genomes. OnLine J. Biol. Sci. 2016, 16, 71–75. [Google Scholar] [CrossRef] [Green Version]
  40. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef]
  41. Nurk, S.; Bankevich, A.; Antipov, D.; Gurevich, A.; Korobeynikov, A.; Lapidus, A.; Prjibelsky, A.; Pyshkin, A.; Sirotkin, A.; Sirotkin, Y. Assembling Genomes and Mini-metagenomes from Highly Chimeric Reads. In Research in Computational Molecular Biology; Deng, M., Jiang, R., Sun, F., Zhang, X., Eds.; RECOMB 2013; Lecture Notes in Computer Science; Springer: Berlin/Heidelberg, Germany, 2013; Volume 7821, pp. 158–170. [Google Scholar] [CrossRef]
  42. Langmead, B.; Salzberg, S. Fast gapped-read alignment with Bowtie 2. Nat. Methods 2012, 9, 357–359. [Google Scholar] [CrossRef] [Green Version]
  43. Li, H. A statistical framework for SNP calling, mutation discovery, association mapping and population genetical parameter estimation from sequencing data. Bioinformatics 2011, 27, 2987–2993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Thorvaldsdóttir, H.; Robinson, J.T.; Mesirov, J.P. Integrative Genomics Viewer (IGV): High-performance genomics data visualization and exploration. Brief. Bioinform. 2013, 14, 178–192. [Google Scholar] [CrossRef] [PubMed]
  45. Alverson, A.J.; Wei, X.; Rice, D.W.; Stern, D.B.; Barry, K.; Palmer, J.D. Insights into the Evolution of Mitochondrial Genome Size from Complete Sequences of Citrullus lanatus and Cucurbita pepo (Cucurbitaceae). Mol. Biol. Evol. 2010, 27, 1436–1448. [Google Scholar] [CrossRef] [PubMed]
  46. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  47. Tillich, M.; Lehwark, P.; Pellizzer, T.; Ulbricht-Jones, E.S.; Fischer, A.; Bock, R.; Greiner, S. GeSeq—Versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 2017, 45, W6–W11. [Google Scholar] [CrossRef] [PubMed]
  48. Greiner, S.; Lehwark, P.; Bock, R. OrganellarGenomeDRAW (OGDRAW) version 1.3.1: Expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Res. 2019, 47, W59–W64. [Google Scholar] [CrossRef]
Figure 1. Mitochondrial genome map of HA89(ANN2) cytoplasmic male sterility (CMS) line of sunflower. Intron containing genes are marked by an asterisk (*) symbol. Trans-spliced genes are presented as the compilation of exons (ex).
Figure 1. Mitochondrial genome map of HA89(ANN2) cytoplasmic male sterility (CMS) line of sunflower. Intron containing genes are marked by an asterisk (*) symbol. Trans-spliced genes are presented as the compilation of exons (ex).
Plants 08 00439 g001
Figure 2. Schematic illustration of homologous regions between mitochondrial genomes of HA89 male-fertile and HA89(ANN2) CMS lines. 1—29,196 bp; 2—557 bp; 3—1245 bp; 4—77,441 bp; 5—41,702 bp; 6—4150 bp; 7—8584 bp; 8—21,433 bp 9—8158 bp; 10—24,687 bp; 11—41,505 bp; 12—6029 bp; 13—12,520 bp; 14—977 bp.
Figure 2. Schematic illustration of homologous regions between mitochondrial genomes of HA89 male-fertile and HA89(ANN2) CMS lines. 1—29,196 bp; 2—557 bp; 3—1245 bp; 4—77,441 bp; 5—41,702 bp; 6—4150 bp; 7—8584 bp; 8—21,433 bp 9—8158 bp; 10—24,687 bp; 11—41,505 bp; 12—6029 bp; 13—12,520 bp; 14—977 bp.
Plants 08 00439 g002
Figure 3. Comparison of transmembrane domains of proteins encoded by cox2 (A) and orf558 (B).
Figure 3. Comparison of transmembrane domains of proteins encoded by cox2 (A) and orf558 (B).
Plants 08 00439 g003
Figure 4. Comparison of orf1197 and atp6 encoded proteins and prediction of transmembrane helices. The amino acid sequence of the orf1197 encoded protein is presented. Amino acids identical to ATP6 are shown in bold. Amino acids forming transmembrane domains are marked by red bars.
Figure 4. Comparison of orf1197 and atp6 encoded proteins and prediction of transmembrane helices. The amino acid sequence of the orf1197 encoded protein is presented. Amino acids identical to ATP6 are shown in bold. Amino acids forming transmembrane domains are marked by red bars.
Plants 08 00439 g004
Figure 5. Reorganizations in CMS lines involving the 265 bp repeats (red stars) shown in the male-fertile mtDNA of HA89.
Figure 5. Reorganizations in CMS lines involving the 265 bp repeats (red stars) shown in the male-fertile mtDNA of HA89.
Plants 08 00439 g005
Table 1. Alignment of the mitochondrial genomes of HA89 and HA89(ANN2) lines.
Table 1. Alignment of the mitochondrial genomes of HA89 and HA89(ANN2) lines.
№ of Alignment RegionThe Alignment Region Length, bpPositions in mtDNA of Male-Fertile HA89Positions in mtDNA of Male-Sterile HA89(ANN2)OrientationSimilarity %Localized Genes
129,1961–29,1961–29,204Plus/Plus99nad2_ex3,4, trnY, trnN, trnC, ccmC, trnT, atp4, nad4L
255733,772–34,32878,343–78,899Plus/Plus99-
3124534,329–35,573148,163–149,411Plus/Minus98-
477,44136,739–114,179217,575–295,553Plus/Plus95atp8, cox3, trnV, rpl5, nad4, trnD, trnK, ccmB, rpl10, trnM, trnG, trnQ, trnH, trnE, nad1_ex1, cox1, nad5_ex3,4, rps11
541,702114,180–155,88235,657–77,315Plus/Minus99atp9, trnM, rps4, rrn26, rrn5, rrn18, rps13, nad1_ex2,3
64150155,883–160,032300,892–305,041Plus/Plus99-
78584160,320–168,903129,358–137,946Plus/Plus99nad2_ex1,2, nad6 *
821,433168,906–190,275171,388–192,871Plus/Minus98nad6 *, trnP, trnF, trnS, trnM, mttB, cob
98158194,543–202,700163,232–171,387Plus/Plus99ccmFC, orf873
1024,687202,701–227,387192,915–217,574Plus/Minus99atp1, ccmFN, nad7
1141,505227,396–268,90087,945–129,446Plus/Plus99nad1_ex4, rps3, rpl16, trnM, matR, nad3, rps12, nad9, trnW, nad5_ex1,2
126029269,217–275,245141,704–147,732Plus/Minus99atp6 *
1312,520275,536–288,055150,723–163,231Plus/Plus99cox2
14977299,971–300,947305,042–306,018Plus/Plus99trnK
* genes, which had impaired sequences as a result of rearrangements.
Table 2. Deletions (>100 bp) localized in the mitochondrial genome of HA89(ANN2) CMS line.
Table 2. Deletions (>100 bp) localized in the mitochondrial genome of HA89(ANN2) CMS line.
Deletion Length, bpPositions in mtDNA of the Male-Fertile Line HA89Deletion Localization according to the Male-Fertile Line HA89 Genetic MapDeleted Genes
287160,032–160,319rps13-nad6-
290275,246–275,535atp6-cox2-
29956,701–56,999nad4-ccmB-
316268,901–269,216nad9-atp6atp6 (partial)
58370,338–70,920rpl10-nad1-
116535,574–36,738nad4L-orf777-atp8orf777
4204190,339–194,542cob-ccmFC-
457529,197–33,771nad4L-orf777-
11,901288,070–299,970cox2-nad2-
Table 3. Localization of insertions (>100 bp) in the mitochondrial genome of HA89(ANN2) CMS line.
Table 3. Localization of insertions (>100 bp) in the mitochondrial genome of HA89(ANN2) CMS line.
Insertion in bpPositions in mtDNA of HA89(ANN2)New ORFs Based on InsertionHomology to
430147,733–148,162orf1197 *atp6
102777,316–78,342orf324orf285 (CMS PET2)
1310149,412–150,722orf345
3757137,947–141,703
5338295,554–300,891orf558cox2
645229,205–35,656
904578,900–87,944orf327, orf933
* appeared as the result of several simultaneous reorganizations of mtDNA structure.
Table 4. Summary of transcriptionally active open reading frames in the mitochondrial genome of isonuclear fertile and CMS lines.
Table 4. Summary of transcriptionally active open reading frames in the mitochondrial genome of isonuclear fertile and CMS lines.
Fertile mtDNACMS ANN2CMS MAX1CMS PET2CMS PET1
---orf228-
---orf285-
--orf306-orf306
-orf324---
-orf327---
-orf345---
--orf480--
----orfH522
-orf558---
--orf645orf645-
orf777---orf777
-orf891---
-orf933---
-orf1197---
--orf1287--
---orf2565-
In bold: ORFs encoding proteins with transmembrane domains.

Share and Cite

MDPI and ACS Style

Makarenko, M.S.; Usatov, A.V.; Tatarinova, T.V.; Azarin, K.V.; Logacheva, M.D.; Gavrilova, V.A.; Kornienko, I.V.; Horn, R. Organization Features of the Mitochondrial Genome of Sunflower (Helianthus annuus L.) with ANN2-Type Male-Sterile Cytoplasm. Plants 2019, 8, 439. https://doi.org/10.3390/plants8110439

AMA Style

Makarenko MS, Usatov AV, Tatarinova TV, Azarin KV, Logacheva MD, Gavrilova VA, Kornienko IV, Horn R. Organization Features of the Mitochondrial Genome of Sunflower (Helianthus annuus L.) with ANN2-Type Male-Sterile Cytoplasm. Plants. 2019; 8(11):439. https://doi.org/10.3390/plants8110439

Chicago/Turabian Style

Makarenko, Maksim S., Alexander V. Usatov, Tatiana V. Tatarinova, Kirill V. Azarin, Maria D. Logacheva, Vera A. Gavrilova, Igor V. Kornienko, and Renate Horn. 2019. "Organization Features of the Mitochondrial Genome of Sunflower (Helianthus annuus L.) with ANN2-Type Male-Sterile Cytoplasm" Plants 8, no. 11: 439. https://doi.org/10.3390/plants8110439

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