Hostname: page-component-848d4c4894-ttngx Total loading time: 0 Render date: 2024-05-01T05:55:49.031Z Has data issue: false hasContentIssue false

Genetic Analysis of Prunus salicina L. by Random Amplified Polymorphic DNA (RAPD) and Intersimple Sequence Repeat (ISSR)

Published online by Cambridge University Press:  01 January 2024

Jun Li
Affiliation:
College of Modern Agriculture, Jiaxing Vocational and Technical College, Jiaxing, Zhejiang 314000, China
Guangchun Gao*
Affiliation:
College of Medicine, Jiaxing University, Jiaxing, Zhejiang 314001, China
Bin Li
Affiliation:
Jiaxing City General Station of Cropping Technical Extension, Jiaxing, Zhejiang 314000, China
Bai Li
Affiliation:
Jiaxing Academy of Agricultural Sciences, Jiaxing, Zhejiang 314016, China
Qihua Lu
Affiliation:
Horticultural Crop and Plum Research Institute, Jiaxing, Zhejiang 314016, China
*
Correspondence should be addressed to Guangchun Gao; gaogcjx@163.com
Rights & Permissions [Opens in a new window]

Abstract

Background. Prunus salicina L. is an important fruit tree species of great economic value which is mainly distributed in the northern hemisphere. Methods. 25 samples of Prunus salicina L. were collected from 8 provinces in China, Japan, USA, and New Zealand. The genetic variations of these samples were characterized by the random amplified polymorphic DNA (RAPD) and intersimple sequence repeat (ISSR) technique, respectively, and in combination. Results. Totally, 257 RAPD bands ranging 200∼2300 bp was found, and 81.59% of these bands were polymorphic. ISSR analysis identified 179 bands ranging 300∼2500 bp, and 87.74% of the bands were polymorphic. ISSR results showed that the similarity coefficient index between samples P10 (Maihuangli in Anhui, Chin) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest, while that between samples P10 (Maihuangli in Anhui, Chin) and P15 (Baili in Japan) was highest. Combined analysis of RAPD and ISSR demonstrated that the similarity coefficient index between samples P4 (Qiepili in Ningbo, Zhejiang, China) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest, while that between samples P19 (Laroda in USA) and P20 (Red heart in USA) was highest. Conclusion. RAPD combined with ISSR analysis can be used for genetic characterization of Prunus L. species.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © 2022 Jun Li et al.

1. Background

Prunus salicina L., belonging to the family of Rosaceae, are one of the most important economical fruit trees and are widely cultivated all over the world. They are mainly distributed in the northern hemisphere, especially in the temperate zone [Reference Chen, Chen and Chen1, Reference Herrera, Lora, Hormaza, Herrero and Rodrigo2]. China is one of the origin and distribution centers of Prunus L. species. Prunus L. species contain more than 430 species and are first segregated into six genera according to the morphology of fruit: Amygdalus L., Armeniaca Scop., Cerasus Mill., Laurocerasus, Padus Mill., Prunus species, and Tourn. ex Duh. However, phylogenetic analysis showed that Cerasus, Laurocerasus, and Padus were not monophyletic [Reference Wen, Berggren and Lee3, Reference Williams, Kubelik, Livak, Rafalski and Tingey4]. Besides, an increasing number of new cultivars from different countries result an important renewal of plant material worldwide [Reference Herrera, Lora, Hormaza, Herrero and Rodrigo2]. It is thus necessary to characterize genetic information of Prunus L. species to cultivate new breed with improved quality characteristics.

DNA polymorphism assay based on the amplification of random DNA segments with single primers of arbitrary nucleotide sequence has been widely used for genetic diversity analysis of species [Reference Pairon, Petitpierre and Campbell5]. Several studies have been devoted to the genetic diversity in Prunus L. species [Reference Fu, Li and Xu6Reference Carrasco, González, Gebauer and García-González8]. Recently, a number of molecular marker techniques including random amplified polymorphic DNA (RAPD), simple sequence repeat (SSR), intersimple sequence repeat (ISSR), and amplified fragment length polymorphism (AFLP) have been developed and widely used in the identification of various organisms [Reference Williams, Kubelik, Livak, Rafalski and Tingey4, Reference Fu, Li and Xu6, Reference Godwin, Aitken and Smith9Reference Liscum, Innis, Gelfand and Sninsky11]. Among these techniques, RAPD and ISSR methods are two PCR-based methods that require only small amounts of DNA sample without involving radioactive labels and therefore have been widely used for genetic characterization [Reference Verma, Ul Haq, Kachhwaha and Kothari12]. RAPD is a technique based on the amplification of the genomic DNA with either a single or multiple short oligonucleotide primers of an arbitrary or random sequence [Reference Verma, Ul Haq, Kachhwaha and Kothari12]. RAPD is simple, cost-efficient, and does not require DNA sequences before application [Reference Tripathi, Chouhan, Saini and Tiwari13]. ISSR is derived from SSR, which is more abundant, informative, highly polymorphic, and efficient [Reference Chaudhary, Kumar and Sharma14]. RAPD and ISSR methods have separately been used for genetic characterization in many species, such as Lonicera japonica Thunb. [Reference Fu, Yang, Khan and Mei15], synthetic hexaploid wheats [Reference Shakeel, Ilyas and Kazi16], Atractylodes lancea [Reference Kohjyouma, Nakajima, Namera, Shimizu, Mizukami and Kohda17], and Ocimum basilicum L. [Reference Giachino, Sönmez and Tonk18]. However, because of their advantages and disadvantages, more studies applied RAPD combined with ISSR to characterize the genetic variation of species, such as Litchi chinensis Sonn. [Reference Long, Cheng and Mei19], Allium species [Reference Mukherjee20], date palm [Reference El Sharabasy and Soliman21], and Cymbopogon [Reference Bishoyi, Sharma, Kavane and Geetha22]. However, only limited studies have been conducted to characterize the genetic relationships among different genus or cultivars of Prunus L. species [Reference Erturk, Ercisli, Maghradze, Orhan and Agar23Reference Baránek, Raddová and Pidra25].

In this study, we applied the RAPD and ISSR technique for the genetic characterization of 25 P. salicina from China and other countries. This study may provide valuable insight into the genetic diversity of P. salicina L. and provide information to cultivate new breed with improved traits.

2. Methods

2.1. Plant Material Collection and DNA Extraction

This study included 25 P. salicina L. which were collected from 14 different regions from China (13 samples), Japan (4 samples), USA (7 samples), and New Zealand (1 sample) (Figure 1 and Table 1). Among them, P1, P2, and P3 are the three lines with different maturity of one cultivar. The flowers of the 25 P. salicina L. are shown in Figure 2.

FIGURE 1: The localities of samples of P. salicina L. from different regions. The spots in black indicate the provinces in China.

TABLE 1: Sources of RAPD and ISSR samples.

FIGURE 2: The flowers of 25 P. salicina L.

The genomic DNA of 25 P. salicina L. was extracted from fresh leaves using a modified cetyl trimethylammonium bromide (CTAB) method as described previously [Reference Fu, Yang, Khan and Mei15, Reference Yang, Fu, Khan, Zeng and Fu26]. DNA integrity was checked by 0.8% agarose gel electrophoresis, and DNA purity was determined by the absorbance ratio at 260 nm : 280 nm on spectrophotometry. The final concentration of DNA samples was adjusted to 10 ng/µl for PCR and stored at −20°C until use.

2.2. Amplification of DNA by RAPD-PCR

The random RAPD primers were selected randomly for PCR amplification (Table 2). The PCR system in 10 μL volume contains 1 μL of 2.5 μmol/L primers, 1 μL of DNA template, 5 μL of 2 × PCR Taq Mastermix (TianGen Biotech Co. Ltd., Beijing), and 3 μL of deionized water. The PCR was executed on Applied Biosystems Veriti 96-Well Thermal Cycler (Thermo Fisher, USA) in the following procedure: initial denaturation at 95°C for 90 s, followed by 40 cycles of 40 s at 94°C, 60 s at 36°C, 90 s at 72°C, and final extension of 5 min at 72°C.

TABLE 2: Sequences of ISSR and RAPD primers.

Note. R = (A/G), Y = (C/T), and D = (A/G/T); aaverage of the column.

2.3. ISSR Amplification

Fifteen ISSR primers were synthesized by Thermo Fisher (USA) (Table 2). ISSR amplification was performed in 10 μL reactions including 1 μL of 2.5 umol/L primers, 1 μL of DNA template, 5 μL of 29 PCR Taq Mastermix (TianGen Biotech Co. Ltd., Beijing), and 3 μL of deionized water. PCR was executed on Applied Biosystems Veriti 96-Well Thermal Cycler using the following procedure: initial denaturation at 95°C for 90 s, followed by 35 cycles of 30 s at 94°C, 30 s at 50°C, 90 s at 72°C, and final extension of 5 min at 72°C [Reference Fu, Yang, Khan and Mei15].

2.4. Agarose Gel Electrophoresis

The amplified PCR products were separated by electrophoresis on 1.8% agarose gel in 1 × TAE buffer. Gels were visualized by 0.5 g/ml ethidium bromide staining, and the images were documented using the ChemiDoc XR (Bio-Rad, USA). Bands that were unambiguous and reproducible in successive amplifications were selected for scoring.

2.5. Data Analysis

All PCRs were repeated five times for each of five samples. Bands in the gel profiles were scored as 1 for present and 0 for absent. The similarity matrix (SM) and the similarity index (SI) were calculated using SM coefficient in Numerical Taxonomy Multivariate Analysis System (NTSYS pc 2.1) software. The dendrogram based on the unweighted pair group method with arithmetic mean algorithm (UPGMA) was generated using the SAHN module in the NTSYS pc 2.1 software.

3. Results

3.1. Amplification of DNA by RAPD and ISSR

A total of nineteen RAPD primers and fifteen ISSR primers were used in this study for the evaluation of DNA polymorphism (Table 2). All RAPD primers and ISSR primers generated evaluable bands. Figure 3 shows the representative reproducible polymorphic amplification bands in these 25 samples generated from ISSR primer UBC807 and RAPD primer S201. For the RAPD primers, a total of 315 bands with an average of 16.58 bands per primer were obtained. Among these bands, 257 (81.59%) bands were polymorphic, and the approximate band size ranged from 200 bp to 2300 bp. The minimum number of bands was 10, which was generated by primer OPA-4 and the maximum was 21, which was produced by primer S43. The total number of polymorphic fragments ranged from 7 (primer OPA-4) to 18 (primer OPA-10). The average polymorphic fragments ratio (PFR) (in %) was 81.60% (min: 65%; max: 94.74%). The other information of the bands generated by RAPD primers, including polymorphism information content (PIC), resolving power (RP), effective multiplex ratio (EMR), and marker index (MI), are presented in Table 3.

FIGURE 3: The representative results of banding profiles obtained by ISSR primer UBC807 (a) and RAPD primer S201 (b). Lanes P1–P25 represented different samples listed in Table 1. Lane “M” represents the DL2000 DNA marker.

TABLE 3: The characteristics of the bands generated by RAPD primers.

Note. aAverage of the column. TF, total number of fragments; PF, number of polymorphic fragments; PFR, polymorphic fragments ratios (%); PIC, polymorphism information content; RP, resolving power; EMR, effective multiplex ratio; MI, marker index.

For the ISSR primers, a total of 204 bands with an average of 13.60 bands per primer were produced; of them, 179 (87.74%) were polymorphic. The approximate range of band size was 300 bp to 2500 bp (Table 4). The minimum number of bands was 8, which was yielded by primer UBC829, and the maximum was 19, which was produced by primer UBC807. The total number of PFs ranged from 8 (primer UBC829) to 15 (primers UBC807, UBC810, UBC846, and UBC881). The average PFR% was 87.80% (min: 69.23%; max: 100%). The other information of the bands generated by ISSR primers, including PIC, RP, EMR, and MI, are presented in Table 4.

TABLE 4: The characteristics of the bands generated by ISSR primers.

Note. aAverage of the column. TF, total number of fragments; PF, number of polymorphic fragments; PFR, polymorphic fragments ratios (%); PIC, polymorphism information content; RP, resolving power; EMR, effective multiplex ratio; MI, marker index.

3.2. Genetic Distance and Cluster Analysis of RAPD and ISSR Markers

Based on the RAPD amplification profiles, cluster dendrogram was obtained using UPGMA (Figure 4). Since P1, P2, and P3 belong to one cultivar, we ignored their coefficients in the following analysis. The dendrogram showed that the similarity coefficients ranged from 0.584 to 0.860. In the RAPD-based dendrogram, the 25 P. salicina samples formed four clusters at a cutoff of 0.692. The similarity coefficient between sample P4 (Qiepili in Ningbo, Zhejiang, China) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest (0.584), while that between sample P19 (Laroda in USA) and P20 (Red heart in USA) was highest (0.860) (Figure 4).

FIGURE 4: Dendrogram of cluster of 25 P. salicina L. based on RAPD markers.

The ISSR analysis showed similar results to the RAPD analysis. The dendrogram showed that the similarity coefficients ranged from 0.558 to 0.892. In the ISSR-based dendrogram, the 25 P. salicina samples were divided into five clusters at a cutoff of 0.692. The similarity coefficient between sample P10 (Maihuangli in Anhui, China) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest (0.558), while that between sample P10 (Maihuangli in Anhui, Chin) and P15 (Baili in Japan) was highest (0.892) (Figure 5).

FIGURE 5: Dendrogram of cluster of 25 P. salicina L. based on ISSR markers.

3.3. Integrating Analysis of RAPD and ISSR Data

The dendrogram results of RAPD combined with ISSR showed that the similarity coefficients ranged from 0.597 to 0.865. Total 519 DNA fragments were yielded, of which 435 (84.7%) were polymorphic. The average number of PF per primer was 12.7. The mean PIC, RP, EMR, and MI values observed for all primers were 0.42, 17.77, 10.80, and 4.54, respectively (Table 5). The similarity coefficients between sample P10 (Maihuangli in Anhui, China) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest (0.597), while that between sample P10 (Maihuangli in Anhui, China) and P15 (Baili in Japan) was highest (0.865) (Table 6).

TABLE 5: Comparative analysis of genetic variability in Prunus L. landraces using ISSR, RAPD, and combined data.

TABLE 6: The similarity matrix of the landraces using Dice’s coefficient based on the ISSR and RAPD bands.

Note. The bold values indicate the maximum and minimum genetic similarity values among the landraces.

3.4. Typical Band Patterns Amplified by ISSR and RAPD Markers

Sixteen primers, including 11 ISSR primers and 5 RAPD primers, could be used as the markers of molecular identification for 25 Prunus L. samples (Table 7). As shown in Table 7, UBC810, UBC834, and UBC836 could be considered as the markers of P1 (Zuili1 in Jiaxing, Zhejiang, China), P2 (Zuili2 in Jiaxing, Zhejiang, China), and P3 (Zuili3 in Jiaxing, Zhejiang, China). S17 could be considered as a marker of P4 (Qiepili in Ningbo, Zhejiang, China). UBC881 might be a marker of P5 (Jintangli in Zhoushan, Zhejiang, China). UBC847 was a marker of P6 (Furongli in Fujian, China). UBC847 and UBC855 could be used to distinguish P7 (Yuhuangli in Hubei, China). UBC848 could be considered as a marker of P8 (Jiuqianli in Guizhou, China). UBC857 might be a potential marker of P9 (Huangguli in Tongxiang, Zhejiang, China). RAPD-1 could be used as a marker of P12 (Niuxinli in Shandong, China). UBC889 could be considered as a marker of P16 (Akihime in Japan). S43 and S1403 might be the markers of P17 (Zhenzhuli in Japan). UBC829 might be a potential marker of P23 (Queen rose in USA). RAPD-5 also might be used a marker of P25 (Misili in New Zealand). The representative banding profiles obtained by ISSR primers UBC834, UBC847, UBC857, and RAPD primer S1403 are shown in Figure 6.

TABLE 7: Typical band patterns amplified by ISSR and RAPD markers.

FIGURE 6: The representative results of banding profiles obtained by ISSR primers UBC834 (a), UBC847 (b), UBC857 (c), and RAPD primer S1403 (d). Lanes P1–P25 represent different samples listed in Table 1. Lane “M” represents the DL2000 DNA marker. The typical bands for molecular identification of P. salicina L. are indicated by a red arrow.

4. Discussion

Illustration of the genetic relationships or characterization of genetic diversity is important to provide genetic guidance for hybrid breeding. In this study, the genetic diversity and relationship among 25 P. salicina L. varieties were evaluated by RAPD and ISSR, respectively, and integrated. The Dice’s similarity coefficient of RAPD ranged from 0.584 to 0.860, and that of ISSR ranged from 0.558 to 0.892. Integrating analysis of RAPD and ISSR indicated the similarity coefficient varied from 0.597 to 0.865. The results indicated high diversity among the 25 varieties.

ISSR and RAPD were widely used for genetic diversity evaluations of Prunus L. species. Tian et al. used ISSR and RAPD for genetic diversity evaluations of 48 Prunus mira L. samples, the high levels of polymorphism, and the results imply that Tibet samples preserved higher genetic diversity and most genetic variations occurred [Reference Tian, Xing and Cao27]. However, the efficiency of RAPD markers and ISSR markers in detecting polymorphism is controversial. Tian et al. demonstrated that ISSR found 77.80% polymorphism, which is higher than that found by RAPD (72.73%). In the study of Kumar et al. the phylogenetic relationships of 36 locally grown P. armeniaca genotypes were analyzed using 20 RAPDs and 11 ISSRs markers. RAPD markers were found more efficient for polymorphism detection, as they detected 97.84% as compared to 96.5% for ISSR markers, and the pattern of clustering of the genotypes remained more or less the same in RAPD and combined data of RAPD + ISSR [Reference Kumar, Mishra, Singh, Kumar, Naik and Singh28]. In our study, the PFR% of RAPD primers was 81.60%, which is lower than that of ISSR primers (87.80%). Our results support the view that ISSR markers are more efficient than RAPD with regards to detecting polymorphism.

The RAPD results showed that the index of similarity coefficient between sample P4 (Qiepili in Ningbo, Zhejiang, China) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest (0.584), while that between sample P19 (Laroda in USA) and P20 (Red heart in USA) was highest (0.860). However, the ISSR results showed that the index of similarity coefficient between sample P10 (Maihuangli in Anhui, Chin) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest (0.558), while that between sample P10 (Maihuangli in Anhui, China) and P15 (Baili in Japan) was highest (0.892). In addition, the analysis of RAPD combined with ISSR showed that the similarity coefficient between sample P10 (Maihuangli in Anhui, China) and P13 (Longyuanqiuli in Heilongjiang, China) was lowest (0.597), while that between sample P10 (Maihuangli in Anhui, Chin) and P15 (Baili in Japan) was highest (0.865), which was consistent with the RAPD analysis. These findings demonstrated that the RAPD technique not only increased the resolution and yield but also was a reliable molecular tool for the genetic characterization of various organisms, which was reported in previous studies [Reference Fu, Li and Xu6, Reference Fu, Yang, Khan and Mei15]. Our RAPD and ISSR analysis showed potentiality to distinguish P. salicina L. from related genus or species.

5. Conclusion

In summary, our study indicates that the RAPD combined with ISSR techniques would be used for the genetic diversity, molecular-assisted breeding, and genetic characterization of P. salicina L. Our results might assist in parental gametophytes selection for hybrid breeding of P. salicina L.

Abbreviations

RAPD: Random amplified polymorphic DNA

SSR: Simple sequence repeat

ISSR: Intersimple sequence repeat

AFLP: Amplified fragment length polymorphism

CTAB: Cetyl trimethylammonium bromide

SM: Similarity matrix

SI: Similarity index

NTSYS: Numerical taxonomy multivariate analysis system

UPGMA: Unweighted pair group method with arithmetic mean algorithm

PIC: Polymorphism information content

RP: Resolving power

EMR: Effective multiplex ratio

MI: Marker index

TF: Total number of fragments

PF: Polymorphic fragments

PFR (%): Polymorphic fragments ratios (%).

Data Availability

The data that support the findings of this study are available on request to the corresponding author.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

Authors’ Contributions

JL and GCG designed experiments. Bin acquired data. Bai and QHL analyzed and interpreted data. GCG obtained the funding. JL is a major contributor in drafting the manuscript. All authors read and approved the final version of the manuscript.

Acknowledgments

This work was supported in part by the National Natural Science Foundation of China (81703667), Basic Public Welfare Research Project of Zhejiang Province (LGN21C150004 and LGF21H280007), Jiaxing Science and Technology Project (2020AY10023), and Special Project of Jiaxing Science and Technology Commissioner (2021K117).

References

Chen, Z.-L., Chen, W.-J., Chen, H. et al., “Prunus pananensis (Rosaceae), a new species from Pan’an of central Zhejiang, China,PLoS One, vol. 8, no. 1, Article ID e54030, 2013.Google ScholarPubMed
Herrera, S., Lora, J., Hormaza, J. I., Herrero, M., and Rodrigo, J., “Optimizing production in the new generation of apricot cultivars: self-incompatibility, S-RNase allele identification, and incompatibility group assignment,Frontiers of Plant Science, vol. 9, p. 527, 2018.CrossRefGoogle ScholarPubMed
Wen, J., Berggren, S., Lee, C.-H. et al., “Phylogenetic inferences in Prunus (Rosaceae) using chloroplast ndhF and ribosomal ITS sequences,Journal of Systematics and Evolution, vol. 46, pp. 322332, 2008.Google Scholar
Williams, J. G. K., Kubelik, A. R., Livak, K. J., Rafalski, J. A., and Tingey, S. V., “DNA polymorphisms amplified by arbitrary primers are useful as genetic markers,Nucleic Acids Research, vol. 18, no. 22, pp. 65316535, 1990.CrossRefGoogle ScholarPubMed
Pairon, M., Petitpierre, B., Campbell, M. et al., “Multiple introductions boosted genetic diversity in the invasive range of black cherry (Prunus serotina; Rosaceae),sAnnals of Botany, vol. 105, no. 6, pp. 881890, 2010.CrossRefGoogle ScholarPubMed
Fu, J., Li, L. Y., Xu, X. et al., “An improved method for increasing the efficiency of the technique of random amplified polymorphic DNA (RAPD),Hereditas, vol. 22, pp. 251252, 2000.Google Scholar
Urrestarazu, J., Errea, P., Miranda, C., Santesteban, L. G., and Pina, A., “Genetic diversity of Spanish Prunus domestica L. germplasm reveals a complex genetic structure underlying,PLoS One, vol. 13, no. 4, Article ID e0195591, 2018.CrossRefGoogle ScholarPubMed
Carrasco, B., González, M., Gebauer, M., and García-González, R. J. H., “Construction of a highly saturated linkage map in Japanese plum (Prunus salicina L.) using GBS for SNP marker calling,PLoS One, vol. 13, no. 12, Article ID e0208032, 2018.CrossRefGoogle ScholarPubMed
Godwin, I. D., Aitken, E. A. B., and Smith, L. W., “Application of inter simple sequence repeat (ISSR) markers to plant genetics,Electrophoresis, vol. 18, no. 9, pp. 15241528, 1997.CrossRefGoogle ScholarPubMed
Agarwal, M., Shrivastava, N., and Padh, H., “Advances in molecular marker techniques and their applications in plant sciences,Plant Cell Reports, vol. 27, no. 4, pp. 617631, 2008.CrossRefGoogle ScholarPubMed
Liscum, E., “Amplified fragment length polymorphism,” in PCR Applications, Innis, M. A., Gelfand, D. H., Sninsky, J. J., Eds., pp. 505519, Academic Press, San Diego, CA, USA, 1999.CrossRefGoogle Scholar
Verma, K. S., Ul Haq, S., Kachhwaha, S., and Kothari, S. L., “RAPD and ISSR marker assessment of genetic diversity in Citrullus colocynthis (L.) schrad: a unique source of germplasm highly adapted to drought and high-temperature stress,3 Biotech, vol. 7, no. 5, p. 288, 2017.CrossRefGoogle ScholarPubMed
Tripathi, N., Chouhan, D. S., Saini, N., and Tiwari, S., “Assessment of genetic variations among highly endangered medicinal plant Bacopa monnieri (L.) from Central India using RAPD and ISSR analysis,3 Biotech, vol. 2, no. 4, pp. 327336, 2012.CrossRefGoogle Scholar
Chaudhary, V., Kumar, M., Sharma, S. et al., “Assessment of genetic diversity and population structure in gladiolus (Gladiolus hybridus Hort.) by ISSR markers,Physiology and Molecular Biology of Plants, vol. 24, no. 3, pp. 493501, 2018.CrossRefGoogle ScholarPubMed
Fu, J., Yang, L., Khan, M. A., and Mei, Z., “Genetic characterization and authentication of Lonicera japonica Thunb. by using improved RAPD analysis,Molecular Biology Reports, vol. 40, no. 10, pp. 59935999, 2013.CrossRefGoogle ScholarPubMed
Shakeel, M., Ilyas, M., and Kazi, M., “Evaluation of synthetic hexaploid wheats (derivative of durum wheats and Aegilops tauschii accessions) for studying genetic diversity using randomly amplified polymorphic DNA (RAPD) markers,Molecular Biology Reports, vol. 40, no. 1, pp. 2126, 2013.CrossRefGoogle ScholarPubMed
Kohjyouma, M., Nakajima, S., Namera, A., Shimizu, R., Mizukami, H., and Kohda, H., “Random amplified polymorphic DNA analysis and variation of essential oil components of atractylodes plants,Biological and Pharmaceutical Bulletin, vol. 20, no. 5, pp. 502506, 1997.CrossRefGoogle ScholarPubMed
Giachino, R. R. A., Sönmez, Ç., Tonk, F. A. et al., “RAPD and essential oil characterization of Turkish basil (Ocimum basilicum L.),Plant Systematics and Evolution, vol. 300, no. 8, pp. 17791791, 2014.CrossRefGoogle Scholar
Long, Y., Cheng, J., Mei, Z. et al., “Genetic analysis of litchi (Litchi chinensis Sonn.) in southern China by improved random amplified polymorphic DNA (RAPD) and inter-simple sequence repeat (ISSR),Molecular Biology Reports, vol. 42, no. 1, pp. 159166, 2015.CrossRefGoogle Scholar
Mukherjee, A., “RAPD and ISSR analysis of some economically important species, varieties and cultivars of the genus Allium (Alliaceae),Turkish Journal of Botany, vol. 37, 2013.Google Scholar
El Sharabasy, S. F. and Soliman, K. A., “Molecular analysis of date palm genetic diversity using random amplified polymorphic DNA (RAPD) and inter-simple sequence repeats (ISSRs),Methods in Molecular Biology, vol. 1638, pp. 143152, 2017.CrossRefGoogle Scholar
Bishoyi, A. K., Sharma, A., Kavane, A., and Geetha, K. A., “Varietal discrimination and genetic variability analysis of cymbopogon using RAPD and ISSR markers analysis,Applied Biochemistry and Biotechnology, vol. 179, no. 4, pp. 659670, 2016.CrossRefGoogle ScholarPubMed
Erturk, Y., Ercisli, S., Maghradze, D., Orhan, E., and Agar, G., “An assessment of genetic variability and relationships among wild-grown blackthorn (Prunus spinosa L.) plants based on RAPD markers,Genetics and Molecular Research, vol. 8, no. 4, pp. 12381244, 2009.CrossRefGoogle Scholar
Fernandezi Marti, A. I., Athanson, B., Koepke, T., Font, I. F., Amit, D., and Oraguzie, N., “Genetic diversity and relatedness of sweet cherry (Prunus avium L.) cultivars based on single nucleotide polymorphic markers,Frontiers of Plant Science, vol. 3, p. 116, 2012.Google Scholar
Baránek, M., Raddová, J., and Pidra, M., “Comparative analysis of genetic diversity in Prunus L. as revealed by RAPD and SSR markers,Scientia Horticulturae, vol. 108, no. 3, pp. 253259, 2006.CrossRefGoogle Scholar
Yang, L., Fu, S., Khan, M. A., Zeng, W., and Fu, J., “Molecular cloning and development of RAPD-SCAR markers for dimocarpus longan variety authentication,SpringerPlus, vol. 2, no. 1, p. 501, 2013.CrossRefGoogle ScholarPubMed
Tian, Y., Xing, C., Cao, Y. et al., “Evaluation of genetic diversity on prunus mira koehne by using ISSR and RAPD markers,Biotechnology and Biotechnological Equipment, vol. 29, no. 6, pp. 10531061, 2015.CrossRefGoogle Scholar
Kumar, M., Mishra, G. P., Singh, R., Kumar, J., Naik, P. K., and Singh, S. B., “Correspondence of ISSR and RAPD markers for comparative analysis of genetic diversity among different apricot genotypes from cold arid deserts of trans-Himalayas,Physiology and Molecular Biology of Plants, vol. 15, no. 3, pp. 225236, 2009.CrossRefGoogle ScholarPubMed
Figure 0

FIGURE 1: The localities of samples of P. salicina L. from different regions. The spots in black indicate the provinces in China.

Figure 1

TABLE 1: Sources of RAPD and ISSR samples.

Figure 2

FIGURE 2: The flowers of 25 P. salicina L.

Figure 3

TABLE 2: Sequences of ISSR and RAPD primers.

Figure 4

FIGURE 3: The representative results of banding profiles obtained by ISSR primer UBC807 (a) and RAPD primer S201 (b). Lanes P1–P25 represented different samples listed in Table 1. Lane “M” represents the DL2000 DNA marker.

Figure 5

TABLE 3: The characteristics of the bands generated by RAPD primers.

Figure 6

TABLE 4: The characteristics of the bands generated by ISSR primers.

Figure 7

FIGURE 4: Dendrogram of cluster of 25 P. salicina L. based on RAPD markers.

Figure 8

FIGURE 5: Dendrogram of cluster of 25 P. salicina L. based on ISSR markers.

Figure 9

TABLE 5: Comparative analysis of genetic variability in Prunus L. landraces using ISSR, RAPD, and combined data.

Figure 10

TABLE 6: The similarity matrix of the landraces using Dice’s coefficient based on the ISSR and RAPD bands.

Figure 11

TABLE 7: Typical band patterns amplified by ISSR and RAPD markers.

Figure 12

FIGURE 6: The representative results of banding profiles obtained by ISSR primers UBC834 (a), UBC847 (b), UBC857 (c), and RAPD primer S1403 (d). Lanes P1–P25 represent different samples listed in Table 1. Lane “M” represents the DL2000 DNA marker. The typical bands for molecular identification of P. salicina L. are indicated by a red arrow.