Next Article in Journal
Nitrogen Acquisition and Transport in the Ectomycorrhizal Symbiosis—Insights from the Interaction between an Oak Tree and Pisolithus tinctorius
Next Article in Special Issue
Antiviral Activity of Selected Essential Oils against Cucumber Mosaic Virus
Previous Article in Journal
The Influence of Different Crop Mulches on Weed Infestation, Soil Properties and Productivity of Wheat under Conventional and Conservation Production Systems
Previous Article in Special Issue
Antimicrobial, Antibiofilm, and Antioxidant Properties of Essential Oil of Foeniculum vulgare Mill. Leaves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Healthy Zerumbone: From Natural Sources to Strategies to Improve Its Bioavailability and Oral Administration

by
María Dolores Ibáñez
1,
Noelia M. Sánchez-Ballester
2,3 and
María Amparo Blázquez
1,*
1
Departament de Farmacologia, Facultat de Farmàcia, Universitat de València, Avd. Vicent Andrés Estellés s/n, Burjassot, 46100 València, Spain
2
ICGM, Département Chimie et Matériaux Moléculaires, University of Montpellier, CNRS, ENSCM, 34090 Montpellier, France
3
Department of Pharmacy, Nîmes University Hospital, 30900 Nîmes, France
*
Author to whom correspondence should be addressed.
Submission received: 17 November 2022 / Revised: 6 December 2022 / Accepted: 15 December 2022 / Published: 20 December 2022

Abstract

:
Zerumbone is a multifunctional compound with antimicrobial, antitumor, hyperalgesic, antioxidant and anti-inflammatory applications, and constitutes a point molecule for the future synthesis of derivatives with improved efficiency. This monocyclic sesquiterpenoid is found in high content in wild ginger (Zingiber zerumbet Smith), a perennial herb with economic importance as an ornamental as well as a medicinal plant. The presence of zerumbone is a distinctive feature that allows identification and differentiation from other species, not only in Zingiber, but also in Curcuma, Alpinia, Boesenbergia, Ethlingera and Ammomum spp., as well as related families (Costaceaee). To successfully use zerumbone in areas such as medicine, food and agriculture, further research on improving its low solubility and bioavailability, as well as its preservation, is a major current priority. In addition, despite its promising pharmacological activities, preclinical and clinical studies are required to demonstrate and evaluate the in vivo efficacy of zerumbone.

1. Introduction

Zerumbone (2,6,9,9-tetramethyl-(2E,6E,10E)-cycloundeca-2,6,10-trien-1-one) is a highly volatile monocyclic sesquiterpene compound (C15H22O) found in large amounts in the rhizome of wild edible ginger, Zingiber zerumbet (L.) Roscoe ex Sm. essential oil [1,2]. It is mainly distributed in tropical and subtropical regions, including Sri Lanka, Nepal, Bangladesh, Malaysia, India, and southwest China (commonly known as hong qiu jiang). Traditionally, this rhizomatous herbaceous species, belonging to the Zingiberaceae family, is also known as bitter ginger, shampoo ginger and pinecone ginger. It has applications as a flavouring and appetizing agent and for treating pain in folk medicine [3,4]. The genus Zingiber comprises about 141 species (12 species native to China), mainly cultivated in Asia, Central, South America and Africa, and is an important source of essential oil widely used in the perfumery, cosmetics, and pharmaceutical industries. Although the different members of this genus have similar morphology, they differ significantly in their pharmacological and therapeutic properties [5,6,7]. Among these species, many studies have focused on five species: Zingiber officinale Rosc (ginger); Zingiber zerumbet (L.) Smith, Zingiber corallinum Hance, Zingiber mioga (Thunb.) Rosc and Zingiber striolatum Diels. The main compounds obtained from the essential oils of these species are considerably different, with α-zingiberene, zerumbone and palmitic acid extracted as the main component of Z. officinale, Z. zerumbet and Z. striolatum, respectively. Focusing our attention on Z. zerumbet, the main bioactive compound found in the ethanolic extract and essential oil is zerumbone [8]. However, the chemical composition of Z. zerumbet rhizome essential oil is different depending on the geographical origin and the drying processes used, which can reduce the biological activity. In this sense, the zerumbone content in essential oils obtained from fresh rhizomes from different geographical origins varies from 8.1 to 84.8%, whereas in essential oils from dried rhizomes, the zerumbone content ranges from 1.2 to 35.5% [6]. Zerumbone (69.9%), followed by α-humulene (12.9%), ae also the main compounds identified in the essential oil of a variety of Z. zerumbet, identified as Z. zerumbet var. darcyi, which grows wild and is used against coughs and colds by the Manipur tribes in northeast India [9]. A similar composition of zerumbone (72.9%) and α-humulene (7.1%), was found in fresh rhizomes of Z. zerumbet collected in an experimental culture in Uttarakhand [2]. However, the amount of zerumbone (40.2%) was lower when the essential oil was obtained from the rhizome collected in Xishuangbanna (Yunnan Province, China) [10]. Zerumbone is a multifunctional compound that exhibits various biological activities, such as antimicrobial, antioxidant, anti-inflammatory, antitumor or antihyperalgesic, attributed to the structure of this bioactive compound biosynthesized from α-humulene (Figure 1) [11]. This naturally occurring sesquiterpene has a 65.3 °C melting point, contains an 11-membered ring, an α, β-unsaturated carbonyl group, and three double bonds (two conjugated and one isolated) at C6, C2 and C10 as part of a cross-conjugated system, which result in its pharmacological activity. It is interesting to note that α-humulene is a versatile starting material for conversion to other derivatives and therapeutic compounds such as paclitaxel, which is widely used in breast cancer treatment.
In the zerumbone biosynthetic path, the enzyme cytochrome P450 hydroxylates α-humulene at the C8 position, and the dehydrogenation of 8-hydroxy humulene is catalyzed by a short-chain dehydrogenase. Zerumbone has also been biosynthesized from metabolically modified Saccharomyces cerevisiae by introducing α-humulene synthase, α-humulene 8-hydroxylase and zerumbone synthase variant together with Arabidopsis thaliana cytochrome P reductase into a yeast strain [12]. The α, β-unsaturated carbonyl group of zerumbone is thought to be responsible for its cell membrane rupturing effect, explaining the mechanism of action of zerumbone against methicillin-resistant Staphylococcus aureus (MRSA). In fact, zerumbone exerts anti-MRSA effects by causing membrane depolarization, increasing membrane permeability, and ultimately disrupting the cell membrane and killing the bacteria [13]. Zerumbone possesses significant antimicrobial activity against Aspergillus niger, Bacillus cereus, B. subtilis, Bacteroides fragilis, Enterococcus faecalis, Escherichia coli, Helicobacter pylori, Proteus vulgaris, Salmonella choleraesuis, S. typhimurium, S. aureus, S. epidermitis, Streptococcus mutans and Yersinia enterocolitica. It has an anti-virulence effect by inhibiting the biofilm formation and hyphal growth of Candida albicans in a concentration-dependent manner. In addition, it exhibits significant cytotoxic activity against lung, colon, leukemia, ovarian, skin, liver and breast cancers, with selective cytotoxic activity on human tumor cell lines, not found with the structural analogue, α-humulene, that lacks the carbonyl group (Figure 1) [4,5,6,13,14]. Several studies have reported the antitumor activity of zerumbone in vitro, as well as in vivo models such as breast, colorectal, cervical, lung, renal cell carcinoma and skin cancers [4]. Furthermore, zerumbone is able to resensitize breast cancer cells to paclitaxel, increasing oxidative stress [15]. The combination of zerumbone-vinblastine and zerumbone-paclitaxel was strongly synergistic in the inhibition of HeLa cell proliferation, showing a similar mechanism of action to other clinically used chemotherapeutic drugs [16], and the combination with cisplatin showed a dose-dependent induction of cytotoxicity and apoptosis in head and neck squamous cell line carcinoma [17]. The anti-biofilm and antimicrobial effects of zerumbone have been studied against the main anaerobic opportunistic pathogen of ulcerative colitis and other inflammatory bowel diseases, B. fragilis. This oxygenated sesquiterpene significantly inhibited biofilm formation, disrupted established biofilms as well as microbial growth of B. fragilis, and is an antipathogenic compound capable of preventing chronic inflammatory diseases of this microorganism [18]. Similar effects were observed at sub-inhibitory doses of zerumbone for the multidrug-resistant opportunistic human pathogen Acinetobacter baumannii [19]. Zerumbone is also, in a concentration-dependent manner, an antimicrobial and antibiofilm compound for the management of nosocomial infections against mono-species and dual-species biofilms of C. albicans and S. aureus [20]. Furthermore, zerumbone is not only capable of influencing microbial richness and diversity but can also improve the balance of intestinal microbial composition. In other words, zerumbone can restore the composition of the intestinal microbiota associated with various diseases such as inflammatory bowel diseases and colon cancer [21]. The anti-inflammatory activity of zerumbone is mainly due to the inhibition of pro-inflammatory genes and antioxidant pathways. Zerumbone was demonstrated to inhibit cyclooxygenase, prostaglandin and interleukin and was able to protect the gastric mucosa via antioxidant mechanisms [22]. Regarding hyperalgesic properties, zerumbone significantly alleviated tactile and cold allodynia, as well as mechanical and thermal hyperalgesia [23]. Zerumbone was able to exert antiallodynic and anti-hyperalgesic effects in a mouse model of neuropathic pain through inhibition of the serotonergic system, specifically 5-HT receptor subtypes 1A 1B, 2A, 3, 6 and 7 [24]. More recent studies indicate that in the antiallodynic and anti-hyperalgesic effects of zerumbone, K+ channels and µ-, δ- and κ-opioid receptor subtypes [25], as well as cannabinoids and peroxisome proliferator-activated receptors (PPARs), are involved [26] in its modulation of neuropathic pain. In addition, the anti-hyperglycemic and anti-hyperlipidemic effects of zerumbone have beneficial effects in reducing the risk of diabetic retinopathy [27], and the dual acetylcholinesterase and butyrylcholinesterase inhibitory activity of zerumbone with high permeability across the blood-brain barrier supports its benefits as a preventive agent in Alzheimer’s disease [28].The pharmacokinetic properties of zerumbone explain the low bioavailability and distribution in the body of this water-insoluble compound. To achieve therapeutic effects in topical applications, incorporation of zerumbone into liquid or semi-solid lipid-based carriers, such as nanostructured lipid gels, improves its solubility and availability, promoting wound healing [29,30]. The physicochemical properties of zerumbone-loaded nanostructured lipid carriers were shown to be relatively stable when stored at 4 °C, 25 °C or 40 °C for at least one month, making them an interesting sustained-release drug carrier for the treatment of leukemia [31]. Nanosuspensions are also used to improve the oral bioavailability of poorly aqueous soluble drugs. In this sense, a nanosuspension formulated using high-pressure homogenization with sodium dodecyl sulphate and hydroxypropyl methylcellulose improved the saturation solubility and dissolution profile of zerumbone [32], and its inclusion complexes with several cyclodextrin derivatives, were also useful for improving the solubility of zerumbone [33]. A clinical study of the inclusion complex of zerumbone with hydroxypropyl-β-cyclodextrin revealed that 20 µM of zerumbone injected with 0.15 g of ranitidine orally for a 20-day treatment was quite effective and safe in acute and chronic gastritis, with or without H. pylori infection [34]. Thus, many studies have focused on finding the most suitable matrices to encapsulate zerumbone and improve its low aqueous solubility, poor absorption, and low bioavailability. Therefore, in the present review, we gathered information on the natural sources of this compound, its antimicrobial activity, mainly associated with anti-biofilm action, and the latest strategies reported to improve the bioavailability of zerumbone and maintaining or increasing its activity, with a particular focus on advances in oral administration.

2. Natural Sources of Zerumbone

The monocyclic sesquiterpenoid zerumbone has been found mainly in plants of the Zingiberaceae family, principally Z. zerumbet, which is considered an invasive species in certain regions, such as Taiwan [35,36,37]. Zerumbone is the major component (Table 1) detected in the essential oil [3,38,39,40,41], ethanol (87.4–242.73 mg/g), acetone (64.58%) and methanol extracts (2.767–33.018 mg/g) of Z. zerumbet rhizome [8,42,43], as well as the principal compound detected in many Z. zerumbet varieties such as darcyi (69.9%) and Jor Lab ZB-103 (32.79%) [9,44]. In contrast to rhizome oil, the essential oil extracted from the aerial parts (leaf, stem and inflorescence) of Z. zerumbet is dominated by nerolidol, trans-phytol, β-caryophyllene, linalool, pinenes, with a lesser content of zerumbone [38,40]. In general, both genotype and environment conditions influence the amount of zerumbone in Z. zerumbet rhizomes [44,45,46]. Thus, the essential oil obtained from fresh rhizomes of Z. zerumbet yielded more zerumbone than the oils from dry rhizomes of Z. zerumbet [6]. Optimization of extraction methods is a fundamental step in obtaining natural bioactive compounds [39,47]. In this sense, optimization with microwave-assisted extraction was 17.9, 28.1 and 30.2% higher than sonication, reflux and Soxhlet, respectively [48]. Zerumbone constitutes the predominant compound of the rhizome essential oil of other Zingiber spp. This genus is well-known for its abundant essential oils [49]. This sesquiterpenoid accounts for more than half of the sample of Z. spectabile Griff. (59.1%) [50]. It confers antibacterial activity against multidrug-resistant strains of E. coli, S. aureus and Pseudomonas aeruginosa (0.19–0.38 mg/mL) [50]. The essential oil from the leaves is rich in other sesquiterpenes, β-caryophyllene (21.3%) and β-elemene (12.5%), providing different odor and weaker antibacterial activity [50]. Other authors disagree on the chemical composition of Z. spectabile rhizome oil [51]. In particular, terpinen-4-ol (23.7%), labda-8(17),12-diene-15,16-dial (24.3%), α-terpineol (13.1%) and β-pinene (10.3%) were the major components found in rhizome oil of Z. spectabile from Malaysia [51]. The hexanolic extract of Z. spectabile rhizomes does not contain zerumbone, but 1,1′-ethylenebisdecali predominates instead (41.3%) [52].
Zerumbone is also the principal component of the essential oil extracted from Z. amaricans BL. (40.70–65.06%) and Z. aromaticum Val. (31.45%) [53,54] species. Both share other secondary components, such as α-humulene, β-selinene and caryophyllene oxide [54].
The sesquiterpenoid zerumbone has also been detected in the essential oil of the rhizome of Z. montanum (J. Koenig) Link ex A. Dietr., where it constitutes one of the main components with relevant pharmacological activities (anticancer, anti-inflammatory, antioxidant, antimicrobial, gastric and liver protection) [55]. These results differ from those of other authors who reported a different chemical composition of the essential oil Z. montanum. Particularly, in the volatile oil obtained from fresh rhizomes of Z. montanum in India, the monoterpenes 4-terpineol (38.0%) and sabinene (10.0%) were the most abundant compounds [56]. Similarly, Rajkumari and Sanatombi detected trace amounts (0.0107 mg/g) of zerumbone in the essential oil of in vitro cell cultures and microrhizomes of Z. montanum [57]. This cyclic sesquiterpene was the major compound in the ethanol extract isolated from Z. montanum from Bangladesh, followed by five kaempferol derivatives [58]. The essential oil of Z. ottensii Valenton rhizomes contained principally terpene compounds (Table 1), among which zerumbone was the main constituent (24.73%), followed by terpinen-4-ol (18.75%), sabinene (15.19%) and β-pinene (7.95%) [59]. Although the essential oils rich in zerumbone have countless applications and benefits, especially as anticancer agents, excessive and prolonged use can produce cytotoxic effects, including embryotoxic and teratogenic effects [59,60,61].
Zerumbone is not always one of the main components of Zingiber spp. rhizome essential oils [78]. For instance, zerumbone constitutes a minor component in the rhizome oils of Z. cassumunar Roxb. (0.4%) and Z. officinale Roscoe (1.25%) [79,80]. In other species, this molecule does not even appear. This is the case of essential oil obtained from the fresh rhizome of Z. wrayi var. halabala in Thailand, Thailandese Z. kerrii Craib and Z. barbatum Wall in Myanmar [81,82,83]. These differences between species may be due to the fact that the qualitative and quantitative chemical composition is subject to the origin, location, extraction methods, and degree of dryness of rhizomes [83,84]. As a consequence, the presence of zerumbone in Z. zerumbet rhizome oil constitutes a distinctive feature that allows the identification and differentiation from other ginger oils as well as different parts of the plant [38].
In the Zingiberaceae family, with approximately 53 genera and over 1200 species distributed in all tropical regions of the world, there are also rhizomatous genera such as Curcuma, well-known for their distinctive aroma and numerous bioactivities [85,86,87,88]. This shares certain bioactive compounds and medicinal properties with Zingiber spp [89]. Related to this, Xiang et al. studied the chemical composition and bioactivities of essential oils obtained from Chinese Curcuma spp., detecting zerumbone as one of the major components among them (1.08~15.45%) and the principal one in C. rubescens Roxb rhizome oil (15.45%) [71]. In contrast, Zhang et al. reported zerumbone as the second major compound (6.14 ± 1.42 mg/g) in C. rubescens essential oil, after ar-turmerone (6.88 ± 2.01 mg/g) [70]. Rhizome oil obtained from different Indian cultivars of C. longa L. also yielded high zerumbone content (0.383–8.333 mg/g), especially in dry conditions (3.209–8.333 mg/g), conferring promising immunomodulatory activity [90]. On the other hand, the essential oil of other Curcuma spp. contains lower quantities of zerumbone, such as C. aromatica, C. caesia and C. xanthorrhiza [91]. Other examples are C. kwangsiensis var nanlingensis essential oil that generated 0.32 mg/g of zerumbone in comparison to the main component β-elemenone (5.03 mg/g) [70], the volatile oil of C. angustifolia Roxb. yielded 2.10–2.17% of zerumbone with respect to the major compound epicurzerenone (29.60–31.18%) and rhizome oil of C. zedoaria (Christm.) Rosc. contained traces of the cyclic sesquiterpene (0.03%), while γ–eudesmol acetate was the main molecule (15.65%) [92]. Other abundant compounds in the essential oil of C. zedoaria were curcumin, ethyl p-methoxycinnamate and β-turmerone [93]. Despite being in lower concentration, zerumbone plays an important role in the bioactivity of these species. The minor components enhance the activity of the essential oil by means of synergism between them [72]. Zerumbone was also detected in the eluted extracts of the rhizome of Curcuma spp. It was one of the most significant compounds in the extracts of C. caesia Roxb. and C. zedoaria, showing potential anticancer and antioxidant activity, respectively [94,95].
Alpinia represents the largest, most extensive, and most taxonomically complex genus in Zingiberaceae family, with 230 species distributed throughout tropical and subtropical Asia. These plants live in forests and have been used as ornamentals, food and as a source of medicines due to their antimicrobial, cytotoxic or antioxidant activities [96,97,98]. The most cultivated and well-known species of this group is probably A. galanga (L.) Willd (AG) [96]. It represents a key ingredient in nutraceutical and pharmaceutical products due to the numerous benefits of its phytoconstituents, such as anti-inflammatory, anticancer and antioxidant effects [99]. Zerumbone is one of these functional components. It was the major compound detected in the essential oil of A. galanga from Sri Lanka (44.9%). In fact, it constitutes a new chemotype, different from other accessions in Malaysia, India and Indonesia that contain zerumbone in lower quantities [74,100,101]. The cyclic sesquiterpene was also detected in the methanolic extract in considerable amount (12.6 g) [102]. On the contrary, the essential oils resulting from other Alpinia spp. were richer in oxygenated monoterpenes and monoterpene hydrocarbons, such as 1,8-cineole, ocimenes, terpinen-4-ol and pinenes [103,104].
Zerumbone is also present in other genera included in Zingiberaceae family. The Boesenbergia genus is one of them [105]. In particular, the essential oil of B. quangngaiensis N.S. Ly, has good antibacterial and antifungal activity, and contains considerable concentration of zerumbone (11.4%) [75]. However, the chemical composition varies between species and parts of the plant. For example, zerumbone was absent in the leaf oil of other Boesenbergia spp., such as B. armeniaca and B. stenophylla, which were richer in other oxygenated sesquiterpenes (nerolidol, 42.55%), and oxygenated monoterpenes (linalool, 11.63%) and phenylpropanoids (methyl cinnamate, 83.17%), respectively [106]. Other species included in the genus Boesenbergia are also lacking in zerumbone, such as B. pandurata with mainly monoterpenes, mostly camphor (57.97%) [53]. The ethanolic extract of the Indonesian Etlingera acanthoides A.D. Poulsen also contains zerumbone [107], which accounts for 1.70% of the total essential oil. Together with the other main compounds, ar-turmerone, caryophyllene and caryophyllene oxide, zerumbone plays an important role in the antiviral activity of the plant [107].
Not only is zerumbone widespread in the Zingiberaceae family, but it represents one of the major elements of the essential oil of Cheilocostus speciosus (J. Konig) C. Specht [72,73]. Crape ginger, as it is popularly known, is a species belonging to the Costaceae family, closely related to Zingiberaceae. It is indigenous to peninsular Malaysia in southeast Asia from where it was introduced and naturalized in tropical and subtropical areas [108]. In general, phytochemicals extracted from different parts of this plant have shown numerous beneficial properties, including antioxidant, antidiabetic, antihypercholesterolemic, anticancer, hepatoprotective and anti-inflammatory effects [108]. In particular, the essential oil obtained from an Indian cultivar of C. speciosus showed anticancer activity, as well as considerable antibacterial capacity, against a wide range of Gram-positive (S. aureus, B. subtilis, S. faecalis and S. albus) and Gram-negative (E. coli, P. aeruginosa, Klebsiella aerogenes, P. vulgaris) bacteria. This feature is due to the synergism between the main components zerumbone (55.11%) and α-humulene (20.55%) [72]. Considering the data, it is thought that zerumbone is mainly localized in the rhizome of the plant species. Nevertheless, this monocyclic sesquiterpene has also been detected in different parts of other plants. Specifically, zerumbone was found to be one of the main components of the essential oil hydrodistilled from leaves of Amomum gagnepainii T.L.Wu, K.Larsen and Turland accounting for 16.4%, together with farnesyl acetate (18.5%) and β-caryophyllene (10.5%) [76].

3. Zerumbone Derivative

Zerumbone represents a point molecule for obtaining advantageous derivates. It is a sesquiterpenoid, specifically a cyclic (1E,4E,8E)-alpha-humulene ketone substituted with an oxo group at the carbon atom attached to two double bonds. This arrangement has been of central focus of research due to its unique reactivity in producing multiple structures as well as its numerous advantageous properties [36,109]. For years, researchers have examined and rearranged its structure with the aim of finding novel analogues with equal or greater activity than the original molecule [109,110,111,112]. This is the case with azazerumbone 2, a major Beckmann rearrangement product of E-zerumbone oxime with higher antibacterial and antimutagenic activity than zerumbone, zerumbone oxime, or zerumbol, possessing comparable or better bioactive attributes than the original sesquiterpene [112]. In addition, bio-isosteric replacements in the ortho position of the phenyl ring of zerumbone led to a series of new hydrazones, including 5a-f and 9a-f, with confirmed cytotoxic effects against three human cancer cell lines, HepG-2, SK-LU-1, and MCF-7 [111]. Another derivative, zerumbone epoxide, results from numerous reactions triggered by transition metal/Lewis acid-catalysed 1,4-conjugate addition of zerumbone [113].
In addition to the production of interesting derivatives, the study of zerumbone chemistry can help to expand the knowledge of sesquiterpenoids and their possible biosynthetic pathways [109]. Despite these findings, research on zerumbone production in suspension culture, as well as metabolic engineering, should be optimized to elucidate the biosynthetic pathway of this sesquiterpene and enhance the production of its derivatives [114].

4. Further Uses of Zerumbone

The beneficial properties of zerumbone extend to other areas, such as the agri-food industry [40] (Figure 2). For example, Keerthi et al. observed that the concentration of zerumbone increases in Z. zerumbet when infected with the rot-causative necrotrophic phytopathogen Pythium myriotylum, exhibiting natural resistance against the contamination [115] (Figure 2). Furthermore, this molecule has shown promising antifungal activity against the other phytopathogenic fungi Aspergillus spp. (MIC 150–170 ppm), Rhizoctonia solani (EC50 3906 ppm), Sclerotium rolfsii (EC50 59.3 ppm) and Macrophomina phaseolina (EC50 147.4 ppm) [40,64] (Figure 2). These plant-pathogenic fungi harm plants by secreting virulence factors such as effector proteins, degrading enzymes, toxins and growth regulators that severely affect crop yield and quality, and consequently have a negative impact on agriculture [116,117]. Amongst them, the essential oil of Z. zerumbet, rich in zerumbone (51.3%), has been especially effective against Aspergillus spp. [67]. The contamination of food and feed by Aspergillus has become a global issue with significant worldwide impact [118]. In particular, derived mycotoxins contaminate feed and food chains, being found in dairy products, dried fruits, vegetables and grain products, and representing a serious problem for public health [119,120].
In general, zerumbone and its analogues have demonstrated antibacterial and antimicrobial activities useful in food preservation [112]. B. quangngaiensis rhizome oil rich in zerumbone (11.4%) inhibited the growth of the spoilage bacteria P. aeruginosa, E. coli and S. enterica [75] (Figure 2). These microorganisms are able to contaminate meat, dairy and vegetable products, representing a considerable hazard for the agri-food industry and human health [121,122].
Curcuma-based botanicals have been evaluated for combating pests and pathogens of food crops [123]. Nowadays, they are emerging as sustainable alternatives to synthetic pesticides and natural protectants of crops [124,125,126]. In particular, the insecticidal activity of numerous essential oils has been confirmed [127,128]. Regarding our topic, the rhizome oil of C. speciosus, rich in zerumbone (38.6%), has exhibited interesting ingestion toxicity (LC50 17.16 µg/mL) and ovicidal activity (EC50 59.51 µg/mL) against the Old-World bollworm (Helicoverpa armigera (Lepidoptera:Noctuidae)), a polyphagous pest of increasing interest due to the rapid development of resistance to synthetic insecticides [73] (Figure 2). Furthermore, Wu et al. reported that the zerumbone analogue, α-caryophyllene (8.6%) exhibited higher insecticidal toxicity (LD50 13.1 µg/adult) against Lasioderma serricone, a beetle usually affecting perishable food products such as cereals and legumes, than the own zerumbone (40.2%) (LD50 = 42.4 µg/adult) [10].
Zerumbone has also been reported to have a herbicidal effect (Figure 2). Weeds represent the most important biotic threat to agricultural production worldwide, accounting for the highest losses in crop productivity. This situation is expected to worsen as climate change progresses and resistance increases. Natural and bio-herbicides represent one strategy to combat weeds with less chemical harmful products [129,130,131]. Specifically, the monocyclic sesquiterpene showed an inhibitory effect on weed seed growth against Phalaris minor with no, or low, effect on germination, seed root and shoot length of Triticum aestivum [40,64] (Figure 2).
The traditional uses of plant-based products have been studied over the years, and their numerous biological properties have been verified. In the field of agronomy, these compounds constitute economical, non-toxic and, consequently, environmentally friendly pesticides for sustainable agriculture. They represent a green alternative to synthetic pesticides, which are hazardous to living beings and the environment [132,133,134]. Among green chemicals, zerumbone, the main component of ginger plants, has demonstrated phytotoxic, antimicrobial, and insecticidal effects. Application of plant growth regulators (auxin and cytokine) with elicitors (methyl jasmonate and salicylic acid) to the fresh rhizome has been demonstrated to enhance the concentration of zerumbone and, consequently, its activity [135]. The food and feed preservative properties of zerumbone provide a basis for its use in the agri-food industry, to avoid contamination of harvests and stored functional food.

5. Formulation Strategies to Enhance the Solubility and Bioavailability of Zerumbone for Oral Administration

Even though the interesting properties of zerumbone have been widely demonstrated, zerumbone bioavailability and transport after oral administration is limited due mainly to its poor solubility. Zerumbone is freely soluble in organic solvents such as ethanol and dimethyl sulfoxide, while its solubility in water is about 1.296 mg/L at 25 °C. So, different approaches have been described during the years to improve the pharmacokinetics of low-water soluble drugs, such as particle size reduction, chemical modifications, amorphisation, salt formation, and through different strategies of formulation development such as solid dispersions, inclusion complexes, lipid-based formulations, micelles and nanoparticles [136,137]. These different formulation methods described in the literature to improve the solubility and bioavailability of zerumbone have recently been reviewed [138]. This section will, therefore, focus on updating the latest publications on the subject, with a particular focus on advances in oral administration. Additional information on the systems described in this section is summarized in Table 2.
One of the strategies explored by several researchers to improve the saturation solubility and dissolution profile of water-insoluble compounds such as zerumbone, is the formation of inclusion complexes with cyclodextrin derivatives [22,139]. However, it is worth noting that while the effectiveness of hydroxypropyl-β-cyclodextrin (HP-β-CyD) in improving zerumbone solubilisation and cytotoxic activity in human cancer cell lines has been reported by several authors, there is limited information on the usefulness of other cyclodextrin (CyD) derivatives in improving the pharmaceutical properties of zerumbone [139]. Recently, Hassan et al. evaluated various CyD derivatives, such as zerumbone solubilising agents, along with their release profile and cytotoxic activity [33]. Among the different CyDs tested, sulfobutylether (SBE)-β-CyD showed the highest solubilizing effect for zerumbone, followed by the dimethyl (DM)-β-CyD and methyl (M)-β-CyD. It is interesting to note that they showed a remarkable solubilizing effect beyond complexation with (HP)-β-CyDs, and its supersaturation was maintained for several hours. Supersaturating drug delivery systems have been acknowledged as a promising concept to obtain adequate oral bioavailability [143]. Regarding their cytotoxic activity, while an SBE-β-CyD/Zerumbone complex retained the activity of Zerumbone, M-β-CyD and DM-β-CyD showed toxicity compared to HP-β-CyD. Finally, to better understand the release mechanism of zerumbone from the different CyDs, crystallographic studies and molecular simulations were used to demonstrate the stronger interactions of SBE-β-CyD with zerumbone compared to HP-β-CyD (binding energy of −18.72 kcal/mol for SBE-β-CyD/Zerumbone vs. −16.24 kcal/mol for HP-β-CyD/Zerumbone).
Despite its promising pharmacological activities, preclinical and clinical studies of zerumbone have been limited due to its poor aqueous solubility, poor absorption, and low oral bioavailability. One example of these few studies involved the use of an inclusion complex of zerumbone with hydroxypropyl-β-cyclodextrin to treat acute and chronic gastritis. All patients of the vehicle group were injected with 20 µM doses of zerumbone over a 20-day period [22]. After evaluation through endoscopic studies, histopathological examinations, and urea breath tests, it was concluded that zerumbone treatment was quite effective in acute and chronic gastritis with or without H. pylori infection. This study once again confirms the strong potential of zerumbone as a therapeutic agent for various diseases, even including acute and chronic settings [4,144,145]. However, to our knowledge, no preclinical or clinical trials have been reported on the efficacy of formulations developed for the oral administration of zerumbone. Further work in this area, including investigations of the corresponding pharmacokinetic and pharmacodynamic data in humans, is therefore required.
Loading zerumbone into lipid carriers is another commonly reported approach to improve its solubility and delivery by facilitating transport across several anatomical barriers, thus avoiding mucosal toxicity. Specifically, nanostructure lipid carriers (NLC) have demonstrated highly advantageous characteristics, including low toxicity, controlled release, biodegradability, high drug loading, low cost, and easy upscaling [31,146,147]. Other proposed advantages of using NLC encapsulation as an oral delivery system are that they are transported via the lymphatic system avoiding first-pass metabolism, thus achieving a higher peak concentration after oral administration. In addition, NLC systems have shown greater permeability in the duodenum, jejunum, ileum and colon, improving absorption in the gastrointestinal tract post oral administration [148].
NLC–Zerumbone has been reported to possess in vitro cytotoxicity against several types of cancer cell, and no observed in vivo oral acute toxicity [31,140,149,150]. For example, Hosseinpour et al. evaluated the anticancer effect of zerumbone and zerumbone-loaded NLC on the human mammary gland adenocarcinoma (MDA-MB-231) cell line [147]. The efficacy of not-loaded and NLC-loaded zerumbone in suppressing the proliferation of MDA-MB-231 was confirmed by electron and fluorescent microscopy and flow cytometry (IC50 of 5.96 ± 0.13 and 6.01 ± 0.11 μg/mL for zerumbone and zerumbone-NLC, respectively). Furthermore, both systems significantly induced apoptosis via activation of caspase-9 and caspase-3 inhibition of antiapoptotic protein, and stimulation of proapoptotic protein expressions in a time-dependent manner. Further information on the in vivo effects of orally administrated NLC-zerumbone was reported in a study by Akhtar et al. in which the antitumor, immunostimulatory and anti-inflammatory effects of zerumbone and NLC-zerumbone in in vivo 4T1 challenged mice were compared [8]. NLC-zerumbone was found to further improve the efficacy of zerumbone, effectively controlling tumor growth and metastasis via delaying cancer cell cycle progression and apoptosis. In addition, zerumbone encapsulation was able to better contain the progression of 4T1 breast cancer by controlling cancer cell mitosis, inducing apoptosis, activating anti-tumor immunity, controlling inflammation in the tumor microenvironment, and delaying tumor metastasis.
Zeolite Y-gelatin composites have also been reported to be interesting oral delivery carriers of zerumbone, providing effective sustained release [142]. The ability to release the compound over an extended period of time at a controlled rate is generally more effective and safer, since it overcomes the possible toxicity associated with the drastic accumulation of the drug in the body over a short period. The design and preparation of cross-linked zeolite Y/gelatine composite was developed using glutaraldehyde as cross-linking agent targeting to prolong the duration of zerumbone released for 24 h [141]. The solid state of zerumbone after loading remained crystalline, ensuring no form changes during storage. Furthermore, cross-linking improved the thermal stability and the composite showed resistance to degradation under acidic and basic conditions (pH 1.2 and pH 7.4), confirming the potential of this system for oral administration. Despite the promising in vitro zerumbone controlled release, further in vivo antitumor evaluation of these matrices is necessary.

6. Conclusions

The agri-food industry requires natural pesticides, including essential oils and/or their main compounds, as substitutes for synthetic pesticides that have adverse effects on soils, water, wildlife, human health, and the environment. Natural compounds (monoterpenoids and sesquiterpenoids) present in essential oils could be used as preservatives in the food industry, with additional benefits due to their antioxidant properties, as well as their taste or aroma characteristics. Among the bioactive compounds, sesquiterpenoids have remarkable value due to their prominent presence in spices and herbs, and their wide range of activities due to their structural variability. In this sense, Z. zerumbet, whose main compound is zerumbone, is a commercialized spices widely consumed by humans. However, the bioavailability and biodegradation of these products are important properties influencing the half-life time and possible persistence problems in the human body. In this sense, zerumbone’s encapsulation represents an interesting alternative to solve the problems of solubility, oxidation, and other types of deterioration due to environmental factors such as light, temperature or humidity. Nonetheless, to date, practical applications of encapsulation and zerumbone-controlled delivery systems are scarce. Considering the relevance of this multifunctional sesquiterpene, further preclinical and clinical studies are needed to demonstrate and evaluate its in vivo efficacy, as well as formulation studies to fully understand its potential application as a therapeutic agent, green pesticide and/or food preservative.

Author Contributions

All authors contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kitayama, T. Attractive reactivity of a natural product, zerumbone. Biosci. Biotechnol. Biochem. 2011, 75, 199–207. [Google Scholar] [CrossRef] [Green Version]
  2. Padalia, R.C.; Verma, R.S.; Chauhan, A.; Singh, V.R.; Goswami, P.; Singh, S.; Verma, S.K.; Luqman, S.; Chanotiya, C.S.; Darokar, M.P. Zingiber zerumbet (L.) Roscoe ex Sm. from northern India: Potential source of zerumbone rich essential oil for antiproliferative and antibacterial applications. Ind. Crops Prod. 2018, 112, 749–754. [Google Scholar] [CrossRef]
  3. Moreira Da Silva, T.; Pinheiro, C.D.; Puccinelli Orlandi, P.; Pinheiro, C.C.; Soares Pontes, G. Zerumbone from Zingiber zerumbet (L.) smith: A potential prophylactic and therapeutic agent against the cariogenic bacterium Streptococcus mutans. BMC Complement. Altern. Med. 2018, 18, 301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Girisa, S.; Shabnam, B.; Monisha, J.; Fan, L.; Halim, C.E.; Arfuso, F.; Ahn, K.S.; Sethi, G.; Kunnumakkara, A.B. Potential of zerumbone as an anti-cancer agent. Molecules 2019, 24, 734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Deng, M.; Yun, X.; Ren, S.; Qing, Z.; Luo, F. Plants of the genus Zingiber: A review of their ethnomedicine, phytochemistry and pharmacology. Molecules 2022, 27, 2826. [Google Scholar] [CrossRef] [PubMed]
  6. Tian, M.; Wu, X.; Hong, Y.; Wang, H.; Deng, G.; Zhou, Y. Comparison of chemical composition and bioactivities of essential oils from fresh and dry rhizomes of Zingiber zerumbet (L.) Smith. Biomed Res. Int. 2020, 2020, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Sharifi-Rad, M.; Varoni, E.M.; Salehi, B.; Sharifi-Rad, J.; Matthews, K.R.; Ayatollahi, S.A.; Kobarfard, F.; Ibrahim, S.A.; Mnayer, D.; Zakaria, Z.A.; et al. Plants of the genus Zingiber as a source of bioactive phytochemicals: From tradition to pharmacy. Molecules 2017, 22, 2145. [Google Scholar] [CrossRef] [Green Version]
  8. Akhtar, N.M.Y.; Jantan, I.; Arshad, L.; Haque, M.A. Standardized ethanol extract, essential oil and zerumbone of Zingiber zerumbet rhizome suppress phagocytic activity of human neutrophils. BMC Complement. Altern. Med. 2019, 19, 331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Rana, V.S.; Verdeguer, M.; Blazquez, M.A. Chemical composition of the essential oil of Zingiber zerumbet var darcyi. Nat. Prod. Commun. 2012, 7, 1369–1370. [Google Scholar] [CrossRef] [Green Version]
  10. Wu, Y.; Guo, S.S.; Huang, D.Y.; Wang, C.F.; Wei, J.Y.; Li, Z.H.; Sun, J.S.; Bai, J.F.; Tian, Z.F.; Wang, P.J.; et al. Contact and repellant activities of zerumbone and its analogues from the essential oil of Zingiber zerumbet (L.) smith against Lasioderma serricorne. J. Oleo Sci. 2017, 66, 399–405. [Google Scholar] [CrossRef]
  11. Veena, K.S.; Gopalan, G.; Madhukrishnan, M.; Varughese, S.; Radhakrishnan, K.V.; Lankalapalli, R.S. Putative biomimetic route to 8-oxabicyclo[3.2.1]octane motif from a humulene sesquiterpenoid zerumbone. Org. Lett. 2020, 22, 6409–6413. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, C.; Liu, J.; Zhao, F.; Lu, C.; Zhao, G.R.; Lu, W. Production of sesquiterpenoid zerumbone from metabolic engineered Saccharomyces cerevisiae. Metab. Eng. 2018, 49, 28–35. [Google Scholar] [CrossRef] [PubMed]
  13. Albaayit, S.F.A.; Maharjan, R.; Abdullah, R.; Noor, M.H.M. Evaluation of anti-methicillin-resistant Staphylococcus aureus property of zerumbone. J. Appl. Biomed. 2022, 20, 15–21. [Google Scholar] [CrossRef] [PubMed]
  14. Woo, H.J.; Yang, J.Y.; Lee, P.; Kim, J.B.; Kim, S.H. Zerumbone inhibits Helicobacter pylori urease activity. Molecules 2021, 26, 2663. [Google Scholar] [CrossRef]
  15. Li, J.; Wang, L.; Sun, Y.; Wang, Z.; Qian, Y.; Duraisamy, V.; Antary, T.M.A.I. Zerumbone-induced reactive oxygen species-mediated oxidative stress re-sensitizes breast cancer cells to paclitaxel. Biotechnol. Appl. Biochem. 2022. [Google Scholar] [CrossRef] [PubMed]
  16. Ashraf, S.M.; Sebastian, J.; Rathinasamy, K. Zerumbone, a cyclic sesquiterpene, exerts antimitotic activity in HeLa cells through tubulin binding and exhibits synergistic activity with vinblastine and paclitaxel. Cell Prolif. 2019, 52, e12558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Schnoell, J.; Stanisz, I.; Jank, B.J.; Stanek, V.; Schmid, R.; Brunner, M.; Heiduschka, G.; Kotowski, U. Zerumbone acts as a radiosensitizer in head and neck squamous cell carcinoma. Investig. New Drugs 2022, 40, 224–231. [Google Scholar] [CrossRef]
  18. Kim, H.R.; Rhee, K.J.; Eom, Y. Bin Anti-biofilm and antimicrobial effects of zerumbone against Bacteroides fragilis. Anaerobe 2019, 57, 99–106. [Google Scholar] [CrossRef]
  19. Kim, H.R.; Shin, D.S.; Jang, H.I.; Eom, Y.B. Anti-biofilm and anti-virulence effects of zerumbone against Acinetobacter baumannii. Microbiol. 2020, 166, 717–726. [Google Scholar] [CrossRef]
  20. Shin, D.S.; Eom, Y.B. Efficacy of zerumbone against dual-species biofilms of Candida albicans and Staphylococcus aureus. Microb. Pathog. 2019, 137, 103768. [Google Scholar] [CrossRef] [PubMed]
  21. Cho, H.; Rhee, K.; Eom, Y. Composition in ETBF Colonized AOM/DSS. J. Microbiol. Biotechnol. 2020, 30, 1640–1650. [Google Scholar] [CrossRef] [PubMed]
  22. Li, L.; Kong, L.; Song, H. The therapeutic effect of zerumbone on chronic gastritis via antioxidant mechanisms. Exp. Ther. Med. 2017, 14, 2505–2510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zulazmi, N.A.; Gopalsamy, B.; Omar Farouk, A.A.; Sulaiman, M.R.; Bharatham, B.H.; Perimal, E.K. Antiallodynic and antihyperalgesic effects of zerumbone on a mouse model of chronic constriction injury-induced neuropathic pain. Fitoterapia 2015, 105, 215–221. [Google Scholar] [CrossRef] [PubMed]
  24. Chia, J.S.M.; Omar Farouk, A.A.; Mohamad, A.S.; Sulaiman, M.R.; Perimal, E.K. Zerumbone alleviates chronic constriction injury-induced allodynia and hyperalgesia through serotonin 5-HT receptors. Biomed. Pharmacother. 2016, 83, 1303–1310. [Google Scholar] [CrossRef]
  25. Gopalsamy, B.; Chia, J.S.M.; Farouk, A.A.O.; Sulaiman, M.R.; Perimal, E.K. Zerumbone-induced analgesia modulated via potassium channels and opioid receptors in chronic constriction injury-induced neuropathic pain. Molecules 2020, 25, 3880. [Google Scholar] [CrossRef]
  26. Chia, J.S.M.; Farouk, A.A.O.; Mohamad, T.A.S.T.; Sulaiman, M.R.; Zakaria, H.; Hassan, N.I.; Perimal, E.K. Zerumbone ameliorates neuropathic pain symptoms via cannabinoid and ppar receptors using in vivo and in silico models. Molecules 2021, 26, 3849. [Google Scholar] [CrossRef]
  27. Liu, W.Y.; Tzeng, T.F.; Liu, I.M. Zerumbone, a bioactive sesquiterpene, ameliorates diabetes-induced retinal microvascular damage through inhibition of phospho-p38 mitogen-activated protein kinase and nuclear factor-κB pathways. Molecules 2016, 21, 1708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Hwang, J.; Youn, K.; Ji, Y.; Lee, S.; Lim, G. Biological and computational studies for dual cholinesterases inhibitory effect of zerumbone. Nutrients 2020, 12, 1215. [Google Scholar] [CrossRef] [PubMed]
  29. Mohamad, N.E.; Abu, N.; Rahman, H.S.; Ky, H.; Ho, W.Y.; Lim, K.L.; How, C.W.; Rasedee, A.; Alitheen, N.B.; Yeap, S.K. Nanostructured lipid carrier improved in vivo anti-tumor and immunomodulatory effect of zerumbone in 4T1 challenged mice. RSC Adv. 2015, 5, 22066–22074. [Google Scholar] [CrossRef]
  30. Albaayit, S.F.A.; Rasedee, A.; Abdullah, N. Zerumbone-loaded nanostructured lipid carrier gel facilitates wound healing in rats. Rev. Bras. Farmacogn. 2020, 30, 272–278. [Google Scholar] [CrossRef]
  31. Rahman, H.S.; Rasedee, A.; How, C.W.; Abdul, A.B.; Zeenathul, N.A.; Othman, H.H.; Saeed, M.I.; Yeap, S.K. Zerumbone-loaded nanostructured lipid carriers: Preparation, characterization, and antileukemic effect. Int. J. Nanomed. 2013, 8, 2769–2781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Md, S.; Kit, B.C.M.; Jagdish, S.; David, D.J.P.; Pandey, M.; Chatterjee, L.A. Development and in vitro evaluation of a zerumbone loaded nanosuspension drug delivery system. Crystals 2018, 8, 286. [Google Scholar]
  33. Hassan, M.M.; Mohammed, A.F.A.; Elamin, K.M.; Devkota, H.P.; Ohno, Y.; Motoyama, K.; Higashi, T.; Imai, T. Improvement of pharmaceutical properties of zerumbone, a multifunctional compound, using cyclodextrin derivatives. Chem. Pharm. Bull. 2020, 68, 1117–1120. [Google Scholar] [CrossRef] [PubMed]
  34. Li, P.; Shu, X. Zerumbone with hydroxypropyl-β-cyclodextrin inclusion complex as a potential treatment for gastritis. Int. J. Pharmacol. 2017, 13, 1071–1078. [Google Scholar] [CrossRef] [Green Version]
  35. Zakaria, Z.A.; Yob, N.J.; Jofrry, S.M.; Affandi, M.M.R.M.M.; Teh, L.K.; Salleh, M.Z. Zingiber zerumbet (L.) Smith: A review of its ethnomedicinal, chemical, and pharmacological uses. Evid. Based Complement. Altern. Med. 2011, 2011, 543216. [Google Scholar]
  36. Koga, A.Y.; Beltrame, F.L.; Pereira, A.V. Several aspects of Zingiber zerumbet: A review. Rev. Bras. Farmacogn. 2016, 26, 385–391. [Google Scholar] [CrossRef]
  37. Datiles, M.J.; Acevedo-Rodríguez, P. Datasheet Report for Zingiber zerumbet (Shampoo Ginger). Available online: https://www.cabi.org/isc/datasheet/57539 (accessed on 21 July 2022).
  38. Chane-Ming, J.; Vera, R.; Chalchat, J.C. Chemical composition of the essential oil from rhizomes, leaves and flowers of Zingiber zerumbet smith from Reunion Island. J. Essent. Oil Res. 2003, 15, 202–205. [Google Scholar] [CrossRef]
  39. Azelan, N.A. Optimization of Hydrodistillation Extraction Conditions of Zerumbone from Zingiber zerumbet Rhizome. Master’s Thesis, Universiti Teknologi Malaysia, Johor, Malaysia, 2017. [Google Scholar]
  40. Rana, V.S.; Ahluwalia, V.; Shakil, N.A.; Prasad, L. Essential oil composition, antifungal, and seedling growth inhibitory effects of zerumbone from Zingiber zerumbet Smith. J. Essent. Oil Res. 2017, 29, 320–329. [Google Scholar] [CrossRef]
  41. Raina, A.P.; Misra, R.C. Zingiber zerumbet (L.) Roscoe ex Smith: A potential source of zerumbone rich essential oil. Med. Plants-Int. J. Phytomedicines Relat. Ind. 2022, 14, 162–168. [Google Scholar]
  42. Chavan, J.J.; Kshirsagar, P.R.; Pai, S.R.; Pawar, N.V. Micropropagation, metabolite profiling, antioxidant activities and chromatographic determination of bioactive molecules across in vitro conditions and subsequent field cultivation stages of ‘Shampoo Ginger’ (Zingiber zerumbet L. Roscoe ex Sm). Biocatal. Agric. Biotechnol. 2018, 16, 79–89. [Google Scholar] [CrossRef]
  43. Dash, B.; Sahoo, A.; Ray, A.; Jena, S.; Nayak, S. Identification of chemical constituents of Zingiber zerumbet rhizome extract using GC/MS. J. Biol. Act. Prod. Nat. 2020, 10, 411–417. [Google Scholar] [CrossRef]
  44. Lal, M.; Begum, T.; Munda, S.; Pandey, S.K. Identification of high rhizome and essential oil yielding variety (Jor Lab ZB-103) of Zingiber zerumbet (L.) Roscoe ex Sm. J. Essent. Oil-Bear. Plants 2021, 24, 1010–1025. [Google Scholar] [CrossRef]
  45. Dash, B.; Ray, A.; Sahoo, A.; Kar, B.; Chatterjee, T.; Halder, T.; Chandra Panda, P.; Patnaik, J.; Ghosh, B.; Nayak, S. A combined approach using ISSR and volatile compound analysis for assessment of genetic and phytochemical diversity in Zingiber zerumbet (L.) from Eastern India. J. Essent. Oil-Bear. Plants 2019, 22, 31–49. [Google Scholar] [CrossRef]
  46. Baby, S.; Dan, M.; Thaha, A.R.M.; Johnson, A.J.; Kurup, R.; Balakrishnapillai, P.; Lim, C.K. High content of zerumbone in volatile oils of Zingiber zerumbet from southern India and Malaysia. Flavour Fragr. J. 2009, 24, 301–308. [Google Scholar] [CrossRef]
  47. Bhavya, M.L.; Ravi, R.; Madhava Naidu, M. Development and validation of headspace Solid-Phase microextraction coupled with gas chromatography (HS-SPME-GC) method for the analysis of Zingiber zerumbet L. Nat. Prod. Res. 2019, 35, 1221–1225. [Google Scholar] [CrossRef]
  48. Ghasemzadeh, A.; Jaafar, H.Z.E.; Rahmat, A.; Swamy, M.K. Optimization of microwave-assisted extraction of zerumbone from Zingiber zerumbet L. rhizome and evaluation of antiproliferative activity of optimized extracts. Chem. Cent. J. 2017, 11, 5. [Google Scholar] [CrossRef] [Green Version]
  49. Paramita, A.; Wibowo, I.; Insanu, M. A study of genus Zingiber: The role of condiments in science. Curr. Res. Biosci. Biotechnol. 2021, 3, 186–195. [Google Scholar] [CrossRef]
  50. Sivasothy, Y.; Awang, K.; Ibrahim, H.; Thong, K.L.; Fitrah, N.; Koh, X.P.; Tan, L.K. Chemical composition and antibacterial activities of essential oils from Zingiber spectabile Griff. J. Essent. Oil Res. 2012, 24, 305–313. [Google Scholar] [CrossRef]
  51. Sirat, H.M.; Leh, N.H.N. The rhizome oil of Zingiber spectabile valet. J. Essent. Oil Res. 2001, 13, 256–257. [Google Scholar] [CrossRef]
  52. Mahardika, R.W.; Ibrahim, H.; Nurulhusna, A.H.; Awang, K. Efficacy of four species of Zingiberaceae extract against vectors of dengue, chikungunya and filariasis. Trop. Biomed. 2017, 34, 375–387. [Google Scholar]
  53. Sukari, M.A.; Mohd Sharif, N.W.; Yap, A.L.C.; Tang, S.W.; Neoh, B.K.; Rahmani, M.; Ee, G.C.L.; Taufiq-Yap, Y.H.; Yusof, U.K. Chemical constituent variations of essential oils from rhizomes of four Zingiberaceae species. Malays. J. Anal. Sci. 2008, 12, 638–644. [Google Scholar]
  54. Ariani, S.R.D.; Nurul, S.; Falentina, S. Isolation and identification of essential oils from bitter ginger (Zingiber amaricans BL.), fragrant ginger (Zingiber aromaticum Val.) and shampoo ginger (Zingiber zerumbet (L.) Smith) rhizomes grown in Jumapolo Karaganyar Central Java Indonesia. J. Kim. Dan Pendidik. Kim. 2021, 6, 242–251. [Google Scholar] [CrossRef]
  55. Devkota, H.P.; Paudel, K.R.; Hassan, M.M.; Dirar, A.I.; Das, N.; Adhikari-Devkota, A.; Echeverría, J.; Logesh, R.; Jha, N.K.; Singh, S.K.; et al. Bioactive compounds from Zingiber montanum and their pharmacological activities with focus on zerumbone. Appl. Sci. 2021, 11, 10205. [Google Scholar] [CrossRef]
  56. Bora, P.K.; Saikia, J.; Kemprai, P.; Saikia, S.P.; Banik, D.; Haldar, S. Evaluation of postharvest drying, key odorants, and phytotoxins in Plai (Zingiber montanum) essential oil. J. Agric. Food Chem. 2021, 69, 5500–5509. [Google Scholar] [CrossRef]
  57. Rajkumari, S.; Sanatombi, K. Secondary metabolites content and essential oil composition of in vitro cultures of Zingiber montanum (Koenig) Link ex A. Dietr. Biotechnol. Lett. 2020, 42, 1237–1245. [Google Scholar] [CrossRef] [PubMed]
  58. Hassan, M.M.; Adhikari-Devkota, A.; Imai, T.; Devkota, H.P. Zerumbone and kaempferol derivatives from the rhizomes of Zingiber montanum (J. Koenig) Link ex A. Dietr. from Bangladesh. Separations 2019, 6, 31. [Google Scholar] [CrossRef] [Green Version]
  59. Thitinarongwate, W.; Mektrirat, R.; Nimlamool, W.; Khonsung, P.; Pikulkaew, S.; Okonogi, S.; Kunanusorn, P. Phytochemical and safety evaluations of Zingiber ottensii Valeton essential oil in zebrafish embryos and rats. Toxics 2021, 9, 102. [Google Scholar] [CrossRef]
  60. Ruttanapattanakul, J.; Wikan, N.; Chinda, K.; Jearanaikulvanich, T.; Krisanuruks, N.; Muangcha, M.; Okonogi, S.; Potikanond, S.; Nimlamool, W. Essential oil from Zingiber ottensii induces human cervical cancer cell apoptosis and inhibits MAPK and PI3K/AKT signaling cascades. Plants 2021, 10, 1419. [Google Scholar] [CrossRef]
  61. Panyajai, P.; Chueahongthong, F.; Viriyaadhammaa, N.; Nirachonkul, W.; Tima, S.; Chiampanichayakul, S.; Anuchapreeda, S.; Okonogi, S. Anticancer activity of Zingiber ottensii essential oil and its nanoformulations. PLoS ONE 2022, 17, e0262335. [Google Scholar] [CrossRef]
  62. Zakaria, Z.A.; Mohamad, A.S.; Ahmad, M.S.; Mokhtar, A.F.; Israf, D.A.; Lajis, N.H.; Sulaiman, M.R. Preliminary analysis of the anti-inflammatory activity of essential oils of Zingiber Zerumbet. Biol. Res. Nurs. 2011, 13, 425–432. [Google Scholar] [CrossRef] [PubMed]
  63. Batubara, I.; Suparto, I.H.; Sadiah, S.; Matsuoka, R.; Mitsunaga, T. Effect of Zingiber zerumbet essential oils and zerumbone inhalation on body weight of Sprague dawley rat. Pak. J. Biol. Sci. 2013, 16, 1028–1033. [Google Scholar] [CrossRef] [Green Version]
  64. Madegowda, B.H.; Rameshwaran, P.; Nagaraju, N.P.; Murthy, P.S. In-vitro mycological activity of essential oil from Zingiber zerumbet rhizomes. J. Essent. Oil Res. 2015, 28, 81–88. [Google Scholar] [CrossRef]
  65. Sulaiman, M.R.; Tengku Mohamad, T.A.S.; Shaik Mossadeq, W.M.; Moin, S.; Yusof, M.; Mokhtar, A.F.; Zakaria, Z.A.; Israf, D.A.; Lajis, N. Antinociceptive activity of the essential oil of Zingiber Zerumbet. Planta Med. 2010, 76, 107–112. [Google Scholar] [CrossRef]
  66. Fusi, F.; Durante, M.; Sgaragli, G.; Khanh, P.N.; Son, N.T.; Huong, T.T.; Huong, V.N.; Cuong, N.M. In vitro vasoactivity of zerumbone from Zingiber Zerumbet. Planta Med. 2015, 81, 298–304. [Google Scholar] [CrossRef] [PubMed]
  67. Huong, L.T.; Chinh, H.V.; An, N.T.G.; Viet, N.T.; Hung, N.H.; Thuong, N.T.H.; Giwa-Ajeniya, A.O.; Ogunwande, I.A. Zingiber zerumbet rhizome essential oil: Chemical composition, antimicrobial and mosquito larvicidal activities. Eur. J. Med. Plants 2020, 3, 2019. [Google Scholar] [CrossRef]
  68. Sithara, T.; Dhanya, B.P.; Arun, K.B.; Sini, S.; Dan, M.; Kokkuvayil Vasu, R.; Nisha, P. Zerumbone, a cyclic sesquiterpene from Zingiber zerumbet induces apoptosis, cell cycle arrest, and antimigratory effects in SW480 colorectal cancer cells. J. Agric. Food Chem. 2018, 66, 602–612. [Google Scholar] [CrossRef] [PubMed]
  69. Jantan, I.; Haque, M.A.; Ilangkovan, M.; Arshad, L. Zerumbone from Zingiber zerumbet inhibits innate and adaptive immune responses in Balb/C mice. Int. Immunopharmacol. 2019, 73, 552–559. [Google Scholar] [CrossRef]
  70. Zhang, L.; Yang, Z.; Wei, J.; Su, P.; Chen, D.; Pan, W.; Zhou, W.; Zhang, K.; Zheng, X.; Lin, L.; et al. Contrastive analysis of chemical composition of essential oil from twelve Curcuma species distributed in China. Ind. Crops Prod. 2017, 108, 17–25. [Google Scholar] [CrossRef]
  71. Xiang, H.; Zhang, L.; Xi, L.; Yang, Y.; Wang, X.; Lei, D.; Zheng, X.; Liu, X. Phytochemical profiles and bioactivities of essential oils extracted from seven Curcuma herbs. Ind. Crops Prod. 2018, 111, 298–305. [Google Scholar] [CrossRef]
  72. Thambi, M.; Shafi, M.P. Rhizome essential oil composition of Costus speciosus and its antimicrobial properties. Int. J. Pharm. Res. Allied Sci. 2015, 4, 28–32. [Google Scholar]
  73. Benelli, G.; Govindarajan, M.; Rajeswary, M.; Vaseeharan, B.; Alyahya, S.A.; Alharbi, N.S.; Kadaikunnan, S.; Khaled, J.M.; Maggi, F. Insecticidal activity of camphene, zerumbone and α-humulene from Cheilocostus speciosus rhizome essential oil against the Old-World bollworm, Helicoverpa armigera. Ecotoxicol. Environ. Saf. 2018, 148, 781–786. [Google Scholar] [CrossRef] [PubMed]
  74. Arambewela, L.S.R.; Arawwawala, M.; Owen, N.L.; Jarvis, B. Volatile oil of Alpinia galanga Willd. of Sri Lanka. J. Essent. Oil Res. 2007, 19, 455–456. [Google Scholar] [CrossRef]
  75. Huong, L.T.; Sam, L.N.; Dai, D.N.; Ogunwande, I.A. Investigation into the chemical compositions and antimicrobial activity of essential oil from the rhizomes of Boesenbergia quangngaiensis N.S. Lý from Vietnam. J. Essent. Oil-Bear. Plants 2021, 24, 1125–1133. [Google Scholar] [CrossRef]
  76. Huong, L.T.; Hung, N.V.; Chung, M.V.; Dai, D.N.; Ogunwande, I.A. Essential oils constituents of the leaves of Amomum gagnepainii and Amomum repoense. Nat. Prod. Res. 2018, 32, 316–321. [Google Scholar] [CrossRef] [PubMed]
  77. Chang, K.S.; Jeon, J.H.; Kim, G.H.; Jang, C.W.; Jeong, S.J.; Ju, Y.R.; Ahn, Y.J. Repellency of zerumbone identified in Cyperus rotundus rhizome and other constituents to Blattella germanica. Sci. Rep. 2017, 7, 16643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Huong, L.T.; Chung, N.T.; Huong, T.T.; Sam, L.N.; Hung, N.H.; Ogunwande, I.A.; Dai, D.N.; Linh, L.D.; Setzer, W.N. Essential oils of Zingiber species from Vietnam: Chemical compositions and biological activities. Plants 2020, 9, 1269. [Google Scholar] [CrossRef]
  79. Bin Jantan, I.; Mohd Yassin, M.S.; Chin, C.B.; Chen, L.L.; Sim, N.L. Antifungal activity of the essential oils of nine Zingiberaceae species. Pharm. Biol. 2003, 41, 392–397. [Google Scholar] [CrossRef]
  80. Bucur, L.; Ionus, E.; Moise, G.; Gîrd, C.; Schröder, V. GC-MS analysis and bioactive properties of Zingiberis rhizoma essential oil. Farmacia 2020, 68, 280–287. [Google Scholar] [CrossRef]
  81. Chairgulprasert, V.; Prasertsongskun, S.; Wichaporn, W. Chemical constituents of the essential oil and anti-bacterial activity of Zingiber wrayi var. halabala. Songklanakarin J. Sci. Technol. 2005, 27, 813–818. [Google Scholar]
  82. Pintatum, A.; Laphookhieo, S.; Maneerat, W.; Logie, E.; Berghe, W. Vanden Chemical composition of essential oils from different parts of Zingiber kerrii Craib and their antibacterial, antioxidant, and tyrosinase inhibitory activities. Biomolecules 2020, 10, 228. [Google Scholar] [CrossRef] [Green Version]
  83. Shukurova, M.K.; Asikin, Y.; Chen, Y.; Kusano, M.; Watanabe, K.N. Profiling of volatile organic compounds in wild indigenous medicinal ginger (Zingiber barbatum Wall.) from Myanmar. Metabolites 2020, 10, 248. [Google Scholar] [CrossRef] [PubMed]
  84. Mahboubi, M. Zingiber officinale Rosc. essential oil, a review on its composition and bioactivity. Clin. Phytosci. 2019, 5, 6. [Google Scholar] [CrossRef] [Green Version]
  85. Sommano, S.R.; Tangpao, T. Aromatic profile of rhizomes from the ginger family used in food. In Aromatic Herbs in Food. Bioactive Compounds, Processing, and Applications; Galanakis, C.M., Ed.; Elsevier: London, UK, 2021; pp. 123–165. ISBN 9780128227169. [Google Scholar]
  86. CAB. Curcuma. Available online: https://www.cabi.org/ISC/datasheet/17013 (accessed on 23 July 2022).
  87. Kristianto, S.; Isrianto, P.L.; Marmi, M.; Chamidah, D.; Khan, A.U. Screening phytochemical and study insilico of family Zingiberaceae as anti-inflammatory. J. Biota 2022, 8, 95–114. [Google Scholar] [CrossRef]
  88. Peng, W.; Li, P.; Ling, R.; Wang, Z.; Feng, X.; Liu, J.; Yang, Q.; Yan, J. Diversity of volatile compounds in ten varieties of Zingiberaceae. Molecules 2022, 27, 565. [Google Scholar] [CrossRef]
  89. Abdelkader, N.F.; Moustafa, P.E. Rhizomatous plants: Curcuma longa and Zingiber officinale in affording immunity. In Plants and Phytomolecules for Immunomodulation; Sangwan, N.S., Farag, M.A., Modolo, L.V., Eds.; Springer Nature: Singapore, 2022; pp. 361–390. ISBN 9789811681172. [Google Scholar]
  90. Kshirsagar, P.R.; Pawar, N.V.; Patil, S.P.; Nimbalkar, M.S.; Pai, S.R. Quantification of zerumbone in eleven accessions of Curcuma longa using RP-HPLC. Indian J. Pharm. Sci. 2018, 80, 1159–1165. [Google Scholar] [CrossRef]
  91. Duong, L.; Mentreddy, S.R.; Satyal, R.; Satyal, P.; Setzer, W.N. Essential oil chemotypes of four vietnamese Curcuma species cultivated in North Alabama. Horticulturae 2022, 8, 360. [Google Scholar] [CrossRef]
  92. Jena, S.; Ray, A.; Sahoo, A.; Panda, P.C.; Nayak, S. Deeper insight into the volatile profile of essential oil of two Curcuma species and their antioxidant and antimicrobial activities. Ind. Crops Prod. 2020, 155, 112830. [Google Scholar] [CrossRef]
  93. Gharge, S.; Hiremath, S.I.; Kagawad, P.; Jivaje, K.; Palled, M.S.; Suryawanshi, S.S. Curcuma zedoaria Rosc (Zingiberaceae): A review on its chemical, pharmacological and biological activities. Future J. Pharm. Sci. 2021, 7, 166. [Google Scholar] [CrossRef]
  94. Hamdi, O.A.A.; Ye, L.J.; Kamarudin, M.N.A.; Hazni, H.; Paydar, M.; Looi, C.Y.; Shilpi, J.A.; Kadir, H.A.; Awang, K. Neuroprotective and antioxidant constituents from Curcuma zedoaria rhizomes. Rec. Nat. Prod. 2015, 9, 349–355. [Google Scholar]
  95. Al-Amin, M.; Eltayeb, N.M.; Khairuddean, M.; Salhimi, S.M. Bioactive chemical constituents from Curcuma caesia Roxb. rhizomes and inhibitory effect of curcuzederone on the migration of triple-negative breast cancer cell line MDA-MB-231. Nat. Prod. Res. 2021, 35, 3166–3170. [Google Scholar] [CrossRef] [PubMed]
  96. Rangsiruji, A.; Newman, M.F.; Cronk, Q.C.B. Origin and relationships of Alpinia galanga (Zingiberaceae) based on molecular data. Edinb. J. Bot. 2000, 57, 9–37. [Google Scholar] [CrossRef]
  97. Kress, W.J.; Liu, A.Z.; Newman, M.; Qing-Jun, L.I. The molecular phylogeny of Alpinia (Zingiberaceae): A complex and polyphyletic genus of gingers. Am. J. Bot. 2005, 92, 167–178. [Google Scholar] [CrossRef] [PubMed]
  98. Van, H.T.; Thang, T.D.; Luu, T.N.; Doan, V.D. An overview of the chemical composition and biological activities of essential oils from: Alpinia genus (Zingiberaceae). RSC Adv. 2021, 11, 37767–37783. [Google Scholar] [CrossRef] [PubMed]
  99. Ramanunny, A.K.; Wadhwa, S.; Gulati, M.; Vishwas, S.; Khursheed, R.; Paudel, K.R.; Gupta, S.; Porwal, O.; Alshahrani, S.M.; Jha, N.K.; et al. Journey of Alpinia galanga from kitchen spice to nutraceutical to folk medicine to nanomedicine. J. Ethnopharmacol. 2022, 291, 115144. [Google Scholar] [CrossRef]
  100. Mallavarapu, G.R.; Rao, L.; Ramesh, S.; Dimri, B.P.; Bhaskaruni, R.; Rajeswara, R.; Kaul, P.N.; Bhattacharya, A.K. Composition of the volatile oils of Alpina galanga rhizomes and leaves from India. J. Essent. Oil Res. 2002, 14, 397–399. [Google Scholar] [CrossRef]
  101. Rana, V.S.; Verdeguer, M.; Blazquez, M.A. GC and GC/MS analysis of the volatile constituents of the oils of Alpinia galanga (L.) Willd and A. officinarum Hance rhizomes. J. Essent. Oil Res. 2010, 22, 521–524. [Google Scholar] [CrossRef]
  102. Huang, G.C.; Kao, C.L.; Li, W.J.; Huang, S.T.; Li, H.T.; Chen, C.Y. A new phenylalkanoid from the rhizomes of Alpinia galanga. Chem. Nat. Compd. 2018, 54, 1072–1075. [Google Scholar] [CrossRef]
  103. Bhatt, G.; Nagarkoti, R.S.; Kholiya, S.; Tiwari, A.; Verma, S.K.; Verma, R.S.; Darokar, M.P.; Padalia, R.C. Chemical and antibacterial activity evaluation of Alpinia calcarata and Alpinia zerumbet grown in foothills agroclimatic conditions of Northern India. Open Bioact. Compd. J. 2021, 9, 15–19. [Google Scholar] [CrossRef]
  104. Feng, Y.X.; Zhang, X.; Wang, Y.; Chen, Z.Y.; Lu, X.X.; Du, Y.S.; Du, S.S. The potential contribution of cymene isomers to insecticidal and repellent activities of the essential oil from Alpinia zerumbet. Int. Biodeterior. Biodegrad. 2021, 157, 105138. [Google Scholar] [CrossRef]
  105. Vanijajiva, O.; Sirirugsa, P.; Suvachittanont, W. Confirmation of relationships among Boesenbergia (Zingiberaceae) and related genera by RAPD. Biochem. Syst. Ecol. 2005, 33, 159–170. [Google Scholar] [CrossRef]
  106. Nor, N.A.; Ibrahim, H. Chemical constituents of essential oils of Boesenbergia armeniaca. Pak. J. Biotechnol. 2018, 50, 1917–1922. [Google Scholar]
  107. Zubair, M.S.; Khairunisa, S.Q.; Widodo, A.; Nasronudin; Pitopang, R. Antiviral screening on Alpinia eremochlamys, Etlingera flexuosa, and Etlingera acanthoides extracts against HIV-infected MT-4 cells. Heliyon 2021, 7, e06710. [Google Scholar] [CrossRef] [PubMed]
  108. Lim, T.K. Edible Medicinal and Non-Medicinal Plants; Springer: Dordrecht, The Netherlands, 2014; Volume 7, Flowers; ISBN 978-94-007-7394-3. [Google Scholar]
  109. Kashiwazaki, G.; Watanabe, R.; Tsuzuki, T.; Yamamoto, C.; Nishikawa, A.; Ohtomo, S.; Yoshikawa, T.; Kitamura, Y.; Utaka, Y.; Kawai, Y.; et al. Brønsted acid-induced transannulation of the phytochemical zerumbone. Org. Biomol. Chem. 2021, 19, 10444–10454. [Google Scholar] [CrossRef]
  110. Kitayama, T.; Okamoto, T.; Hill, R.K.; Kawai, Y.; Takahashi, S.; Yonemori, S.; Yamamoto, Y.; Ohe, K.; Uemura, S.; Sawada, S. Chemistry of Zerumbone. 1. Simplified isolation, conjugate addition reactions, and a unique ring contracting transannular reaction of its dibromide. J. Org. Chem. 1999, 64, 2667–2672. [Google Scholar] [CrossRef] [PubMed]
  111. Vu, V.V.; Vu, T.K. Synthesis of new zerumbone hydrazones and their in-vitro anticancer activity. Curr. Bioact. Compd. 2020, 17, 331–338. [Google Scholar] [CrossRef]
  112. Santosh Kumar, S.C.; Srinivas, P.; Negi, P.S.; Bettadaiah, B.K. Antibacterial and antimutagenic activities of novel zerumbone analogues. Food Chem. 2013, 141, 1097–1103. [Google Scholar] [CrossRef]
  113. Biji, M.; Prabha, B.; Lankalapalli, R.S.; Radhakrishnan, K.V. Transition metal/Lewis acid catalyzed reactions of zerumbone for diverse molecular motifs. Chem. Rec. 2021, 21, 3943–3953. [Google Scholar] [CrossRef] [PubMed]
  114. Nandagopal, K.; Halder, M.; Dash, B.; Nayak, S.; Jha, S. Biotechnological approaches for production of anti-cancerous compounds resveratrol, podophyllotoxin and zerumbone. Curr. Med. Chem. 2018, 25, 4693–4717. [Google Scholar] [CrossRef]
  115. Keerthi, D.; Geethu, C.; Nair, R.A.; Pillai, P. Metabolic profiling of Zingiber zerumbet following Pythium myriotylum infection: Investigations on the defensive role of the principal secondary metabolite, zerumbone. Appl. Biochem. Biotechnol. 2014, 172, 2593–2603. [Google Scholar] [CrossRef]
  116. Doehlemann, G.; Ökmen, B.; Zhu, W.; Sharon, A. Plant pathogenic fungi. Microbiol. Spectr. 2017, 5, FUNK-0023-2016. [Google Scholar] [CrossRef]
  117. Salvatore, M.M.; Andolfi, A. Phytopathogenic fungi and toxicity. Toxins 2021, 13, 689. [Google Scholar] [CrossRef] [PubMed]
  118. Sheikh-Ali, S.I.; Ahmad, A.; Mohd-Setapar, S.H.; Zakaria, Z.A.; Abdul-Talib, N.; Khamis, A.K.; Hoque, M.E. The potential hazards of Aspergillus sp. in foods and feeds, and the role of biological treatment: A review. J. Microbiol. 2014, 52, 807–818. [Google Scholar] [CrossRef] [PubMed]
  119. Ráduly, Z.; Szabó, L.; Madar, A.; Pócsi, I.; Csernoch, L. Toxicological and medical aspects of Aspergillus-derived mycotoxins entering the feed and food chain. Front. Microbiol. 2020, 10, 2908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Navale, V.; Vamkudoth, K.R.; Ajmera, S.; Dhuri, V. Aspergillus derived mycotoxins in food and the environment: Prevalence, detection, and toxicity. Toxicol. Rep. 2021, 8, 1008–1030. [Google Scholar] [CrossRef] [PubMed]
  121. Dousset, X.; Jaffrès, E.; Zagorec, M. Spoilage: Bacterial spoilage. Encycl. Food Health 2015, 48, 106–112. [Google Scholar]
  122. Lorenzo, J.M.; Munekata, P.E.; Dominguez, R.; Pateiro, M.; Saraiva, J.A.; Franco, D. Main groups of microorganisms of relevance for food safety and stability: General aspects and overall description. In Innovative Technologies for Food Preservation: Inactivation of Spoilage and Pathogenic Microorganisms; Barba, F.J., Sant’Ana, A.S., Orlien, V., Koubaa, M., Eds.; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 53–107. ISBN 9780128110324. [Google Scholar]
  123. Pandey, A.K.; Silva, A.S.; Varshney, R.; Chávez-González, M.L.; Singh, P. Curcuma-based botanicals as crop protectors: From knowledge to application in food crops. Curr. Res. Biotechnol. 2021, 3, 235–248. [Google Scholar] [CrossRef]
  124. Arnason, J.T.; Sims, S.R.; Scott, I.M. Natural Products from Plants as Insecticides. UNESCO—Encyclopedia Life Support Systems (UNESCO-EOLSS). Available online: http://www.eolss.net/sample-chapters/c06/e6-151-13.pdf (accessed on 3 November 2022).
  125. Pavela, R. History, presence and perspective of using plant extracts as commercial botanical insecticides and farm products for protection against insects-a review. Czech Acad. Agric. Sci. 2016, 52, 229–241. [Google Scholar]
  126. Isman, M.B. Botanical Insecticides in the twenty-first century-fulfilling their. Annu. Rev. Entomol. 2020, 65, 233–249. [Google Scholar] [CrossRef] [Green Version]
  127. Shaaya, E.; Kostjukovski, M.; Eilberg, J.; Sukprakarn, C. Plant oils as fumigants and contact insecticides for the control of stored-product insects. J. Stored Prod. Res. 1997, 33, 7–15. [Google Scholar] [CrossRef]
  128. Pavela, R. Possibilities of botanical insecticide exploitation in plant protection. Pest Technol. 2007, 1, 47–52. [Google Scholar]
  129. Chauhan, B.S. Grand challenges in weed management. Front. Agron. 2020, 1, 3. [Google Scholar] [CrossRef]
  130. Vilà, M.; Beaury, E.M.; Blumenthal, D.M.; Bradley, B.A.; Early, R.; Laginhas, B.B.; Trillo, A.; Dukes, J.S.; Sorte, C.J.B.; Ibáñez, I. Understanding the combined impacts of weeds and climate change on crops. Environ. Res. Lett. 2021, 16, 034043. [Google Scholar] [CrossRef]
  131. Tataridas, A.; Kanatas, P.; Chatzigeorgiou, A.; Zannopoulos, S.; Travlos, I. Sustainable crop and weed management in the era of the EU green deal: A survival guide. Agronomy 2022, 12, 589. [Google Scholar] [CrossRef]
  132. Fierascu, R.C.; Fierascu, I.C.; Dinu-Pirvu, C.E.; Fierascu, I.; Paunescu, A. The application of essential oils as a next-generation of pesticides: Recent developments and future perspectives. Z. Naturforsch. Sect. C J. Biosci. 2020, 75, 183–204. [Google Scholar] [CrossRef] [PubMed]
  133. Bhovi, V.K. Plants based materials as the antifungal and antibacterial agents. Int. J. Plant Based Pharm. 2022, 2, 98–110. [Google Scholar]
  134. Subhash, S.; Raghavendra, K.V.; Balodi, R.; Deepika; Dubey, N.K. Use of green chemicals in pest and disease management. In Sustainable Management of Potato Pests and Diseases; Kumar Chakrabarti, S., Sharma, S., Shah, M.A., Eds.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 495–524. ISBN 9789811676949. [Google Scholar]
  135. Alwakil, N.H.; Mohamad Annuar, M.S.; Jalil, M. Synergistic effects of plant growth regulators and elicitors on α-humulene and zerumbone production in Zingiber zerumbet Smith adventitious root cultures. Molecules 2022, 27, 4744. [Google Scholar] [CrossRef]
  136. Gupta, T.; Singh, J.; Kaur, S.; Sandhu, S.; Singh, G.; Kaur, I.P. Enhancing bioavailability and stability of curcumin using solid lipid nanoparticles (CLEN): A covenant for its effectiveness. Front. Bioeng. Biotechnol. 2020, 8, 879. [Google Scholar] [CrossRef]
  137. Kurien, B.T.; Scofield, R.H. Increasing aqueous solubility of curcumin for improving bioavailability. Trends Pharmacol. Sci. 2009, 30, 334–335. [Google Scholar] [CrossRef]
  138. Kesharwani, S.S.; Jayarama Bhat, G. Formulation and nanotechnology-based approaches for solubility and bioavailability enhancement of zerumbone. Medicina 2020, 56, 557. [Google Scholar] [CrossRef]
  139. Eid, E.E.M.; Abdul, A.B.; Suliman, F.E.O.; Sukari, M.A.; Rasedee, A.; Fatah, S.S. Characterization of the inclusion complex of zerumbone with hydroxypropyl-β-cyclodextrin. Carbohydr. Polym. 2011, 83, 1707–1714. [Google Scholar] [CrossRef]
  140. Rahman, H.S.; Rasedee, A.; Othman, H.H.; Chartrand, M.S.; Namvar, F.; Yeap, S.K.; Abdul Samad, N.; Andas, R.J.; Muhammad Nadzri, N.; Anasamy, T.; et al. Acute toxicity study of zerumbone-loaded nanostructured lipid carrier on BALB/c mice model. Biomed Res. Int. 2014, 2014, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Salleh, N.; Mahat1, M.M.; Yahaya, S.M.; Rosmamuhamadani, R. Synthesis and characterization of cross-linked zerumbone loaded zeolite Y-gelatin for oral controlled release. AIP Conf. Proc. 2020, 2291, 020063. [Google Scholar]
  142. Salleh, N.; Jais, U.; Sarijo, S. Characterization of crosslinked zerumbone loaded gelatin–zeolite Y hybrid composites with glutaraldehyde for controlled release of natural anticancer drug. Malays. J. Anal. Sci. 2015, 19, 1290–1296. [Google Scholar]
  143. Brouwers, J.; Brewster, M.E.; Augustijns, P. Supersaturating drug delivery systems: The answer to solubility-limited oral bioavailability? J. Pharm. Sci. 2009, 98, 2549–2572. [Google Scholar] [CrossRef] [PubMed]
  144. Sidahmed, H.M.A.; Hashim, N.M.; Abdulla, M.A.; Ali, H.M.; Mohan, S.; Abdelwahab, S.I.; Taha, M.M.E.; Fai, L.M.; Vadivelu, J. Antisecretory, gastroprotective, antioxidant and anti-Helicobacter pylori activity of zerumbone from Zingiber zerumbet (L.) smith. PLoS ONE 2015, 10, e0121060. [Google Scholar] [CrossRef] [PubMed]
  145. Tian, M.; Hong, Y.; Wu, X.; Zhang, M.; Lin, B.; Zho, Y. Chemical constituents and cytotoxic activities of essential oils from the flowers, leaves and stems of Zingiber striolatum diels. Rec. Nat. Prod. 2019, 14, 144–149. [Google Scholar] [CrossRef]
  146. Bhatt, D.A.; Pethe, A. Lipid technology—A promising drug delivery system for poorly water soluble drugs. Int. J. Pharma Res. 2010, 3, 1748–1751. [Google Scholar]
  147. Hosseinpour, M.; Abdul, A.B.; Rahman, H.S.; Rasedee, A.; Yeap, S.K.; Ahmadi, N.; Othman, H.H.; Chartrand, M. Comparison of apoptotic inducing effect of zerumbone and zerumbone-loaded nanostructured lipid carrier on human mammary adenocarcinoma MDA-MB-231 cell line. J. Nanomater. 2014, 182, 742738. [Google Scholar] [CrossRef] [Green Version]
  148. Khan, A.A.; Mudassir, J.; Mohtar, N.; Yusrida, D. Advanced drug delivery to the lymphatic system: Lipid-based nanoformulations. Int. J. Nanomed. 2013, 8, 2733–2744. [Google Scholar]
  149. Abdul, A.B.H.; Al-Zubairi, A.S.; Tailan, N.D.; Wahab, S.I.A. Anticancer activity of natural compound (zerumbone) extracted from Zingiber zerumbet in human HeLa cervical cancer cells. Int. J. Pharmacol. 2008, 4, 160–168. [Google Scholar] [CrossRef]
  150. Rahman, H.S.; Rasedee, A.; Yeap, S.K.; Othman, H.H.; Chartrand, M.S.; Namvar, F.; Abdul, A.B.; How, C.W. Biomedical properties of a natural dietary plant metabolite, zerumbone, in cancer therapy and chemoprevention trials. BioMed Res. Int. 2014, 2014, 20. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biosynthetic route of zerumbone from α-humulene catalyzed by a short-chain dehydrogenase/reductase (SDR).
Figure 1. Biosynthetic route of zerumbone from α-humulene catalyzed by a short-chain dehydrogenase/reductase (SDR).
Plants 12 00005 g001
Figure 2. Applications of zerumbone in the agri-food industry.
Figure 2. Applications of zerumbone in the agri-food industry.
Plants 12 00005 g002
Table 1. Natural sources of zerumbone according to the plant species, origin, method of extraction and analysis. RP-HPLC-PDA: reverse phase high performance liquid chromatography, GC-MS: gas chromatography-mass spectrometry, FID: flame-ionization detector, HPLC: high performance liquid chromatography, FTIR: Fourier transform infrared spectroscopy, UHPLC: ultra-high performance liquid chromatography, HRESI: high-resolution electrospray ionization mass, EI-MS: electron ionized-mass spectroscopy, 1H-NMR: (nuclear magnetic resonance) and 13C-NMR.
Table 1. Natural sources of zerumbone according to the plant species, origin, method of extraction and analysis. RP-HPLC-PDA: reverse phase high performance liquid chromatography, GC-MS: gas chromatography-mass spectrometry, FID: flame-ionization detector, HPLC: high performance liquid chromatography, FTIR: Fourier transform infrared spectroscopy, UHPLC: ultra-high performance liquid chromatography, HRESI: high-resolution electrospray ionization mass, EI-MS: electron ionized-mass spectroscopy, 1H-NMR: (nuclear magnetic resonance) and 13C-NMR.
Plant
Species
Plant
Part
OriginMethod of ExtractionAnalysisYieldMain ComponentsActivityRef.
Zingiber zerumbetRhizomeIndiaMethanol extractionRP-HPLC-PDA Zerumbone (33.02 mg/g)Antioxidant[42]
RhizomesIndia GC/FID-
GC/MS
Zerumbone (70.60%)
α-Humulene (5.65%)
Humulene epoxide I (5.21%)
Humulene epoxide II (5.71%)
Camphor (1.90%)
Camphene (3.47%)
Caryophyllene oxide (2.52%)
Antioxidant[41]
RhizomesIndiaHydrodistillation ClevengerGC- GC/MS0.45%Zerumbone (74.82%)
Humulene (6.02%)
β-Copaen-4α-ol (4.32%)
Antifungal[40]
RhizomesBrazilHydrodistillation ClevengerGC/MS5%Zerumbone (87.93%)Antibacterial[3]
RhizomesMalaysiaHydrodistillation 0.25% Anti-inflammatory[62]
RhizomesIndonesiaWater distillationGC/MS0.12%Sabinene (32.96%)
β-Myrcene (13.27%)
Zerumbone (11.05%)
Increase body weight[63]
RhizomesMalaysiaHydrodistillation
Turbo Extractor Distillator
HPLC0.35%Zerumbone (126.54 mg/mL)Antibacterial[39]
RhizomesReunion IslandSteam distillationGC-GC/MS-
GC/FTIR
0.3–0.4%Zerumbone (36%)
α-Humulene (14.4%)
Camphene (13.8%)
Caryophyllene oxide (5.2%)
Camphor (3.8%)
1,8-Cineole (3.2%)
[38]
RhizomesMalaysiaEthanol extractionHPLC Zerumbone (242.73 mg/g)Immunosuppressant[8]
RhizomesMalaysiaMicrowaveUHPLC Zerumbone (4.82 mg/g DM)Antiproliferative[48]
RhizomesIndiaMethanol extractionRP-HPLC Zerumbone (15.598–30.143 mg/g)Antioxidant [42]
Fresh rhizomesChinaHydrodistillation ClevengerGC/MS–
GC/FID
Zerumbone (40.2%)
α-Caryophyllene (8.6%)
Humulene epoxide II (7.3%)
Camphene (5.9%)
Insecticidal, repellent[10]
Fresh rhizomesChinaHydrodistillation ClevengerGC-FID/MS0.65%Zerumbone (75.0%)
α-Humulene (6.5%)
Humulene oxide I (3.8%)
Camphene (3.3%)
Humulene oxide II (2.7%)
Camphor (1.3%)
Caryophyllene oxide (1.3%)
1,8-Cineole (1.2%)
Antibacterial, cytotoxic [6]
Fresh rhizomesIndiaHydrodistillation ClevengerGC/MS1.12%Zerumbone (49.3%)
α-Caryophyllene (20.1%)
Z-Caryophyllene (3.8%)
Antifungal, antimycotoxin[64]
Fresh rhizomesMalaysiaHydrodistillationGC-FID/MS0.25%Zerumbone (36.12%)
Humulene (10.03%)
Humulene oxide I (4.08%)
Humulene oxide II (2.14%)
Caryophyllene oxide II (1.66%)
Caryophyllene oxide I (1.43%)
Antinociceptive[65]
Fresh rhizomesVietnamSteam distillationHPLC0.1%Zerumbone (98%)Weak in vitro vasodilator[66]
Fresh rhizomesIndia GC/MS0.75%Zerumbone (32.79%)
Camphene (19.41%)
Eucalyptol (6.80%)
[44]
Dried rhizomesChinaHydrodistillation ClevengerGC-FID/MS0.39%Zerumbone (41.9%)
α-Humulene (29.4%)
Humulene oxide I (6.0%)
Humulene oxide II (3.9%)
Camphene (3.9%)
β-Caryophyllene (2.5%)
Camphor (2.4%)
Caryophyllene oxide (2.1%)
1,8-Cineole (1.2%)
Antimicrobial, cytotoxic[6]
Dried rhizomesMalaysiaHydrodistillation ClevengerGC-GC/MS0.37%Zerumbone (58.44%)
α-Humulene (12.24%)
Camphene (5.36%)
Immunosuppressant[8]
Air-dry rhizomesVietnamHydrodistillation ClevengerGC/MS0.65%Zerumbone (51.3%)
Humulene epoxide I (6.4%)
Humulene epoxide II (5.5%)
α-Humulene (5.4%)
Camphene (4.1%)
1,8-Cineole (3.2%)
Insecticidal, antifungal[67]
Powdered rhizomesIndiaAcetone extractionHPLC Zerumbone (99.94%)Anticancer[68]
Dried powdered rhizomesIndiaAcetone extraction
Soxhlet
GC/MS2.86%Zerumbone (64.58%)
Diacetone alcohol (10.64%)
α-Humulene (8.93%)
Caryophyllene oxide (5.68%)
Humulene epoxide (3.18%)
[43]
Powdered rhizomesMalaysiaEthanol extractionHRESI/MS Zerumbone (87.4 mg)Immunosuppressant[69]
RootsIndiaMethanol extractionRP-HPLC-PDA Zerumbone
(05.562 mg/g)
Antioxidant[42]
Z. zerumbet var. darcyiFresh rhizomesIndiaHydrodistillation ClevengerGC-GC/MS0.23%Zerumbone (69.9%)
α-Humulene (12.9%)
Humulene epoxide II (2.5%)
Caryophyllene oxide (1.1%)
Camphene (1.9%)
[9]
Z. ottensiiFresh rhizomeThailandHydrodistillation ClevengerGC/MS0.24%Zerumbone (24.73%)
Terpinen-4-ol (18.75%)
Sabinene (15.19%)
β-Pinene (7.95%)
Cytotoxic[59]
Fresh rhizomesThailandHydrodistillationGC/MS0.21%Zerumbone (25.21%)
Sabinene (23.35%)
Terpene-4-ol (15.97%)
Apoptotic[61]
Z. montanumDried rhizomesBangladeshEthanol extraction1H-NMR/13C-NMR ZerumboneFlavonoid derivatives [58]
Z. amaricans MalaysiaHydrodistillationGC-MS Zerumbone (40.70%) [53]
Powdered dried rhizomesIndonesiaSteam distillationGC-MS1.6%Zerumbone (65.06%)
Humulene oxide (9.66%)
α-Humulene (9.41%)
[54]
Z. aromaticumPowdered dried rhizomesIndonesiaSteam distillationGC-MS2.8%Zerumbone (31.45%)
Cyclohexene (13.52%)
Isogeraniol (10.52%)
[54]
Curcuma rubescensFresh rhizomesChinaSteam distillationGC/MS4.36%Zerumbone (6.88%)
Germacrone (4.99%)
[70]
Powdered dried rhizomesChinaHydrodistillation Clevenger GC/MS Zerumbone (15.45%)
ar-Turmerone (13.80%)
Antimicrobial,
antioxidant,
anticancer,
anti-inflammatory
[71]
Cheilocostus speciosusRhizomesIndiaSteam distillationGC-MS/FID0.18 gZerumbone (55.11%)
α-Humulene (20.55%)
Anticancer,
antibacterial
[72]
Fresh rhizomesIndiaHydrodistillation
Clevenger
GC/MS1.9 mL/kgZerumbone (38.6%)
α-Humulene (14.5%)
Camphene (9.3%)
Insecticidal[73]
Alpinia galangaDried crushed rhizomesSri LankaHydrodistillation ClevengerGC/MS0.56%Zerumbone (44.9%) [74]
Boesenbergia quangngaiensisRhizomesVietnamHydrodistillationGC/MS0.16%cis-β-Elemene (18.4%)
Zerumbone (11.40%)
Myrtenyl acetate (10.6%)
Antimicrobial[75]
Amomum gagnepainiiLeavesVietnamHydrodistillationGC-MS/FID0.20%Farnesyl acetate (18.5%)
Zerumbone (16.4%)
β-Caryophyllene (10.5%)
[76]
Cyperus rotundusAir-dried rhizomesSouth KoreaMethanolextractionEI-MS/1H-NMR/13C-NMR Zerumbone
1,8-Cineole
Insecticidal, repellent[77]
Table 2. Additional information on the studies reviewed in this section.
Table 2. Additional information on the studies reviewed in this section.
TechniquePolymer/MatrixSolubility
Enhancement
Key ResultsReference
Inclusion
complexes
HP-β-CyD>30 fold (From 0.0053 mM for free zerumbone to 0.173 mM with 0.01 M HPβCyD at 20 °C)Molecular modelling calculations showed that zerumbone penetrates completely into the cavity of the HPβCyD. Complexation efficiency zerumbone-HPβCyD was of 1.04.
In vitro cell survival assay on human cervical cancer (Hela), breast cancer (MCF7 and MDA-MB 231) and human leukemic (CEMss) cell lines.
[139]
Inclusion
complexes
HP-β-CyD-Clinical study conducted on three groups of 220 patients each) to treat acute and chronic gastritis. Patients injected with 20 µM of zerumbone.[34]
Inclusion
complexes
HP-β-CyD8 mM of zerumbone dissolved in 50 mM of HP-β-CyD
18 mM zerumbone dissolved in 18 mM of SBE-β-CyD
CyD derivatives are useful to maintain the supersaturation state of zerumbone
SBE-β-CyD[33]
Nanostructured lipid carrierPalm oil:Olive oil:Lipoid S100 (7:3:3)Entrapment efficiency of 99.03% Drug loading of 7.92%.Half maximal inhibitory concentration (IC50) of ZER-NLC was 5.64 ± 0.38 μg/mL (free zerumbone IC50 was 5.39 ± 0.43 μg/mL) after 72 h of treatment.
Zerumbone released after 48 h from NLC was 46.7% (0.95 mg) vs. 90.59% (1.81 mg) from pure drug dispersion.
In vitro cytotoxicity tests in Jurkat T-cell line.
[31]
Nanostructured lipid carriersPalm oil:Olive oil:Lipoid S100 (7:3:3)Entrapment efficiency of 99%Acute toxicity study for zerumbone-NLC conducted orally in BALB/c mice (single dose for 14 days. Oral doses of 100 and 200 mg/kg showed no sign of toxicity or mortality. LD50 is higher than 200 mg/kg therefore safe by oral administration.[140]
Nanostructured lipid carriersPalm oil:Olive oil:Lipoid S100 (7:3:3)Entrapment efficiency of 99%Release of environ 50% after 48 h.
In vivo tests conducted on mice challenged with 4T1 breast cancer.
[29]
Pore
Encapsulation
Zeolite Y-gelatinEntrapment efficiency tests were here performed by soaking 2% w/v zeolite Y in 5, 10 and 15% aqueous solution of 100 µM zerumbone. EE% of 93.7, 97.3 and 99.2% for the zerumbone concentration of 5, 10 and 15% respectively.In vitro controlled release tests confirm the interest of this system for oral administration[141,142]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ibáñez, M.D.; Sánchez-Ballester, N.M.; Blázquez, M.A. Healthy Zerumbone: From Natural Sources to Strategies to Improve Its Bioavailability and Oral Administration. Plants 2023, 12, 5. https://doi.org/10.3390/plants12010005

AMA Style

Ibáñez MD, Sánchez-Ballester NM, Blázquez MA. Healthy Zerumbone: From Natural Sources to Strategies to Improve Its Bioavailability and Oral Administration. Plants. 2023; 12(1):5. https://doi.org/10.3390/plants12010005

Chicago/Turabian Style

Ibáñez, María Dolores, Noelia M. Sánchez-Ballester, and María Amparo Blázquez. 2023. "Healthy Zerumbone: From Natural Sources to Strategies to Improve Its Bioavailability and Oral Administration" Plants 12, no. 1: 5. https://doi.org/10.3390/plants12010005

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