Next Article in Journal
Nematicidal, Acaricidal and Plant Growth-Promoting Activity of Enterobacter Endophytic Strains and Identification of Genes Associated with These Biological Activities in the Genomes
Next Article in Special Issue
Steroidal Saponins with Plant Growth Stimulation Effects; Yucca schidigera as a Commercial Source
Previous Article in Journal
Peach Palm (Bactris gasipaes Kunth.): Ancestral Tropical Staple with Future Potential
Previous Article in Special Issue
Energy vs. Nutritional Potential of Virginia Mallow (Sida hermaphrodita L.) and Cup Plant (Silphium perfoliatum L.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Pyranocoumarins: Description, Biosynthesis, Application

1
Laboratory of Cell and Developmental Biology, Federal Scientific Center of East-Asia Terrestrial Biodiversity, Far Eastern Branch of the Russian Academy of Sciences, Stoletiya Vladivostoka Ave. 159, Vladivostok 690022, Russia
2
Laboratory of Biomedical Cell Technologies of the Center for Genomic and Regenerative Medicine, Institute of Life Sciences and Biomedicine, Far Eastern Federal University, FEFU Campus, 10 Ajax Bay, Russky Island, Vladivostok 690922, Russia
3
Department of Biology, Institute of Natural Sciences, M.K. Ammosov North-Eastern Federal University, 58 Belinsky Str., Yakutsk 677000, Russia
*
Author to whom correspondence should be addressed.
Plants 2022, 11(22), 3135; https://doi.org/10.3390/plants11223135
Submission received: 14 October 2022 / Revised: 7 November 2022 / Accepted: 10 November 2022 / Published: 16 November 2022

Abstract

:
This overview article contains information about pyranocoumarins over the last 55 years. The article is based on the authors’ phytochemical and physiological studies in vivo and in vitro as well as search and analysis of data in literature available on Google Scholar, Web of Science, PubMed, and ScienceDirect before January 2022. Pyranocoumarins are synthesized in plants of the Apiaceae, Rutaceae families, and one species in each of the Cornaceae, Calophyllaceae, and Fabaceae families can synthesize this class of compounds. The physiological role of these compounds in plants is not clear. It has been proven that these substances have a wide range of biological activities: anti-cancer, anti-spasmatic, and anticoagulant, and they also inhibit erythrocyte lysis and accumulation of triacylglycerides. The overview generalizes the modern understanding of the classification, structure, and biological activity of natural pyranocoumarins, and summarizes dispersed data into a unified scheme of biosynthesis. The review analyzes data on the localization and productivity of these substances in individual organs and the whole plant. It discusses a link between the unique structure of these substances and their biological activity, as well as new opportunities for pyranocoumarins in pharmacology. The article evaluates the potential of different plant species as producers of pyranocoumarins and considers the possibilities of cell cultures to obtain the end product.

1. Introduction

Throughout the entire written history of the humanity, and certainly long before that, plants have been used as medicines [1]. Even though knowledge of medical plants has been mainly obtained empirically, this resulted in the development of the main pharmaceutical drugs necessary for modern medicine. It has been calculated that almost half of the drugs approved over the last 30 years are of natural origin or are produced from natural compounds, many of which are plants [2]. Some famous examples of medicines include phenol compounds of natural origin: cholagogue flamen produced from Helichrysum arenarium, antimicrobial shikonin produced from Lithospermum erythrorhizon, anticoagulant dicoumarin produced from plants of the Melilótus genus, etc. A big part of plant chemical diversity remains unstudied, and there are more than 28,000 species of medical plants that may potentially be used to develop new drugs [3]. Searching for unique compounds having high biological activity is the basis of development of the global pharmacological industry. At the same time, three species of higher plants become extinct every year, which is 500 times faster than natural extinction [4]. Because of this, humanity irretrievably loses unique structures of natural compounds that may have biological activity and become a solution in treating certain diseases.
Pyranocoumarins are rare secondary metabolites of plants: coumarin containing a pyran core. Coumarins are widely spread compounds of plant origin that are essential for the life of plants and that perform protective and regulatory functions [5]. Depending on the structural modification, they have a number of physiological activities: anti-spasmodic action, capillary-reinforcing, sedative, diuretic, vermicidal, analgesic, and antimicrobial action. Some coumarins stimulate functions of the central nervous system, reduce the blood cholesterol, prevent clot formation in blood vessels and promote their dissolution, increase skin UV-sensitivity, accelerate ulcer healing, stimulate breathing, and increase the arterial pressure [6]. Unlike coumarins, the physiological role of pyranocoumarins in plants is not clear. However, there is data describing various biological activities of these secondary metabolites in animals and humans: anti-cancer, anti-spasmatic, and anticoagulant activities. Moreover, they inhibit erythrocyte lysis and accumulation of triacylglycerides.
Pyranocoumarins represent a narrow group of substances synthesized in a limited number of plant species of the Apiaceae and Rutaceae families, and rarely in other families. Mechanisms of pyranocoumarin synthesis in plant cells and in whole plants are poorly investigated.
Literature presents disparate data for the diversity of pyranocoumarins in flora and their biological activity in various test models. However, these data need to be systematized and aggregated in order to predetermine the areas of future studies. It is necessary to understand the structural diversity of pyranocoumarins and their functions since even small structural changes define physiological properties. Overviews of pyranocoumarins in the global scientific literature are rare. A very interesting overview by Beena and Sathya Pooja in 2022 was dedicated to the heterocycle of pyranocoumarins, namely to important analogs of pyranocoumarin with various biological activities [7].
Generalized experimental results of R&D centers of different countries analyzed in this paper allow us to recreate the complicated biosynthesis path of pyranocoumarins, show the important functional role of these substances, and find the conditions supporting synthesis of this narrow group of substances in various plant species. The article is based on our own phytochemical and physiological studies in vivo and in vitro, as well as a search and analysis of the literature available on Google Scholar, Web of Science, PubMed, and ScienceDirect before January 2022. Thus, this overview is intended to systematize and analyze the information in the available literature concerning this narrow and unique group of substances.

2. Pyranocoumarins Structure and Classification

Pyranocoumarins belong to the class of coumarins, derivatives of 2H-1-benzopyran-2-on [8]. These substances containing a pyran core condensed with coumarin at 4,5; 5,6; 6,7; 7,8 positions and having substitutes in pyran, benzene, and pyridine rings [9,10]. Structurally, depending on how the pyran ring is formed, there are three groups of pyranocoumarins: linear (Figure 1a), angular (Figure 1b,c), and condensed (Figure 1d). Natural linear pyranocoumarins have substitution in the parent benzenoid ring and/or in heteronuclear rings [11]. The structures of the pyranocoumarins overviewed here are represented in Supplementary Materials (Table S2 and Figure S1).

3. Biosynthesis of Pyranocoumarins in Cells

A precursor of all pyranocoumarins is umbelliferone whose biosynthesis is implemented via the phenylpropanoid pathway (Figure 2) [12]. Exposed to phenylalanine ammonia-lyase (PAL)), L-phenylalanine transforms into trans-cinnamic acid followed by hydroxylation by cinnamate-4-hydroxylase (C4H), forming p-coumaric acid. This acid may also be formed from another product of the shikimate pathway—L-tyrosine, when exposed to tyrosine ammonia lyase (TAL). Exposed to 4-coumarat-CoA ligase, p-coumaric acid provides p-coumaroyl-CoA. Then the molecule is once again hydroxylated by cinnamate/coumarate-2-hydroxylase to produce 2,4-dihydroxy acid (umbellic acid) followed by rotation of the double bond neighboring the carbonic acid group. Finally, the intra-molecular attack of the hydroxyl group C2 on the carbonic acid group encloses the ring and forms the phenylpropanoid umbelliferone, precursor of pyranocoumarins [13].
Exposed to prenyltransferases, umbelliferone forms either osthenol, a precursor of 7,8-angular pyranocoumarins, or 7-demethylsuberosin—a precursor of linear pyranocoumarins [14]. In an experiment with carbon label conducted for orange seedlings (Citrus sinensis (L.) Osbeck) and rangpur seedlings (Citrus limonia Osbeck), it has been proven that the coumarin fragment of pyranocoumarins xanthyletin and seselin originate from the shikimate pathway (L-phenylalanine), and the pyran ring is formed via the methyl–erythrite–phosphate pathway [15]. Transformation and coupling of substitutes are most likely related to various transferases. There are suggestive data on the formation of decursinol and its derivatives [12,14]. Zhao with coauthors combined methods of transcriptomics and metabolomics to build the biosynthesis pathway of coumarins in Peucedanum praeruptorum Dunn, including pyranocoumarins, and suggested that genes CYP450 and MDR participated in coumarin biosynthesis [16]. The data about the biosynthesis pathway of other two groups of pyranocoumarins is limited by what is shown in Figure 3. Gómez-Robledo with coauthors made structural analysis of calanolides (dipyranocoumarins) and transcriptomic analysis using RNA sequencing and suggested biosynthesis pathways and unigene candidates in the dataset of transcriptome potentially participating in biosynthesis pathways of umbelliferone and calanolide [13]. Biosynthesis pathways of condensed pyranocoumarins are suggested to be started with Sturangin C, a coumarin characteristic of the Mammea genus, which is the only genus of plants with condensed pyranocoumarins [17,18]. In any case, the precursor must have substitutes of 1-hydroxypropyl and prenyl or geranyl at C4 and C6 positions of coumarin (Figure 3). Therefore, the tandem of transcriptomic and biochemical analyses can give new opportunities in studying biosynthesis pathways of pyranocoumarins. Herewith, experiments may use producing plants without employing model objects.

4. Pyranocoumarins Distribution in Plant World

Pyranocoumarins are found only among the Umbelliferae and Rutaceae families, and in one species from each of Calophyllaceae, Cornaceae, and Leguminosae. Distribution of pyranocoumarin forms within primary families producing these substances shows significant predomination of one of the structural forms within one family. For example, 58% of substances in the Umbellíferae family are represented by angular form and only 23% are characterized by linear form. The Rutaceae family have predominating linear form of structure in 57% of species. Interestingly, the number of species where substances with different molecular structure are found is 18% or less in both families (Figure 4). This distribution of pyranocoumarin forms may be due to the evolutionary early occurrence of species and existence of a pathway from 7-dimethylsuberoside, while the more evolutionary developed Rutaceae have osthenol derivatives as predominating. Groups of substances with the condensed form are described only in Mammea siamensis T. Anders. from the Calophyllaceae family. Distribution of 5,6-angular pyranocoumarins is also poorly studied. It should be noted that this may change both because of the differentiation and determination methods of the substances and the increased number of species and substances studied.
Table S2 contains data for the distribution of pyranocoumarins; however, attention should be paid to the years of the studies. For some of them, no reliable methods for confirming the structure were available; and in other cases, minor compounds were not determined. Modern research largely focuses on determining minor compounds and studying their distribution in organs, but they do not cover all pyranocoumarin-containing plants and give no grounds to reveal the role in the body as a whole. The available data shows that folk medicine uses seeds and roots of pyranocoumarin-containing plants more often, so it can be vaguely suggested that these organs contain more such compounds, which is confirmed by some data in literature. Usually, plants accumulate substances with a protective function in seeds and roots.
Functions of pyranocoumarins in the plant are poorly studied. However, quite accurate data have been found concerning the effects of coumarins on gravitropism of roots [19], distribution of auxin [20], and protection of organs [5]. Based on papers studying their biological activity, we may suggest that pyranocoumarins play a protective role against phytopathogens [21,22]. However, this conclusion is poorly supported, since there are very few papers intended to investigate the functions of pyranocoumarins in plants.

5. Biological Activity of Pyranocoumarins

This paper is intended to analyze data for biological activity of plant extracts containing pyranocoumarins and individual substances from 1994 to 2022. The collected materials give a general overview of study areas and principles of influence of these unique substances on animal and human cell cultures, fungi, bacteria in vitro and on objects in vivo. For convenience, the results are distributed by various types of activity as a table and are not exhaustive (see Table 1).

5.1. Anti-Inflammatory Activity

Inflammation is part of a complicated biological reaction of the body to harmful irritants such as pathogens, toxins, or cell and tissue damage, and it represents a protective mechanism involving immune cells, blood vessels, and molecular mediators [62]. Without an inflammatory reaction that activates the immune system to combat pathogens, people would not be able to survive even weak infections. Therefore, inflammation fulfills a protective function in the body, but only when it is acute. Chronic inflammation ceases to be a physiological factor and becomes a pathogen causing various diseases such as hay fever, periodontium diseases, atherosclerosis, and osteoarthritis. Since inflammation is the cause and/or effect of many diseases, so searching for anti-inflammatory agents is always relevant.
Literature provides a number of papers showing anti-inflammatory mechanisms of pyranocoumarins. Son et al., from Chungnam University (Daejeon, Korea) and Kyungsung University (Busan, Korea) summarized the works and describe how decursin extracted from Angelica gigas suppressed the inflammatory process in mice RAW264.7 cells. The action mechanism was associated with inhibition of NF-κB transcription factor, which led to reduced expression of proteins of matrix metalloproteinase 9, monocyte chemoattractant protein 1, interleukin 8, tumor necrosis factor α, and IL-1β. Decursin increases the differentiation of IκB, a substance in the body that inhibits the activity of NF-κB, and suppresses the inflammatory response by inhibiting the differentiation and migration of NF-κB to the nucleus [23].
Praeruptorin A extracted from Peucedanum praeruptorum was tested for anti-inflammatory activity against RAW264.7 cells stimulated by lipopolysaccharides. Praeruptorin A in cells inhibited the production of nitrogen oxide (NO), interleukine-1β (IL-1β), and tumor necrosis factor α (TNF-α) and suppressed the expression of mRNA and proteins of inducible nitrogen oxide synthase (iNOS), IL-1β and TNF-α. Praeruptorin A reduced cytoplasmic loss of inhibitor protein κB-α (IκB-α) and thereby inhibited translocation of NF-κB from cytoplasm to the nucleus, which led to an anti-inflammatory effect [24,25]. The same effect against the signaling pathway of NF-kB translocation in RAW264.7 cell line followed by further suppression of various inflammation factors was shown by corymbocoumarin extracted from Seseli gummiferum subsp. corymbosum [26]. Reduced production of NO was also observed when administering pyranocoumarin extracted from Clausena emarginata, 10-(3,7-Dimethylocta-1,6-dien-3-yl)-5-methoxy-8,8-dimethylpyranocoumarin (the half maximal inhibitory concentration (IC50) = 8 µg/mL) to the mouse microglia line BV2 [27]. The reasons of such effects have not been described by the authors, but the mechanism is likely to be similar with the one for RAW264.7 cells.
The anti-inflammatory action of praeruptorin A and visnadine extracted from Ligusticum lucidum. Cuneifolium at 115.2 µg/cm2 of the area of inflammation of the skin surface and 116.52 µg/cm2, respectively, reduced edema in laboratory mice. Aural dermatitis in mice was induced by croton oil. Praeruptorin A decreased edema by 22% and visnadine by 43%. Visnadine differed from Praeruptorin A in the content of 3′-β-O-dihydroangelolic and 4′-β-O-acetyl residues in the pyran ring, and the authors suggest that the structure determined the increase in the anti-inflammatory effect [28].
Some artificially synthesized angular pyranocoumarins (structures of 2a, 2b, 3a, 5a are shown in the Figure S1 in Supplementary Materials) at 250 µg/mL inhibited proteinase activity. The action of pyranocoumarins was 30% to 79.72% inhibition of protein denaturation. The highest anti-proteinase activity was shown by 4 pyranocoumarins, and their action was stronger than in control aspirin (45.83 ± 4.21%). Two individual pyranocoumarins (2b and 3a) with the highest activities (79.72 ± 4.51% and 74.68 ± 3.01%) were also tested in vivo on laboratory rats with formaldehyde-induced paw edema. Pyranocoumarins and control indomethacin were administered orally; as a result, paw edema decreased in rats by 29.2% (when 3a was administered) and 6.57% (when 2b was administered). The results of pyranocoumarin introduction gave a significantly lower effect compared with the control indomethacin (up to 72%), but this may be due to the fact that the administered dose of the active ingredient was one tenth of the indomethacin concentration given to rats. The action of pyranocoumarins towards cyclooxygenases (COX-1 and COX-2) was also tested. Cyclooxygenase is an enzyme that is responsible for formation of prostanoids, including thromboxane and prostaglandins. Inhibition of COX can provide relief from the symptoms of inflammation and pain. Most pyranocoumarin derivatives had a better effect against COX-2 (IC50 ≤ 10 µM). Compound 5a showed significant selectivity against COX-2, which was the highest among the tested compounds (SI = 152) compared with standard celecoxib (SI = 243.9). A slightly lower activity was shown by compounds 2a and 2b with SI 64 and 47.3, respectively. The authors suggested that the anti-inflammatory activity of artificially synthesized pyranocoumarins is related to the amino groups in their structure [29].

5.2. Antioxidant Activity

Occurrence of degenerative processes is correlated with an excess of free radicals promoting oxidative processes that are harmful for the body. Free radicals have a high reactivity so they quickly deprive other molecules of electrons trying to restore their stability. Substances in the cells deprived of electrons become free radicals and also need the electron balance to be restored. As a result, a chain reaction in the body disrupts biochemical processes and destroys cells and tissues. Highly antioxidant substances neutralize free radicals. High content of compounds with antioxidant properties capable of scavenge free radicals (carotenoids, phenol, flavonoid and anthocyanin derivatives, unsaturated fatty acids, vitamins, enzymes, and cofactors) in plants has stimulated research into their use in preventive and curative herbal medicine [30].
Grandivittin, agasyllin, and aegelinol extracted from the roots of Ferulago campestris were tested on the cell line of polymorphonuclear neutrophils (PMN), generating active forms of oxygen in white blood cells of human whole blood (WB). Antioxidant activity of pyranocoumarins within the range of concentrations of 0.01 to 100 µg/mL was assessed on the basis of inhibition of chemical luminescence (CL). Better values were obtained for aegelinol having a hydroxyl residue at 3′ position of the pyran ring, followed by agasyllin with angeloyl at the third position of the coumarin ring and grandivittin with residual senesoyl. Inhibiting activity had linear dose-dependence showing faster inhibition in stimulation. The authors suggested that pyranocoumarins absorb the active form of oxygen or intervene with cell activation mechanisms [30].
Superoxide anion, hydrogen peroxide and radical hydroxyl are the most dangerous radicals known. The cell possesses CAT, GPx, and SOD—enzymes that remove ROS—therefore, the increase in these enzymes provides protection against the harmful effects evoked by ROS. Artificially synthesized 3′,4′-di-O-acetyl-cis-khellactone (DOAcK) was tested on laboratory rats with diabetes induced by streptozotocin. Administering DOAcK to rats increased the activity of antioxidant enzymes CAT, GPx and SOD [29]. Similar works on natural pyranocoumarins were not found.

5.3. Antimicrobial Activity

The question of treating bacterial and fungi infections has always been critical in the world. Wide application of antibiotic and antimycotic agents in agricultural and medical practice gave rise to the issue of antibiotic resistance of pathogen microorganisms. Moreover, antibiotics often cause allergic reactions. Finding efficient, safe, high-quality and environmentally clean plant raw materials with antibacterial activity is a topical issue.
Lee, et al., have tested the antibacterial activity of decursin and decursinol angelate extracted from the roots of Angelica gigas on pathogenic bacterial strains—Escherichia coli, Proteus vulgaris, Salmonella typhimurium, Bacillus subtilis, Staphylococcus epidermis, and S. aureus. The better effect was observed towards the B. subtilis strain for the concentration of active ingredients of 50 and 12.5 µg/mL for decursinol angelate and decursin, respectively [32].
Agasyllin and aegelinol extracted from Ferulago campestris were tested for various species of bacteria, showing significant activity against gram-negative and gram-positive bacteria at 16 to 125 µg/mL. The tests showed that the substances suppressed growth in ATCC strains S. aureus, Salmonella thypii, Enterobacter cloacae, and E. earogenes. Minimum inhibitory concentrations (MIC) were 16 and 32 µg/mL for aegelinol and agasyllin, respectively. Antibacterial activity of both coumarins was higher against gram-negative bacteria [33].
Pd-D-V and disenecioyl hellactone extracted from P. decursivum were tested against S. sclerotiorum, B. cinerea, F. graminearum, and C. capsici. Pd-DV at 30 µg/mL in the culture medium inhibited the growth of S. sclerotiorum by 86.4 ± 0.7 (Half maximal effective concentration (EC50) = 13.2 µg/mL) and of B. cinerea, F. graminearum, and C. capsici by almost 50%. Pd-D-V showed not only a wide range of effects against various bacteria, but also the best antimicrobial effect (EC50 > 56 µg/mL). Disenecioyl hellactone showed the lowest activity at the same concentration and inhibited the growth of S. sclerotiorum and B. cinerea by almost 50% (EC50 = 11.0 µg/mL), and of F. framinearum, T. cucumeris and C. capsica by almost 30% [34].
Artificially synthesized 3,4-angular pyranocoumarins (structures of pyranocoumarins 5a and 5c are shown in the Figure S1 in Supplementary Materials) were tested on ATCC bacterial strains: gram-positive—S. aureus, B. subtilis, B. megaterium, gram-negative—E. coli and Pseudomonas aeruginosa; on strains of yeast fungi from the standard collection of cultures NRRL—S. cerevisiae and Candida albicans. Screening was done by the agar diffusion test. The compounds showed high activity against gram-positive bacteria, but only two compounds (5a and 5c) containing non-substituted benzal and 3,4,5-trimethoxybenzal, respectively, showed a strong effect against gram-negative bacteria (zones of inhibition (IZ): 24–26 mm). Of all the compounds, only one (5a) had a prominent antimicrobial activity against most tested microorganisms (IZ 22–27 mm) with a MIC of 250–500 µg/mL [29].
As antibacterial agents, pyranocoumarins can be used both in medicine and agriculture. Below are several successful works against phytopathogens.
Decursin and decursinol angelate were also tested in vitro and in vivo against pathogen fungi—Botrytis cinerea, Colletotricum coccodes, Fusarium oxysporum f. sp. raphani, Magnaporthe oryzae, Rhizoctonia solani, and Sclerotinia sclerotiorum, as well as two Oomycetes: Phytophthora capsici and P. infestans. The methanol extract of A. gigas root efficiently suppressed the development of pyricularyosis caused by M. oryzae, fungus from Askomycetes, by 75% at 100 µg/mL. Despite the low effect on the growth of fungi mycelium, the separately extracted decursin and decursinol angelate strongly inhibited the growth of M. oryzae spores, which was not found when using control fungicide, kasugamycin, for treatment. Among other tested pathogens, B. cinerea was also the most sensitive to the two pyranocoumarins: its mycelium growth was suppressed by 83%, which is similar to the effect of the control fungicide. This was followed by an in vivo test confirming the action of individual pyranocoumarins against pathogens in rice. For the in vivo test, whole rice plants were treated with decursinol angelate. When treated with 300 µg/mL decursinol angelate, the leaves were not damaged by blast. Treatment with each coumarin separately and with a combination thereof suppressed the disease by more than 80% at 100 and 30 µg/mL. A synergetic effect between the two pyranocoumarins against the pathogen was not observed. No phytotoxic consequences of adding pyranocoumarins were found, which indicates that the studied concentrations of these compounds can be safely used against fungal diseases of plants [22].
Extract of Psoralea corylifolia seeds with identified phenyl derivative of pyranocoumarin (PDP) was tested in vitro on F. oxysporum, F. graminearum, and F. moniliforme. The extract showed an inhibitory effect on the growth of individual pathogens at 10 and 1 mg/mL. Molecular modeling using trichothecene-3-O-acetyltransferase indicated affinity of PDP and the target protein, as well as the ability of pyranocoumarin to bind to the protein, thereby inhibiting the protein acetylation mechanism leading to the death of Fusarium sp. The authors noted that this suggestion still has to be confirmed by a more accurate analysis [21].

5.4. Anti-Cancer Activity

According to GLOBOCAN, 19.3 million new cancer cases and almost 9.96 million cancer deaths were recorded in the world in 2020. The global burden of cancer is expected to be 30.2 million cases in 2024, which is 56% higher than in 2020 (https://gco.iarc.fr/tomorrow/en/dataviz/isotype accessed on 15 October 2022). Figures show a high rise in emerging economies (64% to 95%) compared with developing economies (32% to 56%) due to demographic changes, though this may be aggravated by the increasing risk factors related to globalization and economic growth [63]. Cancer remains one of the most important problems of humanity, which is far from being solved. This disease has a different etiology and is characterized by different manifestations, so it is necessary to search for new methods of treatment. Secondary metabolites of plants can become prototypes of new cures for cancer. Rare compounds synthesized by plants can have useful pharmacological properties. There are a lot of potential anti-cancer agents among pyranocoumarins.
Some of the well-studied pyranocoumarins are decursin and decursinol angelate. Their anti-tumor activity was first shown in vivo in laboratory mice inoculated with tumor cells of sarcoma 180. For this purpose, decursin and decursinol angelate were sequentially injected intraperitoneally for 9 days at 50 and 100 mg/kg. Pyranocoumarin therapy of mice increased the longevity and reduced weight and volume of tumors in mice [35]. The extract of A. gigas containing decursin and decursinol angelate exhibited anti-cancer effect on HeLa cells through the significant activation of caspases and the cleavage of PARP. Expression of TRAIL and TRAIL receptors was also increased by two active components of A. gigas, respectively. In particular, two components regulated the expression of apoptosis-related proteins such as Bcl-2, Bcl-x, survivin, cIAP-1, -2, XIAP, etc. These results suggest that TRAIL expression induced by extract of A. gigas stimulates the extrinsic and intrinsic apoptosis pathway by activating caspase-8 and caspase-9, respectively [42]. Another test on a line of androgen-depending LNCaP showed that decursin in the medium stops G1 phase of cancer cells. This happened due to the inhibition of androgen-stimulating nuclear translocation of androgen receptors and decreased amount of androgen receptor protein [36]. An anti-tumor effect was observed in the mammary gland cancer line upon introducing decursin and decursinol angelate. Both pyranocoumarins increased the population of MCF-7 cells in G1 phase and decreased the number of cells in S phase (G1 stop) within 3.28–16.42 µg/mL. Apoptotic DNA fragmentation was found using ELISA in 24 h and using Western blot analysis of split PARP 48 h after treatment. The test showed an increased caspase-mediated apoptosis at 16.42 µg/mL decursin or 16.42 µg/mL decursinol angelate in MCF-7 [38]. Another paper attributed the anti-cancer effect to suppression of the Wnt/β-catenin pathway, counteracting the transcription of β-catenin response. Decursin also suppressed the expression of D1 cyclin and c-myc, which are β-catenin target genes, so that pyranocoumarin inhibited the growth of PC-3 prostate cancer cells [39]. For RC-58T/h/SA#4 prostate cancer cells, decursin reduced the cell proliferation on a dose-dependent basis (6.57; 13.14; 19.70; 26.27 and 32.84 µg/mL) for 48 h. Cancer cells were more sensitive to decursin than normal cells. Stain testing of Hoechst 33,258 cells indicated that decursin promoted condensation of chromatin and apoptotic bodies [40]. Decursin also exhibited a cytotoxic effect in combination with bortezomib in human multiple myeloma cells. Decursin inhibited the viability of U266, MM.1S, and ARH77 cells, but not of peripheral blood mononuclear cells (PBMC). Decursin-induced apoptosis was mediated by activation of caspase-8, -9, and -3 and by inhibition of constitutive STAT3 activation by inhibiting activation of Janus-activated kinase 2 (JAK2) in U266 cells. Moreover, decursin inhibited interleukin-6-induced STAT-3 activation on a dose-dependent basis in MM.1S cells. Due to these properties, decursin significantly amplified the apoptotic properties of bortezomib against U266 cells [41]. A. gigas extract containing decursin and decursinol angelate showed an inhibitory action on HeLa cervical cancer cells with up to 50% cancer cell proliferation at 6.57 µg/mL. An anti-proliferative action of decursin and decursinol angelate on normal cells of the A549 or AGS lines was not observed at the same concentration. This mechanism was attributed to the fact that the extract increases the expression of TRAIL that stimulates extrinsic and intrinsic apoptosis pathways by activating caspase-8 and caspase-9 [42].
Decursinol extracted from Saposhnikovia divaricata suppressed the growth of DU-154 prostate cancer cells (GI50 3.09 µg/mL), was more efficient compared with the standard drug cisplatin, and showed high activity towards MEL-8 melanoma cells (GI50 = 3.46 µg/mL) [43].
Clausarin extracted from Clausena excavate, as well as a number of synthesized pyranocoumarins, were tested for the cytotoxicity against A549, MCF-7, KB, and KB-VIN human tumor cells. Clausarin showed the best anti-tumor effect against A549, KB, and KB-VIN with EC50 from 1.59 to 2.98 µg/mL. Clausenidin and two other synthesized compounds demonstrated a significant level of activity only towards KB-VIN cells with EC50 from 2.25 to 2.87 µg/mL, while the activity was insignificant or none at all against other lines. The study results are interesting since KB-VIN cells has multiple drug resistance [44].
(±)-Praeruptorin A and (±)-praeruptorin B extracted from Peucedanum praeruptorum were tested for SGC7901 human stomach cancer cells. The study showed that (±)-praeruptorin A increased sensitivity to drugs allegedly by inhibiting the expression of P-glycoprotein responsible for drug efflux from cancer cells via ATF-depending pumps. Moreover, the cell growth was decreasing faster in combined treatment with pyranocoumarin and doxorubicin than when using the chemotherapeutic agent alone, which suggests that (±)-praeruptorin A can reduce the doxorubicin dose to achieve the desired effects [45].
To confirm the activity of pyranocoumarins extracted from P. praeruptorum roots, the permeability of cells by calcein-AM was considered relative to MES-SA/Dx5 poly-drug resistant cancer cells. Five angular pyranocoumarins at 10 µM each showed a significantly higher activity compared with known inhibitors (verapamil and ciclosporin A) [25].
Grandivitin extracted from Ferulago macropara, inhibited matrix metalloproteinase 9 (MMP9). Matrix metalloproteinases, a large family of extracellular zinc-dependent endopeptidases, play a major role in the ischemic cell injury. Their targets include other proteinases, proteinase inhibitors, clotting factors, chemotactic molecules, latent growth factors, growth factor-binding proteins, cell surface receptors, cell–cell adhesion molecules, and virtually all structural extra-cellular matrix proteins. Grandivitin suppressed protein fluorescence binding to it in the hydrophobic cavity and thereby changed the secondary and tertiary MMP9 structures. The biological significance of this work is evident because MMP9 serves as a potential target protein for anticancer agents. The binding study of grandivitin with MMP9 is of great importance in pharmacy, pharmacology and biochemistry [45].
Clausenidin was also tested on HepG2 hepatocellular carcinoma cells. After being introduced into the medium, clausenidin increased the caspase-8 activity and expression of protein components in the death-inducing signaling complex (DISC) in HepG2 cells. The ultra-structural analysis found morphological abnormalities typical for apoptosis. Moreover, clausenidin reliably reduced expression of the vascular endothelial growth factor (VEGF) [47]. Another group of researchers have tested the cytotoxicity of clausenidin extracted from C. excavata on the colon adenocarcinoma line HT29 and normal colon cell line CCD-18Co. Clausenidin in DMSO and in combination with hydroxypropyl-β-cyclodextrin (Clu/HPβCD) was used for the study [48]. Cyclodextrins can increase the water-solubility of low-soluble substances and improve drug penetration through biological membranes [64]. The cytotoxic action of clausenidin in combination with HPβCD against cancer cells was stronger than when dissolved in DMSO. When comparing the action with normal cells, the side effect was reduced. The authors noted that the Clu/HPβCD complex initiated cytotoxicity mediated by active oxygen forms in HT29 cells, which resulted in a stopped cell cycle and death caused by apoptosis associated with caspase activation [47].
Anti-cancer activity was primarily shown for linear pyranocoumarins. It should be noted that the search found no papers studying the action mechanism of angular pyranocoumarins. The two forms of angular pyranocoumarins and the kinetics of linear pyranocoumarins are still to be studied. Dimeric pyranocoumarins—calanolides and condensed pyranocoumarins—also show great potential for future use.

5.5. Antivirus Activity

Antiviral activity is not as widely studied for pyranocoumarins as bacterial infections or cancer. However, the latest global events have become an incentive to search for compounds with antiviral activity.
Causenidin and nordentatin extracted from C. excavata were tested on HepG2 cells that were infected with hepatitis B. Among the tested pyranocoumarin analogs, causenidin and nordentatin had anti-HBV values EC50 of 0.62 µg/mL and 2 µg/mL, respectively. Moreover, other pyranocoumarins were chemically synthesized, among which four compounds showed antiviral activity [44].
A number of artificially synthesized pyranocoumarins were tested on the cell line of a kidney of the African green monkey (CV-1) infected with measles. Some were rather specific and strong MV inhibitors with EC50 from 0.2 to 50 µg/mL, while most EC50 value were below 5 µg/mL. The compounds inhibited nine strains of the measles virus, while the drugs did not physically destroy virion in virulence tests in order to inhibit virus replication [49].

5.6. Antihyperglycemic and Antidyslipidemic Activity

Obesity and overweight have increased at an alarming rate in the world during the last three decades. Obesity is a crucial factor in the development of metabolic abnormalities, including glucose intolerance, insulin resistance, metabolic syndrome, low-grade inflammation and oxidative stress [7]. To reduce obesity-associated risk factors, various methods for prevention and treatment are used, including pharmaceuticals. Disadvantages of conventional synthetic drugs include low efficacy, narrow therapeutic range, toxicity, and major side effects. Plant-derived drugs are safer, and many papers show that using plants against obesity is highly effective [65].
Taira, et al., showed that pyranocoumarins extracted from Peucedanum japonicum Thunb. inhibited lipid accumulation and lipogenous expressions of genes in 3T3-L1 adipocytes [50].
Our experiments on 3T3-L1differentiated adipocytes with seven coumarins from P. sibiricus proved that the total content of triacylglycerols in cells treated with dihydrosamidin was lower at 10 µg/mL (22.6% of the control level) and at 20 µg/mL (19.8% of the control level). The primary coumarin of P. sibiricus roots is an active agent against obesity and can be used against overweight [51].
Hossin, et al., evaluated oral consumption of pure dihydropyranocoumarins (DPC) extracted from P. japonicum in mice fed a high-fat diet and studied their activity when nanoparticles are added. Weight gain and accumulation of fat in white fat tissues were significantly suppressed when DPC was consumed with food (1.943 mg/mouse/day). Consumption of dihydropyranocoumarins significantly decreased the average size of adipocytes and increased the mRNA levels of genes associated with thermogenesis. Nanoparticulation of DPCs with polylactic-co-glycolic acid (PLGA) dramatically increased its activity almost 100-fold over that of a non-nanoparticulated form. Wherein, the results showed activity of dihydropyranocoumarins against obesity in vivo and suggested that encapsulation of PLGA nanoparticles was useful to amplify the DPC activity against obesity in order to develop natural and safe anti-obesity agents [52].
(+)-trans-Decursidinol, Pd-C-I and Pd-C-II, Pd-C-III were extracted from whole Angelica decursiva plants and tested on the HepG2 cell line. The presence of these compounds in the medium increased glucose absorption and reduced expression of PTP1B in insulin-resistant HepG2 cells. (+)-trans-decursidinol showed a high inhibitory action towards PTP1B among extracted compounds with IC50 0.61 µg/mL, which exceeded the action of the control ursolic acid three times (IC50 = 3.14 µg/mL). Pd-C-I, Pd-C-II, and Pd-C-III showed inhibitory activity against PTP1B with IC50 values of 1.35, 2.03, and 4.60 µg/mL, respectively. Coumarins inhibited albumin nitration mediated by peroxynitrites (ONOO-) [66].

5.7. Other Activities

Other biologically active properties of pyranocoumarins, which cannot be classified as any of the activities given above, are considered below. Visnadine extracted from Ammi visnaga fruits was tested on guinea pigs with muscular contractions induced by norepinephrine (16.92 µg/mL). Visnadine predominantly suppressed retractive reactions of K+-spasms in aorta segments mediated by Ca2+ penetration via Ca2+ [53].
The group led by Gulyayev and Urbanova has studied neuropharmacological effects of P. sibiricus extracts. An aqueous infusion of P. sibiricus showed anxiolytic properties, supposedly, by activating GABAA receptor. The GABAA receptor is an ionotropic receptor and ligand-gated ion channel. Its endogenous ligand is γ-aminobutyric acid (GABA), the major inhibitory neurotransmitter in the central nervous system [54]. It has been found that course administration of a P. sibiricus infusion at 0.05 mL/kg limits the development of emotional and neurological disturbances in case of cerebral ischemia in rats [55]. Moreover, P. sibiricus extract was found to inhibit hemolysis of red blood cells induced by peroxide solution, thereby showing the membrane-stabilizing action of coumarins. The best inhibitory effect was shown at 1 µg/mL [56]. The extract was also injected into laboratory rats at 100–400 mg/kg, which had an anti-anxiety action, increased the length of thiopental-induced sleep, and had an anticonvulsant action. This indicates the neuromodulatory activity of the extract, supposedly caused by activating the GABA system by coumarin compounds [57].
(±)-Praeruptorin extracted from P. praeruptorum was tested on laboratory mice with drug-induced asthma. Injection of (±)-praeruptorin suppressed inflammation of airways, hyperactivity and remodeling of airways, reduced IL-4 and IL-13 in BALF and IgE in serum, inhibited TGF-β1 and pSmad2/3 expression, increased Smad7 expression in lung tissue, and increased INF-γ levels in BALF. Due to these processes, mice showed improved condition of airways resulted from their reduced inflammation [58].
Out of seven pyranocoumarins extracted from Clausena emarginata, compounds of clauemarmarin C, D and two known analogs at 10 µM showed hepatoprotective activity against DL-galactosamine-induced damage to the diploid epithelial cell line of normal liver in adult rats with phenotype properties of oval cells WB-F344 [59].
Mammeasins E and F extracted from Mammea siamensi were tested relative to 5α-reductase enzyme. Out of all tested pyranocoumarins, mammeasins E and F caused the highest inhibition of 5α-reductase of testosterone [67].
cis-Khellactone, d-laserpitin, isolaserpitin, and octanoyllomatin extracted from Seseli devenyense were administered to zebrafish larvae. There was an anxiolytic effect of pyranocoumarins defined by the decreasing residence time within the living space of larva (reverse thigmotaxis) [60].
From n-hexane extract of the above-ground parts of Seseli petraeum, pyranocoumarin was extracted, which was called 3′-isovaleryl-4′-oxolomatin (or petracoumarin), as well as 12 previously known pyranocoumarins such as: octanoyllomatin, selinidin, anomalin, 3′-isobutyllomatin, 3′-angeloyl-4′-isovaleryl-cis-hellactone, 3′-isovaleryl-4′-angeloyl-cis-hellactone, calipterixin, samidin, 4′-senecioyl-cis-hellactone, 3′-senecioyl-cis-hellactone, cis-hellactone, and angelicin. Several coumarins were tested for their inhibitory activity against α-amylase and α-glycosidase, and were found to show prominent inhibitory activity against α-glycosidase and a low inhibitory potential against α-amylase. Octanoyllomatin was the best inhibitor of a-glycosidase (IC50 = 69.00 ± 0.43 µg/mL) [61].
Thus, pyranocoumarins are promising antioxidants, anticancer agents, and also show antibacterial, antiglycemic, and neuroprotective effects. However, despite a large number of studies on the various biological activities of various pyranocoumarins, there is practically no data on their bioavailability. It is expected that this issue will be the task of subsequent preclinical trials.

6. Prospects for the Pyranocoumarins Study

The study of pyranocoumarin formation in the cultural medium of plants shows great potential. Using cell cultures in vitro helps study the formation of specialized exchange compounds in controlled conditions and define regulatory mechanisms and systems of this process. The fact that there is no deficiency of materials is equally important. There are certain examples of pyranocoumarin identification in a cell culture; however, regulatory mechanisms were not studied.
We obtained callus and suspension cell cultures from aseptic seedlings of Phlojodicarpus sibiricus. The callus cell cultures of P. sibiricus were characterized by white-yellow color, combination of loose and dense aggregations of cells, and satisfactory growth: growth index by dry biomass for a culture of leaf origin was 7–9, hypocotyl origin 10–12, and root origin 11–13. Suspension cell cultures were initiated from callus cultures of leaf and hypocotyl origin, which also had white-yellow color and predominantly consisted of cell aggregates of meristem-like and parenchyma-like types, while the aggregation degree differed for cultures of different origin. Cell viability during the growing cycle was 70–80%. Unlike the initial callus cultures, the highest growth (growth index of about 10) was shown by the line of suspension cell culture obtained from calluses of leaf origin. Preliminary phytochemical screening using UPLC-MS showed the presence of coumarins of the kellactone group in the biomass of primary (1st–3rd growth cycle) callus and suspension cell cultures of P. sibiricus [68]. This was followed by studies of suspension cultures of P. sibiricus in long-term growing In vitro. The highest growth characteristics were shown by the suspension culture of leaf origin: growth indexes by various categories (dry and wet cell biomass, cell concentration) I = 10–14; specific growth rate μ = 0.3–0.4 day−1; maximum dry biomass accumulation M = 9.6 g/L; economic coefficient Y = 0.29. The cell culture of hypocotyl origin had lower growth characteristics: I = 3.6–4.9, μ = 0.12–0.18 day-1, M = 6.6 g/L, Y = 0.16. Differences in the growth of the studied cultures correlate with the aggregation degree of cells: the suspension culture of cells of leaf origin consists mainly of fine aggregates (10 to 30 cells) while the culture of hypocotyl origin is represented by coarse aggregates (no less than 50 cells in the aggregate). A fine suspension culture of P. sibiricus cells of leaf origin was grown in two types of laboratory bioreactors—a bubble column and a stirred tank. It was found that growing in the bubble column bioreactor is associated with improved primary growth characteristics (growth index by dry biomass I = 12.7; productivity by dry biomass P = 0.78 g/L × day, μ = 0.18 day−1, M = 15.8 g/L, Y = 0.49). When using the stirred tank bioreactor, there was a noticeable reduction in growth characteristics, which is evidently due to the damaging action on the cells by agitation devices [69]. Phytochemical analysis showed significant differences in the composition of phenolic compounds in cells in vitro and in roots of the plant. The cultures of P. sibiricus cells predominantly contained polar (hydrophylic) compounds classified as phenol derivatives—coumarin and benzofuran glycosides. The main components in the roots were more hydrophobic metabolites—ethers of pyranocoumarin kellactone. The analysis of the phytochemical composition of suspension cell cultures of P. sibiricus showed significant differences in the composition of phenol compounds in cells in vitro and in roots of the intact plant. Suspensions cell cultures of P. sibiricus contained predominantly polar (hydrophylic) compounds wherein the main components in the roots were hydrophobic metabolites. Most compounds detected in cells in vitro (except for glycoside kellactone [70]) are not characteristic of intact plants of P. sibiricus [51].
Some of the detected compounds (hexoside ether of prenylated coumarin osthenol/7-dimethylsuberosin, malonyl derivatives of coumarin glycosides) can be classified as quite rare phenol derivatives of plants. Other detected compounds (benzofuran glycosides) are widely spread among various taxons [70,71,72,73,74,75]; however, the available literature apparently contains no information about their presence in Phlojodicarpus spp. [69]. Curiously, prenylated coumarins characteristic of the intact plant (visnadine, dihydrosamidine, and other kellactone ethers) were identified in the initial callus cultures of P. sibiricus, but were not found in long-grown suspension cell cultures [68,69]. The loss of pyranocoumarin synthesis by cell cultures is most likely associated with a change in the strategy for the existence of plant cells in vitro. Heterotrophic nutrition and the absence of the need to protect against pathogens lead to a change or reduction in substances of specialized synthesis. However, further experiments in the form of changing the nutrient medium, the use of elicitors, and transformation can give positive results.
These results are well aligned with the available data concerning changes (in comparison with intact plants) of secondary metabolism in plant cells cultivated in vitro. Such trends were noted in the formation of steroid glucosides in cell cultures of Dioscorea deltoidea Wall. and Tribulus terrestris L. Both cultures only showed accumulation of furostanol glycosides promoting cell proliferation, whereas spirostanol forms of steroid glycosides are more characteristic of intact plants [76,77].
The literature shows at least 10 cell cultures and plant organs being pyranocoumarin producers (see Table 2). However, with the exception of our papers and Xu, et al., no pyranocoumarins were found in cell cultures [68,69,78]. For example, the culture of A. visnaga cells was independently obtained by three groups of researchers, but no secondary metabolites except for synthesis of furanochromones were found [79,80,81,82].
Using cell cultures and tissues as a method for obtaining renewable high-quality plant raw materials for producing biologically active compounds from plants, namely, for creating new and safe functional drugs, has great potential, as well as a number of advantages, for example, a radical solution to the deficiency problem. The produced raw materials are standard, irrespective of weather and climatic conditions, and completely free of all types of contaminations, since the biomass of cell and tissue cultures is grown in sterile and controlled conditions of a laboratory and/or factory [101,102].
In recent years, great success was achieved in metabolic engineering of microorganisms for the purpose of producing valuable compounds of plant origin [103,104]. Levopimaradien and other diterpenoids were successfully obtained in Escherichia coli cells [105]; artemisinic acid in Saccharomyces cerevisiae yeasts [106]; dihydroquerticine in E. coli [107], cannabinoids in S. cerevisiae [108], taxadiene in E. coli [109], noscapine in S. cerevisiae [110] and opioids in S. cerevisiae [111]. Studies of the biosynthesis pathways of secondary metabolites in cell cultures have both fundamental and applied significance irrespective of the source of raw materials. First of all, targeted regulation of biosynthesis becomes possible. Second of all, the known enzyme genes of biosynthesis pathways can be transferred to other organisms (bacteria, fungi, and other plants and/or cells thereof). The production of secondary metabolites in microorganisms is an environmentally clean and renewable method for obtaining the target product compared with chemical synthesis. Therefore, creation of producing microbes for the rapid production of secondary synthesis compounds using metabolic engineering and synthetic biology is rather relevant. A compelling argument for using the bacterial system is the phenylpropanoid pathway in the bacterial cell, as well as the significant reinforcement thereof by adding the signal peptide of cytochrome P450 (TT7) from Arabidopsis thaliana [107]. Successful transformation of primary genes of prenyltransferases cytochrome P450 participating in the biosynthesis of pyranocoumarins into cells of S. cerevisiae and E. coli is shown in the paper by Bu, et al., This paper helped identify the enzymes responsible for the formation of the pyran ring [112].

7. Conclusions

The ability to synthesize various substances of specialized metabolism is a unique characteristic of plants that has allowed them to survive and occupy vast areas. Various groups of secondary metabolites provide protection, attract pollinators, and participate in the regulation of life processes. The variety of substances synthesized by plants helped people produce incenses, dyes, spices, and multiple other essential products that have become common today. Medicines are the most important and relevant ones among them. Searching for unique medicinal drugs in flora remains one of the central tasks of the pharmaceutical industry.
Simple coumarins are widely distributed and fulfill a general protective function. Unlike them, pyranocoumarins belong to a rare sub-group of secondary metabolites synthesized in two families only. This narrow evolutionary adaptation is not random, but most probably caused by the need for protection against a certain pathogen. Selective accumulation of pyranocoumarins in seeds and roots indicates their protective function. This hypothesis has been proven by Yoon and Srinivasan, et al. [21,22]. However, this area requires a deeper study due to the lack of data on pyranocoumarin-containing plant species.
It is rather interesting that the biosynthesis pathway of pyranocoumarins, as any other phenol compounds, starts with the phenylpropanoid pathway, and a proven precursor of all pyranocoumarins is umbelliferon, a coumarin widely spread in the flora. The formation of linear pyranocoumarins from 7-dimethyolsuberoside and 7,8-angular pyranocoumarins from osthenol is well-studied. However, not much is known about the other two, quite rare, 4,5- and 5,6-pyranocoumarins, which are the subject matter of future research.
Pyranocoumarins have various biological activities, but the most interesting and relevant today is anticancer activity. Numerous studies have reliably proven the activity of linear coumarins against various cancer lines. For angular pyranocoumarins, there are fewer experimental works on the action mechanisms, but there are data showing that the growth of cancer cells can be slowed down without significant adverse effects on normal cells. These results are a good prerequisite for the identification of action mechanisms and the development of new drugs. Thus, the information presented in this review indicates the demand for natural pyranocoumarins as promising compounds with a variety of effects on the human body deserving further close study. Unfortunately, many plants that have shown high production of pyranocoumarins (Phlojodicarpus sibiricus) or rare pyranocoumarins (as a group of mammeosines) are endangered or endemic species, which complicates pharmaceutical production. Cell cultures can be a good alternative. If substances play a protective role, they are most likely toxic, and it is rather difficult to achieve their high production in cell culture, where the principle of existence is designed only for cell division. To date, the number of studies on the synthesis of pyranocoumarins in cell cultures is insufficient, but is gradually increasing. Accordingly, it is necessary to carry out targeted studies on the induction/cultivation of organ tissue cultures and whole plants producing pyranocoumarins

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants11223135/s1, Table S1: substituents of pyranocoumarins; Figure S1: structures of artificial pyranocoumarins; Table S2: the content of pyranocoumarins in plants. References [113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, M.T.K. and T.Y.G.; validation, M.T.K. and A.K.S.; formal analysis, M.T.K. and A.K.S.; investigation, M.T.K., A.K.S., T.Y.G., N.K.C.; resources, M.T.K. and A.K.S.; data curation, data curation; writing—original draft preparation, M.T.K., A.K.S. and T.Y.G.; writing—review and editing, M.T.K., T.Y.G. and N.K.C.; visualization, M.T.K. and T.Y.G.; supervision, T.Y.G. and N.K.C.; project administration, M.T.K. and N.K.C.; funding acquisition, M.T.K. and N.K.C. All authors have read and agreed to the published version of the manuscript.

Funding

The experimental work of the authors presented in the review was carried out within the framework of the Ministry of Education and Science of Russia (FSRG-2020-0019), and the review was supported by the Russian Federal Academic Leadership Program Priority 2030.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript.

References

  1. Bynum, H.; Bynum, W. Remarkable Plants That Shape Our World; University of Chicago Press: Chicago, IL, USA, 2014; p. 239. ISBN 978-0226204741. [Google Scholar]
  2. Veeresham, C. Natural products derived from plants as a source of drugs. J. Adv. Pharm. Technol. Res. 2012, 3, 200–201. [Google Scholar] [CrossRef]
  3. Soriano-Castell, D.; Liang, Z.; Maher, P.; Currais, A. The search for anti-oxytotic/ferroptotic compounds in the plant world. Br. J. Pharmacol. 2021, 178, 3611–3626. [Google Scholar] [CrossRef]
  4. Ledford, H. World’s largest plant survey reveals alarming extinction rate. Nature 2019, 570, 148–149. [Google Scholar] [CrossRef] [Green Version]
  5. Zobel, A.M.; Brown, S.A. Coumarins in the interections between the plant and its environment. Allelopath. J. 1995, 2, 9–20. [Google Scholar]
  6. Babekov, A.U.; Keneshov, B.M.; Mamytov, T.A.; Tynybekova, G.T. Pharmacological activity of some biologically active substances. Izv. OshTU 2008, 2, 118–121. [Google Scholar]
  7. Beena, K.P.; Sathya Pooja, G. A Review on Pyranocoumarins and its biological importance. Asian J. Res. Chem. 2022, 15, 176–181. [Google Scholar] [CrossRef]
  8. Vogel, A. Darstellung von Benzoesäure aus der Tonka-Bohne und aus den Meliloten—Oder Steinklee—Blumen. Annal. Phyhk. 1820, 64, 161–166. [Google Scholar] [CrossRef] [Green Version]
  9. Smith, E.; Hosansky, N.; Bywater, W.G.; Tamelen, E.E. Constitution of Samidin, Dihydrosamidin and Visnadin. J. Am. Chem. Soc. 1956, 79, 3534–3540. [Google Scholar] [CrossRef]
  10. Lozhkin, A.V.; Sakanyan, E.I. Natural coumarins: Methods of extraction. Chem. Pharm. J. 2006, 40, 47–56. [Google Scholar] [CrossRef]
  11. Joshi, P.P. A Review in Biological Activities of Linear Pyranocoumarins. Int. J. Adv. Study Res. Work. 2021, 4, 20–27. [Google Scholar] [CrossRef]
  12. Reddy, C.S.; Kim, S.; Hur, M.; Kim, Y.; Park, C.; Lee, W.; Jang, J.; Koo, S. Natural Korean Medicine Dang-Gui: Biosynthesis, Effective Extraction and Formulations of Major Active Pyranocoumarins, Their Molecular Action Mechanism in Cancer, and Other Biological Activities. Molecules 2017, 22, 2170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Gómez-Robledo, H.B.; Cruz-Sosa, F.; Bernabé-Antonio, A.; Guerrero-Analco, A.; Olivares-Romero, J.B.; Alexandro Alonso-Sánchez, A.; Emanuel Villafán, E.; Ibarra-Laclette, E. Identification of candidate genes related to calanolide biosynthesis by transcriptome sequencing of Calophyllum brasiliense (Calophyllaceae). BMC Plant Biol. 2016, 16, 177. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, J.; Li, L.; Jiang, C.; Xing, C.; Kim, S.-H.; Lu, J. Anti-cancer and Other Bioactivities of Korean Angelica gigas Nakai (AGN) and Its Major Pyranocoumarin Compounds. Anti-Cancer Agents Med. Chem. 2012, 12, 1239–1254. [Google Scholar] [CrossRef] [PubMed]
  15. Amaral, J.C.; Silva, M.M.; Silva, M.F.G.F.; Alves, T.C.; Ferreira, G.; Forim, M.R.; Fernandes, J.B.; Pina, E.S.; Lopes, A.A.; Pereira, A.M.S.; et al. Advances in the Biosynthesis of Pyranocoumarins: Isolation and 13C-Incorporation Analysis by High-Performance Liquid Chromatography–Ultraviolet–Solid-Phase Extraction–Nuclear Magnetic Resonance Data. J. Nat. Prod. 2020, 83, 1409–1415. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, Y.; Liu, T.; Luo, J.; Zhang, Q.; Xu, S.; Han, C.; Xu, J.; Chen, M.; Chen, Y.; Kong, L. Integration of a Decrescent Transcriptome and Metabolomics Dataset of Peucedanum praeruptorum to Investigate the CYP450 and MDR Genes Involved in Coumarins Biosynthesis and Transport. Front. Plant Sci. 2015, 6, 996. [Google Scholar] [CrossRef] [Green Version]
  17. Mahandru, M.M.; Ravindran, V.K. Surangin C, a Coumarin from Mammea longifolia. Phytochemistry 1986, 25, 555–556. [Google Scholar] [CrossRef]
  18. Prachyawarakorn, V.; Mahidol, C.; Ruchirawat, S. Pyranocoumarins from the twigs of Mammea siamensis. Phytochemistry 2006, 67, 924–928. [Google Scholar] [CrossRef]
  19. Lupini, A.; Araniti, F.; Sunseri, F.; Abenavoli, M.R. Gravitropic response induced by coumarin: Evidences of ROS distribution involvement. Plant Signal. Behav. 2013, 8, e23156. [Google Scholar] [CrossRef] [Green Version]
  20. Li, X.; Gruber, M.Y.; Hegedus, D.D.; Lydiate, D.J.; Gao, M.-J. Effects of a Coumarin Derivative, 4-Methylumbelliferone, on Seed Germination and Seedling Establishment in Arabidopsis. J. Chem. Ecol. 2011, 37, 880. [Google Scholar] [CrossRef]
  21. Srinivasan, S.; Sarada, D.V.L. Antifungal Activity of Phenyl Derivative of Pyranocoumarin from Psoralea corylifolia L. Seeds by Inhibition of Acetylation Activity of Trichothecene 3-O-Acetyltransferase (Tri101). BioMed Res. Int. 2012, 2012, 310850. [Google Scholar] [CrossRef] [Green Version]
  22. Yoon, M.; Kim, Y.; Ryu, S.; Choi, G.; Jang, K.; Han, S.; Kim, J. In vitro and in vivo antifungal activities of decursin and decursinol angelate isolated from Angelica gigas against Magnaporthe oryzae, the causal agent of rice blast. Pestic. Biochem. Physiol. 2011, 101, 118–124. [Google Scholar] [CrossRef]
  23. Son, C.; Baek, I.; Song, G.; Kang, J.; Kwon, K. Pharmacological Effect of Decursin and Decursinol Angelate from Angelica gigas Nakai. Yakhak Hoeji 2009, 53, 303–313. [Google Scholar]
  24. Yu, P.; Jin, H.; Zhang, J.; Wang, G.; Li, J.; Zhu, Z.; Tian, Y.; Wu, S.; Xu, W.; Zhang, J.; et al. Pyranocoumarins Isolated from Peucedanum praeruptorum Dunn Suppress Lipopolysaccharide-Induced Inflammatory Response in Murine Macrophages Through Inhibition of NF-kB and STAT3 Activation. Inflammation 2012, 35, 967–977. [Google Scholar] [CrossRef]
  25. Lee, J.; Lee, Y.; Kim, J.; Bang, O. Pyranocoumarins from Root Extract of Peucedanum praeruptorum Dunn Multidrug Resistance Reversal and Anti-Inflammatory Activities. Molecules 2015, 20, 20967–20978. [Google Scholar] [CrossRef] [Green Version]
  26. Chun, J.; Tosun, A.; Kim, Y.S. Anti-inflammatory effect of corymbocoumarin from Seseli gummiferum subsp. corymbosum through suppression of NF-κB signaling pathway and induction of HO-1 expression in LPS-stimulated RAW 264.7 cells. Int. Immunopharmacol. 2016, 31, 207–215. [Google Scholar] [CrossRef]
  27. Ouyang, G.Q.; Li, C.J.; Yang, J.Z.; Li, L.; Song, X.Y.; Jiang, Y.N.; Chen, N.H.; Ma, J.; Zhang, D.M. Bioactive Coumarins from the Stems of Clausena emarginata. Chem. Biodivers. 2016, 13, 1178–1185. [Google Scholar] [CrossRef]
  28. Menghini, L.; Epifano, F.; Genovese, S.; Marcotullio, M.C.; Sosa, S.; Tubaro, A. Antiinflammatory activity of coumarins from Ligusticum lucidum Mill. subsp. cuneifolium (Guss.) Tammaro (Apiaceae). Phytother. Res. 2010, 24, 1697–1699. [Google Scholar] [CrossRef]
  29. Alshibl, H.M.; Al-Abdullah, E.S.; Haiba, M.E.; Alkahtani, H.M.; Awad, G.E.A.; Mahmoud, A.H.; Ibrahim, B.M.M.; Bari, A.; Villinger, A. Synthesis and Evaluation of New Couamrin as Antioxidant, Antimicrobal, and Anti-Inflammatory Agents. Molecules 2020, 25, 3251. [Google Scholar] [CrossRef]
  30. Munteanu, I.G.; Apetrei, C. Analytical Methods Used in Determining Antioxidant Activity: A Review. Int. J. Mol. Sci. 2021, 22, 73380. [Google Scholar] [CrossRef]
  31. Domínguez-Mendoza, E.A.; Cornejo-Garrido, J.; Burgueno-Tapia, E.; Ordaz-Pichardo, C. Antidiabetic effect, antioxidant activity, and toxicity of 3′,4′-Di-O-acetyl-cis-khellactone in Streptozotocin-induced diabetic rats. Bioorganic Med. Chem. Lett. 2016, 26, 4086–4091. [Google Scholar] [CrossRef]
  32. Lee, S.; Shin, D.S.; Kim, J.S.; Oh, K.B.; Kang, S.S. Antibacterial coumarins from Angelica gigas roots. Arch. Pharm. Res. 2003, 26, 449–452. [Google Scholar] [CrossRef] [PubMed]
  33. Basile, A.; Sorbo, S.; Spadaro, V.; Bruno, M.; Maggio, A.; Faraone, N.; Rosselli, S. Antimicrobial and Antioxidant Activities of Coumarins from the Roots of Ferulago campestris (Apiaceae). Molecules 2009, 14, 939–952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Song, P.P.; Zhao, J.; Liu, Z.L.; Duan, Y.B.; Hou, Y.P.; Zhao, C.Q.; Wu, M.; Wei, M.; Wang, N.H.; Lv, Y.; et al. Evaluation of antifungal activities and structure-activity relationships of coumarin derivatives. Pest. Manag. Sci. 2017, 73, 94–101. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, S.; Lee, Y.S.; Jung, S.H.; Shin, K.H.; Kim, B.K.; Kang, S.S. Anti-tumor activities of decursinol angelate and decursin from Angelica gigas. Arch. Pharm. Res. 2003, 26, 727–730. [Google Scholar] [CrossRef] [PubMed]
  36. Yim, D.; Singh, R.P.; Agarwal, C.; Lee, S.; Chi, H.; Agarwal, R. A Novel Anticancer Agent, Decursin, Induces G1 Arrest and Apoptosis in Human Prostate Carcinoma Cells. Cancer Res. 2005, 65, 1035–1044. [Google Scholar] [CrossRef]
  37. Jiang, C.; Lee, H.; Li, G.; Guo, J.; Malewicz, B.; Zhao, Y.; Lee, E.; Lee, H.; Lee, J.; Lim, M.; et al. Potent Antiandrogen and Androgen Receptor Activities of an Angelica gigas—Containing Herbal Formulation: Identification of Decursin as a Novel and Active Compound with Implications for Prevention and Treatment of Prostate Cancer. Cancer Res. 2006, 66, 453–463. [Google Scholar] [CrossRef] [Green Version]
  38. Jiang, C.; Guo, J.; Wang, Z.; Xiao, B.; Lee, H.; Lee, E.; Kim, S.; Lu, J. Decursin and decursinol angelate inhibit estrogen-stimulated and estrogen-independent growth and survival of breast cancer cells. Breast Cancer Res. 2007, 9, R77. [Google Scholar] [CrossRef] [Green Version]
  39. Song, G.-Y.; Lee, J.-H.; Cho, M.; Park, B.-S.; Kim, D.-E.; Oh, S. Decursin Suppresses Human Androgen-Independent PC3 Prostate Cancer Cell Proliferation by Promoting the Degradation of β-Catenin. Mol. Pharmacol. 2007, 72, 1599–1606. [Google Scholar] [CrossRef] [Green Version]
  40. Choi, S.R.; Lee, J.H.; Kim, J.Y.; Park, K.W.; Jeong, I.Y.; Shim, K.H.; Lee, M.K.; Seo, K.I. Decursin from Angelica gigas Nakai induces apoptosis in RC-58T/h/SA#4 primary human prostate cancer cells via a mitochondria-related caspase pathway. Food Chem. Toxicol. 2011, 49, 2517–2523. [Google Scholar] [CrossRef]
  41. Kim, S.A.; Oh, H.K.; Kim, J.Y.; Hong, J.W.; Cho, S.I. A Review of Pharmacological Effects of Angelica gigas, Angelica sinensis, Angelica acutiloba and their Bioactive Compounds. J. Korean Med. Sci. 2011, 32, 1–24. [Google Scholar]
  42. Yim, N.; Lee, J.; Cho, W.; Yang, M.; Kwak, D.; Ma, J. Decursin and decursinol angelate from Angelica gigas Nakai induce apoptosis via induction of TRAIL expression in cervical cancer cells. Eur. J. Integr. Med. 2011, 3, 299–307. [Google Scholar] [CrossRef]
  43. Urbagarova, B.M.; Shults, E.E.; Taraskin, V.V.; Radnaeva, L.D.; Petrova, T.N.; Rybalova, T.V.; Frolova, T.S.; Pokrovskii, A.G.; Ganbaatar, J. Chromones and coumarins from Saposhnikovia divaricata (Turcz.) Schischk. Growing in Buryatia and Mongolia and their cytotoxicity. J. Ethnopharmacol. 2020, 261, 112517. [Google Scholar] [CrossRef]
  44. Su, C.R.; Yeh, S.F.; Liu, C.M.; Damu, A.G.; Kuo, T.H.; Chiang, P.C.; Bastow, K.F.; Lee, K.H.; Wu, T.S. Anti-HBV and cytotoxic activities of pyranocoumarin derivatives. Bioorg. Med. Chem. 2009, 17, 6137–6143. [Google Scholar] [CrossRef]
  45. Liang, T.; Yue, W.; Li, Q. Chemopreventive Effects of Peucedanum praeruptorum DUNN and Its Major Constituents on SGC7901 Gastric Cancer Cells. Molecules 2010, 15, 8060–8071. [Google Scholar] [CrossRef]
  46. Gheibi, S.; Shokohinia, Y.; Kiani, A.; Sadrjavadi, K.; Nowroozi, A.; Shahlaei, M. Molecular insight into the Grandivitin- matrix metalloproteinase 9 interactions. J. Photochem. Photobiol. B Biol. 2016, 162, 493–499. [Google Scholar] [CrossRef]
  47. Waziri, P.M.; Abdullah, R.; Rosli, R.; Omar, A.R.; Abdul, A.B.; Kassim, N.K.; Malami, I.; Etti, I.C.; Sani, J.M.; Lila, M.A.M.; et al. Clausenidin Induces Caspase 8-Dependent Apoptosis and Suppresses Production of VEGF in Liver Cancer Cells. Asian Pac. J. Cancer Prev. 2018, 19, 917–922. [Google Scholar] [CrossRef]
  48. Al-Abboodi, A.S.; Al-Sheikh, W.M.; Eid, E.E.M.; Azam, F.; Al-Qubaisi, M.S. Inclusion complex of clausenidin with hydroxypropyl-β-cyclodextrin: Improved physicochemical properties and anti-colon cancer activity. Saudi Pharm. J. 2021, 29, 223–235. [Google Scholar] [CrossRef]
  49. Barnard, D.L.; Xu, Z.-Q.; Stowell, V.D.; Yuan, H.; Smee, D.F.; Samy, R.; Sidwell, R.W.; Nielsen, M.K.; Sun, L.; Cao, H.; et al. Coumarins and pyranocoumarins, potential novel pharmacophores for inhibition of measles virus replication. Antivir. Chem. Chemother. 2002, 13, 39–59. [Google Scholar] [CrossRef]
  50. Taira, N.; Nugara, R.N.; Inafuku, M.; Takara, K.; Ogi, T.; Ichiba, T.; Iwasaki, H.; Okabe, T.; Oku, H. In vivo and in vitro anti-obesity activities of dihydropyranocoumarins derivatives from Peucedanum japonicum Thunb. J. Funct. Foods 2017, 29, 19–28. [Google Scholar] [CrossRef]
  51. Olennikov, D.N.; Fedorov, I.A.; Kaschenko, N.I.; Chirikova, N.K.; Vennos, C. Khellactone Derivatives and Other Phenolics of Phlojodicarpus sibiricus (Apiaceae): HPLC-DAD-ESI-QQQ-MS/MS and HPLC-UV Profile, and Antiobesity Potential of Dihydrosamidin. Molecules 2019, 24, 2286. [Google Scholar] [CrossRef] [Green Version]
  52. Hossin, A.Y.; Inafuku, M.; Oku, H. Dihydropyranocoumarins Exerted Anti-Obesity Activity In Vivo and its Activity Was Enhanced by Nanoparticulation with Polylactic-Co-Glycolic Acid. Nutrients 2019, 11, 3053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Rauwald, H.W.; Brehm, O.; Odenthal, K.P. The involvement of a Ca2+ channel blocking mode of action in the pharmacology of Ammi visnaga fruits. Planta Med. 1994, 60, 101–105. [Google Scholar] [CrossRef] [PubMed]
  54. Gulyaev, S.M. Zashhitnoe dejstvie Phlojodicarpus sibiricus pri ishemii golovnogo mozga u krys [Protective effect of Phlojodicarpus sibiricus on cerebral ischemia in rats]. Acta Biomed. Sci. 2009, 67, 172–174. [Google Scholar]
  55. Gulyaev, S.M. Protivotrevozhnoe dejstvie Phlojodicarpus sibiricus [Anti-anxiety effect of Phlojodicarpus sibiricus]. Acta Biomed. Sci. 2009, 67, 175. [Google Scholar]
  56. Urbanova, E.Z.; Guljaev, S.M.; Nikolaev, S.M.; Turtueva, T.A. Membranostabilizirujushhij jeffekt jekstrakta Phlojodicarpus sibiricus [Membrane stabilizing effect of Phlojodicarpus sibiricus extract]. Sib. Med. Zhurnal 2013, 123, 104–105. [Google Scholar]
  57. Urbanova, E.Z.; Guljaev, S.M.; Nikolaev, S.M.; Turtueva, T.A. Nejrofarmakologicheskie jeffekty Phlojodicarpus sibiricus (Steph. ex Spreng.) K.-Pol. [Neuropharmacological effects of Phlojodicarpus sibiricus (Steph. ex Spreng.) K.-Pol.]. Vestn. Buryatskogo Gos. Univ. 2013, 12, 125–128. [Google Scholar]
  58. Xiong, Y.; Wang, J.; Wu, F.; Li, J.; Zhou, L.; Jong, L. Effects of (±)-praeruptorin A on airway inflammation, airway hyperresponsiveness and NF-κB signaling pathway in a mouse model of allergic airway disease. Eur. J. Pharmacol. 2012, 683, 316–324. [Google Scholar] [CrossRef]
  59. Xia, H.; Li, C.; Yang, J.; Ma, J.; Li, Y.; Li, L.; Zhang, D. Hepatoprotective pyranocoumarins from the stems of Clausena emarginata. Phytochemistry 2016, 130, 238–243. [Google Scholar] [CrossRef]
  60. Widelski, J.; Luca, S.V.; Skiba, A.; Maciąg, M.; Budzyńska, B.; Marcourt, L.; Wolfender, J.-L.; Skalicka-Woźniak, K. Coumarins from Seseli devenyense Simonk.: Isolation by Liquid–Liquid Chromatography and Potential Anxiolytic Activity Using an In Vivo Zebrafish Larvae Model. Int. J. Mol. Sci. 2021, 22, 41829. [Google Scholar] [CrossRef]
  61. Onder, A.; Cinar, A.S.; Baran, M.Y.; Kuruüzüm-Uz, A.; Trendafilova, A. Coumarins from Seseli petraeum M. Bieb. (Apiaceae) and their α-glucosidase inhibitory activity. S. Afr. J. Bot. 2022, 144, 458–463. [Google Scholar] [CrossRef]
  62. Ferrero-Miliani, L.; Seiden, J.B.; Nielsen, O.H. Regulation of cytokine production in inflammatory bowel disease. Ugeskr. Laeger 2006, 168, 1487–1850. [Google Scholar]
  63. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  64. Morrison, P.W.J.; Connon, C.J.; Khutoryanskiy, V.V. Cyclodextrin-Mediated Enhancement of Riboflavin Solubility and Corneal Permeability. Mol. Pharm. 2013, 10, 756–762. [Google Scholar] [CrossRef]
  65. Hasani-Ranjbar, S.; Jouyandeh, Z.; Abdollahi, M. A systematic review of anti-obesity medicinal plants—An update. J. Diabetes Metab. Disord. 2013, 12, 28. [Google Scholar] [CrossRef] [Green Version]
  66. Ali, M.Y.; Jannat, S.; Jung, H.; Choi, J. Insulin-Mimetic Dihydroxanthyletin-Type Coumarins from Angelica decursiva with Protein Tyrosine Phosphatase 1B and α-Glucosidase Inhibitory Activities and Docking Studies of Their Molecular Mechanisms. Molecules 2021, 10, 292. [Google Scholar] [CrossRef]
  67. Morikawa, T.; Luo, F.; Manse, Y.; Sugita, H.; Saeki, S.; Chaipech, S.; Pongpiriyadacha, Y.; Muraoka, O.; Nonomiya, K. Geranylated Coumarins from Thai Medicinal Plant Mammea siamensis with Testosterone 5α-Reductase Inhibitory Activity. Front. Chem. 2020, 8, 199. [Google Scholar] [CrossRef] [Green Version]
  68. Khandy, M.T.; Kochkin, D.V.; Tomilova, S.V.; Galishev, B.A.; Klyushin, A.G.; Nosov, M.K. Obtainment and phytochemical screening of callus and suspension cell cultures of Phlojodicarpus sibiricus (Steph. Ex Spreng.) K.-Pol. Biotechnology 2020, 36, 54–61. [Google Scholar] [CrossRef]
  69. Khandy, M.T.; Kochkin, D.V.; Tomilova, S.V.; Nosov, A.M.; Klyushin, A.G.; Galishev, B.A. Growth and Biosynthetic Characteristics of Phlojodicarpus sibiricus Cell Suspension Cultures. Russ. J. Plant. Physiol. 2021, 68, 569–578. [Google Scholar] [CrossRef]
  70. Murray, R.D.H. The Naturally Occurring Coumarins. In Progress in the Chemistry of Organic Natural Products; Herz, W., Falk, H., Kirby, G.W., Moore, R.E., Eds.; Springer: Vienna, Austria, 2002; Volume 83, pp. 1–619. [Google Scholar] [CrossRef]
  71. Wei, S.-P.; Luan, J.-Y.; Lu, L.-N.; Wu, W.-J.; Ji, Z.-Q. A new benzofuran glucoside from Ficus tikoua Bur. Int. J. Mol. Sci. 2011, 12, 84946. [Google Scholar] [CrossRef] [Green Version]
  72. Sarker, S.D.; Nahar, L. Progress in the Chemistry of Naturally Occurring Coumarins. In Progress in the Chemistry of Organic Natural Products; Kinghorn, A.D., Falk, H., Gibbons, S., Kobayashi, J.I., Eds.; Springer: Berlin, Germany, 2017; Volume 106, pp. 1–241. [Google Scholar] [CrossRef]
  73. He, R.; Zhang, Y.; Wu, L.; Nie, H.; Huang, Y.; Liu, B.; Deng, S.; Yang, R.; Huang, S.; Nong, Z.; et al. Benzofuran glycosides and coumarins from the bark of Streblus indicus (Bur.) Corner. Phytochemistry 2017, 138, 170–177. [Google Scholar] [CrossRef]
  74. Mi, J.; Peng, Y.; Zhang, H.; Wang, X.; Huo, Y.; Wang, Z.; Liu, Y.; Gao, Y.; Zhang, H. A new benzofuran derivative glycoside and a new coumarin glycoside from roots of Heracleum dissectum Ledeb. Med. Chem. Res. 2018, 27, 470–475. [Google Scholar] [CrossRef]
  75. Matsuo, Y.; Yamaguchi, E.; Hakamata, R.; Ootomo, K.; Takatori, K.; Fukaya, H.; Mimaki, Y. Benzofuran and coumarin derivatives from the root of Angelica dahurica and their PPAR-γ ligand-binding activity. Phytochemistry 2020, 173, e112301. [Google Scholar] [CrossRef] [PubMed]
  76. Nosov, A.M.; Popova, E.V.; Kochkin, D.V. Isoprenoid Production via Plant Cell Cultures: Biosynthesis, Accumulation and Scaling-Up to Bioreactors. In Production of Biomass and Bioactive Compounds Using Bioreactor Technology; Paek, K.Y., Murthy, H., Zhong, J.J., Eds.; Springer: Berlin, Germany, 2014; pp. 563–623. [Google Scholar] [CrossRef]
  77. Khandy, M.T.; Kochkin, D.V.; Tomilova, S.V.; Galishev, B.A.; Sukhanova, E.S.; Klyushin, A.G.; Ivanov, I.M.; Nosov, A.M. Obtaining and study of callus and suspension plant cell cultures of Tribulus terrestris L., a producer of steroidal glycosides. Appl. Biochem. Microbiol. 2017, 53, 800–806. [Google Scholar] [CrossRef]
  78. Xu, H.; Kim, Y.; Suh, S.; Uddin, M.R.; Lee, S.; Park, S. Decursin Production from Hairy Root Culture of Angelica gigas. J. Korean Soc. App. Biolo. Chem. 2008, 51, 349–351. [Google Scholar] [CrossRef]
  79. Kaul, B.; Staba, E.J. Ammi visnaga (L.) Lam. tissue cultures. Planta Med. 1967, 15, 145–156. [Google Scholar] [CrossRef] [PubMed]
  80. El-Fiky, F.K.; Remmel, R.P.; Staba, E.J. Ammi visnaga: Somatic Embryo and Furanochromone Production in Embryos, Seedlings, and Plants. Planta Med. 1989, 55, 446–451. [Google Scholar] [CrossRef]
  81. Krolicka, A.; Staniszewska, I.; Malinski, E.; Lojkowska, E. Stimulation of furanochromone accumulation in callus cultures of Ammi visnaga L. by addition of elicitors. Pharm. Int. J. Pharm. Sci. 2019, 58, 590–592. [Google Scholar]
  82. Al-Saleh, M.M.; Shibli, R.A.; Al-Qadiri, H.M.; Tahtamouni, R.W.; Darwish, M.M.; Al-Qudah, T.S. Investigating the Antimicrobial Potential of in-vitro Grown Microshoots and Callus Cultures of Ammi visnaga (L.) Lam. Jordan J. Biol. Sci. 2019, 12, 43–48. [Google Scholar]
  83. Ekiert, H.; Gomolka, E. Coumarin compounds in Ammi majus L. callus cultures. Pharmazie 2000, 55, 684–687. [Google Scholar]
  84. Krolicka, A.; Staniszewska, I.; Bielawski, K.; Malinski, E.; Safranek, J.; Lojkowska, E. Establishment of hairy root cultures of Ammi majus. Plant Sci. 2001, 160, 259–264. [Google Scholar] [CrossRef]
  85. Krolicka, A.; Lojkowska, E.; Staniszewska, I.; Malinski, E.; Safranek, J. Identification of secondary metabolites in in vitro culture of Ammi majus treated with elicitors. Acta Hortic. 2001, 560, 255–258. [Google Scholar] [CrossRef]
  86. Staniszewska, I.; Krolicka, A.; Lojkowska, E.; Malinski, E.; Safranek, J. Elicitation od secondary metabolites in in vitro cultures of Ammi majus L. Enzym. Microb. Technol. 2003, 33, 565–568. [Google Scholar] [CrossRef]
  87. Krolicka, A.; Kartanowicz, R.; Wosinski, S.A.; Spitter, A.; Kaminski, M.; Lojkowska, E. Induction of secondary metabolite production in transformed callus of Ammi majus L. grown after electromagnetic treatment of the culture medium. Enzym. Microb. Technol. 2006, 39, 1386–1391. [Google Scholar] [CrossRef]
  88. Magdum, S.; Kumar, S. Organogenesis of Ammi majus via Callus Culture from leaf. Plant Tissue Cult. Biotechnol. 2013, 23, 101–106. [Google Scholar] [CrossRef] [Green Version]
  89. Eeva, M.; Ojala, T.; Tammela, P.; Galambosi, B.; Vuorela, H.; Hiltunen, R.; Fagerstedt, K.; Vuorela, P. Propagation of Angelica archangelica Plants in an Air-Spaged Bioreactor from a Novel Embryogenic Cell Line, and their Production of Coumarins. Biol. Plant. 2003, 46, 343–347. [Google Scholar] [CrossRef]
  90. Eeva, M. Plant secondary metabolites in Peucedanum palustre and Angelica archangelica and their plant cell cultures. Ph.D. Thesis, University of Helsinki, Helsinki, Findland, 2010; p. 77. [Google Scholar]
  91. Ahn, K.; Sim, W.; Kim, H.; Kim, I. Immunostimulating polysaccharide from cell culture of Angelica gigas Nakai. Biotechnol. Lett. 1998, 20, 5–7. [Google Scholar] [CrossRef]
  92. Lee, S.; Cho, W.; Lee, S.; Park, K.; Choi, K.; Hang, H. Effects of Plant Growth Regulators in In Vitro Germination and Organ Formation of Wild Angelica gigas N. Korean J. Plant Res. 2012, 25, 456–464. [Google Scholar] [CrossRef] [Green Version]
  93. Tsay, H.S.; Huang, H.L. Somatic embryo formation and germination from immature embryo-derived suspension-cultured cells of Angelica sinensis (Oliv.) Diels. Plant Cell Rep. 1998, 17, 670–674. [Google Scholar] [CrossRef]
  94. Chen, C.; Agrawal, D.C.; Lee, M.; Lee, R.; Kuo, C.; Wu, C.; Tsay, S. Influence of LED light spectra on in vitro somatic embryogenesis and LC-MC analysis of chlorogenic and rutin Peucedanum japonicum Thumb.: A medicinal herb. Bot. Stud. 2016, 57, 9. [Google Scholar] [CrossRef]
  95. Mitrofanova, I.V.; Mitrofanova, O.V.; Rudenko, M.I.; Lesnikova-Sedoshenko, N.P.; Kuzmina, T.N. In vitro morphogenesis in endangered plant Seseli lehmanii degen. In I International Symposium on Botanical Gardens and Landscapes 1298; Acta Horticulturae: Leuven, Belgium, 2020; pp. 257–264. [Google Scholar] [CrossRef]
  96. Hazeena, M.S.; Sulekha, G.R. Callus induction and plantlet regeneration in Aegle marmelos (L.) Corr. using cotyledon explants. J. Trop. Agric. 2008, 46, 79–84. [Google Scholar]
  97. Mishra, M.R.; Srivastava, R.K.; Akhtar, N. Enhanced Alkaloid Production from Cell Culture System of Catharanthus reseus in Combined effect of Nutrient Salts, Sucrose and Plant Growth Regulators. J. Biotechnol. Biomed. Sci. 2018, 1, 14–34. [Google Scholar] [CrossRef] [Green Version]
  98. Cappelletti, E.M.; Innocenti, G.; Caniato, R.; Filippini, R.; Povan, A. Haplophyllum patavinum (L.) G. Don fil. (Paduan rue): In Vitro Regeneration, and the Production of Coumarin Compounds. Biotechnol. Agric. Foresty 1998, 41, 238–260. [Google Scholar] [CrossRef]
  99. Hussain, M.A.; Nathar, V.N.; Mir, J.I. Gas chromatography-mass spectrometry (GC-MS) analysis in callus extracts of Ruta graveolens L. World J. Pharm. Res. 2017, 6, 1195–1210. [Google Scholar] [CrossRef]
  100. Rout, G.R.; Samantaray, S.; Das, P. In vitro rooting of Psoralea corylifilia Linn: Peroxidase activity as a marker. Plant Growth regul. 2000, 30, 215–219. [Google Scholar] [CrossRef]
  101. Nosov, A.M. Ispol’zovanie kletochnyh tehnologij dlja promyshlennogo poluchenija biologicheski aktivnyh veshhestv rastitel’nogo proishozhdenija [The use of cellular technologies for the industrial production of biologically active substances of plant origin]. Biotehnologija 2010, 5, 8–28. [Google Scholar]
  102. Paek, K.-Y.; Murthy, H.N.; Zhong, J.-J. (Eds.) Production of Biomass and Bioactive Compounds Using Bioreactor Technology; Springer: Dordrecht, The Netherlands, 2014; p. 709. [Google Scholar]
  103. Smanski, M.J.; Zhou, H.; Claesen, J.; Shen, B.; Fischbach, M.A.; Voigt, C.A. Synthetic biology to access and expand nature’s chemical diversity. Nat. Rev. Microbiol. 2016, 14, 135–149. [Google Scholar] [CrossRef] [Green Version]
  104. Lee, N.; Hwang, S.; Kim, W.; Lee, Y.; Kim, J.H.; Cho, S.; Kim, H.U.; Yoon, Y.J.; Oh, M.K.; Palsson, B.O.; et al. Systems and synthetic biology to elucidate secondary metabolite biosynthetic gene clusters encoded in Streptomyces genomes. Nat. Prod. Rep. 2021, 38, 1330–1361. [Google Scholar] [CrossRef]
  105. Leonard, E.; Ajikumar, P.K.; Thayer, K.; Xiao, W.H.; Mo, J.D.; Tidor, B.; Stephanopoulos, G.; Prather, K.L.J. Combining Metabolic and Protein Engineering of a Terpenoid Biosynthetic Pathway for Overproduction and Selectivity Control. Proc. Natl. Acad. Sci. USA 2010, 107, 13654–13659. [Google Scholar] [CrossRef] [Green Version]
  106. Ro, D.K.; Paradise, E.M.; Ouellet, M.; Fisher, K.J.; Newman, K.L.; Ndungu, J.M.; Ho, K.A.; Eachus, R.A.; Ham, T.S.; Kirby, J.; et al. Production of the antimalarial drug precursor artemisinic acid in engineered yeast. Nature 2006, 440, 940–943. [Google Scholar] [CrossRef]
  107. Park, D.; Yang, D.; Ha, D.; Lee, S. Biosynthesis of dihydroquercetin in Escherichia coli from glycerol. bioRxiv 2020. [Google Scholar] [CrossRef]
  108. Luo, X.Z.; Reiter, M.A.; d’Espaux, L.; Wong, J.; Denby, C.M.; Lechner, A.; Zhang, Y.F.; Grzybowski, A.T.; Harth, S.; Lin, W.Y.; et al. Complete Biosynthesis of Cannabinoids and Their Unnatural Analogues in yeast. Nature 2019, 567, 123–126. [Google Scholar] [CrossRef]
  109. Ajikumar, P.K.; Xiao, W.H.; Tyo, K.E.J.; Wang, Y.; Simeon, F.; Leonard, E.; Mucha, O.; Phon, T.H.; Pfeifer, B.; Stephanopoulos, G. Isoprenoid Pathway Optimization for Taxol Precursor Overproduction in Escherichia coli. Science 2010, 330, 70–74. [Google Scholar] [CrossRef] [Green Version]
  110. Li, Y.R.; Smolke, C.D. Engineering Biosynthesis of The Anticancer Alkaloid Noscapine in yeast. Nat. Commun. 2016, 7, 12137. [Google Scholar] [CrossRef] [Green Version]
  111. Galanie, S.; Thodey, K.; Trenchard, I.J.; Filsinger Interrante, M.; Smolke, C.D. Complete Biosynthesis of Opioids in yeast. Science 2015, 349, 1095–1100. [Google Scholar] [CrossRef] [Green Version]
  112. Bu, X.-L.; He, B.-B.; Weng, J.-Y.; Jiang, C.-C.; Zhao, Y.-L.; Li, S.-M.; Xu, J.; Xu, M.-J. Constructing Microbial Hosts for the Production of Benzoheterocyclic Derivatives. ACS Synth. Biol. 2020, 9, 2282–2290. [Google Scholar] [CrossRef]
  113. Malikov, V.M.; Saidkhodzhaev, A.I. Coumarins. Plants, structure, properties. Chem. Nat. Compd. 1998, 34, 345–409. [Google Scholar] [CrossRef]
  114. Sonnenberg, H.; Kaloga, M.; Eisenbach, N.; Froemming, K.K. Isolation and characterization of an angular-type dihydropyranocoumarine glycoside from the fruits of Ammi visnaga (L.) Lam. (Apiaceae). Z. Fur Nat. 1995, 50, 729–731. [Google Scholar] [CrossRef]
  115. Zgorka, G.; Dragan, T.; Glowniak, K.; Basiura, E. Determination of furanochromones and pyranocoumarins in drugs and Ammi visnaga fruits by combined solid-phase extraction-high-performance liquid chromatography and thin-layer chromatography-high-performance liquid chromatography. J. Chromatogr. A 1998, 797, 305–309. [Google Scholar] [CrossRef]
  116. Sellami, H.K.; Napolitano, A.; Masullo, M.; Smiti, S.; Piacente, S.; Pizza, C. Influence of growing conditions on metabolite profile of Ammi visnaga umbels with special reference to bioactive furanochromones and pyranocoumarins. Phytochemistry 2013, 95, 197–206. [Google Scholar] [CrossRef]
  117. Winderl, B.; Schwaiger, S.; Ganzera, M. Fast and improved separation of major coumarins in Ammi visnaga (L.) Lam. by supercritical fluid chromatography. J. Sep. Sci. 2016, 39, 4042–4048. [Google Scholar] [CrossRef]
  118. Ahmed, S.; Fahim, J.; Abdelmohsen, U. Chemical and biological potential of Ammi visnaga (L.) Lam. and Apium graveolens L.: A review (1963–2020). J. Adv. Biomed. Pharm. Sci. 2021, 4, 160–176. [Google Scholar] [CrossRef]
  119. Imanly, G.A.; Serkerov, S.V. Chemical components of roots and aerial parts of the Ammi visnaga (L.) lam. Khimija Rastit. Syr’ja 2017, 3, 237–240. [Google Scholar] [CrossRef] [Green Version]
  120. Avramenko, L.G.; Nikonov, G.K.; Pimenov, M.G. Andelin-A new dihydropyranocoumarin from the roots of Angelica decursiva. Chem. Nat. Compd. 1970, 6, 186–189. [Google Scholar] [CrossRef]
  121. Ali, M.Y.; Seong, S.H.; Jannat, S.; Jung, H.A.; Choi, J.S. Ethnobotany, Phytochemistry, and Pharmacology of Angelica decursiva Fr. et Sav. Nat. Prod. Sci. 2019, 25, 1135168. [Google Scholar] [CrossRef] [Green Version]
  122. Jung, D.J.; Porzel, A.; Huneck, S. Gigasol and other coumarins from Angelica gigas. Phytochemistry 1991, 30, 710–712. [Google Scholar] [CrossRef]
  123. Kang, S.Y.; Lee, K.Y.; Sung, S.H.; Park, M.J.; Kim, Y.C. Coumarins isolated from Angelica gigas inhibit acetylcholinesterase: Structure-activity relationships. J. Nat. Prod. 2001, 64, 683–685. [Google Scholar] [CrossRef]
  124. Park, Y.; Jeong, D.; Sim, S.; Kim, N.; Park, H.; Jeon, G. The Characteristics of Growth and Active Compounds of Angelica gigas Nakai Population in Mt. Jeombong. Korean J. Plant Res. 2019, 32, 9–18. [Google Scholar] [CrossRef]
  125. Geidarov, I.; Serkerov, S. Coumarins from roots of Angelica purpurascens. Chem. Nat. Compd. 2017, 53, 149–150. [Google Scholar] [CrossRef]
  126. Erdelmeier, C.A.J.; Sticher, O. Coumarin Derivatives from Eryngium campestre1. Planta Med. 1985, 51, 407–409. [Google Scholar] [CrossRef]
  127. Ahmadi, F.; Valadbeigi, S.; Sajjadi, S.E.; Shokoohinia, Y.; Azizian, H.; Taheripak, G. Grandivittin as a natural minor groove binder extracted from Ferulago macrocarpa to ct-DNA, experimental and in silico analysis. Chem. Biol. Interact. 2016, 258, 89–101. [Google Scholar] [CrossRef]
  128. Lee, J.W.; Lee, C.; Jin, Q.; Yeon, E.T.; Lee, D.; Kim, S.-Y.; Han, S.B.; Hong, J.T.; Lee, M.K.; Hwang, B.Y. Pyranocoumarins from Glehnia littoralis inhibit the LPS-induced NO production in macrophage RAW 264.7 cells. Bioorganic Med. Chem. Lett. 2014, 24, 2717–2719. [Google Scholar] [CrossRef]
  129. Swagen, T.M.; Cardellina, J.H. Coumarins from Musineon divaricatum☆. Phytochemistry 1985, 24, 805–813. [Google Scholar] [CrossRef]
  130. Sarkhail, P. Traditional uses, phytochemistry and pharmacological properties of the genus Peucedanum: A review. J. Ethnopharmacol. 2014, 156, 235–270. [Google Scholar] [CrossRef]
  131. Liu, R.; Sun, Q.; Shi, Y.; Kong, L. Isolation and purification of coumarin compounds from the root of Peucedanum decursivum (Miq.) Maxim by high-speed counter-current chromatography. J. Chromatogr. A 2005, 1076, 127–132. [Google Scholar] [CrossRef]
  132. Duh, C.-Y.; Wang, S.-K.; Wu, Y.-C. Cytotoxic pyranocoumarins from the aerial parts of Peucedanum japonicum. Phytochemistry 1991, 30, 2812–2814. [Google Scholar] [CrossRef]
  133. Ikeshiro, Y.; Mase, I.; Tomita, Y. Dihydropyranocoumarins from roots of Peucedanum japonicum. Phytochemistry 1992, 31, 4303–4306. [Google Scholar] [CrossRef]
  134. Ikeshiro, Y.; Mase, I.; Tomita, Y. Dihydropyranocoumarins from Peucedanum japonicum. Phytochemistry 1993, 33, 1543–1545. [Google Scholar] [CrossRef]
  135. Lu, M.; Nicoletti, M.; Battinelli, L.; Mazzanti, G. Isolation of praeruptorins A and B from Peucedanum praeruptorum Dunn. and their general pharmacological evaluation in comparison with extracts of the drug. Il Farmaco 2001, 56, 417–420. [Google Scholar] [CrossRef]
  136. Song, Y.; Jing, W.; Yan, R.; Wang, Y. Research progress of the studies on the roots of Peucedanum praeruptorum Dunn (Peucedani radix). Pak. J. Pharm. Sci. 2015, 28, 71–81. [Google Scholar]
  137. Kong, L.-Y.; Wu, X.; Niwa, M. Two Dihydropyranocoumarins from Peucedanum wawrii. Heterocycles 2003, 60, 599–606. [Google Scholar] [CrossRef]
  138. Babilev, F.V.; Nikonov, G.K. Coumarins of the roots of Phloiodicarpus villosus Turcz. Chem. Nat. Compd. 1965, 1, 383–386. [Google Scholar] [CrossRef]
  139. Olennikov, D.N.; Chirikova, N.K. New compounds from flowers of Phlojodicarpus sibiricus. Chem. Nat. Compd. 2020, 56, 628–632. [Google Scholar] [CrossRef]
  140. Valencia-Islas, N.; Abbas, H.; Bye, R.; Toscano, R.; Mata, R. Phytotoxic Compounds from Prionosciadium watsoni. Nat. Prod. 2002, 65, 828–834. [Google Scholar] [CrossRef] [PubMed]
  141. Shagova, M.E.; Florya, V.N.; Kuznetsova, G.A.; Perel’son, M.E. Diesters of khellactone and rutarin (campesenin) from Seseli campestre growing in Moldavia. Chem. Nat. Compd. 1973, 9, 665. [Google Scholar] [CrossRef]
  142. Abyshev, A.Z.; Sidorova, I.P.; Abyshev, D.Z.; Florya, V.I.; Zmeikov, V.P.; Kerimov Yu, B. Comparative characterization of the coumarin composition ofSeseli campestregrowing in Moldavia and in the Caucasus. Chem. Nat. Compd. 1982, 18, 400–407. [Google Scholar] [CrossRef]
  143. Malikov, V.M.; Saidkhodzhaev, A.I. Coumarins: Plants, structures, properties. Chem. Nat. Compd. 1998, 34, 517–548. [Google Scholar] [CrossRef]
  144. Dukhovlinova, L.I.; Sklyar, Y.E.; Pimenov, M.G. Coumarins of the roots of Seseli coronatum. Chem. Nat. Compd. 1974, 10, 802–803. [Google Scholar] [CrossRef]
  145. Bagirov, V.; Yu; Belyi, M.B. An investigation of Seseli peucedanoides. Chem. Nat. Compd. 1981, 17, 585–586. [Google Scholar] [CrossRef]
  146. Razavi, S.M.; Imanzadeh, G.; Jahed, F.S.; Zarrini, G. Pyranocoumarins from Zosima absinthifolia (Vent) link roots. Russ. J. Bioorg. Chem. 2013, 39, 215–217. [Google Scholar] [CrossRef]
  147. Sklyar, Y.E.; Avramenko, L.G.; Pimenov, M.G.; Avetisyan, R.N. Coumarins of Zosima korovinii. Chem. Nat. Compd. 1982, 18, 744. [Google Scholar] [CrossRef] [Green Version]
  148. Gray, A.I.; Waterman, P.G. Coumarins in the Rutaceae. Phytochemistry 1978, 17, 845–864. [Google Scholar] [CrossRef]
  149. Panthong, K.; Srisud, Y.; Rukachaisirikul, V.; Hutadilok-Towatana, N.; Voravuthikunchai, S.P.; Tewtrakul, S. Benzene, coumarin and quinolinone derivatives from roots of Citrus hystrix. Phytochemistry 2013, 88, 79–84. [Google Scholar] [CrossRef]
  150. Ribeiro, A.B.; Abdelnur, P.V.; Garcia, C.F.; Belini, A.; Severino, V.G.P.; da Silva, M.F.G.F.; Fernandes, J.B.; Vieira, P.C.; de Souza, A.A.; Machado, M.A. Chemical Characterization of Citrus sinensis Grafted on C. limonia and the Effect of Some Isolated Compounds on the Growth of Xylella fastidiosa. J. Argic. Food Chem. 2008, 56, 7815–7822. [Google Scholar] [CrossRef]
  151. Waziri, P.M.; Abdullah, R.; Yeap, S.K.; Omar, A.R.; Abdul, A.B.; Kassim, N.K.; Malami, I.; Karunakaran, T.; Imam, M.U. Clausenidin from Clausena excavata induces apoptosis in hepG2 cells via the mitochondrial pathway. J. Ethnopharmacol. 2016, 194, 549–558. [Google Scholar] [CrossRef]
  152. Joshi, B.S.; Kamat, V.N.; Saksena, A.K. Structures of clausenin and clausenidin two new pyranocoumarins from the roots of clausena heptaphylla Wt. & Arn. Tetrahedron 1967, 23, 4785–4789. [Google Scholar] [CrossRef]
  153. Mohammadhosseini, M.; Venditti, A.; Frezza, C.; Serafini, M.; Bianco, A.; Mahdavi, B. The Genus Haplophyllum Juss.: Phytochemistry and Bioactivities—A Review. Molecules 2021, 26, 4664. [Google Scholar] [CrossRef]
  154. Ulubelen, A.; Mericli, A.H.; Mericli, F.; Sonmez, U.; Ilarslan, R. Alkaloids and Coumarins from Haplophyllum Thesioides. Nat. Prod. Lett. 1993, 1, 269–272. [Google Scholar] [CrossRef]
  155. Nikonov, G.K.; Manaeva, Z.A.; Pek, G.Y. Ostruthol from Xanthogallum purpurascens. Khimiya Prir. Soedin. 1966, 2, 360–361. [Google Scholar] [CrossRef]
  156. Chang, Y.; Bai, M.; He, H.; Huang, S.-W.; Song, S.-J. Phytochemical investigation on the fruits of Camptotheca acuminata and their chemotaxonomic significance. Biochem. Syst. Ecol. 2020, 93, 104169. [Google Scholar] [CrossRef]
  157. Erst, A.S.; Chernonosov, A.A.; Petrova, N.V.; Kulikovskiy, M.S.; Maltseva, S.Y.; Wang, W.; Kostikova, V.A. Investigation of Chemical Constituents of Eranthis longistipitata (Ranunculaceae): Coumarins and Furochromones. Int. J. Mol. Sci. 2022, 23, 10406. [Google Scholar] [CrossRef]
Figure 1. Natural groups of pyranocoumarins: (a) linear pyranocoumarin; (b) 5,6-angular pyranocoumarin; (c) 6,7-angular pyranocoumarin; (d) condensed pyranocoumarin.
Figure 1. Natural groups of pyranocoumarins: (a) linear pyranocoumarin; (b) 5,6-angular pyranocoumarin; (c) 6,7-angular pyranocoumarin; (d) condensed pyranocoumarin.
Plants 11 03135 g001
Figure 2. Biosynthesis pathway of umbelliferone, a precursor of pyranocoumarins.
Figure 2. Biosynthesis pathway of umbelliferone, a precursor of pyranocoumarins.
Plants 11 03135 g002
Figure 3. Biosynthesis of pyranocoumarins.
Figure 3. Biosynthesis of pyranocoumarins.
Plants 11 03135 g003
Figure 4. Distribution of linear and angular pyranocoumarins within the Umbelliferae and Rutaceae families. Mix-Ad and L pyranocoumarins; Ad-angular 7,8-pyranocoumarin; L-linear 6,7-pyranocoumarin.
Figure 4. Distribution of linear and angular pyranocoumarins within the Umbelliferae and Rutaceae families. Mix-Ad and L pyranocoumarins; Ad-angular 7,8-pyranocoumarin; L-linear 6,7-pyranocoumarin.
Plants 11 03135 g004
Table 1. Biological activities of pyranocoumarins.
Table 1. Biological activities of pyranocoumarins.
ActivityPure Substance/ExtractPyranocoumarins FormModelMechanism of ActionReferences
The anti-inflammatory activityDecursin from Angelica gigasL *In vitro in macrophagesDecursin suppressed the expression of proteins matrix metalloproteinase 9, monocytic chemoattractant protein 1, interleukin 8, tumor necrosis factor α and IL-1β due to the fact that it inhibited the activity of the NF-kB translocation signaling pathway.[23]
Praeruptorin A from Peucedanum praeruptorumAd *In vitro in murine macrophage of RAW264.7 cell linePyranocoumarin suppressed the cytoplasmic loss of inhibitor κB-α protein and inhibited the translocation of NF-κB from cytoplasm to nucleus.[24]
(+)-Praeruptorin A from P. praeruptorum rootsAdIn vitro in murine macrophage of RAW264.7 cell lineThe presence of (+)-praeruptorin A in medium reduced nitric oxide production in cells.[25]
Corymbocoumarin from Seseli gummiferum subsp. corymbosumAdIn vitro in murine macrophage of RAW264.7 cell lineThe presence of corimbocoumarin in medium suppressed the nuclear factor κB (NF-κB) signaling pathway activation and heme oxygenase (HO)-1 expression induction in cells.[26]
10-(3,7-Dimethylocta-1,6-dien-3-yl)-5-methoxy-8,8-
dimethylpyranocoumarin, nordentain from Clausena emarginata
LIn vitro in murine microglial of BV2 cell lineThe presence of pyranocoumarin in medium reduced the nitric oxide production in cells.[27]
(+)-Praeruptorin A, visnadin from Ligusticum lucidum subsp. cuneifoliumAdIn vivo in mice(+)-Praeruptorin A reduced local edema by 22%, visnadin, by 43%. The anti-inflammatory activity of these compounds appears to be modulated by the substituents on their aromatic ring.[28]
Artificially synthesized 3,4-angular pyranocoumarinsA *Against proteinase enzymesProteinase activity inhibition. The best effect was shown by pyranocoumarins with amino groups in the pyran ring[29]
In vivo in ratsRat paw edema decreased to 29.2%.
Against COX-1 and COX-2Most of the pyranocoumarin derivatives showed themselves against COX-2 (IC50 ≤ 10 µM).
Antioxidant activityGrandivittin,
agasyllin,
aegelinol from Ferulago campestris
LIn vitro in human polymorphonuclear neutrophils (PMN) respiratory burst cells and on human whole blood leukocytes (WB)It is suggested that pyranocoumarins scavenge reactive oxygen species or interfere with cellular activation mechanisms.[30]
Artificially synthesized 3′,4′-Di-O-acetyl-cis-khellactone (DOAcK)AdIn vivo in ratsDOAcK administration increased the activities of Catalase, Glutathione Peroxidase and Super Oxide Dismutase.[31]
Antimicrobial activityDecursin and decursinol angelate from Angelica gigasLIn vitro in Bacillus subtilisThe six-membered ring and senecioylic acid type side chain closely related to the enhanced antibacterial activities of coumarins against B. subtilis. Although a precise mechanism has not yet been clarified, these phenomena might be considered to occur due to the differences in the binding affinities of both compounds on the active sites of the enzymes or receptors from the differences in the position of the side chain moiety.[32]
Agasyllin,
aegelinol from Ferulago campestris
LIn vitro in bacteriaAt a concentration of 16 to 125 µg/mL, the growth of Staphylococcus aureus, Salmonella thypii, Enterobacter cloacae and Enterobacter earogenes was suppressed, from 5 to 25 µg/mL—Helicobacter pylori.[33]
Pd-D-V and Disenecioyl Khellactone from P. decursivumLIn vitro against S. sclerotiorum, B. cinerea, F. graminearum and C. capsiciPd-DV at a concentration of 30 μg/mL in medium inhibited the growth of S. sclerotiorum by 86.4 ± 0.7, and also inhibited the growth of B. cinerea, F. graminearum and C. capsici by almost 50%.[34]
AdIn vitro against S. sclerotiorum, T. cucumeris, B. cinerea and F. graminearumDisenecioyl Khellactone at a concentration of 30 μg/mL inhibited the growth of S. sclerotiorum and B. cinerea by almost 50%, and the growth of F. framinearum, T. cucumeris and C. capsica by almost 30%.
Agasyllin,
aegelinol from Ferulago campestris
LMagnaporthe oryzae in vitro and in vivoPyranocoumarins inhibited M. oryzae spore germination and upressoria at concentrations of 50 and 200 µg/mL and prevented rice disease by more than 80% at 100 and 300 µg/mL.[22]
Phenyl derivative of pyranocoumarin (PDP) from Psoralea corylifolia L.LIn vitro in Fusarium oxysporum, F. graminearum, and F. moniliformeThe ligand
PDP showed bifurcated hydrogen bond interaction with active site residues at TYR 413 and a single hydrogen bond interaction at ARG 402 with a docking score −7.19 and glide energy of −45.78 kcal/mol. This indicated a strong binding of the ligand with the trichothecene 3-O-acetyltransferase, preventing as a result the acetylation of the trichothecene mycotoxin and destruction of the “self-defense mechanism” of the Fusarium sp.
[21]
Artificially synthesized 3,4-angular pyranocoumarinsAIn vitro in bacteria and fungiMost pyranocoumarins exhibited inhibitory activity in zones with diameters in the range of 15–19 mm or more. It is assumed that activity is associated with lipophilicity of molecules.[29]
Anticancer activityDecursin and decursinol from A. gigasLIn vivo in miceTumor size decreased by 40.6% and 45.6% with introduction of decursin and decursinol angelate[35]
In vitro in prostate cancer DU145 and LNCaP cellsDecursin inhibited cell growth by stopping the G1, G2 and S phases at different dosages (25–100 μM)[36]
In vitro in LNCaP cellsDecursin inhibits androgen-stimulated nuclear translocation of the androgen receptor and reduces the amount of androgen receptor protein.[37]
In vitro in MCF-7 cellsDecursin and decursinol angelate exerted growth inhibitory effects on MCF-7 cells through G1 arrest and caspase-mediated apoptosis.[38]
In vitro in prostate cancer PC-3 cellsDecursin suppressed cell proliferation by suppressing Wnt/β-catenin pathway.[39]
In vitro in RC cellsDecursin inhibits cell proliferation by inducing apoptosis, which is mediated by both caspase-dependent and caspase-independent apoptosis pathways.[40]
In vitro in U266, MM.1S, ARH77 cellsDecursin showed a synergistic effect with bortezomib due to inhibition of STAT3 (activation of signal transducers and the activator of transcription 3)-induced proliferative and angiogenic effect in multiple myeloma.[41]
In vitro in HeLa cellsExtract increases expression of TRAIL, which stimulates extrinsic and intrinsic pathways of apoptosis through the activation of caspase-8 and caspase-9, respectively.[42]
(+)-Decursinol from Saposhnikovia divaricataLIn vitro in MEL-8, U-937, DU-145, MDA-MB-231 and BT-474 cell linesIt inhibited growth and proliferation of DU-154 prostate cancer cells and MEL-8 melanoma cells. The mechanism has not been considered.[43]
Clausarin from Clausena excavataLIn vitro in the multi-drug resistant cell line KB-VIN cellsThe mechanism(s) of action should be further investigated.[44]
Artificially synthesized linear pyranocoumarinsL
(±)-Praeruptorin A & (±)-Praeruptorin B from Peucedanum praeruptorumAdIn vitro in SGC7901 cells(±)-Praeruptorin A increases cancer cells drug sensitivity, presumably by inhibiting the expression of P-glycoprotein, which is responsible for the efflux of drugs from cancer cells through ATP-dependent pumps.[45]
Angular pyranocoumarins from Peucedanum praeruptorum rootsAdIn vitro in MES-SA/Dx5Pyranocoumarins inhibited drug efflux via MDR protein, which prevents multidrug-resistant cancer formation.[25]
Grandivitin from Ferulago macroparaLMolecular modeling analysis against MMP9By binding to Matrix metalloproteinase 9 (MMP9), it affected secondary structure and modified tertiary structure of this protein.[46]
Clausenidin from Clausena excavataLIn vitro in HepG2 cellsClausenidin increased the activity of caspase-8 and expression of protein components of the death inducing signaling complex (DISC) in HepG2 cells.[47]
LIn vitro in HT29 and CCD-18CoHT29 colorectal adenocarcinoma cells treated with clausenidin/hydroxypropyl-β-cyclodextrin complex showed cell cycle arrest and death by apoptosis associated with caspase activation.[48]
Antiviral activityCausenidin and nordentatin from C. excavataLIn vitro in HepA2 cellsThe pyranocoumarins suppressed hepatitis.[44]
Artificially synthesized linear pyranocoumarinsLB virus surface antigen in HepA2 cells and had anti-HBV values. Their EC50 values were 1.14, 1.34, 1.64 and 1.63 µM.
Artificially synthesized 5,6-angular pyranocoumarinsAAfrican green monkey (CV-1) kidney cell lines infected with measles virusThe compounds inhibited nine strains of measles virus, and in virucidal tests, drugs did not physically destroy virion to inhibit virus replication.[49]
Antihyperglycemic and antidyslipidemic activityPyranocoumarins isolated from P. japonicumAdIn vitro on 3T3-L1 adipocytesInhibited lipid accumulation and lipogenic gene expression in 3T3-L1 adipocytes.[50]
Root and herb extracts of P. sibiricus, separately isolated fractions of dihydrosamidine, kelactone estersAdIn vitro in 3T3-L1 cellsRoot extracts and esters of kellacton inhibited triacylglycerol accumulation, while activity depended on acyl groups type.[51]
Dihydropyranocoumarins isolated from P. japonicumAdIn vivo in miceConsumption of dihydropyranocumarins significantly reduced average size of adipocytes and increased mRNA levels of genes associated with thermogenesis. Nanoparticulation of DPCs with polylactic-co-glycolic acid (PLGA) dramatically increased its activity almost 100-fold over that of a non-nanoparticulated form.[52]
(+)-trans-decursidinol, Pd-C-I and Pd-C-II, Pd-C-III isolated from Angelica decursivaLIn vitro in HepG2 cellsPyranocoumarins increased glucose uptake and decreased PTP1B expression in insulin resistant HepG2 cells; inhibited albumin nitration mediated by ONOO-, removed peroxynitrite, ROS.[30]
Other activities/effectsVisnadin from Ammi visnagaAdIn vitro in aortic segments of guinea-pigsVisnadin predominantly inhibits contractile responses mediated by penetration of Ca2+ through Ca2+.[53]
Tincture from Phlojodicarpus sibiricusAdIn vivo in ratsTincture showed anxiolytic properties, presumably due to GABAA benzodiazepine receptors activation.[54]
Tincture from P. sibiricusAdIn vivo in ratsThe extract limits the development of emotional and neurological disorders in cerebral ischemia in rats.[55]
Extract of P. sibiricus-In vitro in erythrocitesExtract inhibited hemolysis of erythrocytes, which is presumably due to the membrane-stabilizing properties of coumarins.[56]
Extract of P. sibiricus-In vivo in ratsExtract has a neuromodulatory effect, probably due to activation of GABA (gamma aminobutyric acid)-ergic system by coumarin compounds.[57]
(±)-Praeruptorin from P. praeruptorumAdIn vivo in mice(±)-Praeruptorin A suppressed airway inflammation, airway hyperreactivity and remodeling, reduced serum IL-4 and IL-13 levels in BALF and IgE, inhibited TGF-β1 and pSmad2/3 expression, increased Smad7 expression in lung tissue, and also increased INF-γ levels in BALF.[58]
Clauemarmarin C, D & 5-hydroxy-8,8-dimethyl-10-
(7-hydroxy-3,7-dimethylocta-1,5-dien-3-yl) pyranocoumarin from C. emarginata
LIn vitro in WB-F344 cellsThey showed a hepatoprotective effect, the mechanism was not considered.[59]
Mammeasins E and F from Mammea siamensisC *Against enzyme testosterone 5α-reductaseInhibited testosterone 5α-reductase.[21]
cis-Khellactone, d-laserpitin, isolaserpitin and octanoyllomatin, isolated from Seseli devenyenseAdIn vivo on zebrafish larvaeCoumarins reduced anxiety behavior (anxiolytic activity) in zebrafish larvae. The mechanism was not considered.[60]
Angular pyranocoumarins from Seseli petraeumAdAgainst enzymes α-amylase and α-glucosidaseThe coumarins exhibited notable inhibitory activity against the α-glucosidase enzyme and low inhibitory potential against α-amylase.[61]
* L—linear 6,7-pyranocoumarin; Ad—angular 7,8-pyranocoumarin; A—angular 5,6- or 7,8-pyranocoumarins; C—condensed 4,5-pyranocoumarin.
Table 2. Cell cultures of pyranocoumarin producers.
Table 2. Cell cultures of pyranocoumarin producers.
SpeciesExplantGrowth Conditions, MediaResultsPyranocoumarinsReferences
Umbelliferae (Apiaceae)
Ammi visnaga (L.) LamFruitsMS * with 1.0 mg/L 2,4-D *Callus tissue contained 45 mg visnagin/1100 g dry weight.Not detected[79]
Fruits
1.
MS with 1.0 mg/L 2,4-D
2.
MS with 10 μM L-glutamine and 1.0 mg/L 2,4-D, 0.5 mg/L BA *, or a growth regulator-free MS.
1.
Globular masses.
2.
The globular, heart and torpedo embryos of A. visnaga produced khellin and visnagin.
Not detected[80]
Hypocotyl of sterile seedlingsMS with 2.5 mg/L NAA * and 1.0 mg/L BACallus tissue with visnagin and/or khellin.Not detected[81]
SeedsMS with 1.0 mg/L BA and
2.0 mg/L 2,4-D
Microshoots in vitro and callus culture.Not detected[82]
Ammi majus L.Not specifiedLinsmaier-Skoog’s media with NAA and BATissue and organ cultures containing different concentrations of the linear furanocoumarins psoralen, bergapten, xanthotoxin, isopimpinellin, imperatorin and their precursor umbelliferone.Not detected[83]
PlantletsMS with 2.0 mg/L NAA, 2.0 mg/L BACallus, suspension and hairy roots of A. majus containing umbelliferone were obtained.Not detected[84]
Plantlets [84]Not specifiedThe effect of various elicitors on the synthesis of coumarins (umbelliferone, marmesinin, scopoletin) in hairy roots of A. majus was shown.Not detected[85]
Plantlets [84]MS with various phytohormonsThe effect of various elicitors on the synthesis of secondary metabolites (scopoletin, dehydrogeijerin) of in callus, cell suspension and hairy roots of A. majus by exposing them to elicitors.Not detected[86]
Hairy root cultureMS with 2.5 mg/L NAA, 1.0 mg/L BAElicitation with ADR-4® (electromagnetic treatment) induced also two times higher accumulation of bergapten.Not detected[87]
LeafsMS with 2 mg/L IAA *, 2 mg/L Kin *Callus of A. majus were obtained. About 81% of calluses converted to shoot on medium with 50 mg/L glutamine and 40 mg/L adenine. Plantlets with shoot were transferred to 1/2 MS with different concentration of IBA and glutamine. Maximum rooting of 75.96% was observed on 1/2 MS supplemented with 0.2% of IBA and 100 mg/L glutamine.Not detected[88]
Angelica archangelica (L.) subsp. archangelicaEmbryogenic cell line from seedsHormone-free, modified B5 *Embryogenic cell line of A. archangelica was obtained, after 5 years it did not decrease capacity of embryo formation.Not detected[89]
Plantlets from embryogenic cell line-14 coumarins were identified in the roots of A. archangelica, oxypeucedanin hydrate and oxypeucedanin being the main compounds.Not detected[90]
Angelica gigas NakaiYoung plantsSchenk and Hildebrandt media with 2.0 mg/L 2,4-D, 1.0 mg/L KinAn immunostimulating polysaccharide was produced extracellularly by suspension cell culture of A. gigas.Not detected[91]
Leaves and stemsHormone-free MSHairy roots were induced from leaf and stem explants of A. gigas. Biomass growth and decursin production was faster in hairy roots than in wild type of A. gigas.Decursin[78]
Stems, roots and hypocotyls In vitroMS with NAA and 2,4-D, BA, GA * and TDZ *The highest callus induction rate was obtained from in vitro germinated stem, root and hypocotyl on the MS medium with 1.0 mg/L NAA and 0.5 mg/L BA.Not detected[92]
Angelica sinensis (Oliv.) Diets.Immature embryosMS, B5, WhiteStudy showed that embryogenic callus growth was more rapid on MS basal medium than on B5 or White medium. Suspension culture and somatic embryos were obtained from this callus.Not detected[93]
Peucedanum japonicum Thunb.Root, leaf blade and
petiole parts of seedlings
In vitro
MS with 0.1–5.0 mg/L 2,4-D and 0.1–5.0 mg/L ABA *Embryogenesis was induced in media MS with ABA from callus subculture (MS with 2,4-D). Four-month-old tissue culture plants derived from somatic embryos showed significantly of chlorogenic acid (10.5 mg/g dw).Not detected[94]
Phlojodicarpus sibiricus (Steph. e x Spreng.) K.-Pol.Seeds, leaves, roots and hypocotyls In vitroMS with c 1.0 mg/L 2,4-D and 0.5 mg/L BACallus was obtained from seeds, leaves, roots and hypocotyls in vitro. Dihydrosamidin, visnadin, khellactone derivatives were identified in callus cultures.Dihydrosamidin, visnadin, khellactone derivatives[68]
CallusMS with c 1.0 mg/L 2,4-D and 0.5 mg/L BASuspension culture was obtained from callus.Not detected[69]
Seseli lehmannii DegenCallus and somatic embryosMS with 1.0–2.0 mg/L BA and 0.5–2.0 mg/L NAAOrganogenesis of callus was induced in MS media with 1.0–2.0 mg/L BAP and 0.5–2.0 mg/L NAA.Not detected[95]
Rutaceae
Aegle marmelos L.LeavesMS with 2.26 μM 2,4-D and 2.2 μM benzyl adenineCallus was obtained from leaves on MS with 2,4-D and BA. Derived shoots from callus were rooted in vitro on MS medium supplemented with 12.3 μM indole-3-butyric acid.Not detected[96]
Catharanthus roseus Linn.Leaves, roots, axillary buds, shoot tipsMS and B5 with various concentrations of 2,4-D, Kin, NAACallus culture was obtained. There was no significant difference in growing parameters between cultures in both types of media formulations on agar (MS or B5 salts). However, the alkaloid content was 2–3 times higher in suspension culture compared to agar medium in similar treatments.Not detected[97]
Haplophyllum patavinum (L.) G. Don fil.PlantletsB5 with 3 mg/L IAA and MS with 3 mg/L IAAH. patavinum is not rich in coumarin compounds in vivo, but a selected cell strain exhibited in vitro coumarin biogenetic potentialities stronger than in vivo.Not detected[98]
Ruta graveolens L.Whole plantMS with 1.5 mg/L 2,4-D and 1.5 mg/L NAAMS with 2, 4-D (1.5 mg/L) and NAA (1.5 mg/L) responded well by giving the maximum percentage of callus induction (97.22 ± 2.54). Coumarins were detected by GC-MS in biomass.Not detected[99]
Fabaceae
Psoralea corylifolia L. (Buguchi)Leaves and stemsMS with IAARooting was induced in microshoots of P. corylifolia. Peroxidase activity increased considerably during root induction indicating a key role of peroxidase in rooting of P. corylifolia microshoots in vitro.Not detected[100]
* MS—Murashige-Skoog Medium; 2,4-D—2,4-dichlorophenoxyacetic acid; BA—6-benzylaminopurine; NAA—1-naphthaleneacetic acid; IAA—indole-3-acetic acid; Kin—kinetin; B5—Gamborg B5 medium; GA—gibberellic acid; TDZ—thidiazuron; ABA—abscisic acid.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khandy, M.T.; Sofronova, A.K.; Gorpenchenko, T.Y.; Chirikova, N.K. Plant Pyranocoumarins: Description, Biosynthesis, Application. Plants 2022, 11, 3135. https://doi.org/10.3390/plants11223135

AMA Style

Khandy MT, Sofronova AK, Gorpenchenko TY, Chirikova NK. Plant Pyranocoumarins: Description, Biosynthesis, Application. Plants. 2022; 11(22):3135. https://doi.org/10.3390/plants11223135

Chicago/Turabian Style

Khandy, Maria T., Anastasia K. Sofronova, Tatiana Y. Gorpenchenko, and Nadezhda K. Chirikova. 2022. "Plant Pyranocoumarins: Description, Biosynthesis, Application" Plants 11, no. 22: 3135. https://doi.org/10.3390/plants11223135

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