Skip to main content

Comparative analysis of medicinal plants Scutellaria baicalensis and common adulterants based on chloroplast genome sequencing

Abstract

Background

Scutellaria baicalensis Georgi has been extensively used as a medicinal herb in China for over 2000 years. They may be intentionally or inadvertently substituted or blended with comparable species in the local market, threatening clinical medication safety. Molecular markers are effective tools to prevent misidentification and eliminate doping and falsification among Scutellaria plants. This study screened four highly variable regions to identify Scutellaria and its adulterants. In addition, a phylogenetic analysis was performed using the complete cp genome combined with published Scutellaria species samples. Moreover, a comparative analysis of the cp genomes was conducted to investigate the cp genome evolution of S. baicalensis.

Results

The complete cp genome of five species of Scutellaria was sequenced for the first time, and four previously published Scutellaria species were re-sequenced. They all exhibited a conserved quadripartite structure in their cp genomes, including two distinct regions, namely a small and large single copy region, respectively, and two inverted repeats encompassing the majority of ribosomal RNA genes. Furthermore, the nine species exhibited high conservation from aspects of the genome structure, codon usage, repeat sequences, and gene content. Four highly variable regions (matK-rps16, ndhC-trnV-UAC, psbE-petL, and rps16-trnQ-UUG) may function as potential molecular markers for differentiating S. baicalensis from its adulterants. Additionally, the monophyly of Scutellaria was ascertained and could be reclassified into two subgenera, subgenus Anaspis and subgenus Scutellaria, as evidenced by the phylogenetic analyses on sequences of cp genome and shared protein-coding sequences. According to the molecular clock analysis, it has been inferred that the divergence of Scutellaria occurred at approximately 4.0 Mya during the Pliocene Epoch.

Conclusion

Our study provides an invaluable theoretical basis for further Scutellaria species identification, phylogenetics, and evolution analysis.

Peer Review reports

Background

Scutellaria baicalensis Georgi, a member of the Lamiaceae family and commonly known as "Huang qin" in Chinese, has been listed in the Chinese Pharmacopoeia due to its critical medicinal properties, such as clearing heat, purging fire, eliminating dampness, preventing miscarriage, and maintaining hemostasis [1]. However, previous studies showed that S. baicalensis is often contaminated by other common adulterants, such as S. indica L., S. yunnanensis H.Lév., S. tenax W.W.Sm., S. forrestii Diels, and S. caryopteroides Hand.-Mazz [2]. These substitute herbs are commonly of substandard quality and may even contain harmful constituents [3]. Given their similar morphology and shared vernacular name, distinguishing them solely by sight can be highly challenging and misleading. Currently, the use of molecular markers has made noteworthy advancements in the field of Chinese medicine identification. This approach aims to discriminate individuals from populations by sequencing particular genomic regions [4]. The utilization of universal DNA barcodes, specifically matK, rpl32-trnl, ndhF-rpl32, and trnL-trnF, have been employed to differentiate and verify S. baicalensis from its dopants [5,6,7,8]. However, these investigations did not cover certain prevalent adulterants, and single-locus DNA barcodes have inherent limitations [2, 9]. Subsequent experiments have identified shortcomings in the universal DNA barcodes for identifying common contaminants in the market, as illustrated in Fig. S1 and Fig. S2. Therefore, it is crucial to develop a more accurate and effective method for distinguishing S. baicalensis from its common contaminants.

The chloroplast (cp) is involved in plant photosynthesis and numerous biochemical processes [10, 11]. Unlike the nuclear genome, the cp genome is comparatively stable, exhibiting only minor variations [12, 13], which has been successfully employed for discriminating Amomum, Isodon, and their contaminants [14, 15]. Although the cp genomes of some Scutellaria species have been described [16, 17], these studies mainly characterized a single genome or compared the intraspecific variation. Nonetheless, Scutellaria species and their common adulterants have never been discriminated by comparing their cp genomes. In addition, gene-encoding regions diverge more slowly than non-coding regions, and therefore, they could offer an improved resolution for understanding phylogenetic relationships [18, 19]. Nevertheless, protein-coding genes have not been used to comprehend the evolutionary relationship of Scutellaria.

Here, de novo sequencing and assembly, as well as annotation of the cp genomes of nine distinct Scutellaria species, were conducted to (i) enhance the comprehension of the overall cp genome structure of Scutellaria, (ii) elucidate the phylogenetic relationships of Scutellaria, and (iii) filter candidate molecular markers to differentiate S. baicalensis from its adulterants. Therein, the cp genomes of five related species, including S. indica, S. yunnanensis, S. tenax, S. forrestii, and S. caryopteroides, were published for the first time. The findings enrich the available genomic resources for Scutellaria by offering crucial information to support phylogenetic analysis, distinguish the Scutellaria genus, and facilitate the safe medical applications of S. baicalensis.

Results

Characterization of cp genomes

For each species, clean data totaling approximately 2-5 Gb were acquired. The complete cp genomes were obtained after assembly and splicing (Fig. 1). The annotated results showed that among the nine species, the cp genome of S. likiangensis was the smallest, with 151,654 bp, while that of S. tenax was the largest (152,093 bp) (Table 1). The length of the LSC, SSC, and IR regions varied from 83,741 bp (S. forrestii) to 84,102 bp (S. caryopteroides), 17,326 bp (S. likiangensis) to 17,549 bp (S. tenax), and 25,229 bp (S. caryopteroides) to 25,255 bp (S. amoena), respectively.

Fig. 1
figure 1

Cp genome map of Scutellaria. Genes inside the circle are transcribed clockwise, while those outside are transcribed counterclockwise

Table 1 Summary of cp genome features

From the cp genomes of nine species, 124-132 genes were identified, including 85-87 protein-coding genes, 31-37 tRNA genes, and eight rRNA genes. Therein, 18 genes have two replicates in IR regions (Table 1). Briefly, 12 repeat genes were located in the LSC, including trnK-UUU, trnG-GCC, trnL-UAA, trnV-UAC, clpP, petB, petD, rpl16, rps16, atpF, rpoC1, and ycf3. Four repetitive genes in the IR included ndhB, rpl2, trnI-GAU, and trnA-UGC, whereas ndhA was the only one present in the SSC region. Notably, Scutellaria contained 16 repeated genes involved in photosynthesis and self-replication (Table 2). Additionally, the cp genomes among the studied species displayed high conservation in the GC content at approximately 38.3%.

Table 2 List of genes in the cp genome of the nine Scutellaria species

Codon usage analysis

The amino acid frequency, the number and bias of codon usage, and RSCU were investigated among the nine Scutellaria species cp genomes. The results indicated that 21 different amino acids were encoded in the cp genome, and a comprehensive set of 64 codons were deduced. Among these codons, 30 were frequently utilized in various Scutellaria species (Table S1-S9). 31 codons displayed RSCU values >1, 28 of which have A or U terminal nucleotides. The rest of the 33 codons had RSCU values ≤1, where 30 of these ended in G or C nucleotides. Additionally, the codon UUA had the highest frequency, followed by AGA, while AGC was the least common (Fig. 2). Leucine and cysteine had the highest and least number of codons, respectively. Moreover, unlike other amino acids were encoded by at least two synonymous codons, methionine, and tryptophan were encoded by only one respective codon. The GC content of synonymous third codon positions (GC3s) is correlated with codon bias to evaluate codon usage patterns. In nine species of Scutellaria, the GC3s values ranged from 38.20% to 39.10%, indicating a strong bias towards A/U-ending codons. Besides, both optimal frequency values and the codon adaptation index were less than 0.5, and the effective proportion of codons ranged from 55.47% to 55.90% (Table S10). The codon usage of cp genomes in S. baicalensis, S. likiangensis, and S. amoena was relatively similar (Fig. 2), suggesting that a minor bias existed in codon use across the nine Scutellaria species.

Fig. 2
figure 2

Heatmap of the RSCU values among nine Scutellaria species

Repeat analysis

Repetitive sequences have a significant impact on genome evolution and rearrangements. A total of 273 long repeats, including 144 P repeats, 128 F repeats, and one R repeat, were identified, while C repeats were absent (Fig. S3 A and Table S11). The repeats ranged in length from 30 bp to 129 bp, with 30 bp-39 bp being the most frequent and those longer than 70 bp being the least abundant (Table S11). Moreover, as the Hamming distance decreased from 3 to 1, the number of repeat sequences decreased significantly from 273 to 86 (Table S12).

SSRs are important genetic markers that facilitate the identification of closely related species [20, 21]. A total of 33,44, 34, 31, 44, 36, 33, 37, and 41 SSRs were obtained in the cp genomes of S. likiangensis, S. tenax, S. barbata, S. baicalensis, S. yunnanensis, S. indica, S. caryopteroides, S. forrestii, and S. amoena, respectively (Table S13, Fig. S3 B). In this study, the analysis of SSRs revealed that mononucleotide repeats in identified SSRs proportioned 54.55% - 64.86%, with the A/T motif being the most prevalent. The subsequent most commonly observed SSR types were dinucleotide repeats (12.20% - 18.18%) with a predominance of the AT/TA motif and tetranucleotide repeats (10.81% - 15.91%) with a predominance of the ATTT/AAAT motif. Trinucleotide repeats accounted for 3.22% - 9.76% of the SSRs dominated by the AAT/ATT motif. Finally, hexanucleotide repeats were observed at 0% - 16.13%, while pentanucleotide repeats were present at a frequency of 0% - 6.06%.

Contraction and expansion of IRs

In the cp genome, the structure is characterized by a circular tetrad composed of the LSC, SSC, and IRs regions. The contraction and expansion of the IRs regions lead to variable genome size [20]. As illustrated in Fig. 3, the structure and connection between IRs regions varied slightly in the 24 Scutellaria species. Specifically, a shortened copy of the rps19 gene was detected at the joint point of the IRb and LSC regions from all the species analyzed. The truncated gene originated in the IRb region and was integrated into the LSC region by 4 - 84 bp fragments. Meanwhile, this pseudogene was also identified in S. tenax, S. yunnanensis, S. microviolacea, and S. rehderiana. At the junction of IRb/SSC regions, a truncated copy of ndhF gene was inserted into the IRb region with 15 - 45 bp fragments, while rpl2 was exclusively located in the IRb region. Besides, a truncated copy of ycf1 gene was observed in all species at the IRa/SSC junction. No ycf1 pseudogene was found in S. likiangensis, S. barbata, S. baicalensis, S. indica, S. caryopteroides, S. forrestii, S. tsinyunensis, S. microviolacea, S. meehanioides, and S. rehderiana in the IRb region. In addition, the trnN genes were in the IRa region, while the trnH gene of all species was in the LSC. These results indicated that the cp genomes of nine Scutellaria species exhibited a distinctive pattern of IRs contraction and expansion.

Fig. 3
figure 3

Comparisons of IR, LSC, and SSC regions amongst 24 Scutellaria cp genomes. The numbers above denoted the distance between the gene ends, and the region sites, and the numbers below indicate their responding size. The features are not to scale

Genome comparison and nucleotide diversity

A comparative analysis was performed to assess the extent of divergence. The findings of this study indicated that the cp genomes of nine Scutellaria species were highly conserved, with a greater degree of conservation in the protein-coding regions than in the non-coding regions. Therein, notable mutations were observed in ycf1 and petD genes (Fig. 4). The most differentiation of non-coding regions was observed in the petN-psbM, rps16-trnQ-UUG, ndhC-trnV-UAC, rbcL-accD, accD-psaI, and rpl16-rps3 intergenic spacers (IGSs). Additionally, the sliding window analysis revealed that the average Pi value of 87 protein-coding genes was 0.0061. Three regions, namely matK-rps16, psbE-petL, and trnN-GUU-trnR-ACG, exhibited highly variable Pi values of > 0.02, with psbE-petL displaying the lowest divergence value of 0.026. The trnN-GUU-trnR-ACG regions displayed the highest divergence value of 0.031 (Fig. 5). Nine highly polymorphic regions, including rps16-trnQ-UUG, petN-psbM, ndhC-trnV-UAC, rbcL-accD, accD-psaI, rpl16-rps3, matK-rps16, psbE-petL, and trnN-GUU-trnR-ACG, were identified as promising molecular markers for distinguish S. baicalensis and its contaminants.

Fig. 4
figure 4

Comparisons of nine cp genomes using S. baicalensis (GenBank OP597814) annotation as a reference. The vertical scale ranged from 50% to 100% represents the percentage of identity. Arrows represent the transcriptional direction of each annotated gene

Fig. 5
figure 5

Sliding window analysis of the cp genomes among nine Scutellaria species

Species authentication analysis

The identification of highly variable regions within the cp genomes represents a valuable resource to distinguish closely related species and provides crucial insights for conducting phylogenetic analyses [22, 23]. An initial ML analysis was conducted on nine IGSs separately, with the default parameters. As shown in Fig. S4, S. baicalensis exhibited discernible differentiation from its commonly encountered adulterants based on analysis of matK-rps16, ndhC-trnV-UAC, psbE-petL, and rps16-trnQ-UUG loci, whereas the remaining IGSs could not provide sufficient discrimination with low bootstrap values. Therefore, a more comprehensive ML tree was constructed using a combination of these four IGS regions (Fig. S4). In this tree, S. baicalensis occupied a distinct branch, indicating the efficacy of these loci in accurately distinguishing it from prevalent contaminants.

Additionally, primers were designed for the nine IGSs (Table S14), and amplification and sequencing experiments were conducted. The results showed that the four IGSs regions (rps16-trnQ-UUG, ndhC-trnV-UAC, matK-rps16, and psbE-petL) produced products of the expected sizes in selected Scutellaria species (Fig. S5). Both amplification and sequencing achieved a 100% success rate, and each species displayed distinct base differences (Fig. S6A-D). In contrast, universal DNA barcodes (ITS, psbA-trnH, matK, rbcL, and trnL-trnF) could not identify S. baicalensis from its contaminants (Table S15). These findings confirm that the four IGSs are ideal for distinguishing S. baicalensis from its adulterants.

Phylogenetic and divergence time analyses

Phylogenetic analysis was performed for 28 sequences, including both cp genomes and CDSs, with Holmskioldia sanguinea utilized as the outgroup. Most nodes on the phylogenetic tree showed strong support (Fig. 6 and Fig. S7). ML analysis of cp genomes and CDS sequences gave nearly identical topologies with varying levels of support at a few nodes. The involved species were grouped into clades A and B. The former was further divided into the subgenera Anaspis and Scutellaria, which contained six species, i.e., S. amoena, S. likiangensis, S. baicalensis, S. altaica, S. przewalskii, and S. kingiana. The monophyly of S. baicalensis was strongly supported with 100% bootstrap value, suggesting that S. baicalensis could be differentiated from its adulterants by the cp genome or shared CDS sequences. Clade B included two subgenera, Scutellaria and Scutellariopsis, and encompassed the remaining species, indicating relatively weak support for the current division of three subgenera of Scutellaria. These results contribute significantly to the taxonomy of Scutellaria and may have implications for the identification and conservation of relevant species.

Fig. 6
figure 6

The constructed ML phylogenetic tree among 28 common cp genes from 24 species

The divergence time estimation showed that clades A and B diverged during the Pliocene Epoch at approximately 4.0 Mya (Fig. S8). Additionally, the most recent common ancestor of the extant subgenera Scutellaria and Anaspis dates back to the beginning of the late Pleistocene, approximately 2.52 Mya.

Discussion

Cp genome organizations

The cp genome is a valuable tool for studying species identification, population genetics, phylogenetics, and gene engineering [20, 24]. The de novo assembly and comparative analysis of nine cp genomes of Scutellaria were performed in this study, including five species that are being reported for the first time. The size of cp genomes ranged from 151,654 bp to 152,093 bp, mainly resulting from gene insertion and deletion in the non-coding regions. The cp genome sizes of the nine Scutellaria species fell within the known range of the genus, ranging from 151,654 bp of S. likiangensis to 152,731 bp of S. baicalensis [25]. The cp genomes of all species were highly similar in total length, GC content, and gene composition, consistent with earlier findings [7] and similar to other Lamiaceae species. At the genus level, the cp genome of angiosperms was highly conserved [26]. However, the rpl36, psbJ, trnfM-CAU, trnS-UGA, trnG-UCC, and trnQ-UUG genes/IGSs were lost in S. indica, suggesting that these variations may be specific to this particular species.

Codon usage and comparative analyses

Codons are critical in linking genetic materials, amino acids, and proteins within organisms [13]. We found over 90% (RSCU ≧ 1) of Scutellaria codons terminated in A/U, and the GC3s value ranged from 38.2% to 39.1%. Scutellaria exhibited a high coding efficiency and a strong preference for A/U termination codons, possibly owing to the overall high AT content in the cp genome. This trend was also evident in other angiosperms [13, 15].

Our investigation of Scutellaria genomes revealed that noncoding regions displayed more substantial variation than coding regions. Four highly variable regions in non-coding regions identified in this study effectively distinguished the most common Scutellaria species. These findings aligned with earlier studies of Scutellaria [2]. Non-coding regions evolve rapidly and contain valuable variations for genus phylogenetic analysis [27]. Therefore, the significance of non-coding regions in the cp genome for identifying Scutellaria species should be underscored.

Identification and phylogenetic analysis

Several molecular studies have demonstrated the high identification capability of cp genetic markers in distinguishing Phyllanthus and Isodon species [15, 28]. Four regions (matK-rps16, ndhC-trnV-UAC, psbE-petL, and rps16-trnQ-UUG) have been filtered as candidate molecular markers for discriminating S. baicalensis from its common adulterants in this study. The matK-rps16 region can distinguish Triticum species [29], and ndhC-trnV-UAC, psbE-petL, and rps16-trnQ-UUG have been proposed to identify other species [24, 30]. A possible reason was the different species involved in the screening process, which may have excluded important information. Besides, although universal DNA barcodes, such as matK, psbA-trnH, and ITS, have been demonstrated to differentiate S. baicalensis from closely related species [3, 31, 32], they did not effectively identify common confounding taxa. In contrast, the four identified IGSs were theoretically found to be capable of effectively distinguishing S. baicalensis from contaminants, a finding that was further validated in experiments.

The phylogenetic analyses between the cp genome and CDSs showed that S. baicalensis occupied an autonomous branch and the sister relationship between S. baicalensis and either S. kingiana or S. amoena was strongly supported, which agreed with previous findings about ITS and ITS2 regions [8, 32]. Notably, S. barbata (Genbank OP597813 and Genbank NC059814) individuals did not cluster and were in different branches, suggesting the intraspecific diversity of S. barbata. This finding was reminiscent of a previous study, where the cp genomes of Isodon rubescens from diverse geographical regions showed high intraspecific diversity [15]. The variation in the cp sequence of S. barbata might also be influenced by the geographical area of origin. Besides, S. kingiana and S. tuberifera, classified under subgenus Anaspis and subgenus Scutellariopsis in the Flora of China, respectively, were grouped with subgenus Scutellaria in this study. All studied Scutellaria species formed a well-supported clade (100%), dividing into two subclades. This challenges the three-subgenus classification proposed in the Flora of China. Notably, the monophyly of subgenus Anaspis might be untenable given the limited sample size, consisting of only one species, as aligned with a previous study [7]. Furthermore, most amino acids in CDSs are highly conserved, and phylogenetic analysis based on CDS sequences can be used for phylogenetic studies [33], which agrees with our results. Using cp genome data for classification at the genus or subgenus level has been widely recognized in several plant taxonomic groups [34]. Accordingly, we propose reclassifying the Scutellaria into two subgenera, namely, subgenus Anaspis and subgenus Scutellaria, supported by ITS-based studies [5]. Overall, the monophyly of S. baicalensis was verified based on the cp genomes and shared CDS sequences, which provides a basis for studying the species identification, phylogeny, and taxonomy of Scutellaria.

The divergence time of Scutellaria

The origins of Scutellaria can be traced back to approximately 4.0 Mya based on estimated divergence time, which agreed with previous studies [35, 36]. This timeline coincides with Pliocene, a geological interval from 5.30 to 2.60 Mya, which had a global temperature that was several degrees warmer than current levels. The Pliocene represents a suitable analog for a future anthropogenic greenhouse world [37]. During the early Pliocene, precipitation levels increased significantly, resulting in a warmer and wetter climate that facilitated the expansion of forests in the northern hemisphere and a decrease in global desert areas. These climate changes may influence the migration and diversification of terrestrial plants and lead to an outbreak of species in the ecological niches [38]. Hence, it can be deduced that the climatic conditions prevailing during the Pliocene epoch drove significantly the diversification of Scutellaria.

Conclusion

Here, the cp genomes of nine Scutellaria species were de novo assembled under the Illumina sequencing platform. Therein, the cp genomes of five species, i.e., S. indica, S. yunnanensis, S. tenax, S. forrestii, and S. caryopteroides, were reported for the first time. All species had relatively conserved cp genomes with similar genomic structure and gene content. Notably, four identified distinct and highly variable cp loci, namely matK-rps16, ndhC-trnV-UAC, psbE-petL, and rps16-trnQ-UUG harbor potential as molecular markers to differentiate S. baicalensis from its common dopants. Moreover, the monophyly of Scutellaria was confirmed, and both the cp genome-based phylogeny and CDS-based phylogeny proposed its reclassification into two subgenera, Anaspis and Scutellaria. The divergence time of Scutellaria was approximately 4.0 Mya during the Pliocene, and Pliocene climatic conditions are likely a significant factor contributing to the diversification of the genus. Overall, this study provides valuable insights into the safety, effectiveness, and bioprospecting and conservation of Scutellaria species.

Materials and methods

Plant materials and DNA sequencing

Healthy and fresh leaves from nine Scutellaria species, including S. indica, S. yunnanensis, S. tenax, S. forrestii, S. caryopteroides, S. baicalensis, S. likiangensis, S. barbata, and S. amoena, were gathered from the Germplasm Resource Garden, Kunming Zhifen Biotechnology Co., Ltd. in Yunnan, China (24°49′55′′N, 102°48′58″E). The detailed information for each sample was summarized in Fig. 7 and Table S16. All the voucher samples were authenticated by Professor Baozhong Duan and preserved in the herbarium of Dali University. A commercial kit DP305 (Tiangen, Beijing, China) was used to extract the total plant genome DNA. High-quality DNA was used for further sequencing. It is worth noting that PCR amplification studies were restricted to six species (S1-S6, S. likiangensis, S. tenax, S. barbata, S. baicalensis, S. yunnanensis, S. amoena) due to the lack of remaining DNA in three Scutellaria samples. Additionally, four other species (S10-S13, S. purpureocardia, S. weishanensis, S. teniana, and S. kingiana) (Table S16) were included in the PCR amplification experiment.

Fig. 7
figure 7

The phenotype of nine Scutellaria species

Assembly and annotation of Scutellaria

Approximately 3 μg DNA was employed to establish the shotgun libraries, sequenced via the Illumina NovaSeq system. NGS QC Toolkit v. 2.3.3 software was utilized to obtain clear reads by trimming paired-end sequence reads. De novo assemblies and annotation of the cp genomes were performed by using GetOrganelle v. 1.6.4 and online tools CpGAVAS2 [13, 39] and GeSeq [28, 40] with the parameter sets referred to previous studies [15, 41], respectively. The Apollo rectified intron/exon boundaries and start/stop codons [9, 42]. The cp genomes were mapped through IRscope (https://irscope.shinyapps.io/Chloroplot/). The accession numbers of the annotated cp genome sequences are available on the National Center for Biotechnology Information (NCBI) GenBank database (OP597811-OP597819) (Table 1).

Repeat sequence and codon preference analyses

The Geneious 9.0.2 software [43] was employed to conduct a GC content analysis. Furthermore, the REPuter program (https://bibiserv.cebitec.uni-bielefeld.de/reputer) was utilized to recognize dispersed repeat sequences, including forward (F), complementary (C), palindromic (P), and reverse (R) [44]. With the setting of >30 bp, ≥90% sequence identity, and a Hamming distance at 3. Simple sequence repeats (SSRs) in the cp genomes were analyzed on the MISA-web (http://pgrc.ipk-gatersleben.de/misa/) [45], and those with different repeat units were regarded as hexanucleotide, pentanucleotide, tetranucleotides, trinucleotides, dinucleotides, mononucleotide, respectively.

Additionally, the amino acid usage frequency and relative synonymous codon usage (RSCU) were identified via the CodonW software [46]. Lastly, TBtools, a software that integrates various biological data handling tools [47], generated a heatmap of the RSCU values.

Comparative and phylogenetic analyses

The comparative analysis of cp genomes of nine species was conducted using the mVISTA program in Shuffle-LAGAN mode [48], with S. baicalensis (GenBank OP597814) as the reference genome. The divergence in the contraction and expansion of the inverted repeat (IR) regions in the cp genomes was visualized by using the IRscope tool (https://irscope.shinyapps.io/irapp/) [49]. The cp genomes were aligned via the MAFFT software (https://mafft.cbrc.jp/alignment/server/index.html) [50]. The nucleotide diversity (Pi) was analyzed using the DnaSP v. 6.12.03 software with previously reported parameter settings [14].

A total of 28 complete cp genome sequences of Scutellaria, including 18 reported, were utilized for phylogenetic analysis (Table S17), and five new species were in this study. Holmskioldia sanguinea (Genbank MN128388) was selected as an outgroup taxa. Moreover, shared coding sequences (CDSs) were obtained from the complete plastomes, and S. baicalensis (GenBank OP597814) serves as the reference chloroplast genome. The alignment of all sequences and trimming were conducted by using MAFFT and TrimAl with the default parameters setting, respectively [50]. The maximum likelihood (ML) tree was constructed using IQ-TREE [51].

Identification and validation of barcode for species discrimination

The IGS were obtained from nine Scutellaria species using PhyloSuite v1.2.2 [52]. Primers were designed based on the variable intergenic regions using Snapgene 6.2.1 (Snapgene, Insightful Science, available at http://www.snapgene.com, last used in 2023). PCR amplifications were conducted in a final volume of 20 μL, comprising 10 μL of 2×Taq Plus PCR Master Mix, 1 μL of each primer, 2 μL of template DNA, and 6 μL of ddH2O. All amplifications were performed using a RePure-A PCR system (Applied Biogener, Hangzhou, China) under the following conditions: an initial denaturation at 95 ℃ for 4 min, followed by 40 cycles of 94 ℃ for 30 s, 56 ℃ for 1 min, and 72 ℃ for 1 min, with a final extension at 72 ℃ for 10 min. PCR products were examined by 1% agarose gel electrophoresis to confirm the amplification of the target fragments. The purified PCR products were sequenced in both directions on a 3730XL DNA Sequencer (Applied Biosystems, Waltham, USA) using the same primers at Vazyme Medical Technology (Nanjing, China).

Divergence time analysis

MEGA X was employed to create a molecular clock tree [53]. The relevant divergence time was then estimated through the TimeTree Resource (http://www.timetree.org/) [28, 54]. Three calibration nodes were employed: (F1) a median time of 2.85-3.85 million years ago (Mya) for the most recent common ancestor (MRCA) of S. amoena and S. baicalensis; (F2) a median time of 2.85-3.85 Mya for the common ancestor of S. baicalensis and S. barbata; and (F3) a median time of 0.11-1.11 Mya for the common ancestor of S. barbata and S. indica.

Availability of data and materials

The data provided in the study were submitted to the NCBI (https://www.ncbi.nlm.nih.gov/), and accession numbers are shown in Table 1.

References

  1. Chinese Pharmacopoeia Commission. Pharmacopoeia of the Peoples Republic of China: Beijing: Chin Med Sci Technol Press. 2020.

  2. Jiang D, Zhao ZY, Zhang T, Zhong WH, Liu CS, Yuan QJ, Huang LQ. The chloroplast genome sequence of Scutellaria baicalensis provides insight into intraspecific and interspecific chloroplast genome diversity in Scutellaria. Genes (Basel). 2017;8(9):227.

    Article  PubMed  Google Scholar 

  3. Guo XR, Wang XG, Su WH, Zhang GF, Zhou R. DNA barcodes for discriminating the medicinal plant Scutellaria baicalensis (Lamiaceae) and its adulterants. Biol Pharm Bull. 2011;34(8):1198–203.

    Article  CAS  Google Scholar 

  4. Xiong C, Sun W, Li JJ, Yao H, Shi YH, Wang P, et al. Identifying the species of seeds in traditional Chinese medicine using DNA barcoding. Front Pharmacol. 2018;9:701.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Safikhani K, Jamzad Z, Saeidi H. Phylogenetic relationships in Iranian Scutellaria (Lamiaceae) based on nuclear ribosomal ITS and chloroplast trnL-F DNA data. Plant Syst Evol. 2018;304(9):1077–89.

    Article  Google Scholar 

  6. Wang XG, Hai MR, Liu T. Analysis of trnL-F sequences of the medicinal plants Scutellaria baicalensis and its adulterants. Chin Agric Sci Bull. 2012;28(06):173–7.

    Google Scholar 

  7. Zhao F, Li B, Drew BT, Chen YP, Wang Q, Yu WB, Liu ED, Salmaki Y, Peng H, Xiang CL. Leveraging plastomes for comparative analysis and phylogenomic inference within Scutellarioideae (Lamiaceae). PLoS One. 2020;15(5):e232602.

    Article  Google Scholar 

  8. Xia Z, Feng CY, Gao ZM, Li HM, Zhang HR. Authentication of DNA barcoding of Scutellaria baicalensis and its related species. Chin Trad Her Drugs. 2014;45(01):107–12.

    CAS  Google Scholar 

  9. Lewis SE, Searle SMJ, Harris N, Gibson M, Lyer V, Richter J, et al. Apollo: a sequence annotation editor. Genome Biol. 2002;3:81–2.

    Article  Google Scholar 

  10. Khawla G Alwadani, Jasmine K Janes, Andrew RL. Chloroplast genome analysis of box-ironbark Eucalyptus. Mol Phylogenet Evol. 2019;136:76-86.

  11. Saski C, Lee SB, Fjellheim S, Guda C, Jansen RK, Luo H, Tomkins J, Rognli OA, Daniell H, Clarke JL. Complete chloroplast genome sequences of Hordeum vulgare, Sorghum bicolor and Agrostis stolonifera, and comparative analyses with other grass genomes. Theor Appl Genet. 2007;115(4):591.

    Article  CAS  PubMed  Google Scholar 

  12. Zhang TT, Xing YP, Xu L, Bao GH, Zhan ZL, Yang YY, et al. Comparative analysis of the complete chloroplast genome sequences of six species of Pulsatilla Miller Ranunculaceae. Chin Med. 2019;14:53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Wang J, Qian J, Jiang Y, Chen XC, Zheng BJ, Chen SL, Yang FF, Xu ZC, Duan BZ. Comparative analysis of chloroplast genome and new insights into phylogenetic relationships of Polygonatum and Tribe Polygonateae. Front Plant Sci. 2022;13:882189.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Cui YX, Chen XL, Nie LP, Sun W, Hu HY, Lin YL, Li HT, Zheng XL, Song JY, Yao H. Comparison and phylogenetic analysis of chloroplast genomes of three medicinal and edible Amomum species. Int J Mol Sci. 2019;20(16):4040.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Zhou ZY, Wang J, Pu TT, Dong JJ, Guan Q, Qian J, Shi LC, Duan BZ. Comparative analysis of medicinal plant Isodon rubescens and its common adulterants based on chloroplast genome sequencing. Front Plant Sci. 2022;13:1036277.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Li JY, Wang D, Xue PP, Sun HR, Feng QH, Miao N. The complete chloroplast genome of Scutellaria scordifolia (Labiatae). Mitochondrial DNA B Resour. 2021;6(1):84–5.

    Article  PubMed Central  Google Scholar 

  17. Liu XN, Zuo YW, Le Lin, Li WQ, Wang Q, Deng HP. The complete chloroplast genome of Scutellaria tsinyunensis (Lamiaceae), an endemic species from China. Mitochondrial DNA B Resour. 2020;5(3):2568-2570.

  18. Bock DG, Kane NC, Ebert DP, Rieseberg LH. Genome skimming reveals the origin of the Jerusalem Artichoke tuber crop species: neither from Jerusalem nor an artichoke. New Phytol. 2014;201(3):1021–30.

    Article  CAS  PubMed  Google Scholar 

  19. Shaw J, Shafer HL, Leonard OR, Kovach MJ, Schorr M, Morris AB. Chloroplast DNA sequence utility for the lowest phylogenetic and phylogeographic inferences in angiosperms: the tortoise and the hare IV. Am J Bot. 2014;101(11):1987–2004.

    Article  PubMed  Google Scholar 

  20. Zhao KH, Li LQ, Quan H, Yang JB, Zhang ZR, Liao ZH, Lan XZ. Comparative analyses of chloroplast genomes from 14 Zanthoxylum species: identification of variable DNA markers and phylogenetic relationships within the genus. Front Plant Sci. 2021;11:605793.

    Article  PubMed Central  Google Scholar 

  21. Zheng G, Wei LL, Ma L, Wu ZQ, Gu CH, Chen K. Comparative analyses of chloroplast genomes from 13 Lagerstroemia (Lythraceae) species: identification of highly divergent regions and inference of phylogenetic relationships. Plant Mol Biol. 2020;102(6):659–76.

    Article  CAS  PubMed  Google Scholar 

  22. Wu L, Wu ML, Cui N, Xiang L, Li Y, Li X, Chen SL. Plant super-barcode: a case study on genome-based identification for closely related species of Fritillaria. Chin Med. 2021;16(1):52.

  23. Yi DK, Lee HL, Sun BY, Chung MY, Kim KJ. The complete chloroplast DNA sequence of Eleutherococcus senticosus (Araliaceae); comparative evolutionary analyses with other three asterids. Mol Cells. 2012;33(5):497–508.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Zhou JG, Cui YX, Chen XL, Li Y, Xu ZC, Duan BZ, Li YH, Song JY, Yao H. Complete Chloroplast Genomes of Papaver rhoeas and Papaver orientale: molecular structures, comparative analysis, and phylogenetic analysis. Molecules. 2018;23(2):437.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Daniell H, Lin CS, Yu M, Chang WJ. Chloroplast genomes: diversity, evolution, and applications in genetic engineering. Genome Biol. 2016;17(1):1–29.

    Article  Google Scholar 

  26. Shahzadia I, Abdullaha, Furrukh M, Zain A, Ibrar A, Bushra M. Chloroplast genome sequences of Artemisia maritima and Artemisia absinthium: comparative analyses, mutational hotspots in genus Artemisia and phylogeny in family Asteraceae. Genomics. 2020;2(112):1454-1463.

  27. Shaw J, Lickey E, Schilling E, Small RL. Comparison of whole chloroplast genome sequences to choose noncoding regions for phylogenetic studies in angiosperms: the tortoise and the hare III. Am J Bot. 2007;94(3):275–88.

    Article  CAS  PubMed  Google Scholar 

  28. Fang H, Dai GN, Liao BB, Zhou P, Liu YL. Application of chloroplast genome in the identification of Phyllanthus urinaria and its common adulterants. Front Plant Sci. 2022;13:1099856.

    Article  PubMed  Google Scholar 

  29. Awad M, Fahmy RM, Mosa KA, Helmy M, El-Feky FA. Identification of effective DNA barcodes for Triticum plants through chloroplast genome-wide analysis. Comput Biol Chem. 2017;71:20–31.

    Article  CAS  PubMed  Google Scholar 

  30. Kim GB, Lim CE, Kim JS, Kim K, Lee JH, Yu HJ, Mun JH. Comparative chloroplast genome analysis of Artemisia (Asteraceae) in East Asia: insights into evolutionary divergence and phylogenomic implications. BMC Genomics. 2020;21(1):415.

    Article  CAS  PubMed Central  Google Scholar 

  31. Lv ZL, Zhang YG, Han M, Liu CJ. Sequence analysis of ITS2 and psbA-trnH DNA barcode of wild populations of Scutellaria baicalensis Georgi from different producing Areas. J Jilin Agric Univ. 2017;39(01):42–8.

    Google Scholar 

  32. Chen SY, Zhao R, Xu L, Feng QJ, Wang B, Kang TG. DNA barcoding of Mongolian medicinal plant Scutellaria scordifolia. J Chin Med Mater. 2016;39(07):1483–7.

    Google Scholar 

  33. Hong Z, He WC, Liu XJ, Tembrock LR, Wu ZQ, Xu DP, Liao XZ. Comparative analyses of 35 complete chloroplast genomes from the genus Dalbergia (Fabaceae) and the identification of DNA barcodes for tracking illegal logging and counterfeit rosewood. Forests. 2022;13(4):626.

    Article  Google Scholar 

  34. Yu XY, Zuo LH, Lu DD, Lu B, Yang MS, Wang JM. Comparative analysis of chloroplast genomes of five Robinia species: genome comparative and evolution analysis. Gene. 2019;689:141–51.

    Article  CAS  PubMed  Google Scholar 

  35. Chiang YC, Huang BH, Liao PC. Diversification, biogeographic pattern, and demographic history of Taiwanese Scutellaria species inferred from nuclear and chloroplast DNA. PLoS One. 2012;7(11):e50844.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Dong CC, Wang S, Zhang H, Liu JQ, Li MJ. Karyotype evolution of the Asterids insights from the first genome sequences of the family Cornaceae. DNA Res. 2023;30(1):c51.

    Article  Google Scholar 

  37. Sniderman JMK, Woodhead JD, Hellstrom J, Jordan GJ, Drysdale RN, Tyler JJ, Porch N. Pliocene reversal of late Neogene aridification. Proc Natl Acad Sci U S A. 2016;113(8):1999–2004.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Zuo YJ, Wen J, Zhou SL. Intercontinental and intracontinental biogeography of the eastern Asian-eastern north American disjunct Panax (the ginseng genus, Araliaceae), emphasizing its diversification processes in eastern Asia. Mol Phylogenet Evol. 2017;117:60–74.

    Article  PubMed  Google Scholar 

  39. Shi LC, Chen HM, Jiang M, Wang LQ, Wu X, Huang LF, Liu C. CPGAVAS2, an integrated plastome sequence annotator and analyzer. Nucleic Acids Res. 2019;47(W1):W65–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Tillich M, Lehwark P, Pellizzer T, Ulbricht-Jones ES, Fischer A, Bock R, Greiner S. GeSeq-versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 2017;45(W1):W6–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Jin JJ, Yu WB, Yang JB, Song Y, dePamphilis CW, Yi TS, Li DZ. GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biol. 2020;21(1):241.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Liang CL, Wang L, Lei J, Duan BZ, Ma WS, Xiao SM, et al. A comparative analysis of the chloroplast genomes of four Salvia medicinal plants. Engineering. 2019;5(5):907–15.

    Article  CAS  Google Scholar 

  43. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, et al. Geneious basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28(12):1647–9.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, Giegerich R. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001;29(22):4633–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Beier S, Thiel T, Münch T, Scholz U, Mascher M. MISA-web: a web server for microsatellite prediction. Bioinformatics. 2017;33(16):2583–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Sharp PM, Tuohy TM, Mosurski KR. Codon usage in yeast: cluster analysis clearly differentiates highly and lowly expressed genes. Nucleic Acids Res. 1986;14(13):5125–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Chen CJ, Chen H, Zhang Y, Thomas HR, Frank MH, He YH, Xia R. TBtools: an integrative toolkit developed for interactive analyses of big biological data. Mol Plant. 2020;13(8):1194–202.

    Article  CAS  PubMed  Google Scholar 

  48. Frazer KA, Pachter L, Poliakov A, Rubin EM, Dubchak I. VISTA: computational tools for comparative genomics. Nucleic Acids Res. 2004;32(Web Server):W273-W279.

  49. Amiryousefi A, Hyvönen J, Poczai P. IRscope: an online program to visualize the junction sites of chloroplast genomes. Bioinformatics. 2018;34(17):3030–1.

    Article  CAS  PubMed  Google Scholar 

  50. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Nguyen L, Schmidt HA, von Haeseler A, Minh B. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol. 2015;32(1):268–74.

    Article  CAS  PubMed  Google Scholar 

  52. Zhang D, Gao FL, Jakovlić I, Zou H, Zhang J, Li WX. Wang GT PhyloSuite: an integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol Ecol Res. 2020;20(1):348–55.

    Article  Google Scholar 

  53. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Kumar S, Stecher G, Suleski M, Hedges SB. TimeTree: a resource for timelines, time trees, and divergence times. Mol Biol Evol. 2017;34(7):1812–9.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

We also thank Northeast Forestry University and the China Academy of Chinese Medical Sciences for technical assistance.

Funding

This work was supported by key technology projects in the Yunnan province of China (202002AA100007).

Author information

Authors and Affiliations

Authors

Contributions

Zhen Li, Zhongyu Zhou, Jing Wang, Meihua Yang, and Ying Zhou participated in the conception and design of the research. Jing Wang analyzed and annotated the cp data. Zhongyu Zhou, Hui Fang, and Ying Zhou collected the species. Zhen Li is responsible for analyzing and processing data and writing the manuscript. Baozhong Duan, Conglong Xia, and Ying Zhou revised the paper. All authors agreed to the submitted version of the manuscript.

Corresponding authors

Correspondence to Ying Zhou or Jing Wang.

Ethics declarations

Ethics approval and consent to participate

The collecting of all samples in this study followed the Regulations on the Protection of Wild Plants of China, the IUCN Policy Statement on Research Involving Species at Risk of Extinction, and the Convention on the Trade in Endangered Species of Wild Fauna and Flora. All methods were carried out following relevant guidelines and regulations. In addition, healthy and fresh leaves from nine Scutellaria species, including S. indica, S. yunnanensis, S. tenax, S. forrestii, S. caryopteroides, S. baicalensis, S. likiangensis, S. barbata, and S. amoena, were gathered from the Germplasm Resource Garden, Kunming Zhifen Biotechnology Co., Ltd. in Yunnan, China (24°49′55′′N, 102°48′58″E). The detailed information for each sample was summarized in Fig. 7 and Table S16.

Consent for publication

Not applicable.

Competing interests

The authors declare the absence of competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

 Table S1. Codons in cp genome of S. likiangensis. Table S2. Codons in cp genome of S. tenax. Table S3. Codons in cp genome of S. barbata. Table S4. Codons in cp genome of S. baicalensis. Table S5. Codons in cp genome of S. yunnanensis. Table S6. Codons in cp genome of S. indica. Table S7. Codons in cp genome of S. caryopteroides. Table S8. Codons in cp genome of S. forrestii. Table S9. Codons in cp genome of S. amoena. Table S10. GC content at different positions of CDS sequence codon. Table S11. The number of fowarde (F), reverse (R), complementary (C), and palindromic (P) repeats in the cp genome. Table S12. The large repeated sequences in the nine Scutel laria cp genomes with diferent hamming distance. F: forward (direct) matching; R: reverse matching;C: complement matching; P: palindromic (inverted) matching. Table S13. Number of SSR types in the cp genome. Table S14. Primer design by SnapGene. Table S15. Universal DNA barcodes primers. Table S16. Information about the samples collected. Table S17. Species information downloaded by NCBI. Fig. S1. The gel electrophoresis results of universal DNA barcodes PCR products. Lane M was the marker of DL2000 Plus. The lanes from left to right corresponded: S1. S. likiangensis; S2. S. barbata; S3. S. yunnanensis; S4. S. amoena; S5. S. tenax; S6. S. baicalensis; S10. S. purpureocardia; S11. S. weishanensis; S12. S. teniana; S13. S. kingiana. Fig. S2. Phylogenetic tree created using the NJ technique based on the universal DNA barcodes (A. ITS; B. psbA-trnH; C.matK; D. rbcL; E. trnL-trnF). Fig. S3. A. Repeat sequences detected in Scutellaria cp genome. P, F, C, and R indicate the repeat types: R (Reverse repeats), P (Palindromic repeats), F (Forward repeats), C (Complement repeats); B. The number and type of SSRs in Scutellaria cp genome. Fig. S4. Phylogenetic tree created using the ML technique based on the cp genome's several IGS (A. accD-psaI; B. matK-rps16; C. ndhC-trnV-UAC; D. petN-psbM; E. psbE-petL; F. rbcL-accD; G. rpl16-rps3; H. rps16-trnQ-UUG; I. trnN-GUU-trnR-ACG; J. four IGSs (matK-rps16, ndhC-trnV-UAC, psbE-petL, and rps16-trnQ-UUG)). Fig. S5. The gel electrophoresis results of nine IGS PCR products. Lane M was the marker of DL2000 Plus. The lanes from left to right corresponded: S1. S. likiangensis; S2. S. barbata; S3. S. yunnanensis; S4. S. amoena; S5. S. tenax; S6. S. baicalensis; S10. S. purpureocardia; S11. S. weishanensis; S12. S. teniana; S13. S. kingiana. Fig. S6. Sequencing chromatograms of the psbE-petL barcode in seven Scutellaria specvies. Fig. S7. The ML phylogenetic tree based on shared coding sequences (CDS) of the 24 species. The bootstrap support values are listed at each node. Fig. S8. Divergence times estimation based on cp genomes. The node ages are given for each node.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Z., Duan, B., Zhou, Z. et al. Comparative analysis of medicinal plants Scutellaria baicalensis and common adulterants based on chloroplast genome sequencing. BMC Genomics 25, 39 (2024). https://doi.org/10.1186/s12864-023-09920-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-023-09920-2

Keywords