Next Article in Journal
Fully Biobased Reactive Extrusion of Biocomposites Based on PLA Blends and Hazelnut Shell Powders (HSP)
Previous Article in Journal
Aromaticity of Heterocirculenes
Previous Article in Special Issue
A Practical Laboratory-Scale Synthesis of All Eight Stereoisomeric Forms of Terpene Linalool Oxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Occurrence of Marine Ingredients in Fragrance: Update on the State of Knowledge

Laboratory of Natural Products Chemistry and Biomolecules, Faculty of Sciences, University Blida 1, POB 270, Soumaa Road, Blida 09000, Algeria
*
Author to whom correspondence should be addressed.
Chemistry 2021, 3(4), 1437-1463; https://doi.org/10.3390/chemistry3040103
Submission received: 29 September 2021 / Revised: 12 November 2021 / Accepted: 16 November 2021 / Published: 4 December 2021
(This article belongs to the Special Issue Flavors and Fragrances: Biology, Chemistry and Biotechnology)

Abstract

:
The fragrance field of perfumes has attracted considerable scientific, industrial, cultural, and civilizational interest. The marine odor is characterized by the specific smell of sea breeze, seashore, algae, and oyster, among others. Marine odor is a more recent fragrance and is considered as one of the green and modern fragrances. The smells reproducing the marine environment are described due to their content of Calone 1951 (7-methyl-2H-1,5-benzodioxepin-3(4H)-one), which is a synthetic compound. In addition to the synthetic group of benzodioxepanes, such as Calone 51 and its derivatives, three other groups of chemical compounds seem to represent the marine smell. The first group includes the polyunsaturated cyclic ((+)-Dictyopterene A) and acyclic (giffordene) hydrocarbons, acting as pheromones. The second group corresponds to polyunsaturated aldehydes, such as the (Z,Z)-3,6-nonadienal, (E,Z)-2,6-nonadienal, which are most likely derived from the degradation of polyunsaturated fatty acids. The third group is represented by small molecules such as sulfur compounds and halogenated phenols which are regarded as the main flavor compounds of many types of seafood. This review exposes, most notably, the knowledge state on the occurrence of marine ingredients in fragrance. We also provide a detailed discussion on several aspects of essential oils, which are the most natural ingredients from various marine sources used in fragrance and cosmetics, including synthetic and natural marine ingredients.

Graphical Abstract

1. Introduction

According to the layman, the fragrance world conjures up a portrait of charm, imagination, dreams, and fascination, but also an image of expensive essential oils, extracts, and balm. The world of perfumery brings to mind a fascinating industry, which has led to the birth of a wonderful perfume house. On the other side of this appearance, there is a modern industry combined with a strong scientific basis, whose basic foundation is chemistry [1,2,3].
Perfumes are composed of pure and/or mixed molecules, natural or synthetic, which stimulate our olfactory senses in a pleasant way. It is widely acknowledged that odorant molecules are compounds with a molecular weight lower than 300 Da [4,5]. They are characterized by their high volatility, usually described as the saturated vapor pressure. This is the pressure generated by a liquid (or solid) in a vacuum at a fixed temperature. It is measured by diverse methods and is expressed as mmHg, and could be calculated by different equation models. A compilation of studies of the vapor pressure of odoriferous compounds has been reported by innumerable studies including terpenes, which are the largest group of natural smell compounds found in essential oils, among others [6,7,8]. The fragrant compounds are detected, via a complex olfaction mechanism, by terrestrial organisms when transported by air from a distance through an airborne signaling process, which is considered the counterpart of waterborne signaling. Thus, the perception of odor occurs when the airborne volatile compounds land on specific sites on odorant receptors (ORs) localized on the cilia of the olfactory sensory neurons, which transmit the signals to the brain [9,10]. It should be noticed that, in a terrestrial environment, airborne signaling is controlled by the volatility of molecules that are carried by the air, while, in an aquatic environment, the solubility in water is responsible for the displacement of compounds and therefore it manages the process.
Previous studies have shown that marine organisms, which have a strong odor when out of the water, produce a large number of secondary metabolites, including terpenoids, which are highly hydrophobic substances [11,12,13]. They play an important role in chemical communication and often act as pheromones [14,15]. Many of these metabolites are omnipresent in marine and terrestrial organisms. Therefore, olfaction (the sense of smell) is considered as a distance sense interaction, whereas gustation (the sense of taste) is a contact sense interaction [16,17,18]. So, between the marine and terrestrial environment, a compromised phenomenon between taste and smell occurs. We observe the same hydrophobic metabolites smelled at long and short distances in land and sea, respectively, while the opposite situation occurs in taste [19]. This fact is most likely one of the keys to understanding the evolutionary transition from aquatic to terrestrial life, yet several complex ecological interactions between marine organisms remain misunderstood [20,21,22]. Further studies on chemoreception within marine and terrestrial ecosystems would allow the elucidation of an interaction mechanism between them and lead to useful results, particularly from an evolutionary perspective [23,24,25]. Otherwise, smell (olfaction) and taste (gustation), which are a form of chemoreception, are closely linked to the nature of fragrance and/or aroma and, consequently, to their physicochemical properties [26], and could be impacted by cultural differences (age, gender, etc.). Hence, for a molecule to be odorous, it must fulfill a number of criteria. It must have a low molecular weight, a certain degree of hydrophobicity, and high volatility. The latter has a direct repercussion on the odor threshold [27]. The main odor families covering the olfactory spectrum are: “fruity”, “marine”, “green”, “floral”, “spicy”, “woody”, “amber” and “musky”. The marine scent is distinguished by the specific smells of sea breeze, seashore, seaweed, oysters, and other sea products [26]. It is more recent and is considered to be one of the modern green fragrances.

2. General Aspects of Fragrance Classes

As mentioned above, fragrances can be natural or synthetic, pure compounds or mixtures. The ingredients of natural fragrance include diverse volatile fractions, depending on the extraction method. We can first cite the absolute prepared from fragrant grease obtained by enfleurage and hot maceration, which are virtually obsolete nowadays, for several reasons (low yield, manual handling, etc.) [28,29]. The absolute can also be recovered, via specific treatment, from a crude extract, which is obtained by solvent extraction [30]. Among the natural fragrances, essential oils are the most popular. They are extracted by steam distillation, hydrodistillation or microwave assisted hydrodistillation, from several parts of dry or fresh plants including flowers, leaves, fruits, seeds, berries, bark, wood, root, and rhizomes [31]. However, expression is the most commonly used technique to obtain essential oils from citrus peel, such as orange, lemon, bergamot, and so forth [32]. Essential oils are a very complex mixture, notably composed of several classes of terpenes and other aromatic and/or aliphatic compounds [33] produced in the specialized secretory tissues of natural organisms [34]. They are secondary metabolites and are undoubtedly involved in allelopathic interactions and, generally, in the chemical communication of the species [35,36]. Thus, they play an important role in plant defense against insects, herbivores, microorganisms, and also in the pollination process [37].
Some natural fragrances’ ingredients are difficult to extract in a sufficient quantity; therefore, chemists have successfully produced them synthetically. In the first step, they have isolated pure odoriferous molecules from essential oils, crude extracts and headspace followed by their structural characterization, while the second step was reserved for the chemical synthesis. The chemists developed two methods, leading to two categories of synthetic materials [26,38]:
-
Total synthesis, which consists of producing totally artificial “new” molecules from simple chemical reagents;
-
Hemisynthesis uses a molecule derived from natural products and has a chemical structure close to the one to be reproduced. Thus, the basic reagent (called the precursor) is slightly transformed to improve its scent qualities.
-
There are two categories of synthetic materials [38,39]:
-
The total synthetic materials, conducting a diverse chemical function such as aldehyde methyl ionone, ketone, heterocycle and so forth;
-
The isolates separated from natural products. This is the case of many compounds serving as the raw material in (or produced) hemisynthesis. We can cite the example of coumarin isolated by Vogel in 1820 from the tonka bean (Dipteryx odorata) [40], which was the first synthetic fragrance compound released on the market in 1866 [41], followed by salicylaldehyde (1876) and vanillin. The interesting odoriferous qualities of coumarin and its derivatives have led to more and more extensive and up-to-date research [42,43,44].
The synthetic fragrances could contain natural components in combination with the synthetic ones, or are entirely made up of artificially manufactured molecules. They are largely recognized as strong, more original, with a high tenacity, and cheaper than natural fragrances [45]. In addition to natural and synthetic fragrances, natural animal-based odorants have a privileged place in the world of fragrances. Thus, the scents of musks, castoreum and ambergris are well recognized in the world of perfumery. However, it seems that it is mainly the musky substances that are of particular importance in the formulation of fragrances with musky notes [46], as evidenced by the large number of studies on musk substances, focusing on chemical analysis [47] and synthesis [48,49]. It must be pointed out that musk odors regroup natural and synthetic musks. The natural musk is composed particularly by macrocyclic constituents, isolated from animal secretion, such as muscone, 3-methylcyclopentadecanone and civetone as cycloheptadecen-1-one. Artificial musk is made up of three main classes: nitro, polycyclic and macrocyclic musks [46,50].

3. Aroma Properties of Fragrance

The choice of odoriferous compounds mixed in fragrance products is based on their odor quality but also on their olfaction threshold, which depends on the vapor pressure, temperature, aroma intensity and odor activity values (OAVs). The relationship between odor quality and molecular properties is arguably the most important issue in olfaction. Thus, the odor quality remains a relatively ambiguous concept, the quality measurement techniques must comply with a number of criteria [51], and the use of subjective and objective descriptors for the characterization of scent notes has been suggested [52]. In-depth studies demonstrated the high performance of human odor categorization. However, there are no precise methods for measuring the multidimensional axis of odor profiling [53]. The odor threshold is defined as the lowest concentration of a compound that is just sufficient to be distinguished from pure air or a solution, and is also called the recognition threshold. Threshold values are usually determined by smell (orthonasal value) and by tasting the samples (retronasal value) [54,55]. However, retronasal thresholds are generally lower than orthonasal thresholds [56]. Thus, the odors’ evaluation is usually performed by flavor profile analysis (FPA) according to the methods reported in the literature [57,58,59]. In the liquid phase, the threshold is determined in water, ethanol or in a 46% (v/v) hydroalcoholic solution (sometimes in a 12% (v/v) water-ethanol mixture), with a slight modification, depending on the polarity of the compounds, and is then calculated using the methods reported in the literature [60].
It is important to distinguish between two types of odor thresholds, the absolute and the difference thresholds. The detection and the recognition thresholds are absolute thresholds.
As a general rule, the odor thresholds can be influenced by the matrix, such as water, air, oil or solvent, but also by the measurement procedure and the calculation method [56,61,62]. Furthermore, although several studies have been undertaken to predict the relationship between the olfactory threshold and chemical structure, no successful approach has been obtained. Updated literature data and compilations of artificial musk, made up of three main classes of odor threshold values [63,64,65], are available; it has been shown that thresholds of odorants cover a wide range of concentrations. The aroma intensity is evaluated by using sensory panel and gas chromatography olfactometry coupled with the Osme technique (GCO-Osme) [66,67,68]; it is a quantitative method that measures the perceived odor intensity of a compound [69]. The odor activity values (OAVs) for a volatile compound are calculated (Equation (1)) to evaluate its contribution to the overall aroma of the studied sample [70,71].
OAV = C/OT
OAV: Odor activity value of compound
C: The total concentration of compound in the sample
OT: Odor threshold of compound.
The compound is considered an odor active compound for OAV > 1 [72,73]. However, the use of FVOs to assess the contribution of aromatic compounds assumes that there is no interaction between the different aromatic compounds. In fact, the overall aroma is the result of a very complex process involving all aroma compounds in the mixture [74,75].

4. Essential Oils as Natural Sources of Fragrance Compounds

The principal ingredients of natural fragrances are exclusively essential oils and their isolated compounds. Three hundred essential oils extracted from about 300 plant species, belonging to more than 60 botanical families [76] have been described, are well distributed among the whole plant classes [77,78,79], and are commercially valorized in flavor and fragrance industries [80] as well as in phytomedicine and aromatherapy [81,82]. In this context, the availability of an excellent online database, namely AromaDb (http://bioinfo.cimap.res.in/aromadb/, accessed on 23 March 2021, including 1321 aroma structures, bioactivities of essential oil and aroma compounds, 358 fragrance types, and 164 plants must be noted. In addition, it comprises calculated cheminformatics properties associated with identification, physico-chemical properties, pharmacokinetics, toxicological, and ecological information [83]. This database could be of great practical use for the valorization of essential oils and aromatic plants in various fields, in particular the flavor and fragrance industries.
High-valued essential oils are used as fragrances, including citrus, corn mint, citronella, lavender, eucalyptus, tea tree, cinnamon, peppermint, rosemary, and other floral oils, among others. Linalool, geraniol/nerol, citronellal, citronellol, and citral are some of the most important terpenoids, with much-appreciated odorants components, used in the fragrance industry [33,84]. Except for citral, they are all used as such in perfumes. The alcohols and their esters are particularly important; all of them are key starting materials for other terpenoids. In addition, several essential oils are often characterized by the presence of at least one major compound, so it represents the source of these compounds. For example, among the monoterpenes, we can note [85]: limonene (Citrus peel, Coreopsis tinctoria), menthol (Mentha piperita), β- Linalool (Zanthoxylum armatum), Cinnamaldehyde (Cinnamomum cassia), Camphre (Artemisia herba alba), Thymol (Thymus vulgaris), Geranial (Cymbopogon citratus). In the sesquiterpenes class, we can mention some major compounds [86]: Cedrene and thujopsene (Cedarwood), bulnesene (Patchouli), cubebene (Vetiver), zingiberene (Ginger) and germacrene (Ylang-Ylang). Regarding the economics aspects, essential oils are the flagship products of the fragrance and flavor manufacturers; their sales are reliable indicators of the development of the essential oils market. The global market demand for essential oils was estimated at 247.08 kilotons in 2020, and is expected to grow at a compound annual growth rate (CAGR) of 7.5% from 2020 to 2027 (https://www.reportlinker.com/p05799283/?utm_source=GNW, accessed on 25 March 2021, which is expected to reach USD 14.1 Billion by 2026) (https://www.globenewswire.com/news-release/2020/11/03/2119245/0/en/Essential-Oils-Market-Size-Share-Will-Reach-to-US-14-1-Billion-by-2026-Facts-Factors.html, accessed on 25 March 2021).

5. Chemical Composition of Natural and Synthetic Fragrances

The essential oils are mainly lipophilic terpenoids but also phenylpropanoids, or short-chain aliphatic hydrocarbon derivatives of low molecular weight. The terpenes, also known as terpenoids, are the largest group of secondary metabolites derived from natural sources contributing to the flavor; scents considered as natural fragrances compounds. Their structures originate from the coupling of two to many isoprene units (C5-building unit). Thus, depending on the number of isoprene units, they are classified into hemiterpenes (C5), monoterpenes (C10), sesquiterpenes (C15) and diterpenes (C20) [87]. Terpenes are marked by their enormous chemical diversity and are distinguished from one another by their basic skeleton (allylic, mono-, bi, or tricyclic) and functional groups (ketones, alcohols, oxides, aldehydes, phenols, esters, acids, etc.). Biogenetically, they are generated from the mevalonate and mevalonate-independent (deoxycelulose-phosphate) pathways [88]. Monoterpenes are the main compounds in essential oils (over 80%), followed by sesquiterpenes with a lesser amount [89,90,91,92]. Several essential oils contain phenolic compounds (phenylpropanoids), sometimes in appreciable quantities, which derive from the shikimate pathway [93]; for instance, the phenolic monoterpenes, carvacrol [94] and thymol [95], from Origanum dictamnus and thyme essential oils, respectively. Furthermore, the phenolic compounds are endowed with a broad spectrum of biological activities [96]. Furthermore, the phenolic compounds have been described in several studies to have diverse sensory properties, with high odor potencies and show consistently low odor thresholds in foods and other materials [97]. As a result, about 230 volatiles have been identified as so-called “key food odorants” (KFOs) [98] and about 10% of all known KFOs are volatile phenols [99]. It was shown that KFOs are the best natural agonists for roughly 400 human olfactory receptors (ORs) [100]. As a class, volatile sulfur-containing compounds are some of the strongest odorants; they exhibit sensory potency at low concentrations due to their low aroma and taste thresholds. In almost all foods, the volatiles’ sulphur compounds are involved in the distinctive flavor bouquet [101]. The perception of their odors is closely linked to their concentration as well as diastereomeric and enantiomeric forms. Sometimes, one isomer may display a lower flavor threshold compared to its epimer. In some other cases, the aroma may change completely between both enantiomeric forms; for instance, the methyl ester of (R) and (S) 3-Methylthiohexanol exhibit tropical fruit, sulfur and herbaceous, respectively. At low concentrations, the sulfur volatiles’ compounds exhibit a specific smell note and contribute to flavor character. However, at higher concentrations, their aromas are perceived as sulfurous and unpleasant [102,103]. The simple method for assessing the organoleptic contributions of thiols in the overall flavor is to simply add copper to samples, which leads to the formation of an odorless chemical complex [104].
The occurrence of volatile sulfur compounds (VSCs) in essential oils [105,106] and in food flavors was the subject of several recent reviews dedicated to their importance for the aroma and flavor of food, but also for their great chemical diversity [107,108,109,110]. Through this review, it is difficult to draw up an exhaustive list of volatile sulfur compounds; we give examples of few characteristic compounds. Thus, the 8-mercapto-p-menthan-3-one (37), p-menthene-8-thiol (59), 3-mercapto-hexanol (62) and Oxane (100) are known for having an exceptional natural fruity olfaction quality [111,112,113]. In herbs and seasonings, it was reported that the presence of diallyl disulfide and diallyl thio- sulfinate (allicin) in garlic [114], propyl propane thiosulfonate and propyl methanethiol-sulfonate in raw onion [115,116], allyl isothiocyanate in white and brown mustards [117]. In vegetables, the existence of dimethyl sulfide was reported in asparagus [118], 4-methyl thiobutylisothiocyanate (often named erucin) in broccoli [119] and allyl isothiocyanate in raw cabbage [120]. It should be noted that, among the herbs, garlic (Allium sativum L.) is considered a natural source of volatile sulfur compounds; for many years, it has been employed as a medication for common colds, influenza, and other kinds of infections [121]. The essential oil of garlic has been investigated several times, and has led to the identification of hundreds of volatile sulfur compounds [122]. It has a broad spectrum of biological activities [123,124]. Recently, it has been suggested that the garlic essential oil is a valuable natural antivirus source, which contributes to preventing the invasion of coronavirus into the human body [125]. It should be pointed out that volatile sulfur compounds in the majority of foods are biogenetically derived from the sulfur amino acids cysteine and L. methionine [113,126,127]. Furthermore, the volatile sulfur compounds (VSCs), key food odorants (KFOs), were divided into three main groups, considering their content in KFOs [98]: generalists, intermediaries, or individualists. The generalist group is represented by methional (3-(methylthio)-propanal), found in 54% of the KFOs) obtained through enzymatic degradation of L-methionine. The intermediary group is distinguished by the presence of furfurylthiol (meet in 15% of KFOs) formed via carbohydrate and amino acid as precursors [128]. The last individualist group covers less than 5% of the KFOs. A key compound would be 1-p-menthene-8-thiol generated from amino acids or hydrogen sulfide and isoprenoid precursors [129,130]. This aroma compound is found in just 0.4% of KFOs and is distinctive of the aroma of grapefruit juice [131]. Moreover, the odor thresholds of (S) and (R)-1-p-menthene-8-thiol in air are 6.6 × 10−6 and 9 × 10−5 ng/L, respectively, while those of analogous terpenes (R) and (S)-α-terpineol are 4600 and 570 ng/L, respectively [132]. In the same context, the odor threshold value in air of the (4S,8R)-1-p-menthene-9-thiol is 9.4 × 10−2 ng/L, whereas the respective terpene alcohol (4S,8R)-1-pmenthene-9-ol is 23 ng/L [132]. The difference between the odor threshold of the homologous terpenes alcohol and thiol could be interpreted based on the vibration theory of olfaction (VTO), which suggests that olfaction involves the detection of odorant vibrational frequencies [100]. The net difference in vibrational frequencies of OH (3300 cm−1) and SH (2550 cm−1) is well known and, consequently, the difference between odor thresholds. These examples given above illustrate the high sensorial quality of sulfur compounds compared to their respective terpene classes. As a non-exhaustive illustration, we grouped together in Table 1 the sensory properties of the main volatile products of essential oils (chemical structure given in Figure 1) belonging to the different chemical classes: terpenes, phenols and sulfurs.

6. Marine Fragrance Chemistry

The marine notes, originating from sea spray, iodine and other organisms, such as algae, are released into the marine environment. Fresh and iodized, the marine notes are an essential part of modern and trendy fragrances with an ecological character. These fragrances are formulated into two different pathways. The first way consists of using natural volatile compounds derived from the essential oils of natural sources such as algae and coastal plants like sea fennel.
The second path uses synthetic molecules belonging to 2H-1,5-benzodioxepin-3(4H)-ones (1a) and 4,5-dihydro-1-benzoxepin-3(2H)-ones (1b) skeletons patented by Pfizer scientists, J. J. J. Beereboom, D. P. Cameron, C. R. Stephens, in 1966 [189]. The most famous was Calone or 7-methyl-2H-1,5-benzodioxepin-3(4H)-one (2a), trade-named Calone 1951, also known in the industry as “watermelon ketone” with a very low odor threshold of 3.1 × 10−2 ng/L air [190]. It was undeniably the starting point of the history of synthetic molecules reflecting the marine environment, which revolutionized perfumes of marine origin and was responsible for defining an entire genus [191,192,193]. Consequently, for 30 years (1970–2000), Calone 1951® (3) was the only commercialized compound in this series (Figure 2), it was particularly popular in the 1990s. It was Aramis [185], with its “New West for Her (Calone %: 1.2)” perfume, that created the first daring but innovative aquatic perfume. Other scents based on the use of Calone were made including [124]: Escape for Her (C. Klein, 1991) 0.8%, l’Eau d’Issey pour homme (I. Miyake, 1994) 0.6%, L’Eau d’Eden (Cacharel, 1996) 0.17%, Polo Sport for Women (R. Lauren, 1996) 0.45%, Cool Water for Women (Davidoff, 1997) 0.4%, Aquawoman (Rochas, 2002) 0.42% and Alberto Morillas’ Acqua di Gio released by Giorgio Armani (Fragrantica, 2019) [194]. The good olfactory properties of Calone have allowed the emergence of other important molecules, such as Conoline (2a’), Transluzone (2b), Aldolone (2c), and Azurone (2d), which are currently produced in large quantities by Firmenich, Givaudan and Calchauvet [194,195,196,197,198,199,200], and are used in fine fragrance as well as in body- and home-care products. It must be pointed out that, in perfumery formulation, the marine notes blend easily with many olfactory families, including floral, fruity, citrus, woody and oriental.

6.1. Synthetic Ingredients of Marine Fragrance

The good olfactory properties of Calone and its derivatives allowed the emergence of several other odorant molecules, resulting from the significant advances in the structure–activity relationship (SAR) with benzodioxepinone and/or benzoxepinone as the basic skeleton. Recently, olfactophore models, developed in analogy with pharmacophore models, are increasingly used in SAR studies. They provide, through the description of the relative spatial arrangement of osmophoric and hydrophobic groups, information on the geometry of the olfactory receptors and the types of interaction with the molecules [201,202,203]. Consequently, the olfactophore models allow the generation of a new candidate’s odorants molecules without having the receptor’s structure. Thus, the Calone 1951® olfactophore model was produced using the phase program (version 3.4, Schrödinger, LLC, New York, NY, USA, 2012), and led to the identification of four structural features (Figure 3).
1. The presence of hydrophobic alkyl group seems to be necessary for the marine smell of the compound. It appears that positions 7 and 8 [199] are the most favorable, while the absence of a substituent dramatically reduces the intensity of the marine odor and conducts to different scents such cabbage, hot iron, metallic-phenolic, green floral and balsamic [197,204]. The same study [197] had shown that increasing the size of the methyl group enhanced the olfactory quality of Calone. Thus, by lengthening the methyl by two carbons, it could reduce the “low tide” aspect in Aldolone (2C); however, a branched alkyl group attenuated the “low tide” scent and intensified the floral quality of transluzone (Lilial) (2b) (Figure 2). It is estimated that the optimal size of the alkyl group could range between C3 and C6. A comprehensive review of the literature clearly showed the influence of the size and the chemical structure of the hydrophobic group on the olfactory threshold (Table 2). Thus, it was found that increasing the size of the alkyl group could improve the olfactory threshold (2c, 2d); nevertheless, branching the alkyl chain exerts the opposite effect (2f, 2g). The presence of unsaturation on the chain significantly improves the threshold values (2k, 2l, 2m, 2n, 2p), especially the conjugated unsaturation (2t). The presences of an unsaturation and a branching have two contradictory effects (2k, 2r, 2s). The presence of an alkyl group, comprising an aromatic ring (2u) or a cyclic alkyl (2x, 2y, 2z), disadvantages the olfactory threshold value. Position 6, contrary to positions 7 and 8, is not favorable for improving the olfactory threshold value (2v, 2w).
2. The aromatic ring is present in almost all compounds with a marine odor. Thus, it seems to be a requirement for a beneficial interaction with the receptor. Nevertheless, exceptional compounds reported in the literature, holding an open aromatic ring (3) or a saturated ring (4, 5), displayed a marine odor (Figure 4). Compounds 4 and 5 are stereoisomers that have an olfactory threshold in air of 297 ng/L and 177 ng/L, respectively.
3. The spacer is a seven-membered group connecting the aromatic ring and the carbonyl group; this latter is likely involved as a hydrogen-bond acceptor in the interaction with the olfactory receptor. Already, the parent skeleton 1,4-dioxepan-6-one and the analogue 2-methyl-1,4-dioxepan-6-one display a marine odor [205]. According to the hypothesis [197,206], the seven-membered ring geometry is crucial for the marine odor, which is promoted by pseudo-twisted boat conformations.
A revised olfactophore model [203] has been suggested for a compound with an intense marine odor (2t) and a pseudo-half chair conformation. Taking into consideration certain structural requirements [197,203], it seems that an adapted olfactophore model for the marine odor could be: two oxygen functions and an aromatic ring, one but not more than C3–C6 aliphatic substituent on this ring, and the absence of substituents in the α-position to the carbonyl. Several studies have been carried out to improve the odor intensity based on chemical modification of the seven-membered ring. Thus, the easy reduction has been reported, in very high yield, of the carbonyl [207] by using LiAlH4 for an alcohol which is described as “fruity, without character, very weak”. Another idea, which could also provide interesting new perspectives, was to open the seven-membered ring up to allow some flexibility to the double-bonded oxygen (Figure 5) in order to find the key interaction, since, in the benzodioxepinone series, the geometry of the seven-membered ring imposes the spatial orientation of the carbonyl group [199,203,208].
Therefore, the analogs of bendioxepinone (6), calcone (7) and aldolone (8) (Figure 6) exhibit an olfactory profile of: “green cresolic”, “spicy, isoeugenol, guaiacol, vanillic” and “perspiration, aldehydic, slightly aldehyde muguet, vanillic, slightly metallic, nice, too weak”, respectively [209].
Another alternative to preserve the marine-ozone character of the compounds was to reduce the ring size from seven to five atoms. This would allow a greater flexibility of the molecule. To this end, a better option to improve this flexibility was adopted by attaching the double-bonded oxygen to the five-membered ring instead of its integration [209]. This would provide some control over the flexibility of the molecule. They also wanted to give the double-bonded oxygen some flexibility by attaching the five-membered ring instead of integrating it. Therefore, the 5H-1,4-dioxepin-6(7H)-one moiety could be replaced by a 2,3-dihydrofuran-2-carbaldehyde. Very satisfactory results were achieved since the obtained molecules (9, 10, 11) exhibited a greener Calone with a more prominent watermelon odor (Table 2). In 2010, another analog called Aquozone (12) was developed. It has a strong marine, fruity and green smell [210]. More recently, in 2018, two derivatives with isobutyl (13) and isopentyl (14) hyrophobic radical groups were discovered (Table 2) and have showed particularly interesting marine olfaction properties.

6.2. Other Synthetic Marine Fragrance Ingredients

Other molecules that are not structurally related to Calone and its derivatives have shown very interesting marine olfactory properties. We can mention, in this context, the symmarine or para-decylpyridine (15) (Figure 7), which is distinguished by a marine, oyster, algae and mossy smell [211]. It is also characterized by the presence of a rare pyridine-type nitrogenous heterocycle. This is also the case for floralozone (16), which has a floral, marine, ethereal and green odor; it has a chemical structure with an aromatic ring and a very flexible carbonyl group [212]. In 2005 [213], the world of fragrances of marine origin was revolutionized by the arrival of calypsone (17), a linear methoxy aldehyde with a simple structure, characterized by a floral, marine and aldehydic character. For instance, it was utilized in “Hermann à mes côtés me paraissaît une Ombre” (État Libre d’Orange, 2015) and more recently in “Azzaro pour Homme Wild Mint” (Azzaro, 2019).

6.3. Natural Ingredients of Marine Fragrance

6.3.1. Fragrance of Marine Animals

Among the famous fragrances of marine animal origin is the ambergris or Baltic amber. It is a pathological waxy secretion of the intestine of the sperm whale Physeter macrocephalus syn. Physeter catodon (also called cachalot), belonging to the family Physeteridae [214]. Ambergris, most of which comes from the Baltic Sea, has a subtle odor reminiscent of seaweed, wood and moss, but with a particular sweet yet dry undertone of unsurpassed tenacity [215]. The chemical composition of ambergris includes two major parts, 40–60% of fecal steroids [216,217], which is mainly a cholestanol type steroid, and 25–54% of a triterpenoid known as ambrein (18) (Figure 8) [218,219]. This latter is odorless, and the odor characteristics of ambergris result from oxidative decomposition and are obtained under the influence of environmental exposure to sunlight, seawater and air. The oxidative degradation generates odorous compounds, highly valued in perfumery, known as ambroxides (Figure 8) such as: ambrafuran (19), also called ambroxide (AmbroxTM) and derivatives compounds: ambraoxide (20), methylambraoxide (21), ambracetal (22), ambrinol (23) [220]. One of the most interesting ambergris fragrant compounds is the terpene furan ambrafuran (19) or 3a,6,6,9a-tetramethyl-2,4,5,5a,7,8,9,9b-octahydro-1H-benzo[e][1]benzofuran, which is regarded as the prototype of all ambergris fragrances. It has emerged as the main and most pertinent compound in ambergris, due to its sensitive odor, a typical animal note of the material, and its fixative properties [221,222]. Therefore, it is considered a valuable fragrance ingredient. It should be noted that (−)-ambrafuran has a much stronger odor than its enantiomer (+)-ambrafuran, which has a relatively different odor [223,224,225]. Currently, in the perfume industry, ambrafurans are derived entirely from synthetic or semi-synthetic sources. Several reports on the synthesis of (−)-ambrafuran from naturally occurring terpenoids have been described in the literature. Thus, ambrafuran has been synthesized from monoterpenes, sesquiterpenes and diterpenes [226]. The main monoterpenes used for the synthesis of ambrafuran are carvone [227] and thujone [228]. Regarding sesquiterpenes, nerolidol [229,230], β-farnesene [231] and drimenol [232] have been used as starting reagents. An interesting synthetic route using β-ionone and nerolidol via allylic alcohol 2.3 Sigmatropic rearrangement has also been described [229].
Since (−)-ambrafuran is a tetranorlabdane diterpenoid, consequently, labdane diterpenoids are considered the most suitable starting material for its synthesis [233,234]. However, it seems that the privileged route for the synthesis of Ambrox [219] is the one using sclareol. Ambrox of high purity is marketed under the name of Cetalox by Firmenich and Ambrofix by Givaudin. It should be pointed out that several biotechnological routes have been developed for the synthesis of Ambrox using stem bacteria [235], enzymes [236] and fungi [237]. The importance of ambergris for the culture and economy of the world’s people is clear; its most common application throughout history was as a key ingredient in the production of medicines and perfumes. Ambergris is responsible for various pharmacological activities. It has been reported to exhibit cyctotoxic [238], antidiabetic [239], aphrodisiac [240], analgesic and anti-nociceptive [241] activities. It is also an ingredient used to produce perfumes because it has a smell similar to that of musk. It has traditionally been used for many years in the perfume and medical industries to fix the smell of perfumes (for long durability of odor) or as a tincture in deluxe perfumes [242]. Ambroxan (ambroxide and/or ambrafuran) is distinguished by an elegant ambery note with a rich woody character; it is an autoxidation compound of ambrein currently produced through chemical synthesis and used in perfumery for producing ambergris notes, and as a fixative [243]. Over time, it has become an essential molecule for the perfume industry and is marketed by famous perfume houses such as Firmenich and Givaudin. It is used in the formulation of recent renowned perfumes such as: Sun di Gioia Giorgio Armani (2016), Yves Saint Laurent Mon Paris (2016), Ambrox1D3 (Zara, 2019), Givenchy particular agreement (2020), Zara floating around (2021).

6.3.2. Fragrance of Marine Flora

More than 70% of the Earth’s surfaces are oceans and seas. Thus, it is well known that the marine ecosystem is characterized by a bio- and chemodiversity without an equivalent in terrestrial environment. In 2010, 230,000 marine species were recorded [244]. However, according to taxonomic experts, the range is between 3 and 100 million species [245,246]. Since antiquity, throughout the sea, different fragrances with pleasant and unpleasant odors have been noticed; the exploitation of such marine fragrance substances for the production of perfumes is relatively less familiar to the perfume industry. The marine ocean releases a characteristic odor reflecting the organisms living there. The quintessential marine essence cannot be the result of emissions of only a single compound. The typical “sea breeze” smell at the beach and/or taste in sea food is the result of a complex mixture of molecules from several chemical classes. Thus, the volatile components released by marine organisms comprise a mixture of chemical classes such as terpenes, hydrocarbons, fatty acids, esters, alcohols, aldehydes, ketones [247], C11-hydrocarbons [248,249], polyphenols and derivatives [250,251], halogenated [252] and sulfur compounds [253,254]. Dimethyl sulfide, mainly distributed in Chlorophyta and in some Rhodophyta [255], is one of the most common and best understood components of seashore aroma; it is marked by a very unpleasant odor including green, sulfur, clammy, boiled cabbage and creamed corn. It results from the enzymatic cleavage of dimethyl-2-carboxyethylsulfonium hydroxide, from the green algae species (Enteromorpha intestinalis and Acrosiphonia centralis) [256]. The halogenated organic compounds also possess a characteristic odor. They are mainly produced in marine algae and are emitted into the atmosphere; the highest amounts of brominated compounds were released by Laminaria saccharina [257]. Chemical investigations revealed the presence of 2-bromophenol, 2,4-dibromophenol and 2,4,6-tribromophenol in numerous red, green and brown algaes. The briny odor of wild seafood, including fish, mollusks, oysters, clams, and crabs, is caused in large part by a family of compounds called bromophenols. At low concentrations, these chemicals are reported to have a seal, fish or crab-like odor [257]. At very high concentrations, they exhibit an acrid, iodine-like chemical odor.
In addition to the simple halogenated compounds, an extremely important class should be included, namely the halogenated terpenes. Among all marine algae, the class Rhodophyta has a privileged biosynthetic pathway for organohalogen compounds; a large number have been isolated from the genera of Rhodophyta [258,259]. The genus Laurencia is the most prolific source of sesquiterpenes among all marine macroalgae, including halogenated sesquiterpenes of various chemical skeletons, comprising chamigrane, bisabolane, laurane, snyderane, brasilane as well as some rearranged derivatives [260]. It is interesting to note that bromine is the most prevalent halogen in marine natural products, although its concentration in seawater is lower than that of chlorine. However, the distinctive ocean smell is due to the presence of terpenes (the same as those of terrestrial plants), but particularly to a fraction of acyclic and cyclic non-isoprenoid C11-hydrocarbons and derivatives [261], acting as sexual attractant pheromones and playing an important role in chemical communication [262]. They are hydrocarbons with 11 carbon atoms without halogens, which can be classified according to their chemical structure into four groups [263]: (a) derivatives of cyclopropane; (b) derivatives of cyclopentene; (c) derivatives of cycloheptadiene; and (d) acyclic olefins. The main odorant polyenes of the genus Dictyopteris are grouped in Figure 9, showing dictyopterene A (24), dictyopterene B (25), multifidene (26), caudoxirene (27), dictyopterene C’ (Dictyotene) (28), Dictyopterene D’ (ectocarpene) (29), (E,Z,Z)-1,3,5,8-undecatetraene or finavarrene (30), (E,E)-1,3,5-undecatriene (31), (E,Z)-1,3,5-undecratiene (32), (Z,Z,E,E)-2,4,6,8-undecatetraene or giffordene (33), (Z,E)-1,3,5-undecatriene (34), (E,E,E)-2,4,6-Undecatriene (35). In addition, fucoserratene or (Z)-3-Octene (36) was the only identified volatile hydrocarbon with eight carbon atoms in the brown algae. These metabolites, known in all the species of Phaeophyceae, are not specific to an order or a family (Figure 9). However, it seems to be most abundant in brown algae of the genus Dictyopteris [264]. Among the fragrance marine substances, special attention is paid to (+)-Dictyopterene A, which has a characteristic seawater smell [265]; it was the first odoriferous hydrocarbon isolated from Dictyopteris sp. algae [266]. The (−)-dictyotene isomer, obtained by the Cope rearrangement of Dictyopterene A, is found in the brown algae [267] and exhibits a fishy odor reminiscent of salmon roe. As regards the biosynthesis aspects, the C11-hydrocarbons derive from the metabolism of C20-fatty acids [268]. The biosynthetic steps leading from arachidonic acid generate the divinylpropane, which decomposes, via Cope rearrangement, to various C11-hydrocarbons.
In the same context, the presence of C11 sulfur metabolites, containing a C11 unit attached to a sulfur atom with an oxygen substituent at C-3, in the genus Dictyopteris specifically, should be noted (Figure 9). The sulfur compounds, described in the genus Dictyopteris [269,270], appear to be biosynthetically related to C11 hydrocarbons and may arise from the oxidative degradation of highly unsaturated fatty acids via oxygen intermediates [269].
The sulfur-functionalized C11 produced by Dictyopteris polypodiodes is known to attract spermatozoa to female gametes, and is most likely a product of the fatty acid-related pathway [271]. Sulfur compounds are characterized by their biological activities [272] and act as a chemical defense against herbivores [273]. Unfortunately, to the best of our knowledge, no data are available about the olfactory profile of C11 sulfur compounds. The gorgonians are also described as a source of odoriferous compounds [274].
In addition to C11 hydrocarbons, it is appropriate to add another group represented by polyunsaturated aldehydes, such as (E,Z)-2,6-nonadienal (5) [275], which is the main volatiles produced by cucumber fruit providing the characteristic, pleasant aroma associated with fresh cucumbers; they are enzymatically synthesized from the degradation of polyunsaturated fatty acids [276]. In fact, the autoxidation of polyunstatured fatty acids, mainly eicosapentaneoinc acid (EPA), docosahexaenoic acid (DHA) and α-linolenic Acid (ALA), lead to an interesting group of polyunsaturated aldehydes with a short carbon chain, exhibiting a pleasant bouquet of fragrances, including, among others, the marine scent [277].
The key odorant compounds originating from PFUA autoxidation and the olfactory profile are grouped in Table 3. It is appropriate to note that the isomers of some aldehydes are easily obtained by isomerization reactions via isomerases, for example, the conversion of Z-3-hexenal to E-2-hexenal and (Z,Z)-3,6-nonadienal to (E,Z)-2,6-nonadienal [278].
In terms of practical results, the use of marine sources such as seaweeds for fragrance are less valorized and, so far, few species of seaweeds, such as Ceramium virgatum, Codium reediae, Crithmum maritimum, Cystoseira barbata, Cystoseira crinita, Dictyopteris sp., Dictyota sp. and Gracilaria dura, have distinct odors that mark them for perfumery applications [279,280]. The literature reports the use of various other marine sources in the perfume industry [281] such as seagrass, mangroves [282], and coastal plants [283]. The use of some active ingredients from marine sources in cosmetics has also been reported [284,285,286,287]. However, considering the extremely important marine natural resources, their exploitation in the perfume industry of marine origin remains below expectations.

7. Conclusions and Outlook

In this review, we summarize the research progress on fragrance chemistry and the privileged place occupied by fragrances of marine origin in modern perfumes. It is quite legitimate that the main sources of fragrances are the essential oils of land plants for various reasons. The earth’s environment is man’s natural living environment and plants are therefore a cultural and civilizational heritage that has enabled him to develop the science of aromatherapy, which is very useful in daily life. Terpenic compounds, especially monoterpenes and sesquiterpenes, are the main constituents of essential oils and are therefore responsible for their characteristic odors, which are in fact a synergy of individual odors of the different constituents of these essential oils. It is clearly established that certain major compounds are used in perfume formulations to provide the characteristic odor of the essential oils from which they are derived. In addition to terpene compounds, some essential oils are characterized by the presence of phenolic, volatile organosulphur and other amino compounds. These have been used to provide their specific fragrance note in the development of original formulations. In addition to natural products, synthetic chemistry has played a very important role in the development of the perfume industry, providing a wide range of compounds characterized by a very low odor threshold and an extremely attractive aroma quality. The choice of compounds has been made either by imitating nature through the hemisynthesis of major and specific compounds of essential oils, or by the chemical synthesis of compounds with an original chemical structure.
The chemistry of marine fragrances can be divided into two parts—synthetic and natural products. As far as the synthetic products are concerned, Calone 1951 and its derivatives have revolutionized the marine fragrance industry. A large number of perfumes have been formulated from these products. Marine natural products belong to various chemical classes, including a fraction of non-isoprenic compounds, namely C11 hydrocarbons and their derivatives. They are involved in chemical communication between species. They are found in particular in species of the genus Dictyopteris. C11 hydrocarbons and their derivatives represent the future of marine fragrances. However, a major drawback related to the availability of these products in quantity could hinder the development of this sector. In addition to products from marine flora, marine animal products, such as ambergris and its derivatives, have been used in the formulation of amber fragrances. Marine natural products are a very promising avenue for the formulation of marine-based fragrances.

Author Contributions

Sensory Properties, N.R.; general aspects of fragrance N.R. and M.R.Z.; economics aspects of fragrance, N.B. and Y.D.; olphactophore modeling, O.T.; marine fragrance, M.E.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the thematic agency ATRBSA-DGRSDT (Algeria), grant number 007 of 15 March 2020 “socio-economic impact research project”.

Acknowledgments

We would like to express our sincere thanks to the DGRSDT for the efforts made in the development of scientific research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rodrigues, A.; Nogueira, I.; Faria, R. Perfume and Flavor Engineering: A Chemical Engineering Perspective. Molecules 2021, 26, 3095. [Google Scholar] [CrossRef]
  2. Qader, M.M.; Perera, K.D.S.P. Chemistry of Perfumes. Tri-Annu. Publ. Inst. Chem. Ceylon 2020, 37, 26–29. [Google Scholar]
  3. Fortineau, A.-D. Chemistry Perfumes Your Daily Life. J. Chem. Educ. 2004, 81, 45–50. [Google Scholar] [CrossRef]
  4. Mori, K.; Takahashi, Y.K.; Igarashi, K.M.; Yamaguchi, M. Maps of Odorant Molecular Features in the Mammalian Olfactory Bulb. Physiol. Rev. 2006, 86, 409–433. [Google Scholar] [CrossRef] [Green Version]
  5. Touhara, K.; Vosshall, L.B. Sensing Odorants and Pheromones with Chemosensory Receptors. Annu. Rev. Physiol. 2009, 71, 307–332. [Google Scholar] [CrossRef]
  6. Wagner, Z.; Bendová, M.; Rotrekl, J.; Orvalho, S. Densities, Vapor Pressures, and Surface Tensions of Selected Terpenes. J. Solut. Chem. 2019, 48, 1147–1166. [Google Scholar] [CrossRef]
  7. Vilas-Boas, S.M.; Pokorný, V.; Štejfa, V.; Ferreira, O.; Pinho, S.P.; Růžička, K.; Fulem, M. Vapor Pressure and Thermophysical Properties of Eugenol and (+)-Carvone. Fluid Phase Equilibria 2019, 499, 112248. [Google Scholar] [CrossRef] [Green Version]
  8. Nelson, C.R. Investigation of Vaporization Enthalpies and Vapor Pressures of Organic Compounds by Correlation Gas Chromatography. Ph.D. Thesis, University of Missouri, St. Louis, MO, USA, 2018. [Google Scholar]
  9. Soria-Gómez, E. Special Issue Olfaction: From Genes to Behavior. Genes 2020, 11, 654. [Google Scholar] [CrossRef] [PubMed]
  10. Zak, J.D.; Reddy, G.; Vergassola, M.; Murthy, V.N. Antagonistic Odor Interactions in Olfactory Sensory Neurons are Widespread in Freely Breathing Mice. Nat. Commun. 2020, 11, 3350. [Google Scholar] [CrossRef]
  11. Shilling, A.J.; Von Salm, J.L.; Sanchez, A.R.; Kee, Y.; Amsler, C.D.; McClintock, J.B.; Baker, B.J.; Anverenes, B.-E. New Polyhalogenated Monoterpenes from the Antarctic Red Alga Plocamium cartilagineum. Mar. Drugs 2019, 17, 230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Huang, P.-C.; Lin, W.-S.; Peng, B.-R.; Chang, Y.-C.; Fang, L.-S.; Li, G.-Q.; Hwang, T.-L.; Wen, Z.-H.; Sung, P.-J. New Furanocembranoids from Briareum violaceum. Mar. Drugs 2019, 17, 214. [Google Scholar] [CrossRef] [Green Version]
  13. Avila, C. Terpenoids in Marine Heterobranch Molluscs. Mar. Drugs 2020, 18, 162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Marquet, N.; Hubbard, P.C.; da Silva, J.P.; Afonso, J.; Canário, A.V.M. Chemicals Released by Male Sea Cucumber Mediate Aggregation and Spawning Behaviours. Sci. Rep. 2018, 8, 239. [Google Scholar] [CrossRef]
  15. Li, K.; Buchinger, T.J.; Li, W. Discovery and Characterization of Natural Products that Act as Pheromones in Fish. Nat. Prod. Rep. 2018, 35, 501–513. [Google Scholar] [CrossRef] [Green Version]
  16. Smith, C.U.M. Biology of Sensory Systems, 2nd ed.; John Wiley & Sons: Chichester, UK, 2008; pp. 203–243. [Google Scholar]
  17. Hay, M.E. Marine Chemical Ecology: Chemical Signals and Cues Structure Marine Populations, Communities, and Ecosystems. Annu. Rev. Mar. Sci. 2009, 1, 193–212. [Google Scholar] [CrossRef] [Green Version]
  18. Say, T.E.; Degnan, S.M. Molecular and Behavioural Evidence that Interdependent Photo—and Chemosensory Systems Regulate Larval Settlement in a Marine Sponge. Mol. Ecol. 2019, 29, 247–261. [Google Scholar] [CrossRef] [PubMed]
  19. Mollo, E.; Fontana, A.; Roussis, V.; Polese, G.; Amodeo, P.; Ghiselin, M.T. Sensing Marine Biomolecules: Smell, Taste, and the Evolutionary Transition from Aquatic to Terrestrial Life. Front. Chem. 2014, 2, 92. [Google Scholar] [CrossRef] [Green Version]
  20. Heinrichs, M.E.; Mori, C.; Dlugosch, L. Complex Interactions between Aquatic Organisms and Their Chemical Environment Elucidated from Different Perspectives. In Youmares 9—The Oceans: Our Research, Our Future; Jungblut, S., Liebich, V., Bode-Dalby, M., Eds.; Springer: Cham, Switzerland, 2020; pp. 279–297. [Google Scholar]
  21. Staehr, P.; Thomsen, M.; Wernberg, T.; Staehr, P.; Schiel, D. Ecological Interactions between Marine Plants and Alien Species. In Marine Macrophytes as Foundation Species; Ólafsson, E., Ed.; Taylor & Francis Group, LLC: Boca Raton, FL, USA, 2016; pp. 226–258. [Google Scholar] [CrossRef]
  22. Valdez, S.R.; Zhang, Y.S.; Van der Heide, T.; Vanderklift, M.A.; Tarquinio, F.; Orth, R.J.; Silliman, B.R. Positive Ecological Interactions and the Success of Seagrass Restoration. Front. Mar. Sci. 2020, 20, 1–11. [Google Scholar] [CrossRef] [Green Version]
  23. Bressman, N.R.; Simms, M.; Perlman, B.M.; Ashley-Ross, M.A. Where Do Fish Go When Stranded on Land? Terrestrial Orientation of the Mangrove Rivulus Kryptolebias Marmoratus. J. Fish Biol. 2018, 95, 335–344. [Google Scholar] [CrossRef] [Green Version]
  24. Collins, S.; Boyd, P.W.; Doblin, M.A. Evolution, Microbes, and Changing Ocean Conditions. Annu. Rev. Mar. Sci. 2020, 12, 181–208. [Google Scholar] [CrossRef] [PubMed]
  25. Silva, M.C.; Chibucos, M.; Munro, J.B.; Daugherty, S.; Coelho, M.M.; Silva, J.C. Signature of Adaptive Evolution in Olfactory Receptor Genes in Cory’s Shearwater Supports Molecular Basis for Smell in Procellariform Sea-Birds. Sci. Rep. 2020, 10, 543. [Google Scholar] [CrossRef] [PubMed]
  26. Sell, C.S. Fundamentals of Fragrance Chemistry; Wiley-VCH: Weinheim, Germany, 2019; pp. 243–326. [Google Scholar]
  27. Niu, Y.; Liu, Y.; Xiao, Z. Evaluation of Perceptual Interactions between Ester Aroma Components in Langjiu by GC-MS, GC-O, Sensory Analysis, and Vector Model. Foods 2020, 9, 183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Burger, P.; Plainfossé, H.; Brochet, X.; Chemat, F.; Fernandez, X. Extraction of Natural Fragrance Ingredients: History Overview and Future Trends. Chem. Biodivers. 2019, 16, e1900424. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, Q.W.; Lin, L.G.; Ye, W.C. Techniques for extraction and isolation of natural products: A comprehensive review. Chin. Med. 2018, 13, 1–26. [Google Scholar] [CrossRef] [Green Version]
  30. Fernandez, X.; Antoniotti, S.; Bussotti, E.; Hurel, M.P. Parfum, Chimie et Création. Actual. Chim. 2008, 323–324, 42–51. [Google Scholar]
  31. Jugreet, B.S.; Suroowan, S.; Rengasamy, R.R.K.; Mahomoodally, M.F. Chemistry, Bioactivities, Mode of Action and Industrial Applications of Essential Oils. Trends Food Sci. Technol. 2020, 101, 89–105. [Google Scholar] [CrossRef]
  32. Available online: https://www.doterra.com/US/en/blog/science-research-news-cold-press (accessed on 8 July 2019).
  33. Jugreet, B.S.; Mahomoodally, M.F.; Zengin, G.; Maggi, F. Essential Oils as Natural Sources of Fragrance Compounds for Cosmetics and Cosmeceuticals. Molecules 2021, 26, 666. [Google Scholar]
  34. Kowalski, R.; Kowalska, G.; Jankowska, M.; Nawrocka, A.; Kałwa, K.; Pankiewicz, U.; Włodarczyk-Stasiak, M. Secretory Structures and Essential Oil Composition of Selected Industrial Species of Lamiaceae. Acta Sci. Pol. Hortorum Cultus 2019, 18, 53–69. [Google Scholar] [CrossRef]
  35. Mirmostafaee, S.; Azizi, M.; Fujii, Y. Study of Allelopathic Interaction of Essential Oils from Medicinal and Aromatic Plants on Seed Germination and Seedling Growth of Lettuce. Agronomy 2020, 10, 163. [Google Scholar] [CrossRef] [Green Version]
  36. El Sawi, S.A.; Ibrahim, M.E.; Gad El-Rokiek, K.; Amin Saad El-Din, S. Allelopathic Potential of Essential Oils Isolated from Peels of Three Citrus Species. Ann. Agric. Sci. 2019, 64, 89–94. [Google Scholar] [CrossRef]
  37. Bai, L.; Wang, W.; Hua, J.; Guo, Z.; Luo, S. Defensive Functions of Volatile Organic Compounds and Essential Oils from Northern White-Cedar in China. BMC Plant Biol. 2020, 20, 500. [Google Scholar] [CrossRef]
  38. Surburg, H.; Panten, J. Common Fragrance and Flavor Materials Preparation, Properties and Uses; Wiley-VCH Verlag: Weinheim, Germany, 2016; pp. 177–238. [Google Scholar]
  39. Masango, P. Towards Understanding Steam Distillation of Essential Oils by Differential Quantification of Principal Components Using Capillary Gas Chromatography. Ph.D. Thesis, University of Surrey, Guildford, UK, 2001. [Google Scholar]
  40. Vogel, A. Darstellung von Benzoesaure aus der Tonka-Boline und aus den MeliIoten-oder Steinklee-Blumen. Ann. Phys. 1820, 64, 161–166. [Google Scholar] [CrossRef] [Green Version]
  41. Frater, G.; Bajgrowicz, J.A.; Kraft, P. Fragrance Chemistry. Tetrahedron 1998, 54, 7633–7703. [Google Scholar] [CrossRef]
  42. Stefanachi, A.; Leonetti, F.; Pisani, L.; Catto, M.; Carotti, A. Coumarin: A Natural, Privileged and Versatile Scaffold for Bioactive Compounds. Molecules 2018, 23, 250. [Google Scholar] [CrossRef] [Green Version]
  43. Lončarić, M.; Gašo-Sokač, D.; Jokić, S.; Molnar, M. Recent Advances in the Synthesis of Coumarin Derivatives from Different Starting Materials. Biomolecules 2020, 10, 151. [Google Scholar] [CrossRef] [Green Version]
  44. Zeydi, M.M.; Kalantarian, S.J.; Kazeminejad, Z. Overview on Developed Synthesis Procedures of Coumarin Heterocycles. J. Iran. Chem. Soc. 2020, 17, 3031–3094. [Google Scholar] [CrossRef]
  45. Stepanyuk, A.; Kirschning, A. Synthetic Terpenoids in the World of Fragrances: Iso E Super® is the Showcase. Beilstein J. Org. Chem. 2019, 15, 2590–2602. [Google Scholar] [CrossRef] [Green Version]
  46. Sommer, C. The Role of Musk and Musk Compounds in the Fragrance Industry. Handb. Env. Chem. 2004, 3, 1–16. [Google Scholar]
  47. Zhang, T.; Jin, W.; Yang, S.; Li, Y.; Zhang, M.; Shi, M.; Guo, X.; Li, D.; Zhang, B.; Liu, S.; et al. Study of Compositions of Musks in Different Types Secreted by Forest Musk Deer (Moschus berezovskii). PLoS ONE 2021, 16, e0245677. [Google Scholar] [CrossRef]
  48. David, O.R.P. A Chemical History of Polycyclic Musks. Chem. Eur. J. 2020, 26, 7537–7555. [Google Scholar] [CrossRef]
  49. Morin, É.; Sosoe, J.; Raymond, M.; Amorelli, B.; Boden, R.M.; Collins, S.K. Synthesis of a Renewable Macrocyclic Musk: Evaluation of Batch, Microwave, and Continuous Flow Strategies. Org. Process. Res. Dev. 2019, 23, 283–287. [Google Scholar] [CrossRef]
  50. Hong, J.-H.; Lee, J.-Y.; Ha, H.-J.; Lee, J.-H.; Oh, S.-R.; Lee, Y.-M.; Lee, M.-Y.; Zoh, K.-D. Occurrence and Sources of Synthetic Musk Fragrances in the Sewage Treatment Plants and the Han River, Korea. Water 2021, 13, 392. [Google Scholar] [CrossRef]
  51. Wise, P.M.; Olsson, M.; Cain, W.S. Quantification of Odor Quality. Chem. Senses 2000, 25, 429–443. [Google Scholar] [CrossRef] [Green Version]
  52. Thiboud, M. Empirical Classification of Odours. In Perfumes: Art, Science, and Technology; Müller, P.M., Lamparsky, D., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1991; pp. 253–286. [Google Scholar]
  53. Bae, J.; Yi, J.-Y.; Moon, C. Odor Quality Profile is Partially Influenced by Verbal Cues. PLoS ONE 2019, 14, e0226385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zhang, L.; Mi, S.; Liu, R.B.; Sang, Y.X.; Wang, X.H. Evaluation of Volatile Compounds during the Fermentation Process of Yogurts by Streptococcus thermophilus Based on Odor Activity Value and Heat Map Analysis. Int. J. Anal. Chem. 2020, 2020, 3242854. [Google Scholar] [CrossRef]
  55. Zhang, L.-L.; Zhao, L.; Wang, H.-Y.; Zhi, R.-C.; Shi, B.-L.; Xie, N. Determination of Recognition Threshold and Just Noticeable Difference in the Sensory Perception of Pungency of Zanthoxylum bangeanum. Int. J. Food Prop. 2015, 19, 1044–1052. [Google Scholar] [CrossRef] [Green Version]
  56. Piornos, J.A.; Delgado, R.C.J.; de La Burgade, L.; Methven, D.; Balagiannis, P.; Koussissi, E.; Brouwer, E.; Parker, J.K. Orthonasal and Retronasal Detection Thresholds of 26 Aroma Compounds in a Model Alcohol-Free Beer: Effect of Threshold Calculation Method. Food Res. Int. 2019, 123, 317–326. [Google Scholar] [CrossRef]
  57. Caul, J.F. The Profile Method of Flavor Analysis. Adv. Food Res. 1957, 7, 1–40. [Google Scholar] [CrossRef]
  58. Guld, Z.; Sárdy, D.N.; Gere, A.; Rácz, A. Comparison of Sensory Evaluation Techniques for Hungarian Wines. J. Chemom. 2020, 34, e3219. [Google Scholar] [CrossRef] [Green Version]
  59. Keane, P. The Flavor Profile. In Manual on Descriptive Analysis Testing for Sensory Evaluation; Hootman, R., Ed.; ASTM International: Philadelphia, PA, USA, 2008; pp. 5–14. [Google Scholar]
  60. Czerny, M.; Christlbauer, M.; Christlbauer, M.; Fischer, A.; Granvogl, M.; Hammer, M.; Hartl, C.; Hernandez, N.M.; Schieberle, P. Reinvestigation on Odour Thresholds of Key Food Aroma Compounds and Development of an Aroma Language Based on Odour Qualities of Defined Aqueous Odorant Solutions. Eur. Food Res. Technol. 2008, 228, 265–273. [Google Scholar] [CrossRef]
  61. Kim, M.; Drake, S.; Drake, M. Evaluation of Key Flavor Compounds in Reduced-and Full-Fat Cheddar Cheeses Using Sensory Studies on Model Systems. J. Sens. Stud. 2011, 26, 278–290. [Google Scholar] [CrossRef]
  62. Leksrisompong, P.; Barbano, D.M.; Foegeding, A.E.; Gerard, P.; Drake, M. The Roles of Fat and pH on the Detection Thresholds and Partition Coefficients of Three Compounds: Diacetyl, δ-Decalactone and Furaneol. J. Sens. Stud. 2010, 25, 347–370. [Google Scholar] [CrossRef]
  63. Rychlik, M.; Schieberle, P.; Grosch, W. Compilation of Odor Thresholds, Odor Qualities and Retention Indices of Key Food Odorants; Deutsche Forschungsanstalt für Lebensmittelchemie: Garching, Germany, 1998; pp. 1–63. [Google Scholar]
  64. Fazzalari, F.A. Compilation of Odor and Taste Threshold Values Data. In American Society of Testing and Materials; The Society: Philadelphia, PA, USA, 1978; pp. 1–497. [Google Scholar]
  65. Yuan, F.; He, F.; Qian, Y.L.; Zheng, J.; Qian, M.C. Aroma Stability of Lemon-Flavored Hard Iced Tea Assessed by Chirality and Aroma Extract Dilution Analysis. J. Agric. Food Chem. 2016, 64, 5717–5723. [Google Scholar] [CrossRef]
  66. Kamadia, V.; Yoon, Y.; Schilling, M.W.; Marshall, D.L. Relationships between Odorant Concentration and Aroma Intensity. J. Food Sci. 2006, 71, S193–S197. [Google Scholar] [CrossRef]
  67. Chen, Q.-C.; Zhu, Y.; Yan, H.; Chen, M.; Xie, D.-C.; Wang, M.-Q.; Ni, D.-J.; Lin, Z. Identification of Aroma Composition and Key Odorants Contributing to Aroma Characteristics of White Teas. Molecules 2020, 25, 6050. [Google Scholar] [CrossRef]
  68. Gong, X.; Han, Y.; Zhu, J.; Hong, L.; Zhu, D.; Liu, J.; Zhang, X.; Niu, Y.; Xiao, Z. Identification of the Aroma-Active Compounds in Longjing Tea Characterized by Odor Activity Value, Gas Chromatography-Olfactometry, and Aroma Recombination. Int. J. Food Prop. 2017, 20, S1107–S1121. [Google Scholar] [CrossRef]
  69. McDaniel, M.R.; Miranda-Lopwz, R.; Wason, B.T.; Michaels, N.J.; Libbey, L.M. Pinot Noir Aroma: A Sensory/Gas Chromatographic Approach. Dev. Food Sci. 1990, 24, 23–36. [Google Scholar]
  70. Shu, C.; She, Y.; Xiao, Z.; Xu, L.; Niu, Y.; Zhu, J. Investigations on the Aroma Active Compounds in Fresh and Aged Longjing Tea by SPME/GC-MS/GC-O/OAV. Food Ind. 2016, 37, 279–285. [Google Scholar]
  71. Zhu, Y.; Chen, J.; Chen, X.; Chen, D.; Deng, S. Use of Relative Odor Activity Value (ROAV) to Link Aroma Pro-Files to Volatile Compounds: Application to Fresh and Dried Eel (Muraenesox cinereus). Int. J. Food Prop. 2020, 23, 2257–2270. [Google Scholar] [CrossRef]
  72. Tunick, M.H. Analyzing Volatile Compounds in Dairy Products. J. Sci. Food Agric. 2014, 94, 1701–1705. [Google Scholar] [CrossRef] [PubMed]
  73. Niu, Y.; Kong, J.; Xiao, Z.; Chen, F.; Ma, N.; Zhu, J. Characterization of Odor-Active Compounds of Various Chinese “Wuliangye” Liquors by Gas Chromatography–Olfactometry, Gas Chromatography–Mass Spectrometry and Sensory Evaluation. Int. J. Food Prop. 2017, 20, S735–S745. [Google Scholar] [CrossRef] [Green Version]
  74. Wu, C.; Liu, J.; Yan, L.; Chen, H.; Shao, H.; Meng, T. Assessment of Odor Activity Value Coefficient and Odor Contribution Based on Binary Interaction Effects in Waste Disposal Plant. Atmos. Environ. 2015, 103, 231–237. [Google Scholar] [CrossRef]
  75. Aravisini, L.; Guichard, E. Interactions between Aroma Compounds and Food Matrix. In Flavour: From Food to Perception; Guichard, E., Salles, C., Morzel, M., Le Bon, A.M., Eds.; John Wiley & Sons, Ltd., Blackwell: Oxford, UK, 2016; pp. 208–234. [Google Scholar]
  76. Babita, S.; Sellam, P.; Jayoti, M.; Puja, R. Floral Essential Oils: Importance and Uses for Mankind. HortFlora. Res. Spectr. 2014, 3, 7–13. [Google Scholar]
  77. Franz, C.; Novak, J. Sources of essential oils. In Handbook of Essential Oils and Science, Technology and Applications; Hüsnü Can Baser, K., Buchbauer, G., Eds.; CRC Press Taylor & Francis Group: Boca Raton, FL, USA, 2010; pp. 39–81. [Google Scholar]
  78. Mollova, S.; Fidan, H.; Antonova, D.; Bozhilov, D.; Stanev, S.; Kostova, I.; Stoyanova, A. Chemical composition and antimicrobial and antioxidant activity of Helichrysum italicum (Roth) G. Don subspecies essential oils. Turk. J. Agric. For. 2020, 44, 371–378. [Google Scholar] [CrossRef]
  79. Kesdek, M.; Kordali, S.; Bozhüyük, A.U.; Güdek, M. Larvicidal Effect of Achillea biebersteinii Afan. (Asteraceae) Essential Oil Against Larvae of Pine Processionary Moth, Thaumetopoea pityocampa (Denis & Schiffermüller, 1775) (Lepidoptera: Notodontidae). Turk. J. Agric. For. 2020, 44, 451–460. [Google Scholar] [CrossRef]
  80. Hussain, H.; Al-Harrasi, A.; Green, I.R. Frankincense (Boswellia) Oils. In Essential Oils in Food Preservation, Flavor and Safety; Elsevier: Amsterdam, The Netherlands, 2016; pp. 431–440. [Google Scholar]
  81. Baser, K.H.C.; Buchbauer, G. Handbook of Essential Oils: Science, Technology and Applications; CRC Press: Boca Raton, FL, USA, 2015; pp. 619–659. [Google Scholar]
  82. Ali, B.; Al-Wabel, N.A.; Shams, S.; Ahamad, A.; Khan, S.A.; Anwar, F. Essential Oils Used in Aromatherapy: A Systemic Review. Asian Pac. J. Trop. Biomed. 2015, 5, 601–611. [Google Scholar] [CrossRef] [Green Version]
  83. Kumar, Y.; Prakash, O.; Tripathi, H.; Tandon, S.; Gupta, M.M.; Rahman, L.U.; Lal, R.K.; Semwal, M.; Darokar, M.P.; Khan, F. AromaDb: A Database of Medicinal and Aromatic Plant’s Aroma Molecules with Phytochemistry and Therapeutic Potentials. Front. Plant Sci. 2018, 9, 601–611. [Google Scholar] [CrossRef] [PubMed]
  84. Barbieri, C.; Borsotto, P. Essential Oils: Market and Legislation, Potential of Essential Oils; El-Shemy, H., Ed.; IntechOpen: London, UK, 2018; pp. 107–127. [Google Scholar]
  85. Lingan, K. A Review on Major Constituents of Various Essential Oils and Its Application. Transl. Med. 2018, 8, 1–5. [Google Scholar] [CrossRef]
  86. Bayala, B.; Bassolé, I.; Scifo, R.; Gnoula, C.; Morel, L.; Lobaccaro, J.-M.; Simpore, J. Anticancer Activity of Essential Oils and Their Chemical Components—A Review. Am. J. Cancer Res. 2014, 4, 591–607. [Google Scholar]
  87. Boncan, D.A.T.; Tsang, S.S.K.; Li, C.; Lee, I.H.T.; Lam, H.M.; Chan, T.F.; Hui, J.H.L. Terpenes and Terpenoids in Plants: Interactions with Environment and Insects. Int. J. Mol. Sci. 2020, 21, 7382. [Google Scholar] [CrossRef]
  88. Nagegowda, D.A.; Gupta, P. Advances in Biosynthesis, Regulation, and Metabolic Engineering of Plant Specialized Terpenoids. Plant Sci. 2020, 294, 110457. [Google Scholar] [CrossRef]
  89. Dehsheikh, A.B.; Sourestani, M.M.; Dehsheikh, P.B.; Mottaghipisheh, J.; Vitalini, S.; Iriti, M. Monoterpenes: Essential Oil Components with Valuable Features. Mini Rev. Med. Chem. 2020, 20, 958–974. [Google Scholar] [CrossRef]
  90. Fotsing, F.; Stephane, Y.; Kezetas, B.; Jules, J. Terpenoids as Important Bioactive Constituents of Essential Oils. In Essential Oils—Bioactive Compounds, New Perspectives and Applications; de Oliveira, M.S., da Costa, W.A., Silva, S.G., Eds.; IntechOpen: London, UK, 2020; pp. 1–33. [Google Scholar]
  91. Loza-Tavera, H. Monoterpenes in Essential Oils. In Chemicals via Higher Plant Bioengineering, Advances in Experimental Medicine and Biology; Shahidi, F., Kolodziejczyk, P., Whitaker, J.R., Munguia, A.L., Fuller, G., Eds.; Springer: Boston, MA, USA, 1999; pp. 49–62. [Google Scholar]
  92. Hajdari, A.; Mustafa, B.; Hyseni, L.; Bajrami, A.; Mustafa, G.; Quave, C.L.; Nebija, D. Phytochemical Study of Eight Medicinal Plants of the Lamiaceae Family Traditionally Used as Tea in the Sharri Mountains Region of the Balkans. Sci. World J. 2020, 2020, 4182064. [Google Scholar] [CrossRef] [Green Version]
  93. Santos Sánchez, N.F.; Salas-Coronado, R.; Hernandez-Carlos, B.; Villanueva, C. Shikimic Acid Pathway in Biosynthesis of Phenolic Compounds. In Plant Physiological Aspects of Phenolic Compounds; IntechOpen: London, UK, 2019; pp. 1–16. [Google Scholar]
  94. Amaya Olivas, N.; Villalba Bejarano, C.; Ayala Soto, G.; Zermeño Ortega, M.; Sandoval Salas, F.; Sánchez Chávez, E.; Hernández Ochoa, L. Bioactive Compounds and Antioxidant Activity of Essential Oils of Origanum dictamnus from Mexico. AIMS Agric. Food 2020, 5, 387–394. [Google Scholar] [CrossRef]
  95. Kowalczyk, A.; Przychodna, M.; Sopata, S.; Bodalska, A.; Fecka, I. Thymol and Thyme Essential Oil-New Insights into Selected Therapeutic Applications. Molecules 2020, 25, 4125. [Google Scholar] [CrossRef]
  96. Elik, G.; Kılıç, G.; Kanbolat, Ş.; Şener, S.Ö.; Karaköse, M.; Yaylı, N.; Karaoğlu, Ş.A. Biological Activity, and Volatile and Phenolic Compounds from Five Lamiaceae Species. Flavour Fragr. J. 2021, 36, 223–232. [Google Scholar]
  97. Czerny, M.; Brueckner, R.; Kirchhoff, E.; Schmitt, R.; Buettner, A. The Influence of Molecular Structure on Odor Qualities and Odor Detection Thresholds of Volatile Alkylated Phenols. Chem. Senses 2011, 36, 539–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Dunkel, A.; Steinhaus, M.; Kottho, M.; Nowak, B.; Krautwurst, D.; Schieberle, P.; Hofmann, T. Nature’s Chemical Signatures in Human Olfaction: A Foodborne Perspective for Future Biotechnology. Angew. Chem. Int. Ed. 2014, 53, 2–22. [Google Scholar] [CrossRef] [PubMed]
  99. Schieber, A.; Wüst, M. Volatile Phenols—Important Contributors to the Aroma of Plant-Derived Foods. Molecules 2020, 25, 4529. [Google Scholar] [CrossRef] [PubMed]
  100. Block, E. Molecular Basis of Mammalian Odor Discrimination: A Status Report. J. Agric. Food Chem. 2018, 66, 13346–13366. [Google Scholar] [CrossRef]
  101. Maga, J.A.; Katz, I. The Role of Sulfur Compounds in Food Flavor Part I: Thiazoles, C R C Critical Reviews in Food Science and Nutrition. J. Agric. Food Chem. 1975, 6, 153–176. [Google Scholar]
  102. McGorrin, R.J. Sensory-Directed Flavor Analysis; Marsili, R., Ed.; Taylor & Francis: Boca Raton, FL, USA, 2007; pp. 223–267. [Google Scholar]
  103. Zhu, Y.-L.; Zheng, G.-D.; Gao, D.; Chen, T.-B.; Wu, F.-K.; Niu, M.-J.; Zou, K.-H. Odor Composition Analysis and Odor Indicator Selection during Sewage Sludge Composting. J. Air Waste Manag. Assoc. 2016, 66, 930–940. [Google Scholar] [CrossRef] [PubMed]
  104. Vermeulen, C.; Lecocq, S.; Collin, S. Polyfunctional Thiols and Drinkability of Beer. In Proceedings of the 29th European Brewery Convention Congress, Dublin, Ireland, 17–22 May 2003. [Google Scholar]
  105. Iranshahi, M.A. Review of Volatile Sulfur Containing Compounds from Terrestrial Plants: Biosynthesis, Distribution and Analytical Methods. J. Essent. Oil Res. 2012, 24, 393–434. [Google Scholar] [CrossRef]
  106. Kasaian, J.; Asili, J.; Iranshahi, M. Sulphur-Containing Compounds in the Essential Oil of Ferula alliacea Roots and Their Mass Spectral Fragmentation Patterns. Pharm. Biol. 2016, 54, 2264–2268. [Google Scholar] [CrossRef] [PubMed]
  107. Goeke, A. Sulfur-Containing Odorants in Fragrance Chemistry. Sulfur Rep. 2002, 23, 243–278. [Google Scholar] [CrossRef]
  108. Varlet, V.; Fernandez, X. Review. Sulfur-containing Volatile Compounds in Seafood: Occurrence, Odorant Properties and Mechanisms of Formation. Food Sci. Technol. Int. 2010, 16, 463–503. [Google Scholar] [CrossRef] [PubMed]
  109. McGorrin, R.J. The Significance of Volatile Sulfur Compounds in Food Flavors. In Volatile Sulfur Compounds in Food; Qian, M., Fan, X., Mahattanatawee, K., Eds.; ACS Symposium Series; ACS Publications: Washington, DC, USA, 2011; pp. 1–29. [Google Scholar]
  110. Vermeulen, C.; Gijs, L.; Collin, S. Sensorial Contribution and Formation Pathways of Thiols in Foods: A Review. Food Rev. Int. 2005, 21, 69–137. [Google Scholar] [CrossRef]
  111. Lin, J.; Jella, P.; Rouseff, R.L. Gas Chromatography-Olfactometry and Chemiluminescence Characterization of Grapefruit Juice Volatile Sulfur Compounds. In Heteroatomic Aroma Compounds; Reineccius, G.A., Reineccius, T.A., Eds.; American Chemical Society: Washington, DC, USA, 2002; pp. 102–112. [Google Scholar]
  112. Rouseff, R.; Jabalpurwala, F.; Gurbuz, O. Analysis of Grapefruit Sulphur Volatiles using SPME and Pulsed Flame Photometric Detection. Food Chem. 2010, 120, 296–303. [Google Scholar]
  113. Cannon, R.J.; Ho, C.T. Volatile Sulfur Compounds in Tropical Fruits. J. Food Drug Anal. 2018, 26, 445–468. [Google Scholar] [CrossRef]
  114. Abe, K.; Hori, Y.; Myoda, T. Volatile Compounds of Fresh and Processed Garlic (Review). Exp. Ther. Med. 2020, 19, 1585–1593. [Google Scholar] [CrossRef] [Green Version]
  115. Abad, P.; Arroyo-Manzanares, N.; Gil Martínez, L.; García-Campaña, A.M. Use of Onion Extract as a Dairy Cattle Feed Supplement: Monitoring Propyl Propane Thiosulfonate as a Marker of Its Effect on Milk Attributes. J. Agric. Food Chem. 2017, 65, 793–799. [Google Scholar] [CrossRef] [PubMed]
  116. Sorlozano-Puerto, A.; Albertuz-Crespo, M.; Lopez-Machado, I.; Gil-Martinez, L.; Ariza-Romero, J.J.; Maro-to-Tello, A.; Baños-Arjona, A.; Gutierrez-Fernandez, J. Antibacterial and Antifungal Activity of Propyl-Propane-Thiosulfinate and Propyl-Propane-Thiosulfonate, Two Organosulfur Compounds from Allium cepa: In Vitro Antimicrobial Effect via the Gas Phase. Pharmaceuticals 2021, 14, 21. [Google Scholar] [CrossRef]
  117. Eib, S.; Schneider, D.J.; Hensel, O.; Seuß-Baum, I. Relationship between Mustard Pungency and Allyl-Isothiocyanate Content: A Comparison of Sensory and Chemical Evaluations. J. Food Sci. 2020, 85, 2728–2736. [Google Scholar] [CrossRef] [PubMed]
  118. Pegiou, E.; Mumm, R.; Acharya, P.; de Vos, R.C.H.; Hall, R.D. Green and White Asparagus (Asparagus officinalis): A Source of Developmental, Chemical and Urinary Intrigue. Metabolites 2020, 10, 17. [Google Scholar] [CrossRef] [PubMed]
  119. Lv, C.; Zhang, Y.; Zou, L.; Sun, J.; Song, X.; Mao, J.; Wu, Y. Simultaneous Hydrolysis and Extraction Increased Erucin Yield from Broccoli Seeds. ACS Omega 2021, 6, 6385–6392. [Google Scholar] [CrossRef]
  120. Wermter, N.S.; Rohn, S.; Hanschen, F.S. Seasonal Variation of Glucosinolate Hydrolysis Products in Commercial White and Red Cabbages (Brassica oleracea var. capitata). Foods 2020, 9, 1682. [Google Scholar] [CrossRef]
  121. Romeilah, R.M.; Fayed, S.A.; Mahmoud, G.I. Chemical Compositions, Antiviral and Antioxidant Activities of Seven Essential Oils. J. Appl. Sci. Res. 2010, 6, 50–62. [Google Scholar]
  122. Satyal, P.; Craft, J.D.; Dosoky, N.S.; Setzer, W.N. The Chemical Compositions of the Volatile Oils of Garlic (Allium sativum) and Wild Garlic (Allium vineale). Foods 2017, 6, 63. [Google Scholar] [CrossRef] [Green Version]
  123. El-Sayed, H.S.; Chizzola, R.; Ramadan, A.A.; Edris, A.E. Chemical Composition and Antimicrobial Activity of Garlic Essential Oils Evaluated in Organic Solvent, Emulsifying, and Self-Microemulsifying Water Based Delivery Systems. Food Chem. 2017, 221, 196–204. [Google Scholar] [CrossRef]
  124. Shang, A.; Cao, S.Y.; Xu, X.Y.; Gan, R.Y.; Tang, G.Y.; Corke, H.; Mavumengwana, V.; Li, H.B. Bioactive Compounds and Biological Functions of Garlic (Allium sativum L.). Foods 2019, 8, 246. [Google Scholar] [CrossRef] [Green Version]
  125. Thuy, B.T.P.; My, T.T.A.; Hai, N.T.T.; Hieu, L.T.; Hoa, T.T.; Loan, H.T.P.; Triet, N.T.; Van Anh, T.T.; Quy, P.T.; Tat, P.V.; et al. Investigation into SARS-CoV-2 Resistance of Compounds in Garlic Essential Oil. ACS Omega 2020, 5, 8312–8320. [Google Scholar] [CrossRef]
  126. Perea-Sanz, L.; Peris, D.; Belloch, C.; Flores, M. Debaryomyces Hansenii Metabolism of Sulfur Amino Acids as Precursors of Volatile Sulfur Compounds of Interest in Meat Products. J. Agric. Food Chem. 2019, 67, 9335–9343. [Google Scholar] [CrossRef] [PubMed]
  127. Müller, N.; Rauhut, D. Recent Developments on the Origin and Nature of Reductive Sulfurous Off-Odours in Wine. Fermentation 2018, 4, 62. [Google Scholar] [CrossRef] [Green Version]
  128. Liu, H.; Wang, Z.; Hui, T.; Fang, F.; Zhang, D. New Insight into the Formation Mechanism of 2-Furfurylthiol in the Glucose-Cysteine Reaction with Ribose. Food Res. Int. 2021, 143, 110295. [Google Scholar] [CrossRef]
  129. Janes, J.F.; Marr, I.M.; Unwin, N.; Banthorpe, D.V.; Yusuf, A. Reaction of Monoterpenoids with Hydrogen Sulfide to form Thiols and Epi-Sulfides of Potential Organoleptic Significance. Flavour Fragr. J. 1993, 8, 289–294. [Google Scholar] [CrossRef]
  130. Ferreira, V.; Lopez, R. The Actual and Potential Aroma of Winemaking Grapes. Biomolecules 2019, 9, 818. [Google Scholar] [CrossRef] [Green Version]
  131. Demole, E.; Enggist, P.; Ohloff, G. 1-p-Menthene-8-thiol: A Powerful Flavor Impact Constituent of Grapefruit Juice (Citrus Parodisi MACFAYDEN). Helv. Chim. Acta 2004, 65, 1785–1794. [Google Scholar] [CrossRef]
  132. Schoenauer, S.; Schieberle, P. Structure–Odor Activity Studies on Monoterpenoid Mercaptans Synthesized by Changing the Structural Motifs of the Key Food Odorant 1-p-Menthene-8-Thiol. J. Agric. Food Chem. 2016, 64, 3849–3861. [Google Scholar] [CrossRef]
  133. Ravi, R.; Prakash, M.; Bhat, K.K. Aroma Characterization of Coriander (Coriandrum sativum L.) Oil Samples. Eur. Food Res. Technol. 2007, 225, 367–374. [Google Scholar] [CrossRef]
  134. Niu, Y.; Wang, P.; Xiao, Q.; Xiao, Z.; Mao, H.; Zhang, J. Characterization of Odor-Active Volatiles and Odor Contribution Based on Binary Interaction Effects in Mango and Vodka Cocktail. Molecules 2020, 25, 1083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zeller, A.; Rychlik, M. Impact of Estragole and Other Odorants on the Flavour of Anise and Tarragon. Flavour Fragr. J. 2007, 22, 105–113. [Google Scholar] [CrossRef]
  136. Schieberle, P.; Buettner, A. Influence of the Chain Length on the Aroma Properties of Homologous Epoxyal-Dehydes, Ketones, and Alcohols. In Aroma Active Compounds in Foods: Chemistry and Sensory Properties; Takeoka, G.R., Güntert, M., Engel, K.-H., Eds.; ACS Symposium Series; ACS Publications: Washington, DC, USA, 2001; Volume 794, pp. 109–118. [Google Scholar]
  137. Güntert, M.; Krammer, G.; Lambrecht, S.; Sommer, S.H.; Surburg, H.; Werkhoff, P. Flavor Chemistry of Peppermint Oil (Mentha piperita L.). In Araoma Active Compounds: Chemistry and Sensory Properties; Takeoka, G.R., Güntert, M., Engel, K.-H., Eds.; ACS Symposium Series; ACS Publications: Washington, DC, USA, 2001; Volume 794, pp. 119–137. [Google Scholar]
  138. Moreno, J.A.; Zea, L.; Moyano, L.; Medina, M. Aroma Compounds as Markers of the Changes in Sherry Wines Subjected to Biological Ageing. Food Control 2005, 16, 333–338. [Google Scholar] [CrossRef]
  139. Nollet, L.M.L.; Toldra, F. Handbook of Food Analysis, 3rd ed.; CRC Press: Boca Raton, FL, USA, 2015; Volume 1, pp. 47–64. [Google Scholar]
  140. Takeoka, G. Volatile Constituents of Asafoetida. In Araoma Active Compounds: Chemistry and Sensory Properties; Takeoka, G.R., Güntert, M., Engel, K.-H., Eds.; ACS Symposium Series; ACS Publications: Washington, DC, USA, 2001; Volume 794, pp. 33–44. [Google Scholar]
  141. Tamura, H.; Boonbumrung, S.; Yoshizawa, T.; Varanyanond, W. The Volatile Constituents in the Peel and Pulp of Green Thai Mango, Khieo Sawoei Cultivar (Mangifera indica L.). Food Sci. Technol. Res. 2001, 7, 72–77. [Google Scholar] [CrossRef] [Green Version]
  142. Xiao, Q.; Zhou, X.; Xiao, Z.; Niu, Y. Characterization of the Differences in the Aroma of Cherry Wines from Different Price Segments Using Gas Chromatography–Mass Spectrometry, Odor Activity Values, Sensory Analysis, and Aroma Reconstitution. Food Sci. Biotechnol. 2017, 26, 331–338. [Google Scholar] [CrossRef]
  143. Van Gemert, L.J. Odour Thresholds: Compilations of Odour Threshold Values in Air, Water and Other Media; Oliemans Punter & Partners: Utrecht, The Netherlands, 2011; pp. 1–486. [Google Scholar]
  144. Peng, Y.; Bishop, K.S.; Quek, S.Y. Compositional Analysis and Aroma Evaluation of Feijoa Essential Oils from New Zealand Grown Cultivars. Molecules 2019, 24, 2053. [Google Scholar] [CrossRef] [Green Version]
  145. Pino, J.A.; Mesa, J. Contribution of Volatile Compounds to Mango (Mangifera indica L.) Aroma. Flvour Fragr. J. 2006, 21, 207–213. [Google Scholar] [CrossRef]
  146. Plotto, A.; Margaria, C.A.; Goodner, K.L.; Goodrich, R.; Baldwin, E.A. Odour and Flavour Thresholds for Key Aroma Components in an Orange Juice Matrix: Terpenes and Aldehydes. Flavour Fragr. J. 2004, 19, 491–498. [Google Scholar] [CrossRef]
  147. Wang, L.; Hu, G.; Lei, L.; Lin, L.; Wang, D.; Wu, J. Identification and Aroma Impact of Volatile Terpenes in Moutai Liquor. Int. J. Food Prop. 2016, 19, 1335–1352. [Google Scholar] [CrossRef]
  148. Ferreira, V.; Ardanuy, M.; Lopez, R.; Cacho, J.F. Relationship between Flavor Dilution Values and Odor Unit Values in Hydroalcoholic Solutions: Role of Volatility and a Practical Rule for Its Estimation. J. Agric. Food Chem. 1998, 46, 4341–4346. [Google Scholar] [CrossRef]
  149. Zea, L.; Moyano, L.; Moreno, J.; Cortes, B.; Medina, M. Discrimination of the Aroma Fraction of Sherry Wines Obtained by Oxidative and Biological Ageing. Food Chem. 2001, 75, 79–84. [Google Scholar] [CrossRef]
  150. Moyano, L.; Zea, L.; Moreno, J.; Medina, M. Analytical Study of Aromatic Series in Sherry Wines Subjected to Biological Aging. J. Agric. Food Chem. 2002, 50, 7356–7361. [Google Scholar] [CrossRef]
  151. Seideneck, R.; Schieberle, P. Comparison of the Key Aroma Compounds in Hand-Squeezed and Unpasteurised, Commercial NFC Juices Prepared from Brazilian Pera Rio Oranges. Eur. Food Res. Technol. 2011, 232, 995–1005. [Google Scholar] [CrossRef]
  152. Cometto–Muñiz, J.E.; Cain, W.S.; Abraham, M.H.; Gola, J.M. Chemosensory Detectability of 1-Butanol and 2-Heptanone Singly and in Binary Mixtures. Physiol. Behav. 1999, 67, 269–276. [Google Scholar] [CrossRef] [Green Version]
  153. Ong, P.K.; Acree, T.E.; Lavin, E.H. Characterization of Volatiles in Rambutan Fruit (Nephelium lappaceum L.). J. Agric. Food Chem. 1998, 46, 611–615. [Google Scholar] [CrossRef]
  154. D’Acampora Zellner, B.; Lo Presti, M.; Barata, L.E.S.; Dugo, P.; Dugo, G.; Mondello, L. Evaluation of Leaf-Derived Extracts as an Environmentally Sustainable Source of Essential Oils by Using Gas Chromatography—Mass Spectrometry and Enantioselective Gas Chromatography—Olfactometry. Anal. Chem. 2006, 78, 883–890. [Google Scholar] [CrossRef] [PubMed]
  155. Boonbumrung, S.; Tamura, H.; Mookdasanit, J.; Nakamoto, H.; Ishihara, M.; Yoshizawa, T.; Varanyanond, W. Characteristic Aroma Components of the Volatile Oil of Yellow Keaw Mango Fruits Determined by Limited Odor Unit Method. Food Sci. Technol. Res. 2001, 7, 200–206. [Google Scholar] [CrossRef] [Green Version]
  156. Schieberle, P.; Grosch, W. Identification of Potent Flavor Compounds formed in an Aqueous Lemon Oil/Citric Acid Emulsion. J. Agric. Food Chem. 1998, 36, 797–800. [Google Scholar] [CrossRef]
  157. Dharmawan, J.; Kasapis, S.; Sriramula, P.; Lear, M.J.; Curran, P. Evaluation of Aroma-Active Compounds in Pontianak Orange Peel Oil (Citrus nobilis Lour. Var. microcarpa Hassk.) by Gas Chromatography-Olfactometry, Aroma Reconstitution, and Omission Test. J. Agric. Food Chem. 2009, 57, 239–244. [Google Scholar] [CrossRef]
  158. Sarkic, A.; Stappen, I. Essential Oils and Their Single Compounds in Cosmetics—A Critical Review. Cosmetics 2018, 5, 11. [Google Scholar] [CrossRef] [Green Version]
  159. Voirol, E.; Daget, N. Comparative Study of Nasal and Retronasal Olfactory Perception. Lebensm. Wiss. Technol. 1986, 19, 316–319. [Google Scholar]
  160. Randebrock, R.E. Geruch und konstitution. Eine Beweisführung für die Molekulartheorie des Geruches. Teil IV: Olfaktometerversuche zur Molekulartheorie des Geruches. Parfuem Kosmet 1986, 67, 10–24. [Google Scholar]
  161. Buhr, K.; Köhlnhofer, B.; Heilig, A.; Hinrichs, J.; Schieberle, P. Behaviour of Selected Flavour Compounds in Dairy Matrices: Stability and Release. In Expression of Multidisciplinary Flavour Scienc; Blank, I., Wüst, H., Yeretzian, C., Eds.; ZHAW Zürcher College for Applied Sciences: Winterthur, Switzerland, 2010; pp. 165–168. [Google Scholar]
  162. Langridge, P.; Chalmers, K. The Principle: Identification and Application of Molecular Markers. In Molecular Marker Systems in Plant Breeding and Crop Improvement; Springer: Heidelberg, Germany, 2004; pp. 3–22. [Google Scholar]
  163. El-Zaeddi, H.; Martínez-Tomé, J.; Calín-Sánchez, Á.; Burló, F.; Carbonell-Barrachina, Á.A. Volatile Composition of Essential Oils from Different Aromatic Herbs Grown in Mediterranean Regions of Spain. Foods 2016, 5, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Giri, A.; Osako, K.; Ohshima, T. Identification and Characterization of Headspace Volatiles of Fish Miso, a Japanese Fish Meat Based Fermenented Paste, with Special Emphasis on Effect of Fish Species and Meat Washing. Food Chem. 2010, 120, 621–631. [Google Scholar] [CrossRef]
  165. Averbeck, M.; Schieberle, P. Influence of Different Storage Conditions on Changes in the Key Aroma Compounds of Orange Juice Reconstituted from Concentrate. Eur. Food Res. Technol. 2011, 232, 129–142. [Google Scholar] [CrossRef]
  166. Molhave, L.; Kjaergaard, S.K.; Hempeljorgensen, A.; Juto, J.E.; Andersson, K.; Stridh, G.; Falk, J. The Eye Irritation and Odor Potencies of Four Terpenes Which Are Major Constituents of the Emissions of VOCs from Nordic Soft Woods. Indoor Air 2000, 10, 315–318. [Google Scholar] [CrossRef]
  167. Ribéreau-Gayon, P.; Boidron, J.N.; Terrier, A. Aroma of Muscat Grape Varieties. J. Agric. Food Chem. 1975, 23, 1042–1047. [Google Scholar] [CrossRef]
  168. Elss, S.; Kleinhenz, S.; Schreier, P. Odor and Taste Thresholds of Potential Carry-Over/Off-Flavor Compounds in Orange and Apple Juice. LWT-Food Sci. Technol. 2007, 40, 1826–1831. [Google Scholar] [CrossRef]
  169. Padrayuttawat, A.; Yoshizawa, T.; Tamura, H.; Tokunaga, T. Optical Isomers and Odor Thresholds of Vola-Tile Constituents in Citrus Sudachi. Food Sci. Technol. Int. 1997, 3, 402–408. [Google Scholar]
  170. Christoph, N. Die Anwendung der Gaschromatographischen Sniffing—Technik zur Bestimmung von Geruchsschwellen und Aromawerten. Ph.D. Thesis, Universität München, München, Germany, 1983. [Google Scholar]
  171. Nagata, Y.; Takeuchi, N. Measurement of Odor Threshold by Triangle Odor Bag Method. Odor Meas. Rev. 2003, 118, 118–127. [Google Scholar]
  172. Kirsch, F.; Buettner, A. Odor Qualities and Thresholds of Physiological Metabolites of 1,8-Cineole as an Example for Structure—Activity Relationships Considering Chirality Aspects. Chem. Biodivers. 2013, 10, 1683–1695. [Google Scholar] [CrossRef]
  173. Farina, L.; Boido, E.; Carrau, F.; Versini, G.; Dellacassa, E. Terpene Compounds as Possible Precursors of 1,8-Cineole in Red Grapes and Wines. J. Agric. Food Chem. 2005, 53, 1633–1636. [Google Scholar] [CrossRef]
  174. Tamura, H.; Yang, R.-H.; Sugisawa, H. Aroma Profiles of Peel Oils of Acid Citrus. ACS Symp. Ser. 1993, 525, 121–136. [Google Scholar] [CrossRef]
  175. Yang, R.; Sugisawa, H.; Nakatani, H.; Tamura, H.; Takagi, N. Comparison of Odor Quality in Peel Oils of Acid Citrus. Nippon. Shokuhin Kogyo Gakkaishi 1992, 39, 16–24. [Google Scholar] [CrossRef] [Green Version]
  176. Matsushita, H.; Arito, H.; Suzuki, Y.; Soda, R. Determination of Threshold Values for Olfactory Perception of primary Substances. Ind. Health 1967, 5, 221–237. [Google Scholar] [CrossRef]
  177. Usami, A.; Ono, T.; Marumoto, S.; Miyazawa, M. Comparison of Volatile Compounds with Characteristic Odor in Flowers and Leaves of Nojigiku (Chrysanthemum japonense). J. Oleo Sci. 2013, 62, 631–636. [Google Scholar] [CrossRef] [Green Version]
  178. Tempelaar, H.C.G. Over den invloed van licht op reukstoffen. Ph.D. Thesis, Unversity of Utrecht, Utrecht, Germany, 1913. [Google Scholar]
  179. Buchbauer, G.; Jirovetz, L.; Jager, W.; Planck, C.; Dietrich, H. Fragrance Compounds and Essential Oils with Sedative Effects upon Inhalation. J. Pharm. Sci. 1993, 82, 660–664. [Google Scholar] [CrossRef]
  180. Sugisawa, H.; Takeda, M.; Yang, R.H.; Takagi, N. The Comparison of Odor Quality of Volatiles in Peel Oils of Five Kinds of Navel Oranges. Nippon Shokuhin Kogyo Gakkaishi 1991, 38, 668–674. [Google Scholar] [CrossRef]
  181. Chaves, M.; Zea, L.; Moyano, L.; Medina, M. Changes in Color and Odorant Compounds during Oxidative Aging of Pedro Ximinez Sweet Wines. J. Agric. Food Chem. 2007, 55, 3592–3598. [Google Scholar] [CrossRef]
  182. Tominaga, T.; Niclass, Y.; Frerot, E.; Dubourdieu, D. Stereoisomeric Distribution of 3-Mercaptohexan-1-ol and 3-Mercaptohexyl Acetate in Dry and Sweet White Wines Made from Vitis Vinifera (Var. Sauvignon Blanc and Semillon). J. Agric. Food Chem. 2006, 54, 7251–7255. [Google Scholar] [CrossRef] [PubMed]
  183. Steinhaus, M.; Sinuco, D.; Polster, J.; Osorio, C.; Schieberle, P. Characterization of the Aroma-Active Compounds in Pink Guava (Psidium guajava, L.) by Application of the Aroma Extract Dilution Analysis. J. Agric. Food Chem. 2008, 56, 4120–4127. [Google Scholar] [CrossRef] [PubMed]
  184. Lehmann, D.; Dietrich, A.; Hener, U.; Mosandl, A. Stereoisomeric Flavour Compounds LXX, 1-p-Menthene-8-thiol: Separation and Sensory Evaluation of the Enantiomers by Enantioselective Gas Chromatography/Olfactometry. Phytochem. Anal. 1995, 6, 255–257. [Google Scholar] [CrossRef]
  185. Huber, U.A. Erfassung des Geruchs von Riechstoffen. Seifen Öle Fette Wachse 1984, 110, 448–451. [Google Scholar]
  186. Sell, C.S. Chemistry and the Sense of Smell; John Wiley & Sons: Hoboken, NJ, USA, 2014; pp. 188–209. [Google Scholar]
  187. Burdock, G.A. Fenaroli’s Handbook of Flavor Ingredients, 6th ed.; CRC Press: London, UK, 2009; pp. 227–370. [Google Scholar]
  188. Bitar, A.; Ghaddar, T.; Malek, A.; Haddad, T.; Toufeili, I. Sensory Thresholds of Selected Phenolic Constituents from Thyme and Their Antioxidant Potential in Sunflower Oil. J. Am. Oil Chem. Soc. 2008, 85, 641–646. [Google Scholar] [CrossRef]
  189. Baldus, C. Odour Threshold Study. Ph.D. Thesis, University of Würzburg, Würzburg, Germany, 1936. [Google Scholar]
  190. Beereboom, J.J.; Cameron, D.P.; Stephens, C.R. Pfizer. U.S. Patent 3799892, Prior, 15 August 1966. [Google Scholar]
  191. Kraft, P.; Eichenberger, W. Conception, Characterization and Correlation of New Marine Odorants. Eur. J. Org. Chem. 2003, 19, 3681–3874. [Google Scholar] [CrossRef]
  192. Hügel, H.M.; Drevermann, B.; Lingham, A.R.; Marriott, P.J. Marine Fragrance Chemistry. In Current Topics in Flavor and Fragrance Research; Kraft, P., Swift, K.A.D., Eds.; Verlag Helvetiva Chimica Acta: Zürich, Switzerland; Wiley-VCH: Weinheim, Germany, 2008; pp. 199–209. [Google Scholar]
  193. Helbig, C. Discover Why the Perfume Ingredient Calone Is So Popular. Available online: https://www.liveabout.com/what-is-calone-346209 (accessed on 8 May 2019).
  194. Behnke, M. Olfactory Chemistry: Calone- The Smell of the Sea (and Watermelon). Available online: http://colognoisseur.com/olfactory-chemistry-calone-the-smell-of-the-sea-and-watermelon/ (accessed on 14 August 2015).
  195. Fragrantica. Calone. Available online: https://www.fragrantica.com/notes/Calone-423.html (accessed on 8 May 2019).
  196. Gaudin, J.M.; Blanc, P.A. 7-Propyl-benzodioxepin-3-one and Its Use in Perfumery. Firmenich SA EP 902024 A1, 17 March 1999. [Google Scholar]
  197. Kraft, P. 1,2-Substituted 2,3-dihydro-1H-5,9-dioxacyclohepta[f]inden-7-ones and 7-substituted benzo[b][1,4]dioxepin-3-ones. Givaudan SA EP 1136481 B1, 13 November 2002. [Google Scholar]
  198. Gaudin, J.M.; Nikolaenko, O.; de Saint Laumer, J.Y.; Winter, B.; Blanc, P.A. Structure–Activity Relationship in the Domain of Odorants Having Marine Notes. Helv. Chim. Acta 2007, 90, 1245–1265. [Google Scholar] [CrossRef]
  199. Kraft, P.; Schär, M. Dérivé de 2H-benzo[b][1,4]dioxepin-3 (4h) -one derivative and its use as fragrance. Givaudan SA WO 2010/121981 Al, 19 April 2010. [Google Scholar]
  200. Kraft, P.; Popaj, K.; Müller, P.; Schär, M. ‘Vanilla Oceanics’: Synthesis and Olfactory Properties of (1′E)-7-(Prop-1′-enyl)-2H-benzo[b][1,4]dioxepin-3(4H)-ones and Homologues. Synthesis 2010, 17, 3029–3036. [Google Scholar] [CrossRef]
  201. Kozlov, N.G.; Basalaeva, L.I.; Vyglazov, O.G.; Chuiko, A. Synthesis of Calone Derivatives. Chem. Nat. Compd. 2011, 47, 391–394. [Google Scholar] [CrossRef]
  202. Kowalewski, J.; Ray, A. Predicting Human Olfactory Perception from Activities of Odorant Receptors. iScience 2020, 23, 101361. [Google Scholar] [CrossRef]
  203. Hauser, N.; Kraft, P.; Carreira, E.M. The Serendipitous Discovery of a Rose Odorant. Chimia 2020, 74, 247–251. [Google Scholar] [CrossRef] [PubMed]
  204. Armanino, N.; Charpentier, J.; Flachsmann, F.; Goeke, A.; Liniger, M.; Kraft, P. What’s Hot, What’s Not: The Trends of the Past 20 Years in the Chemistry of Odorants. Angew. Chem. Int. Ed. Engl. 2020, 59, 16310–16344. [Google Scholar] [CrossRef]
  205. Drevermann, B.; Lingham, A.; Hugel, H.; Marriott, P. Synthesis of Benzodioxepinone Analogues via a Novel Synthetic Route with Qualitative Olfactory Evaluation. Helv. Chim. Acta 2007, 90, 1006–1027. [Google Scholar] [CrossRef]
  206. Plummer, C.M.; Kraft, P.; Froese, J.; Hudlicky, T.; Rook, T.J.; Jones, O.A.H.; Hugel, H.M. Synthesis and Olfactory Properties of 2-Substituted and 2,3-Annulated 1,4-Dioxepan-6-ones. Asian J. Org. Chem. 2015, 4, 1075–1084. [Google Scholar] [CrossRef]
  207. Rowe, D.J. Chemistry and Technology of Flavours and Fragrances; Wiley-Blackwell: Poole, UK, 2005; pp. 85–197. [Google Scholar]
  208. Drevermann, B.; Lingham, A.R.; Hügel, H.M.; Marriott, P.J. Synthesis and Qualitative Olfactory Evaluation of Benzodioxepine Analogues. Helv. Chim. Acta 2007, 90, 854–862. [Google Scholar] [CrossRef]
  209. Li, H.; Spannenberg, A.; Neumann, H.; Beller, M.; Wu, X.F. Regioselective Synthesis of 2,3-dihydrobenzodioxepinones from Epoxides and 2-bromophenols via Palladium-Catalyzed Carbonylation. Chem. Commun. 2014, 50, 2114–2116. [Google Scholar] [CrossRef] [PubMed]
  210. Gaudin, J.M.; de Saint Laumer, J.Y. Structure–Activity Relationships in the Domain of Odorants Having Marine Notes. Eur. J. Org. Chem. 2015, 3, 1–12. [Google Scholar] [CrossRef]
  211. Huboux, A.; Gaudin, J.M.; Millet, P.; Robvieux, F. Benzodioxole Derivatives as Watery Odorants. Int. Pat. Appl. Firmenich WO 2012045646A1, 12 April 2012. [Google Scholar]
  212. Oertling, H. 4-Alkyl substituted pyridines as olfactory substances. Eur. Pat. Appl. Symrise EP 2100589A1, 16 September 2009. [Google Scholar]
  213. Liniger, M. Aldehydes for Use as Odorants. Int. Pat. Appl. Givaudan WO 2019092056A1, 16 May 2019. [Google Scholar]
  214. Müller, U. 6-Methoxy-2,6-dimethyloctanal and Its Use as a Fragrance Ingredient. Eur. Pat. Appl. (to Givaudan), EP 1764355A1, 21 March 2007. [Google Scholar]
  215. Dannenfeldt, K.H. Ambergris: The Search for its Origin. J. Hist. Sci. Soc. 1982, 73, 382–387. [Google Scholar] [CrossRef]
  216. Ohloff, G.; Vial, C.; Wolf, H.R.; Job, K.; Jégou, E.; Polonsky, J.; Lederer, E. Stereochemistry-Odor Relationships in Enantiomeric Ambergris Fragrances. Helv. Chim. Acta 1980, 63, 1932–1946. [Google Scholar] [CrossRef]
  217. Rowland, S.J.; Sutton, P.A. Chromatographic and Spectral Studies of Jetsam and Archived Ambergris. Nat. Prod. Res. 2017, 31, 1752–1757. [Google Scholar] [CrossRef]
  218. Rowland, S.J.; Sutton, P.A.; Belt, S.T.; Fitzsimmons-Thoss, V.; Scarlett, A.G. Further Spectral and Chromatographic Studies of Ambergris. Nat. Prod. Res. 2018, 32, 2603–2609. [Google Scholar] [CrossRef] [Green Version]
  219. Rowland, S.J.; Sutton, P.A.; Wolff, G.A. Biosynthesis of Ambrein in Ambergris: Evidence from Isotopic Data and Identification of Possible Intermediates. Nat. Prod. Res. 2019, 35, 1235–1241. [Google Scholar] [CrossRef] [PubMed]
  220. Rowland, S.J.; Sutton, P.A.; Knowles, T.M. The Age of Ambergris. Nat. Prod. Res. 2019, 33, 3134–3142. [Google Scholar] [CrossRef]
  221. Coste-Manière, I.C.; Zahra, J.P.; Waegell, B. Synthesis of Ambergris Fragrance Chemicals from Sclareol, Involving Palladium Catalysed Key Steps. Tetrahedron Lett. 1988, 29, 1017–1020. [Google Scholar] [CrossRef]
  222. Neu, T.R.; Lawrence, J.R. Investigation of Microbial Biofilm Structure by Laser Scanning Microscopy. In Productive Biofilms. Advances in Biochemical Engineering/Biotechnology; Muffler, K., Ulber, R., Eds.; Springer International Publishing: Cham, Switzerland, 2014; Volume 146, pp. 1–52. [Google Scholar]
  223. Zinkel, D.; Toda, J.; Rowe, J. Occurrence of Anticopalic acid in Pinus Monticola. Phytochemistry 1971, 10, 1161–1163. [Google Scholar] [CrossRef]
  224. Cheng, L.P.; Xu, L.; Mao, H.F.; Wang, G.L. Study of Structural and Electronic Origin of Ambergris Odor of Some Compounds. J. Mol. Model 2009, 15, 1–8. [Google Scholar] [CrossRef]
  225. Mookherjee, B.D.; Wilson, R.A. Tobacco Constituents—Their Importance in Flavor and Fragrance Chemistry. Perfum. Flavor 1990, 15, 27–49. [Google Scholar]
  226. Winter, B. Spirocyclic Ethers Related to Ambrox®: Synthesis and Structure-Odor Relationships. Helv. Chim. Acta 2004, 87, 1616–1627. [Google Scholar] [CrossRef]
  227. Ncube, E.N.; Steenkamp, L.; Dubery, I.A. Ambrafuran (AmbroxTM) Synthesis from Natural Plant Product Precursors. Molecules 2020, 25, 3851. [Google Scholar] [CrossRef]
  228. Kozioł, A.; Stryjewska, A.; Librowski, T.; Sałat, K.; Gaweł, M.; Moniczewski, A.; Lochynski, S. An Overview of the Pharmacological Properties and Potential Applications of Natural Monoterpenes. Mini Rev. Med. Chem. 2014, 14, 1156–1168. [Google Scholar] [CrossRef]
  229. Kutney, J.P.; Chen, Y.H. The Chemistry of Thujone. XVII. The Synthesis of Ambergris Fragrances and Related Analogues. Can. J. Chem. 1994, 72, 1570–1581. [Google Scholar] [CrossRef]
  230. Barrero, A.F.; Altarejos, J.; Alvarez-Manzaneda, E.J.; Ramos, J.M.; Salido, S. Synthesis of (±)-Ambrox from (E)-Nerolidol and β-Ionone via Allylic Alcohol [2,3] Sigmatropic Rearrangement. J. Org. Chem. 1996, 61, 2215–2218. [Google Scholar] [CrossRef]
  231. Aoki, T.; Ataka, Y. Process for Production of (+)-3a,6,6,9a-Tetramethyldecahydro-naphtho [2,1-b]furan-2(1H)-one. U.S. Patent 8,153,826 B2, 10 April 2012. [Google Scholar]
  232. Serra, S. Recent Developments in the Synthesis of the Flavors and Fragrances of Terpenoid Origin, Studies in Natural Products Chemistry; Elsevier BV: Amsterdam, The Netherlands, 2015; Volume 46, pp. 201–226. [Google Scholar]
  233. Benites, J.; Lopez, J.; Farias, J.G.; Cortes, M. The Preparation of Oxygenated Derivatives of Ambrox and Isoambrox from Drimenol. J. Chil. Chem. Soc. 2006, 51, 979–981. [Google Scholar] [CrossRef]
  234. Castro, J.M.; Salido, S.; Altarejos, J.; Nogueras, M.; Sánchez, A. Synthesis of Ambrox® from Labdanolic Acid. Tetrahedron 2002, 58, 5941–5949. [Google Scholar] [CrossRef]
  235. Bolster, M.G.; Jansen, B.J.; De Groot, A. The Synthesis of Ambrox®-Like Compounds Starting From (+)-Larixol. Tetrahedron 2001, 57, 5663–5679. [Google Scholar] [CrossRef]
  236. Cheng, T.; Zhao, G.; Xian, M.; Xie, C. Improved cis-Abienol Production through Increasing Precursor Supply in Escherichia coli. Sci. Rep. 2020, 10, 16791. [Google Scholar] [CrossRef]
  237. Zerbe, P.; Bohlmann, J. Enzymes for Synthetic Biology of Ambroxide-Related Diterpenoid Fragrance Compounds. Adv. Biochem. Eng. Biotechnol. 2015, 148, 427–447. [Google Scholar]
  238. Martins, M.P.; Ouazzani, J.; Arcile, G.; Jeller, A.H.; de Lima, J.P.; Seleghim, M.H.; Oliveira, A.L.; Debonsi, H.M.; Venâncio, T.; Yokoya, N.S.; et al. Biohydroxylation of (−)-ambrox®, (−)-sclareol, and (+)-sclareolide by whole Cells of Brazilian Marine-derived Fungi. Mar. Biotechnol. 2015, 17, 211–218. [Google Scholar] [CrossRef]
  239. Shen, Y.C.; Cheng, S.Y.; Kuo, Y.H.; Hwang, T.L.; Chiang, M.Y.; Khalil, A.T. Chemical Transformation and Biological Activities of Ambrein, A Major Product of Ambergris from Physeter macrocephalus (Sperm Whale). J. Nat. Prod. 2007, 70, 147–153. [Google Scholar] [CrossRef]
  240. Al-Khalil, S. A review on Ambergris Perspective and Modern Chemical Composition and Pharmacology. Acad. J. Plants 2020, 8, 96–101. [Google Scholar]
  241. Sandroni, P. Aphrodisiacs Past and Present: A Historical Review. Clin. Auton. Res. 2001, 11, 303–307. [Google Scholar] [CrossRef]
  242. Taha, S.A. Studies on the Mode of Action of Ambrein as a New Antinociceptive Compound. Jpn. J. Pharmacol. 1992, 60, 67–71. [Google Scholar] [CrossRef]
  243. Gode, P.K. History of Ambergris in India between about A.D. 700 and 1900. Chymia 1949, 2, 51–56. [Google Scholar] [CrossRef]
  244. Fahlbusch, K.G.; Hammerschmidt, F.J.; Panten, J.; Pickenhagen, W.; Schatkowski, D.; Bauer, K.; Garbe, D.; Surburg, H. Flavors and Fragrances. In Ullmann’s Encyclopedia of Industrial Chemistry, 7th ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2003; Volume 15, pp. 73–198. [Google Scholar]
  245. Bouchet, P. The Magnitude of Marine Biodiversity. In The Exploration of Marine Biodiversity: Scientific and Technological Challenges; Duarte, C.M., Ed.; Fundación BBVA: Bilbao, Spain, 2006; pp. 31–62. [Google Scholar]
  246. May, R.M. Tropical Arthropod Species, More or Less? Science 2010, 329, 41–42. [Google Scholar] [CrossRef] [PubMed]
  247. Saeedi, H.; Reimer, J.D.; Brandt, M.I.; Dumais, P.; Jażdżewska, A.M.; Jeffery, N.W.; Thielen, P.M.; Costello, M.J. Global marine Biodiversity in the Context of Achieving the Aichi Targets: Ways Forward and Addressing Data Gaps. Peer J. 2019, 7, e7221. [Google Scholar] [CrossRef] [PubMed]
  248. El Hattab, M. Algae Essential Oils: Chemistry, Ecology, and Biological Activities. In Essential Oils—Bioactive Compounds, New Perspectives and Applications; de Oliveira, M.S., Silva, S., Da Costa, W.A., Eds.; IntechOpen: London, UK, 2020; pp. 1–20. [Google Scholar]
  249. Moore, R.E. Volatile Compounds from Marine Algae. Acc. Chem. Res. 1977, 10, 40–47. [Google Scholar] [CrossRef]
  250. El Hattab, M.; Culioli, G.; Piovetti, L.; Chitour, S.E.; Valls, R. Comparison of Various Extraction Methods for Identification and Determination of Volatile Metabolites from the Brown Alga Dictyopteris membranacea. J. Chromatogr. A 2007, 1143, 1–7. [Google Scholar] [CrossRef]
  251. Cotas, J.; Leandro, A.; Monteiro, P.; Pacheco, D.; Figueirinha, A.; Gonçalves, A.; Silva, G.; Pereira, L. Sea-weed Phenolics: From Extraction to Applications. Mar. Drugs 2020, 18, 384. [Google Scholar] [CrossRef]
  252. Hakim, M.M.; Patel, I.C. A Review on Phytoconstituents of Marine Brown Algae. Future J. Pharm. Sci. 2020, 6, 1–11. [Google Scholar] [CrossRef]
  253. Cabrita, M.T.; Vale, C.; Rauter, A.P. Halogenated Compounds from Marine Algae. Mar. Drugs 2010, 8, 2301–2317. [Google Scholar] [CrossRef] [Green Version]
  254. Tang, K. Chemical Diversity and Biochemical Transformation of Biogenic Organic Sulfur in the Ocean. Front. Mar. Sci. 2020, 7, 68. [Google Scholar] [CrossRef] [Green Version]
  255. Francioso, A.; Conrado, A.B.; Mosca, L.; Fontana, M. Chemistry and Biochemistry of Sulfur Natural Compounds: Key Intermediates of Metabolism and Redox Biology. Oxid. Med. Cell Longev. 2020, 2020, 1–27. [Google Scholar] [CrossRef]
  256. Garcia-Jimenez, P.; Brito-Romano, O.; Robaina, R.R. Production of Volatiles by the Red Seaweed Gelidium arbuscula (Rhodophyta): Emission of Ethylene and Dimethyl Sulfide. J. Phycol. 2013, 49, 661–669. [Google Scholar] [CrossRef]
  257. Challenger, F.; Simpson, M.I. Studies on Biological Methylation. Part XII. A Precursor of the Dimethyl Sulphide Evolved by Polysiphonia fastigiata. Dimethyl-2-carboxyethyl Sulphonium Hydroxide and its Salts. J. Chem. Soc. 1948, 159, 1–7. [Google Scholar]
  258. Flodin, C.; Whitfield, F. 4-Hydroxybenzoic acid: A likely precursor of 2,4,6-tribromo phenol in Ulva lactuca. Phytochemistry 1999, 51, 249–255. [Google Scholar] [CrossRef]
  259. Cikoš, A.-M.; Jurin, M.; Čož-Rakovac, R.; Jokić, S.; Jerković, I. Update on Monoterpenes from Red Macroalgae: Isolation, Analysis, and Bioactivity. Mar. Drugs 2019, 17, 537. [Google Scholar] [CrossRef] [Green Version]
  260. Cikoš, A.M.; Jurin, M.; Coz-Rakovac, R.; Gašo-Sokač, D.; Jokic, S.; Jerkovic, I. Update on Sesquiterpenes from Red Macroalgae of the Laurencia Genus and their Biological Activities (2015–2020). Algal. Res. 2021, 56, 102330. [Google Scholar] [CrossRef]
  261. Blunt, J.; Carroll, A.; Copp, B.; Davis, R.; Keyzers, R.; Prinsep, M. Marine Natural Products. Nat. Prod. Rep. 2018, 35, 8–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Boland, W. The Chemistry of Gamete Attraction: Chemical Structures, Biosynthesis, and (a)Biotic Degradation of Algal Pheromones. Proc. Natl. Acad. Sci. USA 1995, 92, 37–43. [Google Scholar] [CrossRef] [Green Version]
  263. Zatelli, G.A.; Philippus, A.C.; Falkenberg, M. An Overview of Odoriferous Marine Seaweeds of the Dictyopteris genus: Insights into their Chemical Diversity, Biological Potential and Ecological Roles. Rev. Bras. Farmacogn. 2018, 28, 243–260. [Google Scholar] [CrossRef]
  264. Maier, I.; Muller, D.G. Sexual Pheromones in Algae. Biol. Bull. 1986, 170, 145–175. [Google Scholar] [CrossRef]
  265. Riad, N.; Reda Zahi, M.R.; Trovato, E.; Bouzidi, N.; Daghbouche, Y.; Utczás, M.; Mondello, L.; El Hattab, M. Chemical Screening and Antibacterial Activity of Essential Oil and Volatile Fraction of Dictyopteris polypodioides. Microchem. J. 2020, 152, 104415. [Google Scholar] [CrossRef]
  266. Baldovini, N.; Chaintreau, A. Identification of Key Odorants in Complex Mixtures Occurring in Nature. Nat. Prod. Rep. 2020, 37, 1589–1626. [Google Scholar] [CrossRef] [PubMed]
  267. Moore, R.E.; Pettus, J.A.; Dictyopterene, A. An Odoriferous Constituent from Algae of the Genus Dictyopteris. Tetrahedron Lett. 1968, 9, 4787–4790. [Google Scholar] [CrossRef]
  268. Ohloff, G.; Pickenhagen, W. Synthese von (±)-dictyopterene. A HeIv. Chim. Acta. 1969, 52, 880–886. [Google Scholar] [CrossRef]
  269. Stratmann, K.; Boland, W.; Müller, D.G. Biosynthesis of Pheromones in Female Gametes of Marine Brown Algae (Phaeophyceae). Tetrahedron 1993, 49, 3755–3766. [Google Scholar] [CrossRef]
  270. Schnitzler, I.; Boland, W.; Hay, M.E. Organic Sulfur Compounds from Dictyopteris spp. Deter feeding by an Herbivorous Amphipod (Ampithoe longimana) but not by an Herbivorous Sea Urchin (Arbacia punctulata). J. Chem. Ecol. 1998, 24, 1715–1732. [Google Scholar] [CrossRef]
  271. Dimou, M.; Ioannou, E.; Daskalaki, M.G.; Tziveleka, L.A.; Kampranis, S.C.; Roussis, V. Disulfides with Anti-Inflammatory Activity from the Brown Alga Dictyopteris membranacea. J. Nat. Prod. 2016, 79, 584–589. [Google Scholar] [CrossRef]
  272. Roller, P.; Kalfred, A.; Moore, R.E. Isolation of S-(3-oxoundecyl)thioacetate, bis-(3-oxoundecyl) disulphide, (−)-3-hexyl-4,5-dithiacycloheptanone, and S-(trans-3-oxoundec-4-enyl) thioacetate from Dictyopteris. Chem. Commun. 1971, 273, 503–504. [Google Scholar] [CrossRef]
  273. Wratten, S.J.; Faulkner, D.J. Cyclic Polysulfides from the Red Alga Chondria Californica. J. Org. Chem. 1976, 41, 2465–2467. [Google Scholar] [CrossRef]
  274. Hay, M.E.; Duffy, J.E.; Fenical, W.; Gustaíson, K. Chemical Defense in the Seaweed Dictyopteris delicatula: Differential Effects Against Reef Fishes and Amphipods. Mar. Ecol. Prog. Ser. 1988, 48, 185–192. [Google Scholar] [CrossRef]
  275. Oigman, S.S.; Fernandes, Y.F.M.; Teles, D.; Maia, L.F.; Epifanio, R.A.; Rezende, C.M. Brazilian Gorgonians: A Source of Odoriferous Compounds? Rev. Bras. Farmacogn. 2015, 25, 612–618. [Google Scholar] [CrossRef] [Green Version]
  276. Triqui, R. Sensory and Flavor Profiles as a Means of Assessing Freshness of Hake (Merluccius merluccius) during Ice Storage. Eur. Food. Res. Technol. 2006, 222, 41–47. [Google Scholar] [CrossRef]
  277. Venkateshwarlu, G.; Let, M.B.; Meyer, A.S.; Jacobsen, C. Modeling the Sensory Impact of Defined Combinations of Volatile Lipid Oxidation Products on Fishy and Metallic Off-Flavors. J. Agric. Food Chem. 2004, 52, 1635–1641. [Google Scholar] [CrossRef]
  278. Hammer, M.; Schieberle, P. Model Studies on the Key Aroma Compounds Formed by an Oxidative Degradation of ω-3 Fatty Acids Initiated by either Copper(II) Ions or Lipoxygenase. J. Agric. Food Chem. 2013, 61, 10891–10900. [Google Scholar] [CrossRef] [PubMed]
  279. Spyropoulou, E.A.; Dekker, H.L.; Steemers, L.; van Maarseveen, J.H.; de Koster, C.G.; Haring, M.A.; Schuurink, R.C.; Allmann, S. Identification and Characterization of (3Z):(2E)-Hexenal Isomerases from Cu-cumber. Front. Plant Sci. 2017, 8, 1342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Kajiwara, T.; Matsui, K.; Akakabe, Y.; Kawai, T.; Ishihara, M. Preparation of Dictyopterene B Isomers and Fragrance Compositions Containing Them Jpn Kokai Tokkyo Koho JP 003137819. Apud Chem Abstr. 2003, 138, 369034. [Google Scholar]
  281. Milchakova, N. Marine Plants of the Black Sea. An Illustrated Field Guide; Digit Print Press: Sevastopol, Ukraine, 2011; pp. 37–120. [Google Scholar]
  282. Ramesh, C.H.; Koushik, S.; Shunmugaraj, T.; Murthy, M.V.R. A Red Alga Portieria hornemannii (Lyngb.) P. C. Silva 1987 (Gigartinales, Rhizophyllidaceae): A Source of Fragrance Ingredient for Perfume Industry. Indian J. Mar. Sci. 2020, 49, 898–902. [Google Scholar]
  283. Allen, J.A.; Duke, N.C. Bruguiera Gymnorrhiza (Large-Leafed Mangrove), ver. 2.1. In Species Profiles for Pacific Island Agroforestry; Elevitch, C.R., Ed.; Permanent Agriculture Resources (PAR): Hōlualoa, HI, USA, 2006; pp. 1–14. [Google Scholar]
  284. Adkar, P.P.; Bhaskar, V.H. Pandanus odoratissimus (Kewda): A review on ethnopharmacology, phytochemistry, and nutritional aspects. Advan. Pharmacol. Sci. 2014, 120895, 1–19. [Google Scholar] [CrossRef] [Green Version]
  285. Uppala, L. A Review on Active Ingredients from Marine Sources used in Cosmetics. SOJ Pharm. Pharm. Sci. 2015, 2, 1–3. [Google Scholar] [CrossRef] [Green Version]
  286. Guillerme, J.-B.; Couteau, C.; Coiffard, L. Applications for Marine Resources in Cosmetics. Cosmetics 2017, 4, 35. [Google Scholar] [CrossRef] [Green Version]
  287. Couteau, C.; Coiffard, L. Phycocosmetics and Other Marine Cosmetics, Specific Cosmetics Formulated Using Marine Resources. Mar. Drugs 2020, 18, 322. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of the main volatile compounds of essential oils.
Figure 1. Chemical structure of the main volatile compounds of essential oils.
Chemistry 03 00103 g001
Figure 2. Structure of benzodioxepinone and benzoxepinone skeletons, Calone and analogues.
Figure 2. Structure of benzodioxepinone and benzoxepinone skeletons, Calone and analogues.
Chemistry 03 00103 g002
Figure 3. Calone 1951 olphactophore model and derivatives.
Figure 3. Calone 1951 olphactophore model and derivatives.
Chemistry 03 00103 g003
Figure 4. Specific compounds exhibiting marine odor.
Figure 4. Specific compounds exhibiting marine odor.
Chemistry 03 00103 g004
Figure 5. Flexiblity of the system in relation to the seven-membered cycle.
Figure 5. Flexiblity of the system in relation to the seven-membered cycle.
Chemistry 03 00103 g005
Figure 6. Bendioxepinone and analog compounds.
Figure 6. Bendioxepinone and analog compounds.
Chemistry 03 00103 g006
Figure 7. Typical marine odour compounds.
Figure 7. Typical marine odour compounds.
Chemistry 03 00103 g007
Figure 8. Amber and the most characteristic ambergris smelling compounds.
Figure 8. Amber and the most characteristic ambergris smelling compounds.
Chemistry 03 00103 g008
Figure 9. Odoriferous C11 hydrocarbons from Dictyopteris species.
Figure 9. Odoriferous C11 hydrocarbons from Dictyopteris species.
Chemistry 03 00103 g009
Table 1. Sensory properties of some volatile compounds.
Table 1. Sensory properties of some volatile compounds.
No.CompoundsOdor QualityThreshold Range Reported in Literature (ng/L)AI(OAV)
WaterAirOther Medium
1β-Myrcene(pleasant floral) [133],
(green, woody) [134]
0.013–0.015 [60,135]0.041–0.15 [136,137]1.0 [138]4.2 [134]42 [139]
36 [139]
2(Z)-β-Ocimene(warm floral, herbal, sweet) [134]0.034–0.055 [140,141]0.001 [137]-3.6 [134]16 [142,143]
3(E)-β-Ocimenesweet and herbal [142]0.034 [141]0.0187 [137]-1.7 [144]
4Linalool(green, floral, sweet) [134], (grassy, pleasant,citrus) [133]0.006 [141,145]-0.067–0.113 [146]11 [147]˂1 [139]
8 [139]
5Nerol(sweet, citrus) [134] 0.049 [148]0.5 [149,150]0.8 [134]˂1 [142,143]
6Geraniol(floral, fruity) [144], (fresh, rose-like) [133]0.0011–0.006 [60,151]0.6 [152]0.005 [153]8 [147]˂1 [142,143]
7Neral(Citric, green) [154]0.053 [155]0.0088 [156]--1330 [157]
8CitralLemony [158]0.04–0.12 [159]0.00015 [160]0.656–1.23 [149]
9Limonene(citrus, sweet) [134], (Strong odor of orange) [158]0.2 [161]0.0539 [162]14.7 [161]4.7 [134]228 [139]
134 [139]
10γ-Terpinene(Woody, lemon, tropical, herbal) [163]1 [145]55 [162]2.39–3.26 [146]2.0 [134]˂1 [142,143]
11p-Cymene(Fresh, citrus, terpene, woody, spice) [163]0.00501 [164]7.2 [152]66 [138]--
12α-Terpineol(Sweet, lilac odor) [133] (fresh and minty) [144]1.2 [165]0.86 [166]0.46 [167]10 [147]-
13Carvone(Minty herbaceous) [154]0.027 [145]0.0002 [162]0.4–0.6 [168] 1620 [157]
14Camphene(woody, herbal) [134]1.86 [169][0–26] [170] 1.0 [134]˂1 [142,143]
15 β-Pinene(dry woody, green) [134]0.14 [145]0.18 [133]37.2–38.7 [146]1.3 [134]˂1 [142,143]
16∆3-Carene(sweet, fruity) [134] (woody turepentine) [133]0.77 [145]9.3 [152]-4.5 [134]124 [133,134]
171,8-Cineole(Eucalyptus-like, fresh/pungent) [171]0.0011 [172]-0.0013 [60]--
18Linalyl acetate(Floral, sweet citrus) [156]1 [173,174]4.0–6.0 [175]10 [172]--
19Bornylacetate(Woody, camphor, mentholic, spicy) [143,176]1.38 [173,174]0.44 [177]0.08 [178]0.1 [134]-
20Farnesol(Flowery, weak-citrus odor) [158]1 [179]-5 [180]5 [147]-
21 β-Caryophyllene(sweet, woody, spice) [134] 0.064 [145]-8 [147]8 [149]14 [142,143]
22(3R)-(−)-3-Mercaptohexan-1-olfruitier, with a zesty aroma reminiscent of grapefruit [181]-0.08 [182]---
23(3S)-(+)-3-Mercaptohexan-1-olpassion fruit [182]-0.07 [181]---
24p-Menthene-8-thiolgrapefruit-like [134]0.000002 [183]0.00000 [184]---
25Carvacrolmedicinal, tarry [185]2.29 [145]0.18 [186]30.97 [187]--
26ThymolSweet, herbal [185]1.7 [145]0.1 [188]124 [187]--
Table 2. Olfactory profile of five-membered ring compounds.
Table 2. Olfactory profile of five-membered ring compounds.
No.CompoundsOlfactory Profile
9 Chemistry 03 00103 i001Aldehydic, green, marine tenacious, powerful
10 Chemistry 03 00103 i002Phenolic, crab, oakmoss, slightly watery
11 Chemistry 03 00103 i003Watermelon, aldehydic, Aldolone, cyclosal, green, oyster, ozone, watery
12 Chemistry 03 00103 i004Aquozone, marine, fruity, green, rubbery
13 Chemistry 03 00103 i005Watery, fatty, floral, aldehydic
14 Chemistry 03 00103 i006Watery, floral, marine, watermelon
Table 3. Key aroma compounds with marine odor formed from autoxidized unsaturated fatty acids.
Table 3. Key aroma compounds with marine odor formed from autoxidized unsaturated fatty acids.
No.Odorant CompoundsOlfactory ProfileOdor Threshold (ng/L in Air)
46(Z)-4-heptenal fishy0.8
47Trans-4,5-epoxy-(E,Z)-2,7-decadienalmetallic, pungent0.01
48(Z,Z)-2,5-octadienalsweet, melon-like0.08
49(E,Z)-2,5-octadienalsweet, melon-like0.7
50(E,E,Z)-2,4,7-decatrienalfatty, pungent0.8
51(E,E,E)-2,4,7-decatrienalfatty, cucumber-like0.2
52(E,Z)-2,6-nonadienalcucumber-like0.01
53(Z,Z)-3,6-nonadienalfresh, watermelon-like-
54(Z,Z)-2,5-octadienalsweet, melon-like0.08
55(Z)-1,5-octadien-3-onegeranium-like9
56(E,E,E)-2,4,7-decatrienalfatty, cucumber-like0.2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Riad, N.; Zahi, M.R.; Bouzidi, N.; Daghbouche, Y.; Touafek, O.; El Hattab, M. Occurrence of Marine Ingredients in Fragrance: Update on the State of Knowledge. Chemistry 2021, 3, 1437-1463. https://doi.org/10.3390/chemistry3040103

AMA Style

Riad N, Zahi MR, Bouzidi N, Daghbouche Y, Touafek O, El Hattab M. Occurrence of Marine Ingredients in Fragrance: Update on the State of Knowledge. Chemistry. 2021; 3(4):1437-1463. https://doi.org/10.3390/chemistry3040103

Chicago/Turabian Style

Riad, Nacera, Mohamed Reda Zahi, Naima Bouzidi, Yasmina Daghbouche, Ouassila Touafek, and Mohamed El Hattab. 2021. "Occurrence of Marine Ingredients in Fragrance: Update on the State of Knowledge" Chemistry 3, no. 4: 1437-1463. https://doi.org/10.3390/chemistry3040103

Article Metrics

Back to TopTop