Breeding Science
Online ISSN : 1347-3735
Print ISSN : 1344-7610
ISSN-L : 1344-7610
Research Papers
Selection criteria for SNP loci to maximize robustness of high-resolution melting analysis for plant breeding
Yoshiyuki YamagataAtsushi YoshimuraToyoaki AnaiSatoshi Watanabe
Author information
JOURNAL FREE ACCESS FULL-TEXT HTML
Supplementary material

2018 Volume 68 Issue 4 Pages 488-498

Details
Abstract

DNA markers are useful for identifying genes and developing new genetic materials for breeding and genetic research. High-resolution melting (HRM) analysis can detect a single nucleotide polymorphism (SNP) in two polymerase chain reaction (PCR) fragments as a melting temperature (Tm) difference without additional experimental steps, such as gel electrophoresis. To design a method for developing reliable HRM markers that discriminate between homozygous alleles containing SNPs, we tested new evaluation indexes related to the thermodynamics of double-stranded DNA to find one that maximizes the difference in Tm values between PCR fragments. We found that differences in the change in Gibbs free energy (Δ) correlated with actual differences in Tm values. Optimization of the nearest neighboring nucleotide (NNN) of a SNP by nucleotide substitution in the primer and reducing the size of the PCR fragment both enlarged the actual differences in Tm. The genetic DNA markers we developed by NNN substitution, termed NNNs-HRM markers, could be precisely mapped within soybean chromosomes by linkage analysis. We developed a Perl script pipeline to enable the automatic design of a massive number of NNNs-HRM markers; these scripts are freely available and would be useful for practical breeding programs for other plant species.

Introduction

DNA markers are an indispensable tool in plant and animal breeding. Polymorphisms based on restriction fragment length (RFLP), simple sequence length (SSLP), amplified fragment length (AFLP), and random amplified polymorphic DNA (RAPD) have been widely used as DNA markers in genetic analysis and marker-assisted selection (MAS) in practical breeding programs (Collard and Mackill 2008). More recently, with the advent of high-throughput genotyping platforms based on third-generation sequencing technology, genome-wide single nucleotide polymorphisms (SNPs) have become easily attainable for use as markers (Nielsen et al. 2011, Pabinger et al. 2014). Utilization of SNPs plays a critical role in cost-effective and rapid MAS programs, especially in breeding using modern elite varieties that are highly genetically related to each other, e.g., Glycine max (soybean), and show genetic bottlenecks during domestication (Hyten et al. 2006). SNPs are grouped into four classes: class I SNPs are non-complementary transitions, e.g., cytosine (C) to thymine (T) and guanine (G) to adenine (A); class II SNPs are non-complementary transversions, e.g., C to A and G to T; class III SNPs are polymorphisms of complementary nucleotides with three hydrogen bonds, e.g., G to C; and class IV SNPs are polymorphisms of complementary nucleotides with two hydrogen bonds, e.g., A to T. In many organisms, including G. max and Oryza sativa (rice), class I and II SNPs are quite frequent, whereas class III and IV SNPs are relatively rare.

Highly multiplexed genotyping platforms for SNPs, such as Infinium BeadChip assay (Illumina, Inc., San Diego, CA, USA), MASS-array systems (Sequenom, Inc., San Diego, CA, USA), and high-throughput sequencers (e.g., HiSeq, Illumina; Ion Torrent, Thermo Fisher Scientific Inc., Waltham, MA, USA; and RSII, Pacific Biosciences, Menlo Park, CA, USA), are particularly useful for genotyping a huge number of loci in a few samples; however, due to their excessive power and cost, they are not ideal for genotyping a few loci in a huge number of samples. A well-designed experimental plan that considers cost-effectiveness is therefore required prior to such large-scale genotyping. In contrast, in the average laboratory, for practical reasons, clinical diagnosis and agricultural breeding studies require rapid and cheap uniplex SNP genotyping methods at a small number of loci. Uniplex SNP genotyping platforms such as TaqMan (Thermo Fisher Scientific) and Molecular Beacon assay (Premier Biosoft International, Inc., Palo Alto, CA, USA; Szemes and Schoen 2003) require fluorescence-labeled probes to be prepared for each target locus. Another uniplex SNP genotyping method, allele-specific polymerase chain reaction (ASPCR; Ugozzoli and Wallace 1991), and its derivatives have been applied in several procedures: the Mismatch Amplification Mutation Assay (MAMA; Cha et al. 1992), Amplification-Refractory Mutation System-PCR (ARMS-PCR; Ye et al. 2001), Kompetitive Allele Specific PCR (KASP; Semagn et al. 2014), and Tm-shift for discrimination of SNPs or wildtype/mutant using two kinds of allele-specific oligos (ASO1 and ASO2; Wang et al. 2005). Since ASPCR is based on the difference in amplification efficiency of a perfectly matched and mismatched primer to template DNA, amplification from mismatched priming is not completely avoided (Kwok et al. 1990, Liu et al. 2012) even with optimization of annealing temperature by use of gradient PCR (Tortajada-Genaro et al. 2017); this frequently causes pseudo-positive detection of mismatched alleles. Introduction of an artificial mismatch within two to five bases closest to the 3′ end of the primers can produce more accurate genotyping (Cha et al. 1992, Hayashi et al. 2004, Kwok et al. 1990, Liu et al. 2012); however, the best patterns of introduced mismatch bases near the 3′ end of primers have not been quantitatively predicted based on thermodynamics so far.

Another genotyping method, high-resolution melting (HRM), has been used to detect SNPs or mutations within target PCR fragments (Fig. 1A; Vossen et al. 2009). HRM uses a double-stranded DNA (dsDNA)–specific fluorescent dye to monitor decreases in fluorescence signals due to thermal dissociation of dsDNA to single-stranded DNA (ssDNA) during highly regulated temperature increases. Large differences between thermal dissociation of two dsDNA fragments containing single nucleotide substitutions, insertions, or deletions allows more accurate discrimination of genotypes in HRM analysis. The most important parameter to predict thermodynamics of dsDNA-to-ssDNA conversion is Tm value (melting temperature calculated from the above decreases in fluorescence signals). Tm values are determined mainly by the number of hydrogen bonds at each Watson–Crick base pairing and the base-stacking interaction between two base pairs of nearest-neighboring (NN) nucleotides (Crothers and Zimm 1964, SantaLucia 1998). The NN model predicts Tm based on two independent parameters: enthalpy (ΔH°) and entropy (ΔS°) at two base pairs of neighbor nucleotides as follows:

Fig. 1

Example calculation of the |ΔΔ3bp| index and an approach to maximize the |ΔΔ3bp| for the nearest neighbor nucleotide substitution in the forward primer. (A) Primer locations of an HRM marker for amplification of dsDNA harboring a class I SNP A/G at position i. (B) Definition of |ΔΔ3bp|. (C) Combination of the nearest neighbor nucleotide and SNP giving highest |ΔΔ3bp| for class I SNPs A/G. Introduction of T nucleotide at the 5′-nearest neighbor nucleotide flanking the SNP site is the best modification. |ΔΔ3bp| scores of all other combinations of 3-bp nucleotides are listed in Supplemental Table 3. (D) Site-directed mutagenesis by PCR for maximization of |ΔΔ3bp|.

  
T m = Δ H ° T o t a l R ln ( C / 4 ) + Δ S ° T o t a l(1)

where R is the gas constant (1.987 cal/K mol), C is total oligonucleotide concentration (C/4 is used because it is assumed that the two single-stranded DNAs resulting from denaturation of each dsDNA fragment are of equal concentration; SantaLucia 1998), T is Kelvin temperature, and ΔTotal and ΔTotal of a single dsDNA with n base pairs are calculated by summing at the i-th neighboring nucleotide:

  
Δ H ° T o t a l = i = 1 n - 1 Δ H ° i and  Δ S ° T o t a l = i = 1 n - 1 Δ S ° i .

Tm is also predicted based on Gibbs free energy (Δ) as follows:

  
T m = Δ G ° T o t a l R ln ( C / 4 )(2)

where Δ G ° T o t a l = i = 1 n - 1 Δ G ° i.

Equation (2) can be derived from Equation (1) through the following relationship:

  
Δ G ° T o t a l = Δ H ° T o t a l - T Δ S ° T o t a l = - R T ln ( K )(3)

where K is the equilibrium constant for DNA fragments (ratio of single-stranded to double-stranded DNA in solution).

Although there have been some reports of the use of HRM markers for genetic fingerprinting, gene mapping, and MAS in plants (Simko 2016), application of HRM analysis to plant research has been limited. Therefore, we expect that the rational design of HRM markers will promote the use of SNPs in a wide range of applications. The aim of the current study was to use the NN model to design reliable HRM markers to discriminate homozygous alleles, especially those containing a SNP within a PCR amplicon.

We proposed some evaluation indexes calculated based on ΔG°, ΔH°, and ΔS° to easily select SNPs with a large difference in predicted Tm values between PCR fragments harboring a SNP, and we evaluated the correlations between these evaluation indexes and the observed difference in Tm values (ΔTmobs) between two PCR fragments (experiment [Exp.] 1). We then designed primers that had a nucleotide substitution compared to the original genome sequence to enlarge the difference in predicted ΔG°, and we evaluated the effects of the nucleotide substitution (Exp. 2) on the difference in Tm values between the resultant two PCR fragments. We also evaluated the correlation between PCR product size and ΔTmobs (Exp. 3). Finally, we evaluated the usefulness of the obtained HRM markers for linkage analysis and for more genetically diversified samples (Exp. 4). Our Perl scripts for the design of primers for HRM markers are freely available for plant breeding and research studies.

Materials and Methods

Evaluation indexes for predicting the difference between Tm values of two PCR fragments

We assume two dsDNA fragments of n bp, dsDNA 1 and 2, derived from alleles 1 and 2, respectively, which share the same sequence except for one nucleotide (1 SNP) at position i (Fig. 1A). The dsDNA fragments are at concentration C in the reaction mixture and have Tm values of Tm1 and Tm2, respectively. According to equation (1), the absolute value of ΔTm = |Tm1Tm2| is defined as follows:

  
Δ T m = | Δ H ° 1 , T o t a l R ln ( C / 4 ) + Δ S ° 1 , T o t a l - Δ H ° 2 , T o t a l R ln ( C / 4 ) + Δ S ° 2 , T o t a l |(4)

where total enthalpies and entropies of the dsDNA 1 and dsDNA 2 are Δ1,Total and Δ2,Total, and Δ1,Total and Δ2,Total, respectively. The entropy correction for salt concentration such as [Na+] in reaction mixtures is given in SantaLucia (1998) by the equation Δ([Na+]) = Δ(1 M NaCl) + 0.368 × N × ln[Na+]. Since accounting for the salt concentration did not improve the ΔTm prediction, we used the unified oligonucleotide Δ and Δ NN parameters in 1 M NaCl from SantaLucia (1998).

According to equation (2), we can also estimate ΔTm as follows:

  
Δ T m = | Δ G ° 1 , T o t a l R ln ( C / 4 ) - Δ G ° 2 , T o t a l R ln ( C / 4 ) | = | ( Δ H ° 1 , T o t a l - T m 1 Δ S 1 , T o t a l ) - ( Δ H ° 2 , T o t a l - T m 2 Δ S 2 , T o t a l ) R ln ( C / 4 ) |(5)

When we assume Tm1 = Tm2 + ΔTm,

  
Δ T m = | ( Δ H ° 1 , T o t a l - T m 2 Δ S ° 1 , T o t a l ) - ( Δ H ° 2 , T o t a l - T m 2 Δ S ° 2 , T o t a l ) R ln ( C / 4 ) + Δ S ° 1 , T o t a l | Δ T m = | ( Δ H ° 1 , T o t a l - Δ S ° 2 , T o t a l ) - T m 2 ( Δ H ° 1 , T o t a l - Δ S ° 2 , T o t a l ) R ln ( C / 4 ) + Δ S ° 1 , T o t a l |(6)

We defined the Δ and Δ of NN between the (i − 1)-th and i-th nucleotides as ΔNN=i−1 and ΔNN=i−1, and that between the i-th and (i + 1)-th nucleotides as ΔNN=i and ΔNN=i (Fig. 1B). We defined Δ, Δ, and Δ of NN at 3-bp (i − 1, i, and i + 1) for dsDNA s (s = {1, 2}) (Fig. 1B) as follows:

  
Δ H ° s , 3 b p = Δ H ° s , N N = i - 1 + Δ H ° s , N N = i Δ S ° s , 3 b p = Δ S ° s , N N = i - 1 + Δ S ° s , N N = i Δ G ° s , 3 b p = Δ H ° s , 3 b p - T Δ S ° s , 3 b p

Since the dsDNA1 and dsDNA2 are the same except for one SNP at position i,

  
Δ H ° 1 , 3 b p - Δ H ° 2 , 3 b p = Δ H ° 1 , T o t a l - Δ H ° 2 , T o t a l Δ S ° 1 , 3 b p - Δ S ° 2 , 3 b p = Δ S ° 1 , T o t a l - Δ S ° 2 , T o t a l

Therefore, equation (6) can be rewritten as follows:

  
Δ T m = | ( Δ H ° 1 , 3 b p - Δ H ° 2 , 3 b p ) - T m 2 ( Δ S ° 1 , 3 b p - Δ S ° 2 , 3 b p ) R ln ( C / 4 ) + Δ S ° 1 , T o t a l | | Δ G ° 1 , 3 b p - Δ G ° 2 , 3 b p | | Δ Δ G ° 3 b p |(7)

In equation (7), ΔTm is proportional to |ΔΔ3bp| at T = Tm2 since Rln(C/4) + Δ1,Total is a constant when the sequence of dsDNA 1 is known. Since |ΔΔ3bp| at T = 37°C is highly correlated with that at 73°C (r = 0.99), we used the list of NN parameters calculated with T = 37°C (SantaLucia 1998) for our prediction.

For ΔTm estimation, equation (4) uses all the sequence information of the PCR fragments, whereas equation (7) uses only Δ3bp and Δ3bp. The |ΔΔ3bp| index uses the three nucleotides around the SNP without any estimation of the ΔTm of two DNA fragments containing a single SNP. In the experiments described below, we used two indexes: (1) ΔTmpred, the difference between the Tm values of PCR fragments obtained from dsDNA 1 (Tm1pred) and 2 (Tm2pred), was predicted using equation (4) as follows: ΔTmpred = |Tm1predTm2pred|; and (2) |ΔΔ3bp| was used as prediction index for the observed differences of Tm between dsDNA 1 and dsDNA 2 (ΔTmobs) in HRM analysis. Tmobs (see below) was defined as the maximum peak of −Δf/Δt during dissociation, and the absolute difference in the Tmobs values of the two PCR fragments was used for comparison.

Plant materials

Glycine max cultivars, ‘Fukuyutaka’ (accession no., JP29668 in Genebank of the National Agriculture and Food Research Organization [NARO]), ‘Toyoshirome’ (not available in Genebank; bred at the Kyushu Okinawa Agricultural Research Center, NARO), and Chinese G. max landrace ‘Aokimame’ (accession no., JP28298) were used as parental lines. The F7 population containing 96 plants derived from a cross between ‘Fukuyutaka’ and ‘Aokimame’ (hereafter, AFF7) was used as the mapping population. O. sativa Japonica rice cultivar ‘Nipponbare’ and a wild rice AA-genome species Oryza barthii (accession no., W0652) were used. W0652 was kindly provided by the National Institute of Genetics, Mishima, Japan.

SNP data

To develop the HRM DNA markers, we used two sets of soybean SNP data obtained with restriction-site-associated DNA sequencing (RAD-seq) analysis: the first set was between ‘Aokimame’ and ‘Fukuyutaka’ with HaeIII digestion (hereafter, AF SNPs); and the second set was between ‘Toyoshirome’ and ‘Fukuyutaka’ with PstI–MspI–CviQI digestion (TF SNPs). Our previous study (Watanabe et al. 2017) shows details of the RAD-seq analyses; lists of SNPs identified with high confidence (vcf format) containing information about 8372 loci for AF SNPs and 6142 loci for TF SNPs were used as input files. SNP data between ‘Nipponbare’ and W0652 (hereafter, SB SNPs) were obtained from variant analysis of whole genome sequences of W0652 and the reference sequence of ‘Nipponbare’ (Os-Nipponbare-Reference-IRGSP-1.0) by using Burrows–Wheeler Aligner (Li and Durbin 2009) and Samtools mpileup version 1.1 (Li et al. 2009) with a filtering quality score of ≥100. Variant whole genome sequences of W0652 were downloaded from DDBJ sequence archives (accession no., DRP003801).

Primer design

In Exp. 1, we selected 48 HRM primer pairs for AF SNPs by 1) retrieving 201-bp genomic sequences around SNPs (SNP located at the center of sequence) from G. max reference genome sequence (cultivar ‘Williams82’, Glyma1.0 version); 2) designing primers with the stand-alone Primer 3 program (Untergasser et al. 2012) with maximum length of PCR product set to 80 bp; and 3) conducting BLASTN searches to count the number of loci showing identical sequence with the designed primer sequence in the G. max genome (Camacho et al. 2009). We chose primer pairs in which one primer matched a single identical sequence in the soybean genome database and the other matched two or less identical sequences in the database. After designing the primers, we calculated two evaluation indexes, ΔTmpred and |ΔΔ3bp|, for each predicted PCR fragment. These steps for designing primers for each SNP were conducted automatically with Perl scripts (we provide a flow chart and all scripts used in this study as Supplemental Fig. 1 and Supplemental Text 1). We then calculated the Pearson’s correlation coefficient between each of these proposed indexes and the ΔTmobs.

In Exp. 2, we designed HRM primers with or without consideration of nearest neighboring nucleotide (NNN) substitution; the resultant primers were termed ‘NNNs-HRM markers’ and ‘non-NNNs-HRM markers’, respectively. Taking into account the nucleotides at positions i − 1 and i + 1, where i is the position of the SNP, we identified the two-nucleotide combination giving the highest |ΔΔ3bp| index (Fig. 1C); using this information, we designed a forward primer for site-directed mutagenesis to convert the 5′ or 3′-NNN of the SNP site to the optimal nucleotide (Fig. 1D). The reverse primer was designed by using the Primer 3 program. We designed two sets of 16 combinations of forward and reverse primers; in the first set, forward primers contained a nucleotide substitution just upstream of the SNP (M65 to M80 in Supplemental Table 1), compared with the original sequence; in the other set, all primers (M49 to M64) contained sequence identical to the reference genome sequence.

In Exp. 3, we designed 24 pairs of NNNs-HRM primers producing 50–100 bp PCR products with |ΔΔ3bp| index thresholds of >1.43 kcal/mol from TF SNPs data sets on G. max chromosome 10. All primers used in this study are listed in Supplemental Table 1. For more genetically divergent cases, we calculated |ΔΔG°| by subtraction of two ΔG° scores calculated from the sequence of dsDNA 1 and dsDNA2 (Supplemental Text 2) according to equation (3).

DNA marker analysis

Total genomic DNAs of young trifoliate leaflets from parental lines and the AFF7 population, and adult leaves from ‘Nipponbare’ and W0652 plants were isolated by using the cetyl trimethylammonium bromide method (Murray and Thompson 1980). HRM analysis was conducted using quantitative PCR equipment (LightCycler 96 System; Roche Diagnostics K.K., Minato, Tokyo, Japan). PCR mixtures contained 1× PCR buffer (400 mM Tricine-KOH [pH 9.2], 150 mM Sodium acetate [pH 6.0], 35 mM Magnesium acetate, 37.5 μg/mL bovine serum albumin, 0.5% (v/v) Nonidet P-40, 0.5% (v/v) Tween 20), 0.2 mM each dNTP, 20–30 ng genomic DNA, 0.25 units of home-made recombinant Thermus aquaticus (Taq) DNA polymerase (activity roughly equivalent to 0.25 units of Takara Ex Taq Hot Start Version from Takara, Kusatsu, Shiga, Japan), 0.5× fluorescent dye EvaGreen (Biotium, Inc., Fremont, CA, USA), and 0.1 pmol each HRM primer. To prepare the home-made recombinant Taq DNA polymerase, we used the pET-vector protein expression system (Merck Japan, Meguro, Tokyo, Japan) as described briefly below. The gene encoding DNA polymerase I (accession number J04639) of Taq strain JCM 10724 (provided by National Bio Resource Project) was cloned into pET-28a (+); the resulting vector was used to transform Escherichia coli strain XL21-AI (Thermo Fisher Scientific). Production of recombinant Taq polymerase was induced according to the manufacturer’s protocols. The pellet of E. coli cells from a 1 L culture with 0.8 OD was suspended in 30 mL buffer B (20 mM Tris-HCl [pH 7.2], 500 mM NaCl, 5 mM imidazole, 0.1% Triton X-100) and homogenized with an Ultrasonic Generator 201M (Kubota, Hongo, Tokyo, Japan). The cell lysate was incubated at 75°C for 30 min and then centrifuged at 16,000 ×g for 20 min to eliminate insoluble proteins. Polyhistidine-tag fusion recombinant Taq polymerase was purified with 4 mL Ni Sepharose 6 Fast Flow (GE Healthcare Japan, Shinjuku, Tokyo, Japan) according to the manufacturer’s instructions. After elution of the recombinant Taq polymerase with buffer B containing 100 mM imidazole (instead of 5 mM imidazole), the buffer was substituted with the storage buffer (20 mM Tris-HCl [pH 8.0], 100 mM KCl, 0.1 mM EDTA, 1 mM dithiothreitol, 0.5% (v/v) Nonidet-P40, and 0.5% (v/v) Tween 20) by dialysis. Enzyme activity was roughly estimated by comparing the PCR results of commercial Taq DNA polymerase (Takara Ex Taq Hot Start Version) with that of a dilution series of purified recombinant Taq DNA polymerase. PCR conditions were pre-incubation at 95°C for 5 min, and 40 cycles of 95°C for 15 s, 60°C for 15 s, and 72°C for 20 s; melting analysis was then performed at 95°C for 10 s, 65°C for 1 min, and then increasing temperatures to 97°C with a slope of 0.1°C/s with continuous measurement of the fluorescence of the PCR product. The reproducibility of genotypes obtained by HRM analysis was evaluated with 7 NNNs-HRM markers (AF_M65, AF_M67, AF_M69, AF_M73, AF_M75, AF_M76, and AF_M77) with the same conditions mentioned above except that the DNA polymerase and PCR buffer were provided by Takara. The post-run analysis of HRM data was conducted with software provided by the LightCycler 96 System, and the Tmobs of each genotype was calculated as the peak value of −Δft, where Δf is change in fluorescence and Δt is change in temperature. Single band amplification for all PCR amplicons was confirmed with 12% polyacrylamide gel electrophoresis.

Data analysis

Linkage map analysis of the AFF7 population was performed with Antmap software with default parameters for recombinant inbred lines (Iwata and Ninomiya 2006). Correlation analysis and paired t-tests (alternative hypotheses, Tmobs of non-NNNs HRM markers < that of NNNs-HRM markers) were conducted with R (ver. 3.3.1) software (R Development Core Team 2008).

Results

Correlation between |ΔΔG°3bp| and observed Tm

We investigated SNP data sets obtained from intraspecific comparison of G. max data (AF SNPs and TF SNPs) and from interspecific comparison of genus Oryza data (SB SNPs) (Table 1). The class I and II SNPs displayed the first and second highest frequencies in all data sets, whereas class III and IV SNPs showed the lowest frequencies; only class I and II SNPs were sufficiently abundant to be used as HRM markers.

Table 1 Frequencies of SNP classes in soybean and rice SNP datasets
Dataset Genus SNP class Total No. of SNPs
I II III IV
AF Glycine 64.2% 17.7% 6.2% 16.4% 8372
TF Glycine 56.9% 19.8% 6.9% 11.3% 6142
SB Oryza 71.3% 15.1% 5.8% 7.8% 4138822

From the 48 primer pairs designed from the AF SNP dataset (see Materials and Methods), we prepared 38 (79.2%) primer pairs that each produced single-band amplification of a different fragment that was polymorphic between the parents, ‘Aokimame’ and ‘Fukuyutaka’ (Exp. 1). For each primer pair, absolute differences in the Tmobs values of the PCR products obtained from two varieties (i.e., |Tm1obsTm2obs| = ΔTmobs) were calculated (Supplemental Table 2). The ΔTmobs and ΔTmpred values (predicted using whole sequences of dsDNA fragments) were positively correlated (r = 0.648, P < 0.01; Fig. 2A), as were the ΔTmobs and |ΔΔ3bp| scores (r = 0.647, P < 0.01; Fig. 2B). Because the correlation coefficients for the above relationships were almost identical, these results demonstrate that |ΔΔ3bp| score has an equivalent prediction power to ΔTmpred. Selection for primer pairs with the highest |ΔΔ3bp| scores provides us a selection criterion for those that would detect the largest differences in Tmobs values between two variants at an HRM marker.

Fig. 2

Correlation between ΔTmobs and evaluation indexes. (A) Correlation between ΔTmobs and ΔTmpred (r = 0.648). (B) Correlation between ΔTmobs and |ΔΔ3bp| (r = 0.647).

Optimization of HRM markers

Since larger ΔTmobs values in HRM analysis reduce genotyping errors due to factors such as variance in reaction volumes and amplification efficiency across wells, we attempted to optimize the design of the HRM markers for more reliable genotyping by conducting site-directed mutagenesis of the 5′ or 3′ NNN of the SNP with forward primer (Fig. 1C, 1D).

To determine which, if any, 5′ or 3′ NNNs maximize |ΔΔ3bp|, we calculated |ΔΔ3bp| for various types of SNPs (Figs. 1C, 1D, 3). For C/T (Y) SNPs (an example of class I SNPs), |ΔΔ3bp| was maximized by the presence of G or A nucleotide at the 5′ or 3′ NNN (upstream or downstream of SNP, respectively), respectively; for G/A (R) SNPs, |ΔΔ3bp| was maximized by the presence of T or C nucleotide at the 5′ or 3′ NNN, respectively (Fig. 3). Among class II SNPs, G nucleotides at 5′ or 3′ NNN for C/A (M) SNPs and C nucleotides at 5′ or 3′ NNN for G/T (K) SNPs provided maximum |ΔΔ3bp|. These data suggest that nucleotide substitutions immediately upstream or downstream of the SNP are critical determinants of |ΔΔ3bp| at class I and class II SNP sites. However, neither 5′ nor 3′ NNNs at class III and IV SNP sites showed a clear trend for maximizing |ΔΔ3bp|. In Supplemental Table 3, we provide a quick reference matrix for practical guidelines to select suitable NNNs for any SNP.

Fig. 3

|ΔΔ3bp| for class I, II, III, and IV SNPs sorted by 5′ or 3′ nearest neighbor nucleotides (NNN) relative to the SNP. In the groups of three nucleotides, first the 5′ NNNs are arranged A, T, C, G, and then the 3′ NNNs are arranged A, T, C, G. In each panel, SNPs with the highest |ΔΔ3bp| are marked by rectangles.

We then developed a pipeline to automatically design primers to maximize |ΔΔ3bp| scores in HRM analysis. Of the 32 primer pairs, which comprised NNNs-HRM markers (M65 to M80, Supplemental Table 1) and their original non-NNNs-HRM markers (M49 to M64) at 16 SNP loci from the AF SNP dataset, 24 (75.0%) showed single amplification and polymorphism between parental lines. For example, the non-NNNs-HRM marker, M60, and the NNNs-HRM marker, M76, amplified genomic sequences containing the class II SNP (5′-AKT-3′) in DNA strands identical to reference sequence of G. max at 42.7 Mbp on chromosome 1 (Gm01). The results in Fig. 3 suggest that, to maximize |ΔΔ3bp|, the A nucleotide at i-1th position should be converted to C by PCR-mediated site-directed mutagenesis. After this substitution, the |ΔΔ3bp| scores improved from 0.84 to 1.34 kcal/mol and the ΔTmobs values improved from 0.53 to 0.64°C (Supplemental Table 4, Supplemental Fig. 2). Even for the non-NNNs-HRM marker, M64, an example of a class IV SNP (A/T), which had an extremely low |ΔΔ3bp| of 0.05 kcal/mol, the NNN substitution improved ΔTmobs values and provided successful genotyping of the AFF7 population (marker M80_AF, Supplemental Table 4, Supplemental Fig. 2; see section below). Consequently, the average ΔTmobs values of non-NNNs-HRM and NNNs-HRM markers were 0.43°C and 0.63°C, respectively (Supplemental Table 4), and the effect of NNN substitution in primer to increase the ΔTmobs values was significant (11 combinations of non-NNNs and NNNs-HRM markers were used in Supplemental Table 4; df = 10, t value = 7.6, P < 0.01, paired t-test), demonstrating that the concept of |ΔΔ3bp| allowed rational design of HRM markers via NNN substitution.

Next, we investigated the correlation between PCR product size and ΔTmobs. We designed 24 NNNs-HRM markers that displayed |ΔΔ3bp| of >1.43 kcal/mol (Supplemental Table 5) with PCR product sizes of 54–100 bp, and then used them for HRM analysis. Out of 24 primers, 3 showed no polymorphism between parents and another 3 showed a dominant segregation pattern; these primers were not analyzed further. The remaining 18 primers (75%) showed co-dominant segregation among three genotypes. The ΔTmobs values between parental lines (mean, 0.74°C; Supplemental Table 5) were significantly negatively correlated with the length of PCR product (r = −0.75, P < 0.01; Fig. 4). This result indicates that not only maximization of |ΔΔ3bp| but also minimization of PCR product length increased ΔTmobs. We also evaluated reproducibility of genotyping by using different commercial Taq polymerases with 7 NNNs-HRM markers: 653 out of 672 genotypes (97.2%) showed identical genotype results regardless of the source of Taq polymerase (Supplemental Fig. 3, Supplemental Table 6).

Fig. 4

Correlation between the ΔTmobs values for NNNs-HRM markers and size of PCR fragment.

A drawback of thresholding by |ΔΔ3bp| and limiting PCR product length is the decreased flexibility in PCR primer design. Without consideration of NNN substitution, when the ΔTmpred of PCR products was thresholded at >0.6°C and the PCR product length was limited to <70 bp in the AF SNP dataset, we succeeded in designing primers at only 545 (8.0%) of 6830 loci (Fig. 5A). However, under the same conditions, consideration of NNN substitution greatly improved the percentage of primer design such that it increased to 1779 (31.4%) of 5665 loci (Fig. 5B). Primer3_core was used to calculate the melting temperatures of oligos and the tendencies of oligos to form hairpins or dimers (Untergasser et al. 2012). We fixed the position of primer including nucleotide substitution, and optimization of the substituted nucleotide decreased the variations of nucleotides and increased the percentage of NNN markers having hairpin structures (e.g., GC and CG) at the 3′ end of the sequence: 3.5% in non-NNNs and 13.2% in NNNs for AF SNP (data not shown). Optimization of NNNs is the likely cause of the low number of successfully designed primer pairs (6860 to 5665 loci). However, 5269 SNP loci were commonly used to design non-NNNs and NNNs-HRM markers, and even if we use a threshold for primer inclusion of maximum number of loci for a single primer of ≤2 according to BLASTN search, close to 1000 NNNs-HRM markers would be available for analysis with AF and TF SNPs (Supplemental Tables 7, 8).

Fig. 5

Effect of optimization for NNN substitution flanking SNPs. Scatter plot of ΔTmpred values of two PCR fragments and PCR product size without nucleotide substitution (A) and with nucleotide substitution (B). All AF SNPs were used to design primer pairs and simulate ΔTmpred of two PCR amplicons containing a SNP. The square boxes include the NNNs-HRM markers with the ΔTmpred of >0.6 and PCR product size of <70 bp. The numbers of markers (SNPs) in the square boxes are indicated in the figure.

Validation of genotypes obtained with NNNs-HRM markers

We further validated that the NNNs-HRM markers mapped at precise loci on the soybean genetic linkage map. We analyzed the genotype of AFF7 plants by using ten combinations of primer pairs and found that non-NNNs-HRM and NNNs-HRM markers targeting the same SNP mapped on the same genetic locus (Fig. 6A). Moreover, the relationship between genetic and physical positions showed a typical sigmoid-like pattern, with recombination suppressed around the pericentromeric region (Fig. 6B). These results indicate that use of NNNs-HRM markers did not affect the determination of genetic location or genotype.

Fig. 6

Genetic linkage map of soybean chromosome 1 (Gm01) constructed by using the AFF7 population with non-NNNs-HRM markers and NNNs-HRM markers. Non-NNNs-HRM and NNNs-HRM markers detecting the same SNP are arranged next to each other and are indicated with vertical black rectangles, e.g., M49_AF and M65_AF; numbers at the left of the map indicate genetic distance (cM) (A). A scatter plot of the physical positions versus map distances of these HRM markers is shown (B).

Availability of HRM markers for more genetically divergent cases

We tried to extend application of |ΔΔ| for estimation of ΔTmobs to PCR fragments carrying more than one SNP derived from more diverged species: the cultivated rice species O. sativa L. and wild species O. barthii. In a plot of |ΔΔ| against ΔTmobs (Fig. 7, Supplemental Text 2), the values of the PCR products containing single SNPs were close to each other on both the |ΔΔ| and ΔTmobs axes, whereas the values of ones containing more than one SNP were scattered widely. Although this observation is qualitative, |ΔΔ| prediction accuracy for ΔTmobs may be lower when the two PCR products contain more than one SNP.

Fig. 7

Scatter plots of |ΔΔ| versus ΔTmobs for PCR fragments from Oryza sativa and O. barthii. Circles and triangles represent PCR fragments containing single SNPs and more than one SNP, respectively. Digits represent numbers of SNPs in the PCR fragments.

Discussion

Current protocols for choosing which SNPs will be more successful for genotyping in HRM analysis are based on whether the SNPs are class I, II, III, or IV. Here, we suggest a novel criterion that considers the 5′- and 3′-nucleotides flanking the SNP sites, and we make available a Perl script pipeline for the automatic design of a massive number of suitable HRM markers from SNP datasets. For electrophoresis-based DNA markers such as simple sequence repeats (SSR) and sequence-tagged site (STS) markers, large differences of dsDNA length derived from two alleles enable more accurate genotyping. Similarly, as large as possible Tm differences between dsDNAs derived from two alleles are required in HRM analysis. Because two dsDNA fragments containing SNPs optimally have a large ΔTmobs, an easy and practical technique to predict ΔTmobs and thereby select reliable HRM markers would have wide applications in genotyping. Here, we compared two indexes for prediction of Tmobs, i.e., |ΔΔ3bp| and ΔTmpred, and we found that both indexes had equivalent power to predict ΔTmpred. Because calculation of ΔΔ3bp (which requires addition of thermodynamic parameters for three nucleotides) is easier than calculation of ΔTmpred (which requires analysis of the entire nucleotide sequence of the PCR amplicon), we chose to use |ΔΔ3bp| to select SNP sites available for genotyping or to predict the best modification of the 5′- or 3′-NNN at a given SNP site in further experiments. Our empirical observation revealed that a ΔTm value >0.5°C guaranteed accurate genotyping using the Roche LightCycler 96 system (data not shown). All of the Class I and II NNNs-HRM markers with an amplicon size of <90 bp achieved a ΔTmpred of >0.5°C (Fig. 5B), whereas 61.4% (3466/5647 for Class I and II) of non-NNNs-HRM markers did not (Fig. 5A). Furthermore, we noticed that some Class III and IV NNNs-HRM markers with ΔTmobs of 0.2–0.3°C produced accurate genotypes. Using markers that satisfy the criteria of predicted |ΔΔ3bp| of >1.0 kcal/mol and amplicon size of <80 bp would generally guarantee ΔTm of >0.5°C (Fig. 2B). In HRM analysis, NNNs-HRM markers designed using this pipeline showed higher ΔTmobs values, thereby improving the reliability of genotyping regardless of the kind of Taq polymerase. Moreover, considering the size of amplicon for HRM analysis also increased ΔTmobs (Fig. 4). Many manufacturers of equipment for HRM analysis, such as Applied Biosystems (Foster City, CA, USA), Roche, Bio-Rad Laboratories (Hercules, CA, USA), Thermo Fisher Scientific, and QIAGEN (Hilden, Germany), highly recommend using short amplicons (70–150 bp) in their application manuals. Furthermore, Vossen et al. (2009) targeted fragment sizes of 80–100 bp for SNP typing. Our experimental results (Fig. 4) and simulation data (Fig. 5) support their recommendation for amplicon sizes, with <80 bp being preferable for reliable genotyping. In a previous study, we proposed a method for developing dCAPS (derived cleaved amplified polymorphic sequences) markers from TF SNPs (one of the two SNP data sets used here); we found that 2000 out of 6000 SNPs could be used for dCAPS analysis with 17 inexpensive restriction enzymes (Watanabe et al. 2017). By using the current pipeline, we could design over 3700 candidate primers for HRM analysis with the same SNP data set (Supplemental Table 7), thereby increasing the number of loci available for genetic research and breeding programs. The ability to conduct HRM analysis in a single tube (i.e., without any analysis of PCR product or other equipment) would improve the effectiveness (time and labor cost) of genotyping individuals. Using DNA markers tightly linked to a target gene can enhance the accuracy of MAS, and increasing the number of DNA markers linked to a target locus can reduce the size of the donor chromosome segment and thereby reduce linkage drag (Ribaut and Hoisington 1998). In general breeding programs using bulk population breeding methods for self-pollinated crops, such as rice, wheat, and soybean, single plants are selected at the F4 or F5 stage. If we apply MAS to these generations, discrimination of homozygous alleles is necessary. Therefore, increasing the reliability of HRM markers by enlarging the difference of Tm values is relevant.

Designing HRM markers throughout the whole genome would facilitate gene mapping and development of genetic materials such as chromosome segment substitution lines (CSSLs). Tsuda et al. (2015) developed a mutagenized population with the genetic background of Japanese soybean cultivar ‘Enrei’, and screened many mutant lines based on phenotypes such as maturity, yield, and oil and protein contents. We will confirm the effects of each mutated gene by using segregated populations derived from crosses with genetically close cultivars, such as ‘Toyoshirome’ or ‘Fukuyutaka’. These crosses between Japanese elite soybean lines should reduce the introduction of unfavorable traits and allow easy evaluation of the genetic effects of the mutated gene. Bulk segregant analysis (Quarrie et al. 1999) with HRM markers could rapidly identify a chromosome region containing a particular mutated gene. Furthermore, HRM-mapping of segregating populations can provide us with genetic markers linked to the target gene and breeding materials simultaneously. Accumulating melting curve and other data for short DNA fragments containing HRM markers under the same experimental condition could potentially be used to improve the NN parameter table of SantaLucia (1998) for more accurate estimation of DNA thermodynamics. For developing CSSLs, sets of HRM markers detecting SNPs specific to fundamental elite lines could be used to more easily discriminate the donor chromosome, which contains novel genes originating from a genetically diverse germplasm. In rice, many CSSLs with combinations of chromosomes of Japanese rice cultivars, such as ‘Taichung 65’ (T65), and exotic Oryza species have been developed (Yasui et al. 2010) with SSR markers. SNPs specific to T65 could be easily used to develop genome-wide HRM markers that could be used across different donor types to monitor chromosomal segments.

Here, we used two SNP data sets that showed similar trends of naturally occurring nucleotide substitutions to each other. As in humans (Collins and Jukes 1994), in both sets, most SNPs were class I SNPs. Transition-type substitutions occur frequently because of the similarity between the chemical structures of the bases (Collins and Jukes 1994). This means that class I SNPs show redundancy in most species, which probably increases the opportunities to use reliable NNNs-HRM markers. Maintaining the reproducibility of the shape of the melting curve (Tm values) is important (Simko 2016); therefore, future work will consider the need for uniform amplification without unspecific amplicons, the type of dye, and the preparation of DNA samples of uniform concentration. HRM markers could also be extended to include polymorphisms related to insertions and deletions.

While current genotyping techniques, such as genotyping by sequencing and DNA typing arrays, produce a large amount of information on genotypes of individuals, the quick system for developing reliable DNA markers described in this study would accelerate the use of DNA markers for large populations and could also be adapted easily to other crop species to find genes or develop new genetic materials for breeding and genetic research.

Acknowledgements

We are grateful to Tokiko Kitajima (Saga University) for technical support. This study was partially supported by a grant from the Ministry of Agriculture, Forestry, and Fisheries of Japan (SFC1003) to S. W. This research was partially supported by Science and Technology Research Partnership for Sustainable Development (SATREPS), Japan Science and Technology Agency (JST)/Japan International Cooperation Agency (JICA), and by the National Bioresource Project (Rice)/Japan Agency for Medical Research and Development to A. Y.

Literature Cited
 
© 2018 by JAPANESE SOCIETY OF BREEDING
feedback
Top