Next Article in Journal
Integrated Nutrient Management for Rice Yield, Soil Fertility, and Carbon Sequestration
Next Article in Special Issue
Assessment of ITS2 Region Relevance for Taxa Discrimination and Phylogenetic Inference among Pinaceae
Previous Article in Journal
DeepLRR: An Online Webserver for Leucine-Rich-Repeat Containing Protein Characterization Based on Deep Learning
Previous Article in Special Issue
DNA Barcoding of Two Thymelaeaceae Species: Daphne mucronata Royle and Thymelaea hirsuta (L.) Endl
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Challenges in Medicinal and Aromatic Plants DNA Barcoding—Lessons from the Lamiaceae

1
Biomolecular Technology Group, Leicester School of Allied Health Science, Faculty of Health and Life Sciences, De Montfort University, Leicester LE1 9BH, UK
2
Tree of Life Programme, Wellcome Trust Sanger Institute, Wellcome Genome Campus, Cambridge CB10 1SA, UK
*
Authors to whom correspondence should be addressed.
Plants 2022, 11(1), 137; https://doi.org/10.3390/plants11010137
Submission received: 23 November 2021 / Revised: 26 December 2021 / Accepted: 27 December 2021 / Published: 5 January 2022
(This article belongs to the Special Issue DNA Barcoding for Herbal Medicines)

Abstract

:
The potential value of DNA barcoding for the identification of medicinal plants and authentication of traded plant materials has been widely recognized; however, a number of challenges remain before DNA methods are fully accepted as an essential quality control method by industry and regulatory authorities. The successes and limitations of conventional DNA barcoding are considered in relation to important members of the Lamiaceae. The mint family (Lamiaceae) contains over one thousand species recorded as having a medicinal use, with many more exploited in food and cosmetics for their aromatic properties. The family is characterized by a diversity of secondary products, most notably the essential oils (EOs) produced in external glandular structures on the aerial parts of the plant that typify well-known plants of the basil (Ocimum), lavender (Lavandula), mint (Mentha), thyme (Thymus), sage (Salvia) and related genera. This complex, species-rich family includes widely cultivated commercial hybrids and endangered wild-harvested traditional medicines, and examples of potential toxic adulterants within the family are explored in detail. The opportunities provided by next generation sequencing technologies to whole plastome barcoding and nuclear genome sequencing are also discussed with relevant examples.

1. Introduction

1.1. Introducing the Historical Importance and Status of Medicinal Plants

It has been well documented that herbal plants and their derivatives play critical roles in pharmaceutical, cosmetics and food industries, e.g., [1,2,3,4,5,6,7,8,9]. Historically, plants have often been selected for drug development programs because they contain specific classes of compounds, such as alkaloids and terpenoids that are known to be biologically active, or because of their traditional medicinal use [10,11,12,13]. Jumping forward in the history, these compounds have been proven to be antioxidant, antibacterial and antiviral agents with some major advantages over conventional drug therapy and limited side effects [14,15,16,17,18]. Some volatile essential oils have also exhibited a high level of antiviral activity [19,20,21]. The early 1800s was a critical point in the use of medicinal plants. In these years, the discovery and isolation of alkaloids from different plants like poppy (1806), ipecacuanha (1817), Strychnos (1817), quinine (1820), pomegranate (1878), and the discovery of other active substances from medicinal plants such as tannins, saponosides, etheric oils, vitamins, hormones, etc. defined the beginning of scientific pharmacy [22]. This scientific movement away from raw herbal medicines into more refined products containing only the active products created a division between what is called modern medicine and traditional medicine (TM). Nevertheless, the use of TM is still strong, and demand is actually increasing. In the recent outbreak of COVID-19, various traditional herbal plants, including members of the Lamiaceae (Salvia L., Thymus L., Mentha L., Rosmarinus L. and Ocimum L.), have played important roles in the treatment and recovery of individuals with COVID-19, mainly in China and India [23,24,25,26,27,28,29,30].
The WHO reported in 2014 and 2019 [31,32] that an increased number of countries are acknowledging the role of TM in their national health systems, and an increasing number of member states had developed national policies on TM, launching national laws or regulations and implementing regulations on herbal medicines [31,32]. The attention to TM from many countries is deemed to grow further. The global herbal medicines market is anticipated to reach 129 billion US dollar at the compound annual growth rate (CAGR) of 5.88% during 2010–2023 [33].
Currently in the UK, herbal medicines are regulated by the Traditional Herbal Medicines Products Directive, which was devised by the European Union. This Directive requires evidence of a plant’s traditional use as a medicine for 30 years inside the EU or 15 years in the EU and 15 years elsewhere. This has been in place since 2004; however, it came fully into effect on 30 April 2011 [34]. This means that since 2011 all manufactured herbal medicines placed on the UK market are required to have either a Traditional Herbal Registration (THR) or a Marketing Authorization (MA). It is therefore envisaged that, in the near future, all herbal medicines on the market will have to meet the same stringent criteria, satisfying EU requirements applicable to any medicine: a consistently high standard of quality, regular monitoring of safety, and full information for safe and beneficial use of the product provided by in-pack leaflets [35].

1.2. Increasing TM Supply Demand Threatening “Wild Type” Stock

According to recent market research reports, it is estimated that the demand for global herbal medicines will significantly increase in the future [33]. Since ancient times, a variety of products of plant origin have been used in cosmetic products, including vegetable oils, other lipids and essential oils (EOs), and are gaining popularity over synthetic products [8]. To improve the quality of food, herbs and spices have always been recognized as safe, natural preservatives to increase the shelf life of food and are excellent substitutes for chemical additives [36,37,38].
At least 28,187 plant species were recorded as being used medicinally [4]. Unfortunately, the increasing demand for particular herbal products has led to the scarcity of wild populations of the medicinally important species. Another factor is the non-medicinal uses of medicinal plants, including their use as natural dyes, condiments and for ornamental purposes, which is also contributing to the extremely serious threats to certain populations. This in turn increases the frequency of species adulteration—when the target plant species is, deliberately or otherwise, substituted with different species—and thereby threatens consumer safety [39]. Despite significant progress in the herbal drug industry, the quality of herbal products remains a major issue of concern [40,41,42,43,44,45], with the substitution of different species, whether intentional or unintentional, at the forefront.

1.3. Herbal Medicines Quality Assurance Strategies

The currently available morphologic, organoleptic and chemical detection methods such as high-performance thin layer chromatography (HPTLC), high performance liquid chromatography (HPLC), ultraviolet (UV), infrared (IR), mass spectrometry (MS) and nuclear magnetic resonance (NMR) may not be sufficient for complete plant species identification. This occurs particularly when the plant material is in a powered form and where chemical variations occur due to geographic locations and storage conditions [44,46,47]. In addition, chemical identification is not species-specific and cannot distinguish species which share chemical markers [48,49,50]. Plant identification using micro-morphological, chemical and organoleptic methods can be time-consuming, error-prone and requires expertise and reliable references [51,52,53]. In comparison, DNA barcoding is more reliable, is not affected by external factors and can be applied to all tissues [54,55]. Over the past two decades, this approach has been increasingly accepted for the identification of plants.
DNA barcoding provides a way to confirm the authentication of plants and establish a level of quality assurance within the market [52,53,54,55,56,57,58,59]. Since the first DNA barcoding study [60], the “animal barcode”, a portion of the mitochondrial gene cytochrome oxidase 1(CO1), has proved remarkably effective at discriminating among species in diverse groups, such as birds, fishes, and insects [60,61]. However, the low substitution rate of the CO1 gene in plants was considered unsuitable for barcoding [62]. As a result of many plant barcoding studies, e.g., [62,63,64,65,66], in 2009, the Consortium for the Barcode of Life (CBOL) Plant Working Group proposed portions of two coding regions from the plastid genome, maturase K (matK) and ribulose-bisphosphate carboxylase (rbcL), as a standard 2-locus barcode for plants, to be supplemented with additional markers as required [67].
Proposed additional makers include the plastid intergenic spacer region of trnHand photosystem II protein D1 (trnH-psbA) and the internal transcribed spacers of nuclear ribosomal DNA (nrDNA ITS). These have generally been agreed to provide adequate resolution in a multi-locus barcode system [68,69,70,71]. Many other markers of plastid, mitochondrial and nuclear genomes, such as the trnL and trnF intergeneric spacer (trnL-F),RNA polymerase β subunit (rpoB), ATP synthase subunit b-delta (atpF-H), 5S-rRNA and 18S-rRNA have also been tested alone or in combination with respect to their discrimination capacity in plants and found to be appropriate for specific applications [54,72,73,74,75,76]
The nrDNA ITS is the most sequenced region across the plants with the most clearly defined barcode gap between inter- and intra-specific variations [77,78,79,80,81,82,83,84]. Components of the nrDNA ITS are ITS1, 5.8S and ITS2 regions (Figure 1) [85]. The ITS primers, ITS1 and ITS4 [86] were originally designed for fungi and found useful to detect fungal contamination in herbal plant samples [87,88,89,90].Sequences of 18S, 5.8S and 26S rDNA are highly conserved from bacteria, fungi and higher plants, enabling the design of the sequence-complemented universal primers for PCR amplification of ITS [91] across the kingdoms. To improve the quality of ITS sequence information in DNA-barcoding, there are plant-specific ITS primers that can avoid preferential amplification of fungal contaminants or non-plants templates [59,88,92,93]. Due to the decreased length of the ITS2 sequence (<300 bp), it has been proposed as a suitable for DNA barcoding applications in plants [68,94,95,96,97].There are issues, such as paralogy and polymorphic sites, with the ITS repeats [61,98] that make some taxonomists wary of using them, but for authentication purposes, ITS (and particularly ITS2) have advantages that tend to outweigh these issues.
As is evident from the lack of consensus regarding a single universal barcode for plants, it is likely that a more flexible approach will be required in order to make the best use of this technology for the benefit of consumers [99]. The British Pharmacopoeia, when introducing DNA barcoding for plant drugs, recognized this and approached each medicinal plant as a new ‘target species’ inhabiting a particular taxonomic environment. This enabled molecular markers to be selected for each target species, after analysis of each of the standard barcode regions, providing both identification of the target species and differentiation from potential adulterants [58,99]. The BP chose the trnA-psbA region of Ocimumtenuiflorum L. (Holy Basil Leaf) as their first exemplar of DNA barcoding as a tool for botanical identification, and the selection process is described in detail by Sgamma et al. [58] (see Example 1 of the Supporting Information in their publication).
The strategic application of DNA-based identification methods is best applied with a careful consideration of the specific trade, economic and taxonomic environment inhabited by a medicinal plant. The human preference for plant varieties or species based on characteristics that are desired for particular industries exerts a strong selective pressure and skews the material available on the international market toward the leading demand. This presents a challenging situation for those wishing to trade medicinal plants, as this is often not the primary market demand, and the material available may or may not represent the original, traditional, medicinal profile (genetic or phytochemical). These issues are described in this review using various Lamiaceae species as case studies to show the pressures of different markets and how these in turn effect the beneficial application of DNA-based authentication methods.

2. DNA Barcoding—Lessons from the Lamiaceae

The Lamiaceae (mint) is one of the largest families of aromatic plants and contains about 237 genera and 7756 species [4,100]. Of these, 1056 species are used as medicinal plants which is about 13.7% of the family, representing a higher-than-normal proportion [4]. The widely known herbal genera of the family such as Lavandula (lavender), Mentha (mint), Ocimum (basil), Scutellaria (skullcap), Thymus (thyme) have significant medicinal properties and are also major commodities in the food industry [100,101,102]. The Lamiaceae family has great diversity and variety with cosmopolitan distribution and inhabits different natural ecosystems. Some species of the family (e.g., lavender, basil, mint, oregano and thyme) are cultivated due to the high demand for medicines and food from individual species [101]. Many species are known for their aromatic properties due to the production of EOs in the glandular trichomes, one of the significant features of the family [103,104,105,106]. The plants in the family produce an enormous variety of compounds that act as attractants and defence molecules in nature and are also widely used by humans [107,108,109,110]. The EOs typical of the family are rich in terpenoids such as monoterpenes, iridoids and sesquiterpenes which are responsible for many of these functions.
Therefore, herbs and spices from this family are important in the pharmaceutical, flavouring, perfumery and cosmetic industries [111,112,113,114,115,116]. Global supply chains and consumer demand for particular characteristics exert selective pressure, and result in discrete and specific identification and authentication scenarios when attempting to select medicinally relevant material. These issues can be well described using case studies within the complex and species-rich Lamiaceae. Mentha (Mint) exemplifies a scenario of extremely strong consumer demand based around a particular EO, carvone. This skews the global market towards high yield varieties and is further complicated by ready hybridisation and human intervention via widespread cultivation and has led to traders adulterating their products to fulfil market demands. Lavandula (Lavender) is a vital contributor of fragrance industry and most famous for its relaxing aromatic qualities. Increasing demand for lavender extracts in the current market situation is due to two strong economic drivers, scent and horticulture. This dual pressure has resulted in a two-tier trade with varieties selected preferentially for one or the other driver. The rising demand for lavender products and the higher prices charged for English lavender due to its lower oil production per plant have led to lavender adulteration in the market. Ocimum (Basil) is widely used in systems of indigenous medicine and food. Migration of cultures from south Asia to different parts of the world has resulted in basil species becoming intermingled, making DNA authentication assays more difficult to interpret [117]. Origanum (Oregano) as a spice is utilised in numerous regions of the world. The herb has a strong culinary consumer demand and is widely cultivated for this purpose. Oregano is the name used to refer to a great variety of plants. Sixty-one species from seventeen genera in six different families are known as oregano [118], exemplifying the problem of trying to match scientific species with common plant names. Along with the existence of synonymous names, the adulteration of herbs may also be economically motivated and intentional. Scutellaria (skullcap) is mainly used in the pharmaceutical industry and misidentification due to high morphological similarities with its adulterants can lead to serious health issues [119]. Salvia (sage) is the largest genus of the Lamiaceae; most of the plants of this genus are well known for their nutritional components.
Thus, accurate plant identification is essential, in order to reduce the potential risks to the consumers’ well-being and safety. The benefit of DNA-based authentication in these arenas is considered, and how the objectives and approach of work must flex to fit the particular issues faced is discussed.

2.1. A Carvone Focussed Market and Hybridisation: Mentha L.—Mentheae: Nepetioideae

The genus Mentha (mint) is an important aromatic plant and consists of 24 species and 15 hybrids [120] and it is in high demand because of its carvone EO content (Figure 2). Some of the common species of Mentha such as M. aquatica L. (watermint), M. arvensis L. (cornmint), M. longifolia L., M. × piperita L. (peppermint), M. pulegium L., M. × rotundifolia (L.) Huds. and M. spicata L. (spearmint) are commonly grown for the production of EOs and/or utilized as food flavouring and medicinal agents in many countries of Europe, Australia, America, and the Middle East [121,122,123,124].
Based on cpDNA data, the genus is strongly supported as monophyletic (Figure 2), however, a phylogenetic understanding within the Mentha has always been challenging and it may be attributed to a high incidence of polyploidy, variation in base chromosome number, diverse morphology, vegetative propagation, and frequent interspecific hybridization both in wild and cultivated population [125,126,127,128,129]. The basic chromosome number of the genus is x = 12, but complex hybridization processes have led to a large diversity of chromosome numbers from diploid to octoploid [127].
Figure 2. Mintproducts market analysis [130] and phylogenetic relationship among the species.
Figure 2. Mintproducts market analysis [130] and phylogenetic relationship among the species.
Plants 11 00137 g002
M. spicata (2n = 48) has long been thought to originate by hybridization between M. longifolia and M. suaveolens, with a doubling of the chromosome number. However, in a recent study no evidence was found that M. spicata is of hybrid origin and revealed that many cryptic species were underestimated in subsection Spicatae [131]. Spearmint is widely grown throughout all regions of the world and the leaves possess a characteristic aromatic odour and pungent taste. Carvone is the main constituent of spearmint oil [132].
M. × piperita is a hybrid between M. spicata and M. aquatica [133]. The aromatic compounds of the genus, such as menthol, menthofuran, carvone, linalool, and linalyl acetate are frequently used as a part of confectionary, as flavor enhancing agents in toothpastes, chewing gums and beverages, bakery, cosmetics, as oral hygiene products, pharmaceuticals and pesticides [134,135]. Leaves, flowers and stems have been used as herbal teas and spices in many foods to add aroma and flavor [136,137]. The content of aromatic compound differs between species and also depends upon seasonal variations, soil types, etc. [138,139]. Peppermint has a characteristic aromatic odour and taste, with a cooling sensation on the breath, and menthol (35–40%) is the primary constituent of peppermint oil [140,141,142]
Carvone is a very important monoterpene ketone and occurs at high concentrations (70–80%) in spearmint oil and is also the major component responsible for its aroma [143]. Carvone can be used to identify spearmint, but it is also abundant in other species such as caraway (Carum carvi L.) and dill (Anethum graveolens L.), which consequently present major adulteration issues [144]. Another example is peppermint, with a large quantity of global demand produced in US. Peppermint EO has great importance in the flavour and food industries because of its unique sensory and quality properties. Cornmint, a less expensive mint plant is grown in India and is frequently used as a peppermint adulterant [49,129,145]. Another adulterant of peppermint is spearmint EO and L-menthol, which could be identified by using attenuated total reflectance-Fourier transform infrared (ATR-FTIR) spectroscopy coupled with partial least squares regression (PLSR) and principal component regression models, described in a recent study by Taylan et al. [146]. The DNA sequences rbcL, ITS, matK, trnH-psbA, atpB and atpC have been used as an approach to distinguish and identify the complex relationships among Mentha species [129,147,148,149,150]. The whole plastid genomes of M. spicata (Accession no.NC_037247.1), M. longifolia (Accession no. NC_032054.1) and M. × piperita (Accession no. NC_047475.1) have been sequenced and characterised to develop conservation strategies, metabolic engineering, molecular breeding and accurate identification of taxa [151,152]. Due to morphological, genetic plasticity and variation in active components of Eos with respect to geographic origin of Mentha species/subspecies/cultivar, accurate identification is essential for explanation of phylogenetic relatedness and distinctive marker profiles at the DNA level.
There are basically two types of challenges in the correct identification of Mentha species:
(i)
Hybridization or cryptic taxa. Hybridization and polyploidy have indeed most likely played important roles during speciation in mints, which forms one reason the number of taxonomically valid species is a subject of controversy [153,154].The complex genomic networks of taxa with porous genomes, cause phenotypic mosaics that behave dynamically [155]. Indeed, plasticity is highly known in Mentha [156,157], which confounds morphological identification. Complex morphological, chemical, and molecular diversity in mints have already been described in many studies, e.g., [126,131,158,159,160,161,162,163,164,165,166]. Despite the enormous amount of data gathered, however, there is still need of taxonomic revisions within the genus. In the recent revised phylogenetic analysis [131] the origin of M. spicata as hybrid was not supported and hidden cryptic taxa were detected in the genus.
(ii)
Selection of chemical markers. Carvone, a characteristic compound produced by M. spicata is also produced by different species from different plant families [144]. Chemical markers such as carvone in spearmint, and menthofuran and menthol in peppermint are used in practice for authentication of oils regardless of their sources [49].
Therefore, there is need to design a combination of approaches in case of mint, where the misidentification or presence of hidden cryptic species hybridisation makes DNA methods difficult, and production of characteristic compounds in other species makes chemical analysis problematic. Furthermore, in the case of the molecular approach, attempting to use a single, universal DNA barcoding region in these cases would be unproductive, as it would ignore levels of genetic divergence associated with different reproductive strategies. It would be more productive for a DNA authentication approach to target multiple plastid DNA markers to overcome these problems. Obviously, intraspecific plastid DNA polymorphism is highly possible and maternal transmission of the chloroplast species of hybrid origin would not be differentiable from the maternal parental species. Therefore, a multi-level barcoding strategy should be used, testing for both nrITS and multiple plastid markers to increase resolution. Another important aspect, often forgotten in DNA barcoding experiments, is the number of samples analysed. It has been previously suggested that barcoding analysis should use a minimum of 10 individuals per species [167], which could overcome possible ambiguous results.

2.2. Two-Tier Trade Variety Selection for the Fragrance and Horticulture Industries: Lavandula L.—Ocimeae: Nepetioideae

EOs are used frequently in both the flavour/food and fragrance industries and the demand is steadily expanding. The market value of EOs worldwide is expected to grow from around 17 billion U.S. dollars in 2017 to about 27 billion U.S. dollars by 2022 [168]. United Kingdom export of essential oils, perfumes, cosmetics, toiletries was 5.33 billion U.S. dollars during 2020 [169]. The demand for EOs is increasing each year and is expected to grow further in the next few years. The main drivers are growing consumer awareness and a rising demand for high quality natural components in personal care products and in perfumes.
A large quantity of EOs is utilized by the fragrance or flavour industries, with only a small percentage for therapeutic purposes. In order to lower the price of the EOs, adulterants are added to the oils by some producers. It is estimated that approximately 80% of commercially available EOs are adulterated in some way [170]. Major adulterants of EOs are vegetable carrier oils, cheaper oils of the same species but of different geographical origins, EOs extracted from another part of the plant, cheaper EOs from related species, and synthetic compounds [171]. Low quality EOs are prone to produce allergic reactions, irritations, and/or toxic side effects, particularly to young and old populations who are more susceptible [172,173].
The Lavandula (Lavender) are aromatic flowering plants that include 41 species and are widely distributed across Europe, northern and eastern Africa, the Mediterranean, south-west Asia, Arabia, western Iran and India [174,175,176]. Bulgaria is the world largest producer of lavender oil nowadays. However, France and China are among those countries that have also increased their lavender production [177]. The results of phylogenetic studies [178,179] based on cpDNA trnK-matK partial sequences confirmed the monophyly of Lavandula (Figure 3) and the section classification of the genus as defined by Alan [175].
The lavender EOs are applied in a wide range of home and personal care products, perfumery, aromatherapy and alternative medicine [181]. Lavender EOs have a long history of use as fragrance and aromatherapy ingredients. The plant is used in traditional and folk medicines in different parts of the world for the treatment of several gastrointestinal, nervous and rheumatic disorders and is also used for anxiety, stress and insomnia [182,183,184]. Over the years, application of lavender extract, oil and essence in food and beverage products has also increased to a substantial level and is forecasted to grow at an increasing rate in each sector (Figure 3).
Lavender is classified into four categories: L. angustifolia Mill. (English Lavender), L. stoechas L. (French Lavender), L. latifolia Medik. (Mediterranean lavender) and L. × intermedia (lavandin, which is a cross between L. latifolia and L. angustifolia) [185]. Lavender oil, obtained from the flowers of L. angustifoliais chiefly composed of linalyl acetate (3,7-dimethyl-1,6-octadien-3yl acetate), linalool (3,7-dimethylocta-1,6-dien-3-ol), lavandulol, 1,8-cineole, lavandulyl acetate, and camphor [186,187]. English lavender oil is considered to have unique properties that are beneficial for the skin, hence it is used in various skincare products. It is a general view that English lavender is mainly grown for the perfume industry, but they are also grown as scented ornamental plants because of their aroma and attractive blue flowers. The oil from the English lavender plant attracts a high value and is often adulterated with EOs from the much cheaper sterile hybrid, lavandin (L. × intermedia) that produces more oil per plant [188]. Another factor contributing to the adulteration of English lavender with lavandin could be linked to climate change, as lavender production is affected by the weather, with an impact on availability and price [188]. The price is also influenced by the origin of cultivation of the plant, with French grown plants considered to have the oil with the best quality and, therefore, the highest prices [177]. The less valuable lavandin oil is graded accordingly to the origin of production and the hybrid used [188].
Adulteration of lavender can occur in different ways. The Lavender oils could be adulterated with similar oils from different Lavandula species or hybrids, or by the addition of synthetic components with a similar chemical composition, or with non-volatile solvents [189]. Using chemical tests, it is possible to differentiate between Lavandin and lavender oil [188].
In many cases, therefore, intentional adulteration is driven by economic reasons. On the other hand, accidental contamination may occur due to the high level of hybrids. Although lavender oils can be tested and differentiated by chemical fingerprint tests, this is not always reliable as many factors, including environment and developmental stage could alter the oil composition. Therefore, these tests could give us an indication of the oil quality but not always link this to the oil origin. Companies that want to check the quality of their starting material could benefit from DNA barcoding as a faster and more reliable way of testing the authenticity of Lavandula plants before assessing the quality of the lavender oil.
Traditionally, morphological features such as the size and shape of leaves, the presence or absence of non-glandular or glandular trichome and inflorescence were used to distinguish distant lavender species from one another [175]. A number of DNA barcoding studies have been done so far in the case of Lavandula. Hindet al. [190] tested molecular markers such as matK, rbcL, trnH-psbA and ITS to identify important lavender species. The plastid markers rbcL and trnH-psbA alone did not discriminate between L. angustifolia, L. latifolia and L. x intermedia. The ITS concatenated with rbcL, trnH-psbA and rbcL+trnH-psbA were able to discriminate the cultivated L. latifolia from L. angustifolia and L. × intermedia. The matK barcode was not amplified in this study as also reported in previous studies specifically for Lamiaceae taxa [64,191]. In another study the matK gene was successfully applied to differentiate nine Lavandula species along with high-resolution melting (HRM) analysis [192].

2.3. The Diaspora of People and Plants: Ocimum L.—Ocimeae: Nepetioideae

The tremendous increase in migrations and diasporas of human groups in the last century not only bring challenging issues for societies, but also create dramatic changes in traditional knowledge, beliefs and practices related to medicinal use of plants [193]. The discrepancy between traditional and scientific nomenclature often goes unnoticed, and these discrepancies become highly problematic for quality control and consumer protection in the importing countries.
Ocimum is one of the best-known genera of the family for its medicinal properties and economically important aromatic oils (Figure 4). This genus is monophyletic [175], highly variable and possesses a wide range of intra- and inter-specific genetic diversity, comprising more than 65 species distributed all over the world [194,195,196]. Ocimum species and varieties have unique and individual chemical compositions, but their medicinal properties have not been fully explored.Moreover, due to extensive and nonregulated collections, many species have become threatened or endangered [197,198].
Among these, O. tenuiflorum L. (Holy basil or Tulsi), is an important medicinal plant, with religious significance to the Hindu community throughout the world and worshipped for over more than 3000 years due to its healing properties [200,201,202,203]. Tulsi plants are characterised by having a complex chemical composition, containing many biologically active phytochemicals with variable proportions among varieties. The EOs of tulsi contain phenylpropanoids such as eugenol, methyl eugenol, chavicol and estragole (methyl chavicol) [204]. Two chemotypes of O. tenuiflorum are known as ‘Ram’ (white) and ‘Shyam or Krishna’ (black) have been identified based on high or low methyl eugenol:eugenol ratios [200]. As methyl eugenol and methyl chavicol are classed as genotoxic carcinogens, it is important to ensure that the levels of these compounds in herbal products fall below the regulatory thresholds. The genus is known to possess antibacterial, antianaphylactic, antihistaminic, wound healing, antidiabetic, larvicidal, anti-genotoxic, neuro-protective, cardio-protective, hepato-protective, anti-carcinogenic and mast cell stabilization activity [205,206]. O. basilicum L. (Sweet Basil), O. gratissimum L. (African basil or Vana tulsi as some authors claim) and O. tenuiflorum, are frequently cultivated in several countries of East Asia, Europe, America, and Australia for the production of EOs [207,208,209].
An important aspect of globalization of plants is the migration of seeds/plants, and of the traditional knowledge of indigenous medicinal plants along with the migration of people. Tulsi seeds/plants were brought to UK from Africa and India. It was later revealed in a DNA barcoding study [117] that during this migration “Rama tulsi” used by south Asian communities in UK had been substituted with African O. gratissimum. Out of four barcoding markers (matK, rbcL, trnL-F and trnH-psbA) tested by Jurges et al. [117], trnH-psbA was identified as the best marker for commercial application to discriminate different types of Tulsi—“Rama Tulsi” and “Krishna Tulsi” of O. tenuiflorum and “Vana Tulsi” of O. gratissimum. These plastid markers also clarified the phylogenetic relationships mirrored in the chemical differences within the Ocimum [117]. Rama and Krishna appeared within the main clade of O. tenuiflorum and Vana within a different clade as observed in previous studies [210,211]. The trnH-psbA region was introduced as the most suitable candidate barcode into the British Pharmacopoeia [212] to authenticate O. tenuiflorum in industrial quality assurance procedures.
Another approach was adopted by Ríos-Rodríguez et al. [213], who designed a trait-related DNA barcode based on the enzyme eugenol O-methyltransferase (EOMT), responsible for the synthesis of methyl eugenol. The study revealed that a multiplex PCR coupled with trait-related and trait-independent markers can differentiate O. tenuiflorum from other Ocimum species and identify methyl eugenol chemotypes of O. tenuiflorum, even in dried material sold as mixtures, confirming the results of Mali [200]. The high degree of intra- and inter- specific genetic diversity in the genus determines a large number of subspecies, different varieties and forms which produce EOs with varying chemical composition [214]. Some of the Ocimum species are highly similar in apparent vegetative morphology and are hence misidentified. Moreover, the cultural and commercial values associated with the Tulsi plant have also increased the risk of adulteration [215]. Different species are sold mostly as dried powders under the same name, and therefore there exists a great need to develop an accurate method that can prove the authenticity of plant raw material. The existing methods to ensure correct plant species collection and cultivation include organoleptic traits and phytochemical methods [216,217,218], but none of these methods sufficient to guarantee the authenticity of the plant [219].

2.4. Demands of High-Quality Herbal Products in the Food Market: Origanum L.—Mentheae: Nepetioideae

The demand for spices and herbs is increasing globally, and this trend is anticipated to continue in the coming years (Figure 5). The expected growth is forecast to be driven by increasing interest in international ethnic cuisines combined with heathy eating trends. Due to increased awareness and demand, food safety issues such as adulteration of herbs and spices has been recorded frequently, as mentioned previously. Some of the most widely used culinary herbs, such as basil, thyme, mint and oregano are from Lamiaceae. These herbs have been used since ancient times to improve the characteristic of food, as natural preservatives and for their nutritional properties [145,220].
The Origanum genus is comprised of up to 43 species and 16 hybrids characterized by a high morphological and chemical diversity [222,223]. They are all confined to the Mediterranean region except for O. vulgare, which has a native geographical range which extends from Macronesia throughout Europe and eastward to China [224].
The Origanum species have been used since ancient times as culinary and medicinal herbs. Medicinally, O. vulgare (oregano) has been used for thousands of years as a stimulant, carminative, expectorant, and tonic to cure asthma, cough, indigestion, rheumatism, toothache and insect bites and as preservatives in meat storage [225,226,227]. Oregano EO is composed of different compounds. The majority is thymol and carvacrol, but other compounds include p-cymene, thymoquinone, and γ-terpinene [228,229,230]
Oregano is often commercialized as a fine powder or a mixture of small fragments of dried leaves, which makes morphological recognition difficult. Several herbs including oregano leaves/oils are used both in both the food and pharmaceutical industries and the usage is anticipated to rise by a considerable rate (Figure 5). There are many species of the genus used around the world as “oregano”, but variations in their bioactive compounds have been reported in different studies [231,232,233,234]. Geographical distribution and harvest season also effect the chemical composition of the oregano plants [235].
Oregano is the name used to refer to a great variety of plants based on its particular aroma, with at least sixty-one species and seventeen genera belonging to six different families known as oregano [118]. Oregano EOs and spices are frequently adulterated with different genera/species from the same family (e.g., Saturejamontana L. and O. majorana L.) and from different families (e.g., Rubus spp., Cistus ×incanus (Rosaceae), Rhus coriaria (Anacardiaceae), Pimpinella anisum (Apiaceae), Myrtus spp. (Myrtaceae), Corylus avellana L. (Betulaceae), Olea europaea L. (Oleaceae) and Triticum aestivum L. (Poaceae) [236,237,238,239]. The quality of oregano spices is standardised by using protocols based on those specified by European Pharmacopoeia, and only these two species, O. vulgare and O. onites L., can be commercialized as true oregano [239,240]. Within the food market, criteria approved by American Trade Association and ESA for spices are limited to the phytochemical profile of EOs, weight by weight, and the acid-insoluble ash contents. These are time-consuming and not particularly discriminative in the case of oregano, where contamination may be perpetrated with misidentified or cheaper spices belonging to the same genus.
DNA barcoding approaches have been the most effective tools currently used for the authentication of herbal products, particularly when coupled with HRM analysis—a novel analytical approach. The United States Food and Drug Administration (FDA) supports the use of DNA-based technologies in quality assurance of herbal products, among other innovative analytical technologies [241]. In the case of oregano, a universal sequence of the trnL-intron barcode from different Origanum species was identified [226]. When the molecular marker was coupled with HRM analysis, it was found to be an effective method to discriminate Origanum species and genotypes in a fast and simple way [242].

2.5. Rising Demand of Natural Products in Pharma Market: Scutellaria L.—Scutellarioideae

Scutellaria is an herb, commonly known as skullcap, which contains approximately 478 species [243] and has a cosmopolitan distribution [100,244]. Several species have a long history of being used as traditional herbal medicines to treat respiratory, neurological and cardiovascular diseases, hepatic and gastric disorders [245,246,247]. The flavonoids and many other active chemicals derived from S. baicalensis (Huang Qin), S. barbata and S. lateriflora have been found to possess anticancer characteristics [50,79,248,249,250]. Due to the outstanding medicinal value, the chemical composition of the genus has attracted considerable attention in the past ten years. A wide range of chemical components have been discovered from the genus, however, the flavonoids and diterpenes are the two main groups of active constituents in this genus [251].
The main flavonoids are baicalin, baicalein, wogonoside and wogonin, which possess wide pharmacological activities and are produced in high concentration in different parts of different species (Figure 6) [252,253,254,255,256].The flavonoids in the roots of S. baicalensis were found to be high compared to the aerial parts whereas in S. lateriflora the flavonoid content of the aerial part, especially the leaf, was more than in the root [79,257].
The intentional or unintentional adulteration of S. lateriflora herbal products with hepatotoxic Teucrium spp. (Germander), T. canadense and T. chamaedrys, as well as different species from the same genus Scutellaria has been reported since the early 1990s [119,258,259,260]. The genus Teucrium also belongs to the same family Lamiaceae and has high morphological similarities with Scutellaria (Figure 7). Despite these morphological similarities, in the most recent classification of Lamiaceae based on molecular phylogeny, the genera Scutellaria and Teucrim have been placed in different subfamilies; Scrutellarioideae and Ajugoideae, respectively [261]. Phylogenetic analysis based on chloroplast genome sequences suggested that Scutellarioideae is a sister taxon to Lamioideae (Figure 6) [262].
A variety of successful analytical methods for the quality control of skullcap raw material and products were applied to measure the chemical differences between Scutellaria and Teucrium. The genus Scutellaria contained flavonoids, while the major phenolic components of the two Teucrium species (T. canadense and T. chamaedrys) were the phenylethanoids, verbascoside and teucrioside. The phenylethanoid marker was suggested to distinguish between the two genera [245,259,263,264,265,266]. However, these methods require expert analysts and are time consuming.
DNA barcoding has also been tested for authentication of the species. Three candidate DNA barcodes matK, rbcL and the psbA-trnH were sequenced and analysed by Guoet al. [267] to discriminate S. baicalensis and its adulterants (S. amoena, S. rehderiana, and S. viscidula) and this study proposed multilocus barcodes rbcL+ psbA-trnH for the detection of species authentication. We have designed HRM primers (a “two set strategy”) to target SNPs of rbcL and trnH-psbA, that are able to differentiate S. lateriflora from other species of the same genus, and from Teucrium spp. (unpublished data). Our preliminary results also confirmed that rbcL is best suited for discriminating plant taxa at the genus level, while trnH-psbA is a suitable candidate for design of species-specific barcoding tests, confirming the results of Guo et al. [267].

2.6. High Utilisation of Functional or Superfood Food and Complex Taxonomy: Salvia L.—Mentheae: Nepetioideae

The genus Salvia, with about 980 species is the largest genus in the angiosperm family Lamiaceae. It is distributed throughout the subtropical and temperate regions of the Old World and the New World [268,269,270,271,272]. Many species of the genus have been widely utilised in the pharmaceutical, food, cosmetic and horticulture industries [272,273]. The genushas health-healing properties such as antiseptic, antipyretic, analgesic, antimicrobial, antioxidant, anticancer, anticholinesterase and anti-inflammatory characteristics [274]. Different parts of the Salvia plant such as leaves, flowers, roots and seeds may be used for their health benefits and have played an important role in the treatment and recovery of individuals with COVID-19 [275].
S. miltiorrhiza (‘Danshen’ in Chinese) is used in traditional Chinese medicines to treat cardiovascular and cerebrovascular diseases and hyperlipidaemia [272,276]. S. hispanica, commonly known as “Chia”, was initially cultivated by Mesopotamian cultures as staple food and medicinal plant in pre-Columbian times [277]. It was rediscovered in the middle of the 20th century and is now available commercially worldwide as a superfood [278]. Chia seeds contain healthy omega-fatty acids and other nutritional components [272,279,280]. S. divinorum has been used in religious rites by Mazatec shamans to induce hallucinatory visions [281]. In addition, around 150 species are used in the horticulture trade, such as S. officinalis (common sage), S. elegans (pineapple sage), Salvia splendens (scarlet sage) and others (Figure 8) [282].
The genus is well-known for its unusual diverse staminal morphology, in which two fertile stamens are separated by a significantly elongated connective tissue, which form a lever mechanism important in pollination [283]. Based on floral or morphological characters different classification schemes within the genus were proposed, e.g., [284,285,286,287,288,289,290]. On the basis of molecular phylogenetic studies, traditionally defined Salvia is non-monophyletic and is classified into 11 subgenera [268,270,271,272,291]. However, to understand the inter and intra-specific relationships of the genus, it has been suggested in a recent plastomic study that using large single copy and small single copy regions with the exclusion of more rapidly evolving sites could produce the highest resolution in the phylogenetic analysis of Salvia (Figure 8) [292].
Like other species of the Lamiaceae, species of Salvia are under constant threat of economically motivated adulteration. For instance: (i) the roots of S. miltiorrhizaare adulterated with roots of S. przewalskii, S. yunnanensis, and S. trijug, (ii) sage leaves are adulterated witholive leaves, myrtle leaves, sumac, hazelnut leaves, Cistus and Phlomis, strawberry tree leaves and sandalwood [293], (iii) chia oil is expensive to produce and can therefore be easily adulterated with cheaper oils such as corn oil, peanut, soybean and sunflower [294]. Analytical techniques, such as gas chromatography mass spectroscopy (GC-MS) and FTIR, have been used to detect adulterants in Salvia species [293,295], however these techniques require expertise and can be time consuming as described earlier.
Wang et al. [273] conducted a comprehensive DNA barcoding study by using different DNA markers: rbcL, matK, trnL-F, psbA-trnH and ITS1 alone or in different combination for the identification purpose of Salvia species. In this study, ITS1 was found to be superior when compared to other markers for discriminating between species, especially S. miltiorrhiza. In a recent study, DNA barcoding was coupled with chemical analysis by LC-MS profiling and this dual approach proved to be a powerful tool in identification of taxonomically close Salvia species [296]. High-throughput sequencing of chloroplast genomes has also been successfully used for discrimination of species within the genus [275]. Multiple approaches have been tested so far for the authentication of economically important species in Salvia; however, there is still a need to develop quick and simple identification techniques. DNA barcoding can also be used to address conservation issues and germplasm preservation. Identification of plant species is a fundamental component of conservation and management planning, and the benefits of molecular identification include that it can be done any time of the year and from very small tissue samples [297,298]. In the case of Salvia, despite its importance all over the world, a significant number of the species, for example, S. pentstemonoides (Big red sage), S. taraxacifolia and S. miltiorrhiza (red sage) are listed as threatened or endangered [299,300,301]. Particular attention is needed to design conservation strategies for their protection.

3. Evolving DNA Barcoding Technologies

The conventional method of generating DNA barcodes for a species or a specimen are through PCR amplification and Sanger sequencing methods. However, Sanger sequencing technology has been found to be inadequate in some respects when compared to next-generation sequencing (NGS) technology [302,303,304]. The NGS techniques are increasingly used in many fields to obtain huge amounts of data and discover novel and essential information about the genomes. In terms of plant DNA barcoding, different approaches such as transcriptome analysis, whole chloroplast genome sequencing and mini barcoding have been developed by using NGS techniques.
Transcriptome sequence data from plants greatly increases the opportunities for identification of additional loci as DNA barcodes and measuring the phylogenetic relationships among various taxa. Rastogi et al. [196] reported the comprehensive transcriptome analysis of Ocimum species and identified transcriptome SNPs and SSR markers that could be used for the identification of closely related taxa in the genus. Likewise, SNP data was discovered from transcriptome assemblies of Lavandula clones to differentiate between L. angustifolia and its hybrid L. latifolia [305].
The strategy of using the whole chloroplast genome to identify species and reconstructing phylogenetic relationships between closely related species has also been successfully applied to Lamiaceae species. In Mentha, Ocimum, Lavandula, Origanum and Scutellaria, chloroplast genome sequencing is being carried out to understand the complex relationships between species and genera, the function of genes and the medicinal nature of the metabolites synthesized in the plant [152,196,262,306,307,308,309,310,311,312]. Access to the whole chloroplast genome will also provide more informative barcoding sites and has the potential to improve the plant identification process between closely related species. However, the genetic information in angiosperm chloroplasts is mostly inherited maternally, making the chloroplast genome a good indicator only of maternal ancestry [313]. To identify hybrids (e.g., Mentha), the use of chloroplast genome sequences alone are not sufficient and can be concatenated with markers from nuclear genomes to establish a standardised barcoding system in these species.
DNA mini-barcoding, using a smaller length of DNA, 100–250 bp in length with sufficient variable sites could be a solution to overcome the difficulties associated with traditional DNA barcoding [50,313,314,315]. Based on specially designed primers, mini-barcodes can accurately identify targeted species within a genus or family [50]. Moreover, in cases where samples contained different contaminations, identification methods combined with NGS can identify species from multiple taxa [316,317,318,319].
Species adulteration or contamination can cause severe adverse effect on human health, as reported in the cases of Origanum and Scutellaria. The quality control of the plant material is critical and its enforcement seems to be necessary for the protection of the consumer. In addition, global and competing marketplaces added to the decline of the natural habitat of traditional medicinal plants, threaten herbals with extinction. Work to understand the mechanisms of traditional medicines is therefore urgent and must be based on the ‘wild type’ material to conserve the link with thousands of years of traditional knowledge.We know that this is useful based on the number of pharmaceuticals developed from medicinal plants and we risk squandering the collective knowledge.This work is only achievable using a combination of authentication methods.

Author Contributions

Conceptualization, T.S., A.S. and C.H.; investigation, N.N., T.S., A.S. and C.H.; writing—original draft preparation, N.N., T.S., A.S. and C.H.; writing—review and editing, N.N., T.S., A.S. and C.H.; funding acquisition, N.N. and T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants Biotechnology and Biological Sciences Research Council (BBSRC) and Daphne Jackson Trust for the financial support.

Data Availability Statement

The data presented in this study are openly available in [repository name e.g., FigShare] at [doi], reference number [reference number].

Acknowledgments

We would like to thank to BBSRC and Daphne Jackson Trust for the financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaufman, J.; Kalaitzandonakes, N. The economic potential of plant-made pharmaceuticals in the manufacture of biologic pharmaceuticals. J. Commer. Biotechnol. 2011, 17, 173–182. [Google Scholar] [CrossRef]
  2. Srivastava, P.K.; Pandey, A.K. Natural products and derivatives: Biological and pharmacological activities. Biochem. Cell. Arch. 2015, 15, 1–38. [Google Scholar]
  3. Leja, K.B.; Czaczyk, K. The industrial potential of herbs and spices—A mini review. Acta Sci. Pol. Technol. Aliment. 2016, 15, 353–365. [Google Scholar] [CrossRef] [Green Version]
  4. Allkin, B. Useful Plants—Medicines: At Least 28,187 Plant Species Are Currently Recorded as Being of Medicinal Use. In State of the World’s Plants 2017; Willis, K.J., Ed.; Royal Botanic Gardens Kew: London, UK, 2017. [Google Scholar]
  5. González-Minero, F.J.; Bravo-Díaz, L. The use of plants in skin-care products, cosmetics and fragrances: Past and present. Cosmetics 2018, 5, 50. [Google Scholar] [CrossRef] [Green Version]
  6. Marranzano, M.; Rosa, L.R.; Malaguarnera, M.; Palmeri, R.; Tessitori, M.; Barbera, A.C. Polyphenols: Plant Sources and Food Industry Applications. Curr. Pharm. Des. 2018, 24, 4125–4130. [Google Scholar] [CrossRef] [PubMed]
  7. Anand, U.; Jacobo-Herrera, N.; Altemimi, A.; Lakhssassi, N. A comprehensive review on medicinal plants as antimicrobial therapeutics: Potential avenues of biocompatible drug discovery. Metabolites 2019, 9, 258. [Google Scholar] [CrossRef] [Green Version]
  8. Faccio, G. Plant complexity and cosmetic innovation. iScience 2020, 23, 101358. [Google Scholar] [CrossRef]
  9. Salmerón-Manzano, E.; Garrido-Cardenas, J.A.; Manzano-Agugliaro, F. Worldwide research trends on medicinal plants. Int. J. Environ. Res. Public Health 2020, 17, 3376. [Google Scholar] [CrossRef]
  10. Newman, D.J.; Cragg, G.M.; Snader, K.M. Natural products as sources of new drugs over the period 1981–2002. J. Nat. Prod. 2003, 66, 1022–1037. [Google Scholar] [CrossRef]
  11. Li-Weber, M. New therapeutic aspects of flavones: The anticancer properties of Scutellaria and its main active constituents Wogonin, Baicalein and Baicalin. Cancer Treat. Rev. 2009, 35, 57–68. [Google Scholar] [CrossRef]
  12. Cragg, G.M.; Newman, D.J. Natural products: A continuing source of novel drug leads. Biochim. Biophys. Acta 2013, 1830, 3670–3695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Yuan, H.; Ma, Q.; Ye, L.; Piao, G. The traditional medicine and modern medicine from natural products. Molecules 2016, 21, 559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Cowan, M.M. Plant products as antimicrobial agents. Clin. Microbiol. Rev. 1999, 12, 564–582. [Google Scholar] [CrossRef] [Green Version]
  15. Jassim, S.A.; Naji, M.A. Novel antiviral agents: A medicinal plant perspective. J. Appl. Microbiol. 2003, 95, 412–427. [Google Scholar] [CrossRef] [Green Version]
  16. Ganjhu, R.K.; Mudgal, P.P.; Maity, H.; Dowarha, D.; Devadiga, S.; Nag, S.; Arunkumar, G. Herbal plants and plant preparations as remedial approach for viral diseases. Virusdisease 2015, 26, 225–236. [Google Scholar] [CrossRef] [PubMed]
  17. Hussain, H.; Green, I.R.; Ali, I.; Khan, I.A.; Ali, Z.; Al-Sadi, A.M.; Ahmed, I. Ursolic acid derivatives for pharmaceutical use: A patent review (2012–2016). Expert Opin. Ther. Pat. 2017, 27, 1061–1072. [Google Scholar] [CrossRef]
  18. Parham, S.; Kharazi, A.Z.; Bakhsheshi-Rad, H.R.; Nur, H.; Ismail, A.F.; Sharif, S.; Rama Krishna, S.; Berto, F. Antioxidant, antimicrobial and antiviral properties of herbal materials. Antioxidants 2020, 9, 1309. [Google Scholar] [CrossRef]
  19. Brochot, A.; Guilbot, A.; Haddioui, L.; Roques, C. Antibacterial, antifungal, and antiviral effects of three essential oil blends. Microbiologyopen 2017, 6, e00459. [Google Scholar] [CrossRef]
  20. Astani, A.; Reichling, J.; Schnitzler, P. Screening for antiviral activities of isolated compounds from essential oils. Evid. Based Complement. Altern. Med. 2011, 2011, 253643. [Google Scholar] [CrossRef] [Green Version]
  21. Silva, J.K.R.; Figueiredo, P.L.B.; Byler, K.G.; Setzer, W.N. Essential oils as antiviral agents. Potential of essential oils to treat SARS-CoV-2 infection: An in-silico investigation. Int. J. Mol. Sci. 2020, 21, 3426. [Google Scholar] [CrossRef]
  22. Petrovska, B.B. Historical review of medicinal plants’ usage. Pharmacogn. Rev. 2012, 6, 1–5. [Google Scholar] [CrossRef] [Green Version]
  23. Adhikari, B.; Marasini, B.P.; Rayamajhee, B.; Bhattarai, B.R.; Lamichhane, G.; Khadayat, K.; Adhikari, A.; Khanal, S. Potential roles of medicinal plants for the treatment of viral diseases focusing on COVID-19: A review. Phytother. Res. 2020, 1, 15. [Google Scholar] [CrossRef]
  24. Benarba, B.; Pandiella, A. Medicinal plants as sources of active molecules against COVID-19. Front. Pharmacol. 2020, 11, 1189. [Google Scholar] [CrossRef] [PubMed]
  25. Joshi, J.A.; Puthiyedath, R. Outcomes of Ayurvedic care in a COVID-19 patient with hypoxia—A Case Report. J. Ayurveda Integr. Med. 2020, 13, 1–6. [Google Scholar] [CrossRef] [PubMed]
  26. Pan, X.; Dong, L.; Yang, L.; Chen, D.; Peng, C. Potential drugs for the treatment of the novel coronavirus pneumonia (COVID-19) in China. Virus Res. 2020, 286, 198057. [Google Scholar] [CrossRef] [PubMed]
  27. Silveira, D.; Prieto-Garcia, J.M.; Boylan, F.; Estrada, O.; Fonseca-Bazzo, Y.M.; Jamal, C.M.; Magalhães, P.O.; Pereira, E.O.; Tomczyk, M.; Heinrich, M. COVID-19: Is there evidence for the use of herbal medicines as adjuvant symptomatic therapy? Front. Pharmacol. 2020, 11, 581840. [Google Scholar] [CrossRef]
  28. Srivastava, A.K.; Chaurasia, J.P.; Khan, R.; Dhand, C.; Verma, S. Role of medicinal plants of traditional use in recuperating devastating COVID-19 situation. Med. Aromat. Plants 2020, 9, 359. [Google Scholar]
  29. Wen, W.; Su, W.; Tang, H.; Le, W.; Zhang, X.; Zheng, Y.; Liu, X.; Xie, L.; Li, J.; Ye, J.; et al. Immune cell profiling of COVID-19 patients in the recovery stage by single-cell sequencing. Cell Discov. 2020, 6, 31. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, D.; Zhang, B.; Lv, J.T.; Sa, R.N.; Zhang, X.M.; Lin, Z.J. The clinical benefits of Chinese patent medicines against COVID-19 based on current evidence. Pharmacol. Res. 2020, 157, 104882. [Google Scholar] [CrossRef] [PubMed]
  31. WHO Traditional Medicine Strategy: 2014–2023. Available online: https://www.who.int/publications/i/item/9789241506096 (accessed on 7 October 2021).
  32. WHO Global Report on Traditional and Complementary Medic. Available online: https://apps.who.int/iris/rest/bitstreams/1217520/retrieve (accessed on 15 March 2021).
  33. Market Research Future Report. Available online: https://www.globenewswire.com/news-release/2019/04/03/1796359/0/en/Herbal-Medicine-Market-Value-to-Surpass-USD-129-Billion-Revenue-Mark-by-2023-at-5-88-CAGR-Predicts-Market-Research-Future.html (accessed on 21 August 2021).
  34. Directive 2004/24/Ecof The European Parliament and of the Council of 31 March 2004. Official Journal of the European Union. 136/85. Available online: https://eur-lex.europa.eu/legal-content/en/ALL/?uri=CELEX%3A32004L0024 (accessed on 13 November 2021).
  35. Advice on Regulating Herbal Medicines and Practitioners. Available online: https://www.gov.uk/government/publications/advice-on-regulating-herbal-medicines-and-practitioners (accessed on 27 March 2021).
  36. Aliki, X.; Ganopoulos, I.; Kalivas, A.; Osathanunkul, M.; Chatzopoulou, P.; Athanasios, T.; Madesis, P. Multiplex HRM analysis as a tool for rapid molecular authentication of nine herbal teas. Food Control 2015, 60, 10. [Google Scholar]
  37. Gottardi, D.; Bukvicki, D.; Prasad, S.; Tyagi, A.K. Beneficial effects of spices in food preservation and safety. Front. Microbiol. 2016, 7, 1394. [Google Scholar] [CrossRef] [Green Version]
  38. Bondi, M.; Lauková, A.; de Niederhausern, S.; Messi, P.; Papadopoulou, C. Natural preservatives to improve food quality and safety. J. Food Qual. 2017, 2017, 1–3. [Google Scholar] [CrossRef] [Green Version]
  39. Srirama, R.; Kumar, J.U.S.; Seethapathy, G.S.; Newmaster, S.G.; Ragupathy, S.; Ganeshaiah, K.N.; Shaanker, R.U.; Ravikanth, G. Species adulteration in the herbal trade: Causes, consequences and mitigation. Drug Saf. 2017, 40, 1–11. [Google Scholar] [CrossRef] [PubMed]
  40. Patwardhan, B.; Vaidya, A.D.B.; Chorghade, M. Ayurveda and natural products drug discovery. Curr. Sci. 2004, 86, 789–799. [Google Scholar]
  41. Yadav, M.; Chatterji, S.; Gupta, S.K.; Watal, G. Preliminary phytochemical screening of six medicinal plants used in traditional medicine. Int. J. Pharm. Pharm. Sci. 2020, 6, 539–542. [Google Scholar]
  42. Ghosh, D. Quality issues of herbal medicines: Internal and external factors. Int. J. Complement. Altern. Med. 2018, 11, 67–69. [Google Scholar] [CrossRef] [Green Version]
  43. Zhou, X.; Li, C.G.; Chang, D.; Bensoussan, A. Current Status and Major Challenges to the Safety and Efficacy Presented by Chinese Herbal Medicine. Medicines 2019, 6, 14. [Google Scholar] [CrossRef] [Green Version]
  44. Upton, R.; David, B.; Gafner, S.; Glasl, S. Botanical ingredient identification and quality assessment: Strengths and limitations of analytical techniques. Phytochem. Rev. 2020, 19, 1157–1177. [Google Scholar] [CrossRef] [Green Version]
  45. Balekundri, A.; Mannur, V. Quality control of the traditional herbs and herbal products: A review. Future J. Pharm. Sci. 2020, 6, 67. [Google Scholar] [CrossRef]
  46. Parveen, I.; Gafner, S.; Techen, N.; Murch, S.J.; Khan, I.A. DNA barcoding for the identification of botanicals in herbal medicine and dietary supplements: Strengths and limitations. Planta Med. 2016, 82, 1225–1235. [Google Scholar] [CrossRef] [Green Version]
  47. Raclariu, A.C.; Heinrich, M.; Ichim, M.C.; de Boer, H. Benefits and limitations of DNA barcoding and metabarcoding in herbal product authentication. Phytochem.Anal. 2018, 29, 123–128. [Google Scholar] [CrossRef]
  48. de Boer, H.J.; Ichim, M.C.; Newmaster, S.G. DNA barcoding and pharmacovigilance of herbal medicines. Drug Saf. 2015, 38, 611–620. [Google Scholar] [CrossRef]
  49. Adulteration of Essential Oils and Detection Techniques. Available online: https://www.linkedin.com/pulse/adulteration-essential-oils-detection-techniques-dr-sudhi-mestri (accessed on 10 May 2021).
  50. Gao, J.; Yin, W.; Corcoran, O. From Scutellariabarbata to BZL101 in cancer patients: Phytochemistry, Pharmacology, and Clinical Evidence. Nat. Prod. Commun. 2019, 14, 1–12. [Google Scholar]
  51. Schlick-Steiner, B.C.; Steiner, F.M.; Seifert, B.; Stauffer, C.; Christian, E.; Crozier, R.H. Integrative taxonomy: A multisource approach to exploring biodiversity. Annu. Rev. Entomol. 2010, 55, 421–438. [Google Scholar] [CrossRef]
  52. Pawar, R.S.; Handy, S.M.; Cheng, R.; Shyong, N.; Grundel, E. Assessment of the authenticity of herbal dietary supplements: Comparison of chemical and DNA barcoding methods. Planta Med. 2017, 83. [Google Scholar] [CrossRef] [Green Version]
  53. Kazi, T.; Hussain, N.; Bremner, P.; Slater, A.; Howard, C. The application of a DNAbased identification technique to over-the-counter herbal medicines. Fitoterapia 2013, 87, 27–30. [Google Scholar] [CrossRef]
  54. Mishra, P.; Kumar, A.; Nagireddy, A.; Mani, D.N.; Shukla, A.K.; Tiwari, R.; Sundaresan, V. DNA barcoding: An efficient tool to overcome authentication challenges in the herbal market. PlantBiotechnol. J. 2016, 14, 8–21. [Google Scholar] [CrossRef] [PubMed]
  55. Grant, D.M.; Brodnicke, O.B.; Evankow, A.M.; Ferreira, A.O.; Fontes, J.T.; Hansen, A.K.; Jensen, M.R.; Kalaycı, T.E.; Leeper, A.; Patil, S.K.; et al. The future of DNA barcoding: Reflections from early career researchers. Diversity 2021, 13, 313. [Google Scholar] [CrossRef]
  56. Newmaster, S.G.; Grguric, M.; Shanmughanandhan, D.; Ramalingam, S.; Ragupathy, S. DNA barcoding detects contamination and substitution in North American herbal products. BMC Med. 2013, 11, 222. [Google Scholar] [CrossRef] [Green Version]
  57. Ichim, M.C. The DNA-based authentication of commercial herbal products reveals their globally widespread adulteration. Front. Pharmacol. 2019, 10, 1227. [Google Scholar] [CrossRef] [PubMed]
  58. Sgamma, T.; Lockie-Williams, C.; Kreuzer, M.; Williams, S.; Scheyhing, U.; Koch, E.; Slater, A.; Howard, C. DNA barcoding for industrial quality assurance. Planta Med. 2017, 83, 1117–1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Nehal, N.; Choudhary, B.; Nagpure, A.; Gupta, R.K. DNA barcoding: A modern age tool for detection of adulteration in food. Crit. Rev. Biotechnol. 2021, 41, 767–791. [Google Scholar] [CrossRef] [PubMed]
  60. Hebert, P.D.N.; Cywinska, A.; Ball, S.L.; DeWaard, J.R. Biological identifications through DNA barcodes. Proc. R. Soc. Lond. B 2003, 270, 313–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Hollingsworth, P.M. Refining the DNA barcode for land plants. Proc. Natl. Acad. Sci. USA 2011, 108, 9451–9452. [Google Scholar] [CrossRef] [Green Version]
  62. Kress, W.J.; Wurdack, K.J.; Zimmer, E.A.; Weigt, L.A.; Janzen, D.H. Use of DNA barcodes to identify flowering plants. Proc. Natl. Acad. Sci. USA 2005, 102, 8369–8374. [Google Scholar] [CrossRef] [Green Version]
  63. Chase, M.W.; Salamin, N.; Wilkinson, M.; Dunwell, J.M.; Kesanakurthi, R.P.; Haider, N.; Savolainen, V. Land plants and DNA barcodes: Short-term and long-term goals. Philos. Trans. 2005, 360, 1889–1895. [Google Scholar] [CrossRef] [Green Version]
  64. Kress, W.J.; Erickson, D.L. DNA barcodes: Genes, genomics, and bioinformatics. Proc. Natl. Acad. Sci. USA 2007, 105, 2761–2762. [Google Scholar] [CrossRef] [Green Version]
  65. Lahaye, R.; Savolainen, V.; Duthoit, S.; Maurin, O.; van der Bank, M. A test of psbK-psbI and atpF-atpH as potential plant DNA barcodes using the flora of the Kruger National Park (South Africa) as a model system. Nat. Preced. 2008, 1, 1–21. [Google Scholar] [CrossRef]
  66. Newmaster, S.G.; Fazekas, A.J.; Steeves, R.A.D.; Janovec, J. Testing candidate plant barcode regions in the Myristicaceae. Mol. Ecol.Resour. 2008, 8, 480–490. [Google Scholar] [CrossRef]
  67. CBOL Plant Working Group. A DNA barcode for land plants. Proc. Natl. Acad. Sci. USA 2009, 106, 12794–12797. [Google Scholar] [CrossRef] [Green Version]
  68. Chen, S.; Flower, A.; Ritchie, A.; Liuc, J.; Molassiotis, A.; Yue, H.; Lewithe, G. Oral Chinese herbal medicine (CHM) as an adjuvant treatment during chemotherapy for non-small cell lung cancer: A systematic review. Lung Cancer 2010, 68, 137–145. [Google Scholar] [CrossRef]
  69. China Plant BOL Group; Li, D.Z.; Gao, L.M.; Li, H.T.; Wang, H.; Ge, X.J.; Liu, J.Q.; Chen, Z.D.; Zhou, S.L.; Chen, S.L.; et al. Comparative analysis of a large dataset indicates that internal transcribed spacer (ITS) should be incorporated into the core barcode for seed plants. Proc. Natl. Acad. Sci. USA 2011, 108, 19641–19646. [Google Scholar] [PubMed] [Green Version]
  70. Hollingsworth, P.M.; Graham, S.W.; Little, D.P. Choosing and using a plant DNA Barcode. PLoS ONE 2011, 6, e19254. [Google Scholar] [CrossRef]
  71. Li, X.; Yang, Y.; Henry, R.J.; Rossetto, M.; Wang, Y.; Chen, S. Plant DNA barcoding: From gene to genome. Biol. Rev. 2015, 90, 157–166. [Google Scholar] [CrossRef] [PubMed]
  72. Mao, Y.R.; Zhang, Y.H.; Nakamura, K.; Guan, B.C.; Qiu, Y.X. Developing DNA barcodes for species identification in Podophylloideae (Berberidaceae). J. Syst. Evol. 2014, 52, 487–499. [Google Scholar] [CrossRef]
  73. Buddhachat, K.; Osathanunkul, M.; Madesis, P.; Chomdej, S.; Ongchai, S. Authenticity analyses of Phyllanthus amarus using barcoding coupled with HRM analysis to control its quality for medicinal plant product. Gene 2015, 573, 84–90. [Google Scholar] [CrossRef]
  74. Feng, S.; Jiang, M.; Shi, Y.; Jiao, K.; Shen, C.; Lu, J.; Ying, Q. Application of the ribosomal DNA ITS2 region of Physalis (Solanaceae): DNA barcoding and phylogenetic study. Front. Plant Sci. 2016, 7, 1047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Umdale, S.D.; Kshirsagar, P.R.; Lekhak, M.M.; Gaikwaid, N.B. Molecular authentication of the traditional medicinal plant “Lakshman Booti” (Smithiaconferta Sm.) and its adulterants through DNA barcoding. Pharmacogn. Mag. 2017, 13, S224–S229. [Google Scholar]
  76. Yu, J.; Wu, X.; Liu, C.; Newmaster, S.; Ragupathy, S.; Kress, W.J. Progress in the use of DNA barcodes in the identification and classification of medicinal plants. Ecotoxicol. Environ. Saf. 2021, 208, 111691. [Google Scholar] [CrossRef]
  77. Vivas, C.V.; Moraes, R.C.S.; Alves-Araújo, A.; Alves, M.; Mariano-Neto, E.; Berg, C.V.; Gaiotto, F.A. DNA barcoding in Atlantic Forest plants: What is the best marker for Sapotaceae species identification? Genet. Mol. Biol. 2014, 37, 662–670. [Google Scholar] [CrossRef]
  78. Bolson, M.; de Camargo Smidt, E.; Brotto, M.L.; Silva-Pereira, V. ITS and trnH-psbA as efficient DNA Barcodes to identify threatened commercial woody angiosperms from southern Brazilian Atlantic rainforests. PLoS ONE 2015, 10, e0143049. [Google Scholar]
  79. Duan, H.; Wang, W.; Zeng, Y.; Guo, M.; Zhou, Y. The screening and identification of DNA barcode sequences for Rehmannia. Sci. Rep. 2019, 9, 17295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Li, Q.J.; Wang, X.; Wang, J.; Su, N.; Zhang, L.; Ma, Y.; Chang, Z.; Zhao, L.; Potter, D. Efficient identification of Pulsatilla (Ranunculaceae) using DNA barcodes and micro-morphological characters. Front. Plant Sci. 2019, 10, 1196. [Google Scholar] [CrossRef] [PubMed]
  81. Potapova, T.A.; Gerton, J.L. Ribosomal DNA and the nucleolus in the context of genome organization. Chromosome Res. 2019, 27, 109–127. [Google Scholar] [CrossRef] [PubMed]
  82. Banchi, E.; Ametrano, C.G.; Greco, S.; Stanković, D.; Muggia, L.; Pallavicini, A. PLANiTS: A curated sequence reference dataset for plant ITS DNA metabarcoding. Database 2020, 155, 1–9. [Google Scholar] [CrossRef] [Green Version]
  83. Besse, P.; Silva, D.D.; Michel Grisoni, M. Plant DNA barcoding principles and limits: A case study in the genus Vanilla. Methods Mol. Biol. 2020, 2222, 131–148. [Google Scholar]
  84. Masiero, E.; Banik, D.; Abson, J.; Greene, P.; Slater, A.; Sgamma, T. Molecular verification of the UK national collection of cultivated Liriope and Ophiopogon plants. Plants 2020, 9, 558. [Google Scholar] [CrossRef]
  85. Jobes, D.V.; Thien, L.B. A conserved motif in the 5.8S ribosomal RNA (rRNA) gene is a useful diagnostic marker for plant internal transcribed spacer (ITS) sequences. PlantMol. Biol. Rep. 1997, 15, 326–334. [Google Scholar] [CrossRef]
  86. White, T.J.; Bruns, T.D.; Lee, S.B.; Taylor, J.W. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In PCR Protocols; Innis, M.A., Gelfand, D.H., Sninsky, J.J., White, T.J., Eds.; Academic Press: Cambridge, MA, USA, 1990. [Google Scholar]
  87. Manter, D.K.; Vivanco, J.M. Use of the ITS primers, ITSF and ITS4, to characterize fungal abundance and diversity in mixed-template samples by qPCR and length heterogeneity analysis. J. Microbiol. Methods 2007, 7, 7–14. [Google Scholar] [CrossRef]
  88. Cheng, T.; Xu, C.; Lei, L.; Li, C.; Zhang, Y.; Zhou, S. Barcoding the kingdom Plantae: New PCR primers for ITS regions of plants with improved universality and specificity. Mol. Ecol. Resour. 2016, 6, 38–49. [Google Scholar] [CrossRef]
  89. Badotti, F.; de Oliveira, F.S.; Garcia, C.F.; Vaz, A.B.M.; Fonseca, P.L.C.; Nahum, L.A.; Oliveira, G.; Góes-Neto, A. Effectiveness of ITS and sub-regions as DNA barcode markers for the identification of Basidiomycota (Fungi). BMC Microbiol. 2017, 7, 42. [Google Scholar] [CrossRef] [Green Version]
  90. Zheng, R.; Wang, W.; Tan, J.; Xu, H.; Zhan, R.; Chen, W. An investigation of fungal contamination on the surface of medicinal herbs in China. Chin.Med. 2017, 12, 1–8. [Google Scholar] [CrossRef] [Green Version]
  91. Soltis, P.S.; Daniel, L.; Nickrent, D.L.; Leigh, A.; Johnson, L.A.; Hahn, W.J.; Hoot, S.B.; Sweere, J.A.; Kuzoff, R.K.; Kron, K.A.; et al. Angiosperm phylogeny inferred from 8S ribosomal DNA sequences. Ann. Mo. Bot. Gard. 1997, 84, 1–49. [Google Scholar] [CrossRef]
  92. Kim, W.J.; Moon, B.C.; Yang, S.; Han, K.S.; Choi, G.; Lee, A.Y. Rapid authentication of the herbal medicine plant species Aralia continentalis Kitag. and Angelica biserrata C.Q. Yuan and R.H. Shan using ITS2 sequences and multiplex-SCAR markers. Molecules 2016, 21, 270. [Google Scholar] [CrossRef] [Green Version]
  93. Veldman, S.; Ju, Y.; Otieno, J.N.; Abihudi, S.; Posthouwer, C.; Gravendeel, B.; van Andel, T.R.; de Boer, H.J. DNA barcoding augments conventional methods for identification of medicinal plant species traded at Tanzanian markets. J. Ethnopharmacol. 2020, 250, 112495. [Google Scholar] [CrossRef]
  94. Rai, P.D.; Bellampalli, R.; Dobriyal, R.M.; Agarwal, A.; Satyamoorthy, K.; Anantha Narayana, D.A. DNA barcoding of authentic and substitute samples of herb of the family Asparagaceae and Asclepiadaceae based on the ITS2 region. J. Ayurveda Integr. Med. 2012, 3, 36–40. [Google Scholar]
  95. Zhao, S.; Chen, X.; Song, J.; Pang, X.; Chen, S. Internal transcribed spacer 2 barcode: A good tool for identifying Acanthopanacis cortex. Front. Plant Sci. 2015, 6, 840. [Google Scholar] [CrossRef] [Green Version]
  96. Awad, M.; Fahmy, R.M.; Mosa, K.A.; Helmy, M.; El-Feky, F.A. Identification of effective DNA barcodes for Triticum plants through chloroplast genomewide analysis. Comput. Biol. Chem. 2017, 71, 20–31. [Google Scholar] [CrossRef] [PubMed]
  97. Fadzil, N.F.; Wagiran, A.; Salleh, F.M.; Abdullah, S.; Izham, N.H.M. Authenticity testing and detection of Eurycoma longifolia in commercial herbal products using bar-high resolution melting analysis. Genes 2018, 9, 408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Bailey, C.D.; Carr, T.G.; Harris, S.A.; Hughes, C.E. Characterization of angiosperm nrDNA polymorphism, paralogy, and pseudogenes. Mol. Phylogenet. Evol. 2003, 29, 435–455. [Google Scholar] [CrossRef] [PubMed]
  99. Howard, C.; Lockie-Williams, C.; Slater, A. Applied barcoding: The practicalities of DNA testing for herbals. Plants 2020, 9, 1150. [Google Scholar] [CrossRef]
  100. Harley, R.M.; Atkins, S.; Budantsev, A.L.; Cantino, P.D.; Conn, B.J.; Grayer, R.; Harley, M.M.; De Kok, R.; Krestovskaja, T.; Morales, R.; et al. Labiatae. In The Families and Genera of Vascular Plants; Kubitzki, K., Kadereit, J.W., Eds.; Springer: Berlin, Germany, 2004; Volume 7, pp. 167–275. [Google Scholar]
  101. Carović-Stanko, K.; Petek, M.; Grdiša, M.; Pintar, J.; Bedeković, D.; Ćustić, M.H.; Satovic, Z. Medicinal plants of the family Lamiaceae as functional foods—A review. Czech J. Food Sci. 2016, 34, 377–390. [Google Scholar] [CrossRef] [Green Version]
  102. Mamadalieva, N.Z.; Akramov, D.K.; Ovidi, E.; Tiezzi, A.; Nahar, L.; Azimova, S.S.; Sarker, S.D. Aromatic medicinal plants of the Lamiaceae family from Uzbekistan: Ethnopharmacology, essential oils composition, and biological activities. Medicines 2017, 4, 8. [Google Scholar] [CrossRef] [Green Version]
  103. Lawrence, B.M. Basil oil—Perfume. Flavor 1992, 17, 47–50. [Google Scholar]
  104. Ascensão, L.; Marques, N.; Pais, M.S. Glandular trichomes on vegetative and reproductive organs of Leonotisleonurus (Lamiaceae). Ann. Bot. 1995, 75, 619–626. [Google Scholar] [CrossRef]
  105. Giuliani, C.; Maleci-Bini, L. Insight into the structure and chemistry of glandular trichomes of Labiatae, with emphasis on subfamily Lamioideae. Plant Syst. Evol. 2008, 276, 199–208. [Google Scholar] [CrossRef]
  106. Puškárová, A.; Bučková, M.; Kraková, L.; Pangallo, D.; Kozics, K. The antibacterial and antifungal activity of six essential oils and their cyto/genotoxicity to human HEL 12469 cells. Sci. Rep. 2017, 7, 8211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Bozin, B.; Mimica-Dukic, N.; Simin, N.; Anackov, G. Characterization of the volatile composition of essential oils of some Lamiaceae spices and the antimicrobial and antioxidant activities of the entire oils. J. Agric. Food Chem. 2006, 54, 1822–1828. [Google Scholar] [CrossRef] [PubMed]
  108. Birkett, M.A.; Hassanali, A.; Hoglund, S.; Pettersson, J.; Pickett, J.A. Repellent activity of catmint, Nepeta cataria, and iridoid nepetalactone isomers against Afro-tropical mosquitoes, ixodid ticks and red poultry mites. Phytochemistry 2011, 72, 109–114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Chadwick, M.; Trewin, H.; Gawthrop, F.; Wagstaff, C. Sesquiterpenoids lactones: Benefits to plants and people. Int. J. Mol. Sci. 2013, 14, 12780–12805. [Google Scholar] [CrossRef] [Green Version]
  110. Chen, Y.P.; Li, B.; Olmstead, R.G.; Cantino, P.D.; Liu, E.D.; Xiang, C.L. Phylogenetic placement of the enigmatic genus Holocheila (Lamiaceae) inferred from plastid DNA sequences. Taxon 2014, 63, 355–366. [Google Scholar] [CrossRef]
  111. Aburjai, T.; Natsheh, F.M. Plants used in cosmetics. Phytother. Res. 2003, 17, 987–1000. [Google Scholar] [CrossRef]
  112. Grassmann, J.; Elstner, E.F. Essential oils. Properties and uses. In Encyclopedia of Food Science and Nutrition, 2nd ed.; Caballero, B., Trugo, L., Finglas, P., Eds.; Elsevier: New York, NY, USA, 2003; pp. 2177–2184. [Google Scholar]
  113. Burt, S. Essential oils: Their antibacterial properties and potential applications in foods—A review. Int. J. Food Microbiol. 2004, 94, 223–253. [Google Scholar] [CrossRef] [PubMed]
  114. Edris, A.E. Pharmaceutical and therapeutic potentials of essential oils and their individual volatile constituents: A review. Phytother. Res. 2007, 21, 308–323. [Google Scholar] [CrossRef]
  115. Ozkan, M. Glandular and eglandular hairs of Salvia recognita Fisch. & Mey. (Lamiaceae) in Turkey. Bangladesh J. Bot. 2008, 37, 93–95. [Google Scholar]
  116. Rattray, R.D.; Van Wyk, B.E. The Botanical, Chemical and Ethnobotanical Diversity of Southern African Lamiaceae. Molecules 2021, 26, 3712. [Google Scholar] [CrossRef] [PubMed]
  117. Jurges, G.; Sahi, V.; Rodriguez, D.R.; Reich, E.; Bhamra, S.; Howard, C.; Slater, A.; Nick, P. Product authenticity versus globalisation—The Tulsi case. PLoS ONE 2018, 13, e0207763. [Google Scholar] [CrossRef]
  118. Leyva-López, N.; Gutiérrez-Grijalva, E.P.; Vazquez-Olivo, G.; Heredia, J.B. Essential Oils of Oregano: Biological Activity beyond Their Antimicrobial Properties. Molecules 2017, 22, 989. [Google Scholar] [CrossRef] [Green Version]
  119. Foster, S. Adulteration of Skullcap with American Germander; American Botanical Council: Austin, TX, USA, 2016. [Google Scholar]
  120. World Checklist of Lamiaceae (Mentha). Available online: http://wcsp.science.kew.org/namedetail.do?name_id=124319 (accessed on 5 May 2020).
  121. Savithri, B.; Priti, M.; Sushil, K.; Anil, K. Mentha species: In vitro regeneration and genetic Transformation. Mol. Biol. Today 2002, 3, 11–23. [Google Scholar]
  122. Pandey, A.K.; Rai, M.K.; Acharya, D. Chemical composition and antimycotic activity of the essential oils of corn mint (Mentha arvensis) and lemon grass (Cymbopogon flexuosus) against human pathogenic fungi. Pharm. Biol. 2003, 41, 421–425. [Google Scholar] [CrossRef]
  123. Gulluce, M.; Sahin, F.; Sokmen, M. Antimicrobial and antioxidant properties of the essential oils and methanol extract from Mentha longifolia L. ssp. longifolia. Food Chem. 2007, 103, 1449–1456. [Google Scholar] [CrossRef]
  124. Nickavar, B.; Alinaghi, A.; Kamalinejad, M. Evaluation of the Antioxidant Properties of Five Mentha Species. Iran. J. Pharm. Sci. 2008, 7, 203–209. [Google Scholar]
  125. Morton, J.K. The chromosome numbers of the British Menthae. Warsonia 1956, 3, 244–252. [Google Scholar]
  126. Harley, R.M.; Brighton, C.A. Chromosome numbers in the genus Mentha, L. Bot. J. Linn. Soc. 1977, 74, 71–96. [Google Scholar] [CrossRef]
  127. Gobert, V.; Moja, S.; Colson, M.; Taberlet, P. Hybridization in the section Mentha (Lamiaceae) inferred from AFLP markers. Am. J. Bot. 2002, 89, 2017–2023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Tucker, A.; Chambers, H.L. Mentha canadensis L. (Lamiaceae): A relict amphidiploid from the lower tertiary. Taxon 2002, 51, 703. [Google Scholar] [CrossRef]
  129. Testing Hypotheses of Hybridization in Mentha Spicata and M. Canadensis Using Molecular Data. Available online: https://digitalcommons.murraystate.edu/postersatthecapitol/2005/WKU/5/ (accessed on 2 November 2021).
  130. Mint Essential Oil Market Size Worth $330.02 Million by 2025. CAGR: 9.2%: Grand View Research, Inc. Available online: https://www.prnewswire.com/news-releases/mint-essential-oil-market-size-worth-330-02-million-by-2025--cagr-9-2-grand-view-research-inc-300971922.html (accessed on 26 September 2021).
  131. Heylen, O.C.G.; Debortoli, N.; Marescaux, J.; Olofsson, J.K. A Revised Phylogeny of the Mentha spicata Clade Reveals Cryptic Species. Plants 2021, 10, 819. [Google Scholar] [CrossRef]
  132. Snoussi, M.; Noumi, E.; Trabelsi, N.; Flamini, G.; Papetti, A.; De Feo, V. Menthaspicata Essential Oil: Chemical Composition, Antioxidant and Antibacterial Activities against Planktonic and Biofilm Cultures of Vibrio spp. Strains. Molecules 2015, 20, 14402–14424. [Google Scholar] [CrossRef] [PubMed]
  133. Wilson, L. Spices and Flavoring Crops: Leaf and Floral Structures. In Encyclopedia of Food and Health, 1st ed.; Caballero, B., Fingal, P., Toldra, F., Eds.; Academic Press: Cambridge, MA, USA, 2016; pp. 84–92. [Google Scholar]
  134. Khanuja, S.P.S.; Shasany, A.K.; Srivastava, A.; Kumar, S. Assessment of genetic relationships in Mentha species. Euphytica 2000, 111, 121–125. [Google Scholar] [CrossRef]
  135. Shaikh, S.; Yaacob, H.B.; Rahim, Z.H. A Prospective role in treatment of major illnesses and potential benefits as a safe insecticide and natural food preservative of mint (Mentha spp.): A Review. Asian J. Biomed. Pharm. Sci. 2014, 4, 1–12. [Google Scholar] [CrossRef]
  136. Kumar, K.V.; Patra, D. Alteration in yield and chemical composition of essential oil of Mentha piperita L. plant: Effect of fly ash amendments and organic wastes. Ecol. Eng. 2012, 47, 237–241. [Google Scholar] [CrossRef]
  137. Prakash, O.; Chandra, M.; Pant, A.K.; Rawat, D.S. Mint (Mentha spicata L.) Oils. In Essential Oils in Food Preservation, Flavor and Safety; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2016; pp. 561–572. [Google Scholar]
  138. Salehi, B.; Stojanović-Radić, Z.; Matejić, J.; Sharopov, F.; Antolak, H.; Kręgiel, D.; Sen, S.; Sharifi-Rad, M.; Acharya, K.; Sharifi-Rad, R.; et al. Plants of Genus Mentha: From Farm to Food Factory. Plants 2018, 4, 70. [Google Scholar] [CrossRef] [Green Version]
  139. Pandey, J.; Verma, R.; Singh, S. Suitability of aromatic plants for phytoremediation of heavy metal contaminated areas: A review. Int. J. Phytoremediat. 2019, 21, 1–14. [Google Scholar] [CrossRef]
  140. Schmidt, E.; Bail, S.; Buchbauer, G.; Stoilova, I.; Atanasova, T.; Stoyanova, A.; Krastanov, A.; Jirovetz, L. Chemical composition, olfactory evaluation and antioxidant effects of essential oil from Mentha × piperita. Nat. Prod. Commun. 2009, 4, 1107–1112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Afridi, M.S.; Ali, J.; Abbas, S.; Rehman, S.U.; Khan, F.A.; Khan, M.A.; Shahid, M. Essential oil composition of Mentha piperita L. And its antimicrobial effects against common human pathogenic bacterial and fungal strains. Pharmacologyonline 2016, 3, 90–97. [Google Scholar]
  142. Bhattacharya, S. Cultivation of Essential Oils. In Essential Oils in Food Preservation, Flavor and Safety; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2016. [Google Scholar]
  143. Dionísio, A.; Molina, G.; Carvalho, D.; Santos, R.; Bicas, J.; Pastore, G. Natural flavourings from biotechnology for foods and beverages. In Natural Food Additives, Ingredients and Flavourings; Baines, D., Seal, R., Eds.; Woodhead: Sawston, UK, 2012. [Google Scholar]
  144. Morcia, C.; Tumino, G.; Ghizzoni, R.; Terzi, V. Carvone (Mentha spicata L.) oils. In Essential Oils in Food Preservation, Flavor and Safety; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2020. [Google Scholar]
  145. Kokkini, S.; Karousou, R.; Hanlidou, E. Herbs of the Labiatae. In Encyclopedia of Food Sciences and Nutrition; Caballero, B., Ed.; Academic Press: Oxford, UK, 2003; pp. 3082–3090. [Google Scholar]
  146. Taylan, O.; Cebi, N.; Sagdic, O. Rapid screening of Mentha spicata essential oil and l-menthol in Mentha piperita essential oil by ATR-FTIR spectroscopy coupled with multivariate analyses. Foods 2021, 10, 202. [Google Scholar] [CrossRef] [PubMed]
  147. Mattia, F.D.; Bruni, I.; Galimberti, A.; Cattaneo, F.; Casiraghi, M.; Labra, M. A comparative study of different DNA barcoding markers for the identification of some members of Lamiaceae. Food Res. Int. 2011, 44, 693–702. [Google Scholar] [CrossRef]
  148. Attiya, J.; Bin, G.; Bilal, H.A.; Zabta, K.S.; Tariq, M. Phylogenetics of selected Mentha species on the basis of rps8, rps11 and rps14 chloroplast genes. J. Med. PlantRes. 2012, 6, 30–36. [Google Scholar]
  149. Thakur, V.V.; Tiwari, S.; Tripathi, N.; Tiwari, G.; Sapre, S. DNA barcoding and phylogenetic analyses of Mentha species using rbcL sequences. Ann. Phytomed. 2016, 5, 59–62. [Google Scholar]
  150. Ahmed, S.M. Molecular identification of Lavendula dentata L.; Mentha longifolia (L.) Huds. and Mentha × piperita L. by DNA barcodes. Bangladesh J. Plant Taxon 2018, 25, 149–157. [Google Scholar] [CrossRef]
  151. Wang, K.; Li, L.; Hua, Y.; Zhao, M.; Li, S.; Sun, H.; Lv, Y.; Wang, Y. The complete chloroplast genome of Mentha spicata, an endangered species native to South Europe. Mitochondrial DNA B 2017, 2, 907–909. [Google Scholar] [CrossRef] [Green Version]
  152. Vining, K.J.; Johnson, S.R.; Ahkami, A.; Lange, I.; Parrish, A.N.; Trapp, S.C.; Croteau, R.B.; Straub, S.C.K.; Pandelova, I.; Lange, B.M. Draft genome sequence of Mentha longifolia and development of resources for mint cultivar improvement. Mol. Plant 2017, 10, 323–339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Saric-Kundalic, B.; Fialova, S.; Dobes, C.; Ölzant, S.; Tekelova, A.D.; Grancal, D.; Reznicek, G.; Saukel, J. Multivariate numerical taxonomy of Mentha species, hybrids, varieties and cultivars. Sci. Pharm. 2009, 77, 851–876. [Google Scholar] [CrossRef]
  154. Mogosan, C.; Vostinaru, O.; Oprean, R.; Heghes, C.; Filip, L.; Balica, G.; Moldovan, R.I. A comparative analysis of the chemical composition, anti-inflammatory, and antinociceptive effects of the essential oils from three species of mentha cultivated in Romania. Molecules 2017, 22, 263. [Google Scholar] [CrossRef] [PubMed]
  155. Harder, L.D.; Strelin, M.M.; Clocher, I.C.; Kulbaba, M.W.; Aizen, M.A. The dynamic mosaic phenotypes of flowering plants. New Phytol. 2019, 224, 1021–1034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Singh, V.R.; Lal, R.K. Genotype × environment interaction, genetic variability and inheritance pattern in breeding lines including varieties/cultivars of menthol mint (Mentha arvensis L.). J. Med. Aromat. Plant Sci. 2020, 42, 145–156. [Google Scholar]
  157. Mishra, A.; Jain, P.; Lal, R.K.; Dhawan, S.S. Trichomes and Yield Traits in Mentha arvensis: Genotype Performance and Stability Evaluation. J. Herbs SpicesMed. Plants 2018, 24, 1–14. [Google Scholar] [CrossRef]
  158. Bunsawatt, J.; Elliott, N.E.; Hertweck, K.L.; Sproles, E.; Alice, L.A. Phylogenetics of Mentha (Lamiaceae): Evidence from Chloroplast DNA Sequences. Syst. Bot. 2004, 29, 959–964. [Google Scholar] [CrossRef]
  159. Shasany, A.K.; Shukla, A.K.; Soni, G.; Khanuja, S.P.S. Suman. AFLP analysis for genetic relationships among Mentha species. PlantGenet. Resour. Newsl. 2005, 144, 14–19. [Google Scholar]
  160. Celenk, S.; Dirmenci, T.; Malyer, H.; Bicakci, A. A palynological study of the genus Nepeta L. (Lamiaceae). Plant Syst. Evol. 2008, 276, 105–123. [Google Scholar] [CrossRef]
  161. Chengyuan, L.; Wei-Lin, L.; Yong, Z.; Bo Heng, X.; Xiao-Yue, W. Study on Morphological Diversity of Mentha HaplocalyxBriq. Med. Plant 2011, 2, 1–3. [Google Scholar]
  162. Hanafy, D.M.; Prenzler, P.D.; Hill, R.A.; Burrows, G.E. Leaf micromorphology of 19 Mentha Taxa. Aust. J. Bot. 2019, 67, 463. [Google Scholar] [CrossRef]
  163. Vining, K.J.; Pandelova, I.; Hummer, K.E.; Bassil, N.V.; Contreras, R.; Neill, K.; Chen, H.; Parrish, A.N.; Lange, B.M. Genetic diversity survey of Mentha aquatica L. and Mentha suaveolens Ehrh., mint crop ancestors. Genet. Resour. CropEvol. 2019, 66, 825–845. [Google Scholar] [CrossRef]
  164. Bokić, B.S.; Rat, M.M.; Kladar, N.V.; Anačkov, G.T.; Božin, B.N. Chemical diversity of volatile compounds of mints from southern part of Pannonian Plain and Balkan Peninsula—New Data. Chem.Biodivers. 2020, 17, 1–11. [Google Scholar] [CrossRef] [PubMed]
  165. Mamadalieva, N.Z.; Hidayat, H.; Jianbo, X. Recent advances in genus Mentha: Phytochemistry, antimicrobial effects, and food applications. Food Front. 2020, 1, 435–458. [Google Scholar] [CrossRef]
  166. Moshrefi-Araghi, A.; Nemati, H.; Azizi, M.; Moshtaghi, N.; Shoor, M. Study of genetic diversity of some genotypes of Iranian wild mint (Menthalongifolia L.) using ISSRmarker and its correlation with dry yield and essential oil content. Agric. Biotechnol. J. 2020, 12, 117–140. [Google Scholar]
  167. Stoeckle, M.Y.; Hebert, P.D.N. Bar Code of Life: DNA tags help classify animals. Sci.Am. 2008, 299, 82–88. [Google Scholar]
  168. Flavor and Fragrances Market Worldwide—Statistics & Facts. Available online: https://www.statista.com/topics/6300/flavor-and-fragrances-market-worldwide (accessed on 21 November 2021).
  169. United Kingdom Exports of Essential Oils, Perfumes, Cosmetics, Toiletries. Available online: https://tradingeconomics.com/united-kingdom/exports/essential-oils-perfumes-cosmetics-toileteries (accessed on 15 July 2021).
  170. Satyal, P.; Setzer, W.N. Adulteration Analysis in Essential Oils. In Essential Oil Research; Malik, S., Ed.; Springer: Cham, Switzerland, 2019. [Google Scholar]
  171. Ng, T.B.; Fang, E.F.; Bekhit, A.E.-D.A.; Wong, J.H. Methods for the Characterization, Authentication, and Adulteration of Essential Oils. In Essential Oils in Food Preservation, Flavor and Safety; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2016; pp. 11–17. [Google Scholar]
  172. Sarkic, A.; Stappen, I. Essential oils and their single compounds in cosmetics—A critical review. Cosmetics 2018, 5, 11. [Google Scholar] [CrossRef] [Green Version]
  173. Dubnicka, M.; Cromwell, B.; Levine, M. Investigation of the adulteration of essential oils by GC-MS. Curr. Anal. Chem. 2020, 16, 965–969. [Google Scholar] [CrossRef]
  174. Paton, A.J.; Springate, D.; Suddee, S.; Otieno, D.; Grayer, R.J.; Harley, M.M.; Willis, F.; Simmonds, M.S.J.; Powell, M.P.; Savolainen, V. Phylogeny and evolution of basils and allies (Ocimeae, Labiatae) based on three plastid DNA regions. Mol. Phylogenet. Evol. 2004, 31, 277–299. [Google Scholar] [CrossRef]
  175. Alan, P. The Genus Lavandula. Botanical Magazine Monograph; Kew Bulletin: Heidelberg, Germany, 2005; p. 160. [Google Scholar]
  176. World Checklist of Lamiaceae (Lavender). Available online: http://wcsp.science.kew.org/namedetail.do?name_id=108966 (accessed on 19 July 2020).
  177. Giray, F.H. An analysis of world lavender oil markets and lessons for Turkey. J. Essent. Oil Bear. Plants 2018, 21, 1612–1623. [Google Scholar] [CrossRef]
  178. Benabdelkader, T.; Guitton, Y.; Pasquier, B.; Magnard, J.L.; Jullien, F.; Kameli, A.; Legendre, L. Functional characterization of terpene synthases and chemotypic variation in three lavender species of section Stoechas. Physiol. Plant 2015, 153, 43–57. [Google Scholar] [CrossRef] [PubMed]
  179. Moja, S.; Guitton, Y.; Nicolè, F.; Legendre, L.; Pasquier, B.; Upson, T.; Jullien, F. Genome size and plastid trnK-matK markers give new insights into the evolutionary history of the genus Lavandula, L. Plant Biosyst. 2016, 150, 1216–1224. [Google Scholar] [CrossRef]
  180. Lavender Oil Market Size 2021 by Top Countries Data, Industry Analysis by Regions, Revenue, Share, Development, Tendencies and Forecast to 2026. Available online: https://www.yournewsnet.com/story/44196805/lavender-oil-market-size-2021-by-top-countries-data-industry-analysis-by-regions-revenue-share-development-tendencies-and-forecast-to-2026 (accessed on 5 October 2021).
  181. Wells, R.; Truong, F.; Adal, A.M.; Sarker, L.; Mahmoud, S. Lavandula Essential Oils: A Current Review of Applications in Medicinal, Food, and Cosmetic Industries of Lavender. Nat. Prod. Commun. 2018, 13, 1403–1417. [Google Scholar] [CrossRef] [Green Version]
  182. Gören, A.C.; Topçu, G.; Bilsel, G.; Bilsel, M.; Aydoğmuş, Z.; Pezzuto, J.M. The chemical constituents and biological activity of essential oil of Lavandula stoechas ssp. stoechas. Z. Naturforsch. C 2002, 57, 797–800. [Google Scholar] [CrossRef]
  183. Hajhashemi, V.; Ghannadi, A.; Sharif, B. Anti-inflammatory and analgesic properties of the leaf extracts and essential oil of Lavandula angustifolia Mill. J. Ethnopharmacol. 2003, 89, 67–71. [Google Scholar] [CrossRef]
  184. Koulivand, P.H.; Ghadiri, M.K.; Gorji, A. Lavender and the Nervous System. Evid. Based Complement. Altern. Med. 2013, 2013, 681304. [Google Scholar] [CrossRef] [Green Version]
  185. Pokajewicz, K.; Białoń, M.; Svydenko, L.; Fedin, R.; Hudz, N. Chemical composition of the essential oil of the new cultivars of Lavandula angustifolia Mill. Bred in Ukraine. Molecules 2021, 26, 5681. [Google Scholar] [CrossRef]
  186. Cavanagh, H.C.; Jenny, W. Biological activities of lavender essential oil. Phytother. Res. 2002, 16, 301–308. [Google Scholar] [CrossRef]
  187. Woronuk, G.; Demissie, Z.; Rheault, M.; Mahmoud, S. Biosynthesis and therapeutic properties of Lavandula essential oil constituents. Planta Med. 2011, 77, 7–15. [Google Scholar] [CrossRef] [Green Version]
  188. Bejar, E. Adulteration of English Lavender (Lavandula angustifolia) Essential Oil; Botanical Adulterants Prevention Bulletin: Austin, TX, USA, 2020. [Google Scholar]
  189. Satyal, P.; Sorensen, A.C. Authentication of lavender essential oil: Commercial essential oil samples and validity of standard specifications. Int. J. Prof. Hol. Aromather. 2016, 5, 17–22. [Google Scholar]
  190. Hind, K.R.; Adal, A.M.; Upson, T.M.; Mahmoud, S.S. An assessment of plant DNA barcodes for the identification of cultivated Lavandula (Lamiaceae) taxa. Biocatal. Agric. Biotechnol. 2018, 16, 459–466. [Google Scholar] [CrossRef]
  191. Theodoridis, S.; Stefanaki, A.; Tezcan, M.; Aki, C.; Kokkini, S.; Vlachonasios, K.E. DNA barcoding in native plants of the Labiatae (Lamiaceae) family from Chios Island (Greece) and the adjacent Çeşme-Karaburun Peninsula (Turkey). Mol. Ecol. Resour. 2012, 12, 620–633. [Google Scholar] [CrossRef] [PubMed]
  192. Soares, S.; Grazina, L.; Costa, J.; Amaral, J.S.; Oliveira, M.B.P.P.; Mafra, I. Botanical authentication of lavender (Lavandula spp.) honey by a novel DNA-barcoding approach coupled to high resolution melting analysis. Food Control 2018, 86, 367–373. [Google Scholar] [CrossRef] [Green Version]
  193. Pieroni, A.; Vandebroek, I. Travelling Cultures and Plants the Ethnobiology and Ethnopharmacy of Human Migrations; Berghahn Books Ltd.: New York, NY, USA, 2009. [Google Scholar]
  194. Sahoo, Y.; Pattnaik, S.K.; Chand, P.K. In vitro clonal propagation of an aromatic medicinal herb Ocimumbasilicum L. (sweet basil) by axillary shoot proliferation. In Vitro Cell. Dev. Biol.-Plant 1997, 33, 293–296. [Google Scholar] [CrossRef]
  195. Rastogi, S.; Meena, S.; Bhattacharya, A.; Ghosh, S.; Shukla, R.K.; Sangwan, N.S.; Lal, R.K.; Gupta, M.M.; Lavania, U.C.; Gupta, V.; et al. De novo sequencing and comparative analysis of holy and sweet basil transcriptomes. BMC Genom. 2014, 15, 588. [Google Scholar] [CrossRef] [Green Version]
  196. World Checklist of Lamiaceae (Ocimum). Available online: http://wcsp.science.kew.org/namedetail.do?name_id=136790 (accessed on 27 August 2020).
  197. Sudha, G.C.; Seeni, S. In vitro propagation and field establishment of Adhatodabeddomei C. B. Clarke, a rare medicinal plant. Plant Cell Rep. 1994, 13, 203–207. [Google Scholar]
  198. Pandey, A.K.; Singh, P.; Tripathi, N.N. Chemistry and bioactivities of essential oils of some Ocimum species: An overview. Asian Pac. J. Trop. Biomed. 2014, 4, 682–694. [Google Scholar] [CrossRef] [Green Version]
  199. Global Basil Essential Oil Market 2019–2023. Increasing Demand for Organic and Cold-Pressed Basil Essential Oil to Boost Growth | Technavio. Available online: https://www.businesswire.com/news/home/20190411005319/en/Global-Basil-Essential-Oil-Market-2019-2023-Increasing-Demand-for-Organic-and-Cold-Pressed-Basil-Essential-Oil-to-Boost-Growth-Technavio (accessed on 12 October 2021).
  200. Mali, P. Comparison of Methyl Eugenol Levels and Eugenol O-Methyltransferase Gene Structure in Different Ocimum Plant Species. Ph.D. Thesis, De Montfort University, Leicester, UK, 2018. [Google Scholar]
  201. Uma, D.P. Radioprotective, anticarcinogenic and antioxidant properties of the Indian holy basil, Ocimum sanctum (Tulasi). Indian J. Exp. Biol. 2001, 39, 185–190. [Google Scholar]
  202. Gupta, S.K.; Prakash, J.; Srivastava, S. Validation of claim of Tulsi, Ocimum sanctum Linn as a medicinal plant. Indian J. Exp. Biol. 2002, 40, 765–773. [Google Scholar]
  203. Rastogi, S.; Kalra, A.; Gupta, V.; Khan, F.; Lal, R.K.; Tripathi, A.K.; Parameswaran, S.; Gopalakrishnan, C.; Ramaswamy, G.; Shasany, A.K. Unravelling the genome of Holy basil: An “incomparable” “elixir of life” of traditional Indian medicine. BMC Genom. 2015, 16, 413. [Google Scholar] [CrossRef] [Green Version]
  204. Tangpao, T.; Chung, H.H.; Sommano, S.R. Aromatic Profiles of Essential Oils from Five Commonly Used Thai Basils. Foods 2018, 7, 175. [Google Scholar] [CrossRef] [Green Version]
  205. Mondal, S.; Mirdha, B.R.; Mahapatra, S.C. The science behind sacredness of Tulsi (Ocimum sanctum Linn.). Indian J. Physiol. Pharmacol. 2009, 53, 291–302. [Google Scholar] [PubMed]
  206. Rahman, S.; Islam, R.; Kamruzzaman, M.; Alam, M.; Jamal, M.A. Ocimum sanctum L.: A Review of Phytochemical and Pharmacological Profile. Am. J. Drug Discov. Dev. 2011, 154, 455–460. [Google Scholar]
  207. Silva, M.G.D.; Matos, F.J.D.; Lopes, P.R.O.; Silva, F.O.; Holanda, M.T. Composition of essential oils from three Ocimum species obtained by steam and microwave distillation and supercritical CO2 extraction. Arkivoc 2004, 2004, 66–71. [Google Scholar] [CrossRef] [Green Version]
  208. Zheljazkov, V.D.; Cantrell, C.L.; Evans, W.B.; Ebelhar, M.W.; Coker, C. Yield and composition of Ocimumbasilicum L. and Ocimum sanctum L. grown at four Locations. HortScience 2008, 43, 737–741. [Google Scholar] [CrossRef] [Green Version]
  209. Joshi, R.K. Chemical Composition, In Vitro Antimicrobial and Antioxidant Activities of the Essential Oils of Ocimumgratissimum, O. Sanctum and Their Major Constituents. Indian J. Pharm. Sci. 2013, 75, 457–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. British Pharmacopoeia Commission. Deoxyribonucleic Acid (DNA) Based Identification Techniques for Herbal Drugs; British Pharmacopoeia Appendix XI V; British Pharmacopoeia Commisssion: London, UK, 2017. [Google Scholar]
  211. Bhamra, S.; Heinrich, M.; Howard, C.; Johnson, M.; Slater, A. DNA authentication of tulsi (Ocimumtenuiflorum) using the nuclear ribosomal internal transcribed spacer (ITS) and the chloroplast intergenic spacer trnH-psbA. Planta Med. 2015, 81, 20. [Google Scholar] [CrossRef]
  212. British Pharmacopoeia Commission. DNA Barcoding as a Tool for Botanical Identification of Herbal Drugs; British Pharmacopoeia Supplementary Chapter SC VII D; British Pharmacopoeia Commission: London, UK, 2018. [Google Scholar]
  213. Ríos-Rodríguez, D.; Sahi, V.P.; Nick, P. Authentication of holy basil using markers relating to a toxicology-relevant compound. Eur. Food Res. Technol. 2021, 247, 2485–2497. [Google Scholar] [CrossRef]
  214. Khalid, K.; Hendawy, S.F.; El-Gezawy, E. Ocimumbasilicum L. production under organic farming. Res. J. Agric. Biol. Sci. 2006, 2, 25–32. [Google Scholar]
  215. Christina, V.; Arunachalam, A. Nucleotide based validation of Ocimum species by evaluating three candidate barcodes of the chloroplast region. MolEcolResour 2003, 14, 60–68. [Google Scholar]
  216. Fügel, R.; Carle, R.; Schieber, A. Quality and authenticity control of fruit purées, fruit preparations and jams—A review. Trends Food Sci. Technol. 2005, 16, 433–441. [Google Scholar] [CrossRef]
  217. Kurz, C.; Leitenberger, M.; Carle, R.; Schieber, A. Evaluation of fruit authenticity and determination of the fruit content of fruit products using FTNIR spectroscopy of cell wall components. Food Chem. 2010, 119, 806–812. [Google Scholar] [CrossRef]
  218. Klaudija, C.S.; Zlatko, L.; Olivera, P.; Strikić, F.; Kolak, I.; Milos, M.; Satovic, Z. Molecular and chemical characterization of the most widespread Ocimum species. Plant Syst. Evol. 2012, 249, 253–262. [Google Scholar]
  219. Mafra, I.; Ferreira, I.; Oliveira, M. Food authentication by PCR-based methods. Eur. Food Res. Technol. 2008, 227, 649–665. [Google Scholar] [CrossRef]
  220. Nieto, G. Biological Activities of three essential oils of the Lamiaceae family. Medicines 2017, 4, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Global Oregano Oil Market Outlook. Available online: https://www.expertmarketresearch.com/reports/oregano-oil-market (accessed on 18 July 2021).
  222. Skoula, M.; Harborne, J.B. Taxonomy and Chemistry. In Oregano: The Genera Origanum and Lippia; Kintzios, S.E., Ed.; Taylor and Francis CRC Press: Boca Raton, FL, USA, 2002; pp. 67–108. [Google Scholar]
  223. World Checklist of Lamiaceae (Origanum). Available online: http://wcsp.science.kew.org/namedetail.do?name_id=143731 (accessed on 11 August 2020).
  224. Stefanaki, A.; Andel, T. Mediterranean aromatic herbs and their culinary use. In Aromatic Herbs in Foods Bioactive Compounds, Processing, and Application; Galanakis, C., Ed.; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  225. Tucker, A.O.; DeBaggio, T. The Big Book of Herbs: A Comprehensive Illustrated Reference to Herbs of Flavor and Fragrance; Interweave Press: Loveland, CO, USA, 2000. [Google Scholar]
  226. Vannozzi, A.; Lucchin, M.; Barcaccia, G. cpDNA Barcoding by Combined End-Point and Real-Time PCR Analyses to Identify and Quantify the Main Contaminants of Oregano (Origanum vulgare L.) in Commercial Batches. Diversity 2018, 10, 98. [Google Scholar] [CrossRef] [Green Version]
  227. Singletary, K. Oregano: Overview of the literature on health benefits. Nutr. Today 2010, 45, 129–138. [Google Scholar] [CrossRef] [Green Version]
  228. Toncer, O.; Karaman, S.; Diraz, E. An annual variation in essential oil composition of Origanum syriacum from Southeast Anatolia of Turkey. J. Med. Plants Res. 2010, 4, 1059–1064. [Google Scholar]
  229. Lòpez, C.; Martin-Sánchez, A.; Fuentes-Zaragoza, E.; Viuda-Martos, M.; Fernández-López, J.; Sendra, E.; Sayas, E.; Pérez-Álvarez, J. Role of oregano (Origanum vulgare) essential oil as a surface fungus inhibitor on fermented sausages: Evaluation of its effect on microbial and physicochemical characteristics. J. Food Prot. 2012, 75, 104–111. [Google Scholar]
  230. Alwafa, R.A.; Mudalal, S.; Mauriello, G. Origanum syriacum L. (Za’atar), from Raw to Go: A Review. Plants 2021, 10, 1001. [Google Scholar] [CrossRef] [PubMed]
  231. Fleishe, A.; Sneer, N. Oregano spices and Origanum chemotypes. J. Sci. Food Agric. 1982, 33, 441–446. [Google Scholar] [CrossRef]
  232. Hazzit, M.; Baaliouamer, A.; Faleiro, M.L.; Miguel, M.C. Composition of the Essential Oils of Thymus and Origanum Species from Algeria and Their Antioxidant and Antimicrobial Activities. J. Agric. Food Chem. 2006, 54, 6314–6321. [Google Scholar] [CrossRef]
  233. Kordali, S.; Cakir, A.; Ozer, H.; Çakmakçı, R.; Kesdek, M.; Mete, E. Antifungal, phytotoxic and insecticidal properties of essential oil isolated from Turkish Origanumacutidens and its three components, carvacrol, thymol and p-cymene. Bioresour. Technol. 2008, 99, 8788–8795. [Google Scholar] [CrossRef] [PubMed]
  234. Figueredo, G.; Chalchat, J.C.; Pasquier, B. Studies of Mediterranean oregano populations IX: Chemical composition of essential oils of seven species of oregano of various origins. J. Essent. Oil Res. 2006, 18, 411–415. [Google Scholar] [CrossRef]
  235. Napoli, E.; Giovino, A.; Carrubba, A.; How Yuen Siong, V.; Rinoldo, C.; Nina, O.; Ruberto, G. Variations of Essential Oil Constituents in Oregano (Origanum vulgare subsp. viridulum (= O. heracleoticum) over Cultivation Cycles. Plants 2020, 9, 1174. [Google Scholar] [CrossRef]
  236. Marieschi, M.; Torelli, A.; Poli, F.; Sacchetti, G.; Bruni, R. RAPD-based method for the quality control of Mediterranean oregano and its contribution to pharmacognostic techniques. J. Agric. Food Chem. 2009, 57, 1835–1840. [Google Scholar] [CrossRef] [PubMed]
  237. Marieschi, M.; Torelli, A.; Bianchi, A.; Bruni, R. Detecting Saturejamontana L. and Origanummajorana L. by means of SCAR-PCR in commercial samples of Mediterranean oregano. Food Control 2011, 22, 542–548. [Google Scholar] [CrossRef]
  238. Black, C.; Haughey, S.A.; Chevallier, O.P.; Galvin-King, P.; Elliott, C.T. A comprehensive strategy to detect the fraudulent adulteration of herbs: The oregano approach. Food Chem. 2016, 210, 551–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Bejar, E. Adulteration of Oregano Herb, and Essential Oil of Oregano; Botanical Adulterants Prevention Bulletin: Austin, TX, USA, 2019. [Google Scholar]
  240. Węglarz, Z.; Kosakowska, O.; Przybył, J.L.; Pióro-Jabrucka, E.; Bączek, K. The Quality of Greek Oregano (O. vulgare L. Subsp. hirtum (Link) Ietswaart) and Common Oregano (O. vulgare L. Subsp. vulgare) Cultivated in the Temperate Climate of Central Europe. Foods 2020, 9, 1671. [Google Scholar] [CrossRef]
  241. Ekor, M. The growing use of herbal medicines: Issues relating to adverse reactions and challenges in monitoring safety. Front. Pharmacol. 2014, 4, 177. [Google Scholar] [CrossRef] [Green Version]
  242. Papaioannou, C.; Zeliou, K.; Trigas, P.; Papasotiropoulos, V. High Resolution Melting (HRM) Genotyping in the Genus Origanum: Molecular Identification and Discrimination for Authentication Purposes. Biochem. Genet. 2020, 58, 725–737. [Google Scholar] [CrossRef]
  243. Yaprak, A.; Mehmet, C. A new natural hybrid of Scutellaria (Lamiaceae) from Turkey. Phytotaxa 2011, 29, 51–55. [Google Scholar]
  244. Zhao, T.; Tang, H.; Xie, L.; Zheng, Y.; Ma, Z.; Sun, Q. Scutellariabaicalensis Georgi (Lamiaceae): A review of its traditional uses, botany, phytochemistry, pharmacology and toxicology. J. Pharm. Pharmacol. 2019, 71, 1353–1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Brock, C.; Whitehouse, J.; Tewfik, I.; Towell, T. American Skullcap (Scutellarialateriflora): A randomised, double-blind placebo-controlled crossover study of its effects on mood in healthy volunteers. Phytother. Res. 2014, 28, 692–698. [Google Scholar] [CrossRef]
  246. Dong, Q.; Chu, F.; Wu, C.; Huo, Q.; Gan, H.; Li, X.; Liu, H. Scutellariabaicalensis Georgi extract protects against alcohol induced acute liver injury in mice and affects the mechanism of ER stress. Mol. Med. Rep. 2016, 13, 3052–3062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Shen, J.; Li, P.; Liu, S.; Liu, Q.; Li, Y.; Sun, Y.; He, C.; Xiao, P. Traditional uses, ten-years research progress on phytochemistry and pharmacology, and clinical studies of the genus Scutellaria. J. Ethnopharmacol. 2021, 265, 113198. [Google Scholar] [CrossRef]
  248. Cheng, L.; Han, M.; Yang, L.; Li, Y.; Sun, Z.; Zhang, T. Changes in the physiological characteristics and baicalin biosynthesis metabolism of Scutellariabaicalensis Georgi under drought stress. Ind. Crops Prod. 2018, 122, 473–482. [Google Scholar] [CrossRef]
  249. Wang, L.; Zhang, D.; Wang, N.; Li, S.; Tanb, H.-Y.; Feng, Y. Polyphenols of Chinese skullcap roots: From chemical profiles to anticancer effects. R. Soc. Chem. 2019, 9, 25518–25532. [Google Scholar] [CrossRef] [Green Version]
  250. Chen, M.; Xiao, H.; Chen, B.; Bian, Z.; Kwan, H.Y. The advantages of using Scutellariabaicalensis and its flavonoids for the management of non-viral hepatocellular carcinoma. J. Funct. Foods 2020, 78, 104389. [Google Scholar] [CrossRef]
  251. Shang, X.; He, X.; He, X.; Li, M.; Zhang, R.; Fan, P.; Zhang, Q.; Jia, Z. The genus Scutellaria an ethnopharmacological and phytochemical review. J. Ethnopharmacol. 2010, 128, 279–313. [Google Scholar] [CrossRef] [PubMed]
  252. Islam, M.N.; Downey, F.; Ng, C.K.Y. Comparative analysis of bioactive phytochemicals from Scutellariabaicalensis, Scutellarialateriflora, Scutellariaracemosa, Scutellaria tomentosa and Scutellariawrightii by LC-DAD-MS. Metabolomics 2017, 7, 446–453. [Google Scholar] [CrossRef]
  253. Mamadalieva, N.Z.; Herrmann, F.; El-Readi, M.Z.; Tahrani, A.; Hamoud, R.; Egamberdieva, D.R.; Azimova, S.S.; Wink, M. Flavonoids in Scutellariaimmaculata and S. ramosissima(Lamiaceae) and their biological activity. J. Pharm. Pharmacol. 2011, 63, 1346–1357. [Google Scholar] [CrossRef]
  254. He, L.; Sun, F.; Wang, Y.; Zhu, J.; Fang, J.; Zhang, S.; Yu, Q.; Gong, Q.; Ren, B.; Xiang, X.; et al. HMGB1 exacerbates bronchiolitis obliterans syndrome via RAGE/NF-κB/HPSE signaling to enhance latent TGF-β release from ECM. Am. J. Transl. Res. 2016, 8, 1971–1984. [Google Scholar] [PubMed]
  255. Zhao, Q.; Chen, X.Y.; Martin, C. Scutellariabaicalensis, the golden herb from the garden of Chinese medicinal plants. Sci. Bull. 2016, 61, 1391–1398. [Google Scholar] [CrossRef] [Green Version]
  256. Georgieva, Y.; Katsarova, M.; Gercheva, K.; Bozov, P.; Dimitrova, S. HPLC analysis of flavonoids from Scutellariaaltissima. Bulg. Chem. Commun. 2019, 51, 119–123. [Google Scholar]
  257. Makino, T.; Hishida, A.; Goda, Y.; Mizukami, H. Comparison of the major flavonoid content of S. baicalensis, S. lateriflora, and their commercial products. J. Nat. Med. 2008, 62, 294–299. [Google Scholar] [CrossRef]
  258. Sun, J.; Chen, P. A flow-injection mass spectrometry fingerprinting method for authentication and quality assessment of Scutellarialateriflora-based dietary supplements. Anal. Bioanal. Chem. 2011, 401, 1577–1584. [Google Scholar] [CrossRef]
  259. Sandasi, M.; Vermaak, I.; Chen, W.; Viljoen, A.M. Skullcap and Germander: Preventing Potential Toxicity through the Application of Hyperspectral Imaging and Multivariate Image Analysis as a Novel Quality Control Method. Planta Med. 2014, 80, 1329–1339. [Google Scholar] [CrossRef] [Green Version]
  260. Gafner, S. Skullcap Adulteration Laboratory Guidance Document; American Botanical Council: Austin, TX, USA, 2015. [Google Scholar]
  261. Zhao, F.; Chen, Y.P.; Salmaki, Y.; Drew, B.T.; Wilson, T.C.; Scheen, A.C.; Celep, F.; Bräuchler, C.; Bendiksby, M.; Wang, Q.; et al. An updated tribal classification of Lamiaceae based on plastomephylogenomics. BMC Biol. 2021, 19, 2. [Google Scholar] [CrossRef]
  262. Jiang, D.; Zhao, Z.; Zhang, T.; Zhong, W.; Liu, C.; Yuan, Q.; Huang, L. Chloroplast Genome Sequence of Scutellariabaicalensis Provides Insight into Intraspecific and Interspecific Chloroplast Genome Diversity in Scutellaria. Genes 2017, 8, 227. [Google Scholar] [CrossRef]
  263. Gafner, S.; Bergeron, C.; Batcha, L.L.; Reich, J.; Arnason, J.T.; Burdette, J.E.; Pezzuto, J.M.; Angerhofer, C.K. Inhibition of [3H]- LSD binding to 5-HT7 receptors by flavonoids from Scutellarialateriflora. J. Nat. Prod. 2003, 66, 535–537. [Google Scholar] [CrossRef] [PubMed]
  264. Sundaresan, P.R.; Slavoff, S.A.; Grundel, E.; White, K.D.; Mazzola, E.; Koblenz, D.; Rader, J.I. Isolation and characterisation of selected germander diterpenoids from authenticated Teucrium chamaedrys and T. canadense by HPLC, HPLC-MS and NMR. Phytochem. Anal. 2006, 17, 243–250. [Google Scholar] [CrossRef]
  265. Lin, L.-Z.; Harnly, J.M.; Upton, R. Comparison of the phenolic component profiles of skullcap (Scutellarialateriflora) and germander (Teucrium canadense and T. chamaedrys), a potentially hepatotoxic adulterant. Phytochem. Anal. 2009, 20, 298–306. [Google Scholar] [CrossRef] [Green Version]
  266. Chen, S.; Yao, H.; Han, J.; Liu, C.; Song, J.; Shi, L.; Zhu, Y.; Ma, X.; Gao, T.; Pang, X.; et al. Validation of the ITS2 region as a novel DNA barcode for identifying medicinal plant species. PLoS ONE 2010, 5, e8613. [Google Scholar] [CrossRef] [PubMed]
  267. Guo, X.; Wang, X.; Su, W.; Zhang, G.; Zhou, R. DNA barcodes for discriminating the medicinal plant Scutellariabaicalensis (Lamiaceae) and its adulterants. Biol. Pharm. Bull. 2011, 34, 1198–1203. [Google Scholar] [CrossRef] [Green Version]
  268. Walker, J.B.; Sytsma, K.J.; Treutlein, J.; Wink, M. Salvia (Lamiaceae) is not monophyletic: Implications for the systematics, radiation, and ecological specializations of Salvia and tribe Mentheae. Am. J. Bot. 2004, 91, 1115–1125. [Google Scholar] [CrossRef] [PubMed]
  269. Wei, Y.K.; Wang, Q.; Huang, Y.B. Species diversity and distribution of Salvia (Lamiaceae). Biodivers. Sci. 2015, 23, 3–10. [Google Scholar]
  270. Drew, B.; González-Gallegos, J.G.; Xiang, C.L.; Kriebel, R.; Drummond, C.P.; Walked, J.B.; Sytsma, K.J. Salvia united: The greatest good for the greatest number. Taxon 2017, 66, 133–145. [Google Scholar] [CrossRef] [Green Version]
  271. Will, M.; Classen-Bockhoff, R. Time to split Salvia s.l. (Lamiaceae)—New insights from old world Salvia phylogeny. Mol. Phylogenet. Evol. 2017, 109, 33–58. [Google Scholar] [CrossRef]
  272. Hu, G.X.; Takano, A.; Drew, B.T.; Liu, E.D.; Soltis, D.E.; Soltis, P.S.; Peng, H.; Xiang, C.L. Phylogeny and staminal evolution of Salvia (Lamiaceae, Nepetoideae) in East Asia. Ann. Bot. 2018, 122, 649–668. [Google Scholar] [CrossRef]
  273. Wang, M.; Zhao, H.X.; Wang, L.; Wang, T.; Yang, R.W.; Wang, X.L.; Zhou, Y.H.; Ding, C.B.; Zhang, L. Potential use of DNA barcoding for the identification of Salvia based on cpDNA and nrDNA sequences. Gene 2013, 528, 206–215. [Google Scholar] [CrossRef] [PubMed]
  274. Yadav, A.; Joshi, A.; Kothari, S.L.; Kachhwaha, S.; Purohit, S. Medicinal, nutritional and industrial applications of Salvia species: A revisit. Int. J. Pharm. Sci. Rev. Res. 2017, 43, 27–37. [Google Scholar]
  275. Gao, C.; Wu, C.; Zhang, Q.; Zhao, X.; Wu, M.; Chen, R.; Zhao, Y.; Li, Z. Characterization of chloroplast genomes from two Salvia medicinal plants and gene transfer among their mitochondrial and chloroplast genomes. Front. Genet. 2020, 11, 574962. [Google Scholar] [CrossRef] [PubMed]
  276. Wang, B.Q. Salviamiltiorrhiza:Chemicalandpharmacologicalreviewofamedicinalplant. J. Med. Plants Res. 2010, 425, 2813–2820. [Google Scholar]
  277. Ullah, R.; Nadeem, M.; Khalique, A.; Imran, M.; Mehmood, S.; Javid, A.; Hussain, J. NutritionalandtherapeuticperspectivesofChia (Salviahispanica L.): A review. J. Food Sci. Technol. 2016, 53, 1750–1758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Ali, N.M.; Yeap, S.K.; Ho, W.Y.; Beh, B.K.; Tan, S.W.; Tan, S.G. Thepromisingfutureofchia, Salviahispanica L. J. Biomed. Biotechnol. 2012, 2012, 1–9. [Google Scholar] [CrossRef] [Green Version]
  279. Hrnčič, M.K.; Ivanovski, M.; Cör, D.; Knez, Z. ChiaSeeds (Salviahispanica L.): An Overview-Phytochemical Profile, IsolationMethods, and Application. Molecules 2019, 25, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Felemban, L.F.; Attar, A.M.A.; Zeid, I.M.A. MedicinalandNutraceuticalBenefitsofChiaSeed(Salviahispanica). J. Pharm. Res. Int. 2021, 32, 15–26. [Google Scholar] [CrossRef]
  281. Reisfield, A.S. Thebotany of Salviadivinorum (Labiateae). Sida Contrib. Bot. 1993, 15, 349–366. [Google Scholar]
  282. Clebsch, B. The New Book of Salvias: Sages for Every Garden; Timber Press: Portland, ME, USA, 2008. [Google Scholar]
  283. Claßen-Bockhoff, R.; Wester, P.; Tweraser, E. Thestaminallevermechanismin Salvia L.(Lamiaceae)—A review. Plant Biol. 2008, 8, 33–41. [Google Scholar]
  284. Bentham, G. LabiatarumGeneraetSpecies; Ridgeway: London, UK, 1836. [Google Scholar]
  285. Briquet, J. Salvia. In Die NaturlichenPflanzenfamiliennebst IhrerGattungenund WichtigerenArten; Engler, A., Prantl, K., Eds.; WilhelmEngelmann: Leipzig, Germany, 1897; pp. 183–287. [Google Scholar]
  286. Epling, C. The Californiasalvias. A review of Salvia, section Audibertia. Ann. Mo. Bot. Gard. 1938, 25, 95–188. [Google Scholar] [CrossRef]
  287. Epling, C. A RevisionofSalvia, SubgenusCalosphace; VerlagdesRepertoriums: Berlin, Germany, 1939. [Google Scholar]
  288. Pobedimova, E.G. Salvia. In Floraofthe U.S.S.R; Shishkin, B.K., Ed.; Izdatel’stvoAkademiiNauk: Moscow, Russia, 1954; pp. 154–255. [Google Scholar]
  289. Wu, C.Y. Salvia. In FloraReipublicaePopularisSinicae; Wu, C.Y., Ed.; SciencePress: Beijing, China, 1977; pp. 70–196. [Google Scholar]
  290. Murata, G.; Yamazaki, T. Salvia. In FloraofJapan; Iwatsuki, K., Yamazaki, T., Boufford, D., Ohba, H., Eds.; Kodansha: Tokyo, Japan, 1993; pp. 302–307. [Google Scholar]
  291. Kriebel, R.; Drew, B.; Drummond, C.; González-Gallegos, J.; Celep, F.; Malik, M.; Rose, J.; Xiang, C.; Hu, G.; Walker, J.; et al. Tracking temporal shifts in area, biomes, and pollinators in the radiation of Salvia (sages) across continents: Leveraging anchored hybrid enrichment and targeted sequence data. Am. J. Bot. 2019, 106, 573–597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Wu, H.; Ma, P.F.; Li, H.T.; Hu, G.X.; Li, D.Z. Comparative plastomic analysis and insights into the phylogeny of Salvia (Lamiaceae). Plant Divers. 2020, 43, 15–26. [Google Scholar] [CrossRef] [PubMed]
  293. Galvin-King, P.; Haughey, S.A.; Montgomery, H.; Elliott, C.T. The rapid detection of sage adulteration using Fourier Transform Infra-Red (FTIR) Spectroscopy and chemometrics. J. AOAC Int. 2019, 102, 354–362. [Google Scholar] [CrossRef]
  294. Rodríguez, S.D.; Gagneten, M.; Farroni, A.E.; Percibaldi, N.M.; Buera, M.P. FT-IR and untargeted chemometric analysis for adulterant detection in chia and sesame oils. Food Control 2019, 105, 78–85. [Google Scholar] [CrossRef]
  295. Tulukcu, E.; Cebi, N.; Sagdic, O. Chemical fingerprinting of seeds of some Salvia species in Turkey by using GC-MS and FTIR. Foods 2019, 8, 118. [Google Scholar] [CrossRef] [Green Version]
  296. Bielecka, M.; Pencakowski, B.; Stafiniak, M.; Jakubowski, K.; Rahimmalek, M.; Gharibi, S.; Matkowski, A.; Slusarczyk, S. Metabolomics and DNA-Based Authentication of Two Traditional Asian Medicinal and Aromatic Species of Salvia subg. Perovskia. Cells 2021, 10, 112. [Google Scholar] [CrossRef]
  297. Dormontt, E.E.; van Dijk, K.; Bell, K.L.; Biffin, E.; Breed, M.F.; Byrne, M.; Caddy-Retalic, S.; Encinas-Viso, F.; Nevill, P.G.; Shapcott, A.; et al. Advancing DNA Barcoding and Metabarcoding Applications for Plants Requires Systematic Analysis of Herbarium Collections—An Australian Perspective. Front. Ecol. Evol. 2018, 6, 134. [Google Scholar] [CrossRef] [Green Version]
  298. Kress, W.J. Plant DNA barcodes: Applications today and in the future. J. Syst. Evol. 2017, 55, 291–307. [Google Scholar] [CrossRef] [Green Version]
  299. Conserving Genetic Diversity of the Endangered Texas Endemic, Big Red Sage, Salvia pentstemonoides. Available online: https://www.publicgardens.org/events/conserving-genetic-diversity-endangered-texas-endemic-big-red-sage-salvia-pentstemonoides (accessed on 11 December 2021).
  300. Zhou, X.; Zhang, Z.C.; Huang, Y.B.; Xiao, H.W.; Wu, J.J.; Qi, Z.C.; Wei, Y.K. Conservation Genomics of Wild Red Sage (Salvia miltiorrhiza) and Its Endangered Relatives in China: Population Structure and Interspecific Relationships Revealed from 2b-RAD Data. Front. Genet. 2021, 12, 688323. [Google Scholar] [CrossRef]
  301. IUCN Red List of Threatened Species. Available online: https://www.iucnredlist.org/species/139603526/139603531 (accessed on 10 December 2021).
  302. Batovska, J.; Cogan, N.O.; Lynch, S.E.; Blacket, M.J. Using Next-Generation Sequencing for DNA Barcoding: Capturing Allelic Variation in ITS2. G3 Genes Genom. Genet. 2017, 7, 19–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Wilkinson, M.; Szabo, C.; Ford, C.; Yarom, Y.; Croxford, A.E.; Camp, A.; Gooding, P. Replacing Sanger with Next Generation Sequencing to improve coverage and quality of reference DNA barcodes for plants. Sci. Rep. 2017, 7, 46040. [Google Scholar] [CrossRef] [Green Version]
  304. Hebert, P.D.N.; Braukmann, T.W.A.; Prosser, S.W.J.; Ratnasingham, S.; deWaard, J.R.; Ivanova, N.V.; Janzen, D.H.; Hallwachs, W.; Naik, S.; Sones, J.E.; et al. A Sequel to Sanger: Amplicon sequencing that scales. BMC Genom. 2018, 19, 219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Fopa, F.B.; Brunel, D.; Bérard, A.; Rivoal, J.B.; Gallois, P.; Le-Paslier, M.C.; Bouverat-Bernier, J.P. Quick and efficient approach to develop genomic resources in orphan species: Application in Lavandula angustifolia. PLoS ONE 2020, 15, e0243853. [Google Scholar]
  306. Lukas, B.; Novak, J. Thecompletechloroplastgenomeof Origanumvulgare L.(Lamiaceae). Gene 2013, 528, 163–169. [Google Scholar] [CrossRef] [PubMed]
  307. He, S.L.; Yang, Y.; Tian, Y. Characteristic and phylogenetic analyses of chloroplast genome for Mentha haplocalyx (Lamiaceae). Mitochondrial DNA B 2020, 5, 2099–2100. [Google Scholar] [CrossRef] [PubMed]
  308. Huaizhu, L.; Bai, L.; Bai, J.; Wang, P.; Zhou, C.; Lingling, D.; Jiang, J.; Liu, J.; Wang, Q. The complete chloroplast genome sequence of Thymus mongolicus (Labiatae), a special spice plant. Mitochondrial DNA B 2020, 5, 2597–2598. [Google Scholar] [CrossRef]
  309. Balaji, R.; Ravichandiran, K.; Tanuja, P.M. The complete chloroplast genome of Ocimumgratissimum from India—A medicinal plant in the Lamiaceae. Mitochondrial DNA B 2021, 6, 948–950. [Google Scholar] [CrossRef]
  310. Yesuthason, R.S.; Balaji, R.; Tanuja, P.M. The complete chloroplast genome and phylogenetic analysis of Ocimumkilimandscharicum Gurke (Camphor Basil) from India. Mitochondrial DNA B 2021, 6, 2164–2165. [Google Scholar]
  311. Ma, L. The complete chloroplast genome sequence of the fragrant plant Lavandula angustifolia (Lamiaceae). Mitochondrial DNA B 2018, 3, 135–136. [Google Scholar] [CrossRef] [Green Version]
  312. Yen, L.T.; Park, J. The complete chloroplast genome sequence of Origanum majorana L. Mitochondrial DNA B 2021, 6, 1224–1225. [Google Scholar] [CrossRef]
  313. Zhang, C.; Xia, P.; Wu, R.; Mans, D. The complete chloroplast genome of Scutellariameehanioides (Lamiaceae) from Shaanxi Province, China. Mitochondrial DNA B 2021, 6, 1685–1686. [Google Scholar] [CrossRef]
  314. Little, D.P. A DNA mini-barcode for land plants. Mol. Ecol. Resour. 2014, 14, 437–446. [Google Scholar] [CrossRef]
  315. Meusnier, I.; Singer, G.A.; Landry, J.F.; Donal A Hickey, D.A.; Hebert, P.D.N.; Hajibabaei, M. A universal DNA mini-barcode for biodiversity analysis. BMC Genom. 2008, 9, 214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Cheng, X.; Su, X.; Chen, X.; Zhao, H.; Bo, C.; Xu, J.; Bai, H.; Ning, K. Biological ingredient analysis of traditional Chinese medicine preparation based on high-throughput sequencing: The story for LiuweiDihuang Wan. Sci. Rep. 2014, 4, 5147. [Google Scholar] [CrossRef]
  317. Bruno, A.; Sandionigi, A.; Agostinetto, G.; Bernabovi, L.; Frigerio, J.; Casiraghi, M.; Labra, M. Food Tracking Perspective: DNA Metabarcoding to Identify Plant Composition in Complex and Processed Food Products. Genes 2019, 10, 248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Gao, Z.; Liu, Y.; Wang, X.; Wei, X.; Han, J. DNA Mini-Barcoding: A derived barcoding method for herbal molecular identification. Front. Plant Sci. 2019, 10, 987. [Google Scholar] [CrossRef] [PubMed]
  319. Raime, K.; Krjutškov, K.; Remm, M. Method for the identification of plant DNA in food using alignment-free analysis of sequencing reads: A case study on Lupin. Front. Plant Sci. 2020, 11, 646. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of nrDNA region with ITS region’s primer (ITS1/ITS4) localization (arrows). ETS (External transcribed spacers).
Figure 1. Schematic representation of nrDNA region with ITS region’s primer (ITS1/ITS4) localization (arrows). ETS (External transcribed spacers).
Plants 11 00137 g001
Figure 3. Lavender products market analysis [180] and phylogenetic relationship among the species.
Figure 3. Lavender products market analysis [180] and phylogenetic relationship among the species.
Plants 11 00137 g003
Figure 4. Basil products market analysis [199] and phylogenetic relationship among the species.
Figure 4. Basil products market analysis [199] and phylogenetic relationship among the species.
Plants 11 00137 g004
Figure 5. Oregano products market analysis [221] and phylogenetic relationship among the species.
Figure 5. Oregano products market analysis [221] and phylogenetic relationship among the species.
Plants 11 00137 g005
Figure 6. Skullcap products, applications and phylogenetic relationship among the species of Scutellarioideae and other subfamilies of Lamiaceae.
Figure 6. Skullcap products, applications and phylogenetic relationship among the species of Scutellarioideae and other subfamilies of Lamiaceae.
Plants 11 00137 g006
Figure 7. Morphological similarities between Scutellaria and Teucrium.
Figure 7. Morphological similarities between Scutellaria and Teucrium.
Plants 11 00137 g007
Figure 8. Salvia products, applications and phylogenetic relationship among subgenera.
Figure 8. Salvia products, applications and phylogenetic relationship among subgenera.
Plants 11 00137 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nazar, N.; Howard, C.; Slater, A.; Sgamma, T. Challenges in Medicinal and Aromatic Plants DNA Barcoding—Lessons from the Lamiaceae. Plants 2022, 11, 137. https://doi.org/10.3390/plants11010137

AMA Style

Nazar N, Howard C, Slater A, Sgamma T. Challenges in Medicinal and Aromatic Plants DNA Barcoding—Lessons from the Lamiaceae. Plants. 2022; 11(1):137. https://doi.org/10.3390/plants11010137

Chicago/Turabian Style

Nazar, Nazia, Caroline Howard, Adrian Slater, and Tiziana Sgamma. 2022. "Challenges in Medicinal and Aromatic Plants DNA Barcoding—Lessons from the Lamiaceae" Plants 11, no. 1: 137. https://doi.org/10.3390/plants11010137

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