Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

The endogenous transposable element Tgm9 is suitable for generating knockout mutants for functional analyses of soybean genes and genetic improvement in soybean

  • Devinder Sandhu ,

    Roles Conceptualization, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Writing – original draft, Writing – review & editing

    devinder.sandhu@ars.usda.gov (DS); mbhattac@iastate.edu (MKB)

    Affiliation USDA-ARS, US Salinity Laboratory, Riverside, CA, United States of America

  • Jayadri Ghosh,

    Roles Investigation

    Affiliation Department of Agronomy, Iowa State University, Ames, IA, United States of America

  • Callie Johnson,

    Roles Investigation

    Affiliation Department of Biology, University of Wisconsin-Stevens Point, Stevens Point, WI, United States of America

  • Jordan Baumbach,

    Roles Investigation, Writing – review & editing

    Affiliation Department of Agronomy, Iowa State University, Ames, IA, United States of America

  • Eric Baumert,

    Roles Investigation

    Affiliation Department of Biology, University of Wisconsin-Stevens Point, Stevens Point, WI, United States of America

  • Tyler Cina,

    Roles Investigation

    Affiliation Department of Biology, University of Wisconsin-Stevens Point, Stevens Point, WI, United States of America

  • David Grant,

    Roles Data curation, Formal analysis, Investigation, Methodology, Resources, Visualization, Writing – review & editing

    Affiliations Department of Agronomy, Iowa State University, Ames, IA, United States of America, USDA-ARS Corn Insects and Crop Genomics Research Unit, Ames, IA, United States of America

  • Reid G. Palmer †,

    † Deceased.

    Roles Investigation, Resources

    Affiliations Department of Agronomy, Iowa State University, Ames, IA, United States of America, USDA-ARS Corn Insects and Crop Genomics Research Unit, Ames, IA, United States of America

  • Madan K. Bhattacharyya

    Roles Conceptualization, Funding acquisition, Investigation, Methodology, Project administration, Writing – review & editing

    devinder.sandhu@ars.usda.gov (DS); mbhattac@iastate.edu (MKB)

    Affiliation Department of Agronomy, Iowa State University, Ames, IA, United States of America

Abstract

In soybean, variegated flowers can be caused by somatic excision of the CACTA-type transposable element Tgm9 from Intron 2 of the DFR2 gene encoding dihydroflavonol-4-reductase of the anthocyanin pigment biosynthetic pathway. DFR2 was mapped to the W4 locus, where the allele containing Tgm9 was termed w4-m. In this study we have demonstrated that previously identified morphological mutants (three chlorophyll deficient mutants, one male sterile-female fertile mutant, and three partial female sterile mutants) were caused by insertion of Tgm9 following its excision from DFR2. Analyses of Tgm9 insertion sites among 105 independent mutants demonstrated that Tgm9 hops to all 20 soybean chromosomes from its original location on Chromosome 17. Some genomic regions are prone to increased Tgm9-insertions. Tgm9 transposed over 25% of the time into exon or intron sequences. Tgm9 is therefore suitable for generating an indexed insertional mutant collection for functional analyses of most soybean genes. Furthermore, desirable Tgm9-induced stable knockout mutants can be utilized in generating improved traits for commercial soybean cultivars.

1. Introduction

The soybean genome has been sequenced and expression patterns of most soybean genes are known [1, 2]. Although putative genes have been predicted based on the DNA sequence and when possible annotated for inferred function based on protein homology, many soybean genes remain uncharacterized even at this level of annotation. Rapid identification of biological functions of soybean genes will require an indexed insertional mutant collection suitable for reverse genetics studies.

Gene silencing through RNAi has been successfully used in the functional characterization of plant genes [3]. However, the incomplete inactivation of target genes is a common problem with such gene silencing approaches and makes the data interpretation difficult [4]. T-DNA insertion mutagenesis, in which insertion of T-DNA into the coding or promoter sequence of a gene can disrupt its function, has been effectively utilized in Arabidopsis [57]. Gene editing has been shown to be a powerful approach for the functional analyses of genes in plant species including maize [8, 9]. In soybean, the main bottleneck associated with the application of RNAi, T-DNA insertion mutagenesis or gene editing approaches for functional analyses of a large number of genes is the lack of a high-throughput transformation procedure and availability of large greenhouse spaces for growing transformants.

Transposable elements are a major component of the genomes of higher eukaryotes and are widely distributed among plant species [10]. Functional characterization of thousands of soybean genes could be facilitated by knockout mutants induced by transposons [1114]. Active endogenous transposable elements have been identified in several plant species and have been effectively used in functional characterization of plant genes [15, 16].

The utility of active transposons from maize (Ac/Ds), rice (mPing) and tobacco (Tnt1) in tagging soybean genes has been demonstrated [1113]. However, this approach requires genetic transformation. Handling thousands of transgenic lines in the field prior to deregulation is impractical and cumbersome. Furthermore, products generated through genetic transformation are often not well received by a significant proportion of the end users worldwide and the deregulation step could be lengthy and expensive. Tagging of soybean genes using an endogenous transposable element is therefore an attractive solution not only for functional analyses of tens of thousands of soybean genes, but also in generating desirable mutants for rapid genetic improvement of important traits in the commercial soybean cultivars. Since mutants created by an endogenous transposon are not GMOs, the products produced from incorporation of such mutants should be accepted by all consumers.

Five genes, W1, W3, W4, Wm, and Wp, regulate pigmentation in flowers and hypocotyls of soybean [17]. The mutant w4-m allele at the W4 locus is characterized by altered pigment accumulation patterns in flowers and hypocotyls [18]. The soybean line with w4-m allele was added to soybean collection and assigned the genetic type collection number T322 [19]. W4 contains the DFR2 gene encoding a functional dihydroflavonol-4-reductase 2 enzyme. Somatic excision of a 20,548-bp CACTA-type transposable element, Tgm9, from Intron 2 of DFR2 causes variegated flowers and hypocotyls with dark brown sectors [14]. Since Tgm9 is located in an intron, both precise and imprecise excisions of the element lead to restoration of the wild-type phenotype. Excision of the element from the germ tissues of the parent plant results in at least some progenies that carry only purple flowers and are termed germinal revertants. The mutation rates in several genetic loci among the revertants with purple flowers were much higher than the rate of spontaneous mutation [20]. Therefore, it was hypothesized that following excision of Tgm9 from DFR2, the element inserts into new genetic loci.

Tgm9 excises at a high rate with a germinal reversion frequency of about 6% per generation [21]. Several mutant genes, unlinked to the W4 locus, were isolated by screening germinal revertants for morphological mutant phenotypes [20, 22, 23]. For example, male-sterility, female-sterility and root necrotic root mutants were identified by screening thousands of germinal revertants carrying purple flowers [24, 25]. The present study was undertaken to determine if Tgm9 is suitable for creating knockout mutants for conducting large-scale functional analyses in soybean.

2. Materials and methods

2.1. Genetic material

Seeds of chlorophyll deficient mutants (Genetic type collection numbers T323, T325 and T346), male-sterile, female-fertile mutant (T359) and partial female sterile mutants (T364, T365 and T367) were obtained from Dr. Randall Nelson, USDA-Agricultural Research Services.

For generating germinal revertants for this study, we identified T322 (w4-m) plants that showed variegated flowers and their progenies were grown to identify germinal revertants. One germinal revertant per progeny row was selected for this study (S1 Fig). Leaf tissues from each of the selected germinal revertants was harvested for DNA preparation according to a previously described method [26].

2.2. Transposon display

A GenomeWalker Universal kit (Clontech Laboratories, Inc., Mountain View, CA, USA) was used to find the unknown genomic DNA sequences adjacent to a transposable element using the manufacturer’s instructions. DNA (2.5 μg) was independently digested with four restriction enzymes (DraI, EcoRV, PvuII, and StuI) to generate blunt ended fragments (S2 Fig). After phenol:chloroform::1:1 purification, digested genomic DNA was ligated to the GenomeWalker adaptor to generate genomic DNA libraries. The four genomic libraries, generated through digestion of DNA with four restriction endonucleases for each germinal revertant were used for the first PCR (PCR 1) using the outer adaptor primer (AP1) and an outer Tgm9-specific primer (Trans R1) (S2 Fig; S1 Table). The PCR 1 mixture was then diluted to 100 times and used as a template for a second or “nested” PCR (PCR 2). PCR 2 reactions were conducted using the nested adaptor primer (AP2) and a nested Tgm9-specific primer (Trans R2) (S2 Fig). For visualization, the resulting PCR products were separated on a 1.5% agarose gel at 100 V for 1 hour. As there are some residual/truncated copies of Tgm9 in T322, only unique bands of individual mutants were sequenced (S2 Fig). Residual/truncated copies were displayed as common bands among the selected mutants (S3 Fig). Unique PCR products were sequenced and evaluated for presence of over 100 bp Tgm9-end specific sequences.

2.3. Sequencing of the transposon inserted into the soybean male-sterile, female-sterile GmMER3 gene

Long range PCR was performed to amplify a region spanning 2,523 base pairs of the 5’-end of the transposon in two steps using two nested primers from Tgm9 and two from the mutant MER3 gene of mer3 (S1 Table). The PCR products were amplified using Phusion high fidelity DNA polymerase (Thermo Fisher, USA). Bands of correct size were extracted from a 0.8% agarose gel using IBI gel extraction kit (IBI Scientific, USA). The samples were sequenced using the inner Tgm9 and inner MER3 primers at the DNA Facility at Iowa State University. The resulting sequence was aligned with the Tgm9 (GQ344503.1) and Tgmt* (EU190440.1) sequences.

2.4. Identification of Tgm9-insertion sites in the soybean genome

Tgm9-insertion mutants were named using an initial 'T9' with a nomenclature similar to that used for other mutant types in SoyBase. The location of each Tgm9 insertion was identified by using BLAST to find the location of the element's flanking sequence. The location was determined in both the Wm82.a1 and Wm82.a2 soybean genome assemblies (http://soybase.org/aboutgenomenomenclature.php). The proportion of exons and introns in the soybean genome sequences that were mapped to chromosomes were calculated by considering the predicted exonic and intronic sequences in the Wm82.a2 genome assembly (1).

2.5. Calculation of exon and intron sequences of the soybean genome

The data in the GFF file used at SoyBase for the Wm82.a2 genome assembly was parsed into the component parts of the genes. For each gene we used the longest splice variant and determined (i) the full length of the gene, i.e. the region transcribed; (ii) the length of the CDS, i.e. the translated mRNA from ATG to TAG with introns removed; (iii) the length of the 5’ UTR (un-translated region) in the translated mRNA, i.e. with any introns removed; (iv) the length of the 3’ UTR in the translated mRNA, i.e. with any introns removed; (v) the combined length of all introns, i.e. those sequences spliced out of the 5’ UTR and/or the translated region and/or the 3’ UTR. These values were summed for all gene models in the Wm82.a2 genome assembly. We used the sum of the 5’ UTR, CDS and 3’ UTR lengths as the total exon length in the genome. The proportions of exon (11.6%) and intron (10.5%) sequences (S2 Table) are used as the expected values of random Tgm9 insertion into the soybean genome.

2.6. Statistical analysis

We conducted the χ2 analysis to determine if the observed insertion of Tgm9 into exons and introns was random. χ2 values were calculated for both classes of insertions; exon and intron-specific insertions with a degree of freedom of 1.

2.7. Deposition of data and availability of supporting data

Most of the data are presented in the manuscript or as supplemental data files. The data for the Tgm9 insertion sites have been deposited to SoyBase and are available at http://soybase.org/gb2/gbrowse/gmax2.0x/?start=1;stop=56831624;ref=Gm01;width=1024;version=100;flip=0;grid=1;id=28f86f22137273c155fe3e1756259be0;l=tgm9-gene_models_wm82_a2_v1-pericentromere%3Aoverview.

3. Results

We investigated the applicability of Tgm9 in generating knockout mutants for functional analyses of soybean genes as follows. First, we determined if the insertion following excision of Tgm9 from the mutable w4-m allele resulted in mutation among the mutants previously identified from screening germinal revertants, generated from the T322 line [20, 23, 2731] (Table 1). Second, we investigated 124 random germinal revertants to determine the properties of Tgm9 transposition.

thumbnail
Table 1. Genetically mapped soybean mutants identified by screening germinal revertants of the mutable line, T322 (w4-m).

https://doi.org/10.1371/journal.pone.0180732.t001

3.1. Tgm9 induces mutations following excision from DFR2

Several morphological mutants were previously identified from screening thousands of germinal revertants generated from the T322 (w4-m) mutable line. These mutants include chlorophyll deficient mutants (T323, T325 and T346) [23, 27], male-sterile, female-fertile mutant (T359) [28] and partial female sterile (fsp) mutants (T364, T365 and T367) [20] (Table 1). The genes responsible for the mutant phenotypes in these mutants have been previously genetically mapped [2931]. To determine if the mutations in these mutants were caused by Tgm9-insertion, we identified the Tgm9-insertion sites using a transposon display approach for each of the mutants (S2 Fig).

The Tgm9-insertion sites were mapped to single genes in each of the five mutants investigated (Tables 1 and 2; Fig 1). The five Tgm9-induced mutant genes identified in these five mutants are from previously mapped mutant loci or regions. For example, in three chlorophyll deficient mutants (y20 mdh1-n) a Tgm9 insertion was identified in the first intron of Glyma.12G159300 encoding lactate/malate dehydrogenase (Table 2). The y20 Mdh1-n locus was originally mapped to Chromosome 12 and shown to be flanked by microsatellite markers Satt253 and Satt302 [29] (Table 1). Glyma.12G159300 is located in this y20 Mdh1-n genomic region. Similarly, the Tgm9-insertion site in the ms9 mutant was located in the first intron of Glyma.03G152300 that has no functional annotation (Table 2). The gene is in the ms9 region flanked by Satt521 and Satt237 (https://soybase.org) [30, 32] (Tables 1 and 2; Fig 1). The Tgm9-insertion sites in Fsp2 and Fsp3 were detected in the 7th intron and 2nd exon of Glyma.06G174200 and Glyma.08G359000, respectively (Table 2). Glyma.06G174200 encodes a haloacid dehalogenase-like hydrolase; Glyma.08G359000 encodes the embryo-specific protein 3, (ATS3) (Table 2) (https://soybase.org). The Fsp2 locus is flanked by Satt170 and Satt277 on Chromosome 6 [31]. Glyma.06G174200 is located in this Fsp2 region (Tables 1 and 2; Fig 1). Likewise, the Tgm9 insertion sites in the fsp3 and fsp5 mutants were also localized to the genomic regions to which Fsp3 and Fsp5 were genetically mapped (Fig 1; Table 2).

thumbnail
Fig 1. Physical map of five candidate genes that were mutated in Tgm9-insertion mutants.

Flanking molecular markers from the genetic linkage maps for y20 Mdh1-n, ms9, Fsp2, Fsp3 and Fsp5 were placed on the physical map along with Tgm9-insertion sites identified in these mutants [2931]. Scale is represented in million base pairs (Mb) of DNA.

https://doi.org/10.1371/journal.pone.0180732.g001

thumbnail
Table 2. Forward genetics approach showing identification of the Tgm9 insertion sites in known soybean mutants identified from the progenies of mutable line T322 (w4-m).

https://doi.org/10.1371/journal.pone.0180732.t002

3.2. Tgm9 transposes to all 20 soybean chromosomes

To better understand the transposition patterns of Tgm9, we identified 124 germinal revertants bearing only purple flowers, each from individual T322 (w4-m) mutable plants (S1 Fig). Therefore, each revertant was unique. We were able to determine insertion sites from 105 mutants (S3 Table). Whether Tgm9 failed to insert into new genetic loci or we failed to recover the insertion sites among the remaining 19 mutants is unknown.

Physical mapping of the Tgm9 insertion sites among 105 mutants revealed that the element transposes to all 20 chromosomes (Fig 2). Insertion sites per chromosome ranged from 1 to 16 per chromosome (Fig 2). These vast differences in the distribution of insertions among chromosomes could be due to the small sample size of the population studied or due to a preference of the element to transpose into some genomic regions. There were six genomic regions, <1 Mb in size, which were enriched in Tgm9 insertions. Of the 105 identified mutants, 24 were mapped to these six regions: (i) Chromosome 1 (six Tgm9 insertions between 50.06 and 50.50 Mb region (http://www.soybase.org/SequenceIntro.php); (ii) Chromosome 9 (four between 4.07 and 5.07 Mb); (iii) Chromosome 9 (five between 8.77 and 9.06 Mb), (iv) Chromosome 15 (three between 2.58 and 2.87 Mb); (v) Chromosome 17 (three between 40.65 and 40.89 Mb); and (vi) Chromosome 18 (three between 54.31 and 54.51 Mb) (Fig 2). In soybean the euchromatic region constitutes 43% of the genome [1]. In this study, Tgm9 transposed to euchromatic regions 77.1% of the time. The remaining 22.9% of insertions were mapped to the pericentromeric regions (Fig 2).

thumbnail
Fig 2. Physical map of the Tgm9 insertion sites among 105 mutants.

Green arrows represent locations of Tgm9 in genes (exons and introns) and red arrows show Tgm9 insertions in other genomic regions. A purple arrow shows the location of Tgm9 in W4 on Chromosome 17. Centromeres (black rectangles) and heterochromatic regions (grey areas) are shown on individual chromosomes. Scale is represented in million base pairs (Mb) of DNA.

https://doi.org/10.1371/journal.pone.0180732.g002

3.3. Organization of Tgm9-induced mutations in the soybean genome

To study the distribution of transposition events of Tgm9 across the chromosomes, the physical locations of the Tgm9 insertion sites in the soybean genome were investigated. Although Tgm9-insertion in promoters can alter phenotypes, due to the limitation in precisely predicting the promoter sequences we calculated the number of insertions that were localized to only exon- and intron-sequences. These mutations are expected to knock-out gene function, and therefore are desired mutations for functional analyses of soybean genes. Of the 105 insertions studied, 16.2% and 9.5% of the Tgm9-insertions were localized to exons and introns, respectively (Fig 3, S3 Table). Thus, 25.7% of the insertions were generated in gene-sequences. In the soybean genome, gene-sequences cover 22.1% of the genome (S3 Table), which is very comparable to the observed insertion rate of 25.7% in the gene-sequences. The calculated χ2 value (1.92 at df = 1) suggests that the observed frequencies of Tgm9 insertions in exons and introns (Fig 3) is statistically not significant at p = 0.05 from the expected values resulting from random insertion of the element in exon and intron sequences. We therefore conclude based on the study of 105 mutants that the Tgm9 element appears to transpose randomly in the soybean genome (Fig 3, S3 Table).

thumbnail
Fig 3. Percent distribution of the type of Tgm9-induced mutants.

The observed frequency of Tgm9 insertions in exonic and intronic sequences were calculated from 105 Tgm9 insertion loci (S3 Table). The expected frequency of Tgm9 is based on random insertion of the element in the exonic and intronic sequences calculated from the soybean genome sequences mapped to chromosomes (Soybase). The χ2 value (1.7752 at df = 1) calculated for observed Tgm9 insertion frequencies and expected frequencies of random Tgm9 insertions in exon and intron sequences is statistically non-significant at p = 0.05.

https://doi.org/10.1371/journal.pone.0180732.g003

4. Discussion

The availability of the genome sequence has greatly expedited molecular and genomic research in soybean [1]. Reverse genetic approaches such as TILLING and deletion mutants induced by fast neutron irradiation are being exploited for functional characterization of the soybean genome [33, 34]. Heterologous transposable elements have also been successfully applied in functional characterization of soybean genes [1113].

The Ac/Ds transposon system in maize causes insertion mutations in closely linked locations. It has been applied to target specific chromosomal regions in several plant species [35, 36]. In soybean, the feasibility of Ac/Ds was tested by sequencing flanking regions of 200 individual mutants and was validated by isolating a gene involved in male-fertility [13]. Similarly, mPing, a miniature inverted repeat transposable element (MITE) from rice was evaluated in soybean [12]. mPing was shown to transpose to unlinked regions with a strong preference for gene-containing regions [12]. Tnt1, a tobacco retrotransposon, was effectively transformed into soybean and was shown to be re-activated by tissue culture [11].

Transposon-induced mutants are preferred over mutants generated using irradiation or chemical mutagenesis approaches because they typically contain a very few mutations per mutant and therefore are easier to analyze and utilize in breeding programs. However, heterologous transposable elements have to be genetically transformed and at least for the maize Ac/Ds system requires the generation of a large collection of transgenic soybean lines because the element moves only to linked genomic regions [37]. The requirement of tissue culture for the activation of Tnt1 limits its uses in functional analyses of soybean genes because tissue culture itself can generate new genetic and epigenetic mutations, broadly known as somaclonal variation. Furthermore, functional analyses of tens of thousands of soybean genes will require growing a huge number of transgenic soybean lines in the field with proper care not to release transgenes to the environment. Once a product is developed from such a huge effort, the seed industry must then wait to get approval for deregulating the transgenic soybean lines. Then, there is also the issue of non-acceptance of transgenic soybeans by a significant proportion of soybean consumers. Given the complexities associated with the use of heterologous transposons summarized above, endogenous transposable elements are better suited for functional analyses of soybean genes and the generation of useful mutants for breeding desirable cultivars.

Several endogenous transposable elements such as Tgm1, Tgm2, Tgm3, Tgm4, Tgm5, Tgm6, Tgm7, Tgm8, Tgm9, Tgmt* and TgmR* have been identified in soybean [14, 3840]. Of these, Tgm9 has been shown to be active and has recently been used to clone a male-sterile, female-sterile gene (GmMER3) that encodes an ATP dependent DNA helicase [14, 41, 42]. It is possible, although unlikely, to have another Tgm sequence(s) causing mutation in GmMER3 or other genes investigated in this study. To determine if the mutation in the male-sterile, female-sterile (mer3) mutant was caused by Tgm9, and not by any other active elements such as Tgmt*, the 3’-end 2,523 bp region of the transposon in the GmMER3 gene of the mer3 mutant was amplified and sequenced. The sequence matched perfectly with the Tgm9 sequence suggesting that the mer3 mutation was caused by Tgm9 insertion, not by Tgmt* (S4 Fig). Furthermore, sequencing of the T322 genome revealed that Tgm9 is the only intact CACTA-type element identified in this cultivar. In addition to the expected location of Tgm9 in W4, an additional copy of the element was identified in a locus on Chromosome 19. This Tgm9 element was hemizygous in T322 and absent among 15 selected Tgm9-induced mutants. This suggests that most likely the Tgm9 copy on Chromosome 19 was generated from a recent transposition event (J. Baumbach, S. Srivastava and M.K. Bhattacharya, unpublished).

Here we have demonstrated that following excision from dfr2, Tgm9 induces mutations in unlinked soybean genes (Fig 1). We have also demonstrated from analysis of 105 independent Tgm9-induced mutants that the element transposes to unlinked loci on all 20 soybean chromosomes (Fig 2). Since Tgm9 is the only intact CACTA-type element in T322, the mutant genes or sequences identified in this study were resulted from insertion of Tgm9. To our knowledge, Tgm9 is only reported active transposable element in soybean [14].

In this study we have observed that Tgm9 transposes into unlinked loci following excision from the w4-m mutant allele and that such transpositions could cause functional mutations in over 25% of the mutants. There are 15,166 single copy genes in the soybean genome and Tgm9 should be a suitable tool for generating mutations in most of these genes [1]. However, Tgm9 has preference for certain genomic regions (Fig 2). This will reduce the efficiency of Tgm9 in inducing mutations in such a study for functional analyses of soybean genes. Utilizing Tgm9 to advance soybean genetics and biology will therefore require a large indexed Tgm9-induced mutant population. These mutants can be easily identified by taking the advantage of next generation sequencing platforms [7, 43] and displayed in the SoyBase genome browser (Fig 4). The ability of Tgm9 to precisely excise from mutant loci for reconstituting the wild-type function may eliminate the need of complementation analyses of mutants through genetic transformation for identification of novel genes [42].

thumbnail
Fig 4. Display of Tgm9-induced mutants in the SoyBase genome browser.

The track for Tgm9-insertion lines shows the location of two Tgm9-induced insertions (T91200307 and T91200364) on Chromosome 15. Note that in the T91200307 mutant, the Tgm9 insertion site is located in the inter-genic region, and in T91200364, the insertion site is in the 2nd exon of Glyma.15G033300. (http://soybase.org/gb2/gbrowse/gmax2.0x/?start=1;stop=56831624;ref=Gm01;width=1024;version=100;flip=0;grid=1;id=28f86f22137273c155fe3e1756259be0;l=tgm9-gene_models_wm82_a2_v1-pericentromere%3Aoverview {Please use user name: Bhatt; password: tgm9#}).

https://doi.org/10.1371/journal.pone.0180732.g004

The 20,548-bp Tgm9 element has been reported to be fractured during transposition, leading to the generation of stable mutants which could be used in soybean breeding programs [14]. Furthermore, cultivars generated by using such Tgm9-induced mutants can reach consumers rapidly since they do not require the lengthy deregulation process of mutants generated from the use of heterologous transposon systems.

5. Conclusions

In this investigation we have shown that Tgm9 is an active transposable element that can induce mutations in unlinked genes on all 20 soybean chromosomes. It appears based on the study of 105 mutants that the Tgm9 element randomly transposes into genes and over 25% of the mutants are expected to be knockout mutants, suitable for functional analyses of soybean genes. As there is a single active copy of this element in the mutable T322 line, identification of the insertion sites in a large collection of Tgm9-induced mutants is feasible through application of a next-generation sequencing platform. In soybean, the functions of the majority of the genes are still unknown. The generation of an indexed transposon-induced mutant population using Tgm9 and their display in the SoyBase genome browser in the context of the other information available at SoyBase (Fig 4) will likely facilitate the functional characterization of most of the 15,166 single copy soybean genes [1]. Desirable Tgm9-induced mutant genes carrying inactive, fractured Tgm9-elements can also be identified and incorporated into elite breeding programs leading to relatively rapid release of genetically improved cultivars for commercial cultivation.

Supporting information

S1 Fig. Identification of a germinal revertant from a progeny row of a single mutable T322 (w4-m) plant.

Approximately 150 progeny of a mutable plant identified in an earlier experiment were grown in a 15-foot long plot to locate a germinal revertant with only purple flowers.

https://doi.org/10.1371/journal.pone.0180732.s001

(PPTX)

S2 Fig. Flow diagram of the transposon display procedure.

Adaptor has 5’ extended strand with no binding site for primers AP1 (Adaptor Primer 1) or AP2 (Adaptor Primer 2). Binding site for AP1 or AP2 can only be generated by transposon specific primers (TransR1 or TransR2). Exposed 3’ end of the adaptor is blocked by amino group to prevent extension. Unique bands from different lanes are excised and sequenced. The bands that are common to all lanes are most likely ancient transposition events.

https://doi.org/10.1371/journal.pone.0180732.s002

(PPTX)

S3 Fig. Transposon display of two selected progenies from each of six independent plants carrying variegated flowers.

A single progeny row was grown from each of the six independent mutable plants harvested in 2014. From each row, two plants were selected for transposon display: (i) one plant with only green stem; (ii) the other plant with only purple stem (germinal revertant). The PCR fingerprints of each of the six plants with only green stem are shown on lanes 1 through 6; and those for six plants with purple stem on lanes 7 through 12. Note that two plants, Plant # 1 under green stem heading and Plant # 1 under purple stem heading, were descended from the same mutable plant harvested in 2014. Two restriction endonucleases, EcoRV and PvuII, were used in digesting the genomic DNA for generating the transposon displays. White arrows show the amplification of some of the residual insertions; whereas, red arrows indicate the progeny-specific amplification presumably from new Tgm9 insertion sites in distinct loci. DFR2-specific amplification was observed for the plants with purple stems (germinal revertants). Note that in Plant # 5 with purple stem failed to amplify the Tgm9 insertion site at the DFR2 intron II presumably because of simultaneous Tgm9 excision from both DFR2 copies. The sibling plants with green stems failed to amplify DFR2 because of the presence of Tgm9 in both DFR2 copies. Sub-PCR of the two strong PCR amplified ~750 bp fragments in Plant # 3 and 6 with green stem using DFR2 F and TransR2 primers (Supplemental Table 1) indicated that the intense amplified PCR products were from the DFR2 locus (data are not shown).

https://doi.org/10.1371/journal.pone.0180732.s003

(PPTX)

S4 Fig. Tgm9 caused the insertion mutation in the GmMER3 gene of the mer3 mutant.

Transposon insertion sequence from the GmMER3 gene was amplified by conducting long-range PCR and compared to both Tgm9 and Tgmt*, the two highly similar transposons characterized from soybean. A) The orientation of the Tgm9 insertion in the MER3 gene and primers used for nested PCR and sequencing are shown. B) First PCR product was amplified using primers 1 and 2. The amplified PCR product was 2,823 bp. C) The nested PCR product was amplified using primers 3 and 4. The amplified PCR product was 2,758 bp. Primer 3 was used for sequencing the PCR product represented by the dashed line X. Primer 4 was used to sequence the PCR product and produced sequence represented by the dashed line Y. D) Sequence from primer 3 matched the MER3 gene and the start of the 5’ end of the Tgm9 transposon sequence shown with the red font. E) Sequence from primer 4 aligned to the Tgm9 and Tgmt*. The PCR product matches the Tgm9 sequence perfectly. Polymorphic nucleotides between the insertion sequence and Tgmt* sequence are shown in red font.

https://doi.org/10.1371/journal.pone.0180732.s004

(PPTX)

S1 Table. List of primers used in this investigation.

https://doi.org/10.1371/journal.pone.0180732.s005

(DOCX)

S2 Table. Exon and intron sequences of the soybean genome.

https://doi.org/10.1371/journal.pone.0180732.s006

(DOCX)

S3 Table. Tgm9 insertion sites in 105 independent mutants.

https://doi.org/10.1371/journal.pone.0180732.s007

(XLSX)

Acknowledgments

This project was supported by a grant from United Soybean Board and a grant from Undergraduate Research, Scholarship and Creative Activity grant, UWSP.

References

  1. 1. Schmutz J, Cannon SB, Schlueter J, Ma J, Mitros T, Nelson W, et al. Genome sequence of the palaeopolyploid soybean. Nature. 2010;463(7278):178–83. pmid:20075913.
  2. 2. Severin AJ, Woody JL, Bolon YT, Joseph B, Diers BW, Farmer AD, et al. RNA-Seq Atlas of Glycine max: a guide to the soybean transcriptome. BMC Plant Biol. 2010;10:160. pmid:20687943; PubMed Central PMCID: PMCPMC3017786.
  3. 3. Chuang CF, Meyerowitz EM. Specific and heritable genetic interference by double-stranded RNA in Arabidopsis thaliana. Proc Natl Acad Sci USA. 2000;97(9):4985–90. pmid:10781109; PubMed Central PMCID: PMCPMC18344.
  4. 4. Hofgen R, Axelsen KB, Kannangara CG, Schuttke I, Pohlenz HD, Willmitzer L, et al. A visible marker for antisense mRNA expression in plants: inhibition of chlorophyll synthesis with a glutamate-1-semialdehyde aminotransferase antisense gene. Proc Natl Acad Sci USA. 1994;91(5):1726–30. pmid:8127872; PubMed Central PMCID: PMCPMC43236.
  5. 5. McKinney EC, Ali N, Traut A, Feldmann KA, Belostotsky DA, McDowell JM, et al. Sequence-based identification of T-DNA insertion mutations in Arabidopsis: actin mutants act2-1 and act4-1. Plant J. 1995;8(4):613–22. pmid:7496405.
  6. 6. Mabuchi A, Soga K, Wakabayashi K, Hoson T. Phenotypic screening of Arabidopsis T-DNA insertion lines for cell wall mechanical properties revealed ANTHOCYANINLESS2, a cell wall-related gene. J Plant Physiol. 2015;191:29–35. pmid:26720211.
  7. 7. O'Malley RC, Ecker JR. Linking genotype to phenotype using the Arabidopsis unimutant collection. Plant J. 2010;61(6):928–40. pmid:20409268
  8. 8. Svitashev S, Young JK, Schwartz C, Gao HR, Falco SC, Cigan AM. Targeted mutagenesis, precise gene editing, and site-specific gene insertion in maize using Cas9 and guide RNA. Plant Physiol. 2015;169(2):931–45. pmid:26269544
  9. 9. Ran FA, Hsu PD, Wright J, Agarwala V, Scott DA, Zhang F. Genome engineering using the CRISPR-Cas9 system. Nat Protoc. 2013;8(11):2281–308. pmid:24157548
  10. 10. Levin HL, Moran JV. Dynamic interactions between transposable elements and their hosts. Nat Rev Genet. 2011;12(9):615–27. pmid:21850042; PubMed Central PMCID: PMCPMC3192332.
  11. 11. Cui Y, Barampuram S, Stacey MG, Hancock CN, Findley S, Mathieu M, et al. Tnt1 retrotransposon mutagenesis: a tool for soybean functional genomics. Plant Physiol. 2013;161(1):36–47. pmid:23124322; PubMed Central PMCID: PMCPMC3532266.
  12. 12. Hancock CN, Zhang F, Floyd K, Richardson AO, Lafayette P, Tucker D, et al. The rice miniature inverted repeat transposable element mPing is an effective insertional mutagen in soybean. Plant Physiol. 2011;157(2):552–62. pmid:21844309; PubMed Central PMCID: PMCPMC3192579.
  13. 13. Mathieu M, Winters EK, Kong F, Wan J, Wang S, Eckert H, et al. Establishment of a soybean (Glycine max Merr. L) transposon-based mutagenesis repository. Planta. 2009;229(2):279–89. pmid:18855007.
  14. 14. Xu M, Brar HK, Grosic S, Palmer RG, Bhattacharyya MK. Excision of an active CACTA-Like transposable element from DFR2 causes variegated flowers in soybean Glycine max (L.) Merr. Genetics. 2010;184(1):53–63. pmid:19897750
  15. 15. Brutnell TP. Transposon tagging in maize. Funct Integr Genomics. 2002;2(1–2):4–12. pmid:12021846.
  16. 16. Fedoroff NV, Furtek DB, Nelson OE. Cloning of the bronze locus in maize by a simple and generalizable procedure using the transposable controlling element Activator (Ac). Proc Natl Acad Sci USA. 1984;81(12):3825–9. pmid:16593478
  17. 17. Palmer RG, Pfeiffer TW, Buss GR, Kilen TC. Qualitative Genetics. In: Boerma HR, Specht J, editors. Soybeans: Improvement, production, and uses (Agronomy Monograph no 16). 3rd ed. Madison, WI: American Society of Agronomy, Inc.; 2004. p. 137–233.
  18. 18. Groose RW, Weigelt HD, Palmer RG. Somatic analysis of an unstable mutation for anthocyanin pigmentation in soybean. J Hered. 1988;79(4):263–7.
  19. 19. Palmer RG, Groose RW, Weigelt HD, Miller JE. Registration of a genetic stock (w4-m w4-m) for unstable anthocyanin pigmentation in soybean. Crop Sci. 1990;30(6):1376–7.
  20. 20. Palmer RG, Hedges BR, Benavente RS, Groose RW. w4-mutable line in soybean. Dev Genet. 1989;10(6):542–51.
  21. 21. Groose RW, Schulte SM, Palmer RG. Germinal reversion of an unstable mutation for anthocyanin pigmentation in soybean. Theor Appl Genet. 1990;79:161–7. pmid:24226213
  22. 22. Palmer RG, Zhang F, Horner HT. Genetic studies of a male-sterile, female-sterile mutant from a transposon-containing soybean population. Soyb Genet Newsl. 1992;19:112–9. pmid:CABI:19921633145.
  23. 23. Chen XF, Imsande J, Palmer RG. Eight new mutants at the k2 Mdh1-n y20 chromosomal region in soybean. J Hered. 1999;90(3):399–403.
  24. 24. Palmer RG, Sandhu D, Curran K, Bhattacharyya MK. Molecular mapping of 36 soybean male-sterile, female-sterile mutants. Theor Appl Genet. 2008;117(5):711–9. pmid:18592206
  25. 25. Palmer RG, Zhang L, Huang Z, Xu M. Allelism and molecular mapping of soybean necrotic root mutants. Genome. 2008;51(4):243–50. pmid:18356960
  26. 26. Sandhu D, Gao H, Cianzio S, Bhattacharyya MK. Deletion of a disease resistance nucleotide-binding-site leucine-rich-repeat-like sequence is associated with the loss of the Phytophthora resistance gene Rps4 in soybean. Genetics. 2004;168(4):2157–67. Epub 2004/12/22. 168/4/2157 [pii] pmid:15611183; PubMed Central PMCID: PMC1448712.
  27. 27. Hedges BR, Palmer RG. Inheritance of malate-dehydrogenase nulls in soybean. Biochem Genet. 1992;30(9–10):491–502. pmid:1445188
  28. 28. Palmer RG. Genetics of four male-sterile, female-fertile soybean mutants. Crop Sci. 2000;40(1):78–83.
  29. 29. Xu M, Palmer RG. Molecular mapping of k2 Mdh1-n y20, an unstable chromosomal region in soybean [Glycine max (L.) Merr.]. Theor Appl Genet. 2005;111(8):1457–65. pmid:16088398.
  30. 30. Cervantes-Martinez I, Xu M, Zhang L, Huang Z, Kato KK, Horner HT, et al. Molecular mapping of male-sterility loci ms2 and ms9 in soybean. Crop Sci. 2007;47(1):374–9. ISI:000244430100047.
  31. 31. Kato KK, Palmer RG. Molecular mapping of four ovule lethal mutants in soybean. Theor Appl Genet. 2004;108(4):577–85. Epub 2003/11/12. pmid:14610643.
  32. 32. Grant D, Nelson RT, Cannon SB, Shoemaker RC. SoyBase, the USDA-ARS soybean genetics and genomics database. Nucleic Acids Res. 2010;38(Database issue):D843–6. Epub 2009/12/17. pmid:20008513; PubMed Central PMCID: PMCPMC2808871.
  33. 33. Cooper JL, Till BJ, Laport RG, Darlow MC, Kleffner JM, Jamai A, et al. TILLING to detect induced mutations in soybean. BMC Plant Biol. 2008;8:9. pmid:18218134; PubMed Central PMCID: PMCPMC2266751.
  34. 34. Bolon YT, Haun WJ, Xu WW, Grant D, Stacey MG, Nelson RT, et al. Phenotypic and genomic analyses of a fast neutron mutant population resource in soybean. Plant Physiol. 2011;156(1):240–53. pmid:21321255
  35. 35. Jones DA, Thomas CM, Hammond-Kosack KE, Balint-Kurti PJ, Jones JD. Isolation of the tomato Cf-9 gene for resistance to Cladosporium fulvum by transposon tagging. Science. 1994;266(5186):789–93. pmid:7973631.
  36. 36. James DW Jr., Lim E, Keller J, Plooy I, Ralston E, Dooner HK. Directed tagging of the Arabidopsis FATTY ACID ELONGATION1 (FAE1) gene with the maize transposon activator. Plant Cell. 1995;7(3):309–19. pmid:7734965; PubMed Central PMCID: PMCPMC160784.
  37. 37. Dooner HK, Belachew A. Transposition pattern of the maize element Ac from the bz-m2(Ac) allele. Genetics. 1989;122(2):447–57. pmid:17246501
  38. 38. Rhodes PR, Vodkin LO. Organization of the Tgm family of transposable elements in soybean. Genetics. 1988;120(2):597–604. pmid:2848748
  39. 39. Zabala G, Vodkin L. A putative autonomous 20.5 kb-CACTA transposon insertion in an F3'H allele identifies a new CACTA transposon subfamily in Glycine max. BMC Plant Biol. 2008;8:124. pmid:19055742; PubMed Central PMCID: PMCPMC2613891.
  40. 40. Zabala G, Vodkin LO. Methylation affects transposition and splicing of a large CACTA transposon from a MYB transcription factor regulating anthocyanin synthase genes in soybean seed coats. PLOS ONE. 2014;9(11):e111959. pmid:25369033
  41. 41. Raval J, Baumbach J, Ollhoff AR, Pudake RN, Palmer RG, Bhattacharyya MK, et al. A candidate male-fertility female-fertility gene tagged by the soybean endogenous transposon, Tgm9. Funct Integr Genomics. 2013;13(1):67–73. pmid:23184475
  42. 42. Baumbach J, Pudake RN, Johnson C, Kleinhans K, Ollhoff A, Palmer RG, et al. Transposon tagging of a male-sterility, female-sterility gene, St8, revealed that the meiotic MER3 DNA helicase activity is essential for fertility in soybean PLOS ONE 2016;11(3):e0150482. pmid:26930200
  43. 43. O'Malley RC, Alonso JM, Kim CJ, Leisse TJ, Ecker JR. An adapter ligation-mediated PCR method for high-throughput mapping of T-DNA inserts in the Arabidopsis genome. Nat Protoc. 2007;2(11):2910–7. Epub 2007/11/17. pmid:18007627.