Next Article in Journal
Generation of Wheat Near-Isogenic Lines Overexpressing 1Bx7 Glutenin with Increased Protein Contents and SDS-Sedimentation Values
Previous Article in Journal
De Novo Transcriptome Assembly and Comparative Analysis of Differentially Expressed Genes Involved in Cold Acclimation and Freezing Tolerance of the Arctic Moss Aulacomnium turgidum (Wahlenb.) Schwaegr
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Population Structure and Association Mapping for Agronomical and Biochemical Traits of a Large Spanish Apple Germplasm

by
Pierre Mignard
,
Carolina Font i Forcada
,
Rosa Giménez
and
María Ángeles Moreno
*
Department of Pomology, Estación Experimental de Aula Dei, Consejo Superior de Investigaciones Científicas (EEAD-CSIC), P.O. Box 13034, 50080 Zaragoza, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2023, 12(6), 1249; https://doi.org/10.3390/plants12061249
Submission received: 16 February 2023 / Revised: 4 March 2023 / Accepted: 6 March 2023 / Published: 9 March 2023
(This article belongs to the Special Issue Genetic and Metabolic Analysis to Improve Fruit Quality)

Abstract

:
A basic knowledge of linkage disequilibrium and population structure is necessary in order to determine the genetic control and identify significant associations with agronomical and phytochemical compounds in apple (Malus × domestica Borkh). In this study, 186 apple accessions (Pop1), representing both Spanish native accessions (94) and non-Spanish cultivars (92) from the EEAD-CSIC apple core collection, were assessed using 23 SSRs markers. Four populations were considered: Pop1, Pop2, Pop3, and Pop4. The initial Pop1 was divided into 150 diploid (Pop2) and 36 triploid accessions (Pop3), while for the inter-chromosomal linkage disequilibrium and the association mapping analysis, 118 phenotype diploid accessions were considered Pop4. Thus, the average number of alleles per locus and observed heterozygosity for the overall sample set (Pop1) were 15.65 and 0.75, respectively. The population structure analysis identified two subpopulations in the diploid accessions (Pop2 and Pop4) and four in the triploids (Pop3). Regarding the Pop4, the population structure with K = 2 subpopulations segregation was in agreement with the UPGMA cluster analysis according to the genetic pairwise distances. Moreover, the accessions seemed to be segregated by their origin (Spanish/non-Spanish) in the clustering analysis. One of the two subpopulations encountered was quite-exclusively formed by non-Spanish accessions (30 out of 33). Furthermore, agronomical and basic fruit quality parameters, antioxidant traits, individual sugars, and organic acids were assessed for the association mapping analysis. A high level of biodiversity was exhibited in the phenotypic characterization of Pop4, and a total of 126 significant associations were found between the 23 SSR markers and the 21 phenotypic traits evaluated. This study also identified many new marker-locus trait associations for the first time, such as in the antioxidant traits or in sugars and organic acids, which may be useful for predictions and for a better understanding of the apple genome.

1. Introduction

The modern cultivated apple (Malus × domestica Borkh) is the most significant and ancient fruit crop of the Rosaceae in the world as well as in Spain [1]. In 2021, more than 93 M tonnes of apples were produced around the world [2]. In 2021, an average of 617 thousand tonnes was produced in Spain [3]. However, more than 516 thousand tonnes of the national production were dessert apples for fresh consumption and juices, while cider apple production reached only 100 thousand tonnes. Nevertheless, apple production around the world is based on a reduced number of modern bred varieties [1,4,5]. In fact, apple production is dominated by some well-adapted modern cultivars (‘Fuji’, ‘Gala’, ‘Golden’, ‘Granny Smith’, and ‘Delicious’), many of which are genetically linked [6,7]. In addition, the ‘Golden Delicious’ cultivar represented more than 46% of the Spanish apple production for fresh consumption with more than 240 thousand tonnes [3]. As a consequence, many of the traditional and/or locally well-adapted cultivars have been considered outdated, and a dramatic loss of genetic diversity has been noticed in the apple gene pool in the last decades [8,9,10,11,12,13].
All commercial apple cultivars have a basic chromosome number of 17 and are considered to have evolved after an autopolyploidization of a Gillenia-like taxon followed by diploidization [14,15]. Most domesticated apple cultivars are diploid (2n = 2x = 34), even though triploid (2n = 3x = 51) and tetraploid (2n = 4x = 68) cultivars can also be found. The modern apple should be the result of a long evolutionary process spanning hundreds of years, with several species contributing to the domesticated apple gene pool encountered nowadays [16].
Despite the high genetic variability of the apple gene pool, breeding programs over the years have been mainly based on a few organoleptic traits, aesthetic criteria, and resistance to specific diseases [10], resulting in a small number of cultivars and a loss in diversity. Moreover, the breeding programs have limited and reduced the use of the total number of cultivars, such as ‘Cox’s Orange Pippin’, ‘Golden Delicious’, ‘Jonathan’, ‘Red Delicious’, and ‘McIntosh’, although more than 10,000 different apple accessions have been described globally [6,10,17]. Indeed, a great number of new cultivars were obtained from breeding programs during the twentieth century from a reduced number of progenitors and, consequently, shared a high degree of parentage [6,17,18]. For this reason, germplasm evaluation, characterization, and conservation of genetic diversity are crucial for the management of genetic properties [19]. Therefore, there is an increased need for studying and collecting old local apple accessions that could provide a better knowledge of the history of the crop and could be used in future breeding programs aiming to select better adapted cultivars to climate conditions in the current context of global climate change [5,19,20]. In fact, apple allelic diversity should be used to address existing and future biotic and abiotic issues with respect to production [8]. Furthermore, in the last few decades, interest in the genetic and molecular characterization of apples has increased. The diversity found at the genetic level between apple accessions reflects a combination of historical selections and adaptive processes, resulting in extensive genetic variation but a limited population structure [6].
The microsatellite markers or Simple Sequence Repeats (SSRs) have been in apple and other fruit crops, in general, one of the most used markers for molecular characterization together with the Single Nucleotide Polymorphism markers (SNPs). The genome sequence of the domesticated apple has been released and accounts for approximately 750 Mb per haploid [15]. The SSRs analysis is a good technique within crop species, due to their abundance, codominance, high polymorphism, easy use with the PCR conditions, and relatively low cost [21,22]. Indeed, many studies based on these molecular markers have assessed the diversity of Malus accessions [1,8,9,11,14,23,24]. Gross et al. [14] demonstrated that nine SSR loci were sufficient to determine potential duplicate and study differences among Malus accessions. In the present work, 23 microsatellites markers were used for the molecular characterization of a large set of accessions from the northeastern part of Spain. The European Cooperative Program for Plant Genetic Resources (ECPGR) has published several lists of recommended markers, including SSRs that span most of the apple genome [8,25] and have been tested on a set of standard Malus accessions. Many of these mapped SSRs have been associated with quantitative trait loci (QTL) with a great interest in agronomical, morphological, and/or organoleptic traits and could be used as molecular tools for marker assisted selection (MAS) in future plant breeding programs [10,26].
Apart from the recommended SSR, the present work aims to report the identification of the MdSWEET genes by SSRs markers designed by Zhen et al. [27]. In plants, SWEET transporters function as bidirectional uniporters that mediate the influx and efflux of sugars across cell membranes. SWEETs can be divided into four subgroups [27,28]. Clades I, II, and IV of the SWEETs transport predominantly hexoses, whereas clade III of the SWEETs appears to be sucrose transporters (SUTs) [27]. This study will allow a better understanding of the effect of SWEETs on fruit sugar accumulation, and it will also be helpful for the genetic improvement of fruit sugar accumulation in apple-breeding programs.
Moreover, association mapping, also known as linkage disequilibrium (LD) mapping, relies on the strength of association between genetic markers and phenotype. Therefore, LD mapping is an approach that detects and locates genes relative to an existing map of genetic markers [29,30]. Consequently, this method detects relationships between phenotypic variation and gene polymorphism in existing germplasm and in unrelated individuals. Association mapping has been successfully used to identify genes involved in flowering and ripening traits in apples [12], although the bibliography for organoleptic and fruit quality traits is scarce [31,32,33,34,35].
The present work aims to study the genetic characterization of the Malus × domestica Borkh germplasm located at the Experimental Station of Aula Dei (EEAD-CSIC), Spain, to assess the population structure, inter-chromosomal linkage disequilibrium, and association mapping between the local well-adapted accessions compared with the modern and commercial cultivars. This work provides, thus, molecular tools for genetic improvement of fruit quality in apple breeding programs as well as a better knowledge of the apple genetic resources available through a common set of 23 SSR markers. Our results will highlight the importance of genetic variation in germplasm collections for the effective conservation of biodiversity and phytochemical resources in the domesticated apple crop.

2. Materials and Methods

2.1. Plant Material and Field Trial

This research counted 186 accessions (Malus × domestica Borkh) from the apple germplasm bank (Figure 1; Table 1) established at the Experimental Station of Aula Dei (EEAD-CSIC, Zaragoza, NE Spain: 41°43′42.7″ N, 0°48′44.1″ W). The 186 genotypes (Pop1 for population number 1) were grown under Mediterranean soil conditions, typical of the Central Ebro Valley area. This geographic area is characterized by a semi-arid climate with warm and dry summers, high radiation, and large day–night temperature variation [18]. The accessions assessed in this work (Table 1) consisted of a wide range of geographic origins (94 Spanish and 92 non-Spanish accessions). Indeed, most of the non-Spanish accessions were commercial cultivars, whereas autochthonous commercial cultivars or traditional landraces represented the local-Spanish accessions. The accessions were categorized according to Mignard et al. [5,20]. In the field, each accession had three-tree replications established in a single block design. Trees were trained to a low-density system (6 m × 5 m), and the cultural management was carried out as in a commercial plantation. The orchard was flood irrigated every 12 days during the summer.

2.2. Leaf and Fruit Sampling

For the evaluation of the ploidy level for the 186 accessions (Pop1) of this work, newly expanded leaves were collected from each accession and analyzed as described in Reig et al. [18] by flow cytometry. The accessions were consequently classified into diploids (Pop2: 150 accessions) and triploids (Pop3: 36 accessions). According to the fruit sampling, 30 fruits (10 fruits × tree × rep.) were harvested when fruit firmness (FF) attained a value around 70–80 N or when they displayed the representative peel color of each accession. Accessions were harvested during at least three seasons within the 2014–2018 period, and means for each year and accession were calculated [20]. The maturity date fluctuated from late June to early December.

2.3. Phenotypical Evaluation of Biochemical Traits

2.3.1. Basic Fruit Quality

Soluble solids content (SSC) and titratable acidity (TA) were determined on flesh juice as described by Mignard et al. [20]. SSC results were expressed as °Brix, and TA results were expressed as g of malic acid per liter. The ripening index (RI) was thus calculated based on the SSC/TA ratio.

2.3.2. Antioxidant Compounds, Vitamin C, and Relative Antioxidant Capacity

For the analysis of total phenolic content (TPC), total flavonoid content (TFC), vitamin C (ascorbic acid–AsA), and the relative antioxidant capacity (RAC), a flesh sample composite of 5 g of five peeled fruits per replicate was frozen in liquid nitrogen and kept at −20 °C until further analysis. The biochemical compounds were analyzed using a 96-well microplate as described by Font i Forcada et al. [36].
TPC was determined using the Folin-Ciocalteau method [37] with modifications, and the results were expressed in mg of gallic acid equivalent (GAE) per 100 g of fresh weight (FW). TFC was determined using a colorimetric assay based on the method of Zhishen et al. [38] with minor modifications, and the results were expressed in mg catechin equivalent (CE) per 100 g FW. The RAC was measured using the 2,2-diphenyl-1-picrylhydrazyl (DPPH) method adapted from Brand–Williams et al. [39] with modifications. The results were expressed in mg of Trolox per 100 g FW. Finally, vitamin C–ascorbic acid (AsA) was determined using the method for the spectrophotometric determination of AsA as described by Zaharieva and Abadía [40] with modifications. The results were expressed in mg AsA per 100 g FW.

2.3.3. Individual Sugars and Organic Acids

Individual sugars and organic acid contents were assessed by HPLC, as reported by Mignard et al. [5]. Sugars were analyzed using an Aminex HPX-87C column (300 mm × 7.8 mm, Biorad, Hercules, CA, USA) with a refractive index detector at 35 °C (Waters 2410, Waters Corporation, Milford, MA, USA) and milliQ water at 85 °C as mobile phase. Organic acids were assessed with a Rezex™ ROA-Organic Acid H+ (8%) column (300 mm × 7.8 mm, Phenomenex) with a photodiode array detector (Waters 2489, Waters Corporation, Milford, USA) at 210 nm and a sulphuric acid solution (0.005 N) at room temperature as mobile phase. Individual sugars (glucose, fructose, and sucrose), the sugar-alcohol (sorbitol), and main organic acids (citric, malic, oxalic, quinic, succinic, shikimic, and tartaric acids) concentrations were determined by their characteristic retention times following standards and expressed as g per kg of FW.

2.4. Microsatellite Loci Analysis and Genotyping

For genomic DNA extraction, young leaves were collected from each accession, frozen instantly in liquid nitrogen, and stored at −20 °C until use. DNA was isolated using the NucleoSpin® Plant II kit (Macherey-Nagel, Düren, Germany) according to the manufacturer’s instructions. DNA quality was examined on agarose gels (1%), and DNA concentration was determined by spectrophotometry using a NanoDrop (Waltham, MA, USA). Twenty-six simple sequence repeat (SSRs) markers previously described in Malus were tested in the apple population (Table 2), but three of them exhibited weak or no amplification in our plant material. Eleven of these 23 SSRs, which amplified, were recommended by the ‘European Cooperative Program for Plant Genetic Resources (ECPGR) Malus/Pyrus working group’ with a standard methodology proposed to allow comparisons of the same accessions between different laboratories [41]. In addition to these eleven SSRs, four SSRs have also been assessed in apple studies [6], and other eight SSRs were specially designed to amplify in areas of the genome in relation to the metabolic pathway of sugars: the ‘Sugar Will Eventually is Exported Transporters–SWEET’ genes [27].
Six different multiplexed reactions were used for these 23 SSRs. Forward SSR primers were labeled with 5′-fluorescence dyes, including PET, NED, VIC, and 6-FAM. The polymerase chain reactions (PCR) for the multiplexed PCRs were performed in a final volume of 10 μL using 10 ng of DNA template, 0.1 μM of each primer (with the exception of some markers as described in Table 2), and 1× PCR Master Mix of the QIAGEN kit multiplex PCR (Qiagen, Hilden, Germany). Two PCR cycling conditions were used. The first PCR cycling was as follows: pre-incubation for 15 min at 95 °C, followed by 5 cycles using a touchdown amplification program with an annealing temperature reduced by 1 °C per cycle from 65 °C to 60 °C. The next step involved 30 cycles, each consisting of 30 s denaturing at 95 °C, 60 s annealing at 60 °C, and 60 s elongation at 72 °C. The last cycle ended with a final 30 min extension at 72 °C. The second PCR cycling consisted of a 15-min pre-incubation at 95 °C, followed by 7 cycles with a touchdown amplification program from 65 °C to 58 °C, followed by 30 cycles, each consisting of 30 s denaturing at 95 °C, 60 s annealing at 58 °C, and 60 s elongation at 72 °C. The last cycle ended with a final 30 min extension at 72 °C. Fragment analysis and sizing were carried out using Geneious Prime v.2022.0.1 software (Thermo Fisher Scientific, Waltham, MA, USA). The PCR products were diluted with milliQ water at 1:5 (v/v) and mixed with 12 μL Hi-di Formamide (Applied Biosystems) and 0.5 μL size standard Gene Scan 600 Liz® (Applied Biosystems). Finally, the fragment analyses were performed using an Applied Biosystems 3130 DNA Genetic Analyzer (Applied Biosystems, Foster City, CA, USA).

2.5. Data Analysis for the Whole Dataset

2.5.1. Phenotype Statistical Analyses

All statistical analyses were carried out using the R language [42]. A one-way analysis of variance (ANOVA) with a p ≤ 0.05 was run to determine whether there were any statistically significant differences between the means of the evaluated traits. Moreover, Pearson’s bivariate correlations were performed to better determine how biochemical traits contribute to variability among accessions.

2.5.2. Diversity and Variability Assessment

Genetic parameters were carried out with the 23 microsatellites and for the whole dataset (Pop1), which corresponds to the 186 accessions from the EEAD–CSIC germplasm bank (186 accessions in total, divided into two different pools: 150 diploids, Pop2; and 36 triploids, Pop3). No multiloci SSR marker was detected in this study. The number of observed alleles per locus (NA), the effective number of alleles per locus (NE) [43], and rare alleles (NB: alleles with a frequency below 5%) were determined using the Genodive software [44]. Observed heterozygosity (Ho: number of heterozygous accessions/number of accessions assessed), expected heterozygosity (He = 1 − ∑ρi2, where ρi is the frequency of the ith allele) [45], Wright’s fixation index (Fis = 1 − Ho/He), and Shannon’s information index (I) [46] were calculated using the PopGene 1.32 software [47] (http://www.ualberta.ca, accessed on 15 December of 2022).

2.5.3. Analysis of Population Structure

Population structure analysis was first performed on the whole dataset (Pop1). All the accessions were categorized as Spanish or non-Spanish accessions. The program STRUCTURE (version 2.3.4) implements a model-based clustering criterion for inferring population structure using genotypic data from unlinked markers [48]. All kinds of models, including both “ancestry” and “allele frequency” models, were fitted with the selection of admixture and allele frequency correlation, respectively. Furthermore, we also performed ten independent runs per K value with a 10,000 burn-in period and 100,000 MCMC replications, starting with K = 1 to K = 6, under the admixture model. The statistic ΔK was then carried out, where K specifies the number of subpopulations or clusters. This analysis was based on the rate of change in the log probability of the data [49] to select the optimum number of K subpopulations for each population assessed.

2.6. Data Analysis for the Association Mapping for 118 Accessions

Genetic parameters for trait marker association were carried out for the 23 SSRs and according to the Pop4, corresponding to 118 phenotype diploids. Indeed, from the Pop2 (150 diploids), only 126 accessions were phenotyped, and eight resulted in duplicates (126 − 8 = 118 diploids: Pop4). In fact, from these 126 accessions and using the software Cervus v.3.0.7 [50], genetic uniqueness and redundancy were eliminated. The multi-locus DNA profile of all the accessions was compared pairwise under the identity analysis, setting the minimum number of matching loci at 23 and zero mismatches. Eight accessions resulted as duplicated diploid accessions from pairwise comparison of SSR profiles analyzed by Cervus software. The ‘Averdal-1’, ‘Averdal-2’, ‘Evasni’, ‘Red Elstar’, ‘Royal Red Delicious’, ‘Starkrimson-2’, and ‘Topred Delicious’ cultivars were duplicated among them, and thus, only one accession was kept in the study, the ‘Averdal-1’. Moreover, the ‘Galaxy’ and the ‘Regal Prince-1’ were duplicated between them, and the ‘Verde Doncella-1’ and ‘Verde Doncella-2’ were also duplicated between them. The ‘Galaxy’ and ‘Verde Doncella-1’ accessions were kept in the study.

2.6.1. Inter-Chromosomal Linkage Disequilibrium

The analysis of inter-chromosomal linkage disequilibrium (LD) was calculated using the trait analysis software by Association, Evolution, and Linkage (TASSEL, version 3.0.174, http://www.maizegenetics.net, accessed on 21 November of 2022). Alleles with frequencies below 5% were removed (minor allele frequency-MAF). Inter-chromosomal linkage Disequilibrium between pairs of multiallelic loci was calculated using the r2 coefficient, separately for loci in the same or different linkage groups (LG). The statistical r2 gives an indication of both recombination and mutation [51]. The significance level of LD between loci was examined using a permutation test implemented in TASSEL software for multiallelic loci, using the ‘rapid permutation’ option.

2.6.2. Association Mapping

TASSEL (v.3.0.174) was used with the General Linear Model (GLM) option [52] to examine associations between the phenotypic traits and the 23 SSR DNA markers. A structured association approach could be corrected for false associations using a Q-matrix of population membership estimates [52]. Therefore, the population membership estimates obtained from STRUCTURE analyses were fitted as a covariate in the GLM, where phenotype = population structure + marker effect + residual. As permutation methods can provide exact control of false positives and allow the use of non-standard statistics that make only weak assumptions about the data, a standard correction for multiple testing consisting of 10,000 permutations was run before carrying out the GLM.

3. Results

3.1. Phenotypic Evaluation and Pearson’s Correlations

The phenotypic evaluation and the statistical analysis were carried out for the Pop4 (118 diploids) out of the 186 apple accessions described in Table 1 and more extensively by Mignard et al. [20]. The one-way ANOVA analysis showed significant differences (p ≤ 0.001) among the different apple accessions for all traits evaluated (Table 3).
The SSC ranged among apple accessions from 10.14 (‘Bellaguarda Lardero’) to 17.03 (‘Eugenia’) °Brix. Regarding the TA, values varied greatly, ranging from 1.77 (‘Verde Doncella’) to 17.29 (‘Reguard-2’) g of malic acid per liter. The RI values ranged from 0.76 (‘Reguard-2’) to 8.55 (‘Verde Doncella’). The standard deviations for SSC, TA, and RI were fitted at 1.36, 2.92, and 1.34, respectively (Table 3). The TPC varied greatly among accessions, ranging from 15.24 (‘Biscarri-1’) to 98.07 (‘Camuesa Fina de Aragón’) mg GAE/100 g FW. For the TFC, values ranged from 6.00 (‘Biscarri-1’) to 88.95 (‘Camuesa Fina de Aragón’) mg CE/100 g FW. Regarding AsA, values ranged from 1.37 (‘Delgared Infel’) to 5.30 (‘Transparente’) mg AsA/100 g FW. Finally, the RAC values ranged from 5.93 (‘Delgared Infel’) to 30.82 (‘Les-1’) mg trolox/100 g FW. The standard deviations for TPC, TFC, AsA, and RAC, respectively, were fitted at 15.76, 14.67, 0.82, and 5.08 (Table 3).
Total sugar values (sugars) ranged significantly among apple accessions and years from 63.69 (‘Transparente Blanca’) to 115.27 (‘Fuji’) g kg−1 FW. Regarding the main individual sugars, fructose values ranged from 31.39 (‘Baujade’) to 61.41 (‘Akane’) g kg−1 FW, and the alcohol sugar sorbitol values varied from 1.20 (‘Plaona’) to 11.96 (‘Prau Riu-3’) g kg−1 FW. The standard deviations for total sugars, fructose, and sorbitol, were 9.87, 5.17, and 2.43, respectively. Total acid values (Acids) ranged among accessions and years from 3.42 (‘Verde Doncella’) to 11.39 (‘Astrakan Red’) g kg−1 FW. Regarding the main individual acids, malic acid ranged considerably compared with the others, from 2.68 (‘Delciri’) to 9.81 (‘Reguard-2’) g kg−1 FW. The standard deviations for total acids and malate were fitted at 1.82 and 1.69 g kg−1 FW, respectively.
Significant (p ≤ 0.01) bivariate correlations were found between the different traits evaluated except for the peel color and glucose which did not show any correlation (Figure 2). Positive and significant (p ≤ 0.01) correlations were found between phenolics and flavonoids (r = 0.96) and between phenolics, flavonoids, and RAC (r = 0.90 and r = 0.87, respectively). Significant positive correlations were also found between the antioxidant compounds and the organic acids. The individual sugars and the sum of sugars showed significant but low negative correlations with the bioactive compounds (TPC: r = −0.21, TFC: r = −0.26, and RAC: r = −0.25) and the organic acids. Moreover, high and significant correlations were found between harvest date and concentrations of soluble solids, sorbitol, sugars, and, as expected, with the ripening index. Significant and negative correlations were also found between the harvest date and individual/total organic acids, the titratable acidity, the antioxidant compounds (TPC and TFC), and the relative antioxidant capacity (RAC) (Figure 2). According to Pearson’s correlations, the non-Spanish accessions seemed to present lower organic acids and antioxidant compounds (TPC, TFC, and RAC) concentrations compared with the Spanish accessions.

3.2. Genetic Diversity

The mean estimated values for genetic parameters based on the 23 SSR loci assessed are presented in Table 4. According to Pop1 (186 accessions), the 23 microsatellites were all polymorphic, and amplicons could be observed for all of them, for a total of 360 alleles. The average number of alleles per locus (NA) was 15.65, ranging from 4 (MdSWEET2b) to 35 (CH01f02). Nevertheless, the number of effective alleles per locus (NE) was significantly lower (Table 4). Indeed, the average number of effective alleles (NE) overall the loci was 4.90, ranging from 1.5 (MdSWEET2e) to 8.98 (CH02d08). Moreover, rare alleles (NB) were found in all loci, and the number of them increased with the number of alleles per locus (r = 0.96). Alleles with a frequency lower than 5% (NB) varied from 33.3% at locus CH02c11 (5 out of 15) to 86.7% at locus CH04e05 (26 out of 30). Furthermore, all loci except for five (CH04e05, CH01f02, CH01h01, MdSWEET2e, and MdSWEET2d) were not in Hardy-Weinberg equilibrium (p ≤ 0.05). In fact, the 186 accessions assessed in this study do not belong to a panmictic population. While mean observed heterozygosity (Ho) was 0.75, ranging from 0.19 (MdSWEET2e) to 0.96 (MdSWEET7b) (Table 4), average expected heterozygosity (He) was 0.77, varying from 0.41 (MdSWEET2e) to 0.92 (CH05f06). Fis values were positive in 10 primers and negative in the remaining 13 SSRs, indicating a high level of heterozygosis in the genotypes assessed.
Among the Pop2 (150 diploid accessions) and Pop3 (36 triploid accessions), the average number of alleles per locus (NA) was 14.7 and 10, respectively, and the number of effective alleles per locus (NE) was fitted at 5.23 (Pop2) and 4.66 (Pop3). The observed and expected heterozygosities were higher in diploids (Ho = 0.79; He = 0.77) than in triploids (Ho = 0.71; He = 0.73). The Fis value was negative for the Pop2, while it was positive for the Pop3.
Pairwise comparison of multiloci profiles revealed eight duplicated diploid accessions with redundancies, leading to the removal of redundant accessions before further analyses (4.3% of redundancy). Several cases of homonymy (i.e., accessions with the same name but different profiles according to the 23 SSRs) were also found (‘Camosa-1’ and ‘Camosa-2’, ‘Jonathan-1’ and ‘Jonathan-2’, ‘Granny Smith-1’ and ‘Granny Smith-2’, ‘Mañaga-1’ and ‘Mañaga-2’, ‘Starking-1’ and ‘Starking-2’, and ‘Verde Doncella-1’ and ‘Verde Doncella-3’). The final number of unique diploid genotypes further analyzed was therefore 118 (Pop4).

3.3. Population Structure

Bar plots were obtained according to the values of K, the assumed number of subpopulations, corresponding to the number of clusters defined by ΔK [49] (Figure 3). Model-based clustering analyses were used to determine the genetic diversity structure within the 186 (Pop1) assessed Malus × domestica (Borkh) genotypes. A pairwise STRUCTURE analysis based on 23 allelic SSRs molecular data was carried out, and the maximum value of ΔK was observed at K = 2 for Pop2 (150 diploids) and Pop4 (118 diploids), suggesting two genetic clusters. Along with the 36 triploids in the study (Pop3), the maximum value of ΔK was revealed at K = 4, although at K = 2, an increase in ΔK was observed (Figure 3). Genotypes were divided into two or four clusters, respectively, based on their membership coefficients (Q), considering the genotypes as pure when the membership coefficient (qI) was greater than 0.80 and as an admixture or hybrid when the membership coefficient (qI) was lower than 0.80 [48].
The Pop2 (150 diploids) was represented by 72 Spanish and 78 non-Spanish accessions, the Pop3 (36 triploids) by 22 Spanish and 14 non-Spanish, and the Pop4 (118 diploids) by 72 Spanish and 46 non-Spanish accessions. However, the separation at K = 2 was congruent for all the diploid populations assessed, and admixture in each subpopulation was observed, demonstrating allele sharing (Figure 3). For Pop4 (118 diploids), the mean values of genetic differentiation (Fst) among the two subpopulations obtained with K = 2 were 0.25 and 0.0024, respectively. The Fst of the first subpopulation indicated a strong genetic differentiation for the accessions of this group. In addition, the allele frequency divergence among subpopulations was 0.0678 and demonstrated significant genetic differences between subpopulations (>0.05). Furthermore, regardless of the populations studied (Pop2 and Pop4), the first subpopulation was represented mostly by Spanish accessions and the second by the non-Spanish reference cultivars.
Additionally, the dendrogram obtained by the unweighted pair group method with arithmetic mean (UPGMA) analysis was compared with the phenotypic data (Figure 4) and the two subpopulations of the STRUCTURE results for the Pop4 (Figure 5). In the dendrogram obtained from the similarity matrix of pairwise analyses for the 23 SSR markers, there is a clear agreement between clusters representing genetic diversity and subpopulations obtained with STRUCTURE at K = 2. Furthermore, the population assessed can clearly be divided into six clusters/groups from the first two major clusters (cluster 1: groups 1, 2, 3, and 4; cluster 2: groups 5 and 6). According to the Spanish/non-Spanish classification, the first cluster of the UPGMA analysis (groups 1, 2, 3, and 4) included only 16 non-Spanish accessions out of 85. Group 1 included 15 non-Spanish and 13 Spanish accessions. The groups 2 and 3 did not show any non-Spanish, and the group 4 included only one, the ‘Jonadel’ accession. The second cluster (groups 5 and 6) was represented by 33 accessions, and 30 of them were characterized as non-Spanish accessions. The three Spanish accessions observed in this second cluster were ‘Biscarri-1’, ‘De Valdés’, and ‘Valsaina’ (Figure 4 and Figure 5).
The 19 phenotypic traits were divided into two groups of variables (Figure 4). Firstly, the TA, TPC, TFC, RAC, AsA, tartrate, oxalate, citrate, quinate, malate, and the sum of organic acids were clustered. On the other hand, succinate + shikimate, SSC, RI, sucrose, glucose, fructose, sorbitol, and the sum of individual sugars were grouped. Indeed, clusters would group variables that tended to behave similarly across the accessions. The first group of traits was dominated by the antioxidant parameters, the TA, and the individual organic acids (except the sum of shikimate and succinate), while the second corresponded to the individual sugars, the SSC, and the RI traits. Cluster 1 (groups 1, 2, 3, and 4) of accessions seemed to present a more acidic profile, while groups 5 and 6 showed a sweeter profile. Nevertheless, the more acidic cultivar assessed was non-Spanish (‘Astrakan Red’) and was in Cluster 1—group 1. The cluster 1 showed more antioxidant compound values than the groups 5 and 6. Indeed, the ‘Camuesa Fina de Aragón’ accession exhibited the highest TPC values with 98.1 mg GAE 100 g FW−1, far away from the first non-Spanish accession (‘Akane’) with 55.59 mg GAE 100 g FW−1. The ‘Les-1’ accession was the one with the highest relative antioxidant capacity (RAC) found among all the Pop4 data set (30.8 mg Trolox 100 g FW−1) while the ‘Delgared Infel’ (non-Spanish) was the one with the lowest RAC observed (5.9 mg Trolox 100 g FW−1). According to the harvest date, ranging from 170 (‘Mañaga-2’) to 316 (‘Bossost-4’), Julian days, the first cluster (groups 1, 2, 3, and 4) seemed to cluster accessions that ripened later than the accessions of the second cluster (groups 5 and 6). The peel color did not seem to influence clustering in the Pop4. Indeed, six accessions were characterized as ‘Red’ and they were in different groups: ‘Redaphough’ belonged to group number 1; ‘Averdal’, ‘Nueva Starking’, and ‘Starkrimson’ were in group 5, while ‘Delgared Infel’ and ‘Prima’ belonged to group 6.

3.4. Inter-Chromosomal Linkage Disequilibrium and Association Mapping

The inter-chromosomal linkage disequilibrium (LD) arrays of all 253 pairwise combinations of the 23 SSRs were assessed using TASSEL (Figure 6). The LD r2 values varied greatly from 0.00 to 0.77, and r2 = 1 indicates that two different markers provide exactly the same information. The highest LD value was recorded between the CH02c11 and CH02c06 markers. The significance cut-off threshold values from the distribution of LD were assigned at r2 = 0.2. Indeed, 15 pairwise combinations of the 23 SSRs assessed recorded values above the cut-off. Thus, CH01h10 and CH02c11 markers were considered to be in linkage disequilibrium, as well as markers such as CH02c06 with Hi02c07, CH05f06, CH02c09, and CH02c11; Ch02c09 with CH02c11; CH02c11 with CH05f06; MdSWEET2d with CH02c09, CH02c06, and CH02c11; and finally, MdSWEET7b with CH01h10, CH02c11, and CH02c06 and CH02c09 markers.
The results showed that 126 significant associations were reached between the 23 SSR markers and the 21 traits evaluated, using a modeling coefficient of membership (Q) value estimate from structure as a covariate (Table 5). In total, 20 SSRs (out of 23) contributing to the phenotypic variation were significantly associated with at least one trait studied. Moreover, only fructose was not associated with any of the 23 SSRs assessed. Furthermore, the SWEET gene markers MdSWEET2a, MdSWEET2d, MdSWEET7b, MdSWEET2b, MdSWEET7a, and MdSWEET12a showed associations (p ≤ 0.01) with 17 out of the 21 phenotypic traits assessed (harvest, peel color, RAC, TPC, TFC, SSC, TA, RI, sucrose, sorbitol, sum of sugars, oxalate, citrate, tartrate, malate, quinate, and sum of organic acids).
According to the agronomical parameters, five significant associations were found with the harvest date on LG5 (CH05f06 and MdSWEET2d), LG9 (CH01f03b), and LG14 (CH04c07 and MdSWEET12a), and two with the peel color and CH02c11 and MdSWEET7a markers on LG10 and 11. Regarding the antioxidant compounds, 23 associations were found between them, with 11 SSRs on LG2, 3, 4, 5, 6, 7, 10, 11, 12, 14, and 17. At a level of significance of p ≤ 0.0001, the relative antioxidant capacity (RAC) was associated with the MdSWEET7b marker, the TPC with the CH05f06 and MdSWEET7b markers, and finally, the TFC was linked with two markers (CH02c06 and CH05f06). For the basic fruit quality parameters, 25 associations were found between them, with 15 SSRs on LG1, 2, 3, 5, 6, 7, 9, 11, 12, 14, 15, and 17. At the level of significance (p ≤ 0.0001), the RI was associated with the CH04c07 marker and the TA with the CH01f02 marker. Concerning the sugar compounds, 15 associations have been observed between them, along with 10 SSR markers on LG1, 2, 4, 5, 11, 12, 13, 14, and 17. At the level of significance of p ≤ 0.0001, four significant associations were found between sucrose content and the CH-Vf1 marker, between sorbitol content and the CH01f02 and CH02c06 markers, and between the sum of the individual sugars and the MdSWEET12a marker. Lastly, according to the individual organic acids, 53 associations were found between them, with 17 SSRs in all LGs except the LG10 and LG13. At the level of significance p ≤ 0.0001, seven associations were found between the succinate and shikimate acids and the CH01f03b, CH02c06, and CH04e05 markers; the citric and quinic acids with the CH05f06 marker; and the oxalate acid with the CH02c06 and MdSWEET2a markers.

4. Discussion

4.1. Phenotypic Characterization

As expected, a large variability for the different traits assessed was found among accessions, and the concentrations of biochemical compounds observed in the present study were in line with previously reported values [18,27,53,54,55,56,57,58,59]. Nonetheless, the larger sample size of this study (118 accessions) resulted in a greater range of concentrations with significant differences in the values of the traits assessed.
Moreover, positive and significant correlations were found, as previously reported in other studies, between total phenolics and total flavonoids, or RAC [20,60,61]. Many of the correlations observed could be explained by photosynthetic activity [56,62]. Indeed, the photosynthetic products will act as substrates for many of the metabolic pathways. The sum of individual sugars and main products of the primary metabolism, photosynthesis and SSC, have shown significant and negative correlations with bioactive compounds (TPC and TFC) and positive correlations with all the organic acids, substrates, and products of the secondary metabolism in plants [63]. These correlations could thus be explained as a possible response to phenolic compound biosynthesis. Indeed, carbohydrates such as fructose increase the erythose-4-phosphate productions that, together with phosphoenolpyruvate (PEP), constitute a substrate for phenolic compounds through the shikimate pathway [64]. Furthermore, positive and significant correlations were also found between the antioxidants and the organic acids, as also reported by Mignard et al. [5] when they studied 155 apple accessions. The decarboxylation of dicarboxylates such as malate and other organic acids is linked to the degradation of organic acids, and thus, this decarboxylation permits PEP production. The PEP is closely associated with the activation of gluconeogenesis and thus results in glucose production in fruits [64]. These correlations showed the linkage between the primary and secondary metabolisms [65]. In fact, the secondary metabolism (antioxidant compounds) is connected to the primary metabolism (sugars and acids) as substrates are supplied from primary pathways and drove into the secondary biosynthetic routes. The positive and significant correlations between the different individual sugars could be explained through the polyol or sorbitol-aldose reductase pathway, a two-step mechanism converting glucose into fructose.
Moreover, high and significant correlations were found between the harvest date and the concentrations of SSC, sorbitol, sugars, and, as expected, with the RI. These results show that when fruits are harvested late in the season but each one is at its optimum commercial maturity, they are, in general, sweeter. In contrast, significant and negative correlations were found between the harvest date and individual/total organic acids, TA, the antioxidant compounds (TPC and TFC), and the relative antioxidant capacity (RAC) [65]. Furthermore, the non-Spanish accessions seemed to present lower organic acid concentrations and fewer antioxidant compounds (TPC, TFC, and RAC) compared with the Spanish well adapted accessions such as ‘Camuesa Fina de Aragón’, ‘Reguard–2’, ‘Transparente’, and ‘Verde Doncella’, as previously reported [5,20]. These results highlight the importance of germplasm characterization [66] with the aim of boosting autochthone accessions and phytogenetic resources for breeding programs.

4.2. Genetic Identity and Overall Diversity

The successful amplification and polymorphism obtained using 23 SSR markers covering the apple genome used to screen the 186 apple accessions (Pop1) confirmed previously reported results for cultivar identification and genetic mapping in apple trees [1,10,11,27,67,68,69,70].
Although some ancient apple Spanish accessions, relevant in the past in Spain [71], were assessed in this study, there was no proof that they were true-to-type accessions [1,11]. Indeed, the different accessions labeled ‘Camosa-1’/‘Camosa-2’, ‘Reguard-2’/‘Reguard-4’, ‘Mañaga-1’/‘Mañaga-2’ or ‘Verde Doncella-1’/‘Verde Doncella-3’ did not show the same SSR profiles. In contrast, ‘Verde Doncella-2’ and ‘Verde Doncella-3’ exhibited exactly the same allelic profiles. Moreover, the apple genetic resources analysis should be studied with attention because of the incidence of mutations, the genome structure variations, or the epigenetic alterations that could engender phenotypic modifications that are not distinguishable using only SSR markers [8,11,12]. Consequently, the phenotypic and molecular characterizations of apple accessions, both complementary approaches, could determine whether the accessions with the same SSR profiles would be the same accessions.
Among the 23 different SSR markers, the level of genetic diversity and expected/observed heterozygosity were relatively high. As apple is a self-incompatible cross pollinating specie [11], the high diversity observed indicated that the apple genetic resources, and thus the accessions preserved in the germplasm bank of the EEAD–CSIC, were highly diverse. The high genetic diversity found was in agreement with the variability exhibited in other apple studies [1,8,11,23,68,70,72,73]. All SSR loci analyzed in this study displayed a high degree of polymorphism, with 4 to 35 alleles per locus and a 1.5 to 8.98 effective number of alleles per locus. Indeed, the mean value found in this study was a 15.65 number of alleles per locus for Pop1, which is slightly lower than the 16.69 observed by Urrestarazu et al. [1] studying 493 accessions and the 18.62 reported by Pereira-Lorenzo et al. [11] with 1453 accessions in Spain and sharing respectively 16 SSRs (13 markers out of 16 were used in the present study) and 13 SSR markers (12 out of 13 were included in the present work). Moreover, 18 alleles on average per locus were found in a Turkish germplasm collection of 206 accessions [67] and 19.5 in a collection of 2163 accessions in France [8]. Nevertheless, Marconi et al. [10], assessing 175 accessions in Italy, and Ferreira et al. [68], studying 87 accessions in Portugal, observed a mean value for the number of alleles per locus of 14.6 and 11.5 respectively. The observed heterozygosity averaged for the entire dataset (0.75) over the 23 SSR loci was exactly the same as reported by Meland et al. [74], studying 171 accessions and using seven SSR markers also used in the present study (CH02c06, GD12, CH01h10, CH02c11, CH02d08, CH02c09, and CH01h01 markers). Moreover, the H0 was higher than the 0.67 observed by Ferreira et al. [68] but slightly lower than reported values of 0.78 [10], 0.83 [8], 0.76 [67], and 0.81 [1]. The differences found in these studies could be due to the different plant material, the ploidy level, or even the number of accessions assessed in each study.
According to the triploid accessions, it is worth mentioning that the amplification of three alleles in a single locus is not evidence for characterizing an accession as triploid [23,68]. In fact, in diploid accessions, a third fragment could be amplified as the result of duplication or a somatic mutation [23]. However, if more than two alleles were found at several loci (several SSR markers), the accessions were thus characterized as triploids. Furthermore, in the present study, a confirmation of the ploidy level by flow cytometry for the whole dataset was assessed, as described by Reig et al. [18].

4.3. Population Structure

The structure results showed that there were two main subpopulations (K = 2) with some degree of admixture within both of them (Spanish and non-Spanish accessions). Similar studies in apples also reported two unstructured populations, indicating a strong subpopulation structure using respectively 1453 [11] and 493 accessions [1]. Indeed, Pereira-Lorenzo et al. [11] reported that the analysis of 1453 apple accessions, conserved in Spanish collections, permitted the discrimination of an Iberian genepool of apple accessions separated from an extensive set of non-Spanish reference modern cultivars.
The UPGMA cluster analysis grouped all the 118 apple accessions into a dendrogram. The cluster analysis was able to group all genotypes into two large clusters, with the first one (cluster 1) containing four subgroups (groups 1 to 4) and the second one (cluster 2) containing two subgroups (groups 5 and 6). The UPGMA analysis also showed that one of the two subpopulations encountered was quite exclusively formed by non-Spanish accessions (30 out of 33). Indeed, the cluster 1 included 16 non-Spanish cultivars (‘Astrakan Red’, ‘Baujade’, ‘Belleza de Roma’, ‘Blackjon’, ‘Bofla’, ‘Florina’, ‘Fuji’, ‘Granny Smith-1’, ‘Granny Smith-2’, ‘Idared’, ‘Jonadel’, ‘Jonathan-1’, ‘Jonathan-2’, ‘MacIntosh’, ‘Red Rome Beauty’, and ‘Redaphough’) and the cluster 2 included all the 30 remaining non-Spanish reference cultivars and only three Spanish accessions (‘Biscarri-1’, ‘De Valdés’, and ‘Valsaina’). Nevertheless, in the cluster 1, 15 non-Spanish accessions were included in the subgroup 1, and only one was in the subgroup 4 (‘Jonadel’). The groups 2 and 3 did not show any non-Spanish accession and could thus be referred to as the Iberian genepool described by Pereira-Lorenzo et al. [11].
The results for genetic subpopulation obtained suggested that the non-Spanish reference cultivars were slightly more similar among them than with the Spanish accessions. In fact, several studies have shown that the European germplasm core collections shared a lot of the non-Spanish reference plant material conserved [1,6,10,11,16,66,75,76]. Moreover, according to the autochthone accessions from the different germplasm collections around the world, domestication and breeding could have caused diversity loss. Nevertheless, regardless of the many decades of domestication of Malus × domestica and its clonal propagation, there is no proof that domesticated apples have shown a genetic bottleneck in contrast with Malus × sieversii [77,78].
It is interesting to note that the same Spanish/non-Spanish cluster segregation was also distinguishable in previous studies regarding the influence of climate parameters on basic fruit quality (SSC, TA, and RI), antioxidant traits (TPC, TFC, AsA, and RAC), individual sugars (glucose, fructose, sucrose, and sorbitol), and organic acids (citric, malic, oxalic, quinic, succinic, shikimic, and tartaric) contents of 155 accessions out of the 186 accessions included in Pop1 assessed in this study. Moreover, the range of the results was larger for the Spanish accessions than for the non-Spanish ones [5,20], highlighting more similar profiles according to the non-Spanish cultivars. Indeed, Spanish autochthone accessions reported a higher biodiversity and, in general, higher contents for some basic fruit quality traits, antioxidants, individual sugars, and organic acids than non-Spanish accessions over a period of five years of study. Higher contents of antioxidants and organic acids were observed, in general, in the groups 1, 2, 3, and 4 of the clustering analysis, while the sugar profile was more heterogeneous according to the accession’s origins. These results strengthen the awareness of the importance of autochthone phytogenetic resources and underline the high biodiversity found in germplasm core collections [66]. Moreover, climatic traits such as precipitation, solar radiation, and temperatures strongly influenced the antioxidant and metabolite profiles of the accessions studied and depended on their origin [5,18,20,55,56,65]. Nevertheless, Pereira-Lorenzo et al. [11] reported the same segregation with a total of 1453 accessions, including part of the accessions of the present work and different germplasm collections from Spain. Moreover, Mignard et al. [5,20] showed segregation between Spanish local accessions and modern non-Spanish cultivars according to the biochemical contents and the climatic factors influence (high solar radiation and low temperatures). This highlighted the fact that, further than the climatic parameters influence on the metabolite profiles of the apple fruits [79], the genetics have a stronger influence on the biochemical contents and explained why two subpopulations were found in previous fruit quality studies [5,20].

4.4. Association Mapping

To our knowledge, the present assay is the first study reported in apples concerning association mapping with agronomical and biochemical traits in a large germplasm collection using both Spanish and non-Spanish apple genetic resources. Twenty-one traits of interest, such as basic fruit quality, antioxidant parameters, individual sugars, and organic acids, were assessed by 23 SSR molecular markers. Regarding the bibliography, scarce studies of phenotypic associations in apples using SSRs can be found [80]. The most recent studies analyzed several parameters, such as flowering time, harvest date, flesh firmness, ripening index, or polyphenols, but based only on SNP markers [12,32,33,34,35,75,76]. However, Tsykun et al. [81] reported that the multi-allelic SSRs markers seemed to be best suited for detecting genetic structure than SNPs markers because the SSRs markers had a higher discrimination power than bi-allelic SNP markers [82].
Zhen et al. [27] reported five associations between the SWEET SSR markers and four individual sugars (sucrose, glucose, fructose, and sorbitol). Indeed, the MdSWEET2e marker was significantly associated with sucrose, fructose, and total sugars [27], but it was not in the present study. The MdSWEET9b marker showed significant association with the contents of sorbitol and was not linked significantly with fructose and total sugars, as reported by Zhen et al. [27]. Nevertheless, the MdSWEET9b marker also showed associations with RAC, oxalate, and quinate. In fact, Zhen et al. [27] did not find associations between the other SWEET genes SSR markers (MdSWEET7b, MdSWEET2d, MdSWEET2b, MdSWEET2a, MdSWEET12a, and MdSWEET17a) and the individual sugars. In the present study, more phenotypic traits were assessed, and thus, more significant associations were found.
Apart from the MdSWEET markers, the SSR markers assessed in the present work have not been used for association mapping with biochemical traits in previous studies. Nevertheless, QTLs with different phenotypic traits have been encountered, and these results may be very useful because many of the associated markers were located in common regions where major genes or QTLs have been previously identified on the apple genome [35,83,84,85,86,87,88,89].
According to the agronomical parameters, harvest date showed a significant association with the CH01f03b marker. Kenis et al. [86] reported a QTL with harvest date located in the LG9 as well as with the CH01f03b marker. Other studies [34], have shown associations between harvest date and chromosome 3, instead of LG9, using SNP markers, but they failed to find any candidate gene associated with this character. These differences in the results may be due to differences in plant material used and in the coverage density between SNPs and SSRs. The peel color trait was associated with the CH02c11 marker at LG10 and with the MdSWEET7a marker at LG11. Howard et al. [85] found many QTLs in LGs 2, 5, 8, and 9 regarding these agronomical traits. Moreover, Oh et al. [90] found QTLs for the apple skin color in LGs 9, 10, and 13 using the ‘Jonathan’ cultivar. A higher number of significant QTLs related to the basic fruit quality traits were found in previous works [84,86,87,89].
Significant associations between SSC and TA with the CH02c06 marker at LG2 were found. Guan et al. [84], Kunihisa et al. [87], and Zhang & Han [89] reported the same QTL for SSC in the LG2. Indeed, SSC was linked to the CH02c06 (LG2), the CH05f06 (LG5), the MdSWEET12a (LG14), and the CH02c09 marker (LG15), while TA was associated with SSR markers at LGs 1, 2, 3, 5, 9, 11, 12, and 14. Regarding the ripening index, the associations found by Urrestarazu et al. [12] in chromosomes 3, 10, and 16 were different from our outcomes. Fruit ripening trait is a complex character that is quantitatively inherited in most fruit tree species [91], and its control involves the regulation of many metabolic pathways such as starch, acidity, firmness, and changes in color, among others [92]. In any case, our results with SSRs would be a good approximation since Urrestarazu et al. [12] found two NAC transcription factors on LG3, close to the CH02c06 SSR marker on LG2, as candidate genes for the control of this trait. Indeed, the NAC genes constitute one of the largest families of plant-specific transcription factors and are present in a wide range of species [93]. In the same way, a strong micro-synteny was identified between Malus and Prunus, identifying a major locus on chromosome 4 controlling the maturity date of peaches. The NAC transcription factors controlling fruit ripening traits have also been described in tomato [94] and kiwifruit [95].
Antioxidant compounds were highly linked to several SSR markers assessed. Twenty six significant associations were found between the antioxidant compounds and SSR markers in LGs 2, 3, 4, 5, 6, 7, 10, 11, 12, 14, and 17. Chagné et al. [83] showed associations between antioxidant parameters and QTLs at LGs 3, 5, and 14, in agreement with those observed in the present study. In addition, McClure et al. [35] using GWAS identified a candidate gene (LAR1) within chromosome 16 for catechin production and several transcription factors of different classes (MYB, bHLH, bZIP, AP2). These results showed that the polyphenol content in apples may suggest that breeders may be able to improve the nutritional value of apples through marker assisted breeding (MAB) or gene editing. The high content of total phenols was correlated with high radiation and low temperatures, as shown by Mignard et al. [20]. These results are of vital importance, taking into account that extreme weather conditions will affect the fruit quality, with the Ebro Valley, in Spain, being one of the most vulnerable areas for apple production and also one of the most affected Mediterranean regions influenced by these changes.
According to the individual sugars, sucrose was highly linked to the CH-Vf1 marker (LG1), which is consistent with the findings of Larsen et al. [33] between sucrose and the chrl:30221387 SNP. Furthermore, these authors identified that VIN1 is linked to the chrl:30221387 SNP, making it a good candidate gene for this association. Other authors described different vacuolar invertase genes, including VIN1 [96], which play an important role in sugar metabolism in fruits. Further, Guan et al. [84] and Sun et al. [88] reported a QTL in the LG1 associated with sucrose and also other QTLs for fructose, glucose, and sorbitol but in different LGs.
Regarding the organic acids, a significant association with the CH04c07 marker (LG14) was found in the present work. Chagné et al. [83] reported a QTL in the LG14 linked to the quinic acid content. Moreover, for organic acids, significant associations between CH04e05 (LG7), CH01h10 (LG8), and CH02c09 markers (LG15) with oxalate, citrate, succinate, shikimate, and tartrate were found. Sun et al. [88] showed several QTLs in LGs 7, 8, and 15, which are the same LGs as the SSR markers linked to organic acids in the present study. Oh et al. [90] found a fruit acidity-related QTL in the LG13 of the ‘Jonathan’ cultivar, although no significant association was found in the present work.
The results obtained provided a strong base for further association mapping with agronomical and biochemical traits that could be applied to other species because of the synteny within the Rosaceae family [12,97,98]. Moreover, the 126 significant trait-marker associations found in the present study could provide potential information for effective marker assisted selection (MAS) in apple breeding programs. Despite being the first approximation made to date in apples between biochemical traits and SSR markers for the 21 parameters assessed, significant associations have been found and will be of great help in further work. New studies must be performed mapping thousands of SNPs (9K IRSC SNP arrays) to facilitate genome-wide scans and validate marker–locus–trait associations for application in breeding.

5. Conclusions

The present study provided new details about the EEAD-CSIC core collection population structure according to their ploidy level and their origin (Spanish/non-Spanish). The population structure analysis showed two subpopulations in agreement with the phenotypic segregation observed in previous studies. These results highlight the importance of understanding the genetic architecture of important fruit agronomical and biochemical traits because of the intrinsic correlation between the genetics and the metabolite profiles of the apple accessions. A total of 126 significant associations were observed between the 23 SSR markers assessed and the 21 phenotypic traits evaluated. These results would help in breeding programs with marker assisted selection (MAS) for fruit quality traits in apples, and in particular, the possible simultaneous selection for agronomical and biochemical parameters. The content of some nutritional compounds, such as total phenol content, also associated with some SSR markers, has turned out to be correlated with extreme climatic factors, making this a point to take into account in future fruit quality breeding programs. This study also identified many new marker-locus trait associations, such as with antioxidants or organic acid compounds, which may be useful for predictions and for a better understanding of the apple genome. Finding specific regions of the genome will provide further information regarding candidate genes involved in apple fruit quality.

Future Perspectives

With the purpose of moving faster in the association mapping studies, future studies may be required in this area with a view to candidate gene identification and fine mapping using a large-scale of SNP markers. Recent advances in genomic tools such as genome wide association mapping (GWAS) and next-generation sequencing (NGS) have allowed the development of new approaches for mapping important and complex traits and facilitating SNP discovery. The identification of causal genes underlying specific traits is a major goal in plant breeding, subsequently offering opportunities to develop genomic selection tools. Moreover, due to the abundance of SNPs within a genome and the availability of high-throughput sequencing methods, SNPs are increasingly becoming one of the most commonly used markers for genotyping horticultural species.

Author Contributions

Conceptualization, M.Á.M.; methodology, P.M., C.F.i.F. and R.G.; software, P.M. and C.F.i.F.; validation, P.M., C.F.i.F. and M.Á.M.; formal analysis, P.M. and C.F.i.F.; investigation, P.M., C.F.i.F. and M.Á.M.; resources, M.Á.M.; data curation, P.M., C.F.i.F. and M.Á.M.; writing—original draft preparation, P.M. and C.F.i.F.; writing—review and editing, P.M., C.F.i.F., R.G. and M.Á.M.; visualization, P.M., C.F.i.F. and M.Á.M.; supervision, M.Á.M.; project administration, M.Á.M.; funding acquisition, M.Á.M. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Spanish Ministry of Science, Innovation, and Universities grants RFP 2012-00020, RFP 2015-00019, and PIE201640E070, and ‘ERDF A way of making Europe’, and the Regional Government of Aragon (A44, T07_17R, A22_20R, and the P. Mignard grant).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank María Pilar Soteras, Fernando Gómez, and Gemma Reig for their technical assistance in fruit quality analysis and plant management in the orchard. The authors are grateful to Jorge Urrestarazu and Carlos Lahoz for providing equipment access and advice.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Urrestarazu, J.; Miranda, C.; Santesteban, L.G.; Royo, J.B. Genetic diversity and structure of local apple cultivars from Northeastern Spain assessed by microsatellite markers. Tree Genet. Genomes 2012, 8, 1163–1180. [Google Scholar] [CrossRef]
  2. FAOSTAT, 2021. Food and Agricultural Organization. Available online: http://faostat.fao.org (accessed on 31 January 2023).
  3. MAPAMA-España Ministerio de Agricultura, Pesca y Alimentación. Available online: https://www.mapa.gob.es/es/ (accessed on 22 October 2022).
  4. Mignard, P.; Beguería, S.; Reig, G.; i Forcada, C.F.; Moreno, M. Phenotypic analysis of fruit quality traits and effect of climate in an apple (Malus × domestica Borkh) germplasm bank of Aragón, Spain. Acta Hortic. 2021, 1307, 109–114. [Google Scholar] [CrossRef]
  5. Mignard, P.; Beguería, S.; Giménez, R.; i Forcada, C.F.; Reig, G.; Moreno, M. Effect of Genetics and Climate on Apple Sugars and Organic Acids Profiles. Agronomy 2022, 12, 827. [Google Scholar] [CrossRef]
  6. Urrestarazu, J.; Denancé, C.; Ravon, E.; Guyader, A.; Guisnel, R.; Feugey, L.; Poncet, C.; Lateur, M.; Houben, P.; Ordidge, M.; et al. Analysis of the genetic diversity and structure across a wide range of germplasm reveals prominent gene flow in apple at the European level. BMC Plant Biol. 2016, 16, 130. [Google Scholar] [CrossRef] [PubMed]
  7. Ordidge, M.; Kirdwichai, P.; Baksh, M.F.; Venison, E.P.; Gibbings, J.G.; Dunwell, J.M. Genetic analysis of a major international collection of cultivated apple varieties reveals previously unknown historic heteroploid and inbred relationships. PLoS ONE 2018, 13, e0202405. [Google Scholar] [CrossRef] [Green Version]
  8. Lassois, L.; Denancé, C.; Ravon, E.; Guyader, A.; Guisnel, R.; Hibrand-Saint-Oyant, L.; Poncet, C.; Lasserre-Zuber, P.; Feugey, L.; Durel, C.-E. Genetic Diversity, Population Structure, Parentage Analysis, and Construction of Core Collections in the French Apple Germplasm Based on SSR Markers. Plant Mol. Biol. Rep. 2016, 34, 827–844. [Google Scholar] [CrossRef] [Green Version]
  9. Gao, Y.; Liu, F.; Wang, K.; Wang, D.; Gong, X.; Liu, L.; Richards, C.M.; Henk, A.D.; Volk, G.M. Genetic diversity of Malus cultivars and wild relatives in the Chinese National Repository of Apple Germplasm Resources. Tree Genet. Genomes 2015, 11, 106. [Google Scholar] [CrossRef]
  10. Marconi, G.; Ferradini, N.; Russi, L.; Concezzi, L.; Veronesi, F.; Albertini, E. Genetic Characterization of the Apple Germplasm Collection in Central Italy: The Value of Local Varieties. Front. Plant Sci. 2018, 9, 1460. [Google Scholar] [CrossRef] [Green Version]
  11. Pereira-Lorenzo, S.; Urrestarazu, J.; Ramos-Cabrer, A.; Miranda, C.; Pina, A.; Dapena, E.; Moreno, M.; Errea, P.; Llamero, N.; Díaz-Hernández, M.; et al. Analysis of the genetic diversity and structure of the Spanish apple genetic resources suggests the existence of an Iberian genepool. Ann. Appl. Biol. 2017, 171, 424–440. [Google Scholar] [CrossRef]
  12. Urrestarazu, J.; Muranty, H.; Denancé, C.; Leforestier, D.; Ravon, E.; Guyader, A.; Guisnel, R.; Feugey, L.; Aubourg, S.; Celton, J.-M.; et al. Genome-Wide Association Mapping of Flowering and Ripening Periods in Apple. Front. Plant Sci. 2017, 8, 1923. [Google Scholar] [CrossRef] [Green Version]
  13. Vanderzande, S.; Micheletti, D.; Troggio, M.; Davey, M.W.; Keulemans, J. Genetic diversity, population structure, and linkage disequilibrium of elite and local apple accessions from Belgium using the IRSC array. Tree Genet. Genomes 2017, 13, 125. [Google Scholar] [CrossRef]
  14. Gross, B.L.; Volk, G.M.; Richards, C.M.; Forsline, P.L.; Fazio, G.; Chao, C.T. Identification of “Duplicate” Accessions within the USDA-ARS National Plant Germplasm System Malus Collection. J. Am. Soc. Hortic. Sci. 2012, 137, 333–342. [Google Scholar] [CrossRef] [Green Version]
  15. Velasco, R.; Zharkikh, A.; Affourtit, J.; Dhingra, A.; Cestaro, A.; Kalyanaraman, A.; Fontana, P.; Bhatnagar, S.K.; Troggio, M.; Pruss, D.; et al. The genome of the domesticated apple (Malus × domestica Borkh.). Nat. Genet. 2010, 42, 833–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Cornille, A.; Gladieux, P.; Smulders, M.J.M.; Roldán-Ruiz, I.; Laurens, F.; Le Cam, B.; Nersesyan, A.; Clavel, J.; Olonova, M.; Feugey, L.; et al. New Insight into the History of Domesticated Apple: Secondary Contribution of the European Wild Apple to the Genome of Cultivated Varieties. PLOS Genet. 2012, 8, e1002703. [Google Scholar] [CrossRef] [Green Version]
  17. Noiton, D.A.; Alspach, P.A. Founding Clones, Inbreeding, Coancestry, and Status Number of Modern Apple Cultivars. J. Am. Soc. Hortic. Sci. 1996, 121, 773–782. [Google Scholar] [CrossRef] [Green Version]
  18. Reig, G.; Blanco, Á.; Castillo, A.M.; Gogorcena, Y.; Moreno, M.Á. Phenotypic diversity of Spanish apple (Malus × domestica Borkh) accessions grown at the vulnerable climatic conditions of the Ebro Valley, Spain. Sci. Hortic. 2015, 185, 200–210. [Google Scholar] [CrossRef] [Green Version]
  19. Fernández i Martí, A.; Font i Forcada, C.; Company, R.S.I.; Rubio-Cabetas, M.J. Genetic relationships and population structure of local olive tree accessions from Northeastern Spain revealed by SSR markers. Acta Physiol. Plant. 2015, 37, 1726. [Google Scholar] [CrossRef]
  20. Mignard, P.; Beguería, S.; Reig, G.; i Forcada, C.F.; Moreno, M. Genetic origin and climate determine fruit quality and antioxidant traits on apple (Malus × domestica Borkh). Sci. Hortic. 2021, 285, 110142. [Google Scholar] [CrossRef]
  21. Fernández i Martí, A.; Font i Forcada, C.; Kamali, K.; Rubio-Cabetas, M.J.; Wirthensohn, M.; i Company, R.S. Molecular analyses of evolution and population structure in a worldwide almond [Prunus dulcis (Mill.) D.A. Webb syn. P. amygdalus Batsch] pool assessed by microsatellite markers. Genet. Resour. Crop. Evol. 2014, 62, 205–219. [Google Scholar] [CrossRef]
  22. Urrestarazu, J.; Errea, P.; Miranda, C.; Santesteban, L.G.; Pina, A. Genetic diversity of Spanish Prunus domestica L. germplasm reveals a complex genetic structure underlying. PLoS ONE 2018, 13, e0195591. [Google Scholar] [CrossRef] [Green Version]
  23. Liang, W.; Dondini, L.; De Franceschi, P.; Paris, R.; Sansavini, S.; Tartarini, S. Genetic Diversity, Population Structure and Construction of a Core Collection of Apple Cultivars from Italian Germplasm. Plant Mol. Biol. Rep. 2014, 33, 458–473. [Google Scholar] [CrossRef]
  24. Vanderzande, S.; Howard, N.P.; Cai, L.; Linge, C.D.S.; Antanaviciute, L.; Bink, M.C.A.M.; Kruisselbrink, J.W.; Bassil, N.; Gasic, K.; Iezzoni, A.; et al. High-quality, genome-wide SNP genotypic data for pedigreed germplasm of the diploid outbreeding species apple, peach, and sweet cherry through a common workflow. PLoS ONE 2019, 14, e0210928. [Google Scholar] [CrossRef] [Green Version]
  25. Patocchi, A.; Fernández-Fernández, F.; Evans, K.; Gobbin, D.; Rezzonico, F.; Boudichevskaia, A.; Dunemann, F.; Stankiewicz-Kosyl, M.; Mathis-Jeanneteau, F.; Durel, C.E.; et al. Development and test of 21 multiplex PCRs composed of SSRs spanning most of the apple genome. Tree Genet. Genomes 2008, 5, 211–223. [Google Scholar] [CrossRef] [Green Version]
  26. Liu, Z.; Bao, D.; Liu, D.; Zhang, Y.; Ashraf, M.A.; Chen, X. Construction of a Genetic Linkage Map and QTL Analysis of Fruit-related Traits in an F1 Red Fuji × Hongrou Apple Hybrid. Open Life Sci. 2016, 11, 487–497. [Google Scholar] [CrossRef] [Green Version]
  27. Zhen, Q.; Fang, T.; Peng, Q.; Liao, L.; Zhao, L.; Owiti, A.; Han, Y. Developing gene-tagged molecular markers for evaluation of genetic association of apple SWEET genes with fruit sugar accumulation. Hortic. Res. 2018, 5, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Eom, J.-S.; Chen, L.-Q.; Sosso, D.; Julius, B.T.; Lin, I.; Qu, X.-Q.; Braun, D.M.; Frommer, W.B. SWEETs, transporters for intracellular and intercellular sugar translocation. Curr. Opin. Plant Biol. 2015, 25, 53–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Font i Forcada, C.; Oraguzie, N.; Igartua, E.; Moreno, M.; Gogorcena, Y. Population structure and marker–trait associations for pomological traits in peach and nectarine cultivars. Tree Genet. Genomes 2012, 9, 331–349. [Google Scholar] [CrossRef] [Green Version]
  30. Font i Forcada, C.; Oraguzie, N.; Reyes-Chin-Wo, S.; Espiau, M.T.; Socias i Company, R.; Fernandez i Marti, A. Identification of Genetic Loci Associated with Quality Traits in Almond via Association Mapping. PLoS ONE 2015, 10, e0127656. [Google Scholar] [CrossRef] [Green Version]
  31. Amyotte, B.; Bowen, A.J.; Banks, T.; Rajcan, I.; Somers, D.J. Mapping the sensory perception of apple using descriptive sensory evaluation in a genome wide association study. PLoS ONE 2017, 12, e0171710. [Google Scholar] [CrossRef]
  32. Gutierrez, B.L.; Arro, J.; Zhong, G.-Y.; Brown, S.K. Linkage and association analysis of dihydrochalcones phloridzin, sieboldin, and trilobatin in Malus. Tree Genet. Genomes 2018, 14, 91. [Google Scholar] [CrossRef]
  33. Larsen, B.; Migicovsky, Z.; Jeppesen, A.A.; Gardner, K.M.; Toldam-Andersen, T.B.; Myles, S.; Ørgaard, M.; Petersen, M.A.; Pedersen, C. Genome-Wide Association Studies in Apple Reveal Loci for Aroma Volatiles, Sugar Composition, and Harvest Date. Plant Genome 2019, 12, 180104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. McClure, K.A.; Gardner, K.M.; Douglas, G.M.; Song, J.; Forney, C.F.; DeLong, J.; Fan, L.; Du, L.; Toivonen, P.M.; Somers, D.J.; et al. A Genome-Wide Association Study of Apple Quality and Scab Resistance. Plant Genome 2018, 11, 170075. [Google Scholar] [CrossRef] [Green Version]
  35. McClure, K.A.; Gong, Y.; Song, J.; Vinqvist-Tymchuk, M.; Palmer, L.C.; Fan, L.; Burgher-MacLellan, K.; Zhang, Z.; Celton, J.-M.; Forney, C.F.; et al. Genome-wide association studies in apple reveal loci of large effect controlling apple polyphenols. Hortic. Res. 2019, 6, 107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Font i Forcada, C.; Reig, G.; Giménez, R.; Mignard, P.; Mestre, L.; Moreno, M. Sugars and organic acids profile and antioxidant compounds of nectarine fruits influenced by different rootstocks. Sci. Hortic. 2019, 248, 145–153. [Google Scholar] [CrossRef]
  37. Singleton, V.L.; Rossi, J.A. Colorimetry of total phenolics with phosphomolybdic-phosphotungsticacid reagents. Am. J. Enol. Vitic. 1965, 16, 144–158. [Google Scholar]
  38. Zhishen, J.; Mengcheng, T.; Jianming, W. The determination of flavonoid contents in mulberry and their scavenging effects on superoxide radicals. Food Chem. 1999, 64, 555–559. [Google Scholar] [CrossRef]
  39. Brand-Williams, W.; Cuvelier, M.E.; Berset, C. Use of a free radical method to evaluate antioxidant activity. LWT Food Sci. Technol. 1995, 28, 25–30. [Google Scholar] [CrossRef]
  40. Zaharieva, T.B.; Abadía, J. Iron deficiency enhances the levels of ascorbate, glutathione, and related enzymes in sugar beet roots. Protoplasma 2003, 221, 269–275. [Google Scholar] [CrossRef]
  41. Lateur, M.; Ordidge, J.; Engels, J.; Lipman, E. Report of a Working Group on Malus/Pyrus: Fourth Meeting, 7–9 March 2012, Weggis, Switzerland; Bioversity International: Rome, Italy, 2013. [Google Scholar]
  42. R Core Team. R: A Language and Environment for Statistical Computing; R Foundation for Statistical Computing: Vienna, Austria, 2019; Available online: https://www.R-project.org/ (accessed on 15 December 2022).
  43. Kimura, M.; Crow, J.F. The Number of Alleles that Can Be Maintained in a Finite Population. Genetics 1964, 49, 725–738. [Google Scholar] [CrossRef]
  44. Meirmans, P.G. genodiveversion 3.0: Easy-to-use software for the analysis of genetic data of diploids and polyploids. Mol. Ecol. Resour. 2020, 20, 1126–1131. [Google Scholar] [CrossRef] [Green Version]
  45. Nei, M. Analysis of Gene Diversity in Subdivided Populations. Proc. Natl. Acad. Sci. USA 1973, 70, 3321–3323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Lewontin, R.C. The Apportionment of Human Diversity. In Evolutionary Biology; Springer: New York, NY, USA, 1972; pp. 381–398. [Google Scholar] [CrossRef]
  47. Yeh, F.C.; Yang, R.C.; Boyle, T.B.; Ye, Z.H.; Mao, J.X. POPGENE, The User-Friendly Shareware for Population Genetic Analysis; Molecular Biology and Biotechnology Center, University of Alberta: Edmonton, AB, Canada, 1997; Volume 10, pp. 295–301. [Google Scholar]
  48. Pritchard, J.K.; Stephens, M.; Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 2000, 155, 945–959. [Google Scholar] [CrossRef]
  49. Evanno, G.; Regnaut, S.; Goudet, J. Detecting the number of clusters of individuals using the software structure: A simulation study. Mol. Ecol. 2005, 14, 2611–2620. [Google Scholar] [CrossRef] [Green Version]
  50. Kalinowski, S.T.; Taper, M.L.; Marshall, T.C. Revising how the computer program cervus accommodates genotyping error increases success in paternity assignment. Mol. Ecol. 2007, 16, 1099–1106. [Google Scholar] [CrossRef]
  51. Flint-Garcia, S.A.; Thornsberry, J.M.; Buckler, E.S. Structure of Linkage Disequilibrium in Plants. Annu. Rev. Plant Biol. 2003, 54, 357–374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Yu, J.; Buckler, E.S. Genetic association mapping and genome organization of maize. Curr. Opin. Biotechnol. 2006, 17, 155–160. [Google Scholar] [CrossRef] [PubMed]
  53. Aprea, E.; Charles, M.; Endrizzi, I.; Corollaro, M.L.; Betta, E.; Biasioli, F.; Gasperi, F. Sweet taste in apple: The role of sorbitol, individual sugars, organic acids and volatile compounds. Sci. Rep. 2017, 7, 44950. [Google Scholar] [CrossRef] [Green Version]
  54. Castel, L.; Pina, A.; Irisarri, P.; Errea, P. Sugar content and organic acid profiles of local apple cultivars recovered from mountain zones. J. Appl. Bot. Food Qual. 2020, 93, 217–224. [Google Scholar] [CrossRef]
  55. Donahue, D.; Córdoba, G.R.; Elone, S.; Wallis, A.; Basedow, M. ‘Honeycrisp’ Bitter Pit Response to Rootstock and Region under Eastern New York Climatic Conditions. Plants 2021, 10, 983. [Google Scholar] [CrossRef]
  56. Kistechok, A.; Wrona, D.; Krupa, T. Quality and Nutritional Value of ‘Chopin’ and Clone ‘JB’ in Relation to Popular Apples Growing in Poland. Agriculture 2022, 12, 1876. [Google Scholar] [CrossRef]
  57. Slatnar, A.; Kwiecinska, I.; Licznar-Malanczuk, M.; Veberic, R. The effect of green cover within rows on the qualitative and quantitative fruit parameters of full-cropping apple trees. Hortic. Environ. Biotechnol. 2019, 61, 41–49. [Google Scholar] [CrossRef]
  58. Yang, S.; Meng, Z.; Li, Y.; Chen, R.; Yang, Y.; Zhao, Z. Evaluation of Physiological Characteristics, Soluble Sugars, Organic Acids and Volatile Compounds in ‘Orin’ Apples (Malus domestica) at Different Ripening Stages. Molecules 2021, 26, 807. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, L.; Wang, C.; Jia, R.; Yang, N.; Jin, L.; Zhu, L.; Ma, B.; Yao, Y.-X.; Ma, F.; Li, M. Malate metabolism mediated by the cytoplasmic malate dehydrogenase gene MdcyMDH affects sucrose synthesis in apple fruit. Hortic. Res. 2022, 9, uhac194. [Google Scholar] [CrossRef] [PubMed]
  60. Preti, R.; Tarola, A.M. Study of polyphenols, antioxidant capacity and minerals for the valorisation of ancient apple cultivars from Northeast Italy. Eur. Food Res. Technol. 2020, 247, 273–283. [Google Scholar] [CrossRef]
  61. Raudone, L.; Raudonis, R.; Liaudanskas, M.; Janulis, V.; Viskelis, P. Phenolic antioxidant profiles in the whole fruit, flesh and peel of apple cultivars grown in Lithuania. Sci. Hortic. 2017, 216, 186–192. [Google Scholar] [CrossRef]
  62. Pandey, V.P.; Awasthi, M.; Singh, S.; Tiwari, S.; Dwivedi, U.N. A Comprehensive Review on Function and Application of Plant Peroxidases. Biochem. Anal. Biochem. 2017, 6, 308. [Google Scholar] [CrossRef]
  63. Ruan, Y.L. Sucrose metabolism: Gateway to diverse carbon use and sugar signaling. Annu. Rev. Plant Biol. 2014, 65, 33–67. [Google Scholar] [CrossRef]
  64. Vallarino, J.G.; Osorio, S. Organic acids. In Postharvest Physiology and Biochemistry of Fruits and Vegetables; Woodhead Publishing: Cambridge, UK, 2019; pp. 207–224. [Google Scholar] [CrossRef]
  65. Li, M.; Feng, F.; Cheng, L. Expression Patterns of Genes Involved in Sugar Metabolism and Accumulation during Apple Fruit Development. PLoS ONE 2012, 7, e33055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Swarup, S.; Cargill, E.J.; Crosby, K.; Flagel, L.; Kniskern, J.; Glenn, K.C. Genetic diversity is indispensable for plant breeding to improve crops. Crop. Sci. 2020, 61, 839–852. [Google Scholar] [CrossRef]
  67. Bakɪr, M.; Dumanoglu, H.; Aygun, A.; Erdogan, V.; Dost, S.E.; Gülsen, O.; Serdar, U.; Kalkisim, O.; Bastas, K. Genetic diversity and population structure of apple germplasm from Eastern Black Sea region of Turkey by SSRs. Sci. Hortic. 2021, 294, 110793. [Google Scholar] [CrossRef]
  68. Ferreira, V.; Ramos-Cabrer, A.M.; Carnide, V.; Pinto-Carnide, O.; Assunção, A.; Marreiros, A.; Rodrigues, R.; Pereira-Lorenzo, S.; Castro, I. Genetic pool structure of local apple cultivars from Portugal assessed by microsatellites. Tree Genet. Genomes 2016, 12, 36. [Google Scholar] [CrossRef]
  69. Khachtib, Y.; Zinelabidine, L.H.; Bouda, S.; Hamdali, H.; Hammada, S.; Haddioui, A. Genetic characterization of cultivated apple (Malus × domestica Borkh.) in Morocco using microsatellite (SSR) markers. Ecol. Genet. Genom. 2022, 23, 100122. [Google Scholar] [CrossRef]
  70. Van Treuren, R.; Kemp, H.; Ernsting, G.; Jongejans, B.; Houtman, H.; Visser, L. Microsatellite genotyping of apple (Malus × domestica Borkh.) genetic resources in the Netherlands: Application in collection management and variety identification. Genet. Resour. Crop. Evol. 2010, 57, 853–865. [Google Scholar] [CrossRef] [Green Version]
  71. Herrero, J.; Cambra, M.; Tabuenca, M.C. Cartografía de Frutales de Hueso y Pepita; Consta de 12 Tomos, más Uno por Cada Provincia Española; Dpto. de Pomología, Estación Experimental de Aula Dei (CSIC): Zaragoza, Spain, 1964. [Google Scholar]
  72. Bühlmann, A.; Gassmann, J.; Ingenfeld, A.; Hunziker, K.; Kellerhals, M.; Frey, J.E. Molecular Characterisation of the Swiss Fruit Genetic Resources. Erwerbs-Obstbau 2015, 57, 29–34. [Google Scholar] [CrossRef]
  73. Patzak, J.; Paprštein, F.; Henychová, A.; Sedlák, J. Comparison of genetic diversity structure analyses of SSR molecular marker data within apple (Malus × domestica) genetic resources. Genome 2012, 55, 647–665. [Google Scholar] [CrossRef]
  74. Meland, M.; Aksic, M.F.; Frøynes, O.; Konjic, A.; Lasic, L.; Pojskic, N.; Gasi, F. Genetic Identity and Diversity of Apple Accessions within a Candidate Collection for the Norwegian National Clonal Germplasm Repository. Horticulturae 2022, 8, 630. [Google Scholar] [CrossRef]
  75. Jung, M.; Roth, M.; Aranzana, M.J.; Auwerkerken, A.; Bink, M.; Denancé, C.; Dujak, C.; Durel, C.-E.; i Forcada, C.F.; Cantin, C.M.; et al. The apple REFPOP—A reference population for genomics-assisted breeding in apple. Hortic. Res. 2020, 7, 189. [Google Scholar] [CrossRef]
  76. Jung, M.; Keller, B.; Roth, M.; Aranzana, M.J.; Auwerkerken, A.; Guerra, W.; Al-Rifaï, M.; Lewandowski, M.; Sanin, N.; Ry-menants, M.; et al. Genetic architecture and genomic predictive ability of apple quantitative traits across environments. Hortic. Res. 2022, 9, uhac028. [Google Scholar] [CrossRef]
  77. Gross, B.L.; Henk, A.D.; Richards, C.M.; Fazio, G.; Volk, G.M. Genetic diversity in Malus × domestica (Rosaceae) through time in response to domestication. Am. J. Bot. 2014, 101, 1770–1779. [Google Scholar] [CrossRef]
  78. Wedger, M.J.; Schumann, A.C.; Gross, B.L. Candidate genes and signatures of directional selection on fruit quality traits during apple domestication. Am. J. Bot. 2021, 108, 616–627. [Google Scholar] [CrossRef]
  79. Li, M.; Guo, J.; He, J.; Xu, C.; Li, J.; Mi, C.; Tao, S. Possible impact of climate change on apple yield in Northwest China. Theor. Appl. Clim. 2019, 139, 191–203. [Google Scholar] [CrossRef] [Green Version]
  80. Cevik, V.; Ryder, C.D.; Popovich, A.; Manning, K.; King, G.J.; Seymour, G.B. A FRUITFULL-like gene is associated with genetic variation for fruit flesh firmness in apple (Malus domestica Borkh.). Tree Genet. Genomes 2009, 6, 271–279. [Google Scholar] [CrossRef]
  81. Tsykun, T.; Rellstab, C.; Dutech, C.; Sipos, G.; Prospero, S. Comparative assessment of SSR and SNP markers for inferring the population genetic structure of the common fungus Armillaria cepistipes. Heredity 2017, 119, 371–380. [Google Scholar] [CrossRef] [Green Version]
  82. Guichoux, E.; Lagache, L.; Wagner, S.; Chaumeil, P.; Léger, P.; Lepais, O.; Lepoittevin, C.; Malausa, T.; Revardel, E.; Salin, F.; et al. Current trends in microsatellite genotyping. Mol. Ecol. Resour. 2011, 11, 591–611. [Google Scholar] [CrossRef]
  83. Chagné, D.; Krieger, C.; Rassam, M.; Sullivan, M.; Fraser, J.; André, C.; Pindo, M.; Troggio, M.; Gardiner, S.E.; Henry, R.A.; et al. QTL and candidate gene mapping for polyphenolic composition in apple fruit. BMC Plant Biol. 2012, 12, 12. [Google Scholar] [CrossRef]
  84. Guan, Y.; Peace, C.; Rudell, D.; Verma, S.; Evans, K. QTLs detected for individual sugars and soluble solids content in apple. Mol. Breed. 2015, 35, 135. [Google Scholar] [CrossRef]
  85. Howard, N.P.; Tillman, J.; Vanderzande, S.; Luby, J.J. Genetics of zonal leaf chlorosis and genetic linkage to a major gene regulating skin anthocyanin production (MdMYB1) in the apple (Malus × domestica) cultivar Honeycrisp. PLoS ONE 2019, 14, e0210611. [Google Scholar] [CrossRef]
  86. Kenis, K.; Keulemans, J.; Davey, M.W. Identification and stability of QTLs for fruit quality traits in apple. Tree Genet. Genomes 2008, 4, 647–661. [Google Scholar] [CrossRef]
  87. Kunihisa, M.; Moriya, S.; Abe, K.; Okada, K.; Haji, T.; Hayashi, T.; Kim, H.; Nishitani, C.; Terakami, S.; Yamamoto, T. Identification of QTLs for fruit quality traits in Japanese apples: QTLs for early ripening are tightly related to preharvest fruit drop. Breed. Sci. 2014, 64, 240–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Sun, R.; Chang, Y.; Yang, F.; Wang, Y.; Li, H.; Zhao, Y.; Chen, D.; Wu, T.; Zhang, X.; Han, Z. A dense SNP genetic map constructed using restriction site-associated DNA sequencing enables detection of QTLs controlling apple fruit quality. BMC Genom. 2015, 16, 747. [Google Scholar] [CrossRef] [Green Version]
  89. Zhang, B.; Han, Y. Genomics of Fruit Acidity and Sugar Content in Apple. In The Apple Genome; Springer: Cham, Switzerland, 2021; pp. 297–309. [Google Scholar]
  90. Oh, S.; Ahn, S.; Han, H.; Kim, K.; Kim, S.A.; Kim, D. Genetic linkage maps and QTLs associated with fruit skin color and acidity in apple (Malus × domestica). Hortic. Environ. Biotechnol. 2023, 1–12. [Google Scholar] [CrossRef]
  91. Celton, J.-M.; Martinez, S.; Jammes, M.-J.; Bechti, A.; Salvi, S.; Legave, J.-M.; Costes, E. Deciphering the genetic determinism of bud phenology in apple progenies: A new insight into chilling and heat requirement effects on flowering dates and positional candidate genes. New Phytol. 2011, 192, 378–392. [Google Scholar] [CrossRef] [Green Version]
  92. Chagné, D.; Dayatilake, D.; Diack, R.; Oliver, M.; Ireland, H.; Watson, A.E.; Gardiner, S.; Johnston, J.W.; Schaffer, R.J.; Tustin, S. Genetic and environmental control of fruit maturation, dry matter and firmness in apple (Malus × domestica Borkh.). Hortic. Res. 2014, 1, 14046. [Google Scholar] [CrossRef] [Green Version]
  93. Liu, G.-S.; Li, H.-L.; Grierson, D.; Fu, D.-Q. NAC Transcription Factor Family Regulation of Fruit Ripening and Quality: A Review. Cells 2022, 11, 525. [Google Scholar] [CrossRef]
  94. Zhu, M.; Chen, G.; Zhou, S.; Tu, Y.; Wang, Y.; Dong, T.; Hu, Z. A New Tomato NAC (NAM/ATAF1/2/CUC2) Transcription Factor, SlNAC4, Functions as a Positive Regulator of Fruit Ripening and Carotenoid Accumulation. Plant Cell Physiol. 2013, 55, 119–135. [Google Scholar] [CrossRef] [Green Version]
  95. Nieuwenhuizen, N.J.; Chen, X.; Wang, M.Y.; Matich, A.J.; Perez, R.L.; Allan, A.C.; Green, S.A.; Atkinson, R.G. Natural Variation in Monoterpene Synthesis in Kiwifruit: Transcriptional Regulation of Terpene Synthases by NAC and ETHYLENE-INSENSITIVE3-Like Transcription Factors. Plant Physiol. 2015, 167, 1243–1258. [Google Scholar] [CrossRef] [Green Version]
  96. Hyun, T.K.; Eom, S.H.; Kim, J.S. Genomic analysis and gene structure of the two invertase families in the domesticated apple (Malus domestica Borkh). Plant Omics 2011, 4, 391–399. [Google Scholar]
  97. Gharghani, A.; Zamani, Z.; Talaie, A.; Oraguzie, N.C.; Fatahi, R.; Hajnajari, H.; Wiedow, C.; Gardiner, S.E. Genetic identity and relationships of Iranian apple (Malus × domestica Borkh.) cultivars and landraces, wild Malus species and representative old apple cultivars based on simple sequence repeat (SSR) marker analysis. Genet. Resour. Crop. Evol. 2009, 56, 829–842. [Google Scholar] [CrossRef]
  98. Raja, W.H.; Yousuf, N.; Qureshi, I.; Sharma, O.C.; Singh, D.B.; Kumawat, K.L.; Nabi, S.U.; Mir, J.I.; Sheikh, M.A.; Kirmani, S.N.; et al. Morpho-molecular characterization and genetic diversity analysis across wild apple (Malus baccata) accessions using simple sequence repeat markers. South Afr. J. Bot. 2021, 145, 378–385. [Google Scholar] [CrossRef]
Figure 1. Phenotypic fruit diversity encompassed by the 186 accessions of the germplasm bank established at the EEAD-CSIC, Zaragoza, Spain. (a) ‘Reineta Gris’; (b) ‘Reineta Blanca del Canadá’; (c) ‘Granny Smith’; (d) ‘Baujade’; (e) ‘Morro de Liebre’; (f) ‘Golden Paradise’; (g) ‘Cripps Pink’; (h) ‘Reneta’; (i) ‘Solafuente’; (j) ‘Evasni’; (k) 15 different accessions from the EEAD-CSIC germplasm bank exhibiting differences in color and shape.
Figure 1. Phenotypic fruit diversity encompassed by the 186 accessions of the germplasm bank established at the EEAD-CSIC, Zaragoza, Spain. (a) ‘Reineta Gris’; (b) ‘Reineta Blanca del Canadá’; (c) ‘Granny Smith’; (d) ‘Baujade’; (e) ‘Morro de Liebre’; (f) ‘Golden Paradise’; (g) ‘Cripps Pink’; (h) ‘Reneta’; (i) ‘Solafuente’; (j) ‘Evasni’; (k) 15 different accessions from the EEAD-CSIC germplasm bank exhibiting differences in color and shape.
Plants 12 01249 g001
Figure 2. Pearson’s bivariate correlations among the different traits assessed over the 118 accessions phenotyped (Pop4). Abbreviations: SSC, soluble solids content; TA, titratable acidity; RI, ripening index; RAC, relative antioxidant content; TPC, total phenolics content; TFC, total flavonoids content; AsA, Ascorbic acid; Sugars, sum of individual sugars; Acids, sum of individual organic acids. The size of the circle for each correlation and the colour depicts the significance and the magnitude of the correlation coefficient, respectively.
Figure 2. Pearson’s bivariate correlations among the different traits assessed over the 118 accessions phenotyped (Pop4). Abbreviations: SSC, soluble solids content; TA, titratable acidity; RI, ripening index; RAC, relative antioxidant content; TPC, total phenolics content; TFC, total flavonoids content; AsA, Ascorbic acid; Sugars, sum of individual sugars; Acids, sum of individual organic acids. The size of the circle for each correlation and the colour depicts the significance and the magnitude of the correlation coefficient, respectively.
Plants 12 01249 g002
Figure 3. Estimation of the Malus × domestica Borkh collection using LnP (D) derived ΔK for K from 1 to 5 based on: (a) Pop2 (150 diploids apple accessions); (b) Pop3 (36 triploids); and (c) Pop4 (118 diploids) and STRUCTURE bar plots based on: (A) Pop2 at K = 2; (B) Pop3 at K = 2 and K = 4; and (C) Pop4 at K = 2, sorting by subpopulation (Spanish/Non-Spanish) and the coefficient of membership (Q).
Figure 3. Estimation of the Malus × domestica Borkh collection using LnP (D) derived ΔK for K from 1 to 5 based on: (a) Pop2 (150 diploids apple accessions); (b) Pop3 (36 triploids); and (c) Pop4 (118 diploids) and STRUCTURE bar plots based on: (A) Pop2 at K = 2; (B) Pop3 at K = 2 and K = 4; and (C) Pop4 at K = 2, sorting by subpopulation (Spanish/Non-Spanish) and the coefficient of membership (Q).
Plants 12 01249 g003
Figure 4. Pheatmap of 118 apple accessions (Pop4) based on pairwise genetic distances with 23 SSR markers and fruit biochemical characteristics as basic fruit quality (soluble solids content, SSC; titratable acidity, TA; ripening index, RI), antioxidants (total phenolics content, TPC; total flavonoids content, TFC; vitamin C-AsA), individual sugars (glucose, fructose, sucrose and sorbitol), and organic acids (malate, quinate, citrate, tartrate, oxalate, succinate, and shikimate). Annotations as group (n = 6 groups), harvest date, peel color and origin (Spanish/Non-Spanish) for each accession are shown.
Figure 4. Pheatmap of 118 apple accessions (Pop4) based on pairwise genetic distances with 23 SSR markers and fruit biochemical characteristics as basic fruit quality (soluble solids content, SSC; titratable acidity, TA; ripening index, RI), antioxidants (total phenolics content, TPC; total flavonoids content, TFC; vitamin C-AsA), individual sugars (glucose, fructose, sucrose and sorbitol), and organic acids (malate, quinate, citrate, tartrate, oxalate, succinate, and shikimate). Annotations as group (n = 6 groups), harvest date, peel color and origin (Spanish/Non-Spanish) for each accession are shown.
Plants 12 01249 g004
Figure 5. Dendrogram of Pop4, corresponding to 118 diploids apple accessions, based on pairwise genetic distances according to 23 SSR markers and STRUCTURE bar plot at K = 2, optimum number of K subpopulations for Pop4. Annotations as group (n = 6 groups) and origin (Spanish/Non-Spanish) for each accession are shown.
Figure 5. Dendrogram of Pop4, corresponding to 118 diploids apple accessions, based on pairwise genetic distances according to 23 SSR markers and STRUCTURE bar plot at K = 2, optimum number of K subpopulations for Pop4. Annotations as group (n = 6 groups) and origin (Spanish/Non-Spanish) for each accession are shown.
Plants 12 01249 g005
Figure 6. Inter-chromosomal linkage disequilibrium (LD) plot based on 23 SSR markers screened among Pop4 (118 apple accessions). The r2 values are shown in the upper right of the plot and the p-values are represented in lower left of the plot.
Figure 6. Inter-chromosomal linkage disequilibrium (LD) plot based on 23 SSR markers screened among Pop4 (118 apple accessions). The r2 values are shown in the upper right of the plot and the p-values are represented in lower left of the plot.
Plants 12 01249 g006
Table 1. Basic information of the 186 apple accessions (Pop1) assessed in this study.
Table 1. Basic information of the 186 apple accessions (Pop1) assessed in this study.
AccessionEEAD CodePloidyOrigin
1Aciprés *33392nSpanish
2Akane *29022nnon-Spanish
3Almenar-2 *35552nSpanish
4Ascara-134233nSpanish
5Ascara-2 *34242nSpanish
6Astrakan Red *33782nnon-Spanish
7Audiena de Oroz *33752nSpanish
8Augüenta *33352nSpanish
9Averdal-1 *8820212nnon-Spanish
10Averdal-28923402nnon-Spanish
11Baujade *9232842nnon-Spanish
12Belgolden31932nnon-Spanish
13Bellaguarda Lardero *35472nSpanish
14Belleza de Roma *6382nnon-Spanish
15Biscarri-1 *37262nSpanish
16Blackjon *26902nnon-Spanish
17Blacktayman24903nnon-Spanish
18Bofla *34182nSpanish
19Boluaga33403nSpanish
20Boskoop Rouge28983nnon-Spanish
21Bossost-136263nSpanish
22Bossost-236273nSpanish
23Bossost-4 *36292nSpanish
24Bost Kantoia *33412nSpanish
25Cabdellà-2 *36132nSpanish
26Cabello de Angel *32552nSpanish
27Calvilla de San Salvador *33422nSpanish
28Camosa-1 *35532nSpanish
29Camosa-2 *36202nSpanish
30Camuesa de Daroca *33712nSpanish
31Camuesa de Llobregat *13422nSpanish
32Camuesa Fina de Aragón *33722nSpanish
33Carapanón36343nSpanish
34Carrió36363nSpanish
35Cella *25122nSpanish
36Cepiland8819672nnon-Spanish
37Charden 3033nnon-Spanish
38Ciri Blanc *34022nSpanish
39Cirio *36152nSpanish
40Cox’s Orange Pippin *28892nnon-Spanish
41Cripps Pink *9335402nnon-Spanish
42Crispin30803nnon-Spanish
43Cuallarga *34672nSpanish
44Cul de Cirio *35512nSpanish
45De Agosto *36192nSpanish
46De Pera *34162nSpanish
47De Valdés *36322nSpanish
48Delbar Estivale32622nnon-Spanish
49Delciri *34132nSpanish
50Delcon *28962nnon-Spanish
51Delgared Infel *9027082nnon-Spanish
52Deljeni *8513052nnon-Spanish
53Delkistar *9232732nnon-Spanish
54Delorgue Festival *9130442nnon-Spanish
55Democrat28923nnon-Spanish
56Elista *9128832nnon-Spanish
57Esperiega *34202nSpanish
58Esperiega de Olba *37252nSpanish
59Eugenia *34682nSpanish
60Evasni (Scarlet Spur)9335542nnon-Spanish
61Florina *36332nnon-Spanish
62Fortuna Delicious27022nnon-Spanish
63Freyberg26112nnon-Spanish
64Fuji *34882nnon-Spanish
65Fukutami28952nnon-Spanish
66Gala *31972nnon-Spanish
67Galaxy *8924512nnon-Spanish
68Gloster 6931402nnon-Spanish
69Golden Auvil Spur24022nnon-Spanish
70Golden Delicious *6752nnon-Spanish
71Golden Delicious Infel *24912nnon-Spanish
72Golden Paradise *37392nnon-Spanish
73Golden Smoothee *32862nnon-Spanish
74Granny Smith-1 *31962nnon-Spanish
75Granny Smith-2 *26142nnon-Spanish
76Gravenstein31093nnon-Spanish
77Guillemes *34112nSpanish
78Hared *8922322nnon-Spanish
79Harrold Red Delicious28992nnon-Spanish
80Helada *33682nSpanish
81Hierro *33742nSpanish
82Idared *24842nnon-Spanish
83Irgo-2 *36222nSpanish
84Jerseymac31412nnon-Spanish
85Jonadel *26502nnon-Spanish
86Jonagored8820013nnon-Spanish
87Jonathan-1 *24952nnon-Spanish
88Jonathan-2 *30962nnon-Spanish
89Jubilee *8513042nnon-Spanish
90Kidd’s Orange Red28882nnon-Spanish
91Kinrei29002nnon-Spanish
92Lancer8819682nnon-Spanish
93Landetxo *33432nSpanish
94Les-1 *36242nSpanish
95Les-236253nSpanish
96MacIntosh *31922nnon-Spanish
97Mañaga-1 *4692nSpanish
98Mañaga-2 *35542nSpanish
99Marinera *34122nSpanish
100Marquiñez34193nSpanish
101Médulas-1 *35482nSpanish
102Melrose *24842nnon-Spanish
103Merrigold *8513072nnon-Spanish
104Montcada-1 *36312nSpanish
105Morro de Liebre *32562nSpanish
106Mutsu24873nnon-Spanish
107Nesple *34102nSpanish
108Normanda32523nSpanish
109Nueva Starking *18992nnon-Spanish
110Ortell-14133nSpanish
111Ortell-2 *35462nSpanish
112Ozark Gold31752nnon-Spanish
113Pera 2 *34172nSpanish
114Pera de Sangüesa33793nSpanish
115Pero Pardo33693nSpanish
116Peromingan *11582nSpanish
117Peruco de Caparroso *33732nSpanish
118Plaona *9232832nnon-Spanish
119Poma de San Juan *35562nSpanish
120Prau Riu-3 *34912nSpanish
121Prau Riu-434923nSpanish
122Prau Riu-5 *34932nSpanish
123Prima *8513062nnon-Spanish
124Prime Gold31982nnon-Spanish
125Rebellón *33702nSpanish
126Red Delicious *30852nnon-Spanish
127Red Elstar8820022nnon-Spanish
128Red King Delicious26882nnon-Spanish
129Red Rome Beauty *28972nnon-Spanish
130Redaphough *9334112nnon-Spanish
131RedChief *8513082nnon-Spanish
132Redspur Delicious30822nnon-Spanish
133Regal Prince-18820222nnon-Spanish
134Regal Prince-2 *8923412nnon-Spanish
135Reguard-2 *36172nSpanish
136Reguard-4 *36182nSpanish
137Reina de Reinetas24883nnon-Spanish
138Reineta Blanca del Canadá-13083nnon-Spanish
139Reineta Blanca del Canadá-231113nnon-Spanish
140Reineta Blanca del Canadá-331943nnon-Spanish
141Reineta Encarnada *36352nSpanish
142Reineta Gris28833nnon-Spanish
143Reineta Inesita Asua25433nSpanish
144Reineta Regil34663nSpanish
145Reneta *34082nSpanish
146Richared Delicious24812nnon-Spanish
147Roja Valle de Benejama *10382nSpanish
148Roser de la Reula *35522nSpanish
149Royal Red Delicious23632nnon-Spanish
150Rubinete *8615262nnon-Spanish
151Ruixou-1 *36142nSpanish
152San Felipe *33762nSpanish
153San Miguel *25792nSpanish
154Sandía *33362nSpanish
155Sant Jaume34703nSpanish
156Sant Joan *34092nSpanish
157Santa Margarida34013nSpanish
158Shelred28932nnon-Spanish
159Signatillis *34032nSpanish
160Solafuente 35593nSpanish
161Spartan24832nnon-Spanish
162Starking-1 *29642nnon-Spanish
163Starking-2 *6322nnon-Spanish
164Starkrimson-1 *31952nnon-Spanish
165Starkrimson-219042nnon-Spanish
166Stayman Waynesap31103nnon-Spanish
167Taüll-1 *36232nSpanish
168Telamon *33982nnon-Spanish
169Tempera *33342nSpanish
170Terrera34693nSpanish
171Topred Delicious26512nnon-Spanish
172Totxa *34712nSpanish
173Trajan33962nnon-Spanish
174Transparente *33772nSpanish
175Transparente Blanca *33442nSpanish
176Turley Winnesap28843nnon-Spanish
177Tuscan33972nnon-Spanish
178Urarte34153nSpanish
179Urtebete *33452nSpanish
180Valsaina *35582nSpanish
181Vance Delicious26472nnon-Spanish
182Verde Doncella-1 *21252nSpanish
183Verde Doncella-23102nSpanish
184Verde Doncella-3 *35492nSpanish
185Vinçada Tardía36213nSpanish
186Wellspur Delicious30812nnon-Spanish
* Accessions assessed for the association study.
Table 2. Characteristics of the 26 SSR markers used in this study with indication of the corresponding multiplex and dye.
Table 2. Characteristics of the 26 SSR markers used in this study with indication of the corresponding multiplex and dye.
LocusLGMultiplexDyeSize Range (bp)Forward Primer Sequence
(5′→3′)
Reverse Primer Sequence
(5′→3′)
Primer ConcentrationReference
CH-Vf11MP5VIC130–171ATCACCACCAGCAGCAAAGCATACAAATCAAAGCACAACCC[0.1 µM][1]
Hi02c071MP3VIC98–146AGAGCTACGGGGATCCAAATGTTTAAGCATCCCGATTGAAAGG[0.1 µM][1]
CH02c062MP3PET201–261TGACGAAATCCACTACTAATGCAGATTGCGCGCTTTTTAACAT[0.4 µM][1]
GD123MP3NED139–189TTGAGGTGTTTCTCCCATTGGACTAACGAAGCCGCCATTTCTTT[0.1 µM][1]
MdSWEET2a3MP6VIC331–357ATACCGAGGAACTGTAGGACCAAGCCTCCACACTAAACAACCAGAAAGCA[0.1 µM][27]
MdSWEET9b4MP46-FAM336–360GCGCCAATGTAAGACCCTTTACTTTCTGACCTTGTCCTTCTTGGATGCGTA[0.1 µM][27]
CH05f065MP2NED161–189TTAGATCCGGTCACTCTCCACTTGGAGGAAGACGAAGAAGAAAG[0.1 µM][1]
MdSWEET2d5MP6PET265–289CATTCAATTTATTCGACCGGACGACTGGGTTCATCCCTCACTTTCACTCA[0.1 µM][27]
MdSWEET7b6MP4VIC230–267GGGTTTTGAGAATCTTGAGGGTAGGTTTGATGGGTTGGACTGTAACTTGC[0.1 µM][27]
CH03d076MP3VICn.a.CAAATCAATGCAAAACTGTCAGGCTTCTGGCCATGATTTTA[0.1 µM][1]
CH04e057MP1PET163–228AGGCTAACAGAAATGTGGTTTGATGGCTCCTATTGCCATCAT[0.1 µM][1]
CH01h108MP4PET81–120TGCAAAGATAGGTAGATATATGCCAAGGAGGGATTGTTTGTGCAC[0.1 µM][1]
CH01f03b9MP5NED127–177GAGAAGCAAATGCAAAACCCCTCCCCGGCTCCTATTCTAC[0.1 µM][1]
CH01h029MP1NEDn.a.AGAGCTTCGAGCTTCGTTTGATCTTTTGGTGCTCCCACAC[0.1 µM][1]
CH02c1110MP2PET208–238TGAAGGCAATCACTCTGTGCTTCCGAGAATCCTCTTCGAC[0.15 µM][1]
MdSWEET2e10MP4NED205–243GTGAGCCCACAACTAATCCCATCTTGTGCGTAGGAATCCCGATA[0.1 µM][27]
CH02d0811MP1VIC196–256TCCAAAATGGCGTACCTCTCGCAGACACTCACTCACTATCTCTC[0.1 µM][1]
MdSWEET2b11MP56-FAM249–263TGAGGCAGAAACAATCATAAGGGTCGAGCACGGAATTTGAAGCTGTAAAA[0.1 µM][27]
MdSWEET7a11MP5PET340–376TTCTATCTCCCCTTCCCAAACTTCCGCTAAACAGTGCCACTGCATAAGGT[0.1 µM][27]
CH01f0212MP16-FAM155–212ACCACATTAGAGCAGTTGAGGCTGGTTTGTTTTCCTCCAGC[0.1 µM][1]
GD14713MP3PET124–158TCCCGCCATTTCTCTGCGTTTAAACCGCTGCTGCTGAAC[0.1 µM][1]
CH04c0714MP2VIC93–139GGCCTTCCATGTCTCAGAAGCCTCATGCCCTCCACTAACA[0.1 µM][1]
MdSWEET12a14MP6NED223–253ATGACAGGGCAACTTCAGGGTCGTAATAGTCCTTTGCCCTCC[0.1 µM][27]
CH02c0915MP2VIC203–254TTATGTACCAACTTTGCTAACCTCAGAAGCAGCAGAGGAGGATG[0.1 µM][1]
CH04f1016MP36-FAMn.a.GTAATGGAAATACAGTTTCACAATTAAATGCTTGGTGTGTTTTGC[0.1 µM][1]
CH01h0117MP26-FAM92–130GAAAGACTTGCAGTGGGAGCGGAGTGGGTTTGAGAAGGTT[0.1 µM][1]
Abbreviations: n.a. = no or weak amplification.
Table 3. Basic statistics of phenotypical traits over the Pop4 (118 diploid accessions) during the 2014–2018 period: units, minimum, maximum, mean values, standard deviation, and standard error of the mean.
Table 3. Basic statistics of phenotypical traits over the Pop4 (118 diploid accessions) during the 2014–2018 period: units, minimum, maximum, mean values, standard deviation, and standard error of the mean.
TraitUnitsMinimumMaximumMeanSDSEANOVA
SSC°Brix10.1417.0313.401.360.47***
TAg malic acid L−11.7717.296.612.921.35***
RI-0.768.552.621.341.45***
TPCmg GAE 100 g FW−115.2498.0739.5415.760.08***
TFCmg CE 100 g FW−16.0088.9522.8414.670.12***
AsAmg AsA 100 g FW−11.375.302.830.820.27***
RACmg Trolox 100 g FW−15.9330.8215.445.080.12***
Sucroseg.kg−110.2942.1425.527.010.64***
Glucoseg.kg−16.2324.2913.174.120.38***
Fructoseg.kg−131.3961.4145.555.170.48***
Sorbitolg.kg−11.2011.964.552.430.22***
Sugarsg.kg−163.69115.2788.769.870.91***
Oxalicg.kg−10.01360.01760.01470.00060.0001***
Citricg.kg−10.01780.14820.05560.02790.0026***
Tartaricg.kg−10.02600.09100.04800.01460.0013***
Malicg.kg−12.68319.81445.51821.69300.1559***
Quinicg.kg−10.23500.80050.42310.11510.0106***
Succinic + Shikimicg.kg−10.19481.67660.55700.25200.0232***
Acidsg.kg−13.424211.38896.60851.82180.1677***
ANOVA: significant differences at p ≤ 0.001 (***) among the different apple accessions. Abbreviations: SSC, soluble solids content; TA, titratable acidity; RI, ripening index; TPC, total phenolics content; TFC, total flavonoids content; AsA, Ascorbic acid; RAC, relative antioxidant content; Sugars, sum of individual sugars; Acids, sum of organic acids.
Table 4. Mean estimated values for different genetic parameters of the 186 apple accessions (Pop1) based on 23 SSRs loci.
Table 4. Mean estimated values for different genetic parameters of the 186 apple accessions (Pop1) based on 23 SSRs loci.
SSRNANENBHoHeFis
CH-Vf1133.50100.750.73−0.03
Hi02c0793.4040.680.700.03
CH02c06218.54140.930.90−0.03
GD12172.85130.710.69−0.03
MdSWEET2a124.2880.830.81−0.02
MdSWEET9b92.5350.660.63−0.05
CH05f06145.7970.860.920.07
MdSWEET2d134.8880.440.810.46
MdSWEET7b155.1790.960.87−0.10
CH04e05303.47260.670.730.08
CH01h10183.23140.780.72−0.08
CH01f03b155.24110.920.82−0.12
CH02c11158.7250.890.900.01
MdSWEET2e121.50100.190.410.54
CH02d08258.98180.850.900.06
MdSWEET2b41.9620.510.50−0.02
MdSWEET7a123.6180.720.740.03
CH01f02358.60290.810.900.10
GD147154.2990.830.78−0.06
CH04c07197.06110.920.91−0.01
MdSWEET12a82.4240.600.58−0.03
CH02c09136.9650.930.87−0.07
CH01h01165.80100.750.830.10
Mean-186 accessions (Pop1)15.654.9010.430.750.770.03
Mean diploids-150 accessions (Pop2)14.705.239.220.790.77−0.03
Mean triploids-36 accessions (Pop3)10.004.665.260.710.730.03
NA: observed number of alleles per locus; NE: effective number of alleles per locus; NB: number of rare alleles; Ho: observed heterozygosity; He: expected heterozygosity, Fis: Wright’s fixation index; I: Shannon’s information index; Pop: population.
Table 5. Significance (p-value) of association between 23 SSRs polymorphic loci and biochemical traits in 118 apple accessions (Pop4).
Table 5. Significance (p-value) of association between 23 SSRs polymorphic loci and biochemical traits in 118 apple accessions (Pop4).
AgronomicalAntioxidantsBasic Fruit QualityIndividual SugarsIndividual Organic Acids
LGMarker NameHarvestPeel ColorRACTFCTPCAsASSCTARISucGluFru 1SorSugOxaCitTarMalQuiSucc-ShiAcids
1CH-Vf1 *** * **
1Hi02c07 **** * *
2CH02c06 ***** ****** ********* ****
3GD12 1
3MdSWEET2a **** * ******** *
4MdSWEET9b * * ** *
5CH05f06* ******** ***** * ************
5MdSWEET2d* * ** ** *** **
6MdSWEET7b ******** * *
7CH04e05 ** ** * ***
8CH01h10 **
9CH01f03b* *** ** ********
10CH02c11 **** **
10MdSWEET2e 1
11CH02d08 1
11MdSWEET2b * ***** *
11MdSWEET7a * * *** * **
12CH01f02 * ** **** *** * *
13GD147 **
14CH04c07* ***** ***** * ** ****
14MdSWEET12a*** ** * *****
15CH02c09 * * **
17CH01h01 * * ** **
1 No association found for this SSR or trait. Statistical significance at *: p ≤ 0.01; **: p ≤ 0.001; ***: p ≤ 0.0001; Abbreviations: RAC, relative antioxidant content; TFC, total flavonoids content; TPC, total phenolics content; AsA, Ascorbic acid; SSC, soluble solids content; TA, titratable acidity; RI, ripening index; Suc, sucrose; Glu, glucose; Fru, fructose; Sor, sorbitol; Sug, total sugars; Oxa, oxalate; Cit, citrate; Tar, tartrate; Mal, malate; Qui, Quinate; Succ-Shi, succinate and shikimate; Acids, total organic acids.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mignard, P.; Font i Forcada, C.; Giménez, R.; Moreno, M.Á. Population Structure and Association Mapping for Agronomical and Biochemical Traits of a Large Spanish Apple Germplasm. Plants 2023, 12, 1249. https://doi.org/10.3390/plants12061249

AMA Style

Mignard P, Font i Forcada C, Giménez R, Moreno MÁ. Population Structure and Association Mapping for Agronomical and Biochemical Traits of a Large Spanish Apple Germplasm. Plants. 2023; 12(6):1249. https://doi.org/10.3390/plants12061249

Chicago/Turabian Style

Mignard, Pierre, Carolina Font i Forcada, Rosa Giménez, and María Ángeles Moreno. 2023. "Population Structure and Association Mapping for Agronomical and Biochemical Traits of a Large Spanish Apple Germplasm" Plants 12, no. 6: 1249. https://doi.org/10.3390/plants12061249

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