Next Article in Journal
Protective Application of Chinese Herbal Compounds and Formulae in Intestinal Inflammation in Humans and Animals
Previous Article in Journal
Primary Products from Fast Co-Pyrolysis of Palm Kernel Shell and Sawdust
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of Spirits Made from Sugarcane Juice

by
Claudine Corbion
1,
Juliette Smith-Ravin
2,
Odile Marcelin
2 and
Jalloul Bouajila
1,*
1
Laboratoire de Génie Chimique, Université de Toulouse, CNRS-INPT-UPS, 31062 Toulouse, France
2
Groupe BIOSPHERES, Campus de Schoelcher, 97275 Schoelcher, Martinique, France
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(19), 6810; https://doi.org/10.3390/molecules28196810
Submission received: 1 August 2023 / Revised: 5 September 2023 / Accepted: 23 September 2023 / Published: 26 September 2023
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
Among the family of sugarcane spirits, those made from juice are diverse and often produced in a traditional way. They must be distinguished from other sugarcane spirits, which are more widely produced and made from other sugarcane derivatives, such as molasses. These alcoholic beverages contribute significantly to the socio-economic development of many countries. However, despite ancestral know-how, there is a lack of contemporary data required to characterize some sugarcane juice spirits (SCJSs) and to overcome the current and future threats that producers will have to face. While preserving their authenticity and specificity, SCJS producers expect to improve and ensure sufficient yield and a superior quality product. Even if the scientific knowledge on these spirits is not comparable, the available data could help identify the critical points to be improved in the making process. This review aims to present the main SCJSs encountered worldwide, defining their specific features through some important aspects with, notably, references to the complex notion of terroir. To continue, we discuss the main steps of the SCJS process from harvesting to aging. Finally, we expose an inventory of SCJS’s chemical compositions and of their sensory description that define the specific organoleptic properties of these spirits.

1. Introduction

The term ‘spirit’ refers to any alcoholic beverage made by the fermentation and distillation of cereals, fruits or their sub-products. The history of spirits is related to the development of distillation techniques throughout the centuries. Originally used during ancient times by Persians, Egyptians and Sumerians to obtain perfume and essential oils, this technique was used during the Middle Ages to produce alcohol, called aqua vitae, employed in both the medicine and the pharmaceutical sectors to fight against diseases [1]. As a result of the development and industrialization of distillation techniques during the 19th century, the production of spirits expanded across the world with the raw materials that were available [1]. Therefore, the world of spirits is very large and includes baijiu, brandies, rum, shōchū, tequila, vodka and whiskies (Table 1). According to the European Union (EU) regulation 2019/787, a spirit is an alcoholic beverage that possesses particular organoleptic qualities. It is obtained after the distillation of agricultural raw material with a minimum alcohol content of 15% alcohol by volume (ABV) [2]. In 2021, the spirit market worldwide reached 473.6 billion USD and it is expected to increase annually by almost 5% over the next five years [3]. The world’s top spirits are made from fermented cereals or grains such as baijiu from China. The grape-based spirit category is mainly found in European countries, whereas those made from sugarcane and its derivatives (juice, molasses, honey or syrup) are produced in tropical areas (Table 1).
Many countries worldwide have established quality schemes to protect the consumer’s interest and to promote their local and traditional products. In the EU, protected geographical indication helps producers to fight against counterfeiting and ensure that the consumer knows the true origin of the product. In each country, competent authorities control those products comply with quality standards. In France, a Protected Designation of Origin (PDO) called AOC, which stands for ‘Appellation d’Origine Contrôlée’ goes further in the protection of spirits by highlighting the know-how and the specific characteristics linked to the agricultural properties (terroir) of products. The only spirit made from sugarcane juice in a French overseas department with an AOC is the agricultural rum, called ‘Rhum Agricole’, from Martinique [4,5].
The alcoholic beverages discussed in this review are spirits made from fresh sugarcane juice (SCJS). These spirits need to be distinguished from other sugarcane spirits more widely produced, which are made from other sugarcane derivatives, such as molasses [6]. The objective is to gather the information currently available on SCJS and to define their chemical and sensory profiles.
This review deals with the sugarcane, which is the first element to ensure SCJS’s quality. Then, the notion of terroir will be discussed to understand the characteristics that explain the differences or similarities between the SCJSs presented here through their manufacturing process and traditional practices and uses. Finally, the last section will emphasize the chemical and sensorial characteristics of SCJS. A better knowledge of the chemical composition and organoleptic properties of the spirit is essential to promote continuous quality improvement. This strategy has been used for several decades by some producers. As was the case for Rhum Agricole from Martinique during the 1990s, recent studies conducted on cachaça have helped to characterize the profile of this Brazilian SCJS [7,8]. Improving the quality and yield of SCJSs is an economic development strategy that has led to the growth of ‘spiritourism’ [9,10]. Using the current information available in the literature can help producers respond to contemporary challenges (‘drink less but better’ and other environmental and social issues) and promote their traditional products.
An extensive literature is currently available on cachaça. It could serve as a basis for studying the chemical and sensory characterization of other SCJSs.

2. From the Raw Material to the Spirit

The geographical origin of sugarcane (Saccharum spp.) is still unclear [11]. A current model suggests that modern sugarcane came from the New Guinea area around 8000 BP and was spread to India and China where the respective cultivars S. barberi and S. sinense were engendered by the hybridization of S. officinarum with S. spontaneum [12,13,14] (Table 2). However, a new model indicates the existence of a new cryptic species called S. cultum, around 650,000 BP in East Asia, which could be the ancestor of modern cultivars [15].
Human activities led to the migration of these sugarcane varieties across the Pacific and Asian continent around 6000 BP [19], reaching the West around 2400 BP [15]. The crop was introduced to the island of Madeira around 1450 [20], reached the American continent after the second voyage of Christopher Columbus in Santo Domingo [21] and was introduced to the other regions of America such as Brazil, Guadeloupe and Martinique, respectively, in 1520, 1645 and 1650 [20]. It was introduced recently in the 17th century to the Reunion and Mauritius islands and in the 18–19th century to South Africa and Australia [19].
Nowadays, sugarcane is still an important food crop in tropical and subtropical climatic zones such as Central and South America, the Caribbean, India, Africa and Asia-Pacific. In 2021, this crop was cultivated by nearly 100 countries and almost 2 billion tons of cane were harvested. The top 3 producers were Brazil (715 million t), India (405 million t) and China (213 million t) [22].

2.1. Agronomic Characteristics and Agricultural Concerns

Modern sugarcane is a polyploid hybrid of species such as S. officinarum, S. spontaneum, S. robustum, S. barberi and S. sinensi [17,19]. This crop is one of the most genetically complex species due to its large chromosome numbers (Table 2). This perennial C4 plant produces efficient photosynthesis, especially at high temperatures, [23]. The stems grow in clumps, and each unbranched stem is a succession of nodes and internodes. The diameter of the stalk is about 2.5–5 cm and the length is 2–5 m. The leaves reach a length of about 30–100 cm and are alternatively attached to the stem at the base of the nodes (see leaf scar of Figure 1). The root system is composed of adventitious roots emerging from the stem and permanent fasciculated roots. The inflorescence, also called the arrow, is a ramified panicle of 25–30 cm in length.
Sugarcane cultivation starts with planting a stalk piece in the soil. Several stems grow to form a clump that ripens after 12 to 18 months [20]. The cane is mature, and therefore ready to harvest, when the sugar content at the top is almost as high as at the bottom. The plant–cane cycle ends after the first harvest, and the roots and lower parts of the plant are left in the field to allow the growth of new systems. After four or five ratoon–cane cycles, the crop is eradicated from the field by ploughing [11,16].
To fight pests (borer, nematodes, termites or beetles) and diseases, the breeding programs produced new resistant cultivars [23,25]. The other threats to sugarcane cultivation are weeds. For example, in Bangladesh, weed competition leads to yield losses of 40%, compared to 20% by pests [26]. The crop suffers from competition for water and nutrients from weeds, which provide a breeding ground for disease-causing insects.
To control invasive plants, the application of phytopharmaceutical products is widely favored. However, those practices depend on the regulations of the country. Over the last decade, in the EU, the list of approved herbicides, such as Atrazine, Asulam-based herbicides and glyphosate-based products, has decreased, triggering some difficulties in weed control [27,28,29]. The current general trend is to reduce the use of herbicides by adopting new weed control strategies, including innovative cultural and biological technology [25,30].
Pre-harvest burning of sugarcane is the most widely used weed control method, but it has a negative effect on SCJS’s quality. The study of Galinaro et al. indicates that the concentration of total polycyclic aromatic hydrocarbons (PAHs), restricted in food in the EU because of their toxicity, is more than 10 times higher in cachaça produced from burned sugarcane crops (21.1 μg/L) than from unburned crops (1.9 μg/L) [31]. Another study from Thai and Doherty reports that cane juice from unburned cane has a higher proportion of soluble inorganic ions, organic acids, proteins and polysaccharides than that from burned cane [32]. The concentrations of furfural and 5-hydroxymethylfurfural (5-HMF) are also higher in cachaça from burned sugarcane [8]. By contrast, manual weed control, mostly used by smaller producers, is effective, but it can be difficult for farmers to find the necessary workforce for this laborious task [11].
To summarize, weed management requires more diverse and effective methods to respond to the needs of all types of farmers [33]. Moreover, the other greatest current threats to sugarcane cultivation are related to climate change [34]. The lack of water and the increase in temperatures could negatively impact the sugarcane quality [30] and therefore threaten the production of SCJSs. For example, weeds such as Rottboellia cochinchinensis, Ipomea plebeia and Digitaria sanguinalis are expected to increase under high-temperature conditions [34].

2.2. General Uses and Chemical Composition of Sugarcane

Sugarcane is widely used industrially and both fractions obtained from the milling, juice and bagasse, have a wide range of applications (Figure 2). Cane juice is a nutritious energy drink, extracted simply by chewing [35] or with a mechanical crusher [36]. It represents 80–90% of the cane and it is composed of 70–80% of water, 15–25% of soluble solids composed of fermentable carbohydrates, 0.8 to 2% of other organic compounds (wax, amino acids and starch) and 0.8 to 2% of inorganics compounds (Figure 2) [37]. The fibrous residue (10–20%) called bagasse and containing at least 49% moisture, is composed of approximately 33–45% of cellulose, 18–29% of hemicellulose and 19–32% of lignin [31].
Sugarcane provides 60–70% of the sugar produced in the world and other various valuable products like jaggery, syrup, cane honey and molasses [35]. It is important to acknowledge the specificity of small territories such as islands where there is strong competition between producers using the same raw material. This leads to a reduction in the availability of sugarcane for SCJS’s production.
Sugarcane juice has a color ranging from light gray to dark green due to plant pigments (anthocyanins and chlorophyll), polyphenolic compounds, the degradation products of sugars condensed with amino derivatives, and the reaction of ferric salts with the tannins. Colloidal matter found in the fresh juice is responsible for turbidity [36].
The main compound is sucrose [37]. Its concentration increases from the stem bottom to the top during the growth. Starch is found in the stalk and in green leaves but is absent in roots. Its content varies during the day and it is converted into sugar during the night [19]. In the case of alcoholic beverage production, unlike sugar production, the non-sugar composition (vitamins, nitrogenous compounds, etc.) must be suitable for alcoholic fermentation, and this will impact the fermentation yield [20] (Table 3).
As reported by Chen et al., the main aroma characteristics of sugarcane are green and grass odor [39] as contributed by 1-hexanol, trans-2-octenal, nonanal and trans-2-nonenal. Floral and honey aromas are associated with phenylethyl alcohol and 2-ethyl-1-hexanol. The content of carboxylic acids, alcohols and esters increases in the juice, and some compounds not detected in sugarcane are present such as caprylic alcohol, d-limonene, ethyl palmitate and trans-2-nonenal. This latter compound could appear due to the microbial and enzymatic effects during the juice extraction (Table 4).

3. Styles and Methods of Production of SCJS

3.1. The Notion of Terroir

Terroir is a French word that has the Latin root Terra meaning ‘earth’ or ‘soil’ [40]. Frequently used in the wine industry, the notion of terroir refers to the natural specification linked to the geographical origin of the product [41]. This concept makes it possible to define characteristics that influence or explain the unique sensory and chemical profiles of a wine and has been also revealed for several spirits such as whiskey [42], cognac [43] or cachaça [44]. Producers of agricultural rum from Martinique have made use of this concept to differentiate their products thanks to the AOC designation [5].
Hence, the diversity of the SCJS sensory profile depends not only on the terroir-related elements but also on each step of the production process (Figure 3). For a freshly distilled spirit (non-aged SCJS), the factors connected to the sugarcane can greatly influence the flavor [31]. Furthermore, fermentation [45,46,47,48] and distillation [49,50,51] parameters play an important role in the organoleptic characteristics of the spirit. For aged SCJSs, the different conditions of maturation in wooden barrels define the quality of the product [10,52,53,54,55].

3.2. Sugarcane Juice Spirits in the World

The large family of sugarcane spirits includes those made from the thermal processing of the juice (Figure 4). The area of production determines the designation of ‘rhum’, ‘ron’ or rum according to whether the production style is French, Spanish or English [56,57]. Regarding the other category of sugarcane spirit, those made from the pure juice are also diverse with their own denominations and specific organoleptic properties.
Nowadays, SCJSs are produced worldwide. Figure 5 presents only those produced with pure sugarcane juice (non-heated). This list was compiled by searching spirits vendor websites, producer websites and books. This narrows the list considerably, as some producers blend SCJSs with molasses. The most famous SCJSs are cachaça, a Brazilian spirit and rum produced mainly in the Caribbean with hundreds of artisanal or industrial producers. Clairin, from Haiti, is produced with hundreds of units. On the other side of the Atlantic Ocean, in Cape Verde, the grogue with only pure cane juice is produced by two distilleries. We can see that the emergence of new SCJS producers in areas dominated by molasses-based rums is slowly changing the market. However, for the moment, in the Asia/Oceania area, 13 distilleries producing pure cane juice alcohol are reported with 3 distilleries in Japan and Vietnam; 2 distilleries in Thailand, the 5th largest producer of sugarcane in the world, as well as Australia; and 1 distillery each in India, Cambodia and Laos. In the Indian Ocean, the islands of Mauritius (three), Reunion (three) and Seychelles (one) accounted for seven distilleries. Only one distillery producing pure cane juice rum could be identified in South Africa. On the American continent, there are, respectively, three and two distilleries located in the USA and in Mexico. And finally, in the Pacific Ocean, four distilleries in Hawaii (one) and Tahiti (three) produce SCJS.
The following chapters briefly present two well-documented SCJSs, Rhum Agricole and cachaça, and three other SCJSs, which are less documented in the literature: clairin, grogue and ‘aguardiente de cana’.

3.3. Rum and Rhum Agricole

According to European law, rum is exclusively produced by the distillation of fermented material issues of sugarcane (Saccharum officinarum spp.) such as the molasses, the syrup, or the juice. It is distilled at less than 96% ABV and has the specific organoleptic characteristics of rum. The rum is not flavored and the alcoholic strength is at least 37.5% ABV [58]. The term ‘traditional’ could be added for an unsweetened rum obtained by distillation at less than 90% ABV with a non-ethanol volatile substance content minimum of 225 grams per hectoliter of pure alcohol (HPA). The definition of rum by the CARICOM standard (Caribbean Community) is very closely related to the EU standard [59]. Even for the definition from the US, rum is a spirit distilled from fermented sugarcane material such as juice, syrup, molasses or other by-products or the cane sugar itself [60].
The history of rum started with the huge trade based on cane cultivation and the sugar industry in the 16th century [60]. The ancestor of rum, made from molasses, bears different names over time and location: ‘Rumbullion’, ‘Rumbustion’ and then ‘kill-devil’ in the English-speaking regions [60,61]. In the Spanish-speaking regions, the spirit was called ‘aguardiente’ or ‘guaro’. The term ‘Guildive’ or ‘Tafia’ was used in the French-speaking areas [62]. The predominant raw materials used to produce this spirit were the molasses and syrups imported from the neighboring Caribbean islands and the USA [63]. Halfway through the 19th century, a new method of rum production was developed. Some farmers isolated from the production units choose to use their own cane to produce rum from the juice instead of the molasses. This is the ‘z’habitant’ rum (creole word), the ancestor of Rhum Agricole [62]. The term ‘Agricole’, which means ‘agricultural’, can only be used as a geographical indication for traditional rum produced from the sugarcane juice of French overseas departments (Guadeloupe, Martinique, French Guyana and Reunion) or the autonomous region of Madeira [10]. Agricultural rum is characterized by sugarcane fresh notes, not found in rum made from molasses or syrup [57]. Martinican distillers protect their spirit with the AOC label. The decree, which was enacted on 5 November 1996, describes the conditions to be respected, presented in Table 5, so that Rhum Agricole can obtain the AOC Martinique label. This spirit is produced with pure and fresh sugarcane juice without the addition of syrup and molasses. The aged rum is obtained exclusively in oak barrels and the addition of caramel is prohibited.
This alcoholic beverage is economically important for French overseas departments such as Martinique and Guadeloupe, which produced, respectively, 100,000 and 80,000 hectoliters of pure alcohol (HPA) in 2021. French rum exportation is flourishing with an increase of 24 million euros between 2020 and 2022 [65]. This positive outcome has not been the case for rum producers in other parts of the world after the two years of pandemic. Other producers worldwide have registered a decrease in sales of 70% in 2020 [66].

3.4. Cachaça

The white cachaça (non-aged), called also ‘pinga’, is produced by about 40,000 Brazilian distilleries, but only 955 with a valid registration [8] This spirit is produced either by an industrial method (75%) with modern scientific techniques using stainless columns or by an artisanal method (25%) with ancestral techniques with a copper alembic [6].
According to Brazilian legislation, the term cachaça is a sugarcane spirit containing 38–48% ethanol (v/v) at 20 °C obtained by distillation, with a column or pot still, of fresh fermented sugarcane juice and produced only in Brazil [6]. The aged cachaça is matured in wooden barrels (maximum capacity of 700 L) for at least one year [67]. Oak as well as tropical local woods such as amburana, bálsamo, jatobá or jequitibá are used for maturation [68,69,70]. This very popular traditional drink in Brazil has recently begun to be exported internationally and Brazilian producers are seeking to increase the export rate, rising from 1% of exportation 20 years ago [71] to 50% today [72]. Over the past 20 years, Brazilian scientific researchers have been acquiring data to provide producers with valuable new tools to optimize their production methods [7]. Brazilian legislation defines the rules to produce cachaça (Table 6) and MAPA (‘Ministério da Agricultura Pecuária e Abastecimento’) suggests the use of analytical techniques suitable for sugarcane spirits for quality control.
Twenty years ago the annual production of cachaça was estimated at 1.3 billion liters per year [73], and nowadays it reaches about 2 billion liters, representing an economic activity estimated to be USD 6 billion [74].

3.5. Other SCJS

Unlike other spirits, such as whiskey or cognac, sugarcane spirit factories often lack technical knowledge; production can be empirical and rudimentary. Nevertheless, the situation is changing with the development of spiritourism as an economic actor for developing countries. For the following spirits, which are less commonly described in the literature, some producers are seeking to maintain their traditional production methods while striving to ensure superior product quality.
Aguardiente de cana refers to those based on sugarcane, whether juice or by-products, in Spanish-speaking regions. Aguardiente, meaning firewater, is produced in artisanal distilleries in Central and South America with only sugarcane juice [75,76]. These pure cane juice rums are less found on the international market.
Clairin is a traditional rum made according to traditional methods in Haiti by hundreds of rural distilleries. The spontaneous fermentation is realized by indigenous yeasts in wooden vats and lasts several days with fresh cane juice or cane syrup [77]. Moreover, aromatic plants or fruits can be added. The distillation is performed with a pot still called Charentais copper alembic. The first distillate is between 25 and 35% ABV and can undergo a second distillation to reach a higher alcohol content [20].
Grogue called also, ‘grogu’ or ‘grog’ is a traditional rum from Cape Verde. On the archipelago, the island of Santo Antão produces a large quantity of sugarcane on field terraces. Cutting and cane juice extraction is often manual. The fermentation is achieved in days (8 to 15) to weeks. The distillate with a 38–54% ABV is usually obtained using a copper pot still and is mostly consumed locally unaged. However, after a maturation period of 2 to 3 months, the grogue can be aged in wooden barrels [78].
The lack of legal requirements regarding the spirit-making practices in microdistilleries can lead to the manufacture of poor-quality products [79] that can even be harmful to health (high acetone or methanol content). The study of Pereira [80], based on Brazilian legislation, shows the difference in chemical profile between grogue and cachaça from artisanal and industrial production. Thus, copper levels above the Brazilian limit of 5 mg/L were reported in several artisanal production units. Moreover, the chemometric analysis (PCA analysis) performed on both spirits in 2012 shows a greater dispersion of the chemical profiles of grogue, probably due to the lack of regulation at that time [81]. Since then, Cape Verde began to regulate the production of this traditional spirit. The legislation of 12 August 2015, and the national program Vagrog II, is aimed at obtaining standardization and promoting grogue on the international market [82].

3.6. Traditional Customs and Uses

Because of their high alcohol content, distillates were first used medicinally [1]. Medicinal plants are used in maceration with spirit as an excellent hydroalcoholic solution for the extraction of natural compounds [83]. For example, the Bay-rum, an anti-rheumatism lotion, used for local application in the Caribbean, is obtained by maceration of West Indian bay-tree (Pimenta racemose) leaves soaked in rum for several weeks under the sun [84]. In Haiti, clairin is used to make ‘Trempé’, a maceration of wood extracts and spices in 22% alcohol strength clairin [85] and in Voodoo rituals [86]. Another more widely known example is the grog made of rum and hot water with lemon and cinnamon as a flu treatment.
Over the past decade, there has been a renewed interest in traditional medicines. To guarantee the safe use of these homemade medicines, studies have been carried out on the therapeutic and toxic properties of plants and their extracts in hydroalcoholic solutions such as rum [87]. Maceration of aromatic plants in spirit can provide the beverage with several positive characteristics when it is consumed in small quantities thanks to plant phenolic compounds as a source of antioxidants [88] and extracted volatile and biologically active compounds.
SCJSs as a regional cultural heritage are widely used in gastronomy, to flavor cakes and pastries, and are consumed neat, on ice and in traditional cocktails. The ‘Caipirinha’, made with cachaça, sugar, lime juice and crushed ice [89], the ‘Ti’Punch’ (Rhum Agricole, sugar and lime zest) from Martinique, or the ‘Punch’ (Grogue, honey, lemon and cinnamon) from Cape Verde [63] are emblematic beverages representing the country’s image. In accordance with the concept of food pairing, marketers approach new consumers by introducing them to new flavors.

3.7. The Sugarcane Juice Spirit Process

The quality and sensory profile of SCJSs depends upon the different steps of the process. The type of harvest (manual or mechanical), the pre- and post-treatments of the cane during the milling, the alcoholic fermentation conditions, and the type of distillation apparatus will define the chemical profile of non-aged SCJSs. Finally, the maturation in wood barrels is the last step that transforms, improves and enhances the sugarcane distillate. For the different SCJSs cited here, the diversity of flavor can be explained by the specificities of the making process presented in Table 7.

3.7.1. From the Field to the Mill

Once mature, sugarcane stalks are harvested manually with a cutting instrument called a machete or mechanically with a cane harvester. Burning cane pre-harvest facilitates manual cutting by reducing worker injuries from sharp leaves, insects, and snakes, but this agricultural practice has an impact on the atmospheric environment [90] and on the juice’s chemical composition. The burning of sugarcane is seldom practiced today because it reduces the quality of the stalks and destroys the biological balance in the fields [90]. During the harvest, cane top and leaves are separated from the stalk because they decrease the sucrose yield [19]. The mechanization of sugarcane harvest has advantages, reducing the effort of manual harvest and increasing the yield of homogeneous stalks for easier grinding. However, some disadvantages must be considered: a higher deterioration in cut cane and lower quality caused by non-cane materials not deliberately harvested with the cane (leaves, invasive plants and mud) [91]. The whole cane stalks (manual cutting) or in sections (mechanical cutting) are then rapidly transported to the distillery to reduce the impact of post-harvest sugar loss.
The juice extraction efficiency has greatly increased since the very first two-roller mill was created in India in the 1500s [92]. During the 17th century, the Brazilians improved their mills to increase grinding efficiency and to reduce the loss of the slave workforce. In the East, the first mills with two vertical gears, developed by the Chinese, were introduced in Japan in 1610. The origin of the three vertical gear mills is believed to be in Peru and in Mexico back in the 1700s. Nowadays, the animal-driven models with vertical gears can still be found among the small-scale producers, whereas the mills with three horizontal gears are still mainly used in modern distilleries. Motor-driven extraction can be improved using a cane kicker, composed of sets of knives, which breaks up the stalk and promotes regular feed on the conveyor. Then, a shredder, composed of a feed drum and a rotor with hammers, achieves the cellular rupture [93].
The addition of water, called imbibition water, during grinding, improves the efficiency of sugar extraction reaching 92 to 96% for the modern mills [94]. This dilution allows the control of the sugar content and, therefore, defines the alcohol content in the cane wine. Limiting the quantity of sugar available in the fermentation medium helps avoid reaching an alcohol concentration that inhibits yeast growth [95]. Particular attention must be paid to its chemical and microbiological composition because this water is part of the composition of the wort and will therefore have an impact on yeast fermentation efficiency. It is recommended not to use water with a high mineral content that may contain a higher bacterial charge [96]. Composed imbibition refers to the return of the juice from the last mill to the first mill. This technique has the advantage of improving the total extraction of juice [97]. The hygiene of the crushing equipment also has an impact because the microorganisms found on the surface of the rolls will be part of the wort microflora.
Motorized extraction requires a significant input of energy. According to Inskip, the energy requirement is estimated between 28.6 and 43.7 MJ for a milling rate of 400 t/h and a 15% fiber cane [93].
To eliminate the fine bagasse particles found in the juice, post-grinding filtration or decantation techniques such as a strainer or rotary vacuum-drum filter can be used. Then, to perform the fermentation, the sugarcane juice can be supplemented with several components and is, thus, called wort or must. Sulfuric acid or citric acid can be added to adjust the pH and thus regulate the aerobic bacterial flora [20]. With the same aim, disinfectants or antibiotics such as hydrofluoric acid, sodium fluoride and sodium penicillin G can be added [20]. In the case of grogue, the addition of sugar of less than 6 g/L is authorized in the wort [78].

3.7.2. The Key Step of Fermentation

This is the step where secondary products, responsible for sensorial characteristics of SCJSs, are formed in addition to the alcohol. Yeasts, generally belonging to the genus Saccharomyces [98], perform the biochemical reaction of transforming carbohydrates (sugars) into ethanol and carbon dioxide to produce energy released as heat. Sucrose, the main sugar in sugarcane, cannot be catabolized directly by the yeast, so it is first converted into glucose and fructose (by consuming ATP). Once activated in phosphate form, glucose and fructose enter the glycolysis pathway cycle (Embden–Meyerhof) to give pyruvic acid [20]. Under anaerobic conditions, this pyruvic acid is reduced to lactic acid or ethanol after decarboxylation. In addition to the ethanol produced, by-products are released into the fermented juice, including acids, which leads to the acidification of the medium [98]. During a 24-h fermentation, most of the sugars are consumed in the first 12 h [99]. However, the fermentation can last days or weeks, especially with wild yeasts [20]. Vats used for fermentation are usually open and the site is ventilated to facilitate the removal of carbon dioxide. The most common vat material used in modern units is stainless steel, but wood or mild steel are also used [100].
In the case of spontaneous fermentation, no selected yeast strains are added. The heterogenous microbiota (yeasts and bacteria) from the cane, as well as those from the equipment, represent the ecosystem liable for flavor formation [46]. Thus, contamination by tools such as the mill, should also be taken into consideration. The surface bacterial microflora of the stalks, which can be found in the wort, is composed of aerobic lactic acid bacteria (Corynebacteria, Micrococcus, Enterobacteria, Bacillus, and Pseudomonas) and microaerophilic bacteria (Lactobacillus, Leuconostoc) [20]. These microorganisms have access to the sugarcane during stalk cutting and after insect attack. They proliferate if the cut cane is not crushed quickly [96]. The quality and quantity of bacteria in the wort flora depend on the sanitary status of the cane and the wort components. In the case of spontaneous fermentation, the conditions must be favorable to the dominance of yeasts in the wort. Among the bacteria encountered in cane wort, three families are often responsible for a decrease in spirit quality. Lactic acid bacteria (LAB) compete with yeast for sugar and nutrients [45,99]. An excessive proliferation of these bacteria will negatively influence the quality of the spirit because of the important production of 2,3-butandione, acids (lactic, acetic and formic) and acetoin. Bacillus, which forms a pellicle on the vat surface during long fermentation, adversely affects the yield by consuming ethanol and releasing higher alcohol and fatty acids [11]. At the end of the fermentation, when the carbon dioxide is at a low level in the wort, acetic acid bacteria can develop [45].
Nowadays, to ensure better reproducibility and control of fermentation, yeasts are added in different possible forms: mixed or pure culture, in dehydrated form, or as liquid ferment. In general, producers select yeasts with good fermentation yields, low hydrogen sulfide (H2S) production, killer activity (toxin production that eliminates sensitive strains), and the ability to flocculate and produce a high level of aromatic compounds [45]. The main yeast species used is Saccharomyces cerevisiae. Back in the 1970s, in the French West Indies, dehydrated baker’s yeast was used to allow better control of fermentation at a reduced cost. Another way to inoculate the wort consists of selecting and isolating endogenous strains that are adapted to production conditions [95,101], providing an inoculum also related to the terroir. Whatever the inoculation method, in the case of continuous fermentation, the sedimented inoculum can be recycled for other fermentative processes [102]. On the other hand, when fermentation is discontinuous (batch fermentation), the wort must be inoculated at each new vat filling. This method is less economical and substrate inhibition can occur with a high sugar concentration. Although the yeast species mainly used is S. cerevisiae, it is also possible to use non-Saccharomyces yeast. This could provide a different sensory profile. Amorim et al. evaluated that the addition of Meyerozyma caribbica, known as Pichia [103], influences positively the chemical and sensory quality of cachaça [58].
Besides the microflora impact, fermentation efficiency also depends on many physicochemical parameters such as temperature, pH, oxygen and nutrient availability to maintain the growth of yeasts and their viable population, until the consumption of all sugars contained in the wort (Table 8). The nitrogen content available for yeast metabolism in cane juice depends on the cane cultivar and agronomic conditions [104]. When free amino acids are not sufficient for optimal yeast growth, nutrients such as ammonium sulfate can be added to the wort [102]. However, excessive nitrogen content and, conversely, very low nitrogen concentration can induce a high production of higher alcohols [104]. For cachaça, a fermented mixture of crushed corn, lemon and rice is added as a source of supplementary substrates [101], and yeast extract as a source of nitrogen [102].
High temperature due to exothermic reactions during fermentation promotes yeast autolysis, the growth of thermophilic bacteria such as lactic and acetic bacteria [20], and decreases the survival rate for yeast in an environment containing a high ethanol concentration when the temperature is higher than 35 °C [99]. To control the fermentation temperature, moderate-volume vats must be installed in a cool and airy space. The use of a cooling system is also possible. A low temperature can induce sluggish fermentation because of a decrease in membrane fluidity and temperature-dependent enzyme activity as permease [105]. Temperature control is even more important when fermentation is spontaneous to avoid the production of undesirable compounds such as acrolein [8].
Oxygen must be provided at the beginning of the fermentation cycle or before the pitching of the wort (preoxygenation) [106] to promote the biosynthesis of unsaturated fatty acids and sterols responsible for the fluidity and permeability of the cell membrane. Yeasts are then better protected against oxidative stress. However, an excess of oxygen can cause the formation of reactive oxygen species (ROS) that will cause oxidative damage to the cell [106].
The juice extraction and fermentation stages are considered to be those responsible for the highest losses in alcohol yield (between 25 and 50%) [20,96]. A better understanding and control of factors influencing fermentation would improve alcohol yield and spirit quality.

3.7.3. The Art of Distillation

Distillation is a physical process that can separate a mixture of liquids based on the differences in boiling point and volatility of compounds [107]. The process involves heating the liquid mixture, the cane wine, and then condensing the vapor. These two successive steps, evaporation and condensation, can be performed as a continuous or discontinuous process using a column still or a pot still (alembics) [8]. Throughout the industrialization era (18–19th centuries), many major improvements were performed in the Caribbean and South America, increasing the production and quality of sugarcane spirit [1,95]. From the 20th century onwards, both continuous and discontinuous distillations were perfected.
Depending on the still and its characteristics (copper or stainless steel material, size, design and distillation parameters), the spirit will have a different sensory profile [95].
The discontinuous distillation in a batch or pot still starts by charging the pot with wine and heating it. Alcohol vapors, richer in volatile compounds, reach the top of the pot (Figure 6). The reflux stream returns to the pot and the distillate is collected in a storage tank. Three fractions or cuts are collected in separative receivers depending on the boiling point of substances [6]: the head, the heart, and the tail [108]. The heart fraction corresponds to 80–85% of the total distillate volume [95]. Only the heart fraction, which contains the more suitable volatile compounds, is kept. The head fraction (1 to 5%), containing mostly methanol, and the tail fraction (15%), composed of higher alcohols such as propanol, butanol, and isobutanol (fusel oil) and aldehydes, are discarded [95]. The double distillation allows lighter aroma spirit with a higher alcohol content to be obtained [108]. It is also used to reduce the concentration of ethyl carbamate above the limit of 210 μg/L (Table 6). In this case, the first distillate is not split and the three fractions are separated during the second distillation [6].
The column still contains several trays, and the wine is introduced continuously to the column. The column is divided into two sections: the rectifying section, above the entry of the wine inlet, and the stripping section, below (Figure 6). Steam injected on the bottom tray reaches the top of the column through the bubble-cap trays. The vapor at the top of the column, richer in alcohol and other volatile components, is condensed in the condenser [8]. A part of the condensed vapor, called reflux (rectification), returns to the column [1], and a degassing system can be added at the top of the column. This method of distillation, favored by medium and large-size distilleries, has the advantages of being energy-saving and improving the standardization of distillate [95].
During the distillation process, in addition to the separation of compounds according to their boiling point and their concentration, other reactions contribute to the development of unique organoleptic properties. Esterification reactions between alcohol and acids, Maillard reactions, and oxidation reactions of sulfur compounds participate in the formation of aromas [20]. Moreover, copper contributes to superior product quality by forming insoluble copper salt and neutralizing the unpleasant flavor of sulfur compounds [7,8].
The distillate is stored in a tank, whereas the vinasse is discarded or treated to limit its impact on the environment [109]. In the case of efficient distillation, the output vinasse should contain a low ethanol content of less than 0.6% [110]. The fresh distillate has a harsh flavor that can be reduced after a few weeks of resting in an inert container (stainless steel) [6]. Stirring could help eliminate some small off-flavor volatile compounds [20].

3.7.4. Aging

The physicochemical profile (color and flavor) is modified during aging due to the biochemical reactions between the spirit and the wood [67]. The wood widely used to age SCJS is oak (Quercus spp.). However, for cachaça, in particular, many local woods are used, thus bringing aromatic profile diversity [53,111]. The different compounds of wood, cellulose, hemicellulose and lignin interact with the alcohol and generate new compounds [112]. Depending on the duration and intensity of toasting, the spirit’s aromatic profiles will be different [53,112]. The ‘char’ treatment corresponds to an intense but short toasting of the wood [113], whereas the ‘toast’ treatment is less drastic with several levels of intensity.
The first elements that appear after filling the barrel are color and tannins [114]. During the aging process, different biochemical reactions occur depending on three main parameters [67,115]: the barrel features, the duration and conditions of storage [112] and the initial composition of distillate [116].
Aged spirit profile varies according to the wood species (permeability, porosity and density), size and geometry, thermal treatment and number of uses of the barrel [117]. Depending on the grain of the wood, more or less tannins will be released [118]. The coarser and wider the grain, the more tannins will be found in the alcohol. The porosity of the wood allows the interaction of oxygen with spirit compounds. The resulting oxidation removes the harshness of the alcohol, increases the fruitiness (esterification) and enhances the complexity. The cellulose is not normally degraded by heat treatment during barrel making [52]. Thermal degradation of hemicellulose and lignin generates new compounds as presented in Table 9. The evaporation rate of the liquid will be influenced by the size of the barrel, the temperature, the humidity level, the barrel pre-treatment and the air circulation in the cellar.
The climate (temperature, humidity and atmospheric pressure), sometimes difficult to control, plays an important role in maturation, especially in the evolution of the alcohol content in the barrel. The alcohol concentration decreases during the aging process because of its passage through the pores of the barrel and the reactions of ethanol with secondary compounds. Douglas Santiago et al. indicate that during an aging process in high humidity conditions, the quantity of alcohol decreases, whereas when the humidity is low, it is rather the water that evaporates and the alcohol content of the spirit is higher at the end of storage in the barrel [119]. This study states that 3 to 4% of alcohol is lost annually in the barrels in Brazil. In tropical areas, this evaporation rate, called the angel’s share, can reach up to 10% in cellars under uncontrolled temperatures where the temperature ranges from 25 to 35 °C with a rather high relative humidity between 65 to 90% [117,119]. Moreover, in these favorable conditions to alcohol evaporation, the biochemical reactions are faster. Thus, the extraction of phenolic compounds is accelerated and spirits mature 3 to 4 times faster than in temperate climatic zones [67]. Therefore, it is important to study the kinetics of maturation-related congener formation under real production conditions [10]. According to the daily and seasonal temperature changes, the liquid contained in the barrel will undergo expansions and contractions, which will participate in the elaboration of the flavors [119]. The study of underwater aging of bottled agricultural rum by Aguiar et al. reports the positive impact on sensory characteristics without the participation of compounds from wood [10].
During the aging process, three types of mechanisms occur: subtractive, additive and interactive (Table 9). Subtractive mechanisms correspond to the capacity of the carbonized wood to purify the spirit. The carbon removes immature characteristics from the distillate, such as off-flavor sulfur derivatives (e.g., H2S, dimethyl disulfide) [120]. Additive mechanisms allow the modifications of flavor, texture and color of the spirit by extraction of volatile compounds, phenols, sugars, glycerol, non-volatile organic acids and tannins from the wood. Interactive mechanisms refer to polymerization, esterification, hydrolysis and oxidation reactions. The wood improves the organoleptic characteristics of the spirit by reducing its aggressive and bitter character.
The cellar master can blend different batches to obtain the desired sensory profile using, for example, the solera method, which corresponds to blending different ages. In the cellar, the casks are classified into levels based on the age of the spirit. The first level (Criadera) receives the fresh distillate and the last levels contain the oldest spirit (Solera) [56,121].
Table 9. Biochemical phenomena occurring during the aging. A: additive; I: interactive and S: substractive mechanisms.
Table 9. Biochemical phenomena occurring during the aging. A: additive; I: interactive and S: substractive mechanisms.
CompoundEvolutionBiochemical Reaction InvolvedRef.AIS
Sulfuric compoundsInteract with copper—High volatility[122] X
CopperAbsorbed or adsorbed by the wood. Reaction with phenolic compounds[67] XX
EthanolEvaporation, oxidation, reaction with secondary compounds[67] XX
Aldehyde (Ethanal)Oxidation of amino acids, alcohols, and fatty acids[67] X
Esters (Ethyl acetate)Esterification of fatty acids with ethanol[67] X
Acetic acidDue to ethanol oxidation—Degradation of hemicellulose[67]XX
Volatile acidityAcetic acidity increasing and transfer of non-volatile organic acids from the barrel to the beverage[67]XX
MethanolDegradation of hemicellulose (pectin)[67]XX
Ethyl carbamateConversion of cyanide to cyanate by the action of peroxides[67] X
TanninExtraction from the wood[123]X
Gallic acidHydrolysis of wood tannins → increase viscosity[67]X
Furfural and 5-HMFThermal breakdown of pentoses and hexoses from hemicellulose[67]X
Cinammic aldehydesConiferaldehyde and sinapaldehyde generated from lignin[10]X
Vanillin, syringaldehydeDerived from lignin; oxidation of sinapaldehyde and coniferaldehyde[10] X
Vanillic acid, syringic acidDerived from lignin; oxidation of vanillin and syringaldehyde[123] X
Phenolic compoundsSimple extraction (low molecular weight) and extraction and hydroalcoholic degradation of lignin[123]XX
Sugars (glucose, arabinose, sucrose, maltol, xylose, etc.)Hydrolysis of tannins, extracted from the wood, hemicellulose degradation[20,123]XX
Glycerol and fatty acidsExtraction and/or hydrolyzed of resins and triglycerides from cell walls of wood[123]XX
CoumarinFormed by the cyclization of o-coumaric acid[123] X

3.7.5. Quality Control during the SCJS Making Process

Most spirit producers obtained their expertise through many years of successful and failed experiments. Hence, the determinants of a spirit’s flavor can be random without standardization. For this reason, quality control must be carried out at every stage of spirit production to guarantee long-lasting quality and yield. However, depending on the technical, human and financial resources, small to large-size producers do not have the same capacities to ensure this control. Table 10 lists the analysis and associated techniques for monitoring the production from the field to the bottle. The method’s global cost includes the price of equipment, reagents and laboratory consumable items needed. The technical skill necessary is also estimated.
Sugar and nitrogen contents are good indicators for evaluating the cane’s ripening. During plant growth, the sugar concentration increases progressively from the bottom to the top while the amino acid concentration decreases. Evaluation of the sugar content of the juice can be performed quickly with a Brix refractometer. Brix value can also be used to monitor the fermentation process and determine the mill efficiency by analyzing the juices from the successive rolls. Bagasse humidity is an indicator of potential fuel value and mill extraction efficiency. The juice pH, slightly acidic due to organic acids, is measured as an indicator of freshness when the cane arrives in the distillery [5]. A significant decrease in pH can be a sign of the beginning of fermentation and, therefore, a warning of bacterial contamination.
The physicochemical conditions of fermentation can be followed by performing different analyses on both wort and wine. These elements will indicate if the fermentation is over (density, Brix, total reducing sugar and nitrogen) and detect bacterial contamination (acidity and pH) [99,102]. Total acidity (volatile and fixed) is a crucial parameter in control quality [8]. Thus, according to Brazilian legislation, the volatile acidity of a wine, expressed as acetic acid concentration, must be below 150 mg/100 mL aa (anhydrous alcohol). An evaluation of the health of the yeast population (adaptability, growth, etc.) can be defined by counting the number of cells, the rate of viability and their division rate (budding index). The assessment of fermentation efficiency and yield are determined by calculating sugars and alcohol contents in wort and wine using high-performance liquid chromatography with a refractive index detector (HPLC-RI) [48,124].
Distillation performance is determined according to three criteria: the chemical profile of the spirit, and the alcohol content in both the spirit and vinasse. It can be useful to define the composition of the organic and inorganic compounds of the vinasse, especially when it is intended to be applied in the cane field [109].
Several analytical techniques can be applied to distillate characterization. The determination of volatile organic compounds (higher alcohols, esters and aldehydes) is mostly performed by gas chromatography (GC) with a flame ionization detector (FID) or mass detector (MS), with or without derivatization [8,10,67,125,126,127], whereas the maturation-related congeners are analyzed by HPLC with ultraviolet (UV), diode array (DAD) or MS detectors [67,128,129]. These analyses estimate the aromatic richness and the presence of contaminants such as methanol and ethyl carbamate. A high ratio of total esters to ethyl acetate may be indicative of superior spirit quality [108]. Spectrophotometric analysis is useful to determine the colorimetric characteristics of the product and to follow the extraction of phenolic compounds (total phenol index). Not all analytical methods are equal in terms of precision and reliability. For example, the Folin–Ciocalteu method for phenolic compounds is cost-effective and requires little advanced equipment, but it has low specificity because reactions can occur with other reducing substances [123].
Inorganic compounds are determined by ion chromatography (IC), inductively coupled plasma (ICP) or atomic absorption spectroscopy (AAS) [51,80,130].
Table 10. Main methods of analysis used in control quality in a SCJS distillery.
Table 10. Main methods of analysis used in control quality in a SCJS distillery.
Step of ProcessMatrix to Be AnalyzedAnalysisUnitAnalytical Method/ApparatusCostTechnical SkillsReferences
Harvest/
Grinding
Cane juiceBrix%Refractometer—Brix spindle++[131,132]
Sucrose%Polarimeter+++++[131,132]
Free amino acidμL/mLSpectrophotometer-UV+++++[133]
Purity%Calculation: P% = (Sucrose/Brix) × 100++++++[132]
GrindingCrushed caneFiber%Weight and calculation+++[132]
BagasseMoisture%Moisture analyzer+++[132]
FermentationWort
Wine
Density-Mustimeter/Densitometer++-
pH-pH-meter++[134]
Temperature°CThermometer++-
Total aciditymg/LTitration method+++[134]
Volatile aciditymg of acetic acid/100 mL aa *Titration method++++[130]
Budding index%Microscope++[99]
Enumeration of yeast and bacteriaUFC/mLMicrobiological analysis+++[99,100,101,102,103]
Cell countCell/mLMicroscope/Neubauer chamber/Malassez cell+++[102]
Cell viability%Eritrosin/methylene blue+++[102]
Alcohol content% (v/v)Ebulliometer/Distiller++++[61]
Total nitrogenmg/LKjeldahl method++++[102]
Sugar: glucose, fructose, sucrosemg/LHPLC-RI++++++[99]
Total reducing sugar (TRS)g/LDNS method++++[135]
DistillationVinasseAlcohol loss%GC-FID++++[110]
DCO, DBO5, N, P, Kg/L or mg/LMostly spectrophotometry methods++++++[136]
Distillation
Maturation Aging
Distillate (SCJS)Inorganic components (anion, cation)mg/LIC/ICP OES/AAS+++++[80,137,138]
Ethanol content% (v/v)Densitometer/Hydrostatic balance/Pycnometer/spectrophotometric method++ to ++++ to +++[73,102,131,139]
Volatile aciditymEq/L acetic acidTitration method++++[102]
VOCs contentμg to mg/LGC with FID/MS+++[67,127]
Ethyl carbamateμg/LGC MS/HPLC MS++++++[74,140]
Matured-related congenersμg/L to mg/LHPLC—fluorescence/UV/DAD/MS detector+++[52,67,127,130]
Antioxidant activity% inhibitionDPPH+++[141,142]
Antibacterial activity of phenolic extractμg/mlBacteria culture. Minimal inhibitory concentration++++[141]
Dry extractg/LDrying under stirring+++[116]
Chromatic characteristics-CIELab method/Spectrophotometer+++++[114,116,139,143]
Total phenolic indexOD/gSpectrophotometer at 280 nm++[114,116]
Total phenolic compoundsmg/LFolin–Ciocalteu method ++++[123]
Dissolved oxygenmg/LFluorometer+++++[114]
Sensory EvaluationSCJSOdor activity-Odor thresholds—GC-olfactometry+++++[71,144,145]
Descriptive tests-Panel of experts++++[74,122,146,147]
Hedonic test-Consumers++[102,103,122]
Difference test: triangle test-Consumers or experts++[71,148]
* aa: anhydrous alcohol; Estimated cost, +: EUR < 1000; ++: EUR 1000 to 5000; +++: EUR > 10,000, Technical skills required; +: very easy; ++: basic technical knowledge; +++: important technical knowledge.
It is essential to define the sensory profile of the spirit. Relating the chemical composition to the aroma profile is complex due to masking, additions and synergy phenomena. The sensory test is the easiest and fastest way to identify flaws [95]. Defining the visual characteristics is important because the golden color of aged spirits has an attractive impact on consumers. The flavor description is essential to guide them to choose a product according to their preferences.
Depending on the type of sensory test, it is necessary to define a common vocabulary to describe aromas [146] with an expert panel, to facilitate the descriptive analysis [149]. Aroma references can be prepared in a hydroalcoholic solution with an alcohol strength equivalent to the samples to be tasted (between 30 and 55% ABV) [150]. The orthonasal odor thresholds can be determined using reference standards (Table 11) [144,145,151]. The fresh-distilled SCJS is characterized by herbal, fruity (citrus, tropical fruit) and floral aromas. In aged SCJSs, the main aromas are spicy (clove, cinnamon and vanilla), empyreumatic (toasted, smoky, chocolate and coffee) and woody. Aromas perceived as flaws can give clues on the causes (bacteria contamination, incorrect distillation parameters) of poor sensory quality.

4. Chemical Composition and Their Contribution to the Flavor of SCJSs

This hydroalcoholic matrix includes a large variety of organic and inorganic compounds that are important in quality evaluation or in fraud control. The study of Franitza et al. allows the classification of 100% of SCJSs and 76% of spirits made from molasses using chemometric analysis [154].
With the aim of preserving the specific features of their manufacturing process while improving product quality, SCJS producers can validate their empirical knowledge using modern analysis techniques. Scientific data can help ensure reproducibility, quality and product authenticity. This chapter presents the chemical compounds identified in cachaça, agricultural rum and grogue. They are divided into two groups according to the production stage at which they are formed. The fermentation congeners group includes organic fatty acids, alcohols, esters, aldehydes and ketones; the maturation-related congeners group contains phenolic and furanic compounds, lactones and terpenoids. And some compounds belong to both groups.

4.1. Fermentation Congeners

4.1.1. The Aliphatic Acids

Organic acids such as myristic, palmitic, acetic or caproic acids are already present in cane juice (Table 3) but most, especially the short-chain fatty acids (<C5), are produced by the yeast metabolism during the alcoholic fermentation [155]. The highest acid, acetic acid (Table 12), comes from the transformation of sugar into acetic acid instead of ethanol in the presence of oxygen during the fermentation [156]. It is also formed following the oxidation of ethanal or from the hydrolysis of acetyl coA and is produced by lactic and acetic bacteria. The second highest, lactic acid is synthesized by Lactobacillus while Bacillus are responsible for the formation of propionic and butyric acids.
The production of acids depends on three factors. First, the predominant yeast strain in the wort. For example, yeasts belonging to the genus Schizosaccharomyces produce fewer fatty acids than baker’s yeast (S. cerevisiae) [20]. The second and third factors are fermentation duration and distillation temperature [8].
The short-chain saturated aliphatic acids, such as formic, acetic, propionic and butyric acids, are associated with unpleasant aromas (pungent, vinegar, milk/butter and sour), whereas the long-chain fatty acids are related to oily, waxy and soapy aromas. They contribute to giving a negative flavor when they are in high concentration in SCJSs. Moreover, they can increase the solubilization of metallic copper during the distillation and thus induce a higher concentration of this contaminant in spirits [8,74]. However, these organic acids have a positive impact on the spirit when they are in small amounts because they are precursors to the formation of esters during distillation and maturation [8,20]. Acetic acid concentration tends to increase during the aging process due he ethanol oxidation [7] or hemicellulose degradation [67]. The slightly acidic pH of SCJSs, around four, can be explained by the presence of these organic acids.

4.1.2. The Alcohols

Alcohols are the major volatile compounds in unaged spirits (>80%) [54]. With an alcohol strength between 40% and 50% ABV, as in cachaça or agricultural rum, ethanol contributes to the warming and burning perception of the spirit [157]. Among the 26 other alcohols identified in Table 11, isoamyl alcohol, propanol and isobutanol predominate in non-aged and aged SCJS. The higher alcohols, containing more than two carbon atoms, are derived either from amino acids via the Ehrlich catabolic pathway or from pyruvate, which is favored in cases of amino acid deficiency [8,104]. The formation of higher alcohols varies according to the fermentation conditions (strain and amount of yeast, temperature, pH, wort nitrogen content, and duration) [7,8,10,158]. Even if 46% of these alcohols contribute to the fruity aroma and 31% to the sweet flavor, they are also associated with unpleasant aromas such as fusel oil (26%) and solvent (19%). The long-chain fatty alcohols containing up to eight carbon atoms, such as capric and lauric alcohol, are related to waxy/fatty aroma. Depending on their concentration, higher alcohols could be considered as flavoring compounds or contaminants [10]. Hexanol, found also in the raw material, contributes to a herbal aroma, characteristic of freshly distilled SCJS, whereas the presence of 1- and 2-butanol indicates bacterial activity during the alcoholic fermentation [7]. Their concentration is respectively limited in cachaça to 3 and 10 mg/100 mL aa. Glycerol, produced by yeast promotes osmoregulation and redox balance, providing a sweet taste and viscosity [147]. Its content increases in aged SCJSs from the thermal breakdown of wood triglycerides or trans-esterification of triglycerides and ethanol [120]. The concentration of higher alcohols tends to increase during the aging in wooden barrels [67], but it can also decrease because of their oxidation to form aldehydes or esterification with acids [143].
The lowest molecular weight alcohol, methanol, is toxic [8]. For this reason, its concentration is limited to 20 mg/100 mL aa in cachaça and 30 mg/100 mL aa in spirit according to the EU regulation 2019/787 [2]. This compound is produced during the fermentation process by enzymatic demethylation of the methoxy groups of the pectins of the bagasse performed by pectin methylesterase [108]. Filtration, following juice extraction, decreases the fine particles of bagasse in the wort. With a very high perception threshold (1200 mg/L), methanol impacts the flavor negatively with alcoholic and solvent notes [108].

4.1.3. The Aldehydes and Ketones

They originate from the raw material or from the oxidation of fatty acids, alcohols and amino acids [10] and are intermediates to form higher alcohols. The most abundant aldehyde, representing up to 90% of total aldehydes, is ethanal or acetaldehyde (Table 12) [8,108]. This compound, which causes a hangover, derives from the decarboxylation of pyruvate or air-oxidation of ethanol [110,144]. To limit its formation, aeration should be avoided at the end of fermentation [130] and distillation should proceed as soon as possible [108]. This intermediate compound of alcoholic fermentation has a lower boiling point than ethanol and is accumulated in the head fraction during distillation [7]. Hence, ethanal concentration must be below the limitation of 30 mg/100 mL aa for cachaça and 50 mg/100 mL aa for other spirits, according to the EU regulation 2019/787 [2,8].
Aldehydes with less than three carbon atoms, such as acrolein, provide ethereal and pungent aspects to the spirit. This toxic unsaturated aldehyde, produced by bacteria from glycerol and acetaldehyde during fermentation accidents, causes irritation to the skin and mucous membranes and is highly tear-producing. Moreover, it is an intermediate for the acetal, 1,1,3-triethoxy-propane, which is associated with mushroom and solvent aromas. As is the case for ethanal, acrolein is limited to 5 mg/100 mL aa in cachaça. Despite these negative aspects being linked to high concentrations, aldehydes are also known to provide vegetal and fruity aromas (Table 12) [130,159]. Nonanal, already present in the sugarcane juice, and hexanal contribute to green and grass odors [10]. During the aging process, few changes have been observed in the aldehyde concentration, except for aromatic aldehydes, due to the oxidation into acids [67,119], or the formation of acetal from the reaction of ethanal and ethanol in acid medium.
Regarding ketones, acetone and diacetyl are the highest in terms of concentration. The latter is also produced from pyruvate, as well as ethanal, but under oxidative decarboxylation, by LAB [160]. It is a sub-product in the biosynthesis of amino acids (valine and leucine) with an unpleasant flavor (pungent, buttery aroma).
Table 12. Fermentation congeners in SCJS (AR: agricultural rum; C: cachaça; G: Grogue). Concentration range in mg/L or mg/100 mL aa and their sensory descriptor.
Table 12. Fermentation congeners in SCJS (AR: agricultural rum; C: cachaça; G: Grogue). Concentration range in mg/L or mg/100 mL aa and their sensory descriptor.
ORGANIC ACIDS
NameN° CASConcentration (mg/L)Type of SCJSPungentVinegarMilk/ButterRancid/SourSweatyOily/FattyWaxySoapyReferences
Non-AgedAged
Short-chain aliphatic acids (≤C5)
Formic acid64-18-61.90–4.505.9–29.90C [54,161]
Acetic acid64-19-714.60–1211125–848C-AR [51,71,122,125,144,145]
Glycolic acid79-14-1<LOD to 0.16<LOD to 6.71C [122,162]
Propionic acid79-09-40.0001-0.00396 a-C [54,103,163,164]
Pyruvic acid127-17-3<LOD to 0.40<LOD to 3.91C [122,162]
Lactic acid50-21-50.32–8834.89–64.50C [122]
Isobutyric acid79-31-20.43-C [163,164,165]
Butyric acid107-92-60.00096–2.300.563C-AR [54,103,144,163,164,165]
Succinic acid110-15-6<LOD to 0.05<LOD to 0.61C [122,162]
(S)-2-Methylbutanoic acid1730-91-2-0.865AR [144]
Isovaleric acid503-74-2--AR [144]
Valeric acid109-52-4-0.89 aAR [54,166]
2-Methylvaleric acid,97-61-0--C [166]
3-Methylvaleric acid105-43-1--C [54]
Citramalic acid597-44-4<LOD to 0.10<LOD to 0.80C [122]
Medium-chain saturated fatty acids (C6 to C10)
Caproic acid142-62-10.0016–0.0047 a-C [54,103,163,164,166]
Enanthic acid111-14-80.00011–0.00093 a-C [54,103]
Caprylic acid124-07-20.001–0.0040–0.29C [54,103,144,163,164,167]
Pelargonic acid112-05-00.00055–0.00175 a-C [103,166]
Capric acid334-48-50–0.910–11.40C-AR [54,103,122,144,163,164,167]
Long-chain saturated fatty acids (>C10)
Lauric acid143-07-70.00032–0.480–2.94C [54,54,103,122,167]
Myristic acid544-63-80.01–0.37<LOD to 5.16C [54,122]
Pentadecanoic acid1002-84-2--C [54]
Palmitic acid57-10-30.003–0.730.10–1.44C [54,103,122,164]
Octadecadienoic acid59404-49-8--C [54,103,122]
Octadecenoic acid2825-79-8--C [54]
ALDEHYDES & KETONES
Methanol67-56-11.07–660–100C [122,125,152,163,167]
Propanol71-23-821.99 b–3293.76–409C-G-AR [10,73,74,89,122,163,164,166]
Propylene glycol57-55-60.00008–0.010 a-C [103,162]
Glycerol56-81-5<LOD to 5.30<LOD to 66C [127,162]
1,3-Butanediol107-88-00.00004–0.00062-C [103]
Saturated fatty alcohols (>C4)
1-Butanol71-36-3<LOD to 10.300–9C-AR [10,54,122,145,147,164,167]
Isobutanol78-83-12.88–2560–452C-G-AR [51,54,81,122,144,166,167,168]
2-Butanol78-92-2<LOD to 78.20<LOD to 269C [122,125,162]
Amyl alcohol71-41-00.05–0.20 c-C [163,168]
Isoamyl alcohol123-51-3152–937450–1290C-G-AR [51,54,81,122,144,147,163,167,168]
2-Methyl-1-butanol137-32-648.02-C [48,144,162,163,166]
(S)-2-Methyl-1-butanol1565-80-6-245 aAR [144,162]
1-Hexanol111-27-30.00008–9.300.644–18.1C-AR [10,103,122,154,163]
3-Methyl-1-pentanol589-35-50.00042 a–0.00118 a-C [103,162]
Enanthic alcohol111-70-60.00031–0.00064-C [103,163]
2-Heptanol543-49-7-3.38–4.54AR [10,166]
Caprylic alcohol111-87-50.00086–0.001638.05–8.46C-AR [10,103,163]
3-Octanol589-98-0-0.32–0.35AR [10,162]
4-Octanol589-62-80.00042 a–0.00373 a-C [103]
Pelargonic alcohol143-08-8-3.40–50AR [10,162]
Capric alcohol112-30-1-0.018–12.12AR [10,154,162]
Lauryl alcohol112-53-80.00045–0.044.15–8.67C-AR [10,103,162,168]
Myristyl alcohol112-72-10.164.45–7.50AR [10,162,168]
Cetyl alcohol36653-82-46.13 c0.31–1.05C-AR [10,162,168]
Unsaturated fatty alcohols
trans-3-Hexen-1-ol928-97-20.00008–0.0007-C [103,162]
3-Decen-1-ol10339-60-3-2.92–3.64AR [10]
Saturated aldehydes
Methanal50-00-00–20.402.04–16.70C [51,122,130,169]
Acetaldehyde (ethanal)75-07-05.7–2001.55–439C-AR [10,51,74,80,89,122,164,165,167,169]
Propionaldehyde123-38-60–0.55 b<LOD to 0.92C [51,122,125,130,169]
Butyraldehyde123-72-8<LOD to 22<LOD to 1.34C [122,169]
Acetoin513-86-00.914-C [163,165]
Valeraldehyde110-62-3<LOD to 0.31 c-C [51,169]
(R)-2-Methylbutyraldehyde33204-48-7-0.0189 aAR [144]
(S)-2-Methylbutyraldehyde1730-97-8-0.0274 aAR [144]
Isovaleraldehyde590-86-3<LOD to 0.090.233 aC-AR [51,125,144,169]
Hexanaldehyde (hexanal)66-25-1<LOD to 0.77<LOD to 1.07C-AR [10,122,144]
Heptaldehyde (heptanal)111-71-7--C [71,162]
Caprylaldehyde (octanal)124-13-0-0.77–0.93AR [10,145,165]
Pelargonaldehyde (nonanal)124-19-6-2.44–2.72AR [10]
Capraldehyde (decanal)112-31-2-0.68–0.87AR [10]
Unsaturated aldehydes
Acrolein/acrylaldehyde107-02-80–5.80 b<LOD to 0.36C [122,130,169]
Crotonaldehyde123-73-9<LOQ to 5.32 c-C [51]
2,4-Nonadienal5910-87-2--C [71]
trans-2-Nonenal18829-56-6--C [71]
trans-trans-2,4-Decadienal25152-84-5-0.00115 aAR [144]
Ketones
Acetone67-64-10–23<LOD to 6.90C [122,125,130,167]
Diacetyl431-03-80–9.770.0437C-AR [71,125,130,144,145,167,170]
Cyclopentanone120-92-30–2.87-C [51,167]
1-Octen-3-one4312-99-6--C [71,162]
2-Nonanone821-55-60–0.00028 a0.87–2.55C-AR [10,103]
2-Undecanone112-12-9-0.94–1.72AR [10]
2-Pentadecanone2345-28-0-0.97–2.14AR [10]
Acetals
Acetal105-57-7-25 aAR [71,144,164,171]
1,1,3-Triethoxypropane7789-92-6--C [172,173]
ESTERS
Ethyl formate109-94-4--C [54]
Methyl propionate554-12-10.08-C [162,168]
Ethyl acetate141-78-61.56 c–623 c15.5–1160C-AR [10,51,54,74,80,122,163,167,168,169]
Ethyl acrylate140-88-5--C [71]
Ethyl lactate97-64-3<LOD to 75.902.21–120C-AR [10,122,125,130]
Ethyl isobutyrate97-62-1--C [71,171]
Ethyl butyrate105-54-4<LOD to 19.70 c<LOD to 3.20C-AR [54,71,122,144,163,171,174]
Isoamyl acetate123-92-2-1.06 a–13.83AR [10,54,71,144,163,166,171]
Ethyl valerate539-82-2-0.0515–2.02AR [10,144]
Ethyl isovalerate108-64-5-0.211 a–2.56AR [10,144,154,166]
Ethyl 2-methylbutyrate7452-79-1--C-AR [71,144,171]
Ethyl(S)-2-methylbutanoate10307-61-6-0.23AR [144]
Propyl butyrate105-66-80.08-C [162,168]
Ethyl caproate123-66-0<LOD to 9.84 b0.55–66.42C-AR [10,54,103,122,130,144,163,164,171]
Diethyl succinate123-25-10.00308–0.0069-C [103,164]
Amyl propionate624-54-40.02 c-C [168]
2-Hexyl acetate5953-49-1--C [54,162]
Ethyl enanthate106-30-90.05 c3.29–4.07C-AR [10,162,168]
Ethyl cyclohexanoate3289-28-9-0.00103 aAR [144,162]
Ethyl caprylate106-32-10–16.28586.18–723.19C-AR [10,54,103,130,163,164,171,174]
Isoamyl valerate2050-09-10.01 c-C [162,168]
Ethyl pelargonate123-29-50.1508–0.55215.47–21.45C-AR [10,166,174]
Methyl caprate110-42-9-0.00004–0.00033 aAR [154]
Ethyl caprate110-38-30.132 c–13.01 c-C [10,54,130,170,171,174]
Isoamyl octanoate2035-99-60.0666 c–0.234 c-C [162,174]
Ethyl laurate106-33-20.015 a–4.03 c-C [10,54,103,174]
Ethyl myristate124-06-10.00042 a–0.00706 a-C [103,162]
Ethyl palmitate628-97-7--C [54,162]
Ethyl 9-hexadecenoate54546-22-40.00088–0.00175 a-C [103]
Methyl linoleate112-63-0--C [54,162]
Ethyl margarate14010-23-2--C [54]
Ethyl linolate544-35-4--C [54,162]
Ethyl stearate111-61-5--C [54,162]
a Value in μg/L in the reference and converted in mg/L, b value in mg/100 mL aa and converted in mg/L with alcohol strength indicated in the article, c value in mg/100 mL aa, - value not found in the literature.

4.1.4. The Esters

Esters are key compounds in fermentation congeners because a correlation has been established between their content and the aromatic quality of spirits. They have a strong odorant power due to a low olfactory perception threshold [108]. According to the Brazilian legislation, cachaça must contain less than 200 mg/100 mL aa esters expressed in ethyl acetate, whereas this limit is fixed to 130 mg/100 mL aa in the EU [2,8]. These compounds are produced either through the condensation of acetyl CoA with ethanol (acetate esters) or through the condensation of a fatty acid with ethanol (acyl esters) [104]. Thus, ethyl esters of fatty acids and acetates of higher alcohols are produced during alcoholic fermentation by enzymatic reaction by yeasts [117] but also by lactic bacteria during malolactic fermentation [160]. The ester rate depends on three elements: the variety of cane, the fermentation and distillation conditions [20]. First, cane wax is a source of fatty acids that favors the content of ethyl esters. Secondly, the oxygen availability and the yeast strain impact esters formation whereas high temperature negatively affects their formation during the fermentation [20,54]. Third, the ethyl acetate concentration is different according to which distillation apparatus is used (pot or column still) [108]. Esters are known to bring a fruity and sweet flavor to spirits (Table 12). Ethyl acetate, the predominant ester in distillate (80% of total esters) [8,143], contributes to a pleasant flavor at low concentrations. On the contrary, an important amount in the spirit, due to an unwanted acetic bacteria activity [108], has a sensory negative impact (solvent, acidic). Esters most favorable to aromatic richness (‘bouquet’) are those with middle and long-chain fatty acids (≥C5). From 8 carbon atoms, esters contribute to fatty/oily aromas whereas the fatty acid esters with at least 10 carbon atoms give waxy or soap notes. During maturation in wooden barrels, their concentration increases thanks to esterification reactions between alcohols and fatty acids [143]. According to studies on cachaça, esters concentration depends on wood species. Oak is the one providing the highest content after 6 months [119] or 3 years [143].

4.2. Maturation-Related Congeners

Phenolic compounds bring an important contribution to the sensory profile of aged spirit (Table 13). This chemical family contains one or several hydroxylated benzene rings [128]. Some of them are already found in the sugarcane juice such as benzaldehyde, eugenol or 4-vinylguaiacol (Table 4). As a result of all physical and chemical reactions occurring during the aging process (Table 10), the spirit acquires color, new aromas and roundness making its organoleptic profile more harmonious. The aged spirit color ranges from light yellow to dark brown depending, for the most, on the duration of storage in the barrel, the number of uses [74] and the wood species [143] thanks to both phenolic and furanic compounds extraction [67,118,123]. Alike the color, texture evolves during the aging as gallic acid, sugars and glycerol concentration increase [10,67]. Besides, mouthfeel and taste, crucial for consumer acceptance, can be pleasant with the sweetness. On the other hand, they can be unpleasant because of astringency due to high content of tannins and its derivative (gallic acid) [117]. However, oxidation and polymerization of tannins during the wooden aging lead to a reduction in the astringency strength [175]. The tannin family is divided into two groups: hydrolyzable tannins (ellagitanins, gallotanins) or condensed tannins (catechin, gallocatechin) [175,176]. The second group is at the origin of flavonoids, which are odorless. With a molecular structure rich in alcohol function (OH) and double carbon bonds (C=C), these phenolic compounds have a more or less antioxidant capacity. (+)-Catechin seems to possess a better efficient antioxidant activity than quercetin and kaempferol [142].
While fresh distillate features freshness, fruity, floral, and green aromas, the wooden aged distillate is more widely defined by woody, spicy and empyreumatic notes (Table 13). Among the simple phenolic compounds (one aromatic ring) some are noteworthy for their contribution to sensory profile. Eugenol, as well as its precursor, guaiacol, brings spicy (clove) and smoky aromas with low olfactory thresholds [128]. The other lignin-derivative, guaiacol- (vanillin, vanillic acid, coniferaldehyde) and syringol- (sinapaldehyde, syringic acid and syringaldehyde) compounds are widely studied because they are aging markers [129]. The study of Castro et al. indicates that American oak barrel (Q. alba) releases in cachaça a higher amount of benzoic aldehydes (syringaldehyde and vanillin) and benzoic acids (syringic acid and vanillic acid) compared to French oak barrels (Q. petraea) [129]. Conversely, Madrera et al. found in brandy a higher content of these compounds in French oak (Q. sessilis) [177]. It is relevant to note that the aging environment conditions were certainly different (Europe/Brazil).
The γ-lactones, so-called oak or whiskey lactones, derive from the oxidation of lipids by LAB during lactic fermentation and from glycosidic compounds during the heat treatment [117,175]. These volatile compounds are key odorants in aged spirits bringing coconut aroma (Table 13). Among the two isomeric cis- and trans-forms of oak lactones, the first contributes the most because of its low perception threshold [117].
Coumarin and scopoletin are considered markers of aging because these compounds are found in the spirit after a long period of storage in a barrel [178]. However, the concentration of coumarin ought to be monitored because of its toxicity [123] considering that American oak (Q. alba) provides a higher scopoletin amount than French oak (Q. sessillis) in the case of brandy [115,177].
The thermal degradation of hemicellulose from the wood barrel induces the formation of furanic aldehydes, such as 5-HMF and furfural. Slightly found in unaged spirits except when the juice comes from burned cane [119], the sum of furfural and 5-HMF is limited to 5 mg/100 mL aa in cachaça. These C4 heterocyclic-based compounds are formed by the dehydration of pentoses caused by high temperatures under acidic conditions and/or by the Maillard reaction during the barrel-making process [67]. These compounds bring toasted, woody and toasted almond aromas to the spirit and a peculiar color as mentioned below (Table 13) [67].
Terpenes and terpenoids or isoprenoids, are a large group of organic compounds found in plants. Terpenoids include acyclic (e.g., citronellol, linalool), cyclic (e.g., menthol) compounds, composed of isoprene unit (C5), and C13-carotenoid-derived compounds, called norisoprenoids (e.g., ionone and damascenone) (Table 14). Their concentration in unaged spirits depends on the cane variety [10]. Some free or glycosylated odorless hydrocarbons, already present in the raw material, may be transformed during the distillation process [179]. During the maturation process, their formation can be related to the thermal or oxidative degradation of carotenoids [10] and their presence in aged distillate varies depending on the wood species [119]. The most abundant terpene, d-limonene, already present in the raw material is associated with floral aroma. The C13-norisoprenoids provide fruity notes; and sesquiterpenes (C15) contribute to a woody aroma. Terpenoids provide a pleasant aroma to spirits (floral, citrus and vegetal) [54,145]. β-ionone is a marker used to differentiate SCJS from those made from molasses [154].
Table 13. Aromatic compounds in SCJS (AR: agricultural rum; C: cachaça). Concentration range in mg/L or mg/100 mL aa and their sensory descriptor.
Table 13. Aromatic compounds in SCJS (AR: agricultural rum; C: cachaça). Concentration range in mg/L or mg/100 mL aa and their sensory descriptor.
N° CASConcentration (mg/L)Type of SCJS FruitySweetWoodyPhenolBalsamicFloralSpicyEmpyreu 1ChemicalReferences
Non-AgedAged
Phenolic compounds and derivatives
Phenol108-95-20.0005 a–0.0041 a0.0002 a–0.014 aC [142,162]
o-Cresol95-48-70.001 a–0.005 a0–0.097 aC [142,162]
p-Cresol106-44-5-0.00341 aAR [144,162]
Guaiacol90-05-10.0002–0.0030.0005–0.047C-AR [54,71,103,142,144,171]
Protocatechuic acid99-50-3-2.65–2.83AR [10,162]
Gallic acid149-91-7-0-12.97C-AR [10,123,167]
2-Ethylphenol90-00-60–0.0027 a0.0009 a–0.124 aC [142,162]
3,5-Xylenol108-68-90–0.0058 a0–0.065 aC [142,162,163]
4-Ethylphenol123-07-90–0.021 a0.0023 a–0.117 aC-AR [142,144,162]
Vanillin121-33-5-0–3.29C-AR [10,54,71,144,144,162,167,171]
Syringol91-10-1--C [54,162]
Vanillic acid121-34-6-0–2.24C-AR [10,54,162,167]
4-Vinylguaiacol7786-61-0--C [54,71,162]
4-Ethylguaiacol2785-89-90.005 a–0.052 a0.0006 a–0.298 aC-AR [142,144,162]
p-Coumaric acid7400-08-0-0–7.09C-AR [10,167]
Caffeic acid331-39-5-0–1.39AR [10]
Syringaldehyde134-96-3-0–6.11C-AR [10,162,167]
Syringic acid530-57-4-0–4.57C-AR [10,167]
Anethole104-46-1-12.62–17.61AR [10,162]
Eugenol97-53-00.0012–0.0880.002–0.1183C-AR [71,142,144,162,171]
Dihydroeugenol2785-87-7-0.0193AR [144,144,162,171]
Coniferaldehyde458-36-6-0–0.25C [162,167]
Scopoletin92-61-50–0.0300–1.31C [142]
Methoxyeugenol6627-88-9--C [54,162]
Sinapaldehyde4206-58-0-0–1.20C [167]
Ellagic acid476-66-4-0–4.51C-AR [10,167]
Flavonoids
Kaempferol520-18-300–1.95AR [10,142]
Epicatechin490-46-00–0.220–1.53C [142,167]
Catechin154-23-40–0.150–0.42C [142]
Quercetin117-39-500–2.30C-AR [10,142]
Myricetin529-44-2-0–3.48AR [10]
Rutin153-18-4-4.34–7.99AR [10]
Aromatic alcohols and acids
Benzyl alcohol100-51-6--C [54,162]
Benzoic acid65-85-0--C [54,162]
Phenethyl alcohol60-12-80.02 c6.66–17.10C-AR [10,54,71,144,163,168,171]
Cinnamic alcohol104-54-16.95 c-C [168]
Aromatic esters
Phenyl acetate122-79-20–0.00014-C [103,162]
Ethyl benzoate93-89-01.84-C [162,168]
Phenethyl acetate103-45-70.00016 a–0.00029 a-C [71,103,162,163,164,171]
Ethyl hydrocinnamate2021-28-5-0.00354 aAR [144,162]
Lactones
β-Angelica lactone591-11-7--C [54]
Sotolon28664-35-9-0.0162AR [144,162]
Coumarin91-64-5--C [54,162]
Hydrocoumarin119-84-6--C [54,162]
cis-Whiskey lactone80041-00-5-1.12AR [144,171]
trans-Whiskey lactone80041-01-6-0.124AR [144,171]
γ-Nonalactone104-61-0--AR [144,171]
γ-Dodecalactone2305-05-7-0.0037–0.0070AR [154,162]
Aromatic aldehydes and ketones
Benzoic aldehyde100-52-70–8.571.77–8.61C [10,51,122,130,147,162,169]
Acetophenone98-86-20–2.90-C [51,167]
Furanic compounds
3-Furaldehyde498-60-2-0.00108–0.0128 aAR [154]
Furfural (2-Furaldehyde)98-01-10–4.57 c1.12 b–24.59 bC-AR [10,51,67,130,163,169,170,180,181]
Furfuryl alcohol98-00-00-1.30-C [103,162]
5-HMF67-47-0<LOQ to 7.23 c0.8–19.40C-AR [10,51,67,125,162,169,170]
Furfuryl acetate623-17-60.00029 a–0.00105 a-C [103,162]
Aromatic hydrocarbons
o-Xylene95-47-6-0.71–1.68AR [10]
p-Cymene99-87-6-14.7–18.25AR [10]
Cumene98-82-8-2.12–3.93AR [10]
TDN 230364-38-6-3.09–4.32AR [10]
1 Empyreumatic aromas; 2 1,1,6-Trimethyl-1,2-dihydronaphthalene, a value in μg/L in the reference and converted in mg/L, b value in mg/100 mL aa and converted in mg/L with alcohol strength indicated in the article, c value in mg/100 mL aa, - value not found in the literature.

4.3. The Other Organic Compounds

Acetate, the anion of acetic acid, was quantified by IC in cachaça and grogue by Pereira et al. (Table 15). It is a precursor of esters and acids [80].
The sulfur compounds in spirit are known to degrade the sensory profile (sulfurous flavor) [162]. The major sulfur compound, dimethylsulfide, is probably generated from the degradation of sulfur-containing amino acids or from the reduction of sulfate salts in the wort during fermentation [20,182]. These compounds form insoluble copper salts during the distillation, which are found in the vinasse [117].
A noteworthy ester is ethyl carbamate (EC), also known as urethane, with no sensory contribution but with potentially carcinogenic properties [183]. Countries such as Canada and Brazil restrict the maximum acceptable concentration of EC to, respectively, 150 μg/L and 210 μg/L in spirits, whereas, in Europe, a recommendation from the European Commission encourages producers to limit this concentration in stone fruit spirits (e.g., kirsch, plum spirits, apricot spirits, etc.) to 1000 μg/L [8,159].
Many factors can promote EC formation throughout the SCJS-making process. A better control of the fermentation (temperature, nitrogen supply) and distillation (reflux system, wine feed rate and temperature) parameters could avoid the formation of EC precursors, namely, hydrocyanic acid [7,184]. For cachaça, the double distillation helps to reduce the concentration of EC. However, in the case of Rhum Agricole AOC from Martinique, this practice is prohibited [185].

4.4. Inorganic Compounds

The level of inorganic compounds has not been investigated for agricultural rums (Table 16). Nevertheless, these chemical species are important for the quality of the spirit and its flavor [80,117]. Found in sugarcane juice, depending on the variety, they are essential for yeast fermentation. During the distillation step, their concentration varies in the distillate according to the inner materials used in a column or pot still.
Copper, extracted into the spirit from the still, has a positive action on the aroma because it reduces the unpleasant taste due to sulfur compounds [184]. However, it is considered an environmental pollutant when it is found in high concentration in the vinasse. Moreover, copper catalyzes the formation of EC from cyanide, [67]. So, producers try to reduce its concentration in fresh distillate.

5. Conclusions

The family of SCJSs is a diverse category presenting styles that are the reflection of the terroir, tradition and expertise acquired over generations. Economic sectors related to this spirit are currently developing, boosted by spiritourism. Producers strive to improve the quality thanks to the sensory and chemical characterization of their spirits. The current scientific data available on most of these SCJS are incomplete. For example, the specificities of aging under tropical conditions are less commonly studied than those under temperate climates. Nevertheless, additional data are required to enhance the process and help develop new outstanding products. Hence, investments in improving quality will lead to more innovation and optimization and finally open the way to premiumization.
Some producers have put forward unique production methods, such as organic farming, mono-variety production, eco-friendly practices (use of recycled bottles, packaging), finishing in barrels that have contained other alcoholic beverages (wine, Porto, whiskey, etc.), and aging with woods other than oak.
Altogether, the future of SCJSs seems promising, especially if producers succeed in preserving their distinctive features while facing new challenges. In addition to improving yield and quality, these industries are concerned with reducing their carbon footprint and impact on fauna and flora (discharge of vinasse, cultivation practices).

Author Contributions

Conceptualization, J.B., J.S.-R. and O.M.; methodology, J.B., J.S.-R. and O.M.; validation, J.B., J.S.-R. and O.M.; writing—original draft preparation, C.C.; writing—review and editing, J.B., J.S.-R. and O.M.; supervision, J.B., J.S.-R. and O.M.; project administration, J.B., J.S.-R. and O.M.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gorak, A.; Sorensen, E. Distillation: Fundamentals and Principles; Elsevier: Oxford, UK, 2014; ISBN 978-0-12-386547-2. [Google Scholar]
  2. European Parliament and of the Council of 17 April 2019 Regulation (EU) 2019/787. Available online: https://www.legislation.gov.uk/eur/2019/787 (accessed on 19 October 2022).
  3. Statistica.com. Topic: Spirits Market Worldwide. Available online: https://www.statista.com/topics/6788/spirits-market-worldwide/ (accessed on 18 October 2022).
  4. Décret n° 2009–1350 du 29 Octobre 2009 Relatif Aux Appellations d’Origine Contrôlées «Martinique», «Marc d’Alsace» Suivie de la Dénomination «Gewurztraminer», «Calvados», «Calvados Domfrontais», «Calvados Pays d’Auge», «Cornouaille», «Domfront», «Pays d’Auge», «Pommeau du Maine» et «Pommeau de Normandie». Available online: https://www.legifrance.gouv.fr/loda/article_lc/LEGIARTI000021220378/2009-11-02 (accessed on 1 August 2023).
  5. François, R. Appellations of origin and local customs: The need for innovation “Le Rhum de la Martinique”: An example of AOC recognition of a product in an industrialised context. In Proceedings of the European Association of Agricultural Economists 67th Seminar, LeMans, France, 28–30 October 1999. [Google Scholar]
  6. Medeiros, A.; Matos, M.E.; Monteiro, A.; De Carvalho, J.; Soccol, C. Cachaça and Rum. In Current Developments in Biotechnology and Bioengineering: Food and Beverages Industry; Elsevier: Amsterdam, The Netherlands, 2017; pp. 451–468. ISBN 978-0-444-63666-9. [Google Scholar]
  7. Bortoletto, A.M.; Silvello, G.C.; Alcarde, A.R. Good Manufacturing Practices, Hazard Analysis and Critical Control Point plan proposal for distilleries of cachaça. Sci. Agric. 2018, 75, 432–443. [Google Scholar] [CrossRef]
  8. Lima, C.M.G.; Benoso, P.; Pierezan, M.D.; Santana, R.F.; Hassemer, G.d.S.; da Rocha, R.A.; Dalla Nora, F.M.; Verruck, S.; Caetano, D.; Simal-Gandara, J. A state-of-the-art review of the chemical composition of sugarcane spirits and current advances in quality control. J. Food Compos. Anal. 2022, 106, 104338. [Google Scholar] [CrossRef]
  9. Mbolidi-Baron, H. Les Conditions de Durabilité de la Production de la Canne à Sucre à la Martinique: Une Approche Territoriale. Ph.D. Thesis, University of Toulouse, Toulouse, France, 2002. [Google Scholar]
  10. Aguiar, D.; Pereira, A.C.; Marques, J.C. Agricultural Rum of Madeira matured on the seafloor: Improved physicochemical changes induced by a pioneering seafloor ageing process. Eur. Food Res. Technol. 2021, 247, 3023–3035. [Google Scholar] [CrossRef]
  11. James, G. (Ed.) Sugarcane, 2nd ed.; World Agriculture Series; Blackwell Science: Oxford, UK, 2004; ISBN 978-0-632-05476-3. [Google Scholar]
  12. Moore, P.H.; Paterson, A.H.; Tew, T. Sugarcane: The Crop, the Plant, and Domestication. In Sugarcane: Physiology, Biochemistry, and Functional Biology; Moore, P.H., Botha, F.C., Eds.; John Wiley & Sons Ltd: Chichester, UK, 2013; pp. 1–17. ISBN 978-1-118-77128-0. [Google Scholar]
  13. Artschwager, E.; Brandes, E.W. Sugarcane (Saccharum officinarum L.): Origin, Classification, Characteristics, and Descriptions of Representative Clones; United States Department of Agriculture: Washington, DC, USA, 1958. [Google Scholar]
  14. Pompidor, N.; Charron, C.; Hervouet, C.; Bocs, S.; Droc, G.; Rivallan, R.; Manez, A.; Mitros, T.; Swaminathan, K.; Glaszmann, J.-C.; et al. Three founding ancestral genomes involved in the origin of sugarcane. Ann. Bot. 2021, 127, 827–840. [Google Scholar] [CrossRef]
  15. Evans, D.L.; Joshi, S.V. Complete chloroplast genomes of Saccharum spontaneum, Saccharum officinarum and Miscanthus floridulus (Panicoideae: Andropogoneae) reveal the plastid view on sugarcane origins. Syst. Biodivers. 2016, 14, 548–571. [Google Scholar] [CrossRef]
  16. Cheavegatti-Gianotto, A.; de Abreu, H.M.C.; Arruda, P.; Bespalhok Filho, J.C.; Burnquist, W.L.; Creste, S.; di Ciero, L.; Ferro, J.A.; de Oliveira Figueira, A.V.; de Sousa Filgueiras, T.; et al. Sugarcane (Saccharum X officinarum): A Reference Study for the Regulation of Genetically Modified Cultivars in Brazil. Trop. Plant Biol. 2011, 4, 62–89. [Google Scholar] [CrossRef]
  17. Premachandran, M.N.; Prathima, P.T.; Lekshmi, M. Sugarcane and Polyploidy—A Review. J. Sugarcane Res. 2011, 1, 1–15. [Google Scholar]
  18. Wang, K.; Zhang, H.; Khurshid, H.; Esh, A.; Wu, C.; Wang, Q.; Piperidis, N. Past and recent advances in sugarcane cytogenetics. Crop J. 2023, 11, 1–8. [Google Scholar] [CrossRef]
  19. Chen, J.C.; Chou, C.C. Cane Sugar Handbook: A Manual for Cane Sugar Manufacturers and Their Chemists; John Wiley & Sons: New York, NY, USA, 1993; ISBN 0-471-53037-9. [Google Scholar]
  20. Fahrasmane, L.; Ganou-Parfait, B. De la Canne au Rhum; Editions Quae; Institut National de la Recherche Agronomique: Paris, France, 1997; ISBN 2-7380-0728-7. [Google Scholar]
  21. Smith, F.H. Caribbean Rum: A Social and Economic History; University Press of Florida: Gainesville, FL, USA, 2005; ISBN 978-0-8130-2867-5. [Google Scholar]
  22. FAOSTAT. Available online: https://www.fao.org/faostat/en/#data/QCL (accessed on 14 February 2023).
  23. Henry, R.J.; Kole, C. (Eds.) Genetics, Genomics and Breeding of Sugarcane; CRC Press: Boca Raton, FL, USA, 2010; ISBN 978-1-57808-684-9. [Google Scholar]
  24. Hébert, A.; Goebel, F.-R.; Domaingue, R.; Lebourgeois, V.; Fovet-Rabot, C.; Bégué, A.; Chabalier, P.; Chopart, J.-L.; Dagallier, J.-C.; Della Mussia, S.; et al. (Eds.) La Canne Aux Multiples Ressources! CIRAD: Montpellier, France, 2008; Available online: https://agritrop.cirad.fr/542892/ (accessed on 26 July 2023).
  25. De Morais, L.K.; de Aguiar, M.S.; de Albuquerque e Silva, P.; Câmara, T.M.M.; Cursi, D.E.; Fernandes, A.R., Jr.; Chapola, R.G.; Carneiro, M.S.; Bespalhok Filho, J.C. Breeding of Sugarcane. In Industrial Crops: Breeding for BioEnergy and Bioproducts; Cruz, V.M.V., Dierig, D.A., Eds.; Handbook of Plant Breeding Series; Springer: New York, NY, USA, 2015; pp. 29–42. ISBN 978-1-4939-1447-0. [Google Scholar]
  26. Tabriz, S.S.; Kader, M.A.; Rokonuzzaman, M.; Hossen, M.S.; Awal, M.A. Prospects and challenges of conservation agriculture in Bangladesh for sustainable sugarcane cultivation. Environ. Dev. Sustain. 2021, 23, 15667–15694. [Google Scholar] [CrossRef]
  27. Hong, J.; Boussetta, N.; Enderlin, G.; Merlier, F.; Grimi, N. Degradation of Residual Herbicide Atrazine in Agri-Food and Washing Water. Foods 2022, 11, 2416. [Google Scholar] [CrossRef]
  28. Sass, J.B.; Colangelo, A. European Union bans atrazine, while the United States negotiates continued use. Int. J. Occup. Environ. Health 2006, 12, 260–267. [Google Scholar] [CrossRef] [PubMed]
  29. Kogevinas, M. Probable carcinogenicity of glyphosate. BMJ 2019, 365, l1613. [Google Scholar] [CrossRef] [PubMed]
  30. Meena, M.R.; Kumar, R.; Chinnaswamy, A.; Karuppaiyan, R.; Kulshreshtha, N.; Ram, B. Current breeding and genomic approaches to enhance the cane and sugar productivity under abiotic stress conditions. 3 Biotech 2020, 10, 440. [Google Scholar] [CrossRef] [PubMed]
  31. Galinaro, C.A.; Cardoso, D.R.; Franco, D.W. Profiles of Polycyclic Aromatic Hydrocarbons in Brazilian Sugar Cane Spirits: Discrimination between Cachaças Produced from Nonburned and Burned Sugar Cane Crops. J. Agric. Food Chem. 2007, 55, 3141–3147. [Google Scholar] [CrossRef]
  32. Thai, C.; Doherty, W. The composition of sugarcane juices derived from burnt cane and whole green cane crop. In Proceedings of the 33rd Annual Conference of the Australian Society of Sugar Cane Technologists 2011, Mackay, QLD, Australia, 4–6 May 2011; Bruce, R.C., Ed.; Australian Society of Sugar Cane Technologists: Mackay, QLD, Australia; Curran Associates, Inc.: Red Hook, NY, USA, 2011; pp. 368–376, ISBN 978-1-61782-921-5. [Google Scholar]
  33. Shaner, D.L.; Beckie, H.J. The future for weed control and technology. Pest Manag. Sci. 2014, 70, 1329–1339. [Google Scholar] [CrossRef]
  34. Hussain, S.; Khaliq, A.; Mehmood, U.; Qadir, T.; Saqib, M.; Iqbal, M.A.; Hussain, S.; Hussain, S.; Khaliq, A.; Mehmood, U.; et al. Sugarcane Production under Changing Climate: Effects of Environmental Vulnerabilities on Sugarcane Diseases, Insects and Weeds. In Climate Change and Agriculture; IntechOpen: London, UK, 2018; ISBN 978-1-78985-668-2. [Google Scholar]
  35. Eggleston, G. Positive Aspects of Cane Sugar and Sugar Cane Derived Products in Food and Nutrition. J. Agric. Food Chem. 2018, 66, 4007–4012. [Google Scholar] [CrossRef]
  36. Kaavya, R.; Pandiselvam, R.; Kothakota, A.; Banuu Priya, E.P.; Arun Prasath, V. Sugarcane Juice Preservation: A Critical Review of the State of the Art and Way Forward. Sugar Tech. 2019, 21, 9–19. [Google Scholar] [CrossRef]
  37. Canilha, L.; Chandel, A.K.; Suzane dos Santos Milessi, T.; Antunes, F.A.F.; Luiz da Costa Freitas, W.; das Graças Almeida Felipe, M.; da Silva, S.S. Bioconversion of Sugarcane Biomass into Ethanol: An Overview about Composition, Pretreatment Methods, Detoxification of Hydrolysates, Enzymatic Saccharification, and Ethanol Fermentation. J. Biomed. Biotechnol. 2012, 2012, 989572. [Google Scholar] [CrossRef]
  38. Singh, A.; Lal, U.; Mukhtar, H.; Singh, P.; Shah, G.; Dhawan, R. Phytochemical profile of sugarcane and its potential health aspects. Pharmacogn. Rev. 2015, 9, 45–54. [Google Scholar] [CrossRef]
  39. Chen, E.; Song, H.; Li, Y.; Chen, H.; Wang, B.; Che, X.; Zhang, Y.; Zhao, S. Analysis of aroma components from sugarcane to non-centrifugal cane sugar using GC-O-MS. RSC Adv. 2020, 10, 32276–32289. [Google Scholar] [CrossRef]
  40. Hill, R. Aptitude or adaptation: What lies at the root of terroir? Geogr. J. 2020, 186, 346–350. [Google Scholar] [CrossRef]
  41. Brillante, L.; Bonfante, A.; Bramley, R.G.V.; Tardaguila, J.; Priori, S. Unbiased Scientific Approaches to the Study of Terroir Are Needed! Front. Earth Sci. 2020, 8, 539377. [Google Scholar] [CrossRef]
  42. Kyraleou, M.; Herb, D.; O’Reilly, G.; Conway, N.; Bryan, T.; Kilcawley, K.N. The Impact of Terroir on the Flavour of Single Malt Whisk(e)y New Make Spirit. Foods 2021, 10, 443. [Google Scholar] [CrossRef] [PubMed]
  43. Malfondet, N. La Typicité d’Une Eau de Vie de Cognac: Itinéraire Technologique et Lien au Terroir. Ph.D. Thesis, University of Burgundy, Dijon, France, 2015. [Google Scholar]
  44. Portugal, C.B.; de Silva, A.P.; Bortoletto, A.M.; Alcarde, A.R. How native yeasts may influence the chemical profile of the Brazilian spirit, cachaça? Food Res. Int. 2017, 91, 18–25. [Google Scholar] [CrossRef] [PubMed]
  45. Oliveira, V.A.; Vicente, M.A.; Fietto, L.G.; de Miranda Castro, I.; Coutrim, M.X.; Schüller, D.; Alves, H.; Casal, M.; de Oliveira Santos, J.; Araújo, L.D.; et al. Biochemical and Molecular Characterization of Saccharomyces cerevisiae Strains Obtained from Sugar-Cane Juice Fermentations and Their Impact in Cachaça Production. Appl. Environ. Microbiol. 2008, 74, 693–701. [Google Scholar] [CrossRef] [PubMed]
  46. Portugal, C.B.; Alcarde, A.R.; Bortoletto, A.M.; de Silva, A.P. The role of spontaneous fermentation for the production of cachaça: A study of case. Eur. Food Res. Technol. 2016, 242, 1587–1597. [Google Scholar] [CrossRef]
  47. Varela, C. The impact of non-Saccharomyces yeasts in the production of alcoholic beverages. Appl. Microbiol. Biotechnol. 2016, 100, 9861–9874. [Google Scholar] [CrossRef]
  48. Duarte, W.F.; Amorim, J.C.; Schwan, R.F. The effects of co-culturing non-Saccharomyces yeasts with S. cerevisiae on the sugar cane spirit (cachaça) fermentation process. Antonie Van Leeuwenhoek 2013, 103, 175–194. [Google Scholar] [CrossRef]
  49. Alcarde, A.R.; Souza, L.M.; Bortoletto, A.M. Ethyl carbamate kinetics in double distillation of sugar cane spirit. Part 2: Influence of type of pot still: Ethyl carbamate kinetics in double distillation of sugar cane spirit. J. Inst. Brew. 2012, 118, 352–355. [Google Scholar] [CrossRef]
  50. Bruno, S.; Vaitsman, D.; Kunigami, C.; Brasil, M. Influence of the distillation processes from Rio de Janeiro in the ethyl carbamate formation in Brazilian sugar cane spirits. Food Chem. 2007, 104, 1345–1352. [Google Scholar] [CrossRef]
  51. Reche, R.V.; Leite Neto, A.F.; Da Silva, A.A.; Galinaro, C.A.; De Osti, R.Z.; Franco, D.W. Influence of Type of Distillation Apparatus on Chemical Profiles of Brazilian Cachaças. J. Agric. Food Chem. 2007, 55, 6603–6608. [Google Scholar] [CrossRef]
  52. Bortoletto, A.M.; Alcarde, A.R. Aging marker profile in cachaça is influenced by toasted oak chips. J. Inst. Brew. 2015, 121, 70–77. [Google Scholar] [CrossRef]
  53. Bortoletto, A.M.; Correa, A.C.; Alcarde, A.R. Aging practices influence chemical and sensory quality of cachaça. Food Res. Int. 2016, 86, 46–53. [Google Scholar] [CrossRef]
  54. Bortoletto, A.M.; Silvello, G.C.; Alcarde, A.R. Aromatic profiling of flavor active compounds in sugarcane spirits aged in tropical wooden barrels. Braz. J. Food Technol. 2021, 24, e2019071. [Google Scholar] [CrossRef]
  55. Campos, J.O.S.; De Aquino, F.W.B.; Do Nascimento, R.F.; Da Costa, J.G.M.; Keukeleire, D.D.; De Casimiro, A.R.S. Influence and effect of thermal treatment in elaboration of regional wood extracts for cachaça. J. Food Compos. Anal. 2004, 17, 179–185. [Google Scholar] [CrossRef]
  56. Quesada-Granados, J.J.; Samaniego-Sánchez, C.; Blanca-Herrera, R.M. Rhum–Ron–Rum: Technology and Tradition. In Encyclopedia of Food and Health; Elsevier: Oxford, UK; Waltham, MA, USA, 2016; pp. 618–627. ISBN 978-0-12-384953-3. [Google Scholar]
  57. Stewart, G.G. A short history of rum. In Whisky and Other Spirits; Elsevier: London, UK, 2022; pp. 457–462. [Google Scholar]
  58. Council Regulation (EEC) No 1576/89 of 29 May 1989 Laying Down General Rules on the Definition, Description and Presentation of Spirit Drinks; European Union: Brussels, Belgium, 1989; Volume 160.
  59. CRS 25: 2008; CARICOM Regional Standard for Rum. CARICOM Regional Organisation for Standards and Quality (CROSQ): Belleville, Barbados, 2008.
  60. Mangwanda, T.; Johnson, J.B.; Mani, J.S.; Jackson, S.; Chandra, S.; McKeown, T.; White, S.; Naiker, M. Processes, Challenges and Optimisation of Rum Production from Molasses—A Contemporary Review. Fermentation 2021, 7, 21. [Google Scholar] [CrossRef]
  61. Jacques, K.A.; Lyons, T.P.; Kelsall, D.R. The Alcohol Textbook: A Reference for the Beverage, Fuel and Industrial Alcohol Industries, 4th ed.; Nottingham University Press: Nottingham, UK, 2003; ISBN 978-1-897676-13-4. [Google Scholar]
  62. Zamor, H. The history of Martiniquan rum. Int. J. Humanit. Cult. Stud. 2015, 2, 557–568. [Google Scholar]
  63. Anger, J.-P.; Kintz, P. Is rum toxic? Ann. Toxicol. Anal. 2008, 20, 137–144. [Google Scholar] [CrossRef]
  64. Ministère de l’Agriculture et de l’Alimentation. Arrêté du 29 Décembre 2020 Homologuant le Cahier des Charges de l’Appellation d’Origine Contrôlée «Rhum de la Martinique». J. Off. Repub. Fr. 2021, 316, 165–242. [Google Scholar]
  65. Agro-Média Le SIR 2023 Dévoile une Radiographie Inédite du Marché Mondial du Rhum. Available online: https://www.agro-media.fr/actualite/le-sir-2023-devoile-une-radiographie-inedite-du-marche-mondial-du-rhum-56145.html (accessed on 1 July 2023).
  66. Bates, J. Will Rum Finally Make Its Global Travel Retail Breakthrough?—Category Intel. Available online: https://drinks-intel.com/spirits/will-rum-finally-make-its-global-travel-retail-breakthrough-category-intel/ (accessed on 1 July 2023).
  67. Alcarde, A.R.; Souza, L.M.; Bortoletto, A.M. Formation of volatile and maturation-related congeners during the aging of sugarcane spirit in oak barrels. J. Inst. Brew. 2014, 120, 529–536. [Google Scholar] [CrossRef]
  68. Santiago, W.D.; das Graças Cardoso, M.; da Silva Lunguinho, A.; Barbosa, R.B.; Cravo, F.D.; da Silva Gonçalves, G.; Nelson, D.L. Determination of ethyl carbamate in cachaça stored in newly made oak, amburana, jatobá, balsa and peroba vats and in glass containers. J. Inst. Brew. 2017, 123, 572–578. [Google Scholar] [CrossRef]
  69. Cardoso, D.R.; Frederiksen, A.M.; Ataíde da Silva, A.; Franco, D.W.; Skibsted, L.H. Sugarcane spirit extracts of oak and Brazilian woods: Antioxidant capacity and activity. Eur. Food Res. Technol. 2008, 227, 1109–1116. [Google Scholar] [CrossRef]
  70. Da Silva, A.A.; do Nascimento, E.S.P.; Cardoso, D.R.; Franco, D.W. Coumarins and phenolic fingerprints of oak and Brazilian woods extracted by sugarcane spirit. J. Sep. Sci. 2009, 32, 3681–3691. [Google Scholar] [CrossRef] [PubMed]
  71. De Souza, M.D.C.A.; Vásquez, P.; del Mastro, N.L.; Acree, T.E.; Lavin, E.H. Characterization of Cachaça and Rum Aroma. J. Agric. Food Chem. 2006, 54, 485–488. [Google Scholar] [CrossRef]
  72. Mulling Neutzling, D.; Santos, M.S.D.; Barcellos, M.D.D.; Lauren Land, A. Value Creation from Internationalization of Sugar Cane by-products: A multi-stakeholder view of artisanal cachaça production. RBGN 2015, 17, 890–910. [Google Scholar] [CrossRef]
  73. Oliveira, E.S.; Cardello, H.M.A.B.; Jeronimo, E.M.; Souza, E.L.R.; Serra, G.E. The influence of different yeasts on the fermentation, composition and sensory quality of cachaça. World J. Microbiol. Biotechnol. 2005, 21, 707–715. [Google Scholar] [CrossRef]
  74. Bortoletto, A.M.; Alcarde, A.R. Assessment of chemical quality of Brazilian sugar cane spirits and cachaças. Food Control 2015, 54, 1–6. [Google Scholar] [CrossRef]
  75. Donaires Quispe, L. Fermentación del Mosto de Caña de Azúcar (Saccharum officinarum) Con Uso de Catalizadores Naturales y Comerciales Para Obtener Aguardiente—Pachachaca—Abancay—2018. Ph.D. Thesis, Universidad Tecnológica de los Andes, Abancay, Peru, 2018. [Google Scholar]
  76. Paranubes. Available online: https://www.paranubes.com/tech (accessed on 21 July 2023).
  77. Archive of Pure Cane Juice Rum—Rhum Attitude. Available online: https://www.rhumattitude.com/en/categorie/rhum/rhum-agricole/ (accessed on 21 July 2023).
  78. Ramos, A.C.N. Caracterização de Aguardente de Cana-de-Açúcar “Grogue de Cabo Verde”. Master’s Thesis, Instituto Politecnico de Leiria, Peniche, Portugal, 2015. [Google Scholar]
  79. Craveiro, I.; Alves, D.; Amado, M.; Santos, Z.; Fortes, A.; Delgado, A.; Correia, A.; Gonçalves, L. Determinants, Health Problems, and Food Insecurity in Urban Areas of the Largest City in Cape Verde. Int. J. Environ. Res. Public Health 2016, 13, 1155. [Google Scholar] [CrossRef]
  80. Pereira, R.F.R.; Melo, D.; de Barros, A.L.; Clecius, A.A.; do Nascimento, R.F. Comparative study of chemical constituents in sugar cane spirits from Brazil and Cape Verde. Afr. J. Food Sci. 2012, 6, 427–440. [Google Scholar] [CrossRef]
  81. Pereira, R.F.R.; Vidal, C.B.; de Lima, A.C.A.; Melo, D.Q.; Dantas, A.N.S.; Lopes, G.S.; do Nascimento, R.F.; Gomes, C.L.; da Silva, M.N. Chemometric Characterization of Alembic and Industrial Sugar Cane Spirits from Cape Verde and Ceará, Brazil. Int. J. Anal. Chem. 2012, 2012, 840528. [Google Scholar] [CrossRef]
  82. Moreira, É.A.S. Plano de Internacionalização do “Grogue” Para a Diáspora Cabo-Verdiana. Master’s Thesis, Instituto Politecnico do Porto, Peniche, Portugal, 2016. [Google Scholar]
  83. Rodríguez-Solana, R.; Vázquez-Araújo, L.; Salgado, J.M.; Domínguez, J.M.; Cortés-Diéguez, S. Optimization of the process of aromatic and medicinal plant maceration in grape marc distillates to obtain herbal liqueurs and spirits. J. Sci. Food Agric. 2016, 96, 4760–4771. [Google Scholar] [CrossRef] [PubMed]
  84. Bourgeois, P. Une plante aromatique de la Caraïbe: Le bois d’Inde. J. d’Agric. Tradit. Bota. Appl. 1995, 37, 139–149. [Google Scholar] [CrossRef]
  85. Weniger, B.; Haag-Berrurier, M.; Anton, R. Plants of Haiti used as antifertility agents. J. Ethnopharmacol. 1982, 6, 67–84. [Google Scholar] [CrossRef] [PubMed]
  86. Krippner, J. Historical Dictionary of Haiti. By Michael R. Hall. Lanham. Americas 2013, 69, 402. [Google Scholar] [CrossRef]
  87. Riffault-Valois, L.; Wattez, C.; Langrand, J.; Boucaud-Maitre, D.; Gaslonde, T.; Colas, C.; Nossin, E.; Blateau, A.; Michel, S.; Cachet, X. Health risk associated with the oral consumption of “Chiniy-tref”, a traditional medicinal preparation used in Martinique (French West Indies): Qualitative and quantitative analyses of aristolochic acids contained therein. Toxicon 2019, 172, 53–60. [Google Scholar] [CrossRef]
  88. Tmušić, N.; Ilić, Z.S.; Milenković, L.; Šunić, L.; Lalević, D.; Kevrešan, Ž.; Mastilović, J.; Stanojević, L.; Cvetković, D. Shading of Medical Plants Affects the Phytochemical Quality of Herbal Extracts. Horticulturae 2021, 7, 437. [Google Scholar] [CrossRef]
  89. Granato, D.; de Oliveira, C.C.; Caruso, M.S.F.; Nagato, L.A.F.; Alaburda, J. Feasibility of different chemometric techniques to differentiate commercial Brazilian sugarcane spirits based on chemical markers. Food Res. Int. 2014, 60, 212–217. [Google Scholar] [CrossRef]
  90. Allen, A.G.; Cardoso, A.A.; da Rocha, G.O. Influence of sugar cane burning on aerosol soluble ion composition in Southeastern Brazil. Atmos. Environ. 2004, 38, 5025–5038. [Google Scholar] [CrossRef]
  91. Solomon, S. Post-harvest deterioration of sugarcane. Sugar Tech. 2009, 11, 109–123. [Google Scholar] [CrossRef]
  92. Daniels, J.; Daniels, C. The Origin of the Sugarcane Roller Mill. Technol. Cult. 1988, 29, 493–535. [Google Scholar] [CrossRef]
  93. Inskip, S.T. Cane preparation-optimised technology. In Proceedings of the 27th International Society of Sugar Cane Technologists Congress, Veracruz, Mexico, 7–11 March 2010; Asociación de Técnicos Azucareros de México, AC (ATAM): Mexico City, Mexico, 2010. [Google Scholar]
  94. Agu, K.C.; Oduola, M.K. Kinetic modeling of ethanol production by batch fermentation of sugarcane juice using immobilized yeast. Glob. J. Eng. Technol. Adv. 2021, 7, 124–136. [Google Scholar] [CrossRef]
  95. Faria, J.B. Sugar cane spirits: Cachaça and rum production and sensory properties. In Alcoholic Beverages: Sensory Evaluation and Consumer Research; Woodhead Publishing, Ltd.: Cambridge, UK, 2012; pp. 348–358. ISBN 978-0-85709-051-5. [Google Scholar]
  96. Fahrasmane; Ganou-Parfait. Microbial flora of rum fermentation media. J. Appl. Microbiol. 1998, 84, 921–928. [Google Scholar] [CrossRef]
  97. Chantasiriwan, S. Numerical model of sugar mills with compound Imbibition. In Proceedings of the 35th Conference of the Australian Society of Sugar Cane Technologists, Townsville, QLD, Australia, 16–18 April 2013; Australian Society of Sugar Cane Technologists: Mackay, QLD, Australia, 2013. [Google Scholar]
  98. Walker, G.; Stewart, G. Saccharomyces cerevisiae in the Production of Fermented Beverages. Beverages 2016, 2, 30. [Google Scholar] [CrossRef]
  99. Freitas Schwan, R.; Mendonça, A.T.; da Silva, J.J.; Rodrigues, V.; Wheals, A.E. Microbiology and physiology of Cachaça (Aguardente) fermentations. Antonie Van Leeuwenhoek 2001, 79, 89–96. [Google Scholar] [CrossRef] [PubMed]
  100. Nicol, D.A. Rum. In Fermented Beverage Production; Lea, A.G.H., Piggott, J.R., Eds.; Springer: Boston, MA, USA, 2003; pp. 263–287. ISBN 978-1-4615-0187-9. [Google Scholar]
  101. De Araújo Vicente, M.; Fietto, L.G.; de Miranda Castro, I.; Gonçalves dos Santos, A.N.; Coutrim, M.X.; Brandão, R.L. Isolation of Saccharomyces cerevisiae strains producing higher levels of flavoring compounds for production of “cachaça” the Brazilian sugarcane spirit. Int. J. Food Microbiol. 2006, 108, 51–59. [Google Scholar] [CrossRef] [PubMed]
  102. Jeronimo, E.M.; de Souza Oliveira, E.; Souza, E.L.R.; de Almeida Silva, M.; Serra, G.E. Addition of proteic nitrogen during alcoholic fermentation for the production of cachaça. Sci. Agric. 2008, 65, 161–168. [Google Scholar] [CrossRef]
  103. Amorim, J.C.; Schwan, R.F.; Duarte, W.F. Sugar cane spirit (cachaça): Effects of mixed inoculum of yeasts on the sensory and chemical characteristics. Food Res. Int. 2016, 85, 76–83. [Google Scholar] [CrossRef]
  104. Vidal, E.E.; de Billerbeck, G.M.; Simões, D.A.; Schuler, A.; François, J.M.; de Morais, M.A. Influence of nitrogen supply on the production of higher alcohols/esters and expression of flavour-related genes in cachaça fermentation. Food Chem. 2013, 138, 701–708. [Google Scholar] [CrossRef]
  105. Beltran, G.; Rozès, N.; Mas, A.; Guillamón, J.M. Effect of low-temperature fermentation on yeast nitrogen metabolism. World J. Microbiol. Biotechnol. 2007, 23, 809–815. [Google Scholar] [CrossRef]
  106. Verbelen, P.J.; Saerens, S.M.G.; Van Mulders, S.E.; Delvaux, F.; Delvaux, F.R. The role of oxygen in yeast metabolism during high cell density brewery fermentations. Appl. Microbiol. Biotechnol. 2009, 82, 1143–1156. [Google Scholar] [CrossRef]
  107. Kiss, A.A. Distillation technology—Still young and full of breakthrough opportunities. J. Chem. Technol. Biotechnol. 2014, 89, 479–498. [Google Scholar] [CrossRef]
  108. Spaho, N. Distillation Techniques in the Fruit Spirits Production. In Distillation—Innovative Applications and Modeling; IntechOpen: London, UK, 2017; ISBN 978-953-51-3202-8. [Google Scholar]
  109. Christofoletti, C.A.; Escher, J.P.; Correia, J.E.; Marinho, J.F.U.; Fontanetti, C.S. Sugarcane vinasse: Environmental implications of its use. Waste Manag. 2013, 33, 2752–2761. [Google Scholar] [CrossRef] [PubMed]
  110. Batista, F.R.M.; Meirelles, A.J.A. Computer simulation applied to studying continuous spirit distillation and product quality control. Food Control 2011, 22, 1592–1603. [Google Scholar] [CrossRef]
  111. Faria, J.B.; Cardello, H.M.A.B.; Boscolo, M.; Isique, W.D.; Odello, L.; Franco, D.W. Evaluation of Brazilian woods as an alternative to oak for cachaças aging. Eur. Food Res. Technol. 2003, 218, 83–87. [Google Scholar] [CrossRef]
  112. Mosedale, J.R.; Puech, J.-L. Wood maturation of distilled beverages. Trends Food Sci. Technol. 1998, 9, 95–101. [Google Scholar] [CrossRef]
  113. Clyne, J.; Conner, J.M.; Paterson, A.; Piggott, J.R. The effect of cask charring on Scotch whisky maturation. Int. J. Food Sci. Technol. 1993, 28, 69–81. [Google Scholar] [CrossRef]
  114. Canas, S.; Caldeira, I.; Anjos, O.; Belchior, A.P. Phenolic profile and colour acquired by the wine spirit in the beginning of ageing: Alternative technology using micro-oxygenation vs. traditional technology. LWT 2019, 111, 260–269. [Google Scholar] [CrossRef]
  115. Canas, S. Phenolic Composition and Related Properties of Aged Wine Spirits: Influence of Barrel Characteristics. A Review. Beverages 2017, 3, 55. [Google Scholar] [CrossRef]
  116. Vivas, N.; Picard, M.; Bourden-Nonier, M.-F.; Vivas de Gaulejac, N.; Mouche, C.; Rossy, C. Heartwood dry extract: A key fraction for the quality and the diversity of rums and spirits. J. Inst. Brew. 2021, 127, 59–69. [Google Scholar] [CrossRef]
  117. Stewart, G.; Kellershohn, J.; Russell, I. Whisky and Other Spirits: Technology, Production and Marketing; Academic Press: London, UK, 2021; ISBN 0-323-85360-9. [Google Scholar]
  118. Delgado-González, M.J.; García-Moreno, M.V.; Sánchez-Guillén, M.M.; García-Barroso, C.; Guillén-Sánchez, D.A. Colour evolution kinetics study of spirits in their ageing process in wood casks. Food Control 2021, 119, 107468. [Google Scholar] [CrossRef]
  119. Santiago, W.D.; das Graças Cardoso, M.; de Andrade Santiago, J.; Teixeira, M.L.; Barbosa, R.B.; Zacaroni, L.M.; de Sales, P.F.; Nelson, D.L. Physicochemical profile and determination of volatile compounds in cachaça stored in new oak (Quercus sp.), amburana (Amburana cearensis), jatoba (Hymenaeae carbouril), balsam (Myroxylon peruiferum) and peroba (Paratecoma peroba) casks by SPME-GC–MS. J. Inst. Brew. 2016, 122, 624–634. [Google Scholar] [CrossRef]
  120. Lee, K.-Y.M.; Paterson, A.; Piggott, J.R.; Richardson, G.D. Origins of Flavour in Whiskies and a Revised Flavour Wheel: A Review. J. Inst. Brew. 2001, 107, 287–313. [Google Scholar] [CrossRef]
  121. Zea, L.; Serratosa, M.P.; Mérida, J.; Moyano, L. Acetaldehyde as Key Compound for the Authenticity of Sherry Wines: A Study Covering 5 Decades. Compr. Rev. Food Sci. Food Saf. 2015, 14, 681–693. [Google Scholar] [CrossRef]
  122. Serafim, F.A.T.; Seixas, F.R.F.; Silva, A.A.D.; Galinaro, C.A.; Nascimento, E.S.P.; Buchviser, S.F.; Odello, L.; Franco, D.W. Correlation between Chemical Composition and Sensory Properties of Brazilian Sugarcane Spirits (Cachaças). J. Braz. Chem. Soc. 2013, 24, 973–982. [Google Scholar] [CrossRef]
  123. Dos Anjos, J.P.; das Graças Cardoso, M.; Saczk, A.A.; Dórea, H.S.; Santiago, W.D.; Machado, A.M.R.; Zacaroni, L.M.; Nelson, D.L. Evolution of the concentration of phenolic compounds in cachaça during aging in an oak (Quercus sp.) barrel. J. Braz. Chem. Soc. 2011, 22, 1307–1314. [Google Scholar] [CrossRef]
  124. Liang, L.; Zhang, Y.; Zhang, L.; Zhu, M.; Liang, S.; Huang, Y. Study of sugarcane pieces as yeast supports for ethanol production from sugarcane juice and molasses. J. Ind. Microbiol. Biotechnol. 2008, 35, 1605–1613. [Google Scholar] [CrossRef]
  125. Serafim, F.A.T.; Pereira-Filho, E.R.; Franco, D.W. Chemical data as markers of the geographical origins of sugarcane spirits. Food Chem. 2016, 196, 196–203. [Google Scholar] [CrossRef]
  126. Capobiango, M.; Oliveira, E.S.; Cardeal, Z.L. Evaluation of Methods Used for the Analysis of Volatile Organic Compounds of Sugarcane (Cachaça) and Fruit Spirits. Food Anal. Methods 2013, 6, 978–988. [Google Scholar] [CrossRef]
  127. Garcia, A.C.; Serafim, F.A.T.; Keukeleire, D.D.; Franco, D.W. Evaluation of Glycerol Profiles in Sugarcane Spirits (Cachaças). J. Braz. Chem. Soc. 2015, 26, 57–63. [Google Scholar] [CrossRef]
  128. Zhang, B.; Cai, J.; Duan, C.-Q.; Reeves, M.J.; He, F. A Review of Polyphenolics in Oak Woods. Int. J. Mol. Sci. 2015, 16, 6978–7014. [Google Scholar] [CrossRef]
  129. Castro, M.C.; Bortoletto, A.M.; Silvello, G.C.; Alcarde, A.R. Lignin-derived phenolic compounds in cachaça aged in new barrels made from two oak species. Heliyon 2020, 6, e05586. [Google Scholar] [CrossRef] [PubMed]
  130. Caetano, D.; Gonçalves Lima, C.M.; Lima Sanson, A.; Faria Silva, D.; de Souza Hassemer, G.; Verruck, S.; Gregorio, S.R.; da Silva, G.A.; de Cassia Franco Afonso, R.J.; Xavier Coutrim, M.; et al. Chemical Fingerprint of Non-aged Artisanal Sugarcane Spirits Using Kohonen Artificial Neural Network. Food Anal. Methods 2022, 15, 890–907. [Google Scholar] [CrossRef]
  131. Meade, G.P.; Chen, J.C. Cane Sugar Handbook; John Wiley & Sons: New York, NY, USA, 1977; ISBN 0-471-58996-9. [Google Scholar]
  132. Ali, M.N.H.A.; Jamali, L.A.; Soomro, S.A.; Chattha, S.H.; Ibupoto, K.A.; Abbasi, N.A.; Qumi, N.M. Post-Harvest Losses and Control of Unprocessed Sugarcane. Pak. J. Agric. Res. 2018, 31, 355–360. [Google Scholar] [CrossRef]
  133. Saxena, P.; Srivastava, R.P.; Sharma, M.L. Impact of Cut to Crush Delay and Bio-chemical Changes in Sugarcane. Aust. J. Crop Sci. 2010, 4, 692–699. [Google Scholar]
  134. Gabriel, A.V.M.D.; Verruma-Bernardi, M.R.; Margarido, L.A.C.; Borges, M.T.M.R.; Nassu, R.T.; Lavorenti, N.A.; Ceccato-Antonini, S.R. Effect of the spontaneous fermentation and the ageing on the chemo-sensory quality of Brazilian organic cachaça. Cienc. Rural 2012, 42, 918–925. [Google Scholar] [CrossRef]
  135. Miller, G.L. Use of Dinitrosalicylic Acid Reagent for Determination of Reducing Sugar. Anal. Chem. 1959, 31, 426–428. [Google Scholar] [CrossRef]
  136. Mattos, D.S.; Margarido, L.A.C.; Ceccato-Antonini, S.R. Influence of sugarcane variety and management on the mineral composition of vinasse from alembic cachaça. Acta Scientiarum Technol. 2018, 40, e36581. [Google Scholar] [CrossRef]
  137. Nascimento, R.F.; Bezerra, C.W.B.; Furuya, S.M.B.; Schultz, M.S.; Polastro, L.R.; Lima Neto, B.S.; Franco, D.W. Mineral Profile of Brazilian Cachaças and Other International Spirits. J. Food Compos. Anal. 1999, 12, 17–25. [Google Scholar] [CrossRef]
  138. De Souza, J.C.; Toci, A.T.; Beluomini, M.A.; de Paula Eiras, S. Determination of copper in sugarcane spirit by flame atomic absorption spectrometry using a ternary solvent mixture (water-ethanol-acetone). Eclética Química 2017, 42, 33–39. [Google Scholar] [CrossRef]
  139. Rodrigues, B.U.; da Costa, R.M.; Salvini, R.L.; da Silva Soares, A.; da Silva, F.A.; Caliari, M.; Cardoso, K.C.R.; Ribeiro, T.I.M. Cachaça Classification Using Chemical Features and Computer Vision. Procedia Comput. Sci. 2014, 29, 2024–2033. [Google Scholar] [CrossRef]
  140. Jiao, Z.; Dong, Y.; Chen, Q. Ethyl Carbamate in Fermented Beverages: Presence, Analytical Chemistry, Formation Mechanism, and Mitigation Proposals: Ethyl carbamate in fermented beverages. Compr. Rev. Food Sci. Food Saf. 2014, 13, 611–626. [Google Scholar] [CrossRef]
  141. Rodrigues, L.M.A.; das Graças Cardoso, M.; Batista, L.R.; Santiago, W.D.; Resende, J.M.V.; de Andrade Santiago, J.; de Souza Gomes, M.; Andrade, M.A.; Teixeira, M.L.; Passamani, F.R.F. “Cachaças” (Sugarcane Spirit) Aged Quantitation of Phenolic Compounds, Antibacterial and Antioxidant Activity. Am. J. Plant Sci. 2014, 5, 2935. [Google Scholar] [CrossRef]
  142. Bettin, S.; Isique, W.; Franco, D.; Andersen, M.; Knudsen, S.; Skibsted, L. Phenols and metals in sugar-cane spirits. Quantitative analysis and effect on radical formation and radical scavenging. Eur. Food Res. Technol. 2002, 215, 169–175. [Google Scholar] [CrossRef]
  143. Bortoletto, A.M.; Alcarde, A.R. Congeners in sugar cane spirits aged in casks of different woods. Food Chem. 2013, 139, 695–701. [Google Scholar] [CrossRef]
  144. Franitza, L.; Schieberle, P.; Granvogl, M. Characterization of the Key Aroma Compounds in Rum Made from Sugar Cane Juice by Means of the Sensomics Approach. In Sex, Smoke, and Spirits: The Role of Chemistry; Guthrie, B., Beauchamp, J.D., Buettner, A., Toth, S., Qian, M.C., Eds.; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2019; Volume 1321, pp. 291–309. ISBN 978-0-8412-3467-3. [Google Scholar]
  145. Czerny, M.; Christlbauer, M.; Christlbauer, M.; Fischer, A.; Granvogl, M.; Hammer, M.; Hartl, C.; Hernandez, N.M.; Schieberle, P. Re-investigation 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]
  146. Ickes, C.M.; Cadwallader, K.R. Characterization of Sensory Differences in Mixing and Premium Rums Through the Use of Descriptive Sensory Analysis. J. Food Sci. 2017, 82, 2679–2689. [Google Scholar] [CrossRef]
  147. Maza Gomez, S. Rum Aroma Descriptive Analysis. Master’s Thesis, Louisiana State University, Baton Rouge, LA, USA, 2002. [Google Scholar] [CrossRef]
  148. González-Robles, M.I.W. Flavour Interactions Between the ‘Estery’ and ‘Mature/Woody’ Characters of Whisky, Bourbon & Tequila. Ph.D. Thesis, University of Nottingham, Nottingham, UK, 2018. [Google Scholar]
  149. Ickes, C.M.; Lee, S.-Y.; Cadwallader, K.R. Novel Creation of a Rum Flavor Lexicon Through the Use of Web-Based Material. J. Food Sci. 2017, 82, 1216–1223. [Google Scholar] [CrossRef] [PubMed]
  150. Rota, M.B.; Piggott, J.R.; Faria, J.B. Sensory profile and acceptability of traditional and double-distilled cachaça aged in oak casks: Traditional and double-distilled cachaça aged in oak casks. J. Inst. Brew. 2013, 119, 251–257. [Google Scholar] [CrossRef]
  151. Puentes, C.; Joulia, X.; Vidal, J.-P.; Esteban-Decloux, M. Simulation of spirits distillation for a better understanding of volatile aroma compounds behavior: Application to Armagnac production. Food Bioprod. Process. 2018, 112, 31–62. [Google Scholar] [CrossRef]
  152. Caetano, D.; Lima, C.M.G.; Sanson, A.L.; Silva, D.F.; de Souza Hassemer, G.; Verruck, S.; da Silva, G.A.; de Cássia Franco Afonso, R.J.; Coutrim, M.X.; Gregório, S.R. Descriptive screening and lexicon development of non-aged artisanal cachaça sensorial profile using principal component analysis and Kohonen artificial neural networks. J. Sens. Stud. 2021, 36, e12645. [Google Scholar] [CrossRef]
  153. Vilela, A.F.; de Sousa Conrado Oliveira, L.; Muniz, M.B.; de Mélo, B.C.A.; de Figueiredo, M.J.; de Mélo Vieira Neto, J. Assessment of sensory and physical-chemical quality, and potential for certification of cachaças from the state of Paraíba, Brazil. Food Sci. Technol. 2021, 41, 661–668. [Google Scholar] [CrossRef]
  154. Franitza, L.; Nicolotti, L.; Granvogl, M.; Schieberle, P. Differentiation of Rums Produced from Sugar Cane Juice (Rhum Agricole) from Rums Manufactured from Sugar Cane Molasses by a Metabolomics Approach. J. Agric. Food Chem. 2018, 66, 3038–3045. [Google Scholar] [CrossRef] [PubMed]
  155. Bortoletto, A.M.; Corrêa, A.C.; Alcarde, A.R. Fatty acid profile and glycerol concentration in cachaças aged in different wood barrels. J. Inst. Brew. 2016, 122, 293–298. [Google Scholar] [CrossRef]
  156. Zamora, F. Biochemistry of Alcoholic Fermentation. In Wine Chemistry and Biochemistry; Moreno-Arribas, M.V., Polo, M.C., Eds.; Springer: New York, NY, USA, 2009; pp. 3–26. ISBN 978-0-387-74118-5. [Google Scholar]
  157. Wang, Z.; Ickes, C.M.; Cadwallader, K.R. Influence of Ethanol on Flavor Perception in Distilled Spirits. In Sex, Smoke, and Spirits: The Role of Chemistry; Guthrie, B., Beauchamp, J.D., Buettner, A., Toth, S., Qian, M.C., Eds.; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2019; Volume 1321, pp. 277–290. ISBN 978-0-8412-3467-3. [Google Scholar]
  158. Fahrasmane, L.; Parfait, A.; Jouret, C.; Galzy, P. Production of Higher Alcohols and Short Chain Fatty Acids by Different Yeasts Used in Rum Fermentations. J. Food Sci. 1985, 50, 1427–1436. [Google Scholar] [CrossRef]
  159. Pang, X.-N.; Li, Z.-J.; Chen, J.-Y.; Gao, L.-J.; Han, B.-Z. A Comprehensive Review of Spirit Drink Safety Standards and Regulations from an International Perspective. J. Food Prot. 2017, 80, 431–442. [Google Scholar] [CrossRef]
  160. Inês, A.; Falco, V.; Inês, A.; Falco, V. Lactic Acid Bacteria Contribution to Wine Quality and Safety. In Generation of Aromas and Flavours; IntechOpen: London, UK, 2018; ISBN 978-1-78984-453-5. [Google Scholar]
  161. Azevedo, M.S.; Pirassol, G.; Fett, R.; Micke, G.A.; Vitali, L.; Costa, A.C.O. Screening and determination of aliphatic organic acids in commercial Brazilian sugarcane spirits employing a new method involving capillary electrophoresis and a semi-permanent adsorbed polymer coating. Food Res. Int. 2014, 60, 123–130. [Google Scholar] [CrossRef]
  162. The Good Scents. Available online: http://www.thegoodscentscompany.com/ (accessed on 1 July 2023).
  163. Meilgaard, M.C. Flavor chemistry of beer. II. Flavor and threshold of 239 aroma volatiles. Tech. Quart. Master. Brew. Assoc. Am. 1975, 12, 151–168. [Google Scholar]
  164. Siebert, T.E.; Smyth, H.E.; Capone, D.L.; Neuwöhner, C.; Pardon, K.H.; Skouroumounis, G.K.; Herderich, M.J.; Sefton, M.A.; Pollnitz, A.P. Stable isotope dilution analysis of wine fermentation products by HS-SPME-GC-MS. Anal. Bioanal. Chem. 2005, 381, 937–947. [Google Scholar] [CrossRef]
  165. Duarte, W.F.; de Sousa, M.V.F.; Dias, D.R.; Schwan, R.F. Effect of Co-Inoculation of Saccharomyces cerevisiae and Lactobacillus fermentum on the Quality of the Distilled Sugar Cane Beverage Cachaça. J. Food Sci. 2011, 76, C1307–C1318. [Google Scholar] [CrossRef]
  166. Gracia, I.; Rodríguez, J.F.; García, M.T.; Alvarez, A.; García, A. Isolation of aroma compounds from sugar cane spirits by supercritical CO2. J. Supercrit. Fluids 2007, 43, 37–42. [Google Scholar] [CrossRef]
  167. Cardoso, D.R.; Andrade-Sobrinho, L.G.; Leite-Neto, A.F.; Reche, R.V.; Isique, W.D.; Ferreira, M.M.C.; Lima-Neto, B.S.; Franco, D.W. Comparison between Cachaça and Rum Using Pattern Recognition Methods. J. Agric. Food Chem. 2004, 52, 3429–3433. [Google Scholar] [CrossRef] [PubMed]
  168. Boscolo, M.; Bezerra, C.W.B.; Cardoso, D.R.; Lima Neto, B.S.; Franco, D.W. Identification and dosage by HRGC of minor alcohols and esters in Brazilian sugar-cane spirit. J. Braz. Chem. Soc. 2000, 11, 86–90. [Google Scholar] [CrossRef]
  169. Nascimento, R.F.; Marques, J.C.; Lima Neto, B.S.; De Keukeleire, D.; Franco, D.W. Qualitative and quantitative high-performance liquid chromatographic analysis of aldehydes in Brazilian sugar cane spirits and other distilled alcoholic beverages. J. Chromatogr. A 1997, 782, 13–23. [Google Scholar] [CrossRef] [PubMed]
  170. Rodrigues, L.M.A.; das Graças Cardoso, M.; Santiago, W.D.; Barbosa, R.B.; de Andrade Santiago, J.; Lima, L.M.Z.; Nelson, D.L. Organic contaminants in distilled sugar cane spirits produced by column and copper alembic distillation. Res. Soc. Dev. 2020, 9, e930974879. [Google Scholar] [CrossRef]
  171. Pino, J.A.; Tolle, S.; Gök, R.; Winterhalter, P. Characterisation of odour-active compounds in aged rum. Food Chem. 2012, 132, 1436–1441. [Google Scholar] [CrossRef]
  172. Zhao, Y.P.; Zheng, X.P.; Song, P.; Sun, Z.L.; Tian, T.T. Characterization of Volatiles in the Six Most Well-Known Distilled Spirits. J. Am. Soc. Brew. Chem. 2013, 71, 161–169. [Google Scholar] [CrossRef]
  173. Silveira, A.L.; Machado, A.M.d.R.; de Cássia Oliveira Gomes, F.; Garcia, C.F.; Badotti, F. Profile of volatile organic compounds in musts and cachaças produced by selected Saccharomyces cerevisiae yeast strains. Res. Soc. Dev. 2021, 10, e23410414009. [Google Scholar] [CrossRef]
  174. Nascimento, E.S.P.; Cardoso, D.R.; Franco, D.W. Quantitative Ester Analysis in Cachaça and Distilled Spirits by Gas Chromatography−Mass Spectrometry (GC−MS). J. Agric. Food Chem. 2008, 56, 5488–5493. [Google Scholar] [CrossRef]
  175. Mosedale, J.R. Effects of oak wood on the maturation of alcoholic beverages with particular reference to whisky. For. Int. J. For. Res. 1995, 68, 203–230. [Google Scholar] [CrossRef]
  176. Monties, B. Composition chimique des bois de chêne: Composés phénoliques, relations avec quelques propriétés physiques et chimiques susceptibles d’influencer la qualité des vins et des eaux-de-vie. OENO One 1987, 21, 169–190. [Google Scholar] [CrossRef]
  177. Rodríguez Madrera, R.; Blanco Gomis, D.; Mangas Alonso, J.J. Influence of Distillation System, Oak Wood Type, and Aging Time on Composition of Cider Brandy in Phenolic and Furanic Compounds. J. Agric. Food Chem. 2003, 51, 7969–7973. [Google Scholar] [CrossRef] [PubMed]
  178. De Aquino, F.W.B.; Rodrigues, S.; do Nascimento, R.F.; Casimiro, A.R.S. Simultaneous determination of aging markers in sugar cane spirits. Food Chem. 2006, 98, 569–574. [Google Scholar] [CrossRef]
  179. Thibaud, F.; Courregelongue, M.; Darriet, P. Contribution of Volatile Odorous Terpenoid Compounds to Aged Cognac Spirits Aroma in a Context of Multicomponent Odor Mixtures. J. Agric. Food Chem. 2020, 68, 13310–13318. [Google Scholar] [CrossRef] [PubMed]
  180. Viana, E.J.; de Carvalho Tavares, I.M.; Rodrigues, L.M.A.; das Graças Cardoso, M.; Baffa Júnior, J.C.; Gualberto, S.A.; de Oliveira, C.P. Evaluation of toxic compounds and quality parameters on the aged Brazilian sugarcane spirit. Res. Soc. Dev. 2020, 9, e395985544. [Google Scholar] [CrossRef]
  181. Cravo, F.D.; Santiago, W.D.; Da Silva Lunguinho, A.; Barbosa, R.B.; Da Silva Oliveira, R.E.; Alvarenga, G.F.; Santos, S.D.; Souza, R.H.Z.; De Souza, E.C.; De Almeida, K.J.; et al. Composition of Cachaças Produced from Five Varieties of Sugarcane and the Correlation of the Presence of Dhurrin in the Cane with That of Ethyl Carbamate in the Product. Am. J. Plant Sci. 2019, 10, 339–350. [Google Scholar] [CrossRef]
  182. Cardoso, D.R.; Andrade Sobrinho, L.G.; Lima-Neto, B.S.; Franco, D.W. A rapid and sensitive method for dimethylsulphide analysis in Brazilian sugar cane sugar spirits and other distilled beverages. J. Braz. Chem. Soc. 2004, 15, 277–281. [Google Scholar] [CrossRef]
  183. Lachenmeier, D.W.; Lima, M.C.; Nóbrega, I.C.; Pereira, J.A.; Kerr-Corrêa, F.; Kanteres, F.; Rehm, J. Cancer risk assessment of ethyl carbamate in alcoholic beverages from Brazil with special consideration to the spirits cachaça and tiquira. BMC Cancer 2010, 10, 266. [Google Scholar] [CrossRef]
  184. Aresta, M.; Boscolo, M.; Franco, D.W. Copper(II) Catalysis in Cyanide Conversion into Ethyl Carbamate in Spirits and Relevant Reactions. J. Agric. Food Chem. 2001, 49, 2819–2824. [Google Scholar] [CrossRef]
  185. Article AOC “Martinique”—Décret n° 2009–1350 du 29 Octobre 2009 Relatif Aux Appellations d’Origine Contrôlées «Martinique», «Marc d’Alsace» Suivie de la Dénomination «Gewurztraminer», «Calvados», «Calvados Domfrontais», «Calvados Pays d’Auge», «Cornouaille», «Domfront», «Pays d’Auge», «Pommeau du Maine» et «Pommeau de Normandie»—Légifrance. Available online: https://www.legifrance.gouv.fr/loda/article_lc/LEGIARTI000021220378/2009-11-02 (accessed on 21 June 2023).
  186. Souza, J.C.; Toci, A.T.; Beluomini, M.A.; Eiras, S.P. Spectrophotometric Determination of Copper(II) in Sugarcane Spirit Using 1-(2-pyridylazo)-2-naphthol and a Homogeneous Ternary Mixture of the Solvents Water, Ethanol and Methyl Isobutyl Ketone. Rev. Virtual Quim. 2016, 8, 687–701. [Google Scholar] [CrossRef]
  187. Minez Tábua, M.C.; das Graças Cardoso, M.; Santiago, W.D.; da Silva Gonçalves, G.; Barbosa, R.B.; da Silva Lunguinho, A.; da Silva Oliveira, R.E.; d’Avila, G.B.; Magalhães, M.L.; Nelson, D.L. Physicochemical and Chromatographic Profiles of Distilled Sugarcane Spirits Produced in Mozambique. Am. J. Plant Sci. 2018, 9, 617–627. [Google Scholar] [CrossRef]
  188. Maia, M.V.; Suarez, W.T.; dos Santos, V.B.; de Almeida, J.P.B. Carbon dots on paper for determination of Cu2+ in sugar cane spirits samples for fluorescence digital image-based method. Microchem. J. 2022, 179, 107500. [Google Scholar] [CrossRef]
Figure 1. Representation of a sugarcane plant and a piece of cane stalk [13,24].
Figure 1. Representation of a sugarcane plant and a piece of cane stalk [13,24].
Molecules 28 06810 g001
Figure 2. Composition and general uses (non-exhaustive list) of sugarcane. Created with BioRender.com. https://app.biorender.com/ (accessed on 23 June 2023).
Figure 2. Composition and general uses (non-exhaustive list) of sugarcane. Created with BioRender.com. https://app.biorender.com/ (accessed on 23 June 2023).
Molecules 28 06810 g002
Figure 3. Factors influencing the chemical and sensorial profile of SCJSs.
Figure 3. Factors influencing the chemical and sensorial profile of SCJSs.
Molecules 28 06810 g003
Figure 4. The different spirits made from sugarcane.
Figure 4. The different spirits made from sugarcane.
Molecules 28 06810 g004
Figure 5. Production of spirits made exclusively with pure sugarcane juice over the world. Data were collected in February and July 2023.
Figure 5. Production of spirits made exclusively with pure sugarcane juice over the world. Data were collected in February and July 2023.
Molecules 28 06810 g005
Figure 6. Simplified scheme of copper still (A) and column still (B).
Figure 6. Simplified scheme of copper still (A) and column still (B).
Molecules 28 06810 g006
Table 1. Some well-known spirits in the world, the principal raw material used and the main region of production.
Table 1. Some well-known spirits in the world, the principal raw material used and the main region of production.
SpiritRaw MaterialMain Producer Countries
Armagnac (AOC)White wineFrance
BaijiuSorghum, rice or cerealChina
BrandyWineSpain, Italy, Portugal, Greece
CachaçaSugarcane juiceBrazil
Cognac (AOC)WineFrance
MezcalAgaveMexico
PiscoWineSouth America (Chili; Peru)
Rhum AgricoleSugarcane juiceFrench West Indies
RumSugarcane and its derivativesCaribbean, South America
ShōchūBarley, rice, potatoJapan
SchnapsPotato or cereal + FruitGermany
SojuCereal, potatoSouth Korea
TequilaBlue agave (Agave tequilana)Mexico (State of Jalisco)
VodkaCereal, potato, fruitRussia and Scandinavia
WhiskeyCerealScotland, USA, Canada
Table 2. Sugarcane classification [16], chromosome number of the six species of genus Saccharum [14,17,18].
Table 2. Sugarcane classification [16], chromosome number of the six species of genus Saccharum [14,17,18].
Classification
OrderPoalesSpecies
FamilyPoaceaeWild caneS. spontaneum
S. robustum
2n = 40 to 128
2n = 60 to 200
SubfamilyPanicoideaeNoble caneS. officinarum2n = 80
TribeAndropogoneaeChinese caneS. sinense2n = 116 to 120
Sub-tribeSaccharinaeIndian caneS. barberi2n = 81 to 124
GenusSaccharumEdible caneS. edule2n = 60 to 122
Table 3. Chemical properties and composition of sugarcane.
Table 3. Chemical properties and composition of sugarcane.
Chemical Properties and CompositionReferences
ColorLight gray to dark green[36]
pH4.9–5.5[36]
SugarsSucrose, glucose, fructose[19,20]
Organic acidstrans-/cis-aconitic, malic, oxalic, citric, d-gluconic, succinic, l(+)-lactic acids[32]
Nitrogenous compoundsAmino acids, proteins[36]
Organic compoundsStarch, gums, chlorophyll, anthocyanin, biotin, polysaccharides[36]
Fatty compoundsWaxes, fatty acids, phosphatides, sterols[37,38]
Phenolic acidsChlorogenic, cinnamic, hydroxycinnamic, sinapic and caffeic acids[38]
Phenolic compoundsTannin, flavones (tricin, apigenin, luteolin and glycosides derivatives)[38]
Inorganic compoundK, Cl, Mg, P, S, Ca, Si, Fe, Na, Al, Mn, Zn; SiO2, K2O, P2O5, Fe2O3[36,37]
VitaminsA, C, B1, B2, B3, B5, and B6[35,36]
Table 4. Volatile aroma compounds identified in sugarcane and its juice [39]. nd: not detected.
Table 4. Volatile aroma compounds identified in sugarcane and its juice [39]. nd: not detected.
Volatile Organic CompoundsN° CASOdor DescriptionSugarcaneJuice
Acids
Acetic acid64-19-7SourXX
Isobutyric acid79-31-2Rancid, butter, cheeseXX
Caproic acid142-62-1SweatXX
Myristic acid544-63-8Sweet spicyXX
Pentadecanoic acid1002-84-2WaxyXX
Palmitic acid57-10-3Slightly waxy, fattyXX
Alcohols
2,3-Butanediol513-85-9Fruit, onionXX
cis-2-Penten-1-ol1576-95-0Green, plastic, rubberndX
1-Hexanol111-27-3Resin, flower, greenXX
cis-3-Hexen-1-ol928-96-1GrassXX
2-Heptanol543-49-7MushroomXX
Phenylethyl alcohol60-12-8Honey, spice, rose, lilacXX
1-Octen-3-ol3391-86-4MushroomXX
2-Ethyl-1-hexanol104-76-7Rose, greenXX
Caprylic alcohol111-87-5Chemical, metal, burntndX
Lauryl alcohol112-53-8Fat, waxndX
Aldehydes
Benzaldehyde100-52-7Almond, burnt, sugarXX
2,4-Heptadienal4313-03-5Nut, fatXX
trans -2-Octenal 2548-87-0Green, nut, fatXX
trans -2-Nonenal18829-56-6Cucumber, fat, greenndX
Nonanal124-19-6Fat, citrus, greenXX
Esters
Ethyl palmitate628-97-7WaxndX
Dibutyl phthalate84-74-2Faint odorXX
Hydrocarbons
Styrene100-42-5Balsamic, gasolineXX
m-Xylene108-38-3PlasticXX
d-Limonene5989-27-5Citrus, mintndX
Phenols
4-Vinylguaiacol7786-61-0Clove, curryXX
Eugenol97-53-0Clove, honeyXX
2,4-Di-tert-butylphenol96-76-4PhenolicXX
2,6-Di-tert-butyl-methylphenol128-37-0Mild phenolic camphorXX
Heterocyclic compound
2-Pentylfuran3777-69-3Green bean, butterXX
Table 5. Rhum Agricole AOC Martinique specifications according to the decree of 31 December 2020 [64]. HPA: Hectoliter pure alcohol.
Table 5. Rhum Agricole AOC Martinique specifications according to the decree of 31 December 2020 [64]. HPA: Hectoliter pure alcohol.
Step of ProcessAOC Martinique Technical Specifications
TerritoryGeographical areaDelimited on 23 municipalities of Martinique
Raw materialCane speciesSaccharum officinarum and S. spontaneum hybrids
Cane varieties approved by INAO
Harvest periodFrom January 1st to August 31st
CultureYield ≤ 120 t/acres
Juice qualityBrix ≥ 12% and pH ≥ 4.7
Juice extractionProcessCold mechanical extraction with water and/or juice from the last roll. The juice must be filtered
FermentationTypeDiscontinuous in an open vat of 500 hL max
YeastSaccharomyces
Duration<120 h
Wine alcohol content≤7.5% ABV
DistillationTypeMulti-stage continuous distillation column without reflux, ‘creole’ column
Stripping section15 trays in stainless steel or copper—diameter: 0.7 to 2 m
Rectifying section5 to 9 trays in copper
Rum alcohol content65 to 75% ABV
Non-ethanol content (NEC)≥225 g/HPA
RumsWhite rumStorage at least 6 weeks in stainless steel vat—NEC ≥ 225 g/HPA
Straw rum (‘Élevé-sous-bois’)At least 12 months in an oak vessel and NEC ≥ 250 g/HPA
Aged rumAt least 3 years in an oak vessel and NEC ≥ 325 g/HPA
Final validationEach batch is tasted by a panel of experts who approved the AOC Martinique sensory profile
Table 6. Limits for compounds in cachaça according to Brazilian legislation [8].
Table 6. Limits for compounds in cachaça according to Brazilian legislation [8].
ComponentLimit ValueUnit
Alcohol strength38–48% of ethanol (v/v) at 20 °C
Volatile acidity (expressed in acetic acid)150mg/100 mL anhydrous alcohol
Methanol *20
Total higher alcohols<360
1-butanol *3
2-butanol *10
Total esters (ethyl acetate)200
Ethyl carbamate *210μg/L
Total aldehydes (acetaldehyde)30mg/100 mL anhydrous alcohol
Acrolein *5
Furfural + 5HMF5
Lead *200μg/L
Arsenic *100
Copper *5mg/L
* Contaminants.
Table 7. Proposed differentiation of SCJS according to their specific manufacturing processes. Dashed arrow: optional.
Table 7. Proposed differentiation of SCJS according to their specific manufacturing processes. Dashed arrow: optional.
SCJSMain Production AreaHarvest/MillingFermentationDistillationAging
Molecules 28 06810 i001
Aguardiente de CanaCentral and South AmericaTraditional methodSpontaneous for small-size producersPot stillOak or tropical wood
CachaçaBrazilFrom May to NovemberSpontaneous or with
adding yeast.
Continuously open vessel
Simple or double distillation
Pot or column still
At least 12 months in oak or tropical wood
ClairinHaitiMainly manual harvestingAddition of aromatic plantsMostly copper Charentais alembicWooden barrel
GrogueCape VerdeFrom January to July with ‘trapiche’: equipment powered by oxen or mulesLong fermentation:
days to weeks
Spontaneous or addition of yeast
with sugar 6 g/L
Artisanal pot stillAt least 12 months in wood container
Agricultural rumFrench overseas departments and MadeiraImbibition water or composed juiceImbibition water or composed juicePot or column stillOak barrel
Rhum Agricole AOC
Martinique
MartiniqueFrom January to August. Extraction at ambient temperatureOpen vat
Only Saccharomyces genus
<5 days
Simple, creole column still
From January to August
Only oak barrel
Addition of caramel prohibited
Table 8. Wort physicochemical parameters influencing fermentation efficiency.
Table 8. Wort physicochemical parameters influencing fermentation efficiency.
ConditionLowHigh
pH↓ Yeast activity↑ Bacteria proliferation
Temperature↓ Membrane fluidity → sluggish fermentation↓ Yeast activity
Oxygen↓ Membrane permeability↑ ROS production
Nutrient (nitrogen)Sluggish fermentation ↑ Higher alcohol and ester concentration
Table 11. Descriptors used in SCJS sensory analysis for new distillate or mature spirits.
Table 11. Descriptors used in SCJS sensory analysis for new distillate or mature spirits.
Type of AromaDescriptorReference
Standard
New Distillate
Descriptors
Mature
Descriptors
References
FruityFruity in generalIsoamyl acetate
Ethyl caprylate
XX[71]
AppleEthyl caproateX [71]
MelonDimethylheptanalX [71]
BananaBanana aromaXX[150]
FloralHoneyNatural honeyXX[95,150]
Floral/rose2-PhenylethanolX [71]
SugarcaneSugarcaneX [152]
HerbalGrassy/Vegetablecis-3-Hexen-1-olX [46,71,95,122]
SpicyVanillaEthyl vanillin/vanillaXX[95,146,150]
CloveEugenol/Cloves aroma X[71,150]
AnimalLeathery- X[95]
WoodWoody/OakOak wood chips X[71,95,146,150]
Phenolic- X[95]
EmpyreumaticBurntsnug- X[95]
CaramelMaltol-Caramel syrup X[71,95,146]
SmokyGuaiacol-Smoked bacon X[71,146]
CoffeeDark roast coffee X[146]
DefaultOilyHeptanalXX[71,95]
SulphuryDimethyl sulfideX [71,95]
ButteryDiacetylX [71,95]
SoapyEthyl laurateX [71,95]
Mouldy2,4,6-TrichloroanisoleX [71,95]
Solvent2-Methylpropan-1-olX [71,95]
VinegarAcetic acidXX[71]
MedicineThymolXX[71]
AlcoholicEthanolXX[150]
FermentedCompressed yeastX [150,152]
BagasseBagasseX [152]
Trigeminal
sensation
PungentFormic acidXX[71,95]
AstringentOver-brewed green teaXX[146]
SweetnessSugarXX[95,122]
BitterCaffeine solutionXX[122,150]
Burning-XX[152]
Sour tasteAcetic acidXX[153]
RetronasalPersistence-XX[149]
Table 14. Terpenes and terpenoids in SCJS (AR: agricultural rum; C: cachaça). Concentration range in mg/L or mg/100 mL aa and their sensory descriptor.
Table 14. Terpenes and terpenoids in SCJS (AR: agricultural rum; C: cachaça). Concentration range in mg/L or mg/100 mL aa and their sensory descriptor.
NameN° CASConcentration (mg/L)Type of SCJSResinousTerpenoidVegetalFloralCitrusFruityWoodyMentholReferences
Non-AgedAged
Terpenes
d-Limonene5989-27-5-89.51–99.65AR [10]
Sesquisabinene58319-04-3--C [54]
Terpenoids
Citronellol106-22-90.00113–0.00292-C [103]
Geraniol106-24-10.00015 a–0.21 c-C [103,168]
Linalool78-70-60.00035 a–0.00138 a-C [103]
Menthol1490-04-60.095 a–0.73 c6.85–8.02C-AR [10,103,168]
α-Terpineol98-55-50.00018 a–0.00046 a-C [103,162]
d-Isomenthol23283-97-8-1.44–1.77AR [10,162]
Farnesol106-28-5-1.14–2.09AR [10]
Nerolidol7212-44-4--C [54,119]
α-Cadinol481-34-5--C [54]
α-Elemol639-99-6--C [54]
Torulosol1438-65-9--C [54]
C13-norisoprenoids
β-Ionone79-77-6-0.0061–0.012AR [115,162]
β-Damascenone23696-85-7-0.0437–3.31AR [10,144,162]
a value in μg/L in the reference and converted in mg/L, c value in mg/100 mL aa, - value not found in the literature.
Table 15. Other organic compounds in SCJS (C: cachaça; G: grogue). Concentration in mg/L.
Table 15. Other organic compounds in SCJS (C: cachaça; G: grogue). Concentration in mg/L.
Compound NameN° CASConcentration (mg/L)Type of SCJSReferences
Non-AgedAged
Acetate71-50-1<LOD to 2.93-C-G[80]
Dimethylsulphide75-18-3<LOQ to 52.10 c0.01–0.59C[51,122]
2,3-Butanedione monoxime57-71-6<LOQ to 0.859 c-C[51]
Ethyl carbamate51-79-60–1455.23 c<LOD to 0.138 aC[51,74,122,170,180,181]
Methyl anthranilate134-20-3--C[54,162]
a Value in μg/L in the reference and converted in mg/L, c value in mg/100 mL aa, - value not found in the literature.
Table 16. Inorganic compounds in SCJS (C: cachaça; G: grogue). Concentration in mg/L.
Table 16. Inorganic compounds in SCJS (C: cachaça; G: grogue). Concentration in mg/L.
NameMolecular FormulaN° CASConcentration (mg/L)Type of SCJSReferences
Non-AgedAged
Anions
BicarbonateHCO371-52-3<LOD to 105.15-C[80]
ChlorideCl16887-00-6<LOD to 183-C-G[80]
SulfateSO42−14808-79-8<LOD to 12.03-C-G[80]
Cations
CopperCu7440-50-8<LOQ to 9.700.021–7C[74,122,125,130,142,167,180,181,186,187,188]
CalciumCa7440-70-20–7.700.30–10.70C-G[80,167]
CadmiumCd7440-43-9<LOQ0–0.023C[51,167]
ChromiumCr7440-47-3-0C[167]
IronFe7439-89-6<LOD to 2.200–3.60C-G[51,80,122,125,142,167]
PotassiumK7440-09-7<LOD to 2.20-C-G[80]
LeadPb7439-92-1<LOD to 0.24<LOD to 0.19C[122,125]
MagnesiumMg7439-95-40–11.200.002–0.210C-G[51,80,167]
ManganeseMn7439-96-5-0.035–2.70C[80]
SodiumNa7440-23-51.90–200–36.20C-G[80,167]
NickelNi7440-02-0-0–0.017C[80]
StrontiumSr7440-24-6<LOQ-C[51]
CobaltCo7440-48-4-0C[167]
ZincZn7440-66-6<LOD to 2.600–0.59C-G[80,167]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Corbion, C.; Smith-Ravin, J.; Marcelin, O.; Bouajila, J. An Overview of Spirits Made from Sugarcane Juice. Molecules 2023, 28, 6810. https://doi.org/10.3390/molecules28196810

AMA Style

Corbion C, Smith-Ravin J, Marcelin O, Bouajila J. An Overview of Spirits Made from Sugarcane Juice. Molecules. 2023; 28(19):6810. https://doi.org/10.3390/molecules28196810

Chicago/Turabian Style

Corbion, Claudine, Juliette Smith-Ravin, Odile Marcelin, and Jalloul Bouajila. 2023. "An Overview of Spirits Made from Sugarcane Juice" Molecules 28, no. 19: 6810. https://doi.org/10.3390/molecules28196810

Article Metrics

Back to TopTop