Perilla sp. / Zwarte netel

Nom scientifique : Perilla frutescens

Noms communs : shiso, shiso vert, périlla, pérille, sésame sauvage, tiso japonais, herbe serpent à sonnettes

Noms anglais : wild red basil , purple mint , beefsteak plant , Chinese basil , rattlesnake weed , summer coleus

Classification botanique : famille des lamiacées ( Lamiaceae )

Formes et préparations : feuilles fraîches, poudres, gélules, huile essentielle, tisanes

Propriétés médicinales du shiso

Utilisation interne

Antiallergique ; antiasthmatique ; antibactérien ; antiseptique ; antispasmodique ; carminatif; expectorant; stomachique ; tonique ; minéralisant.

Utilisation externe

Réparateur ; apaisant ; revitalisant.

Indications thérapeutiques usuelles

Le shiso est utilisé pour lutter contre de nombreuses allergies. Il permet de soigner les personnes atteintes de troubles du foie. Il diminue l'agressivité, la violence, le stress, l'anxiété. Le shiso est efficace contre les rhumes, l'asthme, la toux et a un effet préventif contre la grippe.

Autres indications thérapeutiques démontrées

Le shiso agit contre les nausées, les vomissements, les douleurs abdominales et la constipation.

Principes actifs

Polyphénols ; flavonoïdes ; lutéoline ; acide rosmarinique ; chrysoériol ; apigénine ; huile essentielle ; limonène ; linalol ; L-menthol ; alpha-pinène ; élémicine.

Utilisation et posologie du shiso

Dosage

Il n'y a pas de dosage précis pour la consommation du shiso, en dehors des préparations standardisées.

En infusion : remplir 1/4 de tasse avec des feuilles séchées en poudre, recouvrir d'eau bouillante et laisser infuser de dix à quinze minutes. Boire cette préparation tout au long de la journée pour lutter contre la grippe ou les maux de gorge.

Pour diminuer les symptômes de l' asthme : utiliser l'huile de graines de shiso durant quatre semaines.

En inhalation, pour dégager les sinus : faire bouillir des feuilles de shiso et inhaler la vapeur.

Contre-indications

Le shiso est contre-indiqué aux femmes enceintes et à celles qui allaitent. L'huile de shiso est proscrite aux personnes atteintes d'un cancer.

Interactions avec des plantes médicinales ou des compléments

L'effet sédatif du shiso peut s'ajouter à celui d'autres plantes ou compléments.

Interactions avec des médicaments

L'effet sédatif du shiso peut s'ajouter à celui des médicaments. Il ne doit pas être consommé en même temps que des anti-inflammatoires non stéroïdiens ou des anticholestérolémiants.

Perilla frutescens (L.) Britt. (PF)

is an annual herbal medicinal, aromatic, functional food, and ornamental plant that belongs to the mint family, Lamiaceae. The origin of perilla traces back to East Asian countries (China, Japan, Korea, Taiwan, Vietnam, and India), where it has been used as a valuable source of culinary and traditional medicinal uses. The leaves, seeds, and stems of P. frutescens are used for various therapeutic applications in folk medicine. In the absence of a comprehensive review regarding all aspects of perilla, this review aims to present an overview pertaining to the botanical drug, ethnobotany, phytochemistry, and biological activity. It was found that the taxonomic classification of perilla species is quite confused, and the number of species is vague. Perilla has traditionally been prescribed to treat depression-related disease, anxiety, asthma, chest stuffiness, vomiting, coughs, colds, flus, phlegm, tumors, allergies, intoxication, fever, headache, stuffy nose, constipation, abdominal pain, and indigestion, and acts as an analgesic, anti-abortive agent, and a sedative. Until now, 271 natural molecules have been identified in perilla organs including phenolic acids, flavonoids, essential oils, triterpenes, carotenoids, phytosterols, fatty acids, tocopherols, and policosanols. In addition to solvent extracts, these individual compounds (rosmarinic acid, perillaldehyde, luteolin, apigenin, tormentic acid, and isoegomaketone) have attracted researchers’ interest for its pharmacological properties. Perilla showed various biological activities such as antioxidant, antimicrobial, anti-allergic, antidepressant, anti-inflammatory, anticancer, and neuroprotection effects. Although the results are promising in preclinical studies (in vitro and in vivo), clinical studies are insufficient; therefore, further study needs to be done to validate its therapeutic effects and to ensure its safety and efficacy.

Pharmacological Properties of Perilla

As mentioned earlier, the biological activity of perilla is due to the presence of various biochemical compounds that are responsible for producing health benefits for humans. Because of this, many researchers have focused on the isolation and identification of these active ingredients as well as their biological evaluations.

7.1. Antioxidant Activity

Epidemiological, clinical, and nutritional studies show that consumption of so-called functional foods and nutraceuticals may be associated with a lowered risk of cancers, cardiovascular diseases, and metabolic disorders [61]. These benefits are often attributed to the high antioxidant capacity of the drug, and especially to the content of phenolic acids, flavonoids, and carotenoids. It has been reported that extracts from perilla seeds and leaves exhibit concentration-dependent antioxidant activity, based on the 2,2-diphenyl-1-picryl-hydrazyl-hydrate (DPPH) radical assay, and 2,2′-azino-bis(3-ethylbenzothiazoline-6 sulphonic acid) (ABTS) radical cation assay [1]. Isolated rosmarinic acid (RA) and luteolin from the fruit of P. frutescens var. acuta showed significant DPPH scavenging capacity with half-maximal inhibitory concentration (IC50) values of 8.61 and 7.50 µM, respectively [32]. Similarly, among five phenolic compounds, RA and rosmarinic acid-3-O-glucoside were the dominant phenolic antioxidants with strong activity from cold-pressed perilla var. arguta seed flour studied by Zhou et al. [1]. RA isolated by these authors from perilla leaf exhibited DPPH radical scavenging activity of 88.3 ± 0.7% at a concentration of 10 μg/mL with a drug concentration eliciting 50% of the maximum stimulation (SC50) value of 5.5 ± 0.2 μg/mL. Tian et al. [41] proved that the antioxidant activity of perilla essential oil may depend on the location of cultivation. Extracts of drugs harvested from different regions exhibited varying degrees of scavenging ability at 10 mg/mL concentrations with an inhibition percentage of 94.80 ± 0.03%. The 80% methanol extract of perilla seeds exhibited a strong antioxidant property [62]. In vivo, the protective activity of RA from P. frutescens leaf (PFL) was demonstrated on Lipopolysaccharide (LPS)-induced liver injury of d-GalN-sensitized mice owing to the scavenging or reducing activities of superoxide or peroxynitrite rather than to inhibition of tumor necrosis factor (TNF)-α production [63].

The roles of the flavonoid luteolin from the perilla seeds seems to provide significant antioxidant activity for drugs and extracts. This compound significantly reversed hydrogen peroxide-induced cytotoxicity in primary cultured cortical neurons. Whereas, luteolin markedly attenuated the reactive oxygen species (ROS) production, and prevented the decrease in activities of mitochondria, catalase, and glutathione in ROS-insulted primary neurons, for preventing neurodegenerative diseases [64]. In another study, luteolin inhibited the peroxidation of linoleic acid catalyzed using soybean lipoxygenase-1 with an IC50 of 5.0 M (1.43 μg/mL) noncompetitively [65].

The monoterpene perillaldehyde was shown to be a potent thioredoxin inducer as it activates the Nrf2-Keap1 system [66]. It seems that the antioxidant activity of perilla may vary among different accessions. As part of an in vitro study in a human subjects, purple perilla leaves showed a higher antioxidant activity, and prevented the oxidation of low-density lipoprotein (LDL) than the green leaves [67]. Another study revealed that 2′,3′-dihydroxy-4′,6′-dimethoxychalcone (DDC) found in green perilla leaves enhanced cellular resistance to oxidative damage through activation of the Nrf2-antioxidant response element (ARE) pathway [68].

7.2. Antibacterial and Antifungal Activity

Recently, the demand for natural compounds from plant extracts as effective antibacterial agents against a wide range of bacteria is definitely growing to control human infection and for the preservation of food [69]. Perilla seed extract rich in polyphenols was examined for its antibacterial activity against oral cariogenic Streptococci and periodontopathic Porphyromonas gingivalis. The ethyl acetate extracts exhibit strong antibacterial activity against oral Streptococci and various strains of P. gingivalis. On the other hand, the ethanolic extract of defatted perilla seed weakly inhibited the growth of oral pathogenic bacterial strains. Among the polyphenols, luteolin showed marked antibacterial activity against the oral bacteria tested [70]. Additionally, the antibacterial activity of the essential oil from perilla leaves on Gram-positive and Gram-negative bacteria was studied, and the results showed the effectiveness of this essential oil to inhibit the growth of the tested bacteria. The minimum inhibitory concentration (MIC) on Staphylococcus aureus and Escherichia coli were 500 μg/mL and 1250 μg/mL. respectively [71]. The most abundant terpene-type compound, perillaldehyde, moderately inhibits a broad range of both bacteria in the range of 125–1000 pg/mL. This compound was also particularly active against filament fungi, with MIC values for M. mucedo and P. chrysogenum already at a 62.5 pg/mL concentration [72]. Kim and Choi [69,73] determined the antibacterial activity of the leaf ethanol extracts of PF var. acuta against S. aureus and Pseudomonas aeruginosa and detected that the population of P. aeruginosa decreased from 6.660 to 4.060 log CFU/mL, and that of S. aureus from 7.535 to 4.865 log CFU/mL, as well as to 2.600 log CFU/mL via extraction with ethyl acetate.

The fungicidal effects of perilla EO were described against Trichophyton mentagrophytes [74], and they dose-dependently decreased the production of α-toxin, enterotoxins A and B, and toxic shock syndrome toxin 1 (TSST-1) in both methicillin-sensitive S. aureus and methicillin-resistant S. aureus [75]. The antifungal activity of perilla EO distilled from aerial parts of the plant was also tested against phytopathogenic fungi and its activity was demonstrated in the cases of Aspergillus flavus, Aspergillus oryzae, Aspergillus niger, Rhizopus oryzae, and Alternaria alternate [41].

7.3. Anti-Allergic Effect

Studies show that water extracts of PF may inhibit allergic reactions in vivo and in vitro. PF extracts (0.05 to 1 g/kg) greatly inhibited systemic allergic reactions activated by anti-DNP IgE in rats in a dose-dependent manner [76]. Similarly, the water extract of PFL has been shown to have a positive result against atopic dermatitis in an animal model. The anti-allergic effects of PFL on 2,4-dinitrofluorobenzene (DNFB)-induced atopic dermatitis in C57BL/6 mice was evaluated by Heo et al. [77] and the results revealed that an aqueous extract (100 μg/mL) of PFL could significantly inhibit DNFB-induced atopic inflammation by alleviating the expression of MMP-9 and IL-31, as well as augmenting T-bet activity. In another experiment, water extract from PFL significantly suppressed the PCA-reaction, using mice ear-passive cutaneous anaphylaxis (PCA)-reaction, and the authors concluded the role of rosmarinic acid [9]. Application of an ethanol extract from PF, rather than the aqueous extract, suppressed the allergen-specific Th2 responses. Furthermore, airway inflammation and hyperreactivity in an ovoalbumin-sensitized murine model of asthma were alleviated. Based on this, Chen et al. [78] suggested perilla as a potential phytotherapeutic tool for immunomodulation.

Besides using crude extracts, individual compounds as a potential biologically active agent against allergies have also been studied. A novel glycoprotein fraction from the hot water extract of perilla was used and it was found that it moderately inhibited mast cell degranulation and the activities of hyaluronidase (IC50 = 0.42 mg/mL) in a dose-dependent manner [79]. Furthermore, daily oral supplementation with RA (1.5 mg/mouse, orally) from perilla significantly prevented the increase in the numbers of eosinophils in bronchoalveolar lavage fluids and in those around murine airways. Likewise, the expression of IL-4 and IL-5, and eotaxin in the lungs of sensitized mice, together with allergen-specific IgG1, were also inhibited. Due to these findings, the authors revealed RA as an effective intervention against allergic asthma [80]. In other study, perilla extracts enriched with RA could inhibit seasonal allergic rhinoconjunctivitis in humans at least partly via inhibition of polymorphonuclear leukocytes (PMNL) infiltration into the nostrils [81]. The use of a diet supplemented with perilla oil might be effective on asthmatic allergy via decreasing serum lipids and ovalbumin-specific IgG1 and IgA levels in mice [82].

7.4. Anti-Depressant Activity

Numerous studies focusing on the extracts and/or purified compounds of P. frutescens displayed antidepressant-like effects. Phenolic-type constituents of perilla leaf, such as apigenin, at intraperitoneal doses of 12.5 and 25 mg/kg [83], RA (2 mg/kg, i.p.) and caffeic acid (4 mg/kg, i.p.) each led to a considerable reduction of the duration of immobility in the forced swimming test. These compounds are also supposed to inhibit the emotional abnormality produced by stress [84,85], which is possibly mediated by the dopaminergic mechanisms in the mouse brain [83].

Essential oils and perillaldehyde from perilla leaves were also found to show an anti-depressant property in mice with CUMS-induced depression [86,87]. In another study, daily consumption of perillaldehyde (20 mg/kg, oral) demonstrated significant antidepressant-like effects in mice with LPS-induced depression and the authors concluded a potential benefit in inflammation-related depression [88]. Inhaling the same compound (perillaldehyde 0.0965 and 0.965 mg/mouse/day, 9 days) had antidepressant-like properties on a stress-induced, depression-like model in mice during the forced swim test (FST) through the olfactory nervous function [89].

The oil of PF seeds might have an anti-depressant activity too since a seed oil-rich diet during a forced swim test in adult male rats modulated the fatty acid profiles and brain-derived neurotrophic factor (BDNF) expression in the brain [90]. Moreover, perilla seed oil rich in n-3 fatty acids improved cognitive function in rats by generating new hippocampal neural membrane structures as well as by inducing specific protein expression [91].

7.5. Anti-Inflammatory Activity

Luteolin has been isolated from PFL ethanol extracts and was demonstrated to exert beneficial effects on neuro-inflammatory diseases in a dose-dependent manner (IC50 = 6.9 μM) through suppressing the expression of inducible nitric oxide synthase (iNOS) in BV-2 microglial cells [43].

The ethanol extract of PFL was identified to display significant anti-inflammatory activity in LPS-induced Raw 264.7 mouse macrophages through the inhibition of the expression of pro-inflammatory cytokines, inhibition of mitogen-activated protein kinase (MAPK) activation, and of nuclear factor-kappa (NF-κB) nuclear translocation in response to LPS [92]. The seed oil from PF showed a great protective effect against reflux esophagitis and this could be attributed to the antisecretory (anticholinergic, antihistaminic), antioxidant, and lipoxygenase inhibitory activities due to the presence of α-Linolenic acid (ALA) (18:3, n-3) [93]. Furthermore, RA isolated from PFL could inhibit the release of high mobility group box 1 protein (HMGB1) and down-regulated HMGB1-dependent inflammatory responses in human endothelial cells, HMGB1-mediated hyperpermeability, and leukocyte migration in mice, as well as reduced cecal ligation and puncture (CLP)-induced HMGB1 release and sepsis-related mortality. This could be a potential remediation for various vascular inflammatory diseases, such as sepsis and septic shock, via inhibition of the HMGB1 signaling pathway [94]. Lipophilic triterpene acids from ethanol extracts of red and green PFL were demonstrated to have remarkable anti-inflammatory activity on 12-O-tetradecanoylphorbol-13-acetate (TPA)-induced inflammation in mice (ID50: 0.09–0.3 mg per ear), and on the Epstein–Barr virus early antigen (EBV-EA) activation (91–93% inhibition at 1 × 103 mol ratio/TPA), [54]. A recent study in mice showed that PF extract ameliorated inflammatory bowel disease (IBD) via protection of dextran sulfate sodium-induced murine colitis, with NF-κB and signal transduction and activator of transcription 3 (STAT3) as putative targets [95]. A perillaketone-type and alkaloid isolated from aerial parts of perilla showed the remarkable inhibitory effect on pro-inflammatory cytokines (TNF-α and/or IL-6) and inflammatory mediator (NO) in LPS-stimulated RAW264.7 cells, indicating that these compounds might be active components for inflammatory disorders [96].

7.6. Antitumor Effect

A number of in vivo and in vitro studies have reported the potential anticancer activity of PF. Tormentic acid, a lipophilic triterpene acid from ethanol extracts of red and green PFL, remarkably blocked carcinogenenesis in an in vivo, two-stage mouse skin model [54]. Similarly, in an in vivo carcinogenesis model, topical application of perilla-derived fraction (2.0 mg/mouse) led to a significant reduction of 7,12-dimethylbenz(a)anthracene (DMBA)-initiated and TPA-promoted tumorgenesis. This is probably based on two independent effects: inhibition of oxidative DNA injury and inhibition of adhesion molecule, chemokine, and eicosanoid synthesis [97].

In addition, Lin et al. [98] evaluated the inhibitory effects of PFL and they found that it effectively induces apoptosis-related genes and could inhibit cell proliferation in human hepatoma HepG2 cells. They also observed that the inhibitory effect of PFL was much higher than the same dose of commercially available RA and luteolin compounds.

In another study, the application of ethanol extract of PFL resulted in induced apoptosis through the combinations of death receptor-mediated, mitochondrial, and endoplasmic reticulum stress-induced pathways, and substantially suppressed the cell proliferation via p21-mediated G1 phase arrest in human leukemia HL-60 cells [99]. Isoegomaketone (IK), an essential oil component of PF, was found to be another potential agent possessing anti-cancer activity. IK induces apoptosis through caspase-dependent and caspase-independent pathways in human colon adenocarcinoma DLD-1 cells [100].

7.7. Miscellaneous Effects

In addition to the pharmacological activities described above, different extracts, seed oil, and some individual phenolic compounds of perilla have been found to exhibit other special physiological activities indicating further therapeutic utilizations.

An aqueous extract of PF showed potent anti-HIV-1 activity via inhibition giant cell formation in co-culture of Molt-4 cells with and without HIV-1 infection showing inhibitory activity against HIV-1 reverse transcriptase [101]. A very recent study indicates the importance of PFL leaf extracts as a potential anti-aging agent for skin, as it showed effectiveness against UV-induced dermal matrix damage in vitro and in vivo [102]. The in vitro neuroprotection activity of unsaturated fatty acids of PF seed oil have been reported by Eckert et al. [103]. Perilla seed oil might be useful for other complaints too. Deng et al. [104] described in vitro and in vivo anti-asthmatic effects of perilla seed oil in the guinea pig and concluded that the oil may ameliorate lung function in asthma by regulating eicosanoid production and suppressing leukotriene (LT) generation. Zhao et al. [105] supposed a possible anti-ischemic activity of luteolin extracted from PFL, likely through a rebalancing of pro-oxidant/antioxidant status.

In vivo, the protective activity of RA from PFL was demonstrated on LPS-induced liver injury of d-GalN-sensitized mice. The treatments significantly reduced the elevation of plasma aspartate aminotransferase (AST) and alanine transaminase (ALT) levels, as well as anti-TNF and sphincter of Oddi dysfunction (SOD) treatment, compared with controls [63]. In one investigation, the hepatoprotective effects of sucrose-treated perilla leaves, other than untreated leaves, exhibited the best result in vitro and in vivo [10].

References

1. Zhou X.J., Yan L.L., Yin P.P., Shi L.L., Zhang J.H., Liu Y.J., Ma C. Structural characterisation and antioxidant activity evaluation of phenolic compounds from cold-pressed Perilla frutescens var. arguta seed flour. Food Chem. 2014;164:150–157. doi: 10.1016/j.foodchem.2014.05.062. [PubMed] [CrossRef] [Google Scholar]

2. Pandey A., Bhatt K.C. Diversity distribution and collection of genetic resources of cultivated and weedy type in Perilla frutescens (L.) Britton var. frutescens and their uses in Indian Himalaya. Genet. Resour. Crop. Evol. 2008;55:883–892. doi: 10.1007/s10722-007-9293-7. [CrossRef] [Google Scholar]

3. Asif M. Health effects of omega-3, 6, 9 fatty acids: Perilla frutescens is a good example of plant oils. Orient. Pharm. Exp. Med. 2011;11:51–59. doi: 10.1007/s13596-011-0002-x. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

4. Yu H., Qiu J.F., Ma L.J., Hu Y.J., Li P., Wan J.B. Phytochemical and phytopharmacological review of Perilla frutescens L. (Labiatae), a traditional edible-medicinal herb in China. Food Chem. Toxicol. 2017;108:375–391. doi: 10.1016/j.fct.2016.11.023. [PubMed] [CrossRef] [Google Scholar]

5. Chen Y.P. Application and prescriptions of Perilla in traditional Chinese medicine. In: Kosuna K., Haga M., Yu H.C., editors. Perilla: The genus Perilla. Harwood Academic Publishers; Amsterdam, The Netherlands: 1997. pp. 37–45. [Google Scholar]

6. Igarashi M., Miyazaki Y. A review on bioactivities of perilla: Progress in research on the functions of perilla as medicine and food. Evid. Based Complement. Altern. Med. 2013;2013 doi: 10.1155/2013/925342. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

7. Liu J., Wan Y., Zhao Z., Chen H. Determination of the content of rosmarinic acid by HPLC and analytical comparison of volatile constituents by GC-MS in different parts of Perilla frutescens (L.) Britt. Chem. Cent. J. 2013;7 doi: 10.1186/1752-153X-7-61. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

8. Ha T.J., Lee J.H., Lee M.H., Lee B.W., Kwon H.S., Park C.H., Shim K.B., Kim H.T., Baek I.Y., Jang D.S. Isolation and identification of phenolic compounds from the seeds of Perilla frutescens (L.) and their inhibitory activities against α-glucosidase and aldose reductase. Food Chem. 2012;135:1397–1403. doi: 10.1016/j.foodchem.2012.05.104. [PubMed] [CrossRef] [Google Scholar]

9. Makino T., Furuta Y., Wakushima H., Fujii H., Saito K.I., Kano Y. Anti-allergic effect of Perilla frutescens and its active constituents. Phytother. Res. 2003;17:240–243. doi: 10.1002/ptr.1115. [PubMed] [CrossRef] [Google Scholar]

10. Yang S.Y., Hong C.O., Lee H., Park S.Y., Park B.G., Lee K.W. Protective effect of extracts of Perilla frutescens treated with sucrose on tert-butyl hydroperoxide-induced oxidative hepatotoxicity in vitro and in vivo. Food Chem. 2012;133:337–343. doi: 10.1016/j.foodchem.2012.01.037. [PubMed] [CrossRef] [Google Scholar]

11. Meng L., Lozano Y.F., Gaydou E.M., Li B. Antioxidant activities of polyphenols extracted from Perilla frutescens varieties. Molecules. 2008;14:133–140. doi: 10.3390/molecules14010133. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

12. Meng L., Lozano Y., Bombarda I., Gaydou E.M., Li B. Polyphenol extraction from eight Perilla frutescens cultivars. C. R. Chim. 2009;12:602–611. doi: 10.1016/j.crci.2008.04.011. [CrossRef] [Google Scholar]

13. Nitta M., Lee J.K., Ohnishi O. Asian Perilla crops and their weedy forms: Their cultivation, utilization and genetic relationships. Econ. Bot. 2003;57:245–253. doi: 10.1663/0013-0001(2003)057[0245:APCATW]2.0.CO;2. [CrossRef] [Google Scholar]

14. Brenner D.M. Perilla: Botany, Uses and Genetic Resources. Wiley; New York, NY, USA: 1993. pp. 322–328. [Google Scholar]

15. Sa K.J., Choi S.H., Ueno M., Park K.C., Park Y.J., Ma K.H., Lee J.K. Identification of genetic variations of cultivated and weedy types of Perilla species in Korea and Japan using morphological and SSR markers. Genes Genom. 2013;35:649–659. doi: 10.1007/s13258-013-0117-1. [CrossRef] [Google Scholar]

16. Bachheti R.K., Joshi A., Ahmed T. A phytopharmacological overview of perilla frutescens. Int. J. Pharm. Sci. Rev. Res. 2014;26:55–61. [Google Scholar]

17. Hu Y., Sun L.W., Zhang Y.X., Wen C.X., Xie X.L., Liu Y.J. Primary identifications and palynological observations of Perilla in China. J. Syst. Evol. 2010;48:133–145. doi: 10.1111/j.1759-6831.2010.00067.x. [CrossRef] [Google Scholar]

18. Lee Y.J., Yang C.M. Growth behavior and perillaldehyde concentration of primary leaves Kwakof Perilla frutescens (L.) Britton grown in different seasons. Crop. Environ. Biol. 2006;3:135–146. [Google Scholar]

19. Negi V.S., Rawat L.S., Phondani P.C., Chandra A. Perilla frutescens in transition: A medicinal and oil yielding plant need instant conservation, a case study from Central Himalaya, India. Int. J. Environ. Sci. Technol. 2011;6:193–200. [Google Scholar]

20. Rawat D.S., Kharwal A.D. Ethnobotanical studies of weed flora in Shivalik Hills, Himachal Pradesh, India. Int. J. Adv. Res. 2014;2:218–226. [Google Scholar]

21. Bharali P., Sharma M., Sharma C.L., Singh B. Ethnobotanical survey of spices and condiments used by some tribes of Arunachal Pradesh. J. Med. Plants Stud. 2017;5:101–109. [Google Scholar]

22. Prakash A. Uses of some threatened and potential ethnomedicinal plants among the tribals of Uttar Pradesh and Uttrakhand in India; Proceedings of the National Conference on Forest Biodiversity—Earth’s Living Treasure; Lucknow, India. 22 May 2011; pp. 93–99. [Google Scholar]

23. Dhami N. Medicinal Plants in Nepal: An Anthology of Contemporary Research. Ecological Society; Kathmandu, Nepal: 2008. Ethnomedicinal uses of plants is Western Terai of Nepal: A case study of Dekhatbhuli VDC of Kanchanpur district; pp. 165–177. [Google Scholar]

24. Tiwari J.K., Ballabha R., Tiwari P. Some promising wild edible plants of Srinagar and its adjacent area in Alaknanda valley of Garhwal Himalaya, India. J. Am. Sci. 2010;6:167–174. [Google Scholar]

25. Yakang B., Gajurel P.R., Potsangbam S., Bhuyan L.R. Account of common and traditional non-timber forest products used by Apatani tribe of Arunachal Pradesh, India. Pleione. 2013;7:514–529. [Google Scholar]

26. Vlkova M., Polesny Z., Verner V., Banout J., Dvorak M., Havlik J., Krausova J. Ethnobotanical knowledge and agrobiodiversity in subsistence farming: Case study of home gardens in Phong My commune, central Vietnam. Genet. Resour. Crop. Evol. 2011;58:629–644. doi: 10.1007/s10722-010-9603-3. [CrossRef] [Google Scholar]

27. Inta A., Shengji P., Balslev H., Wangpakapattanawong P., Trisonthi C. A comparative study on medicinal plants used in Akha’s traditional medicine in China and Thailand, cultural coherence or ecological divergence? J. Ethnopharmacol. 2008;116:508–517. doi: 10.1016/j.jep.2007.12.015. [PubMed] [CrossRef] [Google Scholar]

28. Suneetha J., Seetharami R.T.V.V. Ethnomedicine for Rheumatoid Arthritis by the Tribes of East Godavari District, Andhra Pradesh. J. Med. Plant Res. 2016;6:1–5. [Google Scholar]

29. Ito M., Honda G., Sydara K. Perilla frutescens var. frutescens in northern Laos. J. Nat. Med. 2008;62:251–258. doi: 10.1007/s11418-007-0213-0. [PubMed] [CrossRef] [Google Scholar]

30. Saklani S., Chandra S., Gautam A.K. Phytochemical investigation and contribution of Perilla frutescence as spices in traditional health care system. Int. J. Pharm. Technol. 2011;3:3543–3554. [Google Scholar]

31. Peng Y., Ye J., Kong J. Determination of phenolic compounds in Perilla frutescens L. by capillary electrophoresis with electrochemical detection. J. Agric. Food Chem. 2005;53:8141–8147. doi: 10.1021/jf051360e. [PubMed] [CrossRef] [Google Scholar]

32. Gu L., Wu T., Wang Z. TLC bioautography-guided isolation of antioxidants from fruit of Perilla frutescens var. acuta. LWT Food Sci. Technol. 2009;42:131–136. doi: 10.1016/j.lwt.2008.04.006. [CrossRef] [Google Scholar]

33. Gai F., Peiretti P.G., Karamać M., Amarowicz R. Changes in the Total Polyphenolic Content and Antioxidant Capacities of Perilla (Perilla frutescens L.) Plant Extracts during the Growth Cycle. J. Food Qual. 2017;2017 doi: 10.1155/2017/7214747. [CrossRef] [Google Scholar]

34. Yamazaki M., Nakajima J.I., Yamanashi M., Sugiyama M., Makita Y., Springob K., Awazuhara M., Saito K. Metabolomics and differential gene expression in anthocyanin chemo-varietal forms of Perilla frutescens. Phytochemistry. 2003;62:987–995. doi: 10.1016/S0031-9422(02)00721-5. [PubMed] [CrossRef] [Google Scholar]

35. Kang N.S., Lee J.H. Characterisation of phenolic phytochemicals and quality changes related to the harvest times from the leaves of Korean purple perilla (Perilla frutescens) Food Chem. 2011;124:556–562. doi: 10.1016/j.foodchem.2010.06.071. [CrossRef] [Google Scholar]

36. Raut J.S., Karuppayil S.M. A status review on the medicinal properties of essential oils. Ind. Crop. Prod. 2014;62:250–264. doi: 10.1016/j.indcrop.2014.05.055. [CrossRef] [Google Scholar]

37. Ito M., Toyoda M., Kamakura S., Honda G. A new type of essential oil from Perilla frutescens from Thailand. J. Essent. Oil Res. 2002;14:416–419. doi: 10.1080/10412905.2002.9699907. [CrossRef] [Google Scholar]

38. Zhang X., Wu W., Zheng Y., Chen L., Qianrong C. Essential oil variations in different Perilla L. accessions: Chemotaxonomic implications. Plant Syst. Evol. 2009;281:1–10. doi: 10.1007/s00606-009-0152-1. [CrossRef] [Google Scholar]

39. Ghimire B.K., Yoo J.H., Yu C.Y., Chung I.M. GC–MS analysis of volatile compounds of Perilla frutescens Britton var. Japonica accessions: Morphological and seasonal variability. Asian Pac. J. Trop. Med. 2017;10:643–651. doi: 10.1016/j.apjtm.2017.07.004. [PubMed] [CrossRef] [Google Scholar]

40. Huang B., Lei Y., Tang Y., Zhang J., Qin L., Liu J. Comparison of HS-SPME with hydrodistillation and SFE for the analysis of the volatile compounds of Zisu and Baisu, two varietal species of Perilla frutescens of Chinese origin. Food Chem. 2011;125:268–275. doi: 10.1016/j.foodchem.2010.08.043. [CrossRef] [Google Scholar]

41. Tian J., Zeng X., Zhang S., Wang Y., Zhang P., Lü A., Peng X. Regional variation in components and antioxidant and antifungal activities of Perilla frutescens essential oils in China. Ind. Crop. Prod. 2014;59:69–79. doi: 10.1016/j.indcrop.2014.04.048. [CrossRef] [Google Scholar]

42. Ahmed H.M., Tavaszi-Sarosi S. Identification and quantification of essential oil content and composition, total polyphenols and antioxidant capacity of Perilla frutescens (L.) Britt. Food Chem. 2019;275:730–738. doi: 10.1016/j.foodchem.2018.09.155. [CrossRef] [Google Scholar]

43. Omer E.A., Khattab M.E., Ibrahim M.E. First cultivation trial of Perilla frutescens L. in Egypt. Flavour Fragr. J. 1998;13:221–225. doi: 10.1002/(SICI)1099-1026(1998070)13:4<221::AID-FFJ716>3.0.CO;2-Z. [CrossRef] [Google Scholar]

44. Başer K.H.C., Demirci B., Dönmez A.A. Composition of the essential oil of Perilla frutescens (L.) Britton from Turkey. Flavour Fragr. J. 2003;18:122–123. doi: 10.1002/ffj.1174. [CrossRef] [Google Scholar]

45. Bumblauskiene L., Jakstas V., Janulis V., Mazdzieriene R., Ragazinskiene O. Preliminary analysis on essential oil composition of Perilla L. cultivated in Lithuania. Acta Pol. Pharm. 2009;66:409–413. [PubMed] [Google Scholar]

46. Seo W.H., Baek H.H. Characteristic aroma-active compounds of Korean perilla (Perilla frutescens Britton) leaf. J. Agric. Food Chem. 2009;57:11537–11542. doi: 10.1021/jf902669d. [PubMed] [CrossRef] [Google Scholar]

47. Liu R.L., Zhang J., Mou Z.L., Hao S.L., Zhang Z.Q. Microwave-assisted one-step extraction-derivatization for rapid analysis of fatty acids profile in herbal medicine by gas chromatography-mass spectrometry. Analyst. 2012;137:5135–5143. doi: 10.1039/c2an36178g. [PubMed] [CrossRef] [Google Scholar]

48. Müller-Waldeck F., Sitzmann J., Schnitzler W.H., Graßmann J. Determination of toxic perilla ketone, secondary plant metabolites and antioxidative capacity in five Perilla frutescens L. varieties. Food Chem. Toxicol. 2010;48:264–270. doi: 10.1016/j.fct.2009.10.009. [PubMed] [CrossRef] [Google Scholar]

49. Duelund L., Amiot A., Fillon A., Mouritsen O.G. Influence of the active compounds of Perilla frutescens leaves on lipid membranes. J. Nat. Prod. 2012;75:160–166. doi: 10.1021/np200713q. [PubMed] [CrossRef] [Google Scholar]

50. Chen J.H., Xia Z.H., Tan R.X. High-performance liquid chromatographic analysis of bioactive triterpenes in Perilla frutescens. J. Pharm. Biomed. Anal. 2003;32:1175–1179. doi: 10.1016/S0731-7085(03)00160-2. [PubMed] [CrossRef] [Google Scholar]

51. Kim J.K., Park S.Y., Na J.K., Seong E.S., Yu C.Y. Metabolite profiling based on lipophilic compounds for quality assessment of perilla (Perilla frutescens) cultivars. J. Agric. Food Chem. 2012;60:2257–2263. doi: 10.1021/jf204977x. [PubMed] [CrossRef] [Google Scholar]

52. Ding Y., Hu Y., Shi L., Chao M.A., Liu Y.J. Characterization of fatty acid composition from five perilla seed oils in China and its relationship to annual growth temperature. J. Med. Plant Res. 2012;6:1645–1651. [Google Scholar]

53. Longvah T., Deosthale Y.G., Kumar P.U. Nutritional and short term toxicological evaluation of Perilla seed oil. Food Chem. 2000;70:13–16. doi: 10.1016/S0308-8146(99)00263-0. [CrossRef] [Google Scholar]

54. Banno N., Akihisa T., Tokuda H., Yasukawa K., Higashihara H., Ukiya M., Nishino H. Triterpene acids from the leaves of Perilla frutescens and their anti-inflammatory and antitumor-promoting effects. Biosci. Biotechnol. Biochem. 2004;68:85–90. doi: 10.1271/bbb.68.85. [PubMed] [CrossRef] [Google Scholar]

55. Ogawa Y., Takahashi S., Kitagawa R. Oxalate content in common Japanese foods. Acta Urol. J. Hinyokika Kiyo. 1984;30:305–310. [PubMed] [Google Scholar]

56. Adhikari P., Hwang K.T., Park J.N., Kim C.K. Policosanol content and composition in perilla seeds. J. Agric. Food Chem. 2006;54:5359–5362. doi: 10.1021/jf060688k. [PubMed] [CrossRef] [Google Scholar]

57. Kim K.S., Park S.H., Choung M.G. Nondestructive determination of oil content and fatty acid composition in perilla seeds by near-infrared spectroscopy. J. Agric. Food Chem. 2007;55:1679–1685. doi: 10.1021/jf0631070. [PubMed] [CrossRef] [Google Scholar]

58. Schantz M.M., Sander L.C., Sharpless K.E., Wise S.A., Yen J.H., NguyenPho A., Betz J.M. Development of botanical and fish oil standard reference materials for fatty acids. Anal. Bioanal. Chem. 2013;405:4531–4538. doi: 10.1007/s00216-013-6747-y. [PubMed] [CrossRef] [Google Scholar]

59. Jung D.M., Lee M.J., Yoon S.H., Jung M.Y. A Gas Chromatography-Tandem Quadrupole Mass Spectrometric Analysis of Policosanols in Commercial Vegetable Oils. J. Food Sci. 2011;76:C891–C899. doi: 10.1111/j.1750-3841.2011.02232.x. [PubMed] [CrossRef] [Google Scholar]

60. Longvah T., Deosthale Y.G. Effect of dehulling, cooking and roasting on the protein quality of Perilla frutescens seed. Food Chem. 1998;63:519–523. doi: 10.1016/S0308-8146(98)00030-2. [CrossRef] [Google Scholar]

61. Zhang H., Tsao R. Dietary polyphenols, oxidative stress and antioxidant and anti-inflammatory effects. Curr. Opin. Food Sci. 2016;8:33–42. doi: 10.1016/j.cofs.2016.02.002. [CrossRef] [Google Scholar]

62. Lee J.H., Park K.H., Lee M.H., Kim H.T., Seo W.D., Kim J.Y., Baek I.Y., Jang D.S., Ha T.J. Identification, characterisation, and quantification of phenolic compounds in the antioxidant activity-containing fraction from the seeds of Korean perilla (Perilla frutescens) cultivars. Food Chem. 2013;136:843–852. doi: 10.1016/j.foodchem.2012.08.057. [PubMed] [CrossRef] [Google Scholar]

63. Osakabe N., Yasuda A., Natsume M., Sanbongi C., Kato Y., Osawa T., Yoshikawa T. Rosmarinic acid, a major polyphenolic component of Perilla frutescens, reduces lipopolysaccharide (LPS)-induced liver injury in D-galactosamine (D-GalN)-sensitized mice. Free Radic. Biol. Med. 2002;33:798–806. doi: 10.1016/S0891-5849(02)00970-X. [PubMed] [CrossRef] [Google Scholar]

64. Zhao G., Yao-Yue C., Qin G.W., Guo L.H. Luteolin from Purple Perilla mitigates ROS insult particularly in primary neurons. Neurobiol. Aging. 2012;33:176–186. doi: 10.1016/j.neurobiolaging.2010.02.013. [PubMed] [CrossRef] [Google Scholar]

65. Ha T.J., Lee M.H., Kim H.T., Kwon H.S., Baek I.Y., Kubo I., Jang D.S.B. Slow-binding inhibition of soybean lipoxygenase-1 by luteolin. Arch. Pharmacal. Res. 2012;35:1811–1816. doi: 10.1007/s12272-012-1014-x. [PubMed] [CrossRef] [Google Scholar]

66. Masutani H., Otsuki R., Yamaguchi Y., Takenaka M., Kanoh N., Takatera K., Yodoi J. Fragrant unsaturated aldehydes elicit activation of the Keap1/Nrf2 system leading to the upregulation of thioredoxin expression and protection against oxidative stress. Antioxid. Redox Signal. 2009;11:949–962. doi: 10.1089/ars.2008.2292. [PubMed] [CrossRef] [Google Scholar]

67. Saita E., Kishimoto Y., Tani M., Iizuka M., Toyozaki M., Sugihara N., Kondo K. Antioxidant activities of Perilla frutescens against low-density lipoprotein oxidation in vitro and in human subjects. J. Oleo Sci. 2012;61:113–120. doi: 10.5650/jos.61.113. [PubMed] [CrossRef] [Google Scholar]

68. Izumi Y., Matsumura A., Wakita S., Akagi K.I., Fukuda H., Kume T., Akaike A. Isolation, identification, and biological evaluation of Nrf2-ARE activator from the leaves of green perilla (Perilla frutescens var. crispa f. viridis) Free Radic. Biol. Med. 2012;53:669–679. doi: 10.1016/j.freeradbiomed.2012.06.021. [PubMed] [CrossRef] [Google Scholar]

69. Kim D.H., Kim Y.C., Choi U.K. Optimization of antibacterial activity of Perilla frutescens var. acuta leaf against Staphylococcus aureus using evolutionary operation factorial design technique. Int. J. Mol. Sci. 2011;12:2395–2407. doi: 10.3390/ijms12042395. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

70. Yamamoto H., Ogawa T. Antimicrobial activity of perilla seed polyphenols against oral pathogenic bacteria. Biosci. Biotechnol. Biochem. 2002;66:921–924. doi: 10.1271/bbb.66.921. [PubMed] [CrossRef] [Google Scholar]

71. Qunqun G. Antibacterial activity of Perilla Frutescens leaf essential oil. Sci. Technol. Food Ind. 2003;9:019. [Google Scholar]

72. Kang R., Helms R., Stout M.J., Jaber H., Chen Z., Nakatsu T. Antimicrobial activity of the volatile constituents of Perilla frutescens and its synergistic effects with polygodial. J. Agric. Food Chem. 1992;40:2328–2330. doi: 10.1021/jf00023a054. [CrossRef] [Google Scholar]

73. Choi U.K., Lee O.H., Lim S.I., Kim Y.C. Optimization of antibacterial activity of Perilla frutescens var. acuta leaf against Pseudomonas aeruginosa using the evolutionary operation-factorial design technique. Int. J. Mol. Sci. 2010;11:3922–3932. doi: 10.3390/ijms11103922. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

74. Inouye S., Nishiyama Y., Hasumi Y., Yamaguchi H., Abe S., Uchida K. The vapor activity of oregano, perilla, tea tree, lavender, clove, and geranium oils against a Trichophyton mentagrophytes in a closed box. J. Infect. Chemother. 2006;12:349–354. doi: 10.1007/s10156-006-0474-7. [PubMed] [CrossRef] [Google Scholar]

75. Qiu J., Zhang X., Luo M., Li H., Dong J., Wang J., Deng X. Subinhibitory concentrations of perilla oil affect the expression of secreted virulence factor genes in Staphylococcus aureus. PLoS ONE. 2011;6 doi: 10.1371/journal.pone.0016160. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

76. Shin T.Y., Kim S.H., Kim S.H., Kim Y.K., Park H.J., Chae B.S., Kim H.M. Inhibitory effect of mast cell-mediated immediate-type allergic reactions in rats by Perilla frutescens. Immunopharmacol. Immunotoxicol. 2000;22:489–500. doi: 10.3109/08923970009026007. [PubMed] [CrossRef] [Google Scholar]

77. Heo J.C., Nam D.Y., Seo M.S., Lee S.H. Alleviation of atopic dermatitis-related symptoms by Perilla frutescens Britton. Int. J. Mol. Med. 2011;28:733–737. [PubMed] [Google Scholar]

78. Chen M.L., Wu C.H., Hung L.S., Lin B.F. Ethanol extract of Perilla frutescens suppresses allergen-specific Th2 responses and alleviates airway inflammation and hyperreactivity in ovalbumin-sensitized murine model of asthma. J. Evid. Based Complement. Altern. Med. 2015;2015 doi: 10.1155/2015/324265. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

79. Asada M., Fukumori Y., Inoue M., Nakagomi K., Sugie M., Fujita Y., Oka S. Glycoprotein derived from the hot water extract of mint plant, Perilla frutescens Britton. J. Agric. Food Chem. 1999;47:468–472. doi: 10.1021/jf9802777. [PubMed] [CrossRef] [Google Scholar]

80. Sanbongi C., Takano H., Osakabe N., Sasa N., Natsume M., Yanagisawa R., Yoshikawa T. Rosmarinic acid in perilla extract inhibits allergic inflammation induced by mite allergen, in a mouse model. Clin. Exp. Allergy. 2004;34:971–977. doi: 10.1111/j.1365-2222.2004.01979.x. [PubMed] [CrossRef] [Google Scholar]

81. Takano H., Osakabe N., Sanbongi C., Yanagisawa R., Inoue K.I., Yasuda A., Yoshikawa T. Extract of Perilla frutescens enriched for rosmarinic acid, a polyphenolic phytochemical, inhibits seasonal allergic rhinoconjunctivitis in humans. Exp. Biol. Med. 2004;229:247–254. doi: 10.1177/153537020422900305. [PubMed] [CrossRef] [Google Scholar]

82. Chang H.H., Chen C.S., Lin J.Y. Dietary perilla oil lowers serum lipids and ovalbumin-specific IgG1, but increases total IgE levels in ovalbumin-challenged mice. Food Chem. Toxicol. 2009;47:848–854. doi: 10.1016/j.fct.2009.01.017. [PubMed] [CrossRef] [Google Scholar]

83. Nakazawa T., Yasuda T., Ueda J., Ohsawa K. Antidepressant-like effects of apigenin and 2, 4, 5-trimethoxycinnamic acid from Perilla frutescens in the forced swimming test. Biol. Pharm. Bull. 2003;26:474–480. doi: 10.1248/bpb.26.474. [PubMed] [CrossRef] [Google Scholar]

84. Takeda H., Tsuji M., Inazu M., Egashira T., Matsumiya T. Rosmarinic acid and caffeic acid produce antidepressive-like effect in the forced swimming test in mice. Eur. J. Pharmacol. 2002;449:261–267. doi: 10.1016/S0014-2999(02)02037-X. [PubMed] [CrossRef] [Google Scholar]

85. Takeda H., Tsuji M., Miyamoto J., Matsumiya T. Rosmarinic acid and caffeic acid reduce the defensive freezing behavior of mice exposed to conditioned fear stress. Psychopharmacology. 2002;164:233–235. doi: 10.1007/s00213-002-1253-5. [PubMed] [CrossRef] [Google Scholar]

86. Yi L.T., Li J., Geng D., Liu B.B., Fu Y., Tu J.Q., Weng L.J. Essential oil of Perilla frutescens-induced change in hippocampal expression of brain-derived neurotrophic factor in chronic unpredictable mild stress in mice. J. Ethnopharmacol. 2013;147:245–253. doi: 10.1016/j.jep.2013.03.015. [PubMed] [CrossRef] [Google Scholar]

87. Ji W.W., Li R.P., Li M., Wang S.Y., Zhang X., Niu X.X., Li W., Yan L., Wang Y., Fu Q., et al. Antidepressant-like effect of essential oil of Perilla frutescens in a chronic, unpredictable, mild stress-induced depression model mice. Chin. J. Nat. Med. 2014;12:753–759. doi: 10.1016/S1875-5364(14)60115-1. [PubMed] [CrossRef] [Google Scholar]

88. Ji W.W., Wang S.Y., Ma Z.Q., Li R.P., Li S.S., Xue J.S., Fu Q. Effects of perillaldehyde on alternations in serum cytokines and depressive-like behavior in mice after lipopolysaccharide administration. Pharmacol. Biochem. Behav. 2014;116:1–8. doi: 10.1016/j.pbb.2013.10.026. [PubMed] [CrossRef] [Google Scholar]

89. Ito N., Nagai T., Oikawa T., Yamada H., Hanawa T. Antidepressant-like effect of l-perillaldehyde in stress-induced depression-like model mice through regulation of the olfactory nervous system. J. Evid. Based Complement. Altern. Med. 2011;2011 doi: 10.1093/ecam/nen045. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

90. Lee H.C., Ko H.K., Huang B.E.G., Chu Y.H., Huang S.Y. Antidepressant-like effects of Perilla frutescens seed oil during a forced swimming test. Food Funct. 2014;5:990–996. doi: 10.1039/c3fo60717h. [PubMed] [CrossRef] [Google Scholar]

91. Lee J., Park S., Lee J.Y., Yeo Y.K., Kim J.S., Lim J. Improved spatial learning and memory by perilla diet is correlated with immunoreactivities to neurofilament and α-synuclein in hilus of dentate gyrus. Proteome Sci. 2012;10 doi: 10.1186/1477-5956-10-72. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

92. Huang B.P., Lin C.H., Chen Y.C., Kao S.H. Anti-inflammatory effects of Perilla frutescens leaf extract on lipopolysaccharide-stimulated RAW264. 7 cells. Mol. Med. Rep. 2014;10:1077–1083. doi: 10.3892/mmr.2014.2298. [PubMed] [CrossRef] [Google Scholar]

93. Arya E., Saha S., Saraf S.A., Kaithwas G. Effect of Perilla frutescens fixed oil on experimental esophagitis in albino Wistar rats. BioMed Res. Int. 2013;2013 doi: 10.1155/2013/981372. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

94. Yang E.J., Ku S.K., Lee W., Lee S., Lee T., Song K.S., Bae J.S. Barrier protective effects of rosmarinic acid on HMGB1-induced inflammatory responses in vitro and in vivo. J. Cell. Physiol. 2013;228:975–982. doi: 10.1002/jcp.24243. [PubMed] [CrossRef] [Google Scholar]

95. Park D.D., Yum H.W., Zhong X., Kim S.H., Kim S.H., Kim D.H., Surh Y.J. Perilla frutescens extracts protects against dextran sulfate sodium-induced murine colitis: NF-κB, STAT3, and Nrf2 as putative targets. Front. Pharmacol. 2017;8 doi: 10.3389/fphar.2017.00482. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

96. Wang X.F., Li H., Jiang K., Wang Q.Q., Zheng Y.H., Tang W., Tan C.H. Anti-inflammatory constituents from Perilla frutescens on lipopolysaccharide-stimulated RAW264. 7 cells. Fitoterapia. 2018;130:61–65. doi: 10.1016/j.fitote.2018.08.006. [PubMed] [CrossRef] [Google Scholar]

97. Osakabe N., Yasuda A., Natsume M., Yoshikawa T. Rosmarinic acid inhibits epidermal inflammatory responses: Anticarcinogenic effect of Perilla frutescens extract in the murine two-stage skin model. Carcinogenesis. 2004;25:549–557. doi: 10.1093/carcin/bgh034. [PubMed] [CrossRef] [Google Scholar]

98. Lin C.S., Kuo C.L., Wang J.P., Cheng J.S., Huang Z.W., Chen C.F. Growth inhibitory and apoptosis inducing effect of Perilla frutescens extract on human hepatoma HepG2 cells. J. Ethnopharmacol. 2007;112:557–567. doi: 10.1016/j.jep.2007.05.008. [PubMed] [CrossRef] [Google Scholar]

99. Kwak C.S., Yeo E.J., Moon S.C., Kim Y.W., Ahn H.J., Park S.C. Perilla leaf, Perilla frutescens, induces apoptosis and G1 phase arrest in human leukemia HL-60 cells through the combinations of death receptor-mediated, mitochondrial, and endoplasmic reticulum stress-induced pathways. J. Med. Food. 2009;12:508–517. doi: 10.1089/jmf.2008.1103. [PubMed] [CrossRef] [Google Scholar]

100. Cho B.O., Jin C.H., Park Y.D., Ryu H.W., Byun M.W., Seo K.I., Jeong I.Y. Isoegomaketone induces apoptosis through caspase-dependent and caspase-independent pathways in human DLD1 cells. Biosci. Biotechnol. Biochem. 2011;75:1306–1311. doi: 10.1271/bbb.110088. [PubMed] [CrossRef] [Google Scholar]

101. Yamasaki K., Nakano M., Kawahata T., Mori H., Otake T., Ueda N., Murata H. Anti-HIV-1 activity of herbs in Labiatae. Biol. Pharm. Bull. 1998;21:829–833. doi: 10.1248/bpb.21.829. [PubMed] [CrossRef] [Google Scholar]

102. Bae J.S., Han M., Shin H.S., Kim M.K., Shin C.Y., Lee D.H., Chung J.H. Perilla frutescens leaves extract ameliorates ultraviolet radiation-induced extracellular matrix damage in human dermal fibroblasts and hairless mice skin. J. Ethnopharmacol. 2017;195:334–342. doi: 10.1016/j.jep.2016.11.039. [PubMed] [CrossRef] [Google Scholar]

103. Eckert G.P., Franke C., Nöldner M., Rau O., Wurglics M., Schubert-Zsilavecz M., Müller W.E. Plant derived omega-3-fatty acids protect mitochondrial function in the brain. Pharmacol. Res. 2010;61:234–241. doi: 10.1016/j.phrs.2010.01.005. [PubMed] [CrossRef] [Google Scholar]

104. Deng Y.M., Xie Q.M., Zhang S.J., Yao H.Y., Zhang H. Anti-asthmatic effects of perilla seed oil in the guinea pig in vitro and in vivo. Planta Med. 2007;73:53–58. doi: 10.1055/s-2006-957062. [PubMed] [CrossRef] [Google Scholar]

105. Zhao G., Zang S.Y., Jiang Z.H., Chen Y.Y., Ji X.H., Lu B.F., Guo L.H. Postischemic administration of liposome-encapsulated luteolin prevents against ischemia-reperfusion injury in a rat middle cerebral artery occlusion model. J. Nutr. Biochem. 2011;22:929–936. doi: 10.1016/j.jnutbio.2010.07.014. [PubMed] [CrossRef] [Google Scholar]

106. Jung S., Lee W.Y., Yong S.J., Shin K.C., Kim C.W., Lee J.H., Kim S.H. Occupational asthma caused by inhaling smoke from roasting perilla seeds. Allergy Asthma Respir. Dis. 2013;1:90–93. doi: 10.4168/aard.2013.1.1.90. [CrossRef] [Google Scholar]

107. Jeong Y.Y., Park H.S., Choi J.H., Kim S.H., Min K.U. Two cases of anaphylaxis caused by perilla seed. J. Allergy Clin. Immunol. 2006;117:1505–1506. doi: 10.1016/j.jaci.2006.02.044. [PubMed] [CrossRef] [Google Scholar]

108. Kerr L.A., Johnson B.J., Burrows G.E. Intoxication of cattle by Perilla frutescens (purple mint) Vet. Hum. Toxicol. 1986;28:412–416. [PubMed] [Google Scholar]

109. Bassoli A., Borgonovo G., Morini G., De Petrocellis L., Moriello A.S., Di Marzo V. Analogues of perillaketone as highly potent agonists of TRPA1 channel. Food Chem. 2013;141:2044–2051. doi: 10.1016/j.foodchem.2013.05.063. [PubMed] [CrossRef] [Google Scholar]