Next Article in Journal
Functional Characterization and in Silico Analysis of Phytoene Synthase Family Genes Responsible for Carotenoid Biosynthesis in Watermelon (Citrullus lanatus L.)
Next Article in Special Issue
Isolate-Dependent Inheritance of Resistance Against Pseudoperonospora cubensis in Cucumber
Previous Article in Journal
Phenotyping a Diverse Collection of Forage Sorghum Genotypes for Chilling Tolerance
Previous Article in Special Issue
Analysis of Proteomic Profile of Contrasting Phosphorus Responsive Rice Cultivars Grown under Phosphorus Deficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Variability of Puroindoline-Encoding Alleles and Their Influence on Grain Hardness in Modern Wheat Cultivars Cultivated in Poland, Breeding Lines and Polish Old Landraces (Triticum aestivum L.)

by
Mateusz Przyborowski
1,*,
Sebastian Gasparis
1,
Maciej Kała
1,
Wacław Orczyk
2 and
Anna Nadolska-Orczyk
1
1
Department of Functional Genomics, Plant Breeding and Acclimatization Institute–National Research Institute, Radzików, 05-870 Błonie, Poland
2
Department of Genetic Engineering, Plant Breeding and Acclimatization Institute–National Research Institute, Radzików, 05-870 Błonie, Poland
*
Author to whom correspondence should be addressed.
Agronomy 2020, 10(8), 1075; https://doi.org/10.3390/agronomy10081075
Submission received: 19 June 2020 / Revised: 18 July 2020 / Accepted: 22 July 2020 / Published: 25 July 2020
(This article belongs to the Special Issue Analysis of Crop Genetic and Germplasm Diversity)

Abstract

:
Wheat (Triticum aestivum L.) grain hardness is determined mainly by variations in puroindoline genes (Pina-D1 and Pinb-D1), which are located on the short arm of chromosome 5D. This trait has a direct effect on the technological properties of the flour and the final product quality. The objective of the study was to analyze the mutation frequency in both Pin genes and their influence on grain hardness in 118 modern bread wheat cultivars and breeding lines cultivated in Poland, and 80 landraces from Poland. The PCR products containing the Pin gene coding sequences were sequenced by the Sanger method. Based on detected the SNPs (single-nucleotide polymorphisms) we designed CAPS (cleaved amplified polymorphic sequence) markers for the fast screening of Pinb alleles in a large number of genotypes. All analyzed cultivars, breeding lines, and landraces possess the wild-type Pina-D1a allele. Allelic variation was observed within the Pinb gene. The most frequently occurring allele in modern wheat cultivars and breeding lines (over 50%) was Pinb-D1b. The contribution of the remaining alleles (Pinb-D1a, Pinb-D1c, and Pinb-D1d) was much less (approx. 15% each). In landraces, the most frequent allele was Pinb-D1a (over 70%), followed by Pinb-D1b (21% frequency). Pinb-D1c and Pinb-D1g were found in individual varieties. SKCS (single-kernel characterization system) analysis revealed that grain hardness was strictly connected with Pinb gene allelic variation in most tested cultivars. The mean grain hardness values were significantly greater in cultivars with mutant Pinb variants as compared to those with the wild-type Pinb-D1a allele. Based on grain hardness measured by SKCS, we classified the analyzed cultivars and lines into different classes according to a previously proposed classification system.

1. Introduction

In Europe, common wheat (Triticum aestivum L.) is cultivated on almost 61 million hectares of land, and, in Poland, on 2.416 million hectares, which accounts for about 31% of all cereals cultivated in Poland [1]. Wheat is an important component of food and feed. Wheat grains, depending on their technological properties, are used for various food products [2]. Grain hardness, one of the most important technological traits, is considered as significant as gluten protein composition [3]. It has a direct effect on the grinding and baking characteristics, flour particle sizes, and water absorption ability, which represent crucial technological flour properties [3,4]. In common wheat, grain hardness is primarily controlled by two puroindoline genes, Pina and Pinb. Both genes are intronless, with a 447 bp coding sequence and are located in locus Ha on the short arm of chromosome 5D [5]. The Pin gene coding sequences are 70.2% similar. In the course of wheat evolution, puroindoline genes located on chromosomes 5A and 5B were lost after wheat ancestor polyploidization [6]. Pina and Pinb encode the following puroindoline proteins: puroindoline a (PINA), and puroindoline b (PINB). Puroindolines are small (~13 kDa), cysteine-rich, lipid-binding proteins consisting of 148 amino acids [7,8]. PIN proteins exhibit a tryptophan-rich domain, which consists of five tryptophan residues in PINA and three residues in PINB [5]. Increased grain hardness is the result of mutations within one or both wild-type Pin alleles, which cause amino acid sequence changes in puroindoline proteins [9]. Thus far, 30 Pina alleles and 26 Pinb alleles have been described in different wheat species [10,11,12,13]. In most cases, the differences between wild-type and mutant alleles are single nucleotide polymorphisms (SNPs), such as in Pinb-D1b and Pinb-D1c, or one base deletion, like in Pinb-D1p and Pinb-D1u. Allele Pinb-D1b, originally described by Giroux and Morris [14], is characterized by guanine to adenine substitution at position 223, resulting in glycine to serine substitution in PINB at position 46 and changes in the tryptophan-rich domain (TRD). It is the most frequent Pin gene mutation among all domesticated wheat cultivars. Pinb-D1c and Pinb-D1d are the second most frequent alleles found in common wheat cultivars [11]. Pinb-D1c is characterized by thymine to cytosine substitution at position 266, leading to leucine to proline change at position 60 [15]. In Pinb-D1d, thymine to adenine substitution at position 217 causes tryptophan to arginine change at position 44 and leads to changes in the TRD [15]. An example of the change in reading frame is the allele Pinb-D1g, where a single nucleotide change at position 255 (cytosine to adenine substitution) causes a cysteine to stop codon change at position 56 [16]. An example of a SNP which does not change the amino acid sequence is the Pinb-D1ae allele, in which thymine is substituted by adenine at position 93 [17]. Less common alleles possess more than one SNP, e.g., two SNPs in Pinb-D1v [18] or one SNP and one base deletion in Pinb-D1s [19]. Pina-D1b is a null mutation resulting from the deletion of a 15.3 Kb chromosome fragment [20,21].
Genetic diversity and grain hardness allelic variation were studied in Australian commercial wheat cultivars [22], the Watkins collection [23], and bread wheat landraces from Andalusia [24]. Additionally, the trait is also easily genetically engineered [25,26,27]. Both Pin genes are considered the main grain hardness determinants. TRD is the most important domain of PIN proteins because it is responsible for their lipid-binding properties. Changes in TRD caused by SNP in Pinb-D1b and Pinb-D1d alleles result in decreased affinity to anionic phospholipids and altered interaction between PINA and PINB proteins [28,29]. TRD is comprised of five tryptophan (Trp) residues in PINA and three Trp residues in PINB [5]. Apart from puroindoline-encoding genes, very similar genes showing over 70% similarity with the Pinb nucleotide sequence have been identified. They are categorized as Pinb-2 variants and are located on chromosomes 7A, 7B, and 7C [30,31,32,33]. However, further studies clearly demonstrated that the Pinb-2 variants expression levels were very low and these genes had no significant effect on grain hardness [31,34].
In this study, we analyzed allelic variation in Pin genes in collection of hexaploid wheat cultivated in Poland, including Polish and European cultivars, as well as in Polish breeding lines and landraces, and we determined how this variation affected grain hardness. The cleaved amplified polymorphic sequence (CAPS) markers adapted to the screening of Pinb alleles proved to be useful.

2. Materials and Methods

2.1. Plant Material

The plant material consisted of 67 wheat cultivars cultivated in Poland in 2015–2018, and 51 breeding lines from Polish breeding companies, and all 80 landraces collected in Poland between 1960 and 1976, and deposited in the National Centre for Plant Genetic Resources (KCRZG) of the Plant Breeding and Acclimatization Institute – National Research Institute. Plants were grown during 2018–19 in a greenhouse, with eight plants in a twelve liter pot using standard agronomic practices including preventative spraying with plant protection products. The full list of wheat genotypes is presented in Supplementary Tables S1 and S2.

2.2. DNA Extraction

The seedlings were grown in beakers with 4 mm glass balls and autoclaved tap water, in darkness at 23 °C. Genomic DNA was extracted from coleoptile fragments and leaves of six-day old seedlings using a modified cetrimonium bromide (CTAB) method devised by Murray and Thompson [35]. DNA was extracted from the pooled tissue samples of at least six individuals from the same variety. The DNA purity and concentration were checked using a Nanodrop 1000 (NanoDrop Technologies) and diluted to 50 ng/μL for further PCR analysis.

2.3. Pina and Pinb PCR Amplification and Sanger Sequencing

Both alleles of the Pin genes were sequenced in all examined (198) wheat genotypes. Two primer pairs were used for PCR amplification and sequencing: Pina forward: 5′-CATCTATTCATCTCCACCTGC-3′; reverse: 5′-GTGACAGTTTATTAGCTAGTC-3′ (product length—524 bp); Pinb forward: 5′-GAGCCTCAACCCATCTATTCATC-3′; reverse: 5′-CAAGGGTGA TTTTATTCATAG-3′ (product length—595 bp) [36]. The PCR mix contained: 100 ng DNA, 0.5 μM of each primer, 200 μM deoxynucleotide triphosphates (dNTPs) mix, 0.4 U Phusion Hot Start II DNA Polymerase (Thermo Scientific), 4 μL 5× Phusion HF buffer, and PCR-grade water up for a 20 μL final volume. PCR was performed under the following conditions: initial denaturation at 98 °C for 30 s, followed by 36 cycles consisting of denaturation at 98 °C for 10 s, annealing at 58 °C (for both primer pairs) for 30 s, DNA elongation at 72 °C for 15 s, and, after 36 cycles, the final extension step at 72 °C for 8 min.
The amplification products were purified using the GeneJET PCR Purification Kit (Thermo Scientific), according to the manufacturer’s instructions, and then sequenced by the Sanger method using the same primers that were used for amplification. The samples were sequenced by Genomed S.A. Sequences, which were aligned using MegAlgin (DNASTAR) software.

2.4. CAPS Analysis

The Pinb gene primer sequences for amplification were as follows: forward: 5′-ATGAAGACCTTATTCC TCCTA-3′; reverse: 5′-AGTAATAGCCACTAGGGAACTT-3′ (product length—442 bp) [5]. The PCR mix was composed of: 100 ng DNA, 0.5 μM of each primer, 200 μM dNTP mix, 0.4 U Phusion Hot Start II DNA Polymerase (Thermo Scientific), 4 μL 5× Phusion HF buffer, and PCR-grade water up to a final volume of 20 μL. PCR was performed under the following conditions: initial denaturation at 98 °C for 30 s followed by 36 cycles consisting of denaturation at 98 °C for 10 s, annealing at 53 °C for 30 s, DNA elongation at 72 °C for 15 s, and, after 36 cycles, the final extension step at 72 °C for 8 min. PCR product was digested with restriction enzymes selected using the SNP2CAPS program [37] to confirm the sequencing results. The digest reaction was composed of: 5 μL PCR product, 0.7 μL restriction enzyme, 2 μL 10× compatible buffer, and PCR grade water up to a final reaction volume of 20 μL. Samples were incubated at 37 °C in a water bath for 3 h. Digested fragments were separated on a 2.5% agarose gel by electrophoresis in SB buffer (Brody and Kern, 2004). MbiI (Thermo Scientific) was used to confirm Pinb-D1b, PvuII (Thermo Scientific) for Pinb-D1c, and MnlI (Thermo Scientific) for Pinb-D1d. The amplicons containing Pinb-D1a, Pinb-D1c and Pinb-D1d alleles cut by MbiI gave two fragments, which were—318 and 129 bp long. The amplicon with Pinb-D1b cut with the same enzyme gave three fragments, of 223, 129 and 95 bp. Amplicons with Pinb-D1a, Pinb-D1b and Pinb-D1d digested with PvuII gave two fragments, which were—264 and 183 bp long, and the amplicon with Pinb-D1c, that was not cut with this enzyme had a fragment that was 442 bp long. Digestion of amplicons with Pinb-D1a, Pinb-D1b, and Pinb-D1c with MnII led to nine fragments, which had the following lengths—137, 112, 66, 45, 29, 25, 20, 5 and 3 bp. Digestion with the same enzyme of Pinb-D1d amplicon gave ten fragments, which had the following lengths—115,112, 66, 45, 29, 25, 22, 20, 5, and 3 bp.

2.5. Grain Hardness Measurement

The grain hardness was measured using the Single Kernel Characterization System (SKCS) 4100 (Perten Instruments). Single kernel characterization was measured according to Approved Method 55–31.01 [38]. The hardness of each wheat kernel was determined by the instrumental measurement of the force required to crush the kernel. The SKCS instrument was calibrated to compute the hardness index including the weight, moisture, and diameter of each kernel. For each sample, 300 individual kernels were analyzed.

2.6. Percentage of Protein, Starch, and Wet Gluten

The percentage of protein, starch, and wet gluten was measured using the Infratec Nova Grain Analyser (Foss). Analyses were performed for landraces in two independent replicates. Samples of 30 g grains were used.

2.7. Statistical Analysis

Statistical analysis was performed using the R v. 3.6.2 and the FSA v. 0.8.27 package [39]. The normality distribution was verified by the Shapiro—Wilk test. The Kruskal—Wallis test was performed to compare variable distributions of Pinb allele groups and hardness index. The Dunn test was performed in order to establish which groups were different from each other. Pearson correlationa between hardness and protein or starch or gluten content were calculated separately for landraces with the wild-type Pinb alleles and mutant alleles, as well as for all landraces, regardless of their allelic status. The chart of hardness index ranges was performed using gglot2 v. 3.2.1 package [40].

3. Results

The allelic variation of Pina and Pinb was determined by Sanger sequencing. All tested cultivars, breeding lines, and landraces had the wild-type allele of the Pina gene (Pina-D1a). The frequencies of Pinb-D1 alleles are shown in Table 1. The most frequent allele among modern cultivars and breeding lines was Pinb-D1b (57%). The wild type Pinb-D1a allele occurred in 19 lines (16%) and the frequencies of the remaining alleles, Pinb-D1c and Pinb-D1d, were at a similar level (12% and 15%, respectively).
In landraces, the Pinb frequencies were different to modern cultivars and breeding lines. Over 71% had the Pinb-D1a allele, while 26% (21) had the Pinb-D1b allele. Pinb-D1c and Pinb-D1g were each found in one tested landrace. Alignment of the nucleotide and amino acid sequences of Pinb alleles with the marked mutations is shown in Figure 1, and chromatogram fragments of five alleles found in tested wheat genotypes are shown in Supplementary Figure S1.
Next, we examined whether the mutations in the Pinb coding sequence could be detected by CAPS analysis. Restriction enzyme digestion of the amplified full length Pinb coding sequences produced allele-specific fragment patterns, which were unique for each enzyme. The Pinb-D1b, Pinb-D1c, and Pinb-D1d alleles can be distinguished from the wild-type Pinb-D1a allele by restriction analysis with MbiI, PvuII, and MnlI, respectively (Figure 2). The SKCS hardness values were grouped by each Pinb allele and by genotype (modern vs. landraces) (Figure 3). In modern cultivars and breeding lines the hardness index for each allele group ranged as follows: for Pinb-D1a, from 6 to 57; for Pinb-D1b, from 35 to 80; for Pinb-D1c, from 42 to 77; for Pinb-D1d from 40 to 82. In landraces, the hardness index for the Pinb-D1a group ranged from 9 to 37 and for Pinb-D1b from 18 to 63. The landrace with Pinb-D1c had a hardness index of 66, and another with Pinb-D1g had a hardness index of 61 (Supplementary Tables S1 and S2).
The tested genotypes were classified, as proposed by Corona et al. [41] into three classes, namely soft, medium hard, and hard types with the SKCS values in the range of 15–40, 41–70, and 71–95 respectively. In the modern cultivars group, 20 genotypes were classified as soft type, 74 as medium hard, and 24 as hard type. In the group of landraces, 61 genotypes were classified as soft and 19 as medium hard.
The protein content in the tested landraces ranged from 13.10% to 23.15%. The starch content ranged from 55.85% to 67.70%, and the wet gluten content ranged from 28.05% to 53.15% (Supplementary Table S3). None of the above values were significantly correlated at p < 0.05 with the hardness measured by SKCS. The correlation coefficient between grain hardness and protein, starch and wet gluten content calculated for all landraces was −0.18, 0.19, and −0.16, respectively. The same correlation coefficients calculated for the wild-type Pinb landraces were −0.18 (SKCS vs. protein content), 0.15 (SKCS vs. starch content), and −0.03 (SKCS vs. wet gluten content), and for the mutant landraces, these values were −0.41, 0.36, and −0.45, respectively. The normality distribution of hardness values in each allele group was verified by the Shapiro–Wilk test. The test was positive only for the Pinb-D1b group in modern wheats. The Kruskal–Wallis test was performed to compare variable distributions between Pinb allele groups with a breakdown into modern wheat and landraces, and hardness, and excluding single observations (Pinb-D1c and Pinb-D1g in landraces). The results proved to be statistically significant (p-value < 2.2 × 10−16). In the next step, the Dunn test was performed in order to establish which groups were different from each other. The SKCS values of modern wheats and landraces with Pinb-D1b, Pinb-D1c, or Pinb-D1d were significantly different compared to genotypes with wild-type Pinb allele (Table 2). There were no significant differences between genotypes carrying the Pinb-D1b, Pinb-D1c, or Pinb-D1d alleles (Table 2).

4. Discussion

We analyzed grain hardness within a collection of various wheat cultivars, breeding lines currently applied in wheat breeding programs and old landraces collected during multiple expeditions between 1960 and 1976 in Poland. Cultivars were tested both at the molecular level (by determining Pin gene allelic status) and at the phenotype level (by measuring wheat kernel SKCS hardness index).
Surprisingly, no allelic variability within the Pina gene was observed, although mutant alleles of this gene were detected in many other European cultivars [15,24,42]. All tested cultivars, breeding lines, and landraces had the wild-type Pina-D1a allele. This allele, originally described by Gautier et al. [5], was proven to be the most common Pina allele among wheat genotypes [10,12]. The second Pina allele found in European cultivars was Pina-D1b, discovered by Giroux and Morris [9] which is extremely rare. Other Pina alleles mainly occurred in landraces from China or India [17,18]. The lack of Pina variability in the European modern wheat cultivars and breeding lines tested here indicates negative mutant gene selection during the breeding process.
More common allelic variability was observed for the Pinb gene. The Pinb-D1b allele, detected in 57% of tested modern cultivars and breeding lines, was the most frequent. The remaining Pinb alleles were the wild-type Pinb-D1a and the mutant Pinb-D1c and Pinb-D1d. They were represented by a similar variant number (11–16%). Similar distribution of the same Pinb alleles was observed previously in European wheat populations [15,24,43]. As for the landraces, the Pinb allele distributions were different. The main, wild-type Pinb-D1a allele was found in over 70% of tested landraces. The next most frequent was Pinb-D1b, detected in 26% of the tested landraces. Pinb-D1c was found only in one variety, and none contained the Pinb-D1d allele. It was surprising for us to find a landrace with the Pinb-D1g allele. To date, this allele has not been observed in European wheat cultivars. It was reported for the first time in historical varieties from North America [16]. We did not find such high variability among Pina and Pinb alleles, which were observed in hard wheat cultivars originating from other continents. For example, among those originating from Asia, seven various Pinb alleles, including four newly described, such as Pinb-D1ad, Pinb-D1ae, Pinb-D1af and Pinb-D1ag, have been found [17]. In another study, nine Pinb alleles were described [44]. Therefore, our observation confirms that these four alleles found among modern wheats cultivated in Poland dominate in modern European varieties. This is probably caused by the narrow genetic variability of breeding lines as well as being the result of breeding pressure.
The most commonly used methods of determination of grain hardness include the particle size index (PSI), near-infrared reflectance (NIR) with a properly prepared library, and the single kernel characterization system (SKCS). We decided to use SKCS due to the high reliability and accuracy of this method, and the possibility of analyzing 300 individual grains in a single run. Based on the SKCS value, wheats can be classified into different categories. The first classification was proposed by Morris et al. [16], who distinguished four groups according to the SKCS hardness index: (I) from 0 to 33, (II) from 34 to 46, (III) from 47 to 59, and (IV) above 60. Slightly different categories were proposed by Corona et al. [41] who divided wheat into three main groups: soft (SKCS 15–40), medium hard (41–70), and hard (71–95). The second wheat classification system, used also by us, is the most common. However, in many regions, other classes are used as well, depending on local needs or regulations. For example, Sharma et al. [45] proposed five classes for a more precise classification of soft and hard cultivars from India. They included very soft (SKCS 34 or less), soft (35–54), medium hard (SKCS 55–74), and hard (90 or more).
It was noted that some of the tested cultivars were classified into groups contrary to ourexpectations based on their allelic status. For example, in case of the Pinb-D1a, a group of six cultivars and breeding lines had the SKCS indexes ranging from 31 to 56, with a mean value of 46, which was higher than usual for wild-type alleles. In previous reports, cultivars with both wild-type Pina and Pinb alleles had low grain hardness values ranging from 7 to 40, with a mean value of 27 [41], or from 15 to 29, with a mean value of 24.5 [16]. Moreover, some cultivars, breeding lines, and landraces with Pinb-D1b or Pinb-D1d mutant alleles, which were expected to have higher grain hardness values, were classified as soft, with a hardness index of 15 to 40. Similarly, one cultivar and one breeding line carrying Pinb-D1d were classified as soft.
Particularly interesting is the landrace with the accession number I31930, (Pinb-D1b) of which the hardness index is 17.43 (±18.59). Such a low value has never been reported in the case of the Pinb-D1b allele. In five breeding lines and seven landraces, the standard deviation of SKCS hardness was above 20, which might suggest that the lines represented a mixture of hard and soft genotypes. This is possible because landraces have been sourced from farmland or local markets and may be contaminated by seeds from other genotypes. In turn, the breeding lines may not yet be genetically homogenous.
Statistical analyses do not show differences in grain hardness between genotypes with the mutant alleles of Pinb in both groups. Statistically significant differences occurred only between genotypes carrying the wild-type allele Pinb-D1a and genotypes with mutated alleles, which proved that the mutation causing the amino acid sequence change leads to increased grain hardness.
Differences in grain hardness between the same alleles may be caused by other biochemical components of the grain [11,43,46]. Therefore, we also investigated the total protein content, starch, and wet gluten measured by the NIR method. Unfortunately, no correlation between any of these factors was found. This contradicts the results reported by Salmanowicz et al. [47], who found a strong correlation between grain hardness and protein content. It should be noted however, that in the aforementioned work the grain hardness was measured using the NIR and the PSI methods, whereas in the current study the SKCS was applied. As mentioned earlier, the method of measurement influences the results because each method defines the grain hardness in a slightly different way [46,48,49]. The influence of protein content on grain hardness was also reported in other papers [11,50,51]. Moreover grain hardness might be affected by other grain components such as pentosans, primarily arabinoxylans and arabinogalactans [11,52], and gliadin compositions [50], as well as gluten proteins [53]. Other reasons for differences in grain hardness may be the result of differences in genetic background other than variability in Pin genes.
Nirmal et al. [34] described significant differences between the expression of Pin genes in hard and soft wheat and concluded that higher expression at earlier stages of grain development causes a softer grain, but this does not fully explain entirety of the variability between the genotypes studied. On the other hand, increased expression of Pin genes in rice, durum wheat, or in common wheat caused a significant decrease in grain hardness [27,54,55]. The opposed relationship was shown by the silencing of the Pin genes, which resulted in increased grain hardness [25,26]. Another known genetic factor, which might affect grain hardness is nucleotide polymorphisms of the transcription activator SPA (storage protein activator) [56]. In addition many QTLs (quantitative trait locus) have been described that affect the grain hardness. The QTLs related to grain hardness are located on chromosomes 1A, 1B, 2B, 2D, 3B, 4A, 4B, 5AL, 5BL, 5D, 6A, 6BL, 7A, and 7B [57,58,59,60,61,62]. It has also been proven that grain hardness is affected by environmental factors [49,50,63]. However, in our investigations all cultivars, breeding lines and landraces grew under the same optimal conditions, so we assume that this factor was eliminated. Another factor, which was proven to influence kernel texture is the grain moisture [11,51]. However, in our research, this parameter was almost identical in all tested lines, with the standard deviation values below 1%.

5. Conclusions

In conclusion, no new Pin alleles have been discovered however, the Pinb-D1g allele, unique to European wheat, has been found among landraces from Poland. In addition, we found many interesting objects whose grain hardness seemed to be determined not only by Pina and Pinb alleles but also by other genetic or biochemical factors. In further studies, we aim to unravel the genetic background of these inconsistences.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4395/10/8/1075/s1, Table S1: Means for kernel hardness, weight, moisture and diameter of modern wheat. Table S2: Means for kernel hardness, weight, moisture and diameter of Polish landraces. Table S3: Means for protein, starch, and wet gluten contents of Polish landraces. Figure S1: Chromatogram fragments of five alleles found in tested wheat genotypes.

Author Contributions

Conceptualization, A.N.-O. and M.P.; methodology, M.P. and S.G.; formal analysis, M.P. and S.G.; investigation, M.P., S.G.; and M.K.; Data curation, M.P. and M.K.; resources, M.P.; writing—original draft preparation, M.P.; writing—review and editing, A.N.-O., S.G., M.P., and W.O.; visualization, M.P. and M.K.; supervision, A.N.-O.; project administration A.N.-O.; funding acquisition A.N.-O., All authors have read and agreed to the published version of the manuscript.

Funding

Ministry of Agriculture and Rural Development Grant PW (2015–2020), Task 2.1.

Acknowledgments

This research was supported by the Ministry of Agriculture and Rural Development Grant PW (2015–2020), Task 2.1. We thank Agnieszka Głowacka, Izabela Skuza and Małgorzata Wojciechowska for their technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. FAO, S. Food and Agriculture Organization of the United Nations. Retrieved 2010, 3, 2012. [Google Scholar]
  2. Igrejas, G.; Gaborit, T.; Oury, F.X.; Chiron, H.; Marion, D.; Branlard, G. Genetic and environmental effects on puroindoline-a and puroindoline-b content and their relationship to technological properties in French bread wheats. J. Cereal Sci. 2001, 34, 37–47. [Google Scholar] [CrossRef]
  3. Ribeiro, M.; Rodriguez-Quijano, M.; Giraldo, P.; Pinto, L.; Vazquez, J.F.; Carrillo, J.M.; Igrejas, G. Effect of allelic variation at glutenin and puroindoline loci on bread-making quality: Favorable combinations occur in less toxic varieties of wheat for celiac patients. Eur. Food Res. Technol. 2017, 243, 743–752. [Google Scholar] [CrossRef]
  4. Morris, C.F. Puroindolines: The molecular genetic basis of wheat grain hardness. Plant Mol. Biol. 2002, 48, 633–647. [Google Scholar] [CrossRef] [PubMed]
  5. Gautier, M.F.; Aleman, M.E.; Guirao, A.; Marion, D.; Joudrier, P. Triticum-aestivum puroindolines, 2 basic cystine-rich seed rroteins-cDNA sequence-analysis and developmental gene-expression. Plant Mol. Biol 1994, 25, 43–57. [Google Scholar] [CrossRef] [PubMed]
  6. Chantret, N.; Salse, J.; Sabot, F.; Rahman, S.; Bellec, A.; Laubin, B.; Dubois, I.; Dossat, C.; Sourdille, P.; Joudrier, P.; et al. Molecular basis of evolutionary events that shaped the hardness locus in diploid and polyploid wheat species (Triticum and aegilops). Plant Cell 2005, 17, 1033–1045. [Google Scholar] [CrossRef] [Green Version]
  7. Blochet, J.E.; Chevalier, C.; Forest, E.; Pebaypeyroula, E.; Gautier, M.F.; Joudrier, P.; Pezolet, M.; Marion, D. Complete amino-acid-sequence of puroindoline, a new basic and cystine-rich protein with a unique tryptophan-rich domain, isolated from wheat endosperm by triton X-114 phase partitioning. FEBS Lett. 1993, 329, 336–340. [Google Scholar] [CrossRef] [Green Version]
  8. Day, L.; Bhandari, D.G.; Greenwell, P.; Leonard, S.A.; Schofield, J.D. Characterization of wheat puroindoline proteins. FEBS J. 2006, 273, 5358–5373. [Google Scholar] [CrossRef]
  9. Giroux, M.J.; Morris, C.F. Wheat grain hardness results from highly conserved mutations in the friabilin components puroindoline a and b. Proc. Natl. Acad. Sci. USA 1998, 95, 6262–6266. [Google Scholar] [CrossRef] [Green Version]
  10. Iftikhar, A.; Ali, I. Kernel softness in wheat is determined by starch granule bound Puroindoline proteins. J. Plant Biochem. Biotechnol. 2017, 26, 247–262. [Google Scholar] [CrossRef]
  11. Bhave, M.; Morris, C.F. Molecular genetics of puroindolines and related genes: Regulation of expression, membrane binding properties and applications. Plant Mol. Biol. 2008, 66, 221–231. [Google Scholar] [CrossRef] [PubMed]
  12. Nadolska-Orczyk, A.; Gasparis, S.; Orczyk, W. The determinants of grain texture in cereals. J. Appl. Genet. 2009, 50, 185–197. [Google Scholar] [CrossRef]
  13. Morris, C.F.; Bhave, M. Reconciliation of D-genome puroindoline allele designations with current DNA sequence data. J. Cereal Sci. 2008, 48, 277–287. [Google Scholar] [CrossRef]
  14. Giroux, M.J.; Morris, C.F. A glycine to serine change in puroindoline b is associated with wheat grain hardness and low levels of starch-surface friabilin. Theor. Appl. Genet. 1997, 95, 857–864. [Google Scholar] [CrossRef]
  15. Lillemo, M.; Morris, C.F. A leucine to proline mutation in puroindoline b is frequently present in hard wheats from Northern Europe. Theor. Appl. Genet. 2000, 100, 1100–1107. [Google Scholar] [CrossRef]
  16. Morris, C.F.; Lillemo, M.; Simeone, M.C.; Giroux, M.J.; Babb, S.L.; Kidwell, K.K. Prevalence of puroindoline grain hardness genotypes among historically significant North American spring and winter wheats. Crop Sci. 2001, 41, 218–228. [Google Scholar] [CrossRef]
  17. Kumar, R.; Arora, S.; Singh, K.; Garg, M. Puroindoline allelic diversity in Indian wheat germplasm and identification of new allelic variants. Breed. Sci 2015, 65, 319–326. [Google Scholar] [CrossRef] [Green Version]
  18. Chang, C.; Zhang, H.P.; Xu, J.; Li, W.H.; Liu, G.T.; You, M.S.; Li, B.Y. Identification of allelic variations of puroindoline genes controlling grain hardness in wheat using a modified denaturing PAGE. Euphytica 2006, 152, 225–234. [Google Scholar] [CrossRef]
  19. Ram, S.; Jain, N.; Shoran, J.; Singh, R. New frame shift mutation in puroindoline b in Indian wheat cultivars Hyb65 and NI5439. J. Plant. Biochem. Biotechnol. 2005, 14, 45–48. [Google Scholar] [CrossRef]
  20. Ikeda, T.M.; Ohnishi, N.; Nagamine, T.; Oda, S.; Hisatomi, T.; Yano, H. Identification of new puroindoline genotypes and their relationship to flour texture among wheat cultivars. J. Cereal Sci. 2005, 41, 1–6. [Google Scholar] [CrossRef]
  21. Tranquilli, G.; Heaton, J.; Chicaiza, O.; Dubcovsky, J. Substitutions and deletions of genes related to grain hardness in wheat and their effect on grain texture. Crop Sci. 2002, 42, 1812–1817. [Google Scholar] [CrossRef] [Green Version]
  22. Pickering, P.A.; Bhave, M. Comprehensive analysis of Australian hard wheat cultivars shows limited puroindoline allele diversity. Plant Sci. 2007, 172, 371–379. [Google Scholar] [CrossRef]
  23. Qamar, Z.U.; Bansal, U.K.; Dong, C.M.; Alfred, R.L.; Bhave, M.; Bariana, H.S. Detection of puroindoline (Pina-D1 and Pinb-D1) allelic variation in wheat landraces. J. Cereal Sci. 2014, 60, 610–616. [Google Scholar] [CrossRef]
  24. Ayala, M.; Guzman, C.; Pena, R.J.; Alvarez, J.B. Genetic diversity and molecular characterization of puroindoline genes (Pina-D1 and Pinb-D1) in bread wheat landraces from Andalusia (Southern Spain). J. Cereal Sci. 2016, 71, 61–65. [Google Scholar] [CrossRef]
  25. Gasparis, S.; Orczyk, W.; Zalewski, W.; Nadolska-Orczyk, A. The RNA-mediated silencing of one of the Pin genes in allohexaploid wheat simultaneously decreases the expression of the other, and increases grain hardness. J. Exp. Bot. 2011, 62, 4025–4036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gasparis, S.; Kała, M.; Przyborowski, M.; Orczyk, W.; Nadolska-Orczyk, A. Artificial MicroRNA-based specific gene silencing of grain hardness genes in polyploid cereals appeared to be not stable over transgenic plant generations. Front. Plant Sci. 2017, 7, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Li, Y.; Mao, X.; Wang, Q.; Zhang, J.; Li, X.; Ma, F.; Sun, F.; Chang, J.; Chen, M.; Wang, Y. Overexpression of Puroindoline a gene in transgenic durum wheat (Triticum turgidum ssp. durum) leads to a medium–hard kernel texture. Mol. Breed. 2014, 33, 545–554. [Google Scholar] [CrossRef]
  28. Alfred, R.L.; Palombo, E.A.; Panozzo, J.F.; Bhave, M. The co-operative interaction of puroindolines in wheat grain texture may involve the hydrophobic domain. J. Cereal Sci. 2014, 60, 323–330. [Google Scholar] [CrossRef]
  29. Clifton, L.A.; Green, R.J.; Frazier, R.A. Puroindoline-b mutations control the lipid binding interactions in mixed puroindoline-a: Puroindoline-b systems. Biochemistry 2007, 46, 13929–13937. [Google Scholar] [CrossRef]
  30. Chen, F.; Beecher, B.S.; Morris, C.F. Physical mapping and a new variant of Puroindoline b-2 genes in wheat. Theor. Appl. Genet. 2010, 120, 745–751. [Google Scholar] [CrossRef]
  31. Giroux, M.J.; Kim, K.H.; Hogg, A.C.; Martin, J.M.; Beecher, B. The puroindoline b-2 variants are expressed at low levels relative to the puroindoline D1 genes in wheat seeds. Crop Sci. 2013, 53, 833–841. [Google Scholar] [CrossRef]
  32. Ramalingam, A.; Palombo, E.A.; Bhave, M. The Pinb-2 genes in wheat comprise a multigene family with great sequence diversity and important variants. J. Cereal Sci. 2012, 56, 171–180. [Google Scholar] [CrossRef]
  33. Wilkinson, M.; Wan, Y.F.; Tosi, P.; Leverington, M.; Snape, J.; Mitchell, R.A.C.; Shewry, P.R. Identification and genetic mapping of variant forms of puroindoline b expressed in developing wheat grain. J. Cereal Sci. 2008, 48, 722–728. [Google Scholar] [CrossRef]
  34. Nirmal, R.C.; Furtado, A.; Wrigley, C.; Henry, R.J. Influence of gene expression on hardness in wheat. PLoS ONE 2016, 11, e0164746. [Google Scholar] [CrossRef] [Green Version]
  35. Murray, M.G.; Thompson, W.F. Rapid isolation of high molecular-weight plant DNA. Nucleic Acids Res. 1980, 8, 4321–4325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Lillemo, M.; Chen, F.; Xia, X.C.; William, M.; Pena, R.J.; Trethowan, R.; He, Z.H. Puroindoline grain hardness alleles in CIMMYT bread wheat germplasm. J. Cereal Sci. 2006, 44, 86–92. [Google Scholar] [CrossRef]
  37. Thiel, T.; Kota, R.; Grosse, I.; Stein, N.; Graner, A. SNP2CAPS: A SNP and INDEL analysis tool for CAPS marker development. Nucleic Acids Res. 2004, 32, e5. [Google Scholar] [CrossRef]
  38. AACC Approved Methods of Analysis, 11th ed.; Method 55-31.01. Single-kernel characterization system for wheat kernel texture; Cereals & Grains Association: St. Paul, MN, USA, 1999. [CrossRef]
  39. Ogle, D.H. FSA: Fisheries stock analysis. R Package Vers. 0.8 2017, 17, 636. [Google Scholar]
  40. Wickham, H. ggplot2: Elegant Graphics for Data Analysis; Springer: New York, NY, USA, 2016. [Google Scholar]
  41. Corona, V.; Gazza, L.; Boggini, G.; Pogna, N.E. Variation in friabilin composition as determined by A-PAGE fractionation and PCR amplification, and its relationship to grain hardness in bread wheat. J. Cereal Sci. 2001, 34, 243–250. [Google Scholar] [CrossRef]
  42. Mohler, V.; Schmolke, M.; Paladey, E.; Seling, S.; Hartl, L. Association analysis of Puroindoline-D1 and Puroindoline b-2 loci with 13 quality traits in European winter wheat (Triticum aestivum L.). J. Cereal Sci. 2012, 56, 623–628. [Google Scholar] [CrossRef]
  43. Huang, X.-Q.; Röder, M.S. Development of SNP assays for genotyping the puroindoline b gene for grain hardness in wheat using pyrosequencing. J. Agric. Food Chem. 2005, 53, 2070–2075. [Google Scholar] [CrossRef] [PubMed]
  44. Li, X.; Li, Y.; Zhang, M.; Yu, X.; Hu, R.; Chang, J.; Yang, G.; Wang, Y.; He, G. Diversity of Puroindoline genes and their association with kernel hardness in Chinese wheat cultivars and landraces. Mol. Breed. 2019, 39, 61. [Google Scholar] [CrossRef]
  45. Sharma, R.; Rawat, A.; Misra, B.; Nagarajan, S. Distribution of grain hardness in Indian wheat varieties and landraces. Wheat Inf. Serv. 2012, 114, 1–8. [Google Scholar]
  46. Dobraszczyk, B.; Whitworth, M.; Vincent, J.; Khan, A. Single kernel wheat hardness and fracture properties in relation to density and the modelling of fracture in wheat endosperm. J. Cereal Sci. 2002, 35, 245–263. [Google Scholar] [CrossRef]
  47. Salmanowicz, B.P.; Adamski, T.; Surma, M.; Kaczmarek, Z.; Karolina, K.; Kuczyńska, A.; Banaszak, Z.; Ługowska, B.; Majcher, M.; Obuchowski, W. The relationship between grain hardness, dough mixing parameters and bread-making quality in winter wheat. Int. J. Mol. Sci. 2012, 13, 4186–4201. [Google Scholar] [CrossRef] [Green Version]
  48. Morris, C.F.; Massa, A.N. Puroindoline genotype of the US national institute of standards & technology reference material 8441, wheat hardness. Cereal Chem. 2003, 80, 674–678. [Google Scholar]
  49. Oury, F.-X.; Lasme, P.; Michelet, C.; Rousset, M.; Abecassis, J.; Lullien-Pellerin, V. Relationships between wheat grain physical characteristics studied through near-isogenic lines with distinct puroindoline-b allele. Theor. Appl. Genet. 2015, 128, 913–929. [Google Scholar] [CrossRef]
  50. Huebner, F.R.; Gaines, C.S. Relation between Wheat Kernel Hardness, Environment, and Gliadin Composition. Cereal Chem. 1992, 69, 148–151. [Google Scholar]
  51. Turnbull, K.M.; Rahman, S. Endosperm texture in wheat. J. Cereal Sci. 2002, 36, 327–337. [Google Scholar] [CrossRef]
  52. Lillemo, M.; Ringlund, K. Impact of puroindoline b alleles on the genetic variation for hardness in soft x hard wheat crosses. Plant Breed. 2002, 121, 210–217. [Google Scholar] [CrossRef]
  53. Quayson, E.T.; Marti, A.; Morris, C.F.; Marengo, M.; Bonomi, F.; Seetharaman, K.; Iametti, S. Structural consequences of the interaction of puroindolines with gluten proteins. Food Chem. 2018, 253, 255–261. [Google Scholar] [CrossRef] [PubMed]
  54. See, D.; Giroux, M.; Gill, B. Effect of multiple copies of puroindoline genes on grain softness. Crop Sci. 2004, 44, 1248–1253. [Google Scholar] [CrossRef] [Green Version]
  55. Krishnamurthy, K.; Giroux, M.J. Expression of wheat puroindoline genes in transgenic rice enhances grain softness. Nat. Biotechnol. 2001, 19, 162–166. [Google Scholar] [CrossRef] [PubMed]
  56. Ravel, C.; Martre, P.; Romeuf, I.; Dardevet, M.; El-Malki, R.; Bordes, J.; Duchateau, N.; Brunel, D.; Balfourier, F.; Charmet, G. Nucleotide polymorphism in the wheat transcriptional activator SPA influences its pattern of expression and has pleiotropic effects on grain protein composition, dough viscoelasticity, and grain hardness. Plant Physiol. 2009, 151, 2133–2144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Boehm, J.D.; Ibba, M.I.; Kiszonas, A.M.; See, D.R.; Skinner, D.Z.; Morris, C.F. Genetic analysis of kernel texture (grain hardness) in a hard red spring wheat (Triticum aestivum L.) bi-parental population. J. Cereal Sci. 2018, 79, 57–65. [Google Scholar] [CrossRef]
  58. Breseghello, F.; Finney, P.L.; Gaines, C.; Andrews, L.; Tanaka, J.; Penner, G.; Sorrells, M.E. Genetic loci related to kernel quality differences between a soft and a hard wheat cultivar. Crop Sci. 2005, 45, 1685–1695. [Google Scholar] [CrossRef]
  59. Campbell, K.G.; Bergman, C.J.; Gualberto, D.G.; Anderson, J.A.; Giroux, M.J.; Hareland, G.; Fulcher, R.G.; Sorrells, M.E.; Finney, P.L. Quantitative trait loci associated with kernel traits in a soft x hard wheat cross. Crop Sci. 1999, 39, 1184–1195. [Google Scholar] [CrossRef]
  60. Carter, A.H.; Garland-Campbell, K.; Morris, C.F.; Kidwell, K.K. Chromosomes 3B and 4D are associated with several milling and baking quality traits in a soft white spring wheat (Triticum aestivum L.) population. Theor. Appl. Genet. 2012, 124, 1079–1096. [Google Scholar] [CrossRef]
  61. Osborne, B.G.; Turnbull, K.M.; Anderssen, R.S.; Rahman, S.; Sharp, P.J.; Appels, R. The hardness locus in Australian wheat lines. Aust. J. Agric. Res. 2001, 52, 1275–1286. [Google Scholar] [CrossRef]
  62. Igrejas, G.; Leroy, P.; Charmet, G.; Gaborit, T.; Marion, D.; Branlard, G. Mapping QTLs for grain hardness and puroindoline content in wheat (Triticum aestivum L.). Theor. Appl. Genet. 2002, 106, 19–27. [Google Scholar] [CrossRef]
  63. Gazza, L.; Taddei, F.; Corbellini, M.; Cacciatori, P.; Pogna, N. Genetic and environmental factors affecting grain texture in common wheat. J. Cereal Sci. 2008, 47, 52–58. [Google Scholar] [CrossRef]
Figure 1. (a) Alignment of the nucleotide and (b) amino acid sequences of Pinb alleles detected in the tested cultivars.
Figure 1. (a) Alignment of the nucleotide and (b) amino acid sequences of Pinb alleles detected in the tested cultivars.
Agronomy 10 01075 g001
Figure 2. Examples of cleaved amplified polymorphic sequence (CAPS) analysis for detection of different Pinb alleles. M—molecular weight marker; A—Pinb-D1a; B—Pinb-D1b; C—Pinb-D1c; D—Pinb-D1d; P—undigested 447 bp PCR product.
Figure 2. Examples of cleaved amplified polymorphic sequence (CAPS) analysis for detection of different Pinb alleles. M—molecular weight marker; A—Pinb-D1a; B—Pinb-D1b; C—Pinb-D1c; D—Pinb-D1d; P—undigested 447 bp PCR product.
Agronomy 10 01075 g002
Figure 3. Hardness index ranges for individual Pinb alleles divided into modern wheats and landraces. The widths of the figures indicates the frequency of occurrence of a given hardness index.
Figure 3. Hardness index ranges for individual Pinb alleles divided into modern wheats and landraces. The widths of the figures indicates the frequency of occurrence of a given hardness index.
Agronomy 10 01075 g003
Table 1. The Pinb allele frequency in the tested genotypes and mean hardness index (HI) in each allele group.
Table 1. The Pinb allele frequency in the tested genotypes and mean hardness index (HI) in each allele group.
Pinb AlleleNo. Cultivars (%)Mean HI (SD)
Cultivars and Breeding Lines
Pinb-D1a19 (16%)31.31 (±19.50)
Pinb-D1b67 (57%)59.36 (±11.73)
Pinb-D1c14 (12%)63.70 (±10.43)
Pinb-D1d18 (15%)59.08 (±13.80)
Landraces
Pinb-D1a57 (71%)23.34 (±6.09)
Pinb-D1b21 (26%)49.23 (±10.97)
Pinb-D1c1 (1%)65.80
Pinb-D1g1 (1%)60.79
Table 2. Dunn test for multiple comparisons of SKCS hardness in six allele groups. (L)—landraces, (M)—modern cultivars and breeding lines.
Table 2. Dunn test for multiple comparisons of SKCS hardness in six allele groups. (L)—landraces, (M)—modern cultivars and breeding lines.
Tested PairsZP.unadjP.adjDifference
Pinb-D1a (L)Pinb-D1b (L)−4.9310.0000.000True
Pinb-D1a (L)Pinb-D1b (M)−9.7620.0000.000True
Pinb-D1a (L)Pinb-D1c (M)−6.4480.0000.000True
Pinb-D1a (L)Pinb-D1d (M)−6.8730.0000.000True
Pinb-D1a (M)Pinb-D1a (L)0.1580.8751.000False
Pinb-D1a (M)Pinb-D1b (L)−3.5880.0000.005True
Pinb-D1a (M)Pinb-D1b (M)−6.0250.0000.000True
Pinb-D1a (M)Pinb-D1d (M)−5.0960.0000.000True
Pinb-D1a (M)Pinb-D1c (M)−5.0530.0000.000True
Pinb-D1b (L)Pinb-D1c (M)−1.9260.0540.812False
Pinb-D1b (L)Pinb-D1d (M)−1.6880.0911.000False
Pinb-D1b (M)Pinb-D1b (L)2.0420.0410.617False
Pinb-D1b (M)Pinb-D1c (M)−0.5160.6061.000False
Pinb-D1b (M)Pinb-D1d (M)−0.0580.9541.000False
Pinb-D1c (M)Pinb-D1d (M)0.3940.6941.000False

Share and Cite

MDPI and ACS Style

Przyborowski, M.; Gasparis, S.; Kała, M.; Orczyk, W.; Nadolska-Orczyk, A. The Variability of Puroindoline-Encoding Alleles and Their Influence on Grain Hardness in Modern Wheat Cultivars Cultivated in Poland, Breeding Lines and Polish Old Landraces (Triticum aestivum L.). Agronomy 2020, 10, 1075. https://doi.org/10.3390/agronomy10081075

AMA Style

Przyborowski M, Gasparis S, Kała M, Orczyk W, Nadolska-Orczyk A. The Variability of Puroindoline-Encoding Alleles and Their Influence on Grain Hardness in Modern Wheat Cultivars Cultivated in Poland, Breeding Lines and Polish Old Landraces (Triticum aestivum L.). Agronomy. 2020; 10(8):1075. https://doi.org/10.3390/agronomy10081075

Chicago/Turabian Style

Przyborowski, Mateusz, Sebastian Gasparis, Maciej Kała, Wacław Orczyk, and Anna Nadolska-Orczyk. 2020. "The Variability of Puroindoline-Encoding Alleles and Their Influence on Grain Hardness in Modern Wheat Cultivars Cultivated in Poland, Breeding Lines and Polish Old Landraces (Triticum aestivum L.)" Agronomy 10, no. 8: 1075. https://doi.org/10.3390/agronomy10081075

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