Skip to main content

Genetic diversity of peanut (Arachis hypogaeaL.) and its wild relatives based on the analysis of hypervariable regions of the genome

Abstract

Background

The genusArachisis native to a region that includes Central Brazil and neighboring countries. Little is known about the genetic variability of the Brazilian cultivated peanut (Arachis hypogaea, genome AABB) germplasm collection at the DNA level. The understanding of the genetic diversity of cultivated and wild species of peanut (Arachisspp.) is essential to develop strategies of collection, conservation and use of the germplasm in variety development. The identity of the ancestor progenitor species of cultivated peanut has also been of great interest. Several species have been suggested as putative AA and BB genome donors to allotetraploidA. hypogaea.Microsatellite or SSR (Simple Sequence Repeat) markers are co-dominant, multiallelic, and highly polymorphic genetic markers, appropriate for genetic diversity studies. Microsatellite markers may also, to some extent, support phylogenetic inferences. Here we report the use of a set of microsatellite markers, including newly developed ones, for phylogenetic inferences and the analysis of genetic variation of accessions ofA. hypogeaand its wild relatives.

Results

A total of 67 new microsatellite markers (mainly TTG motif) were developed forArachis.Only three of these markers, however, were polymorphic in cultivated peanut. These three new markers plus five other markers characterized previously were evaluated for number of alleles per locus and gene diversity using 60 accessions ofA. hypogaea.Genetic relationships among these 60 accessions and a sample of 36 wild accessions representative of sectionArachiswere estimated using allelic variation observed in a selected set of 12 SSR markers. Results showed that the Brazilian peanut germplasm collection has considerable levels of genetic diversity detected by SSR markers. Similarity groups forA. hypogaeaaccessions were established, which is a useful criteria for selecting parental plants for crop improvement. Microsatellite marker transferability was up to 76% for species of the sectionArachis, but only 45% for species from the other eightArachis部分测试。一个新的标记(ah - 041)提出了一个100% transferability and could be used to classify the peanut accessions in AA and non-AA genome carriers.

Conclusion

The level of polymorphism observed among accessions ofA. hypogaeaanalyzed with newly developed microsatellite markers was low, corroborating the accumulated data which show that cultivated peanut presents a relatively reduced variation at the DNA level. A selected panel of SSR markers allowed the classification ofA. hypogaeaaccessions into two major groups. The identification of similarity groups will be useful for the selection of parental plants to be used in breeding programs. Marker transferability is relatively high between accessions of sectionArachis.The possibility of using microsatellite markers developed for one species in genetic evaluation of other species greatly reduces the cost of the analysis, since the development of microsatellite markers is still expensive and time consuming. The SSR markers developed in this study could be very useful for genetic analysis of wild species ofArachis, including comparative genome mapping, population genetic structure and phylogenetic inferences among species.

Background

The genusArachis, family Fabaceae, is native to South America, probably from a region including Central Brazil and Paraguay [1]. There are 69 species described in the genus, assembled into nine sections, according to the morphology, geographic distribution and crossability [2]. Some of these species have been used for forage in South America, but the most important species is the cultivated peanut,Arachis hypogaeaL. This crop is widely grown in more than 80 countries in the Americas, Asia and Africa [3] and is used for both human consumption and as a source of oil. World production is increasing and has reached 37 million tons (in the shell) and 5.8 million tons of oil, according to FAO estimates [4]. Peanut now ranks fifth among the world oil production crops.

Nearly allArachisspecies are diploid, but cultivated peanut is an allotetraploid (genome AABB). It is a member of the sectionArachis, which also includes about 25 diploid and one tetraploid wild species (A. monticola) [2].Arachis hypogaeais classified based on the presence or absence of flowers on the main axis into two subspecies,hypogaeaandfastigiata.These two subspecies were further classified into six botanical varieties based on morphology and growth habits [2]. Subspecieshypogaeawas divided in two botanical varieties,hypogaeaandhirsuta, while ssp.fastigiatain the varietiesfastigiata,vulgaris,aequatorianaandperuviana.The identity of the progenitor species of cultivated peanut has been of great interest. Several species have been suggested as putative A and B genome donors [reviewed by [57]]. RFLP analysis that included 17 diploid species of the sectionArachisand threeA. hypogaeaaccessions suggested a single origin for domesticated peanut and ancestral species related toA. duranensis(A genome) andA. ipaënsis(B genome) as the most likely progenitors ofA. hypogaea[5].In situhybridization analysis of six diploid species and oneA. hypogaeaaccession, and RAPD and ISSR (Inter-simple sequence repeat) analyses of 13A. hypogaeaaccessions and 15 wild species, however, suggestedA. villosa(A genome) andA. ipaënsis(B genome) as the progenitors of cultivated peanut [6,7].

种植花生展览大量of variability for various morphological, physiological, and agronomic traits. However, little variation has been detected at the DNA level using techniques such as RAPDs (Random Amplified Polymorphic DNAs), AFLPs (Amplified Fragment Length Polymorphisms) and RFLPs (Restriction Fragment Length Polymorphisms) [5,815]. The low level of variation in cultivated peanut has been attributed to three causes or to combinations of them: (1) barriers to gene flow from related diploid species to domesticated peanut as a consequence of the polyploidization event [16]; (2) recent polyploidization, from one or a few individual(s) of each diploid parental species, combined with self-pollination [8]; or (3) use of few elite breeding lines and little exotic germplasm in breeding programs, resulting in a narrow genetic base [17,18]. Recently, some studies revealed DNA polymorphism inA. hypogaeausing SSRs (Simple Sequence Repeats), AFLPs, RAPDs and ISSRs [7,1923].

Little is known about the genetic variability of the BrazilianA. hypogaeagermplasm collection at the DNA level. Knowledge of the genetic variation of peanut accessions is important for their efficient use in breeding programs, for studies on crop evolution, and for conservation purposes. Molecular marker analysis, joined to phenotypic evaluation, is a powerful tool for grouping of genotypes based on genetic distance data and for selection of progenitors that might constitute new breeding populations. In the context ofex situconservation, it has been demonstrated that molecular markers are very useful for the management of germplasm collections [24]. SSRs are ideal tools for such studies as they are PCR-based markers, genetically defined, typically co-dominant, multiallelic, and uniformly dispersed throughout plant genomes. SSRs have been used in plants for many genetic applications, including the assessment of genetic variability in germplasm collections or pedigree reconstruction [reviewed by [25]]. InArachis, SSR markers have been recently developed and proved to be useful for accession discrimination and assessment of genetic variation [19,22,23]. Moreover, since little genetic variability has been detected in cultivated peanut, the use of a polymorphic marker, such as SSRs, in addition to distinguishing closely related genotypes, should also be useful for phylogenetic studies, as demonstrated in other crops, such as wheat [26], melon [27], potato [28], and coffee [29].

The objectives of the present work were: (1) to develop new SSR markers for genetic analysis of cultivated peanut (A. hypogaea), (2) to employ a set of SSR markers to analyze the genetic variation among wild and cultivated peanut accessions of the Brazilian germplasm collection, and (3) to evaluate the cross-species transferability of SSR markers and their usefulness in phylogenetic studies of the genusArachis.

Results and Discussion

SSR development and screening of markers

Microsatellite enriched genomic libraries ofA. hypogaeawere constructed in order to develop new SSR markers for the species. Digestion of theA. hypogaeagenomic DNA with five different enzymes (AluI,MseI,RsaI,Sau3AI, andTsp509I) revealed thatTsp509I produced the most adequate profile for library development, with fragments ranging from 200 to 800 bp in size. Four libraries were initially constructed based on trinucleotide repeat motifs (TTG, TGG, ATG, and ATC). Hybridization analysis revealed that the TTG/AAC repeats were more abundant in the peanut genome than the other tested motifs (TTG > TGG > ATG > ATC). This is in agreement with a survey of published DNA sequences in 54 plant species, where the TTG/AAC repeat was one of the most abundant SSRs [30]. Therefore, the TTG library was used for SSR development forA. hypogaea.750年由锚定PCR克隆显示t的筛查hat 162 had SSRs (21.6%) and these were sequenced. Out of the 162 positive clones sequenced, there were 91 unique (non-redundant) sequences (56.2% of the sequenced clones) but only 67 of them were suitable for primer designing (41.4% of the positive clones). The design of primers for the other 24 unique SSRs identified was not possible due to the occurrence of very short tandem repeats (< 5 units) or low GC content of the regions flanking the SSR. The anchored PCR screening prior to sequencing significantly improved the yield of useful clones. Few false-positive clones were obtained (less than 10% of the sequenced clones). The percentage of primers designed, in relation to the number of clones sequenced (41.4%), indicated that the method used was relatively efficient for the discovery and development of SSR markers for peanut.

The 67 SSR markers (seeAdditional File 1) were screened for polymorphism on seven samples, including five varieties ofA. hypogaea, one accession ofA. ipaënsisand one accession ofA. duranensis.这些, 62 markers (92.5%) generated clearly interpretable PCR products, but only three markers (Ah-041, Ah-193 and Ah-558) were polymorphic in cultivated peanut (Table1)。其他四个标记(ah - 075、ah - 229、ah - 522和Ah-657) showed to be invariant for the fiveA. hypogaeaaccessions, but were polymorphic inA. ipaënsisandA. duranensis登记入册。自从祖先物种有关se two species are considered potential progenitors of the AABB genome, these seven markers were included in the analysis of the genetic relationships between accessions of the Brazilian germplasm collection (see below). The other 55 markers were not polymorphic inA. hypogeaaccessions and were not considered for further analysis. They have been useful, however, for studies with diploid wild species of peanut currently under development (data not shown).

Table 1 SSR loci characterization Primer pairs, repeat motifs, range of fragment sizes, total number of alleles (A), and gene diversity (h) estimates based on the analysis of 60Arachis hypogaeaaccessions for the eight polymorphic SSR markers.

SSR marker characterization

Marker loci duplication

Four (Ah-193, Ah-229, Ah-522, and Ah-657) of the seven newly developed markers produced single fragments both in the tetraploid accessions and in the diploid species belonging to theArachissection (seeAdditional File 2), what indicates that each of them amplified alleles at single loci. Two markers (Ah-075 and Ah-558) produced one or two fragments in the diploid species and three (Ah-075) or four (Ah-558) fragments in the tetraploid accessions. These results are consistent with allele amplification on duplicated loci. The presence of the three alleles (150/144/138 bp) amplified by marker Ah-075 in the tetraploid accessions was confirmed by the identification of diploid species homozygous for each of these three alleles (for example, accessions K 30006, Sv 3806, V 6389, V 10309) or heterozygous for the three pairs of alleles (for example, accessions K 30076, K 30097, V 14167). Similar results were observed for marker Ah-558, but one (244 bp) of the six alleles (244/241/235/232/229/226 bp) detected in the tetraploid accessions was not found in any of the tested diploid wild species. The remaining marker (Ah-041) amplified one (292 bp) or two of three alleles (300/292/280 bp) in accessions of cultivated peanut and in the diploid species (Figure1)。This indicates that either the individuals tested are highly heterozygous for this locus or this marker locus is duplicated in the cultivated peanut genome. SinceA. hypogaeais an allotetraploid and preferentially an autogamous species, with only 2.5% outcrossing [31], the latter hypothesis is more probable. Five other markers previously developed [19] were also analyzed for marker locus duplication, and two of them (Ah4-04 and Ah4-24) seem to represent a single locus in the genome. Therefore, five out of 12 markers (41,7%) probably amplified duplicated loci in most of the cultivated peanut accessions. This proportion, although not representative, indicates that locus duplication in peanut might be higher than found for other polyploid species, such as wheat [32], apple [33], cassava [34], and sweet potato [35]. The inclusion of hybrids and parental diploid plants in the study contributed to a better understanding of the genetic basis of these marker loci. Markers that amplified three or four alleles in tetraploid accessions amplified consistently only two alleles in the interspecific hybrids. In depth analyses of peanut segregant populations should elucidate the inheritance of these SSR loci.

Figure 1
figure1

Microsatellite polymorphism (marker Ah-041) inArachisspecies visualized in silver-stained denaturing polyacrilamide gel.Samples are: (1)A. duranensis-V14167; (2)A. ipaënsis; (3)A. magna-V13760; (4)A. batizocoi-K9484; (5)A. cardenasii; (6)A. stenosperma-V10229; (7)A. magna-K30097; (8)A. helodes; (9)A. hoehnei; (10)A. batizocoi-K9484m; (11)A. villosa; (12)A. microsperma; (13)A. simpsonii; (14)A. monticola; (15)A. hypogaea fastigiata fastigiata(16)A. hypogaea fastigiata vulgaris; (17)A. hypogaea fastigiata peruviana;(18)A. hypogaea hypogaea hypogaea;(19)A. hypogaea hypogaea hypogaea,Xingu type;(20)A. stenosperma-V10309; (21–23) Hybrids K7988 × V10309; (24)A. duranensis-K7988; (25)A. diogoi; (26)A. duranensis-G10038; (27)A. kempf-mercadoi; (28)Arachissp. 6389; (29)A. cruziana; (30)A. praecox; (31)A. valida

Marker allelic diversity

Allelic diversity for a sample of 60A. hypogaeaaccessions (Additional File 2) was estimated for the three newly developed polymorphic markers (Ah-041, Ah-193, Ah-558) and five other SSR markers (Ah4-04, Ah4-20, Ah4-24, Ah4-26 e Lec-1), previously developed for cultivated peanut [19] (Table1)。Marker Ah6-125 [19] was not included in the analysis since it was found to be invariant for the 60 accessions of cultivated peanut. The number of alleles detected ranged from 2 to 27 at each of the eight polymorphic loci, with an average of 8.4 alleles per locus. Gene diversity values ranged from 0.468 to 0.929, with an average value of 0.683. For polyploid organisms, gene diversity is an estimate of the probability that two randomly chosen genes from a population are different and it is not related to the expected proportion of heterozygous, as for diploid organisms in Hardy-Weinberg equilibrium [36]. The values obtained in the present study are probably overestimated, since some of the markers amplified alleles on duplicated loci. Thus, they cannot be compared to those obtained for diploid species, but they are similar to values found for other polyploid species such as cassava [34], potato [37] and wheat [38]. Moreover, averagehvalues estimated by RFLP markers were considerably lower forA. hypogaea[10]. These results demonstrate the usefulness of microsatellite markers for genetic diversity analysis of cultivated peanut.

Marker Ah-041 and genome specific alleles

One of the markers developed in the present study (Ah-041) presented interesting banding patterns (Figure1) on AA and non-AA genomes of diploid wild species. Genome AA is characterized by the presence of a pair of chromosomes (called AA) that are considerably smaller than the other chromosomes [39], while genome BB lacks this small chromosome pair. All AA genome species, previously characterized by cytological analysis, showed a 292 bp allele when genotyped at the Ah-041 locus, while all BB genome species andA. glandulifera(DD genome) showed a 280 bp allele.Arachis hypogaeaaccessions showed both alleles (292/280) or they had only the 292 bp allele.Arachis monticola, the other tetraploid species of sectionArachis, had also both 292 and 280 bp alleles. Accessions of five AA genome species(A. stenosperma,A. villosa,A. microsperma,A. simpsoniiandA. kempff-mercadoi), one BB genome species (A. valida) andA. glanduliferashowed a 300 bp band. Marker Ah-041 allowed allele detection on all the 114 accessions tested, with a 100% transferability in the genus. Species from the sectionsErectoides,Extranervosae, andTriseminataehad different alleles (277 and 274 bp). Allele 280 bp was also found in species from the sectionsTrierectoides,Heteranthae,Procumbentes, andRhizomatosae.Only sectionArachisspecies showed the allele 292 bp (AA genome). Although allelic segregation and genome mapping of this marker locus has not yet been examined in controlled crosses, our results suggest that the 292 bp allele is specific to AA genome species. If so, this marker may be valuable for genetic studies of peanut, including phylogenetic inferences in the genusArachisand studies about genome origin (see below). It should be noticed that the absolute transferability of marker Ah-041 to allArachisspecies suggests that this SSR sequence is positioned in or near coding regions. However, search on the National Center for Biotechnology Information (NCBI) databasehttp://www.ncbi.nlm.nih.gov/BLAST/did not find any significantly complete or partially homologous sequence.

Transferability of SSR markers

Twelve SSR markers developed forArachis hypogaea(Ah-041, Ah-193, Ah-558, Ah-075, Ah-229, Ah-522, Ah-657, Ah4-04, Ah4-20, Ah4-24, Ah4-26 and Lec-1) were tested on 54 accessions of wild species ofArachis, representing the nine sections of the genus. The transferability of the twelve markers was up to 76% for species of the sectionArachis, but ranged from 23% (Triseminatae) to 62% (Procumbentes), with an average of 45%, for species of the other eightArachissections (seeAdditional File 3)。Transferability of SSR markers between related species is a consequence of the homology of flanking regions of the microsatellites and the size of the region between the primer pair amenable to amplification by PCR. Other studies have demonstrated the conservation of SSR sequences in plants, as reviewed by Gupta and Varshney [25]. The possibility of using microsatellite markers developed for one species in genetic evaluation of other species greatly reduces the cost of the analysis, since the development of microsatellite markers is still expensive and time consuming. The SSR markers developed in this study could be very useful for genetic analysis of wild species ofArachis, including comparative genome mapping, population genetic structure and phylogenetic inferences among species.

Relationships betweenArachisspp. accessions

Genetic similarities among accessions were estimated by shared allele distance in pairwise comparisons of 60A. hypogaeaaccessions and 36 other accessions of wild species belonging to sectionArachis(accessions 1 to 96 –Additional File 2), using the same set of 12 markers described above. Among the 60A. hypogaeaaccessions, the average genetic distance was 0.336. Some accessions, such as Tatu, Tatu2 e Pd2622, shared all the alleles detected with the 12 SSR markers.

A dendrogram based on the neighbor joining method was constructed for the 96 accessions of the sectionArachis(Figure2)。Considering initially theA. hypogaeaaccessions, two main clusters were evident. Group I contained all 32fastigiata/fastigiata, all fourfastigiata/vulgarisaccessions, the onlyhypogaea/hirsutaaccession (Mf 1538) and one accession not identified to the subspecies level (Wi 632). The twofastigiata/aequatorianaaccessions included in the study (Mf 1678 and Mf1640) were also clustered in this group. The twofastigiata/peruvianaaccessions (Mf 1560 e Sv 429) formed a separate subgroup within Group I. Another subgroup was formed by afastigiata/aequatorianaaccession (Mf 1640), two variant forms ofhypogaea/hypogaeaaccessions (Nambiquarae, accession As 436 and Malhado, accession V 12577) and V 12549, identified as "ssphypogaeaXingu type", because of its morphological differences from the Virginia type. All 45 accessions of Group I had the alleles 292 bp and 280 bp at the Ah-041 locus. The other group (Group II) contained solely the 15 Virginia type (hypogaea/hypogaea) accessions examined in the study. The accessions of this group had only the 292 bp allele (Figure1)。在ah - 041位点多态性可能是原因d by mutation at the primer binding site in the Group II accessions, preventing the detection of the 280 bp PCR product, or by an insertion/deletion event. If the Ah-041 locus is duplicated in allotetraploid peanut, these results could indicate that cultivated peanut did not have a single origin since Group I accessions could be derived from hybridization of AA and non-AA ancestral diploid species followed by genome duplication, while Group II accessions could be derived from straight genome duplication of an AA diploid species. Multiple origin hypotheses, based on cytological, crossing and isozyme studies, have been previously suggested to explain the cultivated peanut origin [4042].

Figure 2
figure2

Dendrogram based on allele sharing genetic distances of 60A. hypogaeaand 36Arachiswild species generated by the neighbor joining method.The letters, after eachA. hypogaeaaccession number, refer to the subspecies, varieties, and types: FF-fastigiata/fastigiata; FV-fastigiata/vulgaris; FA-fastigiata/aequatoriana; FP-fastigiata/peruviana; HH-hypogaea/hypogaea; HHi-hypogaea/hirsuta; HHN-hypogaea/hypogaeaNambiquarae type; HHM-hypogaea/hypogaea"Malhado" type; HX-hypogaea兴谷河类型。倪,对应于一个non-identified accession to the subspecies level. The letters after the wild species accession names refer to the genome type (A or B), according to the patterns obtained with marker Ah-041. The D genome ofA. glanduliferawas defined by Stalker, 1991.

The taxonomic status ofA. hypogaeassp.fastigiatavar.peruvianais controversial. Studies based on AFLP [21] as well as RAPD and ISSR [7] markers have shown that accessions of this variety have greater similarity to ssp.hypogaearather than ssp.fastigiata,to which they are currently thought to belong [2]. In the present study, the twofastigiata/peruvianaaccessions grouped with the otherfastigiataaccessions and not with the 15hypogaea/hypogaeaaccessions clustered in Group II. The data, therefore, corroborates the current classification of this botanical variety. However, the classification ofhypogaea/hirsutaand the three variant forms ofhypogaea/hypogaea(Nambiquarae, Malhado and Xingu types) remains controversial, since these four accessions did not cluster in Group II.

The SSR markers used in the present study detected genetic variability inA. hypogaeaaccessions, despite the known narrow genetic base of cultivated peanut. In the two main clusters (Group I and Group II), some similarity groups, genetically distant from each other, were formed. These groups were not related to collection sites, seed color, or any other known trait. However, the establishment of similarity groups, jointly with morphologic and agronomic data, can be used in the selection and crossing of superior plants, allowing for a more efficient use of the existing genetic variability conserved in the collection.

Considering the wild species, the analysis showed that, in general, the species containing the same genome clustered together. However, accessions ofA. duranensisclustered withA. hypogaeaandA. monticolaaccessions (Figure2)。This result corroborates the hypothesis that an ancestral species related toA. duranensisorA. duranensisitself could be the donor of the AA genome toA. hypogaea.In the group composed of BB genome species there is a cluster of four species (A. ipaënsis,A. magna,A. williamsiiandArachissp. 6389). It has recently been noticed that plants of these four species are morphologically very similar and produce fertile hybrids. A revision of their taxonomic status is currently under evaluation by JFM Valls and CE Simpson.

The genetic distance analysis based on SSR marker polymorphism supports previous estimates of high genetic similarity betweenA. monticolaaccessions andA. hypogaeaaccessions (Figure2)。只有一个(ah - 522) 12个SSR位点的萤火虫ed in the present study showed alleles detected inA. hypogaeanot observed inA. monticola.Other studies with molecular data have also shown the little differentiation between these two taxa [8,9,11,42]. Moreover, hybrids betweenA. monticolaandA. hypogaeaare fertile [2]. These facts have led some researchers to suggest thatA. monticolaandA. hypogaeashould not be considered as separate species [11]. Our results, based on SSR markers and only one accession ofA. monticola, are certainly insufficient to define the taxonomic status ofA. monticola.However, they indicated that this species could be directly related to the allotetraploid ancestral progenitor ofA. hypogaea, as suggested by several studies [6,7,43].

Conclusions

The present study showed the utility of microsatellite markers for the detection of polymorphisms among cultivated peanut accessions and for the genetic relationship analysis betweenA. hypogaeaaccessions and wild species of sectionArachis.New SSR markers were developed for genetic analysis of peanut, but the percentage of markers that detected polymorphism among accessions ofA. hypogaeawas low, corroborating the notion that cultivated peanut presents a relatively reduced variation at the DNA level. However, a set of informative SSR markers detected considerable levels of genetic variability in the Brazilian Peanut Germplasm Collection. Based on this data, cultivated peanut was classified into two major groups, one composed of solely the Virginia type (hypogaea/hypogaea) accessions and another defined byfastigiata/fastigiata, fastigiata/vulgaris, hypogaea/hirsuta, fastigiata/aequatoriana,fastigiata/peruvianaand two variant forms ofhypogaea/hypogaea登记入册。The identification of similarity groups could be useful for the selection of parental plants to be used in the breeding programs. Marker transferability was up to 76% for species of the sectionArachis, but ranged from 23% (Triseminatae) to 62% (Procumbentes), with an average of 45%, for species of the other eightArachissections. The possibility of using microsatellite markers developed for one species in genetic evaluation of other species greatly reduces the cost of the analysis, since the development of microsatellite markers is still expensive and time consuming. The SSR markers developed in this study could be very useful for genetic analysis of wild species of Arachis, including comparative genome mapping, population genetic structure and phylogenetic inferences among species. A marker (Ah-041) was identified that allows the discrimination of AA from non-AA genome accessions of peanut.

Methods

Plant material

A singleA. hypogaeaplant was used for DNA library construction (accession UF 91108). FiveA. hypogaeasamples (accessions 3274, 3294, 3310, 3324, and UF 91108), oneA. ipaënsis(accession 3309) and oneA. duranensis(accession 4454) were used to test the effectiveness of SSR markers to detect polymorphism at the DNA level. These plants were obtained from USDA-ARS Plant Genetic Resources Conservation Unit (Griffin, GA) peanut collection.

A total of 114 accessions of wild and cultivated peanut were included in the study (Additional File 2)。Ninety-six accessions belonging to sectionArachis被用于遗传多样性分析。这些, 60 areA. hypogaea在巴西登记入册,47岁的收集,representing bothA. hypogaeassp.hypogaeaand ssp.fastigiata, and their six varieties. The other 36 accessions belong to 27 wild species, including representatives of four new taxa that are being described by JFM Valls and CE Simpson. The wild species include tetraploidA. monticola(AABB), 13 diploid (2n = 20) species showing the small chromosome pair [44] which are associated to the AA genome ofA. hypogaea, the diploidA. glandulifera, for which a DD genome has been suggested [45], two diploids with 2n = 18 [46] and ten species with 20 chromosomes lacking the small chromosome pair typical of the AA genome, which are classified as BB genome species (Additional File 2)。Other 18 accessions of wild species belonging to eightArachissections were included in the analysis to test the transferability of SSR markers. Three hybrid plants of theA. stenosperma(V 10309) ×A. duranensis(K 7988) cross and one hybrid plant of theA. appressipila(G 10002) ×A. repens(Nc 1579) cross were also used to verify the heritability of alleles in interspecific crosses. The 114 accessions and the four hybrid plants were obtained from the Brazilian Peanut Germplasm Collection, maintained at Embrapa Genetic Resources and Biotechnology – CENARGEN (Brasília-DF, Brazil). All plants were grown from seed or cuttings (sectionsRhizomatosaeandCaulorrhizae) under greenhouse conditions at CENARGEN prior to DNA extraction.

DNA extraction

Total genomic DNA used for library construction was extracted and purified on cesium chloride gradients following standard protocols [47]. For the other samples, total genomic DNA was extracted according to published protocol [48], with some modifications: Proteinase K (20 mg/ml) was added to the extraction buffer and an additional step of polysaccharide precipitation with 2 M NaCl was included.

Library construction

Total genomic DNA libraries enriched for trinucleotide repeats were constructed following a standard protocol with minor modifications [49]. Peanut total genomic DNA was digested with five different enzymes (AluI,MseI,RsaI,Sau3AI, andTsp509I) to identify which one produced the most adequate fragment profile for library development. Fragments ranging in size from 200 to 800 bp were transferred to S&S NA 45 DEAE cellulose membranes, after electrophoresis in 1.5% low-melting agarose gel. DNA was resuspended in TE buffer and ligated toTsp509 I adapters. Four genomic libraries enriched for trinucleotide repeats were constructed. Fragments were selected by hybridization with biotinylated trinucleotide repeats ((TTG)6, (TGG)6, (ATG)6, (ATC)6) and recovered by magnetic beads linked to streptavidine. Enriched fractions were amplified by PCR, using a primer complementary to the adapters. Amplification products were purified using Wizard PCR Preps DNA purification system (Promega, Madison, WI), cloned into λ ZAP II phagemid vector (Stratagene) and then transformed intoE. colistrain XL1-Blue. Transformed cells were grown on LB-Amp plates (50 μg/ml ampicillin) at 37°C, until plaques were between 0.5 and 1.0 mm in diameter. Individual clones were picked up from the plates and phage were eluted in SM buffer (500 μl) with a drop of chloroform. Clone DNAs were then amplified by PCR using primers complementary to the vector DNA flanking the insert (T3 or SK and T7 or KS primers, Stratagene, CA), and electrophoresed in 1.5% agarose gels in 0.5X TBE buffer to determine insert sizes. To determine the position of the SSRs relative to the vector cloning site, an anchored PCR strategy was performed for each positive clone [49], using T3 (or SK) and T7 (or KS) primers separately plus another primer complementary to the target SSR.

DNA sequencing and primer design

Clones containing repeats that were not directly adjacent to the cloning site were amplified, treated withExoI and shrimp alkaline phosphatase enzymes, and PCR products were sequenced on an ABI 377 automated DNA sequencer (Applied Biosystems, CA, USA), using fluorescent dye-terminator chemistry. Redundant sequences were identified using the software Sequencher (Gene Codes Corporation). Primers complementary to unique DNA sequences flanking the SSRs were designed using the computer program Primer 3 (Whitehead Institute of Biomedical Research –http://www-genome.wi.mit.edu/cgi-bin/primer/primer3_www.cgi)。Primer design parameters were as follows: (i) Tmranged from 60°C to 65°C; (ii) 1°C difference in Tmbetween primer pairs; (iii) GC content ranging from 45% to 55%, (iv) absence of complementarity between primers, (v) default values for the other parameters.

Primer screening and PCR amplification of SSR loci

Primer pairs were initially screened for polymorphisms in a set of five cultivated and two wild peanut accessions (3274, 3294, 3310, 3324, UF 91108, 3309, 4454). PCRs were performed in 13 μl volumes, containing 1X PCR buffer (10 mM Tris-HCl pH8.3, 50 mM KCl, 1.5 mM MgCl2), 0.2 mM of each dNTP, 1 unitTaqDNA polymerase, DMSO 50% (1.3 μl), 5 pmol of each primer and 10 ng of genomic DNA. Amplifications were performed using either a 9600 System (Applied Biosystems, CA, USA) or a PTC-100 (MJ Research, MA, USA) thermal cycler, with the following conditions: 96°C for 2 min (1 cycle), 94°C for 1 min, 55–66°C for 1 min, 72°C for 1 min (30 cycles); and 72°C for 7 min (1 cycle). The annealing temperature was optimized for each primer pair to produce clear DNA band amplification, without spurious fragments. PCR products were resolved on 3.5% Metaphor agarose (FMC Bioproducts, ME, USA) gels stained with ethidium bromide. For genotype determination, the amplified products were separated on 4% denaturing polyacrilamide gels stained with silver nitrate [50]. Fragment sizes were estimated by comparison with a 10-bp DNA ladder standard (Gibco/BRL, MD, USA).

Data analysis and transferability of SSR loci

Eight selected SSR markers were characterized for number of alleles per locus and gene diversity [51] using 60 accessions ofArachis hypogaea(accessions 1 to 60 –Additional File 2)。The markers included three developed in the present study and five other SSR markers developed forA. hypogaea[19]. Gene diversity (h) at a marker locus was estimated according to the formula:h= 1 - Σ(pi)2, wherepi2is the frequency of theith allele at this locus [51], using the GDA software [52]. For estimates of genetic distance, 60 accessions of cultivated peanut and 36 accessions from 27 wild species of sectionArachiswere analyzed (accessions 61 to 96 –Additional File 2)。A total of 12 markers were used in this analysis (Ah-041, Ah-193, Ah-558, Ah-075, Ah-229, Ah-522, Ah-657, Ah4-04, Ah4-20, Ah4-24, Ah4-26 e Lec-1). Genetic distances among wild and cultivated peanut accessions based on microsatellite marker polymorphism were estimated by shared allele distance in pairwise comparisons. The estimates were based on the sum of the proportion of common alleles between two peanut accessions examined across loci (PS) divided by twice the number of tested loci [53,54]. Genetic distances were obtained by the parameter [-ln (PS)] using the Genetic Distance Calculator [55]. The diagonal matrix was then submitted to cluster analysis using the neighbor joining method and a genetic distance dendrogram built using the software NTSYS 2.1 [56]. Transferability of the 13 SSR markers (the 12 described above plus marker Ah6-125) was tested using the 36 accessions of sectionArachis(accessions 61 to 96 –Additional File 2) and another 18 wild species accessions belonging to the remainingArachissections (accessions 97 to 114 –Additional File 2)。PCR amplification followed the same protocol used forA. hypogaea.PCR products were visualized on silver-stained 4% denaturing polyacrilamide gels.

References

  1. 1.

    Gregory WC, Krapovickas A, Gregory MP: Structure, variation, evolution, and classification inArachis.In: Advances in Legume Science. Edited by: Summerfield RJ, Bunting AH. Kew, England: Royal Botanical Gardens; 1980, 469-481.

    Google Scholar

  2. 2.

    Krapovickas A, Gregory WC: Taxonomía del géneroArachis(Leguminosae). Bonplandia. 1994, 8: 1-186.

    Google Scholar

  3. 3.

    Singh U, Singh B: Tropical grain legumes as important human foods. Econ Bot. 1992, 46: 310-321.

    ArticleGoogle Scholar

  4. 4.

    FAO (2004). [http://faostat.fao.org/faostat/collections?subset=agriculture].

  5. 5.

    Kochert G, Stalker HT, Gimenes M, Galgaro L, Lopes CR, Moore K: RFLP and cytogenetic evidence on the origin and evolution of allotetraploid domesticated peanut,Arachis hypogaea(Leguminosae). Am J Bot. 1996, 83: 1282-1291.

    CASArticleGoogle Scholar

  6. 6.

    Raina SN, Mukai Y: Genomic in situ hybridization inArachis(Fabaceae) identifies the diploid wild progenitors of cultivated (A. hypogaea) and related wild (A. monticola) peanut species. Pl Syst Evol. 1999, 214: 251-262.

    ArticleGoogle Scholar

  7. 7.

    Raina SN, Rani V, Kojima T, Ogihara Y, Singh KP, Devarumath RM: RAPD and ISSR fingerprints as useful genetic markers for analysis of genetic diversity, varietal identification, and phylogenetic relationships in peanut (Arachis hypogaea) cultivars and wild species. Genome. 2001, 44: 763-772. 10.1139/gen-44-5-763.

    PubMedCASArticleGoogle Scholar

  8. 8.

    Halward TM, Stalker HT, Larue EA, Kochert G: Genetic variation detectable with molecular markers among unadapted germ-plasm resources of cultivated peanut and related wild species. Genome. 1991, 34: 1013-1020.

    CASArticleGoogle Scholar

  9. 9.

    Kochert G, Halward T, Branch WD, Simpson CE: RFLP variability in peanut (Arachis hypogaea) cultivars and wild species. Theor Appl Genet. 1991, 81: 565-570.

    PubMedCASArticleGoogle Scholar

  10. 10.

    Paik-Ro OG, Smith RL, Knauft DA: Restriction fragment length polymorphism evaluation of six peanut species within theArachissection. Theor Appl Genet. 1992, 84: 201-208.

    PubMedCASArticleGoogle Scholar

  11. 11.

    Hilu KW, Stalker HT: Genetic relationships between peanut and wild species ofArachissect.Arachis (Fabaceae): evidence from RAPDs. Pl Syst Evol. 1995, 198: 167-178.

    CASArticleGoogle Scholar

  12. 12.

    He G, Prakash CS: Identification of polymorphic DNA markers in cultivated peanut (Arachis hypogaeaL.). Euphytica. 1997, 97: 143-149. 10.1023/A:1002949813052.

    CASArticleGoogle Scholar

  13. 13.

    Subramanian V, Gurtu S, Nageswara Rao RC, Nigam SN: Identification of DNA polymorphism in cultivated groundnut using random amplified polymorphic DNA (RAPD) assay. Genome. 2000, 43: 656-660. 10.1139/gen-43-4-656.

    PubMedCASArticleGoogle Scholar

  14. 14.

    Gimenes MA, Lopes CR, Valls JFM: Genetic relationships amongArachisspecies based on AFLP. Genet Mol Biol. 2002, 25: 349-353.

    CASArticleGoogle Scholar

  15. 15.

    Herselman L: Genetic variation among Southern African cultivated peanut (A. hypogaeaL.) genotypes as revealed by AFLP analysis. Euphytica. 2003, 133: 319-327. 10.1023/A:1025769212187.

    CASArticleGoogle Scholar

  16. 16.

    Young ND, Weeden NF, Kochert G: Genome mapping in legumes(Family Fabaceae). In: Genome Mapping in Plants Edited by:Paterson AH. Austin, TX: Landes Biomedical Press; 1996:212-227.

    Google Scholar

  17. 17.

    Knauft DA, Gorbet DW: Genetic diversity among peanut cultivars. Crop Sci. 1989, 29: 1417-1422.

    ArticleGoogle Scholar

  18. 18.

    Isleib TG, Wynne JC: Use of plant introductions in peanutimprovement. In: Use of Plant Introductions in Cultivar Development Volume 2. Edited by: Shands HL. Madison: Crop Science Society of America; 1992:75-116.

    Google Scholar

  19. 19.

    Hopkins MS, Casa AM, Wang T, Mitchell SE, Dean R, Kochert GD, Kresovich S: Discovery and Characterization of Polymorphic Simple Sequence Repeats (SSRs) in Peanut. Crop Sci. 1999, 39: 1243-1247.

    CASArticleGoogle Scholar

  20. 20.

    Dwivedi SL, Gurtu S, Chandra S, Yuejin W, Nigam SN: Assessment of genetic diversity among selected groundnut germplasm. I: RAPD analysis. Plant Breeding. 2001, 120: 345-349. 10.1046/j.1439-0523.2001.00613.x.

    CASArticleGoogle Scholar

  21. 21.

    He G, Prakash CS: Evaluation of genetic relationships among botanical varieties of cultivated peanut (Arachis hypogaeaL.) using AFLP markers. Genet Resour Crop Evol. 2001, 48: 347-352. 10.1023/A:1012019600318.

    ArticleGoogle Scholar

  22. 22.

    He G, Meng R, Newman M, Gao G, Pittman RN, Prakash CS: Microsatellites as DNA markers in cultivated peanut (A. hypogaeaL.). BMC Plant Biology. 2003, 3: 3-10.1186/1471-2229-3-3. [http://www.biomedcentral.com/1471-2229/3/3].

    PubMedPubMed CentralArticleGoogle Scholar

  23. 23.

    Ferguson ME, Burow MD, Schulze SR, Bramel PJ, Paterson AH, Kresovich S, Mitchell S: Microsatellite identification and characterization in peanut (A. hypogaeaL.). Theor Appl Genet. 2004, 108: 1064-1070. 10.1007/s00122-003-1535-2.

    PubMedCASArticleGoogle Scholar

  24. 24.

    Westman AL, Kresovich S: Use of molecular markers techniquesfor description of plant genetic variation. In: Plant Genetic Resources: Conservation and Use Edited by: Callow JA, Ford-Lloyd BV, Newbury HJ. Wallingford, UK: CAB International;1997:9-48.

    Google Scholar

  25. 25.

    Gupta PK, Varshney RK: The development and use of microsatellite markers for genetic analysis and plant breeding with emphasis on bread wheat. Euphytica. 2000, 113: 163-185. 10.1023/A:1003910819967.

    CASArticleGoogle Scholar

  26. 26.

    Lelley T, Stachel M, Grausgruber H, Vollmann J: Analysis of relationships betweenAegilops tauschiiand the D genome of wheat utilizing microsatellites. Genome. 2000, 43: 661-668. 10.1139/gen-43-4-661.

    PubMedCASArticleGoogle Scholar

  27. 27.

    Danin-Poleg Y, Reis N, Tzuri G, Katzir N: Development and characterization of microsatellite markers inCucumis.Theor Appl Genet. 2001, 102: 61-72. 10.1007/s001220051618.

    CASArticleGoogle Scholar

  28. 28.

    Ashkenazi V, Chani E, Lavi U, Levy D, Hillel J, Veilleux RE: Development of microsatellite markers in potato and their use in phylogenetic and fingerprinting analyses. Genome. 2001, 44: 50-62. 10.1139/gen-44-1-50.

    PubMedCASArticleGoogle Scholar

  29. 29.

    Anthony F, Combe MC, Astorga C, Bertrand B, Graziosi G, Lashermes P: The origin of cultivatedCoffea arabicaL. varieties revealed by AFLP and SSR markers. Theor Appl Genet. 2002, 104: 894-900. 10.1007/s00122-001-0798-8.

    PubMedCASArticleGoogle Scholar

  30. 30.

    Wang Z, Weber JL, Zhong G, Tanksley SD: Survey of plant short tandem DNA repeats. Theor Appl Genet. 1994, 88: 1-6.

    PubMedCASGoogle Scholar

  31. 31.

    Norden AV: Breeding of the cultivated peanut (Arachis hypogaeaL.). In: Peanuts-cultures and uses. Am Peanut Res Ed Assoc, Stillwater, Oklahoma, 1973:175-208.

    Google Scholar

  32. 32.

    Bryan GJ, Collins AJ, Stephenson P, Orry A, Smith JB, Gale MD: Isolation and characterization of microsatellites from hexaploid bread wheat. Theor Appl Genet. 1997, 94: 557-563. 10.1007/s001220050451.

    CASArticleGoogle Scholar

  33. 33.

    Guilford P, Prakash S, Zhu JM, Rikkerink E, Gardiner S, Basset H, Foster R: Microsatellites inMalus×domestica(apple): abundance, polymorphism and cultivar identification. Theor Appl Genet. 1997, 94: 249-254. 10.1007/s001220050407.

    CASArticleGoogle Scholar

  34. 34.

    Chavarriaga-Aguirre P, Maya MM, Bonierbale MW, Kresovich S, Fregene MA, Tohme J, Kochert G: Microsatellites in cassava (Manihot esculentaCrantz): discovery, inheritance and variability. Theor Appl Genet. 1998, 97: 493-501. 10.1007/s001220050922.

    CASArticleGoogle Scholar

  35. 35.

    Buteler MI, Jarret RL, LaBonte DR: Sequence characterization of microsatellites in diploid and polyploidIpomoea.Theor Appl Genet. 1999, 99: 123-132. 10.1007/s001220051216.

    CASArticleGoogle Scholar

  36. 36.

    Nei M: Molecular Evolutionary Genetics. 1987, New York: Columbia University Press, 512.

    Google Scholar

  37. 37.

    Provan J, Powell W, Waugh R: Microsatellite analysis of relationships within cultivated potato (Solanum tuberosum)。Theor Appl Genet. 1996, 92: 1078-1084. 10.1007/s001220050233.

    PubMedCASArticleGoogle Scholar

  38. 38.

    Prasad M, Varshney RK, Roy JK, Balyan HS, Gupta PK: The use of microsatellites for detecting DNA polymorphism, genotype identification and genetic diversity in wheat. Theor Appl Genet. 2000, 100: 584-592.

    CASGoogle Scholar

  39. 39.

    Husted L: Cytological studies of the peanutArachis.I. Chromosome number and morphology. Cytologia. 1933, 5: 109-117.

    ArticleGoogle Scholar

  40. 40.

    Singh AK, Moss JP: Utilization of wild relatives in genetic improvement ofArachis hypogaeaL. 2.Chromosome complements of species in the sectionArachis.Theor Appl Genet. 1982, 61: 305-314.

    PubMedCASGoogle Scholar

  41. 41.

    Singh AK: Putative genome donors ofArachis hypogaea(Fabaceae), evidence from crosses with synthetic amphidiploids. Plant Syst Evol. 1988, 160: 143-151.

    ArticleGoogle Scholar

  42. 42.

    Lu J, Pickersgill B: Isozyme variation and species relationships in peanut and its wild relatives (ArachisL. – Leguminosae). Theor Appl Genet. 1993, 85: 550-560.

    PubMedCASArticleGoogle Scholar

  43. 43.

    Gregory WC, Gregory MP: Groundnut. In: Evolution of Crop Plants. Edited by: Simmonds NW. 1976, London: Longman, 151-154.

    Google Scholar

  44. 44.

    Fernandez A, Krapovickas A: Cromosomas y evolución enArachis(Leguminosae). Bonplandia. 1994, 8: 187-220.

    Google Scholar

  45. 45.

    Stalker HT: A new species in sectionArachisof peanuts with D genome. Am J Bot. 1991, 78: 630-637.

    ArticleGoogle Scholar

  46. 46.

    Lavia GI: Karyotypes ofArachis palustrisandA.praecox(SectionArachis), two species with basic chromosme number x = 9. Cytologia. 1998, 63: 177-181.

    ArticleGoogle Scholar

  47. 47.

    Reichardt M, Rogers S: Preparation of genomic DNA from plant tissue. In: Current protocols in molecular biology. Edited by: Ausubel FM, Brent R, Kingston RE, Moore DD, Seidman JG, Smith JA, Struhl K. 1997, New York: John Wiley & Sons.

    Google Scholar

  48. 48.

    Grattapaglia D, Sederoff R: Genetic linkage maps ofEucalyptus grandisandEucalyptus urophyllausing a pseudo-testcross: Mapping strategy and RAPD markers. Genetics. 1994, 137: 1121-1137.

    PubMedCASPubMed CentralGoogle Scholar

  49. 49.

    名为《Rafalski JA,傅高义JM Morgante M,鲍威尔W,安德烈C, Tingey SV:Generating and using DNA markers in plants. In: Analysis ofnon-mammalian genomes – a practical guide Edited by: Birren B, Lai E.New York: Academic Press; 1996:75-134.

    Google Scholar

  50. 50.

    Bassam BJ, Caetano-Anolles G, Gresshoff PM: Fast and sensitive silver staining of DNA in polyacrilamide gels. Analyt Biochem. 1991, 196: 80-83.

    PubMedCASArticleGoogle Scholar

  51. 51.

    Nei M: Analysis of gene diversity in subdivided populations. Proc Natl Acad Sci USA. 1973, 70: 3321-3323.

    PubMedCASPubMed CentralArticleGoogle Scholar

  52. 52.

    Lewis PO, Zaykin D: Genetic Data Analysis: Computer program for the analysis of allelic data. Version 1.0 (d16c). 2001, Free program distributed by the authors over the Internet, [http://lewis.eeb.uconn.edu/lewishome/software.html].

    Google Scholar

  53. 53.

    Bowcock AM, Ruiz-Linhares A, Tomfohrde J, Minch E, Kidd JR, Cavalli-Sforza LL: High resolution of human evolutionary trees with polymorphic microsatellites. Nature. 1994, 368: 455-457. 10.1038/368455a0.

    PubMedCASArticleGoogle Scholar

  54. 54.

    Goldstein DB, Ruiz Linares A, Cavalli-Sforza LL, Feldman MW: Genetic absolute dating based on microsatellite and the origin of modern humans. Proc Natl Acad Sci USA. 1995, 92: 6723-6727.

    PubMedCASPubMed CentralArticleGoogle Scholar

  55. 55.

    Genetic Distance Calculator. [http://www2.biology.ualberta.ca/jbrzusto/sharedst.php].

  56. 56.

    Rohlf FJ: NTSYS-PC: numerical taxonomy and multivariate analysis system – version 2.0. 1993, New York: Exeter Software.

    Google Scholar

Download references

Acknowledgements

The authors thank Dr. R. E. Dean for computational assistance and L. Zhang and T. M. Jenkins for their technical laboratory assistance. We also thank M.R. Bertozo for generously providing the DNA samples of some wild species ofArachis.Development of SSR marker research supported by grants from the USDA. Characterization of SSR markers supported by Embrapa Recursos Genéticos e Biotecnologia, Brazil.

Author information

Affiliations

Authors

Corresponding author

Correspondence toMarcio de Carvalho Moretzsohn

Additional information

Authors' contributions

MM participated in the SSR markers development, carried out the analysis of theArachisaccessions and drafted the manuscript. MH participated in the development of the SSR markers. SM and SK coordinated the development of the SSR markers. JV participated in the conception of the project and provided the germplasm. MF participated in the conception of the project, in the data analysis and reviewed the manuscript. All authors read and approved the manuscript.

Electronic supplementary material

List of the 67 microsatellite markers for genetic analysis of

Additional File 1:Arachisspp.Marker sequences, repeat motifs, and annealing temperature (Ta) for the 67 newly developed SSR markers (XLS 24 KB)

spp. accessions analyzed in this study.

Additional File 2:Arachisspp. accessions analyzed in this study.Numbers in parenthesis, after the specific names, correspond to numbers included in the dendrogram (Figure 2). *Genome types ofArachis hypogaeaaccessions were defined according to the genotypes obtained with marker Ah-041. (DOC 170 KB)

Transferability of SSR markers developed for

Additional File 3:A. hypogaeaacross 45 wild species ofArachisThis file contains a table that lists the markers that amplified visible products for each of the 49 accessions of theArachiswild species included in the study. (XLS 27 KB)

Authors’ original submitted files for images

Below are the links to the authors’ original submitted files for images.

Authors’ original file for figure 1

Authors’ original file for figure 2

Rights and permissions

Reprints and Permissions

About this article

Cite this article

Moretzsohn, M.d.C., Hopkins, M.S., Mitchell, S.E.et al.Genetic diversity of peanut (Arachis hypogaeaL.) and its wild relatives based on the analysis of hypervariable regions of the genome.BMC Plant Biol4,11 (2004). https://doi.org/10.1186/1471-2229-4-11

Download citation

Keywords

  • Wild Species
  • Arachis
  • Diploid Species
  • Arachis Species
  • Diploid Wild Species