Next Article in Journal
Environmental Behaviors of Bacillus thuringiensis (Bt) Insecticidal Proteins and Their Effects on Microbial Ecology
Next Article in Special Issue
Simple Sequence Repeat (SSR)-Based Genetic Diversity in Interspecific Plumcot-Type (Prunus salicina × Prunus armeniaca) Hybrids
Previous Article in Journal
Intelligent Identification and Features Attribution of Saline–Alkali-Tolerant Rice Varieties Based on Raman Spectroscopy
Previous Article in Special Issue
Characterizing Tetraploid Populations of Actinidia chinensis for Kiwifruit Genetic Improvement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Study on Supergenus Rubus L.: Edible, Medicinal, and Phylogenetic Characterization

Lushan Botanical Garden, Chinese Academy of Sciences, Jiujiang 332900, China
*
Author to whom correspondence should be addressed.
Plants 2022, 11(9), 1211; https://doi.org/10.3390/plants11091211
Submission received: 14 March 2022 / Revised: 27 April 2022 / Accepted: 28 April 2022 / Published: 29 April 2022

Abstract

:
Rubus L. is one of the most diverse genera belonging to Rosaceae; it consists of more than 700 species with a worldwide distribution. It thus provides an ideal natural “supergenus” for studying the importance of its edible, medicinal, and phylogenetic characteristics for application in our daily lives and fundamental scientific studies. The Rubus genus includes many economically important species, such as blackberry (R. fruticosus L.), red raspberry (R. ideaus L.), black raspberry (R. occidentalis L.), and raspberry (R. chingii Hu), which are widely utilized in the fresh fruit market and the medicinal industry. Although Rubus species have existed in human civilization for hundreds of years, their utilization as fruit and in medicine is still largely inadequate, and many questions on their complex phylogenetic relationships need to be answered. In this review, we briefly summarize the history and progress of studies on Rubus, including its domestication as a source of fresh fruit, its medicinal uses in pharmacology, and its systematic position in the phylogenetic tree. Recent available evidence indicates that (1) thousands of Rubus cultivars were bred via time- and labor-consuming methods from only a few wild species, and new breeding strategies and germplasms were thus limited; (2) many kinds of species in Rubus have been used as medicinal herbs, though only a few species (R. ideaus L., R. chingii Hu, and R. occidentalis L.) have been well studied; (3) the phylogeny of Rubus is very complex, with the main reason for this possibly being the existence of multiple reproductive strategies (apomixis, hybridization, and polyploidization). Our review addresses the utilization of Rubus, summarizing major relevant achievements and proposing core prospects for future application, and thus could serve as a useful roadmap for future elite cultivar breeding and scientific studies.

Graphical Abstract

1. Introduction

Rubus L. is one of the most diverse and largest genera in the Rosaceae family. The genus consists of more than 700 shrubby or herbaceous species mainly distributed throughout the temperate zone of the northern hemisphere, with a few having expanded to the tropics and the southern hemisphere [1,2,3,4,5,6]. Species of the Rubus genus worldwide are classified into 12 subgenera [1,2,3]. However, Lu et al. [6] reclassified them into 8 subgenera, whereby only habitats in China were considered. There are two hypothetical centers of origin for Rubus: one is North America [7,8] and the other is southwestern China [9,10,11]. In addition, the pleasant flavor of the fresh Rubus fruit, its medicinal functions due to the health benefits of its very high secondary metabolite content, and its high genetic diversity and complex phylogeny rendering it suitable for scientific studies, make Rubus an important and ideal genus for breeders as well as scientists [8,12,13] (Figure 1). Furthermore, the rich secondary metabolites and the bark of Rubus are also important raw materials for cosmetics and fiber [14].
Rubus bears aggregate drupetum fruits that have economically important edible and medicinal characteristics [12,13]. They have a pleasant flavor and have been dubbed “superfoods” due to their very high levels of secondary metabolites, such as hydrolyzable tannins, anthocyanins, polyphenols, flavanols, organic acids, and many other organic compounds [12,15,16,17,18,19,20]. In an early investigation, Moyer et al. [15] extracted multiple anthocyanins and phenols from the ripe fruits of Rubus. Based on genomic resequencing, quadrupole time-of-flight liquid chromatography, and mass spectroscopy, 29 hydrolyzable tannins and their candidate chromosomal regions were identified by Wang et al. [20]. Because of these diverse secondary metabolites, the superfood Rubus fruits can provide anti-oxidants as well as anti-cancer, anti-microbial, and anti-complement activities, in addition to having other benefits for humans [21,22,23,24,25].
Since the publication of Darwin’s On the Origin of Species, understanding the genetic basis of adaptation for the arisal of new species has been a central topic in evolutionary biology [26,27,28,29]. In the species-rich genus of Rubus, the diversity of reproductive strategy, such as through the process of hybridization, polyploidization, and apomixis, which enhances the adaptation capacity of Rubus [8,30,31]. Additionally, the various reproductive strategies also bring a huge challenge regarding the taxonomy of the Rubus genus in terms of morphological and molecular systematics [8,9,31,32,33]. However, in order to make better use of the wild germplasm in Rubus, it is necessary to develop a better knowledge of the clear affinity of its phylogeny. Therefore, species of the Rubus genus provide a natural experimental system for studying the fundamental mechanisms of adaptation via diverse reproductive strategies and reticulate evolutionary phylogeny.
Over the past hundred years since Focke [1] published Species Ruborum, our understanding of Rubus L. has improved, including regarding its edibility [4,12,30] and in the medicinal [4,13,15,18,20,34] and phylogenetic fields [7,8,31,35,36] (Figure 1). However, most previous studies and reviews have devoted their attention to metabolic compounds of pharmacologic interest in only a very few specific species [13,19]. This review summarizes the general outcomes for fresh fruit breeding, medicinal components, and studies into phylogenetic relationships of the Rubus genus; based on the latest advances in the fields of omics (genomics, transcriptomics, proteomics, and metabolomics), CRISPR/Cas, and other genome editing technologies, experimental efficiency has improved remarkably. Finally, we point out the important value of Rubus in fruit germplasms, medicinal research, and understanding complex phylogenetic relationships resulting from diverse adaptative reproduction strategies. In short, Rubus L., in the Rosaceae family, is an ideal natural “supergenus” for breeders, pharmacologists, and evolutionary biologists. Our review addresses major questions regarding how to better exploit the wild germplasm in Rubus species and can thus serve as a useful roadmap for future breeding and fundamental scientific studies.

2. Studies of Edible Rubus Species

Rubus L. belongs to the Rosaceae family, from which many palatable fruit cultivars have been bred, such as the woodland strawberry (Fragaria vesca L. var. americana Porter), and the domesticated apple (Malus × domestica Borkh.), pear (Pyrus bretschneideri Rehd.), and peach (Prunus persica (L.) Batsch) [1,9,37]. Rubus species are popular for their pleasant fresh fruits, including the blackberry (R. fruticosus L.), red raspberry (R. ideaus L.), and black raspberry (R. occidentalis L.). Rubus fruits are aggregate drupetum fruits with varied colors, such as red, yellow, purple, and black [12]. Rubus fruits have been called “superfoods” because of the very high levels of beneficial secondary metabolites that they contain, including, e.g., anthocyanins, phenolic acids, flavonoids, tannins, and other essential compounds [19,24,38,39,40] (Figure 2). The chemical composition of Rubus is influenced not only by environmental factors but also internal genetic differences. For example, studies of blackberries found that a temperate climate and higher cumulative rainfall increased the production of phenolic compounds [41]. The concentration of chemical components in Rubus is also related to the storage conditions, growing season or location, and maturity [42,43,44,45]. The genotype differences between species and cultivars in Rubus is the major internal factor that influences the divergence of chemical compositions; for instance, Skrovankova et al. [46] found that different cultivars show significant variations in the production of secondary metabolites, even when they were grown under the same environmental conditions. Furthermore, a series of studies on different Rubus genotypes, cultivars, and species produced results consistent with those of Skrovankova [41,44,47,48,49].
Species of the Rubus genus have been cultivated and appeared in gardens for more than 15 centuries in Europe, for example in Turkey and Rome [18,38]. To date, thousands of cultivars have been bred, and these can mainly be divided into two types: the primocane-fruiting (also called annual-fruiting) type, including Heritage, Amity, Autumn Bliss, Autumn Britten, Dinkum, and Polana cultivars; and the floricane-fruiting (also called biennial-fruiting) type, including Claudia, Emily, Esta, Lauren, and Qualicum cultivars [8,12,38]. Among these, Logan, Boysen, and Marion are three elite cultivars. The major areas for growing these cultivars are Russia (125,000 t), North America (59,123 t), and Europe (43,000 t) [8,38]. Spurred by the human pursuit of fruit quality and increasing consumption, the fruit production of Rubus cultivars has rapidly expanded for the production of fresh fruit for use in jams and fruit juice [12,38,50,51] (Figure 1).
In the early stages of Rubus domestication, breeders commonly selected superior individuals from wild habitats [4,5]. During the nineteenth century, approximately 30 breeding projects were conducted in North America and Europe. The elite cultivars of Preussen, Cuthbert, and Newburgh were bred by crossing between different subspecies of red raspberries [30]. However, the process of domestication has vastly reduced the morphological and genetic diversities of crops [52,53,54,55]. Current cultivars are bred from crossing or the improvement of only a few wild species, i.e., red raspberry (R. ideaus L.), black raspberry (R. occidentalis L.), and blackberry (R. fruticosus L.). In order to meet the demands of visual appeal, higher yield, greater quality, excellent health benefits, and diverse adaptation for breeders, more wild germplasms of Rubus resources and advanced biotechnology should be utilized for Rubus breeding. Based on simple sequence repeat (SSR) and amplified fragment length polymorphism (AFLP) markers, the first linkage map for Rubus was constructed based on the crossing of two red raspberry cultivars, Glen Moy and Latham, from Europe and North America, respectively [56]. Bushakra et al. [57] constructed a genetic linkage map using 1218 markers by the crossing of S1 (R. occidentalis L.) and Latham (R. idaeus L.) and compared it with genomes of other genera in the rosa family, such as Fragaria L., Malus Mill. and Prunus L. That study reported a high consistency of collinearity of genomes between different genera of Rosaceae, and hundreds of new polymorphic genetic markers were found for future quantitative trait loci mapping studies. In recent years, different high-resolution markers and advanced sequencing methods have been applied for trait mapping or new wild germplasm identification in Rubus [58,59,60,61,62].

3. Medicinal Studies of Rubus

Rubus L. is one of the most species-rich genera in the Rosaceae family, but only a few species have been used as medicinal herbs [13,18,63]. According to the records of the ancient pharmacopoeias in Europe and China, Rubus species have been used as medicinal herbs for several centuries. The stems and leaves of blackberry (R. fruticosus L.) were soaked with white wine for use as an astringent poultice for wound healing and for difficulties during childbirth, as suggested by Hippocrates [18]. The dried unripe fruits of “Fu-Pen-Zi” (R. chingii Hu) were used to improve and enhance liver and kidney health [20,64]. More recently, as shown in Figure 2, many kinds of secondary metabolites with remarkable beneficial effects on humans have been extracted. Table 1 exhibits detailed information on the species, concentration, and part from which they were extracted. For instance, anthocyanins with an anti-oxidant function are extracted from the fruits of the blackberry [15,19]; flavonoids with anti-oxidant, anti-cancer, and anti-inflammatory effects are extracted from the fruits of the blackberry or Fu-Pen-Zi [16,65,66,67]. Other organic compounds have been identified, mainly in blackberry or Fu-Pen-Zi, such as hydrolyzable tannin, glycoprotein, organic acid, and phenolic compounds [19,21,61,68,69]. In addition, reports have indicated many other kinds of beneficial effect of these secondary metabolites on humans, for example, improving mitosis and eyesight, treating or preventing cancer, back pain, and frequent urination [23,25,70,71]. To date, most of the secondary metabolites mentioned above are considered safe according to data from limited studies [13,72,73,74,75,76]. For example, the extracted components from R. niveus Thunb. showed no statistically significant toxicity for mice [72]. Based on a cytotoxic experiment on Caco-2 cells, Ke et al. [77] found that metabolites extracted from the fruit of R. chingii Hu were safe and had a favourable effect on anti-cancer cells. Overall, the limited available data on the toxicity and allergenicity of Rubus species indicate they are safe for humans. More in-depth investigations regarding medicinal applications in pharmacology are needed.
In the genomic age, genomic sequencing of Rubus is already lagging behind compared to other major crops and fruits, such as rice, maize, cotton, apple, and pear [53,78,79,80,81,82,83,84,85]. However, benefiting from the quick development of the cost-effective next-generation sequencing (NGS) [86,87] and transcriptome (RNA-seq) technology [88,89], nuclear or plastid genomes have been sequenced for some important medicinal species in the Rubus genus [20,90,91,92,93,94,95]. Utilizing the RNA-seq data of the red raspberry (R. idaeus L.) fruit, Hyun et al. [90] determined the regulated candidate genes for biosynthesis of γ-aminobutyric acid and anthocyanins, which have anti-oxidant activity. Based on unripe Fu-Pen-Zi (R. chingii Hu) fruits, the chromosome-scale reference and genomic regions related to the biosynthetic pathway for hydrolyzable tannin (HT) have been reported [20]. Therefore, using the results of these studies, breeders could modify candidate genes or genomic regions of Rubus cultivars to improve the content of targeted secondary metabolites (HT, anthocyanins, etc.) using site-directed genome editing technologies such as CRISPR/Cas. More recently, in the field of crop breeding, CRISPR/Cas genome-editing technology has been used with encouraging results [96]. However, use of this speedy, proven, and precise genome editing technology has not been reported in programs for breeding Rubus. To date, various studies taking advantage of transcriptomic analysis at different developmental stages (green; green and yellow; yellow, orange, and red) of Fu-Pen-Zi fruits have revealed that flavonoids and anthocyanins are synthesized at an early stage and their levels then decrease during subsequent development [61,97,98]. These studies also indicated that anthocyanins might not be responsible for the reddish color of ripe fruits. Thus, better characterization will require extraction of the pharmacological metabolites at the early stage of Rubus fruit development. Meanwhile, several studies on plastid genomes have focused on the pharmacological components of Rubus, i.e., R. eucalyptus Focke [93], R. rufus Focke [99], R. longisepalus Nakai and R. hirsutus Thunb. [100,101], and R. phoenicolasius Maxim. [94]. However, the detailed genetic basis and biosynthetic pathways of the pharmacological metabolites in Rubus species are still largely unclear, and further future investment and research on different aspects are needed [102,103,104].
Table 1. Major secondary metabolites of Rubus L.
Table 1. Major secondary metabolites of Rubus L.
Secondary
Metabolite
SpeciesConcentration *
(mg/100 g)
PartReferences
AnthocyaninR. chingii2.1~326Leaf[20,39,40,105,106]
R. fruticosusFruit
R. ideaus
R. hirsutus
FlavonoidR. chingii2.8~6Leaf[20,34,40,107]
R. occidentalisFruit
Phenolic
compounds
R. chingii13.7~1541Root[19,20,39,40,48,105,106,108]
R. occidentalisStem
R. setchuenensisLeaf
Flower
Fruit
Organic
acids
R. chingii0.2~52.9Stem[20,109]
R. coreanusLeaf
Fruit
GlycoproteinR. chingii14.6~81.4Fruit[63]
* The concentration data were collected from multiple studies in which the extraction method, part, and species varied, and thus the data presented in the table show minimum and maximum values.

4. Phylogenetic Studies of Rubus

The Rubus genus is one of the most successful models of an adaptive and evolutionary group, with distribution worldwide except for Antarctica [1,9]. As shown in Table 2, Rubus species were classified by Focke into 12 or by Lu into 8 subgenera according to worldwide distribution or distribution in China, respectively. The classification and phylogenetic construction of Rubus is a challenging task due to phenomena such as hybridization, apomixis, polyploidization, and introgression, which happen frequently in this genus. Ploidy levels among different subgenera and species are highly differentiated [7,32,33,35,58] (Table 2). More importantly, the diverse reproductive strategies may have conferred to Rubus species the ability to occupy various habitats worldwide, and thus demonstrate reticulate evolutionary phylogeny. According to previous phylogenetic analysis based on the ndhF gene, Howarth et al. [110] suggested that the Hawaiian Islands species (R. hawaiensis A. Gray and R. macraei A. Gray) originated from different ancestors, in contradiction with the morphological results. The reticulate evolution of Rubus has been indicated in recent studies. Wang et al. [31] used multiple chloroplast and nuclear genes to investigate the phylogenetic relationships of 142 Rubus taxa, which indicated reticulate evolutionary events between different subgenera and species. A study based on approximately 1000 target genes constructed the phylogenetic tree for 87 wild Rubus taxa and three cultivars, concluding that hybridization and incomplete lineage sorting (ILS) were responsible for the low resolution and topological conflicts between different subgenera, which were not caused by insufficient molecular signals [8]. Furthermore, it has been suggested that North America might be the primary center of origin of Rubus, which then expanded into Asia and Europe and finally dispersed to Oceania via birds [8].
Additionally, a series of studies on the phylogeny of Rubus detected conflicts in the phylogenetic affinities between plastid genes and nuclear genes in most cases [36,114,115]. The molecular and morphological topotaxies also appear inconsistent, which may result from the multiple reproductive strategies [36,111,114,115,116,117]. In general, the difficulties of morphological or molecular taxonomy in subgenera and between species are not caused by lack of characteristics or signals; the real reason may be the diverse reproductive patterns that have made this group an ideal genus for investigation of the genetic basis of different reproductive and adaptive patterns.

5. Concluding Remarks and Future Perspectives

The Rubus genus consists of more than 700 species, but only a few of them, such as blackberry (R. fruticosus L.), red raspberry (R. ideaus L.), and black raspberry (R. occidentalis L.), have been domesticated or crossed by breeders to generate elite cultivars with excellent characteristics of strong adaptability, good storage properties, and pest or disease resistance. Unfortunately, there have been relatively few molecular breeding studies on Rubus and fewer genomic resources exist compared to other types of crops and fruits. Rubus breeders can reference those studies in order to improve the breeding methods for elite cultivars of different crops and fruits.
Furtehrmore, the situation for medicinal cultivars of Rubus is worse, and most of the investments in pharmacology have been concerned with R. ideaus L., R. fruticosus L., and R. chingii Hu, rarely involving residual species such as R. eucalyptus Focke, R. occidentalis L., and R. phoenicolasius Maxim., which have also been used as medicinal ingredients for hundreds of years. Notably, there remains a large deficiency in the study of the basic mechanisms and genetics of the active ingredients in medicinal Rubus species. Fortunately, advances in NGS and RNA-seq technologies offer an opportunity for researchers to spend less money and labor investigating the above-mentioned problems and to quickly identify and choose high-quality germplasms.
Finally, reconstructing the phylogenetic relationships for Rubus is a task made challenging by hybridization, polyploidization, apomixis, and introgression. However, researchers can also combine consideration of morphological characteristics with omics technologies (i.e., genomics, transcriptomics, proteomics, and metabolomics) to decipher the phylogenetic and evolutionary puzzles of Rubus. Consumers in the present era are increasingly demanding tastier and healthier fresh fruits. The wild species and elite cultivars of Rubus provide ideal candidates to address this demand due to their pleasant flavor and high concentrations of secondary metabolites. However, only a few wild species have been domesticated and are used in our daily food markets and medical treatment. Therefore, in order to better exploit the abundance of Rubus wild germplasms, information on their phylogenetic relationships and genetic diversity should be clarified. This study reviewed three major topics (edible, medicinal, and phylogenetic properties), but many challenges still exist in the utilization and research of Rubus. Working as a team and applying the latest omics strategies may open the door for developing a series of satisfactory elite germplasms for fruit and medicine, and reveal the central evolutionary phenomenon resulting in reticulate evolution.

Author Contributions

Conceptualization, W.H. and Q.M.; writing—original draft preparation, Q.M.; preparation of diagram and editing, Q.M. and H.M.; writing—review and editing, Q.M. and W.H.; supervision, W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Lushan Botanical Garden, Chinese Academy of Sciences, grant number 2021ZWZX11, and the National Natural Science Foundation of China, grant number 32160099.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Focke, W.O. Species Ruborum Monographiae Generis Rubi Prodromus; Bibliotheca Botanica; E. Schweizerbart: Stuttgart, Germany, 1910. [Google Scholar]
  2. Focke, W.O. Species Ruborum Monographiae Generis Rubi Prodromus; Bibliotheca Botanica; E. Schweizerbart: Stuttgart, Germany, 1911. [Google Scholar]
  3. Focke, W.O. Species Ruborum Monographiae Generis Rubi Prodromus; Bibliotheca Botanica; E. Schweizerbart: Stuttgart, Germany, 1914. [Google Scholar]
  4. Jennings, D. Raspberries and Blackberries: Their Breeding, Disease and Growth; Academic Press: New York, NY, USA, 1988. [Google Scholar]
  5. Janick, J.; Moore, J.N. Fruit Breeding, Vine and Small Fruits; Wiley: New York, NY, USA, 1996; Volume II, pp. 109–190. [Google Scholar]
  6. Lu, L.T.; Gu, C.Z.; Li, C.L.; Alexander, C.; Bartholomew, B.; Brach, A.R.; Boufford, D.E.; Ikeda, H.; Ohba, H.; Robertson, S.A.; et al. Rubus Linnaeus. Flora China 2003, 9, 195–285. [Google Scholar]
  7. Alice, L.A.; Campbell, C.S. Phylogeny of Rubus (rosaceae) based on nuclear ribosomal DNA internal transcribed spacer region sequences. Am. J. Bot. 1999, 86, 81–97. [Google Scholar] [CrossRef] [PubMed]
  8. Carter, K.A.; Liston, A.; Bassil, N.V.; Alice, L.A.; Bushakra, J.M.; Sutherland, B.L.; Mockler, T.C.; Bryant, D.W.; Hummer, K.E. Target Capture Sequencing Unravels Rubus Evolution. Front. Plant Sci. 2019, 10, 1615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Lu, L.T. A study on the genus Rubus of China. Acta Phytotaxon. Sin. 1983, 21, 13–25. [Google Scholar]
  10. Kalkman, C. The phylogeny of the Rosaceae. Bot. J. Linn. Soc. 1988, 98, 37–59. [Google Scholar] [CrossRef]
  11. Gu, Y.; Sun, Z.J.; Cai, J.H.; Huang, Y.S.; He, S.A. Introduction and utilization of small fruits in China, with special refercence to Rubus species. Acta Hortic. 1989, 262, 47–56. [Google Scholar] [CrossRef]
  12. Foster, T.M.; Bassil, N.V.; Dossett, M.; Leigh Worthington, M.; Graham, J. Genetic and genomic resources for Rubus breeding: A roadmap for the future. Hortic. Res. 2019, 6, 116. [Google Scholar] [CrossRef] [Green Version]
  13. Yu, G.; Luo, Z.; Wang, W.; Li, Y.; Zhou, Y.; Shi, Y. Rubus chingii Hu: A Review of the Phytochemistry and Pharmacology. Front. Pharmacol. 2019, 10, 799. [Google Scholar] [CrossRef]
  14. Zhu, Y.A.; Li, C.; Wang, S.; Lei, L.; Xiao, J. The complete chloroplast genome of Rubus setchuenensis, an edible and medicinal dual-purpose wild plant. Mitochondrial DNA Part B Resour. 2022, 7, 228–230. [Google Scholar] [CrossRef]
  15. Moyer, R.A.; Hummer, K.E.; Finn, C.E.; Frei, B.; Wrolstad, R.E. Anthocyanins, phenolics, and antioxidant capacity in diverse small fruits: Vaccinium, rubus, and ribes. J. Agric. Food Chem. 2002, 50, 519–525. [Google Scholar] [CrossRef]
  16. Guo, Q.L.; Gao, J.Y.; Yang, J.S. Analysis of Bioactive Triterpenes from Rubus chingii by Cyclodextrin-Modified Capillary Electrophoresis. Chromatographia 2005, 62, 145–150. [Google Scholar] [CrossRef]
  17. Guo, Q.L.; Yang, J.S.; Liu, J.X. Studies on Chemical Constituents in Fruits of Rubus chingii. Chin. Pharm. J. 2007, 42, 1141–1143. [Google Scholar]
  18. Hummer, K.E.; Janick, J. Rubus Iconography: Antiquity to the Renaissance. Acta Hortic. 2007, 759, 89–106. [Google Scholar] [CrossRef]
  19. Kaume, L.; Howard, L.R.; Devareddy, L. The blackberry fruit: A review on its composition and chemistry, metabolism and bioavailability, and health benefits. J. Agric. Food Chem. 2012, 60, 5716–5727. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, L.; Lei, T.; Han, G.; Yue, J.; Zhang, X.; Yang, Q.; Ruan, H.; Gu, C.; Zhang, Q.; Qian, T.; et al. The chromosome-scale reference genome of Rubus chingii Hu provides insight into the biosynthetic pathway of hydrolyzable tannins. Plant J. 2021, 107, 1466–1477. [Google Scholar] [CrossRef]
  21. Sun, N.; Wang, Y.; Liu, Y.; Guo, M.-L.; Yin, J. A new ent-labdane diterpene saponin from the fruits of Rubus chingii. Chem. Nat. Compd. 2013, 49, 49–53. [Google Scholar] [CrossRef]
  22. Zhang, T.-T.; Yang, L.; Jiang, J.-G. Bioactive comparison of main components from unripe fruits of Rubus chingii Hu and identification of the effective component. Food Funct. 2015, 6, 2205–2214. [Google Scholar] [CrossRef]
  23. Zhong, R.; Guo, Q.; Zhou, G.; Fu, H.; Wan, K. Three new labdane-type diterpene glycosides from fruits of Rubus chingii and their cytotoxic activities against five humor cell lines. Fitoterapia 2015, 102, 23–26. [Google Scholar] [CrossRef]
  24. Li, K.; Xia, X.X.; Ding, Y.H.; Zhou, W.Y.; Xie, B.S.; Cai, J.B.; Xie, Y.H.; Huang, L.P. Effect of Different Extract Parts From Rubi Fructus on Improving Memory Disorder in Mice. Chin. J. Exp. Tradit. Med. Formulae 2016, 22, 142–147. [Google Scholar] [CrossRef]
  25. Li, K.Y.; Zeng, M.L.; Li, Q.L.; Zhou, B.H. Identification of polyphenolic composition in the fruits of Rubus chingii Hu and its antioxidant and antiproliferative activity on human bladder cancer T24 cells. J. Food Meas. Charact. 2018, 13, 51–60. [Google Scholar] [CrossRef]
  26. Rieseberg, L.H.; Willis, J.H. Plant speciation. Science 2007, 317, 910–914. [Google Scholar] [CrossRef] [PubMed]
  27. Jones, F.C.; Grabherr, M.G.; Chan, Y.F.; Russell, P.; Mauceli, E.; Johnson, J.; Swofford, R.; Pirun, M.; Zody, M.C.; White, S.; et al. The genomic basis of adaptive evolution in threespine sticklebacks. Nature 2012, 484, 55–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Theis, A.; Ronco, F.; Indermaur, A.; Salzburger, W.; Egger, B. Adaptive divergence between lake and stream populations of an East African cichlid fish. Mol. Ecol. 2014, 23, 5304–5322. [Google Scholar] [CrossRef] [PubMed]
  29. Ferris, K.G.; Barnett, L.L.; Blackman, B.K.; Willis, J.H. The genetic architecture of local adaptation and reproductive isolation in sympatry within the Mimulus guttatus species complex. Mol. Ecol. 2017, 26, 208–224. [Google Scholar] [CrossRef]
  30. Thompson, M.M. Survey of Chromosome Numbers in Rubus (Rosaceae: Rosoideae). Ann. Mo. Bot. Gard. 1997, 84, 128–164. [Google Scholar] [CrossRef]
  31. Wang, Y.; Chen, Q.; Chen, T.; Tang, H.; Liu, L.; Wang, X. Phylogenetic Insights into Chinese Rubus (Rosaceae) from Multiple Chloroplast and Nuclear DNAs. Front. Plant Sci. 2016, 7, 968. [Google Scholar] [CrossRef] [Green Version]
  32. Tutin, T.G.; Heywood, V.H.; Burges, N.A.; Moore, D.M.; Valentine, D.H.; Walters, S.M.; Webb, D.A. (Eds.) Rubus. Flora Europaea; Cambridge University Press: Cambridge, UK, 1980; Volume 5, 452p. [Google Scholar]
  33. Thompson, M.M.; Zhao, C.M. Chromosome numbers of Rubus Species in southwest China. Acta Hortic. 1993, 352, 493–502. [Google Scholar] [CrossRef]
  34. Zhang, T.-T.; Wang, M.; Yang, L.; Jiang, J.-G.; Zhao, J.-W.; Zhu, W. Flavonoid glycosides from Rubus chingii Hu fruits display anti-inflammatory activity through suppressing MAPKs activation in macrophages. J. Funct. Foods 2015, 18, 235–243. [Google Scholar] [CrossRef]
  35. Alice, L.A.; Eriksson, T.; Eriksen, B.; Campbell, C.S. Hybridization and Gene Flow Between Distantly Related Species of Rubus (Rosaceae): Evidence from Nuclear Ribosomal DNA Internal Transcribed Spacer Region Sequences. Syst. Bot. 2001, 26, 769–778. [Google Scholar] [CrossRef]
  36. Sochor, M.; Vašut, R.J.; Sharbel, T.F.; Trávníček, B. How just a few makes a lot: Speciation via reticulation and apomixis on example of European brambles (Rubus subgen. Rubus, Rosaceae). Mol. Phylogenet. Evol. 2015, 89, 13–27. [Google Scholar] [CrossRef]
  37. Xiang, Y.; Huang, C.-H.; Hu, Y.; Wen, J.; Li, S.; Yi, T.; Chen, H.; Xiang, J.; Ma, H. Evolution of Rosaceae Fruit Types Based on Nuclear Phylogeny in the Context of Geological Times and Genome Duplication. Mol. Biol. Evol. 2017, 34, 262–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Kim, M.J.; Sutton, K.L.; Harris, G.K. Raspberries and Related Fruits. In Encyclopedia of Food and Health; Caballero, B., Finglas, P.M., Toldrá, F., Eds.; Academic Press: Oxford, UK, 2016; pp. 586–591. [Google Scholar]
  39. Siriwoharn, T.; Wrolstad, R.E.; Finn, C.E.; Pereira, C.B. Influence of cultivar, maturity, and sampling on blackberry (Rubus, L. Hybrids) anthocyanins, polyphenolics, and antioxidant properties. J. Agric. Food Chem. 2004, 52, 8021–8030. [Google Scholar] [CrossRef] [PubMed]
  40. Nile, S.H.; Park, S.W. Edible berries: Bioactive components and their effect on human health. Nutrition 2014, 30, 134–144. [Google Scholar] [CrossRef] [PubMed]
  41. Croge, C.P.; Cuquel, F.L.; Pintro, P.; Biasi, L.A.; Bona, C. Antioxidant Capacity and Polyphenolic Compounds of Blackberries Produced in Different Climates. HortScience A Publ. Am. Soc. Hortic. Sci. 2019, 54, 2209–2213. [Google Scholar] [CrossRef]
  42. Ancos, B.d.; González, E.M.; Cano, M.P. Ellagic acid, vitamin C, and total phenolic contents and radical scavenging capacity affected by freezing and frozen storage in raspberry fruit. J. Agric. Food Chem. 2000, 48, 4565–4570. [Google Scholar] [CrossRef] [Green Version]
  43. Haffner, K.; Rosenfeld, H.J.; Skrede, G.; Wang, L. Quality of red raspberry Rubus idaeus L. cultivars after storage in controlled and normal atmospheres. Postharvest Biol. Technol. 2002, 24, 279–289. [Google Scholar] [CrossRef]
  44. Mazur, S.P.; Nes, A.; Wold, A.B.; Remberg, S.F.; Aaby, K. Quality and chemical composition of ten red raspberry (Rubus idaeus L.) genotypes during three harvest seasons. Food Chem. 2014, 160, 233–240. [Google Scholar] [CrossRef]
  45. Wang, S.Y.; Chen, C.-T.; Wang, C.Y. The influence of light and maturity on fruit quality and flavonoid content of red raspberries. Food Chem. 2009, 112, 676–684. [Google Scholar] [CrossRef]
  46. Skrovankova, S.; Sumczynski, D.; Mlcek, J.; Jurikova, T.; Sochor, J. Bioactive Compounds and Antioxidant Activity in Different Types of Berries. Int. J. Mol. Sci. 2015, 16, 24673–24706. [Google Scholar] [CrossRef] [Green Version]
  47. Gülçin, İ.; Topal, F.; Çakmakçı, R.; Bilsel, M.; Gören, A.C.; Erdogan, U. Pomological Features, Nutritional Quality, Polyphenol Content Analysis, and Antioxidant Properties of Domesticated and 3 Wild Ecotype Forms of Raspberries (Rubus idaeus L.). J. Food Sci. 2011, 76, C585–C593. [Google Scholar] [CrossRef]
  48. Bobinaitė, R.; Viškelis, P.; Venskutonis, P.R. Variation of total phenolics, anthocyanins, ellagic acid and radical scavenging capacity in various raspberry (Rubus spp.) cultivars. Food Chem. 2012, 132, 1495–1501. [Google Scholar] [CrossRef] [PubMed]
  49. Verma, R.; Gangrade, T.; Punasiya, R.; Ghulaxe, C. Rubus fruticosus (blackberry) use as an herbal medicine. Pharmacogn. Rev. 2014, 8, 101–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Strik, B.; Clark, J.; Finn, C.; Banados, M. Worldwide Blackberry Production. HortTechnology 2007, 17, 205–213. [Google Scholar] [CrossRef] [Green Version]
  51. Hassani, S.; Shariatpanahi, M.; Tavakoli, F.; Nili-Ahmadabadi, A.; Abdollahi, M. The changes of bioactive ingredients and antioxidant properties in various berries during jam processing International Journal of Biosciences|IJB. Int. J. Biosci. IJB 2015, 6, 172–179. [Google Scholar] [CrossRef]
  52. Tanksley, S.D.; McCouch, S.R. Seed banks and molecular maps: Unlocking genetic potential from the wild. Science 1997, 277, 1063–1066. [Google Scholar] [CrossRef] [Green Version]
  53. Clark, R.M.; Linton, E.; Messing, J.; Doebley, J.F. Pattern of diversity in the genomic region near the maize domestication gene tb1. Proc. Natl. Acad. Sci. USA 2004, 101, 700–707. [Google Scholar] [CrossRef] [Green Version]
  54. Flint-Garcia, S.A. Genetics and consequences of crop domestication. J. Agric. Food Chem. 2013, 61, 8267–8276. [Google Scholar] [CrossRef]
  55. Li, J.; Yuan, D.; Wang, P.; Wang, Q.; Sun, M.; Liu, Z.; Si, H.; Xu, Z.; Ma, Y.; Zhang, B.; et al. Cotton pan-genome retrieves the lost sequences and genes during domestication and selection. Genome Biol. 2021, 22, 119. [Google Scholar] [CrossRef]
  56. Graham, J.; Smith, K.; MacKenzie, K.; Jorgenson, L.; Hackett, C.; Powell, W. The construction of a genetic linkage map of red raspberry (Rubus idaeus subsp. idaeus) based on AFLPs, genomic-SSR and EST-SSR markers. Theor. Appl. Genet. Theor. Angew. Genet. 2004, 109, 740–749. [Google Scholar] [CrossRef]
  57. Bushakra, J.M.; Stephens, M.J.; Atmadjaja, A.N.; Lewers, K.S.; Symonds, V.V.; Udall, J.A.; Chagné, D.; Buck, E.J.; Gardiner, S.E. Construction of black (Rubus occidentalis) and red (R. idaeus) raspberry linkage maps and their comparison to the genomes of strawberry, apple, and peach. Theor. Appl. Genet. Theor. Angew. Genet. 2012, 125, 311–327. [Google Scholar] [CrossRef]
  58. Randell, R.; Howarth, D.; Morden, C. Genetic analysis of natural hybrids between endemic and alien Rubus (Rosaceae) species in Hawai‘i. Conserv. Genet. 2004, 5, 217–230. [Google Scholar] [CrossRef]
  59. Castillo, N.; Reed, B.M.; Graham, J.; Fernández-Fernández, F.; Bassil, N. Microsatellite Markers for Raspberry and Blackberry. J. Am. Soc. Hortic. Sci. 2010, 135, 271–278. [Google Scholar] [CrossRef] [Green Version]
  60. Bushakra, J.M.; Lewers, K.S.; Staton, M.E.; Zhebentyayeva, T.; Saski, C.A. Developing expressed sequence tag libraries and the discovery of simple sequence repeat markers for two species of raspberry (Rubus L.). BMC Plant Biol. 2015, 15, 258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Chen, Z.; Jiang, J.; Shu, L.; Li, X.; Huang, J.; Qian, B.; Wang, X.; Li, X.; Chen, J.; Xu, H. Combined transcriptomic and metabolic analyses reveal potential mechanism for fruit development and quality control of Chinese raspberry (Rubus chingii Hu). Plant Cell Rep. 2021, 40, 1923–1946. [Google Scholar] [CrossRef]
  62. Zhou, Z.M.; Yan, D.M.; Wang, Y.K.; Zhang, T.; Xiao, X.R.; Dai, M.Y.; Zhang, S.W.; Liu, H.N.; Li, F. Discovery of quality markers in Rubus chingii Hu using UPLC-ESI-QTOF-MS. J. Pharm. Biomed. Anal. 2021, 203, 114200. [Google Scholar] [CrossRef] [PubMed]
  63. Zeng, H.J.; Liu, Z.; Wang, Y.P.; Yang, D.; Yang, R.; Qu, L.B. Studies on the anti-aging activity of a glycoprotein isolated from Fupenzi (Rubus chingii Hu.) and its regulation on klotho gene expression in mice kidney. Int. J. Biol. Macromol. 2018, 119, 470–476. [Google Scholar] [CrossRef]
  64. Sheng, J.Y.; Wang, S.Q.; Liu, K.H.; Zhu, B.; Zhang, Q.Y.; Qin, L.P.; Wu, J.J. Rubus chingii Hu: An overview of botany, traditional uses, phytochemistry, and pharmacology. Chin. J. Nat. Med. 2020, 18, 401–416. [Google Scholar] [CrossRef]
  65. Xie, Y.H.; Lian, B.; Gong, J.H.; Tu, L.D.; Zhang, Y.T.; Huang, L.P. Preparation of Magnetic Chitosan Hyamine Microspheres and Separation of Phenolic Acids from Rubus Chingii Hu. Adv. Mater. Res. 2013, 634–638, 1347–1351. [Google Scholar] [CrossRef]
  66. Cai, Y.Q.; Hu, J.H.; Qin, J.; Sun, T.; Li, X.L. Rhododendron Molle (Ericaceae): Phytochemistry, pharmacology, and toxicology. Chin. J. Nat. Med. 2018, 16, 0401–0410. [Google Scholar] [CrossRef]
  67. Li, X.; Sun, J.; Chen, Z.; Jiang, J.; Jackson, A. Characterization of carotenoids and phenolics during fruit ripening of Chinese raspberry (Rubus chingii Hu). RSC Adv. 2021, 11, 10804–10813. [Google Scholar] [CrossRef]
  68. Ding, H.-Y. Extracts and constituents of Rubus chingii with 1,1-diphenyl-2-picrylhydrazyl (DPPH) free radical scavenging activity. Int. J. Mol. Sci. 2011, 12, 3941–3949. [Google Scholar] [CrossRef] [PubMed]
  69. Yao, X.; Zhu, W.R.; Huang, H.L.; Zeng, Y.R.; Yu, W.W. Effective medicinal ingredients and screening of excellent germplasm in Rubus chingii. China J. Chin. Mater. Med. 2021, 46, 575–581. [Google Scholar] [CrossRef]
  70. Madrigal-Gamboa, V.; Jiménez-Arias, J.; Hidalgo, O.; Quesada, S.; Pérez, A.M.; Azofeifa, G. Membrane processing effect of blackberry (Rubus adenotrichos) on cytotoxic and pro-apoptotic activities against cancer cell lines. J. Food Process. Preserv. 2021, 45, e15575. [Google Scholar] [CrossRef]
  71. Xu, W.; Zhao, M.; Fu, X.; Hou, J.; Wang, Y.; Shi, F.; Hu, S. Molecular mechanisms underlying macrophage immunomodulatory activity of Rubus chingii Hu polysaccharides. Int. J. Biol. Macromol. 2021, 185, 907–916. [Google Scholar] [CrossRef]
  72. Tolentino, F.; Araújo, P.A.d.; Marques, E.d.S.; Petreanu, M.; Andrade, S.F.d.; Niero, R.; Perazzo, F.F.; Rosa, P.C.P.; Maistro, E.L. In vivo evaluation of the genetic toxicity of Rubus niveus Thunb. (Rosaceae) extract and initial screening of its potential chemoprevention against doxorubicin-induced DNA damage. J. Ethnopharmacol. 2015, 164, 89–95. [Google Scholar] [CrossRef] [PubMed]
  73. Kanegusuku, M.; Benassi, J.C.; Pedrosa, R.C.; Yunes, R.A.; Filho, V.C.; Maia, A.A.; de Souza, M.M.; Delle Monache, F.; Niero, R. Cytotoxic, hypoglycemic activity and phytochemical analysis of Rubus imperialis (Rosaceae). Z. Naturforschung. C J. Biosci. 2002, 57, 272–276. [Google Scholar] [CrossRef] [PubMed]
  74. Om, A.-S.; Song, Y.-N.; Noh, G.; Kim, H.; Choe, J. Nutrition Composition and Single, 14-Day and 13-Week Repeated Oral Dose Toxicity Studies of the Leaves and Stems of Rubus coreanus Miquel. Molecules 2016, 21, 65. [Google Scholar] [CrossRef] [Green Version]
  75. AlQahtani, F.S.; AlShebly, M.M.; Govindarajan, M.; Senthilmurugan, S.; Vijayan, P.; Benelli, G. Green and facile biosynthesis of silver nanocomposites using the aqueous extract of Rubus ellipticus leaves: Toxicity and oviposition deterrent activity against Zika virus, malaria and filariasis mosquito vectors. J. Asia-Pac. Entomol. 2017, 20, 157–164. [Google Scholar] [CrossRef]
  76. Ali, N.; Shaoib, M.; Shah, S.W.A.; Shah, I.; Shuaib, M. Pharmacological profile of the aerial parts of Rubus ulmifolius Schott. BMC Complement. Altern. Med. 2017, 17, 59. [Google Scholar] [CrossRef] [Green Version]
  77. Ke, H.; Bao, T.; Chen, W. New function of polysaccharide from Rubus chingii Hu: Protective effect against ethyl carbamate induced cytotoxicity. J. Sci. Food Agric. 2021, 101, 3156–3164. [Google Scholar] [CrossRef]
  78. Kawahara, Y.; de la Bastide, M.; Hamilton, J.P.; Kanamori, H.; McCombie, W.R.; Ouyang, S.; Schwartz, D.C.; Tanaka, T.; Wu, J.; Zhou, S.; et al. Improvement of the Oryza sativa Nipponbare reference genome using next generation sequence and optical map data. Rice 2013, 6, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Jiao, Y.; Peluso, P.; Shi, J.; Liang, T.; Stitzer, M.C.; Wang, B.; Campbell, M.S.; Stein, J.C.; Wei, X.; Chin, C.-S.; et al. Improved maize reference genome with single-molecule technologies. Nature 2017, 546, 524–527. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, M.; Tu, L.; Yuan, D.; Zhu, D.; Shen, C.; Li, J.; Liu, F.; Pei, L.; Wang, P.; Zhao, G.; et al. Reference genome sequences of two cultivated allotetraploid cottons, Gossypium hirsutum and Gossypium barbadense. Nat. Genet. 2019, 51, 224–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. 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]
  82. Wu, J.; Wang, Z.; Shi, Z.; Zhang, S.; Ming, R.; Zhu, S.; Khan, M.A.; Tao, S.; Korban, S.S.; Wang, H.; et al. The genome of the pear (Pyrus bretschneideri Rehd.). Genome Res. 2013, 23, 396–408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Benz, B.F. Archaeological evidence of teosinte domestication from Guilá Naquitz, Oaxaca. Proc. Natl. Acad. Sci. USA 2001, 98, 2104–2106. [Google Scholar] [CrossRef] [Green Version]
  84. Huang, X.; Yang, S.; Gong, J.; Zhao, Q.; Feng, Q.; Zhan, Q.; Zhao, Y.; Li, W.; Cheng, B.; Xia, J.; et al. Genomic architecture of heterosis for yield traits in rice. Nature 2016, 537, 629–633. [Google Scholar] [CrossRef]
  85. Chen, P.; Li, Z.; Zhang, D.; Shen, W.; Xie, Y.; Zhang, J.; Jiang, L.; Li, X.; Shen, X.; Geng, D.; et al. Insights into the effect of human civilization on Malus evolution and domestication. Plant Biotechnol. J. 2021, 19, 2206–2220. [Google Scholar] [CrossRef]
  86. Varshney, R.K.; Nayak, S.N.; May, G.D.; Jackson, S.A. Next-generation sequencing technologies and their implications for crop genetics and breeding. Trends Biotechnol. 2009, 27, 522–530. [Google Scholar] [CrossRef] [Green Version]
  87. Ray, S.; Satya, P. Next generation sequencing technologies for next generation plant breeding. Front. Plant Sci. 2014, 5, 367. [Google Scholar] [CrossRef] [Green Version]
  88. Ramirez-Gonzalez, R.H.; Segovia, V.; Bird, N.; Fenwick, P.; Holdgate, S.; Berry, S.; Jack, P.; Caccamo, M.; Uauy, C. RNA-Seq bulked segregant analysis enables the identification of high-resolution genetic markers for breeding in hexaploid wheat. Plant Biotechnol. J. 2015, 13, 613–624. [Google Scholar] [CrossRef]
  89. Julio, E.; Malpica, A.; Cotucheau, J.; Bachet, S.; Volpatti, R.; Decorps, C.; Dorlhac de Borne, F. RNA-Seq analysis of Orobanche resistance in Nicotiana tabacum: Development of molecular markers for breeding recessive tolerance from ‘Wika’ tobacco variety. Euphytica 2019, 216, 6. [Google Scholar] [CrossRef]
  90. Hyun, T.K.; Lee, S.; Kumar, D.; Rim, Y.; Kumar, R.; Lee, S.Y.; Lee, C.H.; Kim, J.Y. RNA-seq analysis of Rubus idaeus cv. Nova: Transcriptome sequencing and de novo assembly for subsequent functional genomics approaches. Plant Cell Rep. 2014, 33, 1617–1628. [Google Scholar] [CrossRef]
  91. VanBuren, R.; Bryant, D.; Bushakra, J.M.; Vining, K.J.; Edger, P.P.; Rowley, E.R.; Priest, H.D.; Michael, T.P.; Lyons, E.; Filichkin, S.A.; et al. The genome of black raspberry (Rubus occidentalis). Plant J. 2016, 87, 535–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. VanBuren, R.; Wai, C.M.; Colle, M.; Wang, J.; Sullivan, S.; Bushakra, J.M.; Liachko, I.; Vining, K.J.; Dossett, M.; Finn, C.E.; et al. A near complete, chromosome-scale assembly of the black raspberry (Rubus occidentalis) genome. GigaScience 2018, 7, giy094. [Google Scholar] [CrossRef] [PubMed]
  93. Zhu, Q.; Tian, Y.; Liang, W. The complete chloroplast genome sequence of Rubus eucalyptus (Rosaceae). Mitochondrial DNA Part B Resour. 2019, 4, 2976–2977. [Google Scholar] [CrossRef] [Green Version]
  94. Zhang, G.; Liu, Y.; Hai, P. The complete chloroplast genome of Tibetan medicinal plant Rubus phoenicolasius Maxim. Mitochondrial DNA Part B Resour. 2021, 6, 886–887. [Google Scholar] [CrossRef]
  95. Davik, J.; Røen, D.; Lysøe, E.; Buti, M.; Rossman, S.; Alsheikh, M.; Aiden, E.L.; Dudchenko, O.; Sargent, D.J. A chromosome-level genome sequence assembly of the red raspberry (Rubus idaeus L.). PLoS ONE 2022, 17, e0265096. [Google Scholar] [CrossRef]
  96. Zhu, J.K. The Future of Gene-Edited Crops in China. Natl. Sci. Rev. 2022, nwac063. [Google Scholar] [CrossRef]
  97. Li, X.; Jiang, J.; Chen, Z.; Jackson, A. Transcriptomic, Proteomic and Metabolomic Analysis of Flavonoid Biosynthesis During Fruit Maturation in Rubus chingii Hu. Front. Plant Sci. 2021, 12, 706667. [Google Scholar] [CrossRef]
  98. Li, X.; Sun, J.; Chen, Z.; Jiang, J.; Jackson, A. Metabolite profile and genes/proteins expression in β-citraturin biosynthesis during fruit ripening in Chinese raspberry (Rubus chingii Hu). Plant Physiol. Biochem. 2021, 163, 76–86. [Google Scholar] [CrossRef] [PubMed]
  99. Huang, W.; Qiao, F.; Guo, W.; Wu, W. Characterization of the complete chloroplast genome sequence of Rubus rufus Focke (Rosaceae). Mitochondrial DNA Part B Resour. 2021, 6, 3093–3094. [Google Scholar] [CrossRef] [PubMed]
  100. Park, Y.S.; Park, J.Y.; Kang, J.H.; Lee, W.H.; Yang, T.J. Diversity and authentication of Rubus accessions revealed by complete plastid genome and rDNA sequences. Mitochondrial DNA Part B Resour. 2021, 6, 1454–1459. [Google Scholar] [CrossRef] [PubMed]
  101. Wang, Q.; Huang, Z.; Gao, C.; Ge, Y.; Cheng, R. The complete chloroplast genome sequence of Rubus hirsutus Thunb. and a comparative analysis within Rubus species. Genetica 2021, 149, 299–311. [Google Scholar] [CrossRef]
  102. Su, X.H.; Duan, R.; Sun, Y.Y.; Wen, J.F.; Kang, D.G.; Lee, H.S.; Cho, K.W.; Jin, S.N. Cardiovascular effects of ethanol extract of Rubus chingii Hu (Rosaceae) in rats: An in vivo and in vitro approach. J. Physiol. Pharmacol. Off. J. Pol. Physiol. Soc. 2014, 65, 417–424. [Google Scholar]
  103. He, Y.; Jin, S.; Ma, Z.; Zhao, J.; Yang, Q.; Zhang, Q.; Zhao, Y.; Yao, B. The antioxidant compounds isolated from the fruits of chinese wild raspberry Rubus Chingii Hu. Nat. Prod. Res. 2018, 34, 872–875. [Google Scholar] [CrossRef]
  104. Chen, Y.; Chen, Z.; Guo, Q.; Gao, X.; Ma, Q.; Xue, Z.; Ferri, N.; Zhang, M.; Chen, H. Identification of Ellagitannins in the Unripe Fruit of Rubus Chingii Hu and Evaluation of its Potential Antidiabetic Activity. J. Agric. Food Chem. 2019, 67, 7025–7039. [Google Scholar] [CrossRef]
  105. Wang, S.Y.; Lin, H.S. Antioxidant activity in fruits and leaves of blackberry, raspberry, and strawberry varies with cultivar and developmental stage. J. Agric. Food Chem. 2000, 48, 140–146. [Google Scholar] [CrossRef]
  106. Fu, Y.; Zhou, X.; Chen, S.; Sun, Y.; Shen, Y.; Ye, X. Chemical composition and antioxidant activity of Chinese wild raspberry (Rubus hirsutus Thunb.). LWT Food Sci. Technol. 2015, 60, 1262–1268. [Google Scholar] [CrossRef]
  107. Han, N.; Gu, Y.; Ye, C.; Cao, Y.; Liu, Z.; Yin, J. Antithrombotic activity of fractions and components obtained from raspberry leaves (Rubus chingii). Food Chem. 2012, 132, 181–185. [Google Scholar] [CrossRef]
  108. Sellappan, S.; Akoh, C.C.; Krewer, G. Phenolic compounds and antioxidant capacity of Georgia-grown blueberries and blackberries. J. Agric. Food Chem. 2002, 50, 2432–2438. [Google Scholar] [CrossRef] [PubMed]
  109. Lee, M.J.; Nam, J.H.; Jeong, J.H.; Rho, I.R. Effect of plant part, extraction method, and harvest time over antioxidant yield of rubus coreanus. Pharmacogn. Mag. 2020, 16, 455–461. [Google Scholar]
  110. Howarth, D.G.; Gardner, D.E.; Morden, C.W. Phylogeny of Rubus subgenus Idaeobatus (Rosaceae) and its implications toward colonization of the Hawaiian Islands. Syst. Bot. 1997, 22, 433–441. [Google Scholar] [CrossRef]
  111. Morden, C.W.; Gardner, D.E.; Weniger, D.A. Phylogeny and biogeography of pacific Rubus subgenus Idaeobatus (Rosaceae) species: Investigating the origin of the endemic Hawaiian raspberry R. macraei. Pac. Sci. 2003, 57, 181–197. [Google Scholar] [CrossRef]
  112. Wang, Y.; Chen, Q.; Chen, T.; Zhang, J.; He, W.; Liu, L.; Luo, Y.; Sun, B.; Zhang, Y.; Tang, H.-r.; et al. Allopolyploid origin in Rubus (Rosaceae) inferred from nuclear granule-bound starch synthase I (GBSSI) sequences. BMC Plant Biol. 2019, 19, 303. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, Y.; Chen, Q.; Tang, H.; Wang, X. Phylogeny of Chinese Rubus (Rosaceae) based on nuclear ribosomal internal transcribed spacer (ITS) sequences. Acta Hortic. 2018, 1208, 13–24. [Google Scholar] [CrossRef]
  114. Alice, L.A.; Dodson, T.M.; Sutherland, B.L. Diversity and relationships of Bhutanese Rubus (Rosaceae). Acta Hortic. 2008, 777, 63–70. [Google Scholar] [CrossRef]
  115. Yang, J.Y.; Pak, J.-H. Phylogeny of korean rubus (rosaceae) based on its (nrDNA) and trnL/F intergenic region (cpdna). J. Plant Biol. 2006, 49, 44–54. [Google Scholar] [CrossRef]
  116. Zhang, L.; Wang, Y.; Chen, Q.; Luo, Y.; Zhang, Y.; Tang, H.R.; Wang, X.R. Phylogenetic Utility of Chinese Rubus (Rosaceae) Based on ndhF Sequence. Acta Hortic. Sin. 2015, 42, 19–30. [Google Scholar] [CrossRef]
  117. Imanishi, H.; Nakahara, K.; Tsuyuzaki, H. Genetic relationships among native and introduced Rubus species in Japan based on rbcL sequence. Acta Hortic. 2008, 769, 195–199. [Google Scholar] [CrossRef]
Figure 1. Different aspects of Rubus species utilization. The attributes of Rubus are primarily utilized for applications involving fruit (fresh fruit, jam, and juice), medicinal compounds (fruit, leaf, and stem), and scientific studies (adaptation, reproduction, polyploidy, and evolution). In addition, among the wild Rubus species, some can also be used to produce cosmetics or fiber products. This figure was created using BioRender software.
Figure 1. Different aspects of Rubus species utilization. The attributes of Rubus are primarily utilized for applications involving fruit (fresh fruit, jam, and juice), medicinal compounds (fruit, leaf, and stem), and scientific studies (adaptation, reproduction, polyploidy, and evolution). In addition, among the wild Rubus species, some can also be used to produce cosmetics or fiber products. This figure was created using BioRender software.
Plants 11 01211 g001
Figure 2. Schematic description of the secondary metabolites of Rubus and their medicinal benefits to humans. The central red rectangle represents the wild Rubus species, the seven green ellipses represent the major secondary metabolites of Rubus, and the outermost ellipses represent the main medicinal functions of secondary metabolites. The various colors show different effects on human health.
Figure 2. Schematic description of the secondary metabolites of Rubus and their medicinal benefits to humans. The central red rectangle represents the wild Rubus species, the seven green ellipses represent the major secondary metabolites of Rubus, and the outermost ellipses represent the main medicinal functions of secondary metabolites. The various colors show different effects on human health.
Plants 11 01211 g002
Table 2. List of Rubus L. subgenera.
Table 2. List of Rubus L. subgenera.
SubgenusCodeSpecies in
Subgenus
Ploidy Level
(x = 7)
References
AnoplobatusAn92x[8,30,111,112]
ChamaebatusCb6 (5)2x, 6x[8,30,31]
ChamaemorusCm1 (1)6x, 8x[8,30]
ComaropsisCo24x[8,30]
CylactisCy18 (8)2x−4x[8,30,31]
DalibardaDa52x[8,30]
DalibardastrumDs15 (10)4x, 6x[8,30,31,111,112,113]
IdaeobatusId125 (83)2x, 3x, 4x, 13x, 18x[8,20,30,31,111,112,113]
LampobatusLa10 (1)4x[8,30]
MalachobatusMa104 (85)4x, 6x, 8x, 14x[30,31,111,112,113]
OrobatusOr166x[8,30]
RubusRu444 (1)2x−12x[8,30,111,112]
The number within brackets is the corresponding number of Rubus species in China.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Meng, Q.; Manghwar, H.; Hu, W. Study on Supergenus Rubus L.: Edible, Medicinal, and Phylogenetic Characterization. Plants 2022, 11, 1211. https://doi.org/10.3390/plants11091211

AMA Style

Meng Q, Manghwar H, Hu W. Study on Supergenus Rubus L.: Edible, Medicinal, and Phylogenetic Characterization. Plants. 2022; 11(9):1211. https://doi.org/10.3390/plants11091211

Chicago/Turabian Style

Meng, Qinglin, Hakim Manghwar, and Weiming Hu. 2022. "Study on Supergenus Rubus L.: Edible, Medicinal, and Phylogenetic Characterization" Plants 11, no. 9: 1211. https://doi.org/10.3390/plants11091211

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