Genus Mentha: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor:

Mint (Mentha species) exhibits multiple health beneficial properties, such as prevention from cancer development and anti-obesity, antimicrobial, anti-inflammatory, anti-diabetic, and cardioprotective effects, as a result of its antioxidant potential, combined with low toxicity and high efficacy. Mentha species are widely used in savory dishes, food, beverages, and confectionary products. Phytochemicals derived from mint also showed anticancer activity against different types of human cancers such as cervix, lung, breast and many others. Mint essential oils show a great cytotoxicity potential, by modulating MAPK and PI3k/Akt pathways; they also induce apoptosis, suppress invasion and migration potential of cancer cells lines along with cell cycle arrest, upregulation of Bax and p53 genes, modulation of TNF, IL-6, IFN-γ, IL-8, and induction of senescence phenotype. Essential oils from mint have also been found to exert antibacterial activities against Bacillus subtilis, Streptococcus aureus, Pseudomonas aeruginosa, and many others.

  • mint
  • Mentha
  • phytochemicals
  • essential oils
  • anticancer
  • antimicrobial

1. Introduction

Medicinal plants and their derived compounds (phytochemicals) have been considered of pharmacological significance since ancient times. The use of plants in medicine dates back to 60,000 years ago, before the birth of civilization [1]. Today, more than 30% of all medicinal drugs (and their derivatives and analogs) derive from plants, and natural products will continue to possess considerable impact in human medicine. Most synthetic bioactive drugs are structurally similar to the phytochemicals of plants from which they were firstly isolated [2][3]. In many developing countries, plant materials have an important role in primary care or disease treatment. In addition, due to contraindications in the usage of chemical drugs, there is a growing interest in the utilization of the plant-derived medicinal products [4], since compared to synthetic drugs some of these natural products have lower toxicity and higher efficacy [5]. Anticancer drugs, antibiotics, anti-inflammatory drugs, immunomodulators are among the most important drugs having herbal origin [6][7][8]. Moreover, some of the aforementioned plant-derived compounds have pleasant taste and odor and can be used in kitchen as flavorings, spices and food [9]. Among the plants with global economic and culinary importance, mint is used worldwide for perfuming sweet and savoury dishes and flavouring tea, in addition to its pharmacological importance.

2. Phytochemical Composition of Mentha

Mentha species are rich in polyphenols [10] and, moreover, contain caffeic acid and its derivatives caftaric acid, cinnamic acid, ferulic acid, and oleanolic acid [11][12][13]. Flavonoids including luteolin and its derivatives apigenin, acacetin, diosmin, salvigenin and thymonin, have also been detected in these plants, accounting for some 10–70 compounds out of the total phenolics, and also flavanols such as catechin, epicatechin and coumarins, including esculetin and scopoletin [14][15][16][17]. With regard to mint compositions, the essential oils represent a major focus. They are colorless, pale yellow or greenish yellow [18] and alcohols, ketones, esters, ethers and oxides are their main components in Mentha species [19][20]. Menthol, menthone, isomenthone, menthyl acetate, linalool, linalyl acetate, lippione, pulegone, carvone, piperitenone oxide and cis-piperitone epoxide are the main constituents of essential oils prepared from different mint species [19][21][22][23]. Table 1 lists the main compounds isolated from different species of this genus.

Table 1. Main chemical compounds isolated from different Mentha species.

3. Properties of the Mentha Genus

Mentha species are characterized by a great chemical diversity and were reported to contain a number of chemical compounds responsible for pharmacological properties (Table 2).

Table 2. Chemical compounds in Mentha genus and their pharmacological properties.

The use of essential oils has a long history being widely exploited in the food, beverage, confectionery and cosmetic industries. In traditional Iranian medicine, it is reported that Mentha species have cooling sensation properties, strengthen the stomach and are effective to relieve digestive symptoms, respiratory tract problems and hemorrhoids [66]. It has been shown that Mentha spicata essential oil can reduce pain caused by Caesarean section [67][68]. Recent investigations have suggested that the transient receptor potential cation channel subfamily M (melastatin) member 8 (TRPM8) and transient receptor potential cation channel, subfamily A, member 1 (TRPA1) are implicated in pain relief (analgesia) and sensation of cooling [67][69]. In Ayurvedic medicine, some Mentha species were used to mitigate skin problems and headaches [70]. In vitro studies showed that peppermint essential oil and menthol acted as smooth muscle relaxants via blocking Ca2+ influx through L-type Ca2+ channels [71][72]. Also peppermint juice led to reduction in total cholesterol levels, triglycerides and could increase HDL levels in the blood of university students [73] (Figure 1).

Figure 1. Summary of the main effects of Mentha species.

4. Clinical Trials

Limited data are available on clinical trials using Mentha species in humans [74], and indeed only two works reported the use of Mentha species in humans related to cancer. First a randomized, double-blind clinical trial was conducted in 200 patients to determine the efficacy of volatile oils of Mentha piperita or Mentha spicata in preventing chemotherapy-induced nausea and vomiting (CINV) in four groups, control, placebo, M. piperita, M. spicata. The results showed a significant reduction in the intensity and number of emetic events in the first 24 h with M. spicata and M. piperita in both treatment groups (p< 0.05) when compared with the control and no adverse effects were reported. The cost of treatment was also reduced when essential oils were used [75]. The second work is another randomized, double blind placebo clinical trial conducted in 60 patients to evaluate the effects of Mentha piperita (and Matricaria recutita) on oral mucositis (OM) in patients undergoing hematopoietic stem cell transplantation (HSCT) [76]. OM is one of the most common side effects of intensive chemotherapy in patients undergoing HSCT. Patients who received herbal mouthwash three times daily for 1 week before HSCT showed significant improvements in pain intensity (p = 0.009), dryness (p = 0.04) and dysphagia (p = 0.009), suggesting a therapeutic role for M. piperita in OM.

5. Adverse Effects of Mentha Species

Although medicinal plants such as Mentha species are commonly believed to be safe, they are not devoid of side effects that can be severe in some cases. Furthermore, allergic reactions can occur with any natural or synthetic compound in sensitive persons. No chronic toxicity studies in humans are available, therefore toxicity of Mentha species are scarcely reported. However, it seems that no adverse effects have been reported after consumption 0.24 mL of pure M. spicata essential oil daily for three continuous weeks in two different clinical studies [77][78]. Leaves of Mentha spicata are known for its contact allergy such as contact cheilitis caused by its essential oil use as toothpaste flavoring [79]. In addition, MPEO is also associated with adverse effects like vomiting, headaches, flushing, heartburn and nausea [80]. Mentha piperita and spearmint tea can deprive the human body of iron and cause anemia if consumed excessively, and carvone and limonene showed to be major allergens [79]. Gürbüz found that pulegone, contained in low concentrations in Mentha piperita oil extracts, is hepatotoxic, and Douros et al. also reported the likely liver injury caused by M. piperita [81][82]. Other research showed that menthol and pulegone could be toxic compounds; in particular, pulegone and its metabolite menthofuran have been suggested as the hepatotoxic compounds in Mentha pulegium and have been also found in smaller quantity in Mentha piperita [83]. Notably, inhalation of menthol can cause apnea and larygospasm in sensitive patients and indeed has been reported that mentholated preparation can be involved in nausea, anorexia, cardiac problems, ataxia and other CNS symptoms [84]. Mentha spicata extracts displayed toxicity to neuronal cells when applied at concentrations which are one order of magnitude higher than those effective for radical scavenging [85]. The positive correlation between the two aforementioned effects suggests that a higher desirable radical scavenging is associated with a higher undesirable toxicity.

Peppermint oil is contraindicated in obstruction of the bile ducts, gallbladder inflammation, and severe liver failure [83]. The American College of Gastroenterology has recommended reducing the peppermint intake as it is a risk factor for gastroesophageal reflux disease (GERD) and lifestyle changes [86][87]. Further, Zong et al. [88] reported that peppermint essential oil, not only stimulated bile fluid secretion, but might be involved in upregulating the bile acid synthesis-related gene, cholesterol 7α-hydroxylase (CYP7A1), and the nuclear bile acid receptor FXR (farnesoid X receptor) mRNA. In addition, peppermint oil could cause heartburn or perianal irritation, bradycardia and muscle tremor, a hypersensitivity reaction, contact dermatitis, abdominal pain and jaundice in newborn babies [89]. A study on a 58-year-old woman who smoked menthol cigarettes also established that she suffered from gastrointestinal upsets with occasional vomiting, hand tremor, mental confusion and depression which were all ascribed to menthol [90]. Similarly, in another case report, a 40-year-old woman with no history of asthma or any other forms of allergy has shown the symptoms of dyspnea, wheezing and nasal after using menthol containing candies and toothpaste, suggesting the development of classical symptoms of asthma [91]. Menthol administered for 28 days at a dose level (≤ 800 mg/kg) in rats caused hepatocellular changes and pulegone (≤ 160 mg/kg) has been reported as hepatotoxic and neurotoxic. Consequently, pulegone caused weight loss, atonia, decreased blood creatinine, histopathological changes in the liver and also in the white matter of the cerebellum [92]. Menthone (≤ 800 mg/kg orally) on the other hand, dose dependently decreased plasma creatinine, but increased alkaline phosphatase and bilirubin along with liver and spleen weight [93]. Menthone administered to rats at a high dose over 28 days did display some signs of hepatotoxicity and cerebellar histopathology [93]. In a later study examining peppermint constituents for their possible induction of encephalopathy, one-month treatment with limonene (≤ 1600 mg/kg) or 1,8-cineole (1000 mg/kg) produced an accumulation of protein droplets containing α-2µ-globulin in proximal tubular epithelial cells, but no encephalopathy in rats [94]. Peppermint and menthol have both been shown to possess Ca2+ channel blocking properties, which might underlie their mechanism of efficacy against irritable bowel syndrome in the clinic [84]. However, in some patients, the use of peppermint is accompanied by oral symptoms like burning mouth syndrome and oral ulceration [87]. Also in this context, direct application of peppermint oil to the chest or nasal area of infants is not recommended due to the risk of apnea, bronchial and/or laryngeal spasms [95].

This entry is adapted from the peer-reviewed paper 10.3390/molecules26041118

References

  1. Solecki, R.S. Shanidar IV, a Neanderthal Flower Burial in Northern Iraq. Sci. 1975, 190, 880–881.
  2. Nirmal, S.A.; Pal, S.C.; Otimenyin, S.O.; Aye, T.; Elachouri, M.; Kundu, S.K.; Thandavarayan, R.A.; Mandal, S.C. Contribution of Herbal Products In Global Market. The pharma review 2013, 95–104.
  3. Sevindik, M. Pharmacological Properties of Mentha Species. J. Tradit. Med. Clin. Naturop. 2018, 7, 1–4.
  4. Anwar, F.; Abbas, A.; Mehmood, T.; Gilani, A.-H.; Rehman, N.-u. Mentha: A genus rich in vital nutra-pharmaceuticals—A review. Phytother. Res. 2019, 33, 2548–2570.
  5. Fabricant, D.S.; Fransworth, N.R. The value of Plants used in traditional medicine for drug discovery. Environ Health Perspect 2001, 109, 69–75.
  6. Rates, S. Plants as source of drugs. Toxicon 2001, 39, 603–613.
  7. Petrovska, B.B. Historical review of medicinal plants’ usage. Phar. Rev. 2012, 6, 1–5.
  8. Balasubramanian, A.; Ramalingam, K.; Krishnan, S.; Ajm, C. Anti-inflammatory Activity of Morus indica Linn. Ira. J. of Phar. Thera. 2005, 4, 13–15.
  9. Salehi, B.; Valussi, M.; Jugran, A.K.; Martorell, M.; Ramírez-Alarcón, K.; Stojanović-Radić, Z.Z.; Antolak, H.; Kręgiel, D.; Mileski, K.S.; Sharifi-Rad, M.; et al. Nepeta species: From farm to food applications and phytotherapy. Trends Food Sci. Technol. 2018, 80, 104–122.
  10. Pereira, O.R.; Cardoso, S.M. Overview on Mentha and Thymus Polyphenols. Cur. Anal. Chem. 2013, 9, 382–396.
  11. Benedec, D.; Vlase, L.; Oniga, I.; Mot, A.C.; Silaghi-Dumitrescu, R.; Hanganu, D.; Tiperciuc, B.; Crişan, G. LC-MS analysis and antioxidant activity of phenolic compounds from two indigenous species of mentha. Farmacia 2013, 61, 262–267.
  12. Taamalli, A.; Arráez-Román, D.; Abaza, L.; Iswaldi, I.; Fernandez-Gutierrez, A.; Zarrouk, M.; SeguraCarretero, A. LC-MS-based metabolite profiling of methanolic extracts from the medicinal and aromatic species Mentha pulegium and Origanum majorana. Phytochem. Anal. 2015, 26, 320–330.
  13. Wu, Z.; Tan, B.; Liu, Y.; Dunn, J.; Guerola, P.M.; Tortajada, M.; Cao, Z.; Ji, P. Chemical Composition and Antioxidant Properties of Essential Oils from Peppermint, Native Spearmint and Scotch Spearmint. Molecules 2019, 24, 2825.
  14. Koşar, M.; Dorman, H.J.D.; Baser, K.H.C.; Hiltunen, R. Screening of Free Radical Scavenging Compounds in Water Extracts ofMenthaSamples Using a Postcolumn Derivatization Method. J. Agric. Food Chem. 2004, 52, 5004–5010.
  15. Bimakr, M.; Rahman, R.A.; Taip, F.S.; Ganjloo, A.; Salleh, L.M.; Selamat, J.; Hamid, A.; Zaidul, I. Comparison of different extraction methods for the extraction of major bioactive flavonoid compounds from spearmint (Mentha spicata L.) leaves. Food Bioprod. Process. 2011, 89, 67–72.
  16. Salin, O.; Törmäkangas, L.; Leinonen, M.; Saario, E.; Hagström, M.; Ketola, R.A.; Saikku, P.; Vuorela, H.; Vuorela, P.M. Corn Mint (Mentha arvensis) Extract Diminishes Acute Chlamydia pneumoniae Infection in Vitro and in Vivo. J. Agric. Food Chem. 2011, 59, 12836–12842.
  17. Fatiha, B.; Didier, H.; Naima, G.; Khodir, M.; Martin, K.; Léocadie, K.; Caroline, S.; Mohamed, C.; Pierre, D. Phenolic composition, in vitro antioxidant effects and tyrosinase inhibitory activity of three Algerian Mentha species: M. spicata (L.), M. pulegium (L.) and M. rotundifolia (L.) Huds (Lamiaceae). Ind. Crop. Prod. 2015, 74, 722–730.
  18. Franz, C.; Novak, J. Sources of Essential Oils; CRC Press/Taylor & Francis Group: Boca Raton, FL, USA, 2010; pp. 45–67.
  19. Lawrence, B.M. Mint. The Genus Mentha; CRC Press: Boca Raton, FL, USA, 2006; pp. 1–56.
  20. Malingré, T.M. Chemotaxonomic study of Mentha arvensis L. Pharm. week. 1971, 106, 165–171.
  21. Maffei, M.; Codignola, A. Photosynthesis, Photorespiration and Herbicide Effect on Terpene Production in Peppermint (Mentha piperitaL.). J. Essent. Oil Res. 1990, 2, 275–286.
  22. Sokovic, M.D.; Vukojevic, J.; Marin, P.D.; Brkic, D.D.; Vajs, V.; van Griensven, L.J. Chemical composition of essential oils of Thymus and Mentha species and their antifungal activities. Molecules 2009, 14, 238–249.
  23. Moghaddam, M.; Pourbaige, M.; Tabar, H.K.; Farhadi, N.; Hosseini, S.M.A. Composition and Antifungal Activity of Peppermint Mentha piperita Essential Oil from Iran. J. of Ess. Oil Bear. Pla. 2013, 16, 506–512.
  24. Andro, A.-R.; Boz, I.; Zamfirache, M.-M.; Burzo, I. Chemical composition of essential oils from Mentha aquatica L. at different moments of the ontogenetic cycle. J. Med. Plant Res. 2013, 7, 470–473.
  25. Dai, D.N.; Thang, T.D.; Emmanuel, E.E.; Abdulkabir, O.O.; Ogunwande, I.A. Study on essential oil of Mentha aquatica L. from Vietnam. Am. J. Essent. Oil. 2015, 2, 12–16.
  26. Getahun, Z.; Asres, K.; Mazumder, A.; Bucar, F. Essential Oil Composition, Antibacterial and Antioxidant Activities of Mentha aquatica Growing in Ethiopia. Ethiop. Pharm. J. 2008, 26, 9–16.
  27. Morteza-Semnani, K.; Saeedi, M.; Akbarzadeh, M. The Essential Oil Composition of Mentha aquatica L. J. Essent. Oil Bear. Plants 2006, 9, 283–286.
  28. Fancello, F.; Zara, S.; Petretto, G.L.; Chessa, M.; Addis, R.; Rourke, J.P.; Pintore, G. Essential oils from three species of Mentha harvested in Sardinia: Chemical characterization and evaluation of their biological activity. Int. J. Food Prop. 2017, 20, 1–11.
  29. Pereira, O.R.; Macias, R.I.R.; Domingues, M.R.M.; Marin, J.J.G.; Cardoso, S.M. Hepatoprotection of Mentha aquatica L., Lavandula dentata L. and Leonurus cardiaca L. Antioxidants 2019, 8, 267.
  30. Barros, A.D.S.; Morais, S.M.d.; Ferreira, P.A.T.; Vieira, Í.G.P.; Craveiro, A.A.; Fontenelle, R.O.D.S.; Menezes, J.E.S.A.D.; Silva, F.W.F.D.; Sousa, H.A.D. Chemical composition and functional properties of essential oils from Mentha species. Ind. Cro. and Prod. 2015, 76, 557–564.
  31. Li, M.; Xu, L.; Li, Z.; Qian, S.; Qin, M. Chemical constituents from Mentha canadensis. Biochem. Syst. Ecol. 2013, 49, 144–147.
  32. Shelepova, O.V.; Voronkova, T.V.; Kondrat’eva, V.V.; Semenova, M.V.; Bidyukova, G.F.; Olehnovich, L.S. Phenotypic and Phytochemical Differences between Mentha arvensis L. and Mentha canadiensis L. Biol. Bull. 2014, 41, 19–23.
  33. Thawkar, B.S.; Jawarkar, A.G.; Kalamkar, P.V.; Pawar, K.P.; Kale, M.K. Phytochemical and pharmacological review of Mentha arvensis. Int. J. Green Pharm. 2016, 10, 71–76.
  34. Abdel-Hameed, E.-S.S.; Salman, M.S.; Fadl, M.A.; Elkhateeb, A.; Hassan, M.M. Chemical Composition and Biological Activity of Mentha longifolia L. Essential Oil Growing in Taif, KSA Extracted by Hydrodistillation, Solvent Free Microwave and Microwave Hydrodistillation. J. Essent. Oil Bear. Plants 2018, 21, 1–14.
  35. Teymouri, M.; Alizadeh, A. Chemical composition and antimicrobial activity of the essential oil of Mentha mozaffarianii Jamzad growing wild and cultivated in Iran. Nat. Prod. Res. 2017, 32, 1320–1323.
  36. Golparvar, A.R.; Hadipanah, A.; Gheisari, M.M.; Salehi, S.; Khaliliazar, R.; Ghasemi, O. Comparative analysis of chemical composition of Mentha longifolia (L.) Huds. J Herb Med 2017, 7, 235–241.
  37. Daneshbakhsh, D.; Asgarpanah, J.; Najafizadeh, P.; Rastegar, T.; Mousavi, Z. Safety Assessment of Mentha mozaffarianii Essential Oil: Acute and Repeated Toxicity Studies. Iran. J. Med. Sci. 2018, 43, 479–486.
  38. Tavakkoli-Khaledi, S.; Asgarpanah, J. Essential Oil Chemical Composition of Mentha mozaffarianii Jamzad Seeds. J. Mex. Chem. Soc. 2017, 60, 19–22.
  39. Alexa, E.; Danciu, C.; Radulov, I.; Obistioiu, D.; Sumalan, R.M.; Morar, A.; Dehelean, C.A. Phytochemical Screening and Biological Activity of Mentha × piperita L. and Lavandula angustifolia Mill. Extracts. Anal. Cell. Pathol. 2018, 2678924.
  40. Goudjil, M.B.; Ladjel, S.; Ben, S.E.; Zighmi, S.; Hamada, D. Chemical Composition, Antibacterial and Antioxidant Activities of the Essential Oil Extracted from the Mentha piperita of Southern Algeria. Res. J. Phytochem. 2015, 9, 79–87.
  41. Marwa, C.; Fikri-Benbrahim, K.; Ou-Yahia, D.; Farah, A. African peppermint (Mentha piperita) from Morocco: Chemical composition and antimicrobial properties of essential oil. J. Adv. Pharm. Technol. Res. 2017, 8, 86–90.
  42. Satmi, F.R.S.; Hossain, M.A. In vitro antimicrobial potential of crude extracts and chemical compositions of essential oils of leaves of Mentha piperita L native to the Sultanate of Oman. Pac. Sci. Rev. A: Nat. Sci. Eng. 2016, 18, 103–106.
  43. Tsai, M.; Wu, C.; Lin, T.; Lin, W.; Huang, Y.; Yang, C. Chemical Composition and Biological Properties of Essential Oils of Two Mint Species. Trop. J. Pharm. Res. 2013, 12, 577–582.
  44. Politeo, O.; Bektašević, M.; Carev, I.; Jurin, M.; Roje, M. Cover Picture: Phytochemical Composition, Antioxidant Potential and Cholinesterase Inhibition Potential of Extracts from Mentha pulegium L. (C&B 12/2018). Chem. Biodivers. 2018, 15, e1800624.
  45. Rahmani, F.; Rezaeian-Doloei, R.; Alimoradi, L. Evaluation of Phytochemical Composition of Mentha pulegium L. Essential Oil and Its Antibacterial Activity against Several Pathogenic Bacteria. Iran. J. Med. Microbiol. 2018, 11, 167–177.
  46. Siham, F.; Rachid, B.; Read, A.-Z.M. Chemical Composition and Antioxidant Effect of Mentha rotundifolia Extracts. Pharmacogn. J. 2019, 11, 521–526.
  47. Derwich, E.; Benziane, Z.; Boukir, A.; Benaabidate, L. GC-MS Analysis of the Leaf Essential Oil of Mentha rotundifolia, a Traditional Herbal Medicine in Morocco. Chem. Bul. “Politehnica” Univ. Timisoara 2009, 54, 85–88.
  48. Riahi, L.; Elferchichi, M.; Ghazghazi, H.; Jebali, J.; Ziadi, S.; Aouadhi, C.; Chograni, H.; Zaouali, Y.; Zoghlami, N.; Mliki, A. Phytochemistry, antioxidant and antimicrobial activities of the essential oils of Mentha rotundifolia L. in Tunisia. Ind. Crop. Prod. 2013, 49, 883–889.
  49. Bahadori, M.B.; Zengin, G.; Bahadori, S.; Dinparast, L.; Movahhedin, N. Phenolic composition and functional properties of wild mint (Mentha longifolia var. calliantha (Stapf) Briq.). Int. J. Food Prop. 2018, 21, 183–193.
  50. Brada, M.; Bezzina, M.; Marlier, M.; Lognay, G.C. Chemical Composition of the Leaf Oil of Mentha rotundifolia (L.) from Algeria. J. Essent. Oil Res. 2006, 18, 663–665.
  51. Yahia, I.B.H.; Jaouadi, R.; Trimech, R.; Boussaid, M.; Zaouali, Y. Variation of chemical composition and antioxidant activity of essential oils of Mentha x rotundifolia (L.) Huds. (Lamiaceae) collected from different bioclimatic areas of Tunisia. Biochem. Syst. Ecol. 2019, 84, 8–16.
  52. Sytar, O.; Hemmerich, I.; Zivcak, M.; Rauh, C.; Brestic, M. Comparative analysis of bioactive phenolic compounds composition from 26 medicinal plants. Saudi J. Biol. Sci. 2018, 25, 631–641.
  53. Bardaweel, S.K.; Bakchiche, B.; Alsalamat, H.A.; Rezzoug, M.; Gherib, A.; Flamini, G. Chemical composition, antioxidant, antimicrobial and Antiproliferative activities of essential oil of Mentha spicata L. (Lamiaceae) from Algerian Saharan atlas. BMC Complement. Altern. Med. 2018, 18, 1–7.
  54. Bayan, Y.; Küsek, M.; Ahi Evran University. Kahramanmaras Sutcu Imam University Chemical Composition and Antifungal and Antibacterial Activity of Mentha spicata L. Volatile Oil. Cienc. Investig. Agrar. 2018, 45, 64–69.
  55. Sevindik, E.; Yamaner, Ç.; Kurtoğlu, C.; Tin, B. Chemical Composition of Mentha spicata L. subsp. tomentosa and M. pulegium L., and their Antimicrobial Activity on Strong Pathogen Microorganisms. Not. Sci. Biol. 2017, 9, 73–76.
  56. Bouyahya, A.; Belmehdi, O.; Abrini, J.; Dakka, N.; Bakri, Y. Chemical composition of Mentha suaveolens and Pinus halepensis essential oils and their antibacterial and antioxidant activities. Asian Pac. J. Trop. Med. 2019, 12.
  57. Božović, M.; Pirolli, A.; Ragno, R. Mentha suaveolens Ehrh. (Lamiaceae) Essential Oil and Its Main Constituent Piperitenone Oxide: Biological Activities and Chemistry. Molecules 2015, 20, 8605–8633.
  58. Rita, I.; Pereira, C.; Barros, L.; Santos-Buelga, C.; Ferreira, I.C.F.R. Mentha spicata L. infusions as sources of antioxidant phenolic compounds: Emerging reserve lots with special harvest requirements. Food Funct. 2016, 7, 4188–4192.
  59. Hussain, A.I.; Anwar, F.; Shahid, M.; Ashraf, M.; Przybylski, R. Chemical Composition, and Antioxidant and Antimicrobial Activities of Essential Oil of Spearmint (Mentha spicata L.) From Pakistan. J. Essent. Oil Res. 2010, 22, 78–84.
  60. Al-Okbi, S.Y.; Fadel, H.H.; Mohamed, D.A. Phytochemical constituents, antioxidant and anticancer activity of Mentha citrata and Mentha longifolia. Res. J. Pharm. Biol. Chem 2015, 6, 739–751.
  61. Shahbazi, Y. Chemical Composition and In Vitro Antibacterial Activity of Mentha spicata Essential Oil against Common Food-Borne Pathogenic Bacteria. J. Pathog. 2015, 2015, 1–5.
  62. Barros, A.D.S.; De Morais, S.M.; Ferreira, P.A.T.; Vieira, Í.G.P.; Craveiro, A.A.; Fontenelle, R.O.D.S.; De Menezes, J.E.S.A.; Da Silva, F.W.F.; De Sousa, H.A. Chemical composition and functional properties of essential oils from Mentha species. Ind. Crop. Prod. 2015, 76, 557–564.
  63. Li, Y.; Liu, Y.; Ma, A.; Bao, Y.; Wang, M.; Sun, Z. In vitro antiviral, anti-inflammatory, and antioxidant activities of the ethanol extract of Mentha piperita L. Food Sci. Biotechnol. 2017, 26, 1675–1683.
  64. Elansary, H.O.; Szopa, A.; Kubica, P.; Ekiert, H.; Klimek-Szczykutowicz, M.; El-Ansary, D.O.; Mahmoud, E.A. Polyphenol Profile and Antimicrobial and Cytotoxic Activities of Natural Mentha × piperita and Mentha longifolia Populations in Northern Saudi Arabia. Processes 2020, 8, 479.
  65. Yassin, M.T.; Mostafa, A.A.; Al-Askar, A.A. Anticandidal and anti-carcinogenic activities of Mentha longifolia (Wild Mint) extracts in vitro. J. King Saud Univ.-Sci. 2020, 32, 2046–2052.
  66. Kee, L.A.; Shori, A.B.; Baba, A.S. Bioactivity and health effects of Mentha spicata. Integr. Food, Nutr. Metab. 2017, 5, 1–2.
  67. Liu, B.; Fan, L.; Balakrishna, S.; Sui, A.; Morris, J.B.; Jordt, S.-E. TRPM8 is the principal mediator of menthol-induced analgesia of acute and inflammatory pain. Pain 2013, 154, 2169–2177.
  68. Uritu, C.M.; Mihai, C.T.; Stanciu, G.-D.; Dodi, G.; Alexa-Stratulat, T.; Luca, A.; Leon-Constantin, M.-M.; Stefanescu, R.; Bild, V.; Melnic, S.; et al. Medicinal Plants of the Family Lamiaceae in Pain Therapy: A Review. Pain Res. Manag. 2018, 2018, 1–44.
  69. Karashima, Y.; Damann, N.; Prenen, J.; Talavera, K.; Segal, A.; Voets, T.; Nilius, B. Bimodal Action of Menthol on the Transient Receptor Potential Channel TRPA1. J. Neurosci. 2007, 27, 9874–9884.
  70. Mahboubi, M. Mentha spicata L. essential oil, phytochemistry and its effectiveness in flatulence. J. Tradit. Complement. Med. 2018, 1–7.
  71. Amato, A.; Liotta, R.; Mulè, F. Effects of menthol on circular smooth muscle of human colon: Analysis of the mechanism of action. Eur. J. Pharmacol. 2014, 740, 295–301.
  72. Grigoleit, H.-G.; Grigoleit, P. Pharmacology and preclinical pharmacokinetics of peppermint oil. Phytomedicine 2005, 12, 612–616.
  73. Brabalho, S.M.; Machado, F.M.V.F.; Oshiiwa, M.; Abreu, M.; Guiger, E.L.; Tomazela, P.; Goulart, R.A. Investiga-tion of the effects of peppermint (Mentha piperita) on the biochemical and anthropometric profile of university students. Ciência Tecnol. Alime. 2011, 31, 584–588.
  74. ClinicalTrials.gov. Clinical Trials. Available online: (accessed on 3 January 2020).
  75. Tayarani-Najaran, Z.; Talasaz-Firoozi, E.; Nasiri, R.; Jalali, N.; Hassanzadeh, M. Antiemetic activity of volatile oil from Mentha spicata and Mentha × piperita in chemotherapy-induced nausea and vomiting. Canc. Med. Sci. 2013, 7, 290.
  76. Tavakoli Ardakani, M.; Ghassemi, S.; Mehdizadeh, M.; Mojab, F.; Salamzadeh, J.; Ghassemi, S.; Hajifathali, A. Evaluating the effect of Matricaria recutita and Mentha piperita herbal mouthwash on management of oral mucositis in patients undergoing hematopoietic stem cell transplantation: A randomized, double blind, placebo controlled clinical trial. Compl. Ther. Med. 2016, 29, 29–34.
  77. Amui Roknabad, M.; Sarafraz, N. Comparison between the Effect of Supermint and Ibuprofen on Primary Dysmenorrheal: A Randomized Clinical Trial. Qom Univ Med. Sci J. 2011, 5, 37–41.
  78. Nasiri, A.; Pakmehr, M.; Shahdadi, H.; Balouchi, A.; Sepehri, Z.; Ghalenov, A.R. A Comparative Study of Dimethicone and Supermint Anti-flatulence Effects on Reducing Flatulence in Patients with Irritable Bowel Syndrome. Der. Pharm. Lett. 2015, 7, 432–436.
  79. Akdoğan, M.; Tamer, M.N.; Cüre, E.; Cüre, M.C.; Köroğlu, B.K.; Delibaş, N. Effect of spearmint (Mentha spicata Labiatae) teas on androgen levels in women with hirsutism. Phytother. Res. 2007, 21, 444–447.
  80. Peixoto, I.T.A.; Furletti, V.F.; Anibal, P.C.; Duarte, M.C.T.; Höfling, J.F. Potential pharmacological and toxicological basis of the essential oil from Mentha spp. Rev. Ciênc. Farm. Básica Apl. 2010, 30, 235–239.
  81. Gürbüz, P. An Overview about Adverse Hepatic Effects of the Plants Used in Turkey. Cerrahpaşa Med J. 2020, 44, 115–124.
  82. Douros, A.; Bronder, E.; Andersohn, F.; Klimpel, A.; Kreutz, R.; Garbe, E.; Bolbrinker, J. Herb-induced liver injury in the Berlin case-control surveillance study. Int J. Mol. Sci 2016, 17, 114.
  83. Balakrishnan, A. Therapeutic uses of peppermint –A review. J. Pharm. Sci. Res. 2015, 7, 474–476.
  84. Hawthorn, M.; Ferrante, J.; Luchowski, E.; Rutledge, A.; Wei, X.Y.; Triggle, D.J. The actions of peppermint oil and menthol on calcium channel dependent processes in intestinal, neuronal and cardiac preparations. Alim. Pharm. Ther. 1988, 2, 101–118.
  85. Amat-ur-Rasool, H.; Symes, F.; Tooth, D.; Schaffert, L.-N.; Elmorsy, E.; Ahmed, M.; Hasnain, S.; Carter, W.G. Potential Nutraceutical Properties of Leaves from Several Commonly Cultivated Plants. Biomolecules 2020, 10, 1556.
  86. Jarosz, M.; Taraszewska, A. Risk factors for gastroesophageal reflux disease: The role of diet. Prz Gastroenterol. 2014, 9, 297.
  87. DeVault, K.R.; Castell, D.O. Updated guidelines for the diagnosis and treatment of gastroesophageal reflux disease. Am. J. Gastroenter. 2005, 100, 190–200.
  88. Zong, L.; Qu, Y.; Luo, D.X.; Zhu, Z.Y.; Zhang, S.; Su, Z.; Shan, J.C.; Gao, X.P.; Lu, L.G. Preliminary experimental research on the mechanism of liver bile secretion stimulated by peppermint oil. J. Dig. Dis. 2011, 12, 295–301.
  89. Sharma, V.; Hussain, S.; Gupta, M.; Saxena, A.K. In vitro anticancer activity of extracts of Mentha spp. against human cancer cells. Indian J. Biochem. Biophys. 2014, 51, 416–419.
  90. Alankar, S. A review on peppermint oil. Asian J. Pharm. Clin. Res. 2009, 2, 27–33.
  91. Dos Santos, M.; CE, S.G. Menthol-induced asthma: A case report. J. Invest. Allerg. Clin. Immun. 2001, 11, 56–58.
  92. Thorup, I.; Würtzen, G.; Carstensen, J.; Olsen, P. Short term toxicity study in rats dosed with pulegone and menthol. Toxicol. Let. 1983, 19, 207–210.
  93. Madsen, C.; Würtzen, G.; Carstensen, J. Short-term toxicity study in rats dosed with menthone. Toxicol. Let. 1986, 32, 147–152.
  94. Kristiansen, E.; Madsen, C. Induction of protein droplet (α2μ-globulin) nephropathy in male rats after short-term dosage with 1, 8-cineole and l-limonene. Toxicol. Let. 1995, 80, 147–152.
  95. Shah, P.P.; Mello, P. A review of medicinal uses and pharmacological effects of Mentha piperita. Nat. Prod. Rad. 2004, 3, 214–221.
More
This entry is offline, you can click here to edit this entry!
Video Production Service