You are on page 1of 54

22 Cider: The Production Technology

V.K. Joshi1, Somesh Sharma1* and Vikas Kumar2


School of Bioengineering and Food Technology, Shoolini University, Solan, Himachal Pradesh – 173229
1 

Food Technology and Nutrition, School of Agriculture, Lovely Professional University, Phagwara,
2 

Punjab – 144411, India

1. Introduction
One of the most important temperate fruit crops of the world is apple (Malus domestica Borkh.). The
major apple-producing countries of the world are China, USA, India, Germany, France, Italy and Turkey
(FAO, 2018). In India, apple is commercially cultivated in the states of Jammu & Kashmir, Himachal
Pradesh, Uttarakhand and Arunachal Pradesh. It is used for both dessert and processing purposes, such
as apple juice concentrate, vinegar, apple sauce, juice, butter, preserve, candy, jam, jellies, canned and
alcoholic beverages (cider, wine, vermouth and brandy).
Out of these beverages, cider is a low alcoholic drink produced from apple, a popular beverage
especially in those countries where grapevine cultivation is not practiced due to the prevailing agro-
climatic conditions. Worldwide, cider varies in alcohol content from less than 1.2 per cent alcohol by
volume (ABV), as found in French cidredoux, to 8.5 per cent ABV in traditional English ciders. In the
United States of America, the legal definition of cider for tax purposes specifies 7 per cent or lower
ABV; anything above 7 per cent ABV falls into a different tax category – it can still be called cider, but
is taxed at a different rate (Bureau of Alcohol, Tobacco and Firearms 1998). France is the world’s largest
cider-producing country. Normandy and Brittany in Northern France is the main apple cider-producing
region and is famous for its traditional sweet cidre. Some restaurants even substitute a bottle of cider
for the usual free bottle of wine (Herrero et al., 2001). The United Kingdom leads the world in hard
cider production and consumption, though the United States of America is catching up. The quantity of
cider produced is second only to the wine produced from grapes. Out of the two, cider has become an
increasingly important commercial product in recent years (Jarvis et al., 1995). It was a common drink at
the time of Roman invasion of England in 55 BC and was drunk throughout Europe in the third century
AD. Commercial cider production started during the 19th century. Today, Bullmers alone makes 65 per
cent of the 5 million hectolitre cider produced in UK.
Recently, low alcoholic beverages have gained importance in preventing cardiovascular diseases.
Wine consumption prevents the formation of LDL and increases HDL levels (having protective effects
against heart diseases) (Joshi et al., 1999). The use of hops in cider and spices has enhanced the
antimicrobial activity and is considered significant (Joshi and Siby, 2002). Here, in this chapter, the word
‘cider’ is used to describe the fermented juice of apple. Details of its production and its quality and factors
effecting it have been described.

2. History
The first record of cider was documented in 1205. Description of cidermaking in the Mediterranean basin
is found in the works of the Roman writer Pliny during the first century AD (Pliny, 1967). Thereafter,
its production appears to have moved toward north, so that cidermaking was well established in France
by the time of Charlemagne (9th century) and probably was introduced in England from Normandy
well before Duke William’s conquest in 1066 (Revier, 1985; Roach, 1985). Cider was established in
the Basque country well before the 12th century and by the 11th and 12th centuries, it was also being
made in Contentin, the area around Caenand in the Paysd’ Auge. During the 13th century, it was found
in Southeast England and by the end of 15th, and the beginning of 16th centuries, its production had

*Corresponding author: someshsharma@shooliniuniversity.com


582 Applied Aspects of Winemaking: Production of Wine

spread to Eastern Normandy, as well as to Brittany, but it remained a drink mainly of poorer classes. The
continuing preference of the people of Brittany for wine during the 16th century is well attested by the
large quantities of wine imported through the Breton ports. By the 17th and 18th centuries, its production
in England had reached something of an art form and had become the subject matter of a number of
learned discourses. In the aftermath of the Napoleonic wars, corn growing and cattle rearing became
more profitable. Naturally, orchards and consequently, cidermaking received less skilled attention. Fallen
trees were replaced with worthless seedlings and the enormous volume of cider produced was of low or
indifferent quality, resulting in depressed prices. Even much of the trade was given to cider merchants,
who bought everything wholesale at their own price; thus, the quality of cider started deteriorating at that
time. During the 19th century, imports of continental and American apples began to the market in Britain.
Cider apples were grafted over the table type and eventually, led to the growth of the specialised table
apple industry. During the 19th century, the popularity and quality of cider declined, until it became a little
more than a cheap source of alcohol for itinerant farm workers and acquired its unfortunate ‘scrumpy’
image. Increasingly, since 1900, however, cider has prospered in new markets and the last decade has
witnessed an increase in sales against a generally static or declining consumption pattern for many other
types of alcoholic drinks. During 19th and 20th centuries, cider and perry were produced and consumed
throughout the world, for more details, see Unwin (1980).

3. Definition and Characteristics of Cider


Definition and characteristics of cider are given in Table 1. Cider is an alternative term used for cyder,
though both the terms have been used since at least 1631, including that in Australia. Nevertheless,
distinction between the two is also made. Cider is an alcoholic beverage, whereas cyder is usually apple
juice or a non-alcoholic beverage (apple juice). The fermented juice, called cider in England, is known as
‘hard cider’ in the USA. In Europe, fermented apple juice is known as cider (France), Sidre (Italy), ‘Sidra’
(Spain), ‘Applewein’ (Germany and Switzerland) where the name for the corresponding unfermented
product is clearly distinguished as apple juice. Figure 1 shows the various classes of cider. Depending
upon the alcohol content, cider could be a soft cider (1-5 per cent) or hard cider (6-7 per cent) (Downing,
1989) and may be made from fresh juice or juice from single cultivar, and may be classified as vintage
cider or white ciders which is made from decolourised apple juice or pale coloured juice (Jarvis, 1993).
The sweet cider has residual sugar from fermentation or is sweetened after fermentation, and still cider is
Table 1. Definition of Cider

1. Definition
Cider: It is a beverage obtained by the complete or partial fermentation of the juice of fresh apples or a
mixture of the juice of fresh apples and fresh pears, with or without the addition of drinking water.
2. Characteristics of Cider
Pure juice cider Other cider
Actual potential alcohol by volume Minimum 5 per cent 4 per cent
Total dry extract Minimum 16 g/l 13 g/l
Without sugar Minimum 1 g/l -
Volatile acid Maximum 1.4 g/l as acetic acid 1.4 g/l as acetic acid
Iron Maximum 150 mg /l 150 mg /l

3. Forbidden Treatments or Practices


Use of colouring matter except caramel, any other practice intended to alter the composition with a view to
deceiving the purchaser of the true nature or origin of the product or to disguise its deterioration is forbidden.
Draft regulation on cider, Resolution 7, Council of Europe, Strasbrough, 30th November 1962 (A.S./Agr / V.Sp
(14)10 Rev)
Source: Beech and Carr, 1977
Cider: The Production Technology 583

with low sugar and without carbon dioxide. Dry cider is without sugar and with an alcohol content of 6-7
per cent, whereas ciders having alcohol content not more than 1.2 per cent alcohol by volume are made by
removing the alcohol from strong cider by thermal evaporation, by reverse osmosis or generally by adding
apple juice to it (Jarvis, 1993). The cider produced by the ‘Methode Champenoise’ is called champagne
cider (Downing, 1989). Sparkling sweet cider is produced by fermenting apple juice containing not more
than 1 per cent alcohol (v/v) and the natural CO2 formed during fermentation is retained. Sparkling cider
has lower sugar and higher alcohol content of 3.5 per cent but with partial retention of CO2 formed during
fermentation. Carbonated cider is charged with commercial CO2 to produce effervescent.

Figure 1. Classification of cider

Usually, the sparkling cider is produced through two different methods:


1. Sparkling cider made from concentrated apple juice and/or fresh apple must with addition of sugars
and carbon dioxide being permitted, as well as the use of different stabilisation processes.
2. ‘Natural cider’ is made according to traditional methods, which imply, among other practices (such
as the prohibition of addition of sugars and CO2), the exclusive use of juices obtained from the
pressing of cider apples (Picinelli et al., 2000).

4. Factors Affecting Fermentation


The quality of cider depends upon various factors, such as cultivars, composition of the fruit, method of
preparation and condition of fruit fermentation.

4.1. Cultivars
Cider makers know from repeated observations that some cultivars have good, stable pressing properties,
while others give lower yields or yields which decrease considerably upon maturation (Guillermina et
al., 2006). Theoretically, cider can be prepared from any apple but choice of right cultivar is one of
the important factors influencing the quality of cider (Downing, 1989). Broadly, the cider apples are
classified into four categories (Cider Advisory Committee, 1956):
1. Bittersweet – high in tannin, low in acid
2. Bittersharp – high in tannin, high in acid
3. Sharp – low in tannin, high in acid
4. Sweet – low in tannin, low in acid
Further, different varieties of apple suitable for cidermaking have also been recommended (Table
2). The type of apple grown in countries like France and England for fermentation into cider are usually
unsuitable for other products because of their very high content of pectin-estrase and total phenol content
than that of dessert apples (Burroughs, 1973). The phenolic compounds increase the tendency to enzymatic
browning during juice extraction and the polymers of the catechins and leucoanthocyanins, which
contribute huge bitterness and astringency to the cider. It has been used for many years as a criterion for
classification of juice for suitability for fermentation into cider. However, to specify which apple variety
584 Applied Aspects of Winemaking: Production of Wine

makes the best cider is difficult (Beech and Carr, 1977). Still the use of apple juice concentrate for cider
production has increased considerably due to several advantages offered by it, such as price stabilisation,
quality maintenance and storage of concentrate for a long time without spoilage (Downing, 1989; Joshi et
al., 1991), though it leads to a loss of development of specific cultivars for cidermaking (Beech and Carr,
1977; Labelle, 1979; Jarvis et al., 1995). Since more than the cultivars, fermentation conditions effect the
quality of cider, the varietal effect is not given much importance these days. In India, cider production
is in infancy and the suitability of Indian varieties for cider production has not been adequately worked
out, though Ambri-Kashmiri, Red Delicious, Golden Pippin, Maharaji apples and crab apples, Golden
Delicious, Red delicious and Rus Pippin have been found suitable for cider making (Joshi et al., 1991;
Joshi et al., 1994). A comparative study of scabbed fruits with normal fruits showed that fruits with less
than 15 spots did not affect the fermentation behaviour or the physico-chemical and sensory qualities
of cider produced (Azad et al., 1987). Guillermin et al. (2006) compared technological and rheological
properties of two cider apple cultivars, Avrolles and Doucecoetligne. Cvr Avrolles had higher juice yield
and rheological characters than to Doucecoetligne. Use of cultivar and pasteurisation also has effect on
formation of alcoholic beverage. In a study done by Hang and Woodams (2009) on the influence of apple
cultivar juice pasteurisation on the methanol content of the wines showed that among the four cultivars
used, crispin apple yielded more methanol in hard cider than Empire, Jonagold or Pacific rose apples.
However, pasteurisation of crispin apple juice reduced the methanol content.

Table 2. Varieties Suitable for Apple Wine and Cidermaking

Varieties suitability Quality characteristics


• Delicious, Cortland, Cider, Rome Beauty Low acid group
• Jonathan, Winesap, Stayman, Higher acid group
Cider Stayman, Northern spy, Rhode
Island, Greening Wealthy, Newton, Wayne
• McIntosh, Gravenstein, Golden Russet, Cox’s Aromatic group
Orange, Roxbury Russet, Wealthy

• Dabinett, Michelin, Chisel Jersey, Harry Masters Jersey, Yarlington Bitter sweet
Mill, Viberie, Medaille, Bedan, Kermerrein
• Breakwells seedling, Backwell Red, Brown’s Apple, Crimson King, Sharp/Bitter sharp
Stoke Red Frederick, Kingston Black
• Sweet Copin, Sweet Alford, Northwood Sweet
Source: Labelle, 1980; Pourlx and Nicholas, 1980; Rana et al., 1986; Vyas and Kochhar, 1986; Jarvis
et al., 1995

4.2. Composition of Fruits


It is certainly an important factor influencing the quality of wine and cider. Carbohydrates are the principal
food constituents in apple (Table 3) but is a poor source of protein. Among the minerals, potassium,
phosphorus and calcium are present in significant amounts in apple fruit. There are large variations
for various physico-chemical and flavour characteristics amongst the various cultivars used to make
cider (Jarvis et al., 1995). The amino acid composition of cider apple includes asparagine, aspartic acid,
glutamic, serine and alanine to be present while other are just in traces (Table 4). Phenolic compounds
constitute a significant component of apple, due to its antioxidant role (Table 5). Different cultivars, like
bittersweet cultivars (particularly French and English) have relatively high concentrations of polyphenols,
conferring bitterness and astringency to the finished beverage. In addition to the procyanidins, other
classes of polyphenols are also present, such as phenolic acids (chlorogenic and p-coumaroylquinic),
together with phloretin glucoside (phloridzin) and the xylo-glucoside (Lea, 1978; Lea, 1982; Lea, 1984).
Levels of all these components in bittersweet cider cultivars may be tenfold higher than in the dessert
apples. Since the polyphenols make a major contribution to flavour, colour, pressability and also have
same antimicrobial properties, therefore their retention in cider is considered useful.
Cider: The Production Technology 585

Table 3. Composition of Apple Fruit

Constituents Average range


Physico-chemical characteristics
Calories (k cal/100 g) 37-46
Water (g) 84.3-85.6
Fiber (g) 2.0-2.4
Total Nitrogen (g/100 g) 0.04-0.05
Protein (%) 0.19
Lipid (%) 0.36
Sugar (per 100 g flesh) 9.2-11.8
Total sugars (%) 10.65-13.23
Reducing sugars (%) 7.05-10.67
Sucrose (%) 1.95-5.02
Pectin (%) 0.32-0.75
Phenolic compounds (%) 0.15-2.4
Vitamin C (mg/100 g) 3.15-5.7

       Source: Gebhardt et al., 1982; Mitra, 1991; Sharma and Joshi, 2005; Upshaw et al., 1978

Table 4. Amino Acids in Cider Apple Juices

Amino acid Concentration range


(mg nitrogen /100 ml)
Asparagine 0.19-17.5
Aspartic acid 0.35-1.5
Glutamic acid 0.21-0.31
Serine 0.04-0.65
Alpha-Alanine 0.03-0.18
Gamma-Aminobutyric acid Trace
Valine Trace
Leucine+isoleucine Trace
          Source: Beech and Carr, 1977

Table 5. Phenolic Compounds Present in Cider Apple Juices

Compound Example
Phenolic acids Chlorogenic acid
Phloretin derivatives Phloridzin
Simple catechins (-) Epicatechin
Condensed procyanidins B2
Flavonol glycosides Anthocyanins

         Source: Gebhardt et al., 1982; Mitra, 1991; Sharma and Joshi, 2005; Upshaw et al., 1978
586 Applied Aspects of Winemaking: Production of Wine

4.3. Yeast Strains for Fermentation


In traditional fermentation, where no yeast is added and no sulphite is used, the first few days are dominated
by the non-Saccharomyces species (Candida pulcherrima, species of Pichia, Torulopsis, Hansenula and
Kloeckera apiculata) which multiply quickly to produce a rapid evolution of gas and alcohol. They also
generate a distinctive range of flavours, characterised by ethyl acetate, butyrate and related esters. As the
alcohol level rises (2-4 per cent), these initial fermenters begin to die out and the microbial succession is
taken over by Saccharomyces uvarum which completes the conversion of sugar to alcohol and generation
of a more wine-like flavour. The yeast cells become sub-lethally damaged by the increasing concentration
of alcohol, but death does not occur until the concentration exceeds 9 per cent alcohol (v/v). The formation
of higher alcohols and esters during alcoholic fermentation is related to a particular yeast strains used
(Castelli, 1973). The desirable characteristics for yeast used in cidermaking include production of
polygalacturonase to breakdown soluble pectins; produce rapid onset of fermentation; relatively resistant
to SO2, low pH value and high ethanol level; low requirements for vitamins, fatty acids and oxygen; ferments
to ‘dryness’ (i.e. no residual fermentable sugar); does not produce excessive foam; efficient utilisation
of sugars; minimal production of SO2; non-producer of H2S and acetic acid; produces required aroma
components, organic acids and glycerol (Beveridge, 1986; Beech 1993). The effect of different yeast starter
cultures on the overall quality of pumpkin wine reveals that a mixed culture of Saccharomyces cerevisiae,
Torulaspora delbrueckii and Zygosaccharomyces rouxii produced a wine with better quality in comparison
to individual culture (Sharma et al., 2018). The number and type of yeast in the apple juice are also
influenced by the method of juice extraction (Table 6). Consequently, the quality of cider is also affected.
Table 6. Comparison of Yeast Composition of Apple Juice Extracted by the Two Methods of Pressing

Number of yeasts/ml

Species Screw press Hydraulic press

Kloeckera apiculata 360.000 160,000

Saccharomyces spp. - 23,000

Pichia spp. 5,000 23,000

Torulopsis spp. - 2,030,000

Candida pulcherrima - 23,000

Carotenoid-containing yeasts - 23,000

Source: Beech, 1977


      

4.4. Nitrogen Source


Nitrogen containing compounds of must are important to the growth of yeast and hence, to fermentation
rate and aroma compounds production. Supplementation with nitrogen source is also essential as in its
absence, the yeast uses the amino acids of the must resulting in the formation of higher alcohols (Amerine
et al., 1980). DAHP @ 0.1 per cent has been used as a yeast food in alcoholic fermentation. Addition
of nitrogen source to the musts made from apple juice concentrate and that from pear for alcoholic
fermentation has also been reported (Joshi and Sandhu, 1994). The effect of nitrogen on the quality of
persimmon wine has similarly been documented (Mahant et al., 2017).

4.5. Must Clarification


Clarification of cider must is an indispensable factor affecting the overall quality. Pectic enzymes are
often added to the juice to hasten and improve the clarification of cider during and after fermentation.
Enzymatically clarified apple wines were rated better in terms of colour, appearance, body and flavour
than non-clarified apple wines (Joshi and Bhutani, 1991). The addition of increasing levels of insoluble
Cider: The Production Technology 587

solids to the apple juice leads to the production of undesirable physico-chemical characteristics on the
apple wine (Joshi et al., 2013). Thus, to prepare quality apple wine, juice without insoluble solids by pre-
settling and clarification using pectolytic enzyme should be practiced.

4.6. Sulphur Dioxide


The treatment of juice with SO2 before fermentation is the most common means for controlling undesirable
micro-organisms as well as to prevent enzymatic and non-enzymatic browning reactions and is a well-
established practice in winemaking (Beech and Carr, 1977; Rana et al., 1986). Table 7 shows the effect of
SO2 and temperature in cider. If SO2 is added immediately after pressing, nearly all the colour (chemically
and visually) is reduced as the sulphite binds to the quinoidal forms. If the sulphite is added at the later
stage, less reduction in colour will take place; presumably the quinones become more tightly cross-linked
and less susceptible to nucleophilic addition and reduction. Sulphur dioxide has also clarifying action and
reduces volatile acidity while exerting solvent effect on anthocyanin pigments (Amerine et al., 1980).
Usually, 100-200 ppm SO2 is added to the musts for cider production.

Table 7. Effect of SO2 and Temperature in Cider and Wine Preparation

Concentration Effects Optimum Low temperature High temperature


temperature
SO2 (50-200 Controls undesirable micro 15-18°C Less activity of Enhanced growth
ppm) organisms bacteria and wild of thermophillic
yeast in the must organisms

Prevents enzymatic Less loss of volatile More loss of volatile


browning of the juice aromatic principles compounds
Has clarifying action, i.e. Greater alcohol yield Slowing down of
it neutralizes negatively fermentation
charged colloids
SO2 has solvent effect on More residual carbon Less carbon dioxide
anthocyanin pigments dioxide production production
Reduces volatile acidity
Increases glycerol
production

Source: Amerine et al., 1980; Frazier and Westhoff, 1995; Jarvis, 1993

4.7. Temperature
The temperature affects the rate of fermentation and nature of metabolites formed. It takes 3 to 4 weeks
to attenuate cider fermentation at temperatures within the range of 20-25°C. Although temperature of
15-18°C is preferred for flavour development in Germany and France, the optimum temperature for cider
fermentation was found to be 15-18°C (Jarvis et al., 1995). Higher temperature increases the rate of
fermentation but enhances the chances of contamination with undesirable thermophillic microorganisms
(Frazier and Westhoff, 1995). The changes in viable cell count, ethanol, glucose, fructose and sucrose
during cider fermentation at 20°C with Saccharomyces cerevisiae have been noted.

4.8. Factors Affecting Flavour of Cider


Flavour is one of the important sensory attribute that determines the overall quality of cider. The various
factors that influence the flavour of cider are described in Table 8.
588 Applied Aspects of Winemaking: Production of Wine

Table 8. Factors Influencing the Flavour of Cider

Specific Variety of fruit(s); Maturity and condition of apple fruit


Apple juice
at pressing
Other ingredients of raw material Fresh juice or concentrate
Condition of concentrate
Type of chaptilization sugar(s)
Quantity of SO2
Yeast nutrients, amelioration of pH by addition of acid
Yeast Natural or Inoculated fermentation
Strain of yeast(s)
Condition of yeast(s) when inoculated
Fermentation Temperature, time of fermentation
Fermenter design and operation Hydrostatic pressure; operational pH/acidity level
Secondary fermentation Natural or induced malo-lactic fermentation secondary yeast
fermentation
Spoilage organisms

Maturation Chemical and enzymatic changes


Processing factors Decolourisation of juice or final cider
Dealcoholisation
Final product make-up Carbonation

Source: Jarvis et al., 1995

5. Cider Production Technology


5.1. Methods of Cidermaking
The cider production process that combines two successive biological fermentations: the first is the
classical alcoholic fermentation of sugar into alcohol conducted by yeast strains, like Saccharomyces
cerevisiae species and the second is malo-lactic fermentation that occurs during maturation process by
lactic acid bacteria. The latter is an important manufacturing step to reduce the acidity of cider and
stabilises it with respect to microbial spoilage through the bacteriostatic effect of the lactic acid produced.
The different methods used to make cider have been reviewed in a systematic way earlier (Amerine et
al., 1980) and are summarised in Table 9. There are some reports on the studies of cider preparation from
India also (Kerni and Shant, 1984).The yeast strain affect the formation of flavour compounds also in
cider (Jarvis et al., 1995). A method for sweet cider making has been developed. Cider with 5 per cent
alcohol, TSS/acid ratio of 25 was found to be the most favoured at laboratory and consumer survey scales
(Rana et al., 1986).

5.2. Raw Materials


Geographic location may have a greater effect upon the finished product than the cultivar itself.
Americans interested in cultivating English and European cider apple cultivars may not grow them in
their particular location; hence different cultivars are recommended in different areas of even small
countries. In traditional orchards, fruits are generally allowed to fall naturally or are shaken from the
trees using long poles (lugs), and are then, picked up either by hand or by machine, but in intensive
bush orchards, mechanically shaking of the tree permits fruits to fall which are collected and washed
immediately after falling and transferred to the mill (Jarvis, 1993). Generally, apples for cider production
are different from culinary apples as they have a higher tannin and sugar content, but are lower in acid.
Dessert and culinary apples lose more body and flavour during fermentation (Smock and Neubert, 1950)
than do the cider apples.
Table 9. Summary of Methods Used in Cider Preparation

Type of Fruit Juice Parameters/additives Fermentation Maturation Others


method
European
Method-1 (a) Some stored for Extracted as usual, SO2 50-100 mg/1 Temperature 4.4-10°C, Secondary Malo-lactic
3-4 days and others cold stabilised at pure yeast in some, mixed fermentation fermentation,
macerated 0-7.8°C in others in casks for several produces CO2 in
months bottles
(b) – – Pectic enzymes for
clarification
Cider: The Production Technology

Method -2 Lower sugar, higher Juice extracted, no Lactic acid added to Pure yeast such as – –
acidity maceration, juice increase acidity, if Steinberg added
centrifuged for needed
bacteria and yeast
removal
Method -3 Sound fruits separated by Juice extracted in – Natural fermentation Storage in concrete Before delivery,
flotation hydraulic press from 1.008-1.005 sp. tanks lined with wooden cider is sweetened
gravity coatings with syrup
Method -4 – – – Fermentation allowed up Stored in wooden casks Carbonated and
to TSS of 5-7.5°B filtered bottled
or centrifuged
American

Method -1 (a) Sound apples are used Juice is extracted Sulphur dioxide Spontaneous fermentation – Clarified by
for cider making in usual press after 100-125 ppm added, may begin during settling bentonite
crushing in a mill glucose added to give treatment, blended
13 per cent alcohol to give 10°B,
filtered, bottled
and pasteurised
(b) Juice extraction Juice made from Sweetened Yeast Champagne, 24.4°C – –
instead of apple juice concentrate temperature was the best
concentrate used
589

Source: Amerine et al., 1980


590 Applied Aspects of Winemaking: Production of Wine

Over the years, cider apples have been classified into various categories (see earlier section on
cultivars) based on the properties of the juice they yield. A recent classification is based on the level of
flavour, acids and tannins (Beech and Carr, 1977). Apples for cidermaking should be mature and free
from starch. Blending has always been an important step in controlling uniformity of the finished product
and suggestions for blending have also been made (Pourlx and Nicholas, 1980). The preferred procedure
is to use fermented stock for blending because the effect of fermentation on the fresh juice cannot always
be accurately predicted.
Juice from apples in the sweet group is generally, considered good for blending with strong flavoured
juices, while that in the bittersweet group gives cider a tangy sensation. Juice for making sparkling sweet
cider should neither be sweet nor too heavy in the body. Astringency is considered less important than
the correct sugar/acid ratio. Campo et al. (2007) showed that in the apple musts with high content of acid
and phenols, having malo-lactic fermentation first followed by alcoholic fermentation, had comparatively
low production of acetic acid.
The sweet, low-acid cultivars, such as Delecious, Cortland, Ben Davis and Rome beauty are
recommended for basic juice, while those of Jonathan, Stayman Winesap, Northern Spy, Rhode Island
Greening, Wayne, Newtown possess higher acid levels and add tartness to the cider. MacIntosh,
Gravenstein, Ribston Pippin, Golden Russet, Delecious are aromatic and add flavour and bouquet to the
cider. The body and flavour can be improved by using astringent apples, such as Red Astrakhan, Lindel,
and crab apples. A good thumb rule is to add less than 10 per cent of astringent cider to an acidic cider
and not more than 20 per cent should be added to any blend.

5.3. Milling and Pressing


Fresh and ripe apple fruits here used to extract the juice are generally stored for a few weeks after harvest
so that all the starch can be converted into sugar. Apples selected for juice processing are then, washed
and inspected for the presence of any foreign materials and decayed fruits, which have adverse effects on
microbiological status and ultimate cider quality. Earlier practice was to empty bulk truckloads or bins
of apples on a de-leafing screen into a tank of water. A circulating pump was used to direct the apples
to an elevating conveyor which discharged to an inspection belt. At this point, inspectors removed any
damaged or decayed fruit and extraneous material. Rinsing with clean water was accomplished at the
scrubber or after inspection. The routine replacement of holding water was necessary. Fruits were also
transferred into the mill using a water flume, which provided an additional advantage of washing of fruits
(Jarvis, 1993).
In preparation for pressing, the apples are ground to a mash using either a hammer or grating mill
and even slicers are required for difficult extraction. A recent development in the production of apple
juice for concentrate is by diffusion extraction (Bump, 1989; Downing, 1989) and the juice can be used
for cidermaking. Several types of equipment have been developed for pressing and extracting apples.
These include hydraulic presses, screw presses, basket presses, belt presses and pneumatic presses. The
factors involved in the choice of equipment are production capacity, product yield, ease of cleaning and
sterilisation and length of production season (Bump, 1989). For many years, rack and frame press has
been used in the apple juice industry which consists of a frame containing a slatted board covered by
a cloth into which a measured amount of mash is transferred and corners of the cloth are folded over
to form an envelope. The frame is removed and next slatted board added, together with the frame and
another cloth; the procedure is repeated 10-20 times. Then, it is pressed hydraulically to remove the juice,
under a ram pressure of 15 Mpa giving a recovery of 80 per cent of juice (Jarvis, 1993).
The advantage of this type of press is that no press aids are required and the apple juice produced
has a low level of solid particles, though it involves high labour cost alongwith the need for great care
in cleaning and sterilising the cloths and racks is required. The other presses used are the stoll press, the
bucher-guyer press, bullmer continuous belt press, atlas-pacific press for the apple juice extraction. The
use of screw presses for the recovery of additional juice for hydraulically pressed pomace in France has
been made for additional recovery of juice from centrifuged apple pulp containing cellulose fibres (Lowe
et al., 1964). Using press aids, the Zenith and Jones presses were among the first continuous presses used
for apple juice extraction in the United States of America. The amount of apple solids in the juice from
the screw presses is much greater than that in juice from hydraulic rack and cloth presses (Bump, 1989).
Cider: The Production Technology 591

A screw press that is well adapted to apple juice production is the Reitz press system. Electroacoustic
dewatering process is also one of the recent methods that employs passing of electric current through
the pulp prior to pressing has been claimed to release a higher yield of juice. Immediately after pressing,
the juice is treated with sulphur dioxide, which acts both as antioxidant to prevent browning of the juice
and a preservative by destroying wild yeast and bacteria (Bump, 1989). From an economic standpoint,
maximum recovery of juice is most important for cheap cider production. Cider can also be produced
from the apple concentrate directly after diluting it to a desired Brix level.

5.4. Controlling Microorganisms Before Fermentation


In earlier times, natural fermentation of apple juice was the common practice. However, for proper
fermentation the microflora of the juice must be controlled before inoculation with yeast to avoid off-
flavour or similar defects in the cider (Beech and Carr, 1977). There are several methods to accomplish
this (Beech and Carr, 1977). In northern France, centrifuging or fining of the juice with gelatine and
tannin followed by filtration to reduce the rate of fermentation is practiced. Another approach is to treat
the juice with pectin hydrolysing enzymes and filter before adding yeast. The danger of bacterial spoilage
is still present with these methods. Use of sulphur dioxide is made extensively for this purpose. The
natural fermentation of apple juice depends upon the ability of naturally occurring yeasts in the juice to
convert the fruit sugars to ethyl alcohol. These yeasts are native to fruit or normal contaminants on the
pressing equipment. Liang et al. (2004) reported that Pulsed Electric Field (PEF) can also be used for
inactivation of spoilage microorganisms. Further, the use of PEF along with clove oil showed additional
reduction in the microbial count.
The Swiss reportedly dilute sterile fourfold concentrate to a specific gravity of 1.050, treat the juice
with 35-40 ppm SO2 and then, pitch inoculate with yeast two days later. However, sterile juices or diluted
concentrates with little or no sulphur dioxide need to be kept in sterile containers to prevent contamination
with spoilage microorganisms.
The treatment of juice with SO2 before fermentation is undoubtedly the most common means for
controlling undesirable microorganisms (Table 10) but the amount required depends on the pH of the
juice as well as on the concentrations of the sulphite-binding compounds that are present in the juice
(Table 11). Since the effectiveness of SO2 is pH-dependent undissociated form (so called molecular SO2),
both SO2 and lysozyme prevent the development of undesirable bacterial fermentations. Addition of
lysozyme and oenological tannis during alcoholic fermentation could represent a promising alternative
to the use of SO2 and for the production of wines with reduced content of SO2 (Sonni et al., 2009). The
cider apple juices should always be brought below a pH of 3.8 by the addition of malic acid before SO2
addition. The SO2-binding compounds produced by acetic acid bacteria are present in greater quantities
in the poor quality fruit. The binding of SO2 is dependent upon the nature and origin of carbonyl group.
Compounds, such as glucose, xylose, and xylosone present in the juice bind with SO2. Such juices will
require addition of higher amounts of SO2 to control the microorganisms effectively. Consequently, all
the additions must be completed immediately after pressing the juice, provided the initial fermentation
is inhibited and further addition of free SO2 can be made during the following 24 hours (Jarvis, 1993).

Table 10. Typical Microorganisms of Freshly Pressed Apple Juice

Microorganisms type Typical species


Yeast Saccharomyces cerevisiae, S. uvarum, Saccharomycodes ludwigii, Kloeckera
apiculata, Candida pulcheriima, Pichia spp., Torulopsis famata, Aureobasidium
pullulans, Rhodotorula spp.
Bacteria Acetobacter xylinum, Pseudomonas spp., Escherichia coli, Salmonella spp.,
Micrococcus spp., Staphylococcus spp., Bacillus spp., Clostridium sp.

Source: Beech and Carr, 1977

The composition of musts may also control the bacterial population. In fact, a pH lower than 3.5 is
recommended for initial musts (Jarvis et al., 1995) and apple varieties rich in phenolic compounds are
usual in cidermaking (Guyot et al., 1998). Juices of pH >3.8 should be brought down to this value by
592 Applied Aspects of Winemaking: Production of Wine

blending or acid addition and then, 150 ppm SO2 is added. However, it has been noted that juices with
pH 3.8 could not be satisfactorily treated within the legal limit of 200 ppm SO2 (Beech, 1972; Burroughs,
1973). After sulphiting, the juice should be allowed to equilibrate for a minimum of six hours before free
SO2 is determined. In countries where legal limits of sulphur dioxide lower than 200 ppm are prescribed,
the best approach would be to use only good microbial quality fruit, maintain good plant sanitation and
monitor the pH of the raw material. In Table 11, the amount of sulphur dioxide to be added to the juice
of particular pH is given.

Table 11. Amount of Sulfur Dioxide to be Added to Juice Based on pH

pH Concentration of sulphur dioxide (ppm)


(Sodium metabisulphite solution)
3.0-3.3 75
3.3-3.5 100
3.5-3.8 150
         Source: Beech, 1972; Burroughs, 1973

5.5. Amelioration
In alcoholic beverage-production terminology, correction of raw material to make a product of consistent
quality is referred to as amelioration, i.e. adjustment of the sugar and/or acid content of the juice, as
regulated by the respective standards. Controlling the sugar content of apple juice is required to maintain
the proper final alcohol content. It is achieved by the addition of water, juice from the second pressing of
the pomace, sugar, or concentrated juice. Initial sugar concentration (ISC) influences the quality of the
cider and its value of 20°B was found optimum (Joshi and Sandhu, 1997). Fortification of apple juice after
dilution from its concentrate with diammonium hydrogen phosphate is essential for rapid fermentation, as
discussed earlier. The must prepared by direct dilution of the concentrate reportedly ferments faster than
that ameliorated with sugar (Joshi et al., 1991; Joshi and Sandhu, 1994).

5.6. Inoculation
The traditional method of cidermaking does not employ any external source of yeast. The indigenous
micro-flora of apple in the order of 5×104 cells per gram of stored fruits carries out spontaneous
fermentation (Lea, 1995). After sulphiting the juice, it is inoculated with the desired yeast culture in case
of inoculated fermentation, wherever employed. The growth of yeasts, acetic acid and lactic acid bacteria
can be excluded by washing and sorting of apples before milling and pressing. High counts of bacteria
(including lactic acid bacteria) were observed during alcoholic fermentation and storage of cider. The
levels of LAB found in musts fermented in small vessels, using acid-washed apples, were low. However,
the must fermented by using unwashed apples but blended with different varieties had a limited number
of microorganisms only. The growth of microorganisms could have been limited by fermentation and
storage temperature of 10°C together with low pH (Ribereau-Gayon et al., 1975). A yeast strain making
a clean-tasting beverage, with a minimum of yeasty flavour and a maximum fermentation rate along with
other desirable characteristics, is selected for the preparation of cider. The use of a mixed inoculum of
S. uvarum and S. bayanus is a widespread practice, on the grounds that the first yeast provides a speedy
start, but the second will cope up better with the fermentation to dryness to produce high alcohol level.
These dried yeasts require no pre-propagation and are simply hydrated in warm water before pitching
directly into the juice. A small quantity of heat-sterilised juice is inoculated with a dry culture or liquid-
nitrogen-frozen culture and after fermentation, the inoculum is added to a larger volume of sulphited juice.
The procedure is continued until a final inoculum of 1 per cent or greater by volume is obtained. It may be
added that not many different cultures are actually used in cidermaking (Beech and Carr, 1977), such as
A.W. Y.350r (Saccharomyces uvarum) (Australian Wine Research Institute, Aldaide), Champagne strains,
Champagne Epernay, Geisenheim GE 1 (Institute for Microbiology and Biochemistry at Geisenheim),
Cider: The Production Technology 593

Pinnacle (No. 729), Montrachet (UCD 522) or Champagne A.Y.D. (Australian Wine Research Institute,
Aldaide Universal Foods Corporation, USA).
A comparison of two methods of cidermaking has also been made with respect to evolution of
microbial population and malo-lactic fermentation (Deunas et al., 1994). The two methods were: the
traditional method where unwashed apples of different varieties were used and in the other, a sole acidic
varieties of apple with temperature control during fermentation was used. The frequency distribution
(%) of yeast species isolated during cidermaking is summarised in Table 12. The occurrence of malo-
lactic fermentation together with alcoholic fermentation is not considered desirable in French and
English ciders (Salih et al., 1990) and degradation of malic acid occurs after alcoholic fermentation.
However, it does not occur until the population of lactic acid bacteria reaches 106 CFU/ml (Deunas et al.,
1994). Interestingly alcoholic fermentation was carried out by Kloeckera apiculata and Saccharomyces
cerevisiae and the distribution was found similar in both the methods. In the traditional method, the malo-
lactic fermentation proceeded at the same time as alcoholic fermentation but in the modified method, no
malo-lactic fermentation occurred, but produced cider with lower volatile acid. Controlled malo-lactic
fermentation in cider using Oenococcus oeni immobilised in alginate beads has been made as a starter
culture (Herrero et al., 2001). Malic acid degradation was similar to that with free cells of Oenococcus.
Immobilised cells synthesised less ethanoic acid and ethyl ethanoate but the profile of evolution of
pyruvic acid, shikmic acid and succinic acid was similar. The immobilised cells produced more ethanol
during earlier four days but it declined during the later periods. The results are promising with respect to
production of better quality cider but may need more research work (Nedovic et al., 2000).
Another interesting approach for continuous production of cider was attempted with the use of Ca-
alginate material to co-immobilised Saccharomyces bayanus and Leuconostoc oenos in one integrated
biocatalyst system which permitted much faster fermentation than traditional cidermaking with better
flavour fermentation. After completion of fermentation, D-lactate was produced while L-lactate progress
of lactic acid bacteria in cidermaking also took place.

Table 12. Frequency (%) of Yeast Species Isolated During Cidermaking

Period of fermentationa
Species A B C
Sacchromyces cerevisiae 1b 12 42 78
Zygosaccharomyces cidri 1 - 2 -
Zygosaccharomyces florentius 1 2 2 -
Kloeckera apiculata 2 68 52 12
Sacchromycodes ludwigii 2 4 - 2
Candida pulcherrima 3 6 - -
Rhodotorula rubra 3 2 - 2
Torulasporadel brueckii 3 2 2 -
Candida vini 3 2 - 2
Pichiamem braaefaciens 3 2 - 4
Source: Deunas et al., 1994
a A: After barrel filling, at beginning of alcoholic fermentation; B: Active fermentation (density at 20°C

between 1.035­1.005); C: at the end of alcoholic fermentation (density at 20°C below 1.005).
b 1: Yeasts with strong fermentative metabolism; 2: apiculata species with low fermentative activity; 3:

species with mainly an oxidative metabolism.

5.7. Fermentation
Mostly stainless steel tanks are used these days for fermentation of cider (Downing, 1989) though
traditionally barrels of oak have been used for this purpose. Wooden barrels or vats of mild steel, with
a ceramic or resin lining, bitumen lined, concrete vats and, more recently, stainless steel and even lined
594 Applied Aspects of Winemaking: Production of Wine

fibre glass-resin vats or tanks have been employed commonly for cider fermentation. The former may
still be employed but today most companies use vertical stainless steel tanks while other cidermakers
use conico-cylindrical vats. These tanks may be equipped with temperature controlled systems, level
indicators and carbon dioxide venting and blanketing systems. Correct sulphiting of the juice and proper
cleaning of all the equipment will help ensure a good start to fermentation.
The best procedure for assuring a good fermentation is to employ larger inocula as all yeast strains
perform in the same manner at the same concentrations. Either juice or cider, if exposed to air during
fermentation, will usually develop a surface film of acetic acid bacteria or yeast. This aerobic spoilage can
be prevented by excluding the air from the vats properly by sulphiting the juice. Treatment of juice before
inoculation influences the fermentation. Heated juice ferments faster than unheated juice but sulphited
juices ferment slower than those not treated with sulphur dioxide. The availability of soluble nitrogen
in the juice affects the rate of fermentation of cider. Fermentation with Schizosaccharomyces pombe
reduced the malic acid in several fruits, including apple (Azad et al., 1986), though with low rate of
alcohol production (Parkand John, 1980). Leuconostoc has also been employed to reduce the acidity
of the fermented product. Simultaneous inoculation of apple juice with Saccharomyces cerevisiae and
Schizosaccharomyces pombe produced cider with acceptable levels of alcohol and acidity. Ion exchange
sponge with tailored surface charge for immobilisation of Saccharomyces cerevisiae encouraged yeast
growth but reduced fermentation (O’Reilly and Scott, 1993). In most of the ciders studied, the malo-lactic
fermentation and the alcoholic fermentation started at the same time.
The best flavoured cider is generally produced by a slow fermentation process. The rate of
fermentation may be controlled by maintaining the temperature around 16°C, by reducing the yeast
population by racking or by the addition of sulphur dioxide. However, if the cider fermentation is too
slow, it may be susceptible to cider sickness, imparting milky white appearance to the cider with a sweet
pungent odour. This condition is encouraged at the elevated temperatures but is reduced in ciders with 0.5
per cent malic acid. Occasionally, fermentation may be slowed down or even stalled. Aeration has been
found useful for restoring yeast activity in such cases (Burroughs and Pollard, 1954). Stuck fermentations
could be restarted if the temperature could be 12-13°C, some fermentable sugar remained and at least
10,000 yeast cells per ml are present in the fermenting musts (Whiting, 1961).
A process was developed, based on alcoholic fermentation of the available carbohydrates present
in ciders. The impact of inhibitors at different pH, size and reuse of inocula and different nutrient
supplementation on the ethanol yield were evaluated. The use of a 0.5 g/l yeast inoculum and corn steep
water as the nutrient source allowed depletion of the sugars in less than 48 hours, which increased the
content of ethanol to more than 70 g/l (Seluy et al., 2018).

5.8. Clarification
Juice and cider can also be clarified and one of the clarification treatments consists in adding gelatine. This
treatment can be used either on the juice before fermentation or on cider before bottling. By removing
selectively high DP procyanidins, this treatment modifies both the total tannin content and the profile of
the residual tannins, and thus, may change the composition and the taste. However, the most common
clarifying method in French cidermaking uses endogenous pectin methyl esterase as a clarifying agent.
Calcium is added to induce a formation of a calcium pectate gel that includes all particles of the cloudy
juice. This gel is then, separated by ‘natural’ flotation due to CO2 bubbles of the beginning fermentation.
This process produces a pectin-free clarified juice with a reduced nitrogen amount and results in slower
fermentation and better stabilisation (Que´re et al., 2006).
After completion of fermentation, the cider is left on the lees for few a days to facilitate the yeast to
autolyse, thereby adding enzymes and amino acids to the cider. The cider will be separated from the lees
and transferred after clarification into the storage vats or storage tanks (Jarvis, 1993). Initial clarification
may be performed by the natural settling of a well-flocculated yeast, by centrifugation, by fining, or by a
combination of all the three. Typical fining agents are bentonite, gelatine, isinglass or chitosan. Gelatine
forms a block with negatively charged tannins in the cider and brings down other suspended materials by
entrapment, and can also be used together with bentonite for similar effect.
Cider: The Production Technology 595

5.9. Aging/Maturation and Secondary Fermentation


After clarification, cider may be bulk-stored or bottled. Extreme care in the sanitation of storage vessels
however, is necessary to prevent contamination with undesirable microorganisms. Stored cider should
be cultured periodically and removed from storage for special processing if unwanted growth occurs.
Storage temperature can be as low as 4°C, but not higher than 10°C. Sparkling or charged ciders have to
be stored in pressure tanks to avoid any loss of CO2. Uncharged cider should be kept in a full, closed tank
with an air trap or under a blanket of carbon dioxide or nitrogen or a mixture of the two (Cant, 1960). If air
is not excluded from the tanks, acetic acid bacteria will produce acetic acid taints. Film yeasts, which may
develop, also produce volatile acids in cider. A temperature of 4°C for bulk storage of cider is desirable.
After fermentation, the cider is racked and filtered. Maturation is an important step in cidermaking during
which most of the suspended material settles down, leaving rest of the liquid clear which may be clarified
with bentonite, casein, gelatine followed by filtration.
Enumeration, isolation and identification of lactic acid bacteria in processing and storage of
Australian cider revealed Leuconostoc oenos as the predominant bacteria (Salih et al., 1990). During
maturation, growth of LAB cultures can occur extensively, especially if wooden vats are used (Table
13). This growth results in malo-lactic fermentation. Such fermentation would convert malic acid into
lactic acid and reduce acidity, impart subtle flavour, which generally improves the flavour of the product.
However, in certain circumstances, metabolites of LAB cultures damage the cider flavour by excessive
production of diacetyl, the butter scotch-like taste (Jarvis, 1993). Since malic acid is a predominant acid
in the apple, so reduction in acidity due to malo-lactic acid fermentation might be detrimental to the
quality of cider (Salih et al., 1990).

Table 13. Progress of Lactic Acid Bacteria During Cidermaking

Period During active During malo-lactic During storage


fermentation fermentation
Sample Total LAB Species Total LAB Species Total LAB Species
count count count
(CFU/ml) (CFU/ml) (CFU/ml)
A11 1.2 × 105 L. oenos 7.4 × 106 L. oenos 7.5 × 106 L. oenos
A2 1.2 × 105 L. oenos 1.2 × 107 L. oenos 1.2 × 107 L. oenos
B1 5.8 × 10 5 L. oenos 6.8 × 10 6 L. oenos 1.0 × 105 L. oenos
Tediococcub
sp.
B2 9.3 × 106 L. oenos 1.2 × 107 L. oenos 1.5 × 106 L. brevis
L. mesenteroides 1.3 × 10 6 L. oenos 6.3 × 106 L. brevis
C1 1.8 × 105 L. oenos L. brevis L. oenos
C2 8.6 × 10 4 L. oenos 7.9 × 10 6 L. oenos 7.6 × 105 L. brevis
LAB = Lactic acid bacteria
Source: Salih et al., 1990

5.10. Biochemical Changes During Aging


Production of aldehydes as one of the flavouring compounds takes place as a result of auto-oxidation of
polyphenolic compounds and oxidation of ethanol by direct chemical reaction with air (Wildenradt and
Singleton, 1974). Alcohols in wine react with organic acids, like tartaric, malic, succinic and lactic acid
to form their respective esters, which have been reported to increase with the aging of wine (Amerine et
al., 1980). The concentration of total volatile compounds also increases during fermentation as well as in
storage. Higher alcohols formation has been found to be closely related to the aroma and taste of wine.
During maturation, a decrease in the tannins due to their complexing with protein and polymerisation and
subsequent, precipitation takes place (Amerine et al., 1980).
596 Applied Aspects of Winemaking: Production of Wine

5.11. Final Treatment and Packaging


The desired product determines the final treatments that cider receives. Different batches of cider,
generally made from a mixture of different juices, are blended to give a specific flavour. To make cider
with no haze, it is desirable to treat the raw cider with fining agents organic or inorganic agents, such as
bentonite, gelatin or chitin, silica solution, albumen, casein, isinglass and tannin, and filtered. A number of
invisible components, such as polymeric carbohydrates, proteins and polyphenols are present in the wine.
If the negative charge in wine is altered by adding positively-charged particles, it becomes neutralised.
Particles combine and are flocculated or coagulated and clear wine can be recovered after removal of
these sediments. The use of cross-flow microfiltration systems has also been used to obviate the need
for fining and reducing the processing time and labour requirement (Jarvis, 1993). Cider can be sold as
a still or sparkling beverage with varying degrees of sweetness and clarity. The amount of carbonation
ranges from saturation for ciders in jars to 2-2.5 volumes of CO2 in most bottled ciders, and up to 5
volumes in Champagne cider. Carbonation pressure ranges between 2.5-3.5 bar, higher pressure is being
used in case of PET bottles (Jarvis, 1993). The sweetening may be from unfermented juice sugars, added
juice or concentrate, sweetening agents, depending upon the appropriate regulations. In terms of clarity,
ciders range from turbid farm cider to brilliantly clear ciders. Majority of commercial cider is filled into
kegs, bottles or cans. Keg cider is carbonated and pasteurised in-line and filled into stainless steel kegs
in a plant which rinses, washes and sterilises the kegs prior to filling. It can also be filled in glass bottles
that may be carbonated and pasteurised after filling. Common container closures are crown caps and
roll-on or plastic stoppers, which have replaced corks. It is then, pasteurised at 60°C for 20-30 min. or
preserved with SO2 as the best practical approach. Sulphiting the cider after finishing the alcoholic and
MLF is a solution which is sometimes employed in order to eliminate the bacteria that cause undesirable
alterations. However, in the elaboration of natural cider in the Basque Country, this practice is rarely used
(Dueñas et al., 2002). It is preferable to keep the addition of chemicals to a minimum in order to maintain
the sensory qualities of the final product.
Results indicate that UV light is effective for reducing pathogen, like E. coli in cider. However, with
the dosages used, additional reduction measures are necessary to achieve the required 5-log reduction
(Wright et al., 2000). The effect of pulsed electric field (PEF) in inactivating naturally occurring
microorganisms (yeast and molds) in freshly squeezed apple cider in a continuous flow system was
investigated by Liang et al. (2006) and reported that the microbial count decreased with an increase of
applied pulses (17.6-58.7 total) and treatment temperature (45-50°C), and a decrease of flow rate (3-10
l/h). At field strength of 27-33 kV/cm (3 mm electrode gap in a concentric chamber), 200 pulses/s, 3 l/h
flow rate, and 50°C process temperature, there was a 3.10 log reduction in microbial counts (Table 14).
Table 14. Effect of PEF Using a Continuous Flow System on Microbial Inactivation in Apple Cider

Treatment Field Flow Pulses Pulse Number of microorganisms


strength rate applied rate 45°C 50°C
(kV/cm) (l/h) (pulse/s)
log10N0 log10N0/N log10N0 log10N0/N
PEF 33 3 58.7 200 4.20±0.73 2.65±0.17a 6.67±0.18 3.1±0.10
6 29.3 200 7.52±0.11 1.90±0.17b 7.44±0.16 2.20±0.19
27 10 17.6 200 — — 4.52±0.16 1.12±0.12c
PEF + nisin- 27 10 17.6 200 — — 5.44±0.09 1.78±0.20b
lyso (1:3) (0.02±0.00)a
PEF +clove 27 10 17.6 200 — — 5.26±0.21 2.88±0.36a
oil, 3 ml/100 (0.06±0.02)
ml b

PEF + clove 27 10 17.6 200 — — 4.36±0.10 3.11±0.23a


oil, 5 ml/100 (0.07±0.03)
ml b

Source: Liang et al., 2006


Cider: The Production Technology 597

Aluminium-spotted caps are not satisfactory for bottling of cider due to excessive corrosion but
vinyl spotted caps are satisfactory for this purpose. Inert gases like CO2, N2 or their mixture can also
be used for storage of cider (Cant, 1960). Locally sold, still cider may be sold in plastic containers used
for or dispensed from a refrigerated bulk container. When refrigerated, cider remains stable for a week.
Carbonated cider is either sterile-filtered or flash-pasteurised before packaging. Various requirements to
be fulfilled for labelling bottled cider as select cider (Champagne) process are enumerated in Table 15.

Table 15. Requirements for Labelling Bottled Cider as Select Cider (Champagne) Process

• To be made from clean sound cider apples only or from a blend (including pears).
• No sweeteners other than cane sugar or beet can be added.
• No additions of concentrate or other fruit juice, foreign acids, artificial essences or artificial carbonation
be done.
• To the undiluted juice or battery diffusion juice may be added not more than 25 per cent of its own
volume of syrup made from pure cane or beet sugar. The original specific gravity of the pure juice and
cider to which the syrup was added must not be less than 1.040 at 15°C.
• No preservative or colouring matter prohibited by the Public Health Regulations be added.
• Acetic acid should not be discernible on the palate and volatile acidity not to exceed 0.15 per cent as
acetic acid.
• To be free of disorders.
• The last stages of fermentation must take place in the bottle and the deposit removed by disgorging

6. Quality of Cider
6.1. Chemical Composition of Cider
The most important compounds formed during fermentation which are considered key products effecting
sensory profile of cider are higher alcohols, esters, organic acids, carbonyl compounds, sugars and tannins
(Table 16). Except for extensive hydrolysis by pectolytic and cellulolytic enzymes, the composition of
fermented products, especially the flavour components, remained almost similar in the products obtained
by mechanical or mild enzymatic extraction process (Poll, 1993).

6.1.1. Ethyl Alcohol


Different types of ciders are classified according to their ethanol content, which varies from 0.05-13.6 per
cent (Amerine et al., 1980; Jarvis et al., 1995).

6.1.2. Acids
The acids are important in maintaining the pH low enough so as to inhibit the growth of many undesirable
bacteria. Like apple, must cider contains a variety of organic acids and their concentration depends on the
maturity and fermentation conditions (Beechand Carr, 1977; Labelle, 1980). Sweet cider could have less
than 0.45 gm acid. In dry ciders made by traditional method of fermentation, i.e. in which the apples are
not washed, have a high amount of volatile acidity (1 g/l) than the ciders made after washing and blending
of apples. In storage, the acetic acid is generally increased. In traditional method of fermentation, malic
acid in must is low (3-3.8 g/l) but in the must made by modern fermentation high concentration (4.8 g/l)
is observed because of acidic apples. The complete degradation of L-malic acid was carried out rapidly
by LAB in all the musts, except that made by modern methods of fermentation, where no MLF occurred.

6.1.3. Higher Alcohols


The formation of higher alcohol is an important criterion to determine the quality of the alcoholic
beverages but varies from strain to strain of yeast, cultivars used and fermentation conditions
employed (Amerine et al., 1980). The biosynthesis of higher alcohols is generally linked with amino
598 Applied Aspects of Winemaking: Production of Wine

Table 16. Analysis of 15 Commercial Ciders

Parameters Range Mean


TSS (oB) 4.60-7.40 6.8
Volatile acidity (g/100 ml) 0.060-0.105 0.926
Tannins (mg/L) 45-100 68
Esters (mg/100 ml) 17.60-28.72 21.76
Acidity (% malic acid) 0.40-0.69 0.55
Sweetness (%, w/v) 1.56-5.58 2.80
Alcohol (%, w/v) 3.2-6.6 4.71
Tannins (%, w/v) 0.028-0.17 10.10
Total sulphur dioxide ( ppm) 64-189 130
Total nitrogen (ppm) 18-63 42
Thiamin (μg/ml) all<0.005 <0.005
Nicotinic acid (μg/ml) 0.03-0.33 0.16
Pantothenate (μg/ml) 0.10-0.80 0.38
Riboflavin (μg/ml) 0.41-4.7 1.35
n-Propyl alcohol (ppm) 4-27 12
Isopropyl alcohol (ppm) 24-82 45
n-Butyl alcohol (ppm) 3-6 5
2- and 3-Methyl butyl alcohol (ppm) 113-176 150
n-Hexyl alcohol (ppm) 2-29 10
2-Phenethanol (ppm) 51-160 79
Magnesium (ppm) 8-41 27
Chloride (ppm) 33-146 112
Phosphate (ppm) 20-195 100
Sulphate (ppm) 120-380 227
Sodium (ppm) 30-275 123
Potassium (ppm) 415-1420 722
Iron (ppm) 0.95-6.73 3.7
Carbon (ppm) 0.10-1.05 0.42
Zinc (ppm) 0.21-1.77 0.56
         Source: Jarvis et al., 1995; Vyas and Kochhar, 1993

acid metabolism. Higher alcohols are formed as by-products of both anabolic (Genevois pathway) and
catabolic metabolism (Ehrlich pathway) and allow the re-equilibrium of the redox balance involving
NAD+/NADH cofactors (Hammond, 1986). Therefore, they may appear via biosynthesis route using the
amino acid biosynthesis pathway of the yeast or by the deamination and decarboxylation of amino acids
present in the substrate. It is also known that higher fusel alcohols are generated from cloudy rather than
clear juice fermentation (Beech, 1993). Table 17 shows the amounts of certain higher alcohols present
before and after fermentation in different varieties of apple.

6.1.4. Tannins
Tannins enhance the sensory qualities of wines by affecting their astringency level, which vary in cider
from 50-100 mg/100 ml (Azad et al., 1987). Polyphenols play an important role in the cider quality as
they are related to the colour, bitterness and astringency, whose balance defines the overall mouthfeel of
the beverage (Guyot et al., 1998; Alonso-Salces et al., 2001; Lea and Drilleau, 2003; Alonso-Salces, et
al., 2004). They may be involved in providing the cider aroma, and as inhibitors of the microorganism
development, controlling the fermentation rates and avoiding some faults that can develop in cider from the
Cider: The Production Technology 599

Table 17. The Major Higher Alcohols in Apple Juices and Ciders

Type of Content (ppm)


higher alcohols Yarlington Sweet Kingston Bramley’s
Mill Coppin Black seedling
Juice Cider Juice Cider Juice Cider Juice Cider
n-Propanol 0 52 0.6 46 0 34 3 44
n-Butanol 35 2 16 1 40 34 8 6
Isobutanol 0 6 1 6 0.5 6 2 25
n-Pentanol 0.05 0.01 0.2 tr. 0.5 0.3 0.1 0.1
2-and 3- 2 96 1 107 9 105 8 90
Methyl butanol
Hexanol 4 4 3 1 11 9 6 4
2-Hexanol 1 0.05 1 0.05 0.1 0.2 0.3 0.1
2-Phenethanol 0 19 0 34 0 30 0 19

action of lactic acid bacteria, such as acidification, mannitol taint, ‘framboisé’, bitterness (Alonso-Salces
et al., 2004). Furthermore, the phenolic compounds participate in the formation of sediments during cider
storage due to their colloidal interaction with the proteins through the van der Waals forces (Siebert et al.,
1996; Kawamoto and Nakatsubo, 1997). They can also inhibit the pre-fermentative clarification enzymes
(Cowan, 1999). The cider-making steps mainly responsible for the extraction and content of the phenols
in the final product are maceration, pressing, enzymatic clarification of the must prior to fermentation,
centrifugation, filtration and fining. During the maceration and pressing time, intensive oxidation of the
polyphenols takes place, due to the activity of the polyphenoloxidase (PPO) and the subsequent coupled
oxidation reactions with other polyphenols. In addition, a large proportion of the procyanidins from the
fruits remain in the pomace after the pressing step because of their adsorption onto the cell-wall matrix
(Renard et al., 2001; Guyot et al., 2003). These lead to musts with lower phenolic content (Siebert et al.,
1996). It has been proved that must oxidation was higher when it was in contact with the apple pulp. The
enzymatic clarification, centrifugation, filtration and fining of the French ciders lead to partial elimination
of procyanidins due to their ability to precipitate proteins and to interact with cell wall polysaccharides
(Alonso-Salce et al., 2004).
The type of polyphenols or tannins found in the bittersweet English cider are listed in Table 18.
No significant change is seen in the content during fermentation although the chlorogenic, caffeic
and p-coumaryl acids may be reduced to dihydroshikimic acid and ethyl catechol (Jarvis, 1993). The
chlorogenic and caffeic acid in apple juice cultivars and ciders correlated very well with total phenols.
The chlorogenic acid constitutes 6.2-10.7 per cent of total phenols and the involvement of these acids is
responsible for non-enzymic auto-oxidative browning reaction (Cilliers et al., 1989; Cilliers et al., 1990).

Table 18. Polyphenols in Bitter Sweet English Cider

Polyphenol type Conc. in cider (mg/100 ml)


Chlorogenic acid 98
Epicatechin 38
Dimericprocyanidin 79
Trimericprocyanidin 26
Tetramericprocyanidin 21
Source: Jarvis, 1993

6.1.5. Carbonyl Compounds


The most important carbonyl compounds formed in cider fermentation are acetaldehydes, diacetyl
and 2,3-pentanedione. Aldehydes, having very low flavour thresholds, tend to be considered as off-
600 Applied Aspects of Winemaking: Production of Wine

flavour (green leaf-like flavour). As intermediates in the formation of ethanol and higher alcohols from
amino acids and sugars, the conditions favouring alcohols production also generate the formation of
small quantities of aldehydes. They are excreted and then, reduced to ethanol during the later stage
of fermentation. Diacetyl makes an important contribution to the flavour of cider and its presence is
considered essential for correct flavour.
Aldehyde is a by-product of alcoholic fermentation and its low amounts are considered responsible
for the flavour and taste of wine. The yeast strains affect the formation of flavour compounds in wine
(Jarvis et al., 1995). Diacetyl and acetaldehyde may also be produced if the process is inhibited by excess
sulphite and/or if controlled lactic fermentation occurs. Methanol is also produced in small quantities (10-
100 ppm) as a result of demethylation of pectin juice (Jarvis, 1993).

6.1.6. Total Esters


Generally, esters are present in smaller concentrations than alcohols (Table 19), with the notable
exception of ethyl acetate and 2- and 3-methyl butyl acetates, which in Yarlington Mill juice, increase
200-fold during fermentation. Esters constitute a major group of desirable compounds. Among the
esters that can be formed, the most significant in fermented beverages are ethyl acetate (fruity), isoamyl
acetate (pear drops), isobutyl acetate (banana like), ethyl hexanoate (apple like), and 2-phenyl acetate
(honey, fruity, flowery). These esters are formed by yeast during fermentation in a reaction between
alcohols, fatty acids, co-enzyme A (COASH) and an esters synthesising enzyme (Nedovic et al., 2000).
The supplementation of amino acids was also found responsible for the production of esters in ciders.
The addition of aspartate, asparagine and glutamate positively influenced the production of esters in the
cider models. In addition, the combination of aspartate and glutamate predicted a higher production. The
optimal suggested concentrations were 43.4 per cent of aspartate and 56.6 per cent of glutamate for 120
mg/L of total nitrogen. The apple must supplemented with these two amino acids resulted in production
of four times more esters than in the same cider without supplementation (Eleutério dos Santos, 2016).

Table 19. Concentration of Major Esters in Apple Juices and Ciders of Different Varieties

Major esters Content (ppm)


Yarlington Mill Sweet Coppin Kingston Black Bramley’s seedling
Juice Cider Juice Cider Juice Cider Juice Cider
Ethyl acetate 2 35 1 20 1 17 2 15

Isobutyl acetate - 0.2 - 0.003 - 0.1 0.03 0.3


Ethyl butyrate 0.3 - 0.01 0.01 0.2 0.4 0.3 0.1
2- and 3- 0.15 30 0.02 3 0.2 4 0.1 0.9
Methyl butyl
acetates
Ethyl-methyl 0.006 - - - 0.01 - 0.04 -
butyrate
Hexyl acetate 0.6 6 0.3 0.1 0. 2 1.5 0.3 0.7
Ethyhexanoate - 2 - 0.02 - 0.6 - 0.9
2- and 3- - 4 - 0.1 - 0.7 - 0.01
Methyl butyl
octanoates

6.2. Sensory Qualities


Appearance, colour, aroma, taste and subtle taste factors, such as flavour constitute the quality of cider.
The aroma and taste are very complex and depend on a number of factors, such as varieties, agricultural
land, vinification practices, fermentation and maturation (Gayon, 1978). The taste of cider is determined
more by the apple composition whereas cider odour is governed by the technological factors and yeast
Cider: The Production Technology 601

employed than the apple varieties used to make cider. Cider with higher juice content was preferred to
that with lower juice content in various sensory quality characteristics (Joshi et al., 1991). The effect of
addition of insoluble solids, pectolytic enzyme and strains of wine yeast has been evaluated using various
descriptors (Joshi et al., 2013). Influence of cider-making technology on low boiling-point compounds
can be clearly seen in preparation of semi-sweet cider (Mangas et al., 1993; Mangas et al.,1993). A
comparison of the concentration of volatile compounds produced in cider made by batch and continuous
fermentation has also been made. The cider flavour is assessed using both by subjective and objective
approaches. In human beings, flavour sensation by taste is limited to sweetness, sourness, bitterness and
astringency together with such tastes as metallic and pungency (Piggott, 1988). Quantitative descriptive
analysis (QDA) has also been applied to cider to profile their flavour analytically (Williams, 1975). Out
of 86 descriptors used, 33 descriptors had greatest meaning to characterise the cider aroma and perry
essence. The development of a cider flavour wheel (Fig. 2) like that for wine and beer flavour profile is
employed in cider industry. In cider, bitterness and astringency are due to the polyphenols, especially
procyanidins which are polymers of catechins. The degree of polymerisation (DP) is the main factor
that influence the ratio between the two sensations: small procyanidins (up to DPn 4) are rather bitter
and higher (DPn 5-9) are rather astringent but both sensations are usually associated. There is often an
interaction with other constituents of the beverage: alcohol and polysaccharides reduce astringency, while
pH can increase it without changing the bitterness. Sugars are also known to reduce acidity and bitterness
(Quére et al., 2006).
At a simple level, a number of general descriptors can be used, such as fruity, acidity, sweetness,
astringency, alcohol, body, bitterness and sulphury, but at the analytical level, the number of descriptors

Figure 2. Cider flavour wheel (Source: National Association of Cider Maker, 1994)
602 Applied Aspects of Winemaking: Production of Wine

are kept large to differentiate the ciders of different types. Another approach which has been applied to
the flavour profiling is sniff analysis, where the effluent from GC is assessed by specially-trained judges.
Capillary gas chromatography (GC) on head-space samples of cider has been made to characterise the
aroma compounds (Fig. 3) (Jarvis et al., 1995). As many as 200 compounds reportedly contribute to the
flavour of ciders, but the key compounds are alcohols, acids, aldehydes, esters and sulphur compounds. The
spicy, aromatic and apple-like are the notes which differentiate the cider from other fermented beverages.

Figure 3. Chromatograph of a synthetic solution of low-boiling point components: 1. acetaldehyde; 2. ethyl


formate; 3. ethyl acetate; 4. ethyl propionate; 5. methanol; 6. ethanol; 7. propanol; 8. isobutanol; 9. butanol; 10.
4-methyl 2-pentanol (internal standard); 11. 2-methyl butanol; 12. 3-methyl butanol (Source: Mangas et al., 1993a)

The colour of cider is determined by the extent of juice oxidation or degradation and, in fact, it is
possible to make water-white high tannin ciders if oxidation is completely inhibited (Lea and Timberlake,
1978; Lea, 1982). During fermentation, however, the initial colour diminishes by around 50 per cent. It is
presumably because of the strong reductive power of yeasts, which readily reduces the keto or carbonyl
groups to hydroxyls with consequent loss of the chromophore (exposure to sterile air after fermentation
will slowly re-generate the colour).
Traditional English and French ciders made from bittersweet fruit have been distinguished by
relatively high levels of bitterness and astringency caused by the procyanidins (tannins). The cultivars and
juice-processing condition (notably oxidation) also play a part in determining the final non-volatile flavour.

6.3. Spoilage of Cider


Some ciders with residual sugar or the sweet cider with pH above 3.8 stored at ambient temperature
develops a defect called ropiness or oiliness. It is caused by certain strains of lactic acid bacteria
(Lactobacillus and Leuconostoc spp.) that produce a polymeric glucan (Carr, 1983, 1987), which
thickens its consistency and when poured, it appears oily in texture with a detectable sheen. At higher
concentration of glucan, the texture thickens so that the cider moves as a slimy ‘rope’ when poured
Cider: The Production Technology 603

from a bottle. Properly sulphited juice with a low pH can correct the defect (Beech and Carr, 1977). A
much simpler approach would be to pasteurise the affected juice. However, the treated cider would need
blending before use. Another defect is referred to as mousiness (Tucknott and Williams, 1973). The exact
cause of this defect is not known but it occurs in unsulphited cider with a high pH that has necessarily
been exposed to air during fermentation. The growth of film-forming yeasts, such as Brettanomyces
spp., Pichia membranefaciens, Candida mycoderma also produce ‘mousy’ flavour (1,4,5,6-tetrahydro-
2-acetopyridine). Sacchromycodes ludwigii is often resistant to SO2 levels (1000-1500 ppm). These can
grow slowly during fermentation and maturation, resulting in production of butyric flavour and formation
of flaky particles, which spoil the appearance of cider. Contamination of final product with Saccharomyces
cerevisiae, S. bailli and S. uvarum increases the concentration of CO2.
Ciders low in acidity, tannin and nitrogen but high in mineral matter occasionally develop an olive
green colour; the fermentation ceases and starch is deposited. If iron in the ciders combines with tannins,
a black or greenish black colour develops. Bottled cider stored at high temperature sometimes produces
a sediment called casse. The action of peroxidase on tannins causes this defect, which can be prevented
by the addition of SO2 after fermentation.
The classical microbiological disorder of stored bulk ciders is known as ‘cider sickness’ or
‘Framboise’ in French (Beech and Carr, 1977; Carr, 1987) which is caused by the bacterium Zymomonas
anaerobia which ferments sugar in bulk sweet ciders stored at pH values greater than 3.7. The features of
cider sickness are a renewed and ‘almost explosive fermentation’ accompanied by a raspberry or banana-
peel aroma and a dense white turbidity in the beverage due to production of acetaldehyde at high levels
by Zymomonas. The acetaldehyde reacts with the ‘tannins’ to produce an insoluble aldehyde-phenol
complex and consequently, turbidity.
Flavour taints in ciders may arise from the presence of naphthalene and related hydrocarbons where
tarred rope had been stored adjacent to a cider keg. A new taint in ciders is caused by indole and is derived
from tryptophan breakdown (Wilkins, 1975) at levels in excess of 200 ppb where its odour becomes
increasingly faecal and unpleasant.

6.4. Waste from Cider Industry


Wastes and wastewater generated during the cider-making process were identified as potential sources
to obtain value added products, such as ethanol, via alcoholic fermentation mediated by yeast. This
wastewater comprises the purges from the fermentation process, cider losses during transfers, products
discarded due to quality policies and products that have returned from the market past the expiration date.
The wastewater also exhibit a high Chemical Oxygen Demand (COD), with values greater than 170,000
mg O2/l, due to its elevated sugar and ethanol content, and usually represents approximately 10 per cent
of the volume of cider produced. Therefore, it must be treated prior to discharge into the environment.

References
Alonso-Salces, R.M., Barranco, A., Abad, B., Berrueta, L.A., Gallo, B. and Vicente, F. (2004).
Polyphenolic profiles of basque cider apple cultivars and their technological properties. Journal of
Agricultural and Food Chemistry 52: 2938-2952.
Alonso-Salces, R.M., Korta, E., Barranco, A., Berrueta, L.A., Gallo, B. and Vicente, F. (2001).
Determination of polyphenolic profiles of Basque cider apple varieties using accelerated solvent
extraction. Journal of Agricultural and Food Chemistry 49: 3761-3767.
Amerine, M.A., Kunkee, R.E., Ough, C.S., Singleton, V.L. and Webb, A.D. (1980). Technology of Wine
Making. AVI Publ. Co., Westport, Connecticut.
Azad, K.C., Vyas, K.K., Joshi, V.K. and Srivastava, M.P. (1986). Deacidification of fruit juices for
alcoholic fermentation. Abst. ICOFOST-86, 53.
Azad, K.C., Vyas, K.K. Joshi, V.K. and Sharma, R.P. (1987). Observations on juice and cider made from
scabbed apple fruit. Indian Food Packer 41(1): 56-61.
604 Applied Aspects of Winemaking: Production of Wine

Babsky, N.E., Toribio, J.L. and Lozano, J.E. (1986). Influence of storage on the composition of clarified
apple juice concentrate. J. Food Sci. 51: 564.
Beech, F.W. (1972). Quick determination of adequate juice sulphiting. J. Inst. Brew. 78: 477.
Beech, F.W. and Carr, J.G. (1977). Cider and perry. p. 139. In: A.H. Rose (Ed.). Economic Microbiology,
vol. VI: Alcoholic Beverages. Academic Press, London.
Beech, F.W. (1993). Yeasts in cidermaking. p. 169. In: A.H. Rose and J.S. Harrison (Eds.). The Yeasts,
second ed., vol. 5: Yeast Technology. Academic Press, London.
Beveridge, T., Franz, K. and Harrison, J.E. (1986). Clarified natural apple juice: Product storage stability
of juice and concentrate. J. Food. Sci. 51: 411.
Bump, V.L. (1989). Apple processing and juice extraction. pp. 53-82. In: Donald, L. Downing (Ed.).
Processed Apple Products. AVI Publishing Co., New York.
Bureau of Alcohol, Tobacco and Firearms (1998). Implementation of Public Law 105-34, Sections 908,
910 and 1415, related to hard cider, semi-generic wine designations, and wholesale liquor dealers’
signs (97-2523). US Department of the Treasury. Federal Register 63(162): 44819-44820, http://
www.atf.gov/regulations-rulings/rulemakings/notices/notice-859.html
Burroughs, L.F. and Pollard, J.P. (1954). Annual Report of the Agricultural and Horticultural Research
Station. Long Ashton, 1953, p. 184, University of Bristol, England.
Burroughs, L.F. (1973). Report. Long Ashton, 1972, p. 124, University of Bristol, England.
Campo, G.D., Berregi, I., Santos, J.I., Duenas, M. and Gud Irastorza, A. (2008). Development of alcoholic
and malo-lactic fermentations in highly acidic and phenolic apple musts. Bioresource Technology
99: 2857-2863.
Cant, R.R. (1960). The effect of nitrogen and carbon dioxide treatment of wine on dissolved oxygen
levels. Am. J. Enol. Vitic. 11: 164.
Carr, J.G. (1983). Microbes I have known. J. Appl. Bacteriol. 55: 383.
Carr, J.G. (1987). Microbiology of wines and ciders. p. 291. In: J.R. Norris and G.L. Pettipher (Eds.).
Essays in Agricultural and Food Microbiology. John Wiley, London.
Castelli, T. (1973). Lecologie Des Levures. Collogue Inter. D’s Conolgie Arcensenas, Vignasvins Mai
19: 25.
Cilliers, J.J.L., Singleton, V.L. and Lamuela-Raventos (1990). Total polyphenols in apples and ciders:
Correlation with chlorogenic acid. J. of Food Sci. 55(5):1458-1459.
Cilliers, J.J.L. and Singleton, V.L. (1991). Non-enzymic autoxidative phenolic browning reactions in a
caffeic acid model system. J. Agric. Food Chem. 37: 1298-1303.
Cowan, M.M (1999). Plant products as antimicrobial agents. Clinical Microbiology Reviews 12: 564-582.
Dinsdale, M.G., Lloyd, D. and Jarvis, B. (1994). Membrane potential studies of Saccharomyces cerevisiae
during cider fermentation. In: Biochemical Society Transactions. 650th Meeting, Cardiff, p. 325.
Downing, Donald, L. (1989). Processed Apple Products. AVI Publishing Company, New York.
Dueñas, M., Irastorza, A., Fernandez, A.B. and Huerta, A. (1994). Microbial populations and malo-lactic
fermentation of apple cider using traditional and modified methods. J. Food Sci. 59(5): 1060-1064.
Dueñas, M., Irastorza, A., Munduate, A., Santos, J.I., Berregi, I. and del Campo, G. (2002). Influence of
enzymatic clarification with a pectin methylesterase on cider fermentation. J. Inst. Brew. 108: 243-
247.
Eleutério dos Santos, M.C., Alberti, A., Arruda Moura Pietrowski, G. de, Ferreira Zielinski, A.A.,
Wosiacki, G., Nogueira, Alessandro and Matos Jorge, R.M. (2016). Supplementation of amino acids
in apple must for the standardisation of volatile compounds in ciders. J. Inst. Brew, DOI 10.1002/
jib.318
FAO (2018). Production Year Book. Food and Agriculture Organisation, Rome, www.fao.org
Frazier, W.C. and Westhoff, D.C. (1995). Food Microbiology. Tata McGraw-Hill Publishing Co. Ltd,
New Delhi.
Garcia, Y.D., Valles, B.S. and Lobo, A.P. (2009). Phenolic and antioxidant composition of by-products
from the cider industry: Apple pomace. Food Chemistry 117: 731-738.
Gary, L.M., Renee, T.T. and Justin, R.M. (2007). Reduction of malic acid in wine using natural and
genetically enhanced microorganisms. Am. J. Enol. Vitic. 58: 341-345.
Cider: The Production Technology 605

Gayon, P.R. (1978).Wine flavour. p. 335. In: G. Charlambous and G.E. Inglett (Eds.). Flavour of Food
and Beverage Chemistry and Technology. Academic Press, Inc, New York, London.
Guillermin, P., Dupont, N., LeMorvan, C., Le Quéré, J.-M., Langlais, C. and Mauget, J.C. (2006).
Rheological and technological properties of two cider apple cultivars. LWT 39: 995-1000.
Gullon, B., Yanez, R., Alonso, J.L. and Parajo, J.C. (2008). L-Lactic acid production from apple pomace
by sequential hydrolysis and fermention. Bioresource Technology 99: 308-319.
Guyot, S., Marnet, N., Laraba, D., Sanoner, P. and Drilleau, J.F. (1998). Reversed-phase HPLC following
thiolysis for quantitative estimation and characterisation of the four main classes of phenolic
compounds in different tissue zones of a French cider apple variety. J. Agric. Food Chem., 46: 1698-
1705.
Guyot, S., Marnet, N., Sanoner, P., Drilleau, J.-F. (2003). Variability of the polyphenolic composition of
cider apple (Malus domestica) fruits and juices. Journal of Agricultural and Food Chemistry 51:
6240-6247.
Gebhardt, S.E., Cutrufelli, R. and Mathews, R.H. (1982). Composition of Foods. Agric. Handbook 8-9,
US Dept of Agri., Washington.
Hang, Y.D. and Woodams, E.E. (2010). Influence of apple cultivar and juice pasteurisation on hard cider
and ean-de-vie methanol content. Bioresource Technology 101: 1396-1398.
Hammond, J. (1986). The contribution of yeast to beer flavor. Brew. Guardian 115: 27-33.
Herrero, M., Laca, A., Garcia, L.A. and Diaz, M. (2001). Controlled malo-lactic fermentation in cider
using Oenococcus oeni immobilised in alginate beads and comparison with free cell fermentation.
Enzy. Microb.Technol. 28: 35-41.
Jarvis, B. (1993). Chemistry and microbiology of cidermaking. In: Encyclopedia: Food Science and
Nutrition, first ed. Coleraine Campus, Cromore Road, Coleraine, BT52ISA, Belfast Coleraine
Jordans town, Magee.
Jarvis, B. (1993). Cider: Hard cider. In: Encyclopedia: Food and Nutrition, first ed., Coleraine Campus,
Cromore Road, Coleraine, BT52ISA, Belfast Coleraine Jordans town, Magee.
Jarvis, B., Foster, M.J. and Kinsella, W.P. (1995). Factors affecting the development of cider flavour. J.
Appl. Bacteriol. Symp. supp., 79: 55.
Joshi, V.K. and Bhutani, V.P. (1991). The influence of enzymatic clarification on fermentation behaviour
and qualities of apple wine. Sci. Des Aliments. 11(3): 491-496.
Joshi, V.K., Sandhu, D.K., Attri, B.L. and Walia, R.K. (1991). Cider preparation from apple juice
concentrate and its consumer acceptability. Indian J. Hort. 48(4): 321-327.
Joshi, V.K. and Sandhu, D.K. (1994). Influence of juice contents on quality of apple wine prepared from
apple juice concentrate. Res. Ind. 39(4): 250-252.
Joshi, V.K. and Sandhu, D.K. (1997). Effect of different concentrations of initial soluble solids on
physicochemical and sensory qualities of apple wine. Indian J. Hort. 54(2): 116-123.
Joshi, V.K., Sandhu, D.K. and Thakur, N.S. (1999). Fruit-based alcoholic beverages. pp. 647-744. In: V.K.
Joshi and Ashok Pandey (Eds.). Biotechnology: Food Fermentation, vol. II. Educational Publishers
and Distributors, New Delhi.
Joshi, V.K. and Siby John (2002). Antimicrobial activity of apple wine against some pathogenic and
microbes of public health significance. Alimentaria November: 31(2): 67-69.
Joshi, V.K., Sandhu, D.K. and Kumar Vikas (2013). Influence of addition of insoluble solids, wine yeast
strains and pectinolytic enzymes on the flavour profile of apple wine. International Journal of Food
and Fermentation Technology 3(1): 57-66.
Kawamoto, H. and Nakatsubo, F. (1997). Effects of environmental factors in two-stage tannin-protein
co-precipitation. Phytochemistry 46: 379-483.
Kerni, P.N. and Shant, P.S. (1984). Commercial Kashmir apple for quality cider. Indian Food Packer
38(1): 78.
Labelle, R.L. (1979). The many faces of (hard) cider. N.Y. State Agric. Exp. Stn. Spec. Rep. 32: 1.
Labelle, R.L. (1980). Apple cultivars tested as naturally fermented cider at Geneva. Memo, N.Y. State
Agric. Exp. Stn. Geneva, New York.
Lea, A.G.H. (1978). Phenolics of cider – Procyanidins. J. Sci. Food Agric. 29: 484-492.
606 Applied Aspects of Winemaking: Production of Wine

Lea, A.G.H. (1982). Analysis of phenolics in oxidising apple juice by HPLC using a pH shift method. J.
Chromatog. 238: 253.
Lea, A.G.H. (1984). Colour and tannins in English cider apples. Flussiges Obst. 51: 356.
Lea, A.G.H. (1995). Cidermaking. p. 66. In: A.G.H. Lee and J.R. Piggott (Eds.). Fermented Beverage
Production. Blackie Academic and Professional, London, UK.
Lea, A.G.H. and Timberlake, C.F. (1974). Phenolics of Ciders – The effect of processing. J. Sci. Fd.
Agric. 25: 1537-1545.
Lea, A. and Drilleau, J.-F. (2003). Cidermaking. pp. 59-87. In: Lea, A. (Ed.). Fermented Beverage
Production. London.
Liang Ziwei, Cheng, Z. and Mittal, G.S. (2006). Inactivation of spoilage microorganisms in apple cider
using a continuous flow-pulsed electric field system. LWT 39: 350-356.
Lowe, E., Durkee, E.L., Hamilton, W.E. and Moyan, A.I. (1964). Bitter apple juice dejuicing through
thick cake extraction. Food Eng. 36(12): 48-50.
Mahant, K., Sharma, S., Sharma, S. and Thakur, A.D. (2017). Effect of nitrogen source and citric acid
addition on wine preparation from Japanese persimmon. Journal of the Institute of Brewing 123(1):
144-150.
Mangas, J., Paz-Gonzallez, M. and Blanco, D. (1993a). Influence of cider-making technology on low-
boiling point volatile compounds. Zest Schrift fur labensmittelutersuchung Forschung, 197(6): 522-
524.
Mangas, J.J., Moreno, J., Cabranes, C., Dapana, E. and Blanco, D. (1993b). Study of semi-sweet cider.
Alimentaria 243: 85.
Mitra, S.K. (1991). Apples. p. 122. In: S.K. Mitra, T.K. Bose and D.S. Rathore (Eds.). Temperate Fruits.
Horticulture and Applied Publ., Calcutta.
Moulton, G.A., Miles, C.A., King and Zimmerman, J.A. (2010). Hard Cider Production & Orchard
Management in the Pacific Northwest. A Pacific Northwest Extension Publication. Washington State
University Extension, Oregon State University Extension Service, University of Idaho Cooperative
Extension System and the US Department of Agriculture Co-operative.
Nedovic, V.A., Durieux, A., Van Nedervelde, L., Rossels, P., Vandegans, J., Plainsant, A.M. and Simon,
J.F. (2000). Continuous cider fermentation by co-immobilised yeast and Leuconostoc oenos cells.
Enz. Micro. Technol. 26: 834.
Nicolini, G., Román, T., Carlin, S., Malacarne, M., Nardin, T., Bertoldi, D. and Larcher, R. (2017).
Characterisation of single variety still ciders produced with dessert apples in the Italian Alps. J. Inst.
Brew., DOI 10.1002/jib.510
Nogueira, A., Guyot, S., Marnet, N., Lequere, J.M., Drilleav, J.F. and Wosiacki, G. (2008). Effect of
alcoholic fermentation in the content of phenolic compound in cider processing. Brazilian Archives
of Biology and Technology 51(5): 1025-1032.
O’ Reilly, A. and Scott, J.A. (1993). Use of an ion-exchange sponge to immobilise yeast in high gravity
apple-based cider alcoholic fermentation. Biotechnol.Letters 15(10): 1061-1069.
Park, Y.J. and John, C.B. (1980). Decomposition of acid in wine by yeast. Res. Rep. Agric. Sci. Technol.
Chungnam Nat, Univ., Daejeon, S. Korea, 7(2): 176.
Picinelli, A., Suarez, B., Moreno, J., Rodrıguez, R., Caso-Garcıa, L.M. and Mangas, J.J. (2000). Chemical
characterisation of Asturian cider. Journal of Agricultural and Food Chemistry 48: 3997-4002.
Piggott, J.R. (1988). Sensory Analysis of Foods, second ed. Elsevier Applied Science, London, New
York.
Pliny (the Elder), Reissue (1967). Natural History. Loeb Parallel text ed., Heinemann, London.
Poll, L. (1993). The effect of pulp holding time and pectolytic enzyme treatment on the acid content of
apple juice. Food Chem. 47(1): 73-75.
Pourlx, A. and Nicholas, L. (1980). Sweet and Hard Cider. Gardenway Publishing Co., Charlotte, V.T.
Quéré Jean-Michel Le, Husson Franc-ois, Catherine, Renard, M.G.C. and Primault, Jo (2006). French
cider characterisation by sensory, technological and chemical evaluations. LWT, 39: 1033-1044.
Rana, R.S., Vyas, K.K. and Joshi, V.K. (1986). Studies on production and acceptability of cider from
Himachal Pradesh apples. Indian Food Packer 40(6): 48-56.
Cider: The Production Technology 607

Renard, C.M.G.C., Baron, A., Guyot, S. and Drilleau, J.F. (2001). Interactions between apple cell walls
and native apple polyphenols: Quantification and some consequences. International Journal of
Biological Macromolecules 29: 115-125.
Revier, M. (Ed.) (1985). Le cidre-heiretaujourd’hui. La Nouvelle Libraire, Paris.
Reedy, D., Mcclactchey, W.C., Smith, C., Lan, Y.H. and Bridges, K.W. (2009). A monthful of diversity:
Knowledge of cider apple cultivars in the United Kingdom and northwest United States. Economic
Botany 63(1): 2-15.
Ribereau-Gayon, J., Penaud, E., Ribereau-Gayon, P. and Sudraud (1975). Traited’oenologie. Sciences
ettechniquies du vin, vol. 2, Dunod, Paris.
Roach, F.A. (1985). Cultivated Fruits in Britain – Their Origin and History. Basil Blackwell, Oxford.
Salih, A.G., Le Quere, J.M., Drilleau, J.F. and Fernandez, J.M. (1990). Lactic acid bacteria and malo-
lactic fermentation in manufacture of Spanish cidermaking. J. Inst. Brew. 96: 369-372.
Seluy, L.G., Comelli, R.N., Benzzo, M.T. and Isla, M.A. (2018). Feasibility of bioethanol production
from cider waste. J. Microbiol. Biotechnol. 28(9): 1493-1501.
Sharma, R.C. and Joshi, V.K. (2005). Apple processing technology. pp. 445-498. In: K.L. Chadha and
R.P. Awasthi (Eds.). The Apple. Malhotra Publish. House, New Delhi.
Sharma, S., Thakur, A.D., Sharma, S. and Attenasova, M. (2018). Effect of different yeast species on the
production of pumpkin wine making. Journal of the Institute of Brewing 124(2): 187-193.
Siebert, K.J., Carrasco, A. and Lynn, P.Y. (1996). Formation of protein-polyphenol haze in beverages.
Journal of Agricultural and Food Chemistry 44: 1997-2005.
Smock, R.M. and Neubert, A.M. (1950). Apple and Apple Products. Interscience Publishers, New York.
Sonni, F., Cejudo, B., Maria, J., Chinnici, F., Natali, Nadia and Riponi Claudio (2009). Replacement of
sulphur dioxide by lysozyme and oenological tannins during fermentation: Influence on volatile
composition of white wines. Science Food Agri. 89(4): 688-696.
Tucknott, O.G. and Williams A.A. (1973). Report. Long Ashton, 1972, p. 150, University of Bristol,
England.
Unwin, T. (1991). Wine and the Vine: An Historical Geography of Viticulture and the Wine Trading.
Routledge, London.
Upshaw, S.C., Lopez, A. and Williams, H.L. (1978). Essential elements in apples and canned apple sauce.
J. Fd. Sci. 43(2): 449-456.
Vyas, K.K. and Kochhar, A.P.S. (1993). Studies on cider and wine from culled apple fruit available in
Himachal Pradesh. Indian Food Packer 47(4): 15. 63-69.
Whiting, G.C. (1961). Annual Report of the Agricultural and Horticultural Research Station. Long
Ashton, 1960, p. 135, University of Bristol, England.
Wildenradt, H.L. and Singleton, V.L. (1974). The production of aldehydes as a result of oxidation of
polyphenolic compounds and its relation to wine aging. Am. J. Enol. Vitic. 25: 119-126.
Wenlai, F., Yan, Xu and Yu, Aimei (2006). Influence of oak chips geographical origin, toast level, dosage
and aging time on volatile compounds of apple cider. J. Inst. Brew. 112(3): 255-263.
Wilkins, C.K. (1990). Analysis of indole and skatol in porcine gut contents. Inter. J. Fd. Sci. Technol. 25:
313-317.
Williams, A.A. (1975). The development of vocabulary and profile and profile assessment method for
evaluating the flavour contribution of cider and perry aroma constituents. J. Sci. Fd. Agric., 26:
567-582.
Wong, M. and Stanton, D.W. (1993). Effect of removal of amino acid and phenolic compounds on non
enzymatic browning of stored kiwifruit juice concentrate. Lebensmittel-Wissenching und technologie
26: 138.
Wright, J.R., Summer, S.S., Hackney, C.R., Pierson, M.D. and Zoechlein, B.W. (2000). Efficacy of
ultraviolet light for reducing Escherichia coli O157:H7 in unpasteurised apple cider. Food Protection
63: 563-567.
Wrolstad, R.E., Spanos, G.A. and Durst, R.W. (1990). Changes in phenolics and amino acid profiles of
apple juice concentration during processing and storage. Berichte Intenational Fruchtsaft-union,
Wissenchaftlich-Technische Kommission 103. www.google.com. www.nhb.org (Hard Honey Cider).
23 Brandy Production: Fundamentals and
Recent Developments
Francisco López
Departament d’Enginyeria Química, Facultat d’Enologia, Universitat Rovira i Virgili Av.


Països Catalans 26, 43007 Tarragona

1. Introduction
Brandy is a spirit obtained through the distillation of wine and generally contains 35-60 per cent v/v of
ethanol. If the name of the brandy is not associated with the type of raw material originating from this
spirit (fruit brandy, grain brandy, pomace brandy, etc.), it is understood that it is made exclusively from
grape wine. The origin of the word ‘brandy’ comes from the Dutch brandewijn, whose meaning is ‘wine
burned’. Some brandies are aged in wooden barrels, some are coloured with caramel to emulate the effect
of aging and/or homogenise the final product, and some brandies are produced using a combination of
aging and colouration (Amerine et al., 1989; Christoph and Bauer-Christoph, 2007).
Different types of brandy are made all over the world from wine. The best known are produced in
France under the appellation of Cognac and Armagnac. In other countries, different types of brandies are
made from wine, for example, brandy of Jerez and brandy of Penedès in Spain, Italian brandy produced
from regional wine grapes and distilled by column stills, although there are also a number of low-scale
producers, which employ pot stills. German brandy, which is called weinbrand (burnt wine), is made
usually from imported wine. The most known South American brandy is Pisco. In Peru, it is made mainly
from Muscat grapes. In Chile, it is made from different grape varieties and is distilled in pot stills. It
is obvious that, worldwide, there are various legal definitions according to the national traditions and
commercial interests (Tsakiris et al., 2014).
The use of the word ‘brandy’ to define a spirit obtained from distillation of wine leads to confusion,
since it is sometimes used as a generic of a product and sometimes to define a specific type of distillate.
European legislation distinguishes wine spirits and brandies (European Union 2008). Among the first are
Cognac and Armagnac. Table 1 shows the characteristics of these two types of spirit drinks.

Table 1. Characteristics of Wine Spirit and Brandy According to European Legislation

Spirit drink Wine spirit Brandy or weinbrand


Maximum distillation alcoholic strength 86% v/v 94.8% v/v
Minimum alcoholic strength 37.5% v/v 36% v/v
Maximum commercial alcoholic strength Not indicated 50% v/v
Maturation time Not indicated, but if it is 1 year oak receptacle
matured as a brandy can be 6 months oak casks <1000 L
labelled as ‘wine spirit’
Volatile substances >125 g/hL p.a. >125 g/hL p.a.
Methanol content <200 g/hL p.a. <200 g/hL p.a.
Flavouring Not, except traditional Not, except traditional
production methods production methods
Caramel addition Only to adapt colour Only to adapt colour
Addition of alcohol No No

E-mail: francisco.lopez@urv.cat
Brandy Production: Fundamentals and Recent Developments 609

In United States, brandies are defined according to the US Code of Federal Regulations, Title 27:
Alcohol, Tobacco and Firearms PART 5, Sub-part C, Standards of Identity for Distilled Spirits, as class 4:
‘Brandy’ is an alcoholic distillate made from fermented juice, mash, or wine of fruit, or from the residue
thereof, produced at less than 190°C, in such a manner that the distillate possesses the taste, aroma and
characteristics generally attributed to the product, and bottled at not less than 80°C. Brandy, or mixtures
thereof, not conforming to any of the standards in paragraphs (d) (1) through (8) of this section shall be
designated as ‘brandy’, and such designation shall be immediately followed by a truthful and adequate
statement of composition. The fruit brandy, derived from grapes, shall be designated as ‘grape brandy’
or ‘brandy’, except that in the case of brandy (other than neutral brandy, pomace brandy, marc brandy
or grappa brandy) distilled from the fermented juice, mash, or wine of grapes, or the residue thereof,
which has been stored in oak containers for less than two years, the statement of class and type shall be
immediately preceded, in the same size and kind of type, by the word ‘immature’. Fruit brandy, other
than grape brandy, derived from one variety of fruit, shall be designated by the word ‘brandy’ qualified
by the name of such fruit (for example, ‘peach brandy’), except that ‘apple brandy’ may be designated
‘applejack’. Fruit brandy derived from more than one variety of fruit shall be designated as a ‘fruit
brandy’ qualified by a truthful and adequate statement of composition.

2. History of Brandy and Economical Aspects


2.1. Some Historical Aspects
Distillation was a technique already used by ancient cultures in China (3000 years BC), India (2500 years
BC), Egypt (2000 BC), Greece (1000 years BC) and Rome (200 BC). Initially, all these cultures produced
a distilled liquid, later called alcohol by the Arabs, for the preparation of medicines and perfumes. In the
7th century, the Arabs invaded Europe, introducing the technique of distillation there. Later, in Christian
Europe, the doctor and theologian Arnau de Vilanova (Valencia, 1238?-1311) published the book Liber
Aqua Vitae, a treatise on wines and spirits, which was a manual of the period for the production of
distillates (Lopez et al., 2017).
Until the end of the 15th century, distilled wine (aqua vitae) was a product used as a medicine
and doctors and apothecaries controlled the distillation technique. From the authorisation to distil to
vinegarmakers by Louis XII and later to the winegrowers by Francisco I, in France, its elaboration
like product of consumption in the 16th century increased. In the wine trade with Holland, to reduce
costs, the wine was transported as distillate, which was then reconstituted with water. However, this
wine concentrate was consumed as such, being called ‘Brandewijn’, whose name evolved to brandy.
Although, the Dutch had provided much of the original capital together with the stills and the technology
necessary for distillation, local producers had also invested in the production of brandy and they were,
therefore, forced to seek new markets for their products. Elsewhere throughout France, but particularly
in Armagnac, other wine producers had also begun to distill their wines, especially those that were of
poor quality. However, brandy gradually came to be made where the raw material was available. Later,
it spread to southern wine-growing lands, such as Andalusia, Catalonia and Languedoc (Dhiman and
Attri, 2011).

2.2. Economical Aspects of Brandy


The Wiseguy report (2018) indicates that despite the slowdown in global economic growth, the spirits
industry has also suffered a certain impact, but still maintained a relatively optimistic growth in the past
four years. Spirits market maintain the average annual growth rate of 1.86 per cent from 218,200 million
$ in 2014 to 230,600 million $ in 2017, the analysts believe that in the next few years, spirits’ market
size will be further expanded. They expect that by 2022, the market size of the spirits will reach 239,200
million $. Figure 1 shows the brand values of spirits worldwide in 2017 (Statista, 2018), with 8.0 per cent
of spirits brand value.
The top markets for brandy are India, Philippines, Russia, Brazil and Germany. Grape-based
brandy-producing countries have a significant-sized wine industry that provides the base wine for brandy
production. Brandy production and consumption tends to be highly regionalised, with the exception of
610 Applied Aspects of Winemaking: Production of Wine

Figure 1. Brand values share of spirits worldwide in 2017, by spirit type (Statista, 2018)

Cognac, that is known and consumed globally. The production methods thus differ significantly (Bougas,
2014; Lambrecths et al., 2016).
However, in recent years the trend towards the consumption of artisanal distillates has increased and
the consequent growth in the number of distilleries, due to a greater perception of new consumers of these
artisanal products which are of a higher quality (Kohlmann, 2016).

3. Techniques Used in Brandy Elaboration


3.1. Wine Characteristics
For the production of one litre of brandy, around 4.5 litres of wine are needed and during the distillation,
the aromas of the wine are concentrated. These aromas are produced mostly during fermentation, which
is why grape varieties with a lower aromatic potential and higher productivity are usually used to make
the wines that are then distilled (Tsakiris et al., 2014). The characteristics of the wines elaborated for their
distillation must minimise the presence of sulphur dioxide and be distilled as soon as possible. The presence
of phenolic compounds should be minimised; therefore, continuous presses or the use of pectolytic
enzymes is not recommended, as they can increase the presence of methanol (Tsakiris et al., 2016). Wines
with lees affect the aromatic profile and their complexity; so sometimes in different types of brandies or
wine distillates, the regulation authorises its presence in wine. Another important aspect is the preservation
of wines before fermentation, since they can pass periods of up to five months between obtaining the
wine (case of Cognac and Armagnac), is to control the presence of bacteria that can impart strange
aromas to wine, which can be increased in the distillate (Amerine et al., 1989; Tsakiris et al., 2014).

3.2. Distillation Systems


For the distillation of wine, different distillation systems are used throughout the world. These systems
range from the simple alembic (batch distillation system) to continuous distillation systems with
distillation columns. Usually batch distillation produces more complex products, since during the
distillation, the distillate varies in a wide range of ethanol concentrations, dragging more variety of minor
compounds to the final distillate, while the continuous distillations are carried out at a fixed and higher
alcoholic strength, so that the final distillate has fewer congeners.
The traditional alembic (Fig. 2), which is formed by a kettle with deposits of the wine to distill.
Over the kettle is placed the hat, in which a small rectification of distillate is produced, which allows
obtaining products with an alcoholic strength higher than that expected by an ideal and simple differential
distillation. The hat is connected to the total condenser by the swan neck, in which the distillate is still
rectified gently. Before the distillate passes to the total condenser, it can go through a preheater of wines,
so that part of the energy supplied in the kettle is used to preheat the wine for a second batch of distillation.
Brandy Production: Fundamentals and Recent Developments 611

Figure 2. Schema of traditional alembic (Charantais type) (Adapted from Bureau National Interprofessionel de
Cognac, http://www.cognac.fr/cognac/_fr/2_cognac/index.aspx?page=distillation)

In the total condenser, the vapours of the distillate are finally condensed and cooled to a low temperature
to collect finally the distillate.
The hat shape can affect the characteristics of the distillate, since the rectification produced will
depend on it surface to generate more or less reflux inside the equipment. Hats with a large surface area
will produce less aromatic distillates and contain fewer compounds with more body, such as the longer
chain fatty acids (Bougass et al., 2014). For example, the hats used in making cognacs are onion-shaped
and therefore small; so these products are more complex and aromatic. Another effect of the lower reflux
generated with small hats is that it minimises the re-concentration of water in the kettle, causing the
concentration of ethanol in the waste to be higher and, therefore, the recovery will be lower and the final
product will have a minor alcoholic strength.
The distillation time will also affect the characteristics of the distillate, since during boiling, it can
produce different reactions of formation or degradation of compounds (Cantagrel, 2003; Bougass et al.,
2014), such as equations 1.1-1.4.
Esterification/Hydrolysis: Acid + Alcohol ßà Ester + Water [1.1]
Acetal formation: Aldehyde + Alcohol ßà Acetal + Water [1.2]
Maillard’s reaction: Sugar + Amino acid à Pyrazine, Furans [1.3]
Strecker degradation: a-Amino acid à Aldehydes àAcetals [1.4]
The operation of batch distillation systems consists in the separation of a first fraction, called ‘heads’,
in which negative volatile compounds, such as ethyl acetate and acetaldehyde, are concentrated. So this
fraction is rejected. Next, a second fraction, called ‘heart’, is collected in which most of the positive
compounds in the distillate are concentrated. Finally, a third fraction, called ‘tails’, is collected, in which
a significant amount of alcohol remains, as well as less compounds, which in general are organoleptically
negative in the final distillation. This fraction and the head fraction are usually reintroduced in subsequent
distillations in order to recover the maximum possible ethanol. The residue left in the kettle is rejected.
With this discontinuous system, the distillate obtained has a variable alcoholic strength between
the beginning and the end that favours the complexity of the final distillate, as previously mentioned.
This discontinuous system is applied to beverages, such as cognac and Peruvian Pisco, using different
strategies.
612 Applied Aspects of Winemaking: Production of Wine

The other classic alternative is the column distillation, operating discontinuous or continuous, as in
the case of Armagnac. Figure 3 presents a scheme of this distillation system.
The distillation strategy operating in discontinuous is similar to the traditional alembic, separating
the three distillation fractions (head, heart and tail). In the column top can be placed a partial condenser
or dephlegmator to generate the reflux necessary to operate the column, which can also be external.
In this case, the reflux generated is reintroduced through the upper part of the column. According to
this distillation system, beverages, such as Armagnac and Chilean Pisco, are obtained with some design
modifications and/or distillation strategies.

Figure 3. Batch column distillation (Adapted from Bureau National Interprofessionel de l’Armagnac,
http://www.armagnac.fr/distillation-et-vieillissement)

In other areas, brandies are distilled in continuous still by means of distillation columns (Fig. 4). In
this case, columns usually have the distillate output before the upper part, since the last trays are used to
concentrate the more volatile compounds, such as acetaldehyde, ethyl acetate and others, corresponding
to the head fraction in discontinuous distillation. At the bottom of the column, the residue is extracted,
in which the water would be concentrated with the less volatile compounds. In this column, lateral
extractions could be carried out, in which the concentration of higher alcohols is higher, so that their
presence in the final distillate can be minimised. This, however, has the disadvantage that the quantity of
first quality brandy obtained is reduced; in addition, this practice complicates the operation of the column
(Guymon, 1974).
These distillation columns usually have a large number of trays with the goal of producing distillates
at 95-96 per cent v/v, when only three or four trays are required to obtain distillates of 85 per cent v/v.
It is also possible to concentrate other undesired compounds, which are eliminated at the top of the
column, while the distillate is extracted in two or three trays below. In general, 60-70 per cent of the trays
are located above the feed tray, while the rest are below. The source of heat is usually by open steam,
although some brandy producers use indirect heating, such as reboilers (Guymon, 1974).
Another alternative is the employment of split column units as shown in Fig. 5, or systems with three
or four columns (Fig. 6) in which one of them concentrates the heads (aldehydes) up to 20 times and the
ethanol, devoid of aldehydes and other components of low point of boiling, is recycled to the main unit
(Bertrand, 2003).
Brandy Production: Fundamentals and Recent Developments 613

Figure 4. Continuous distillation column

Figure 5. Double column distillation


614 Applied Aspects of Winemaking: Production of Wine

Figure 6. Split column with a head concentrating column

3.3. Distillation Strategies


In the elaboration of brandies and distillates of the wine, different distillation strategies are carried out
to obtain a final distillate, which together with the subsequent ripening or aging treatments will return
the commercial product. Given below are some of those currently used in the production of different
brandies.
The first strategy is that in a single distillation in a still with a column or not, the head and tail
fractions are separated as described in Section 1.3.2. The obtained heart is the distillate to be used for the
elaboration of the final product. This strategy is applied, for example, in the production of Peruvian Pisco.
Another strategy consists in the realisation of a double distillation, known as the ‘charantais’ method,
which is used in France to make cognac with a traditional alembic. In the first distillation, a small head
fraction with an alcoholic strength around 60 per cent v/v is separated; then a heart fraction is collected
with an average alcoholic strength of 28 per cent v/v (range of 60-5 per cent v/v in ethanol), called
brouillis, and finally a tail with an average alcoholic strength of 3 per cent v/v (range of 5-0 per cent v/v
in ethanol). The heads and tails of the first distillation are reintroduced in the distillation of later batches,
mixing with the wine to be distilled. Then, this first heart (heart 1) is re-distilled where the heads are
separated (approximately 1 per cent by volume from the initial load and an alcoholic strength around
75 per cent v/v), in which the distillate is collected between 75-60 per cent v/v, obtaining an average
alcoholic strength of 70 per cent v/v. This is followed by heart 2 (called secondes) with an average
alcoholic strength of 30 per cent v/v (range of 60-5 per cent v/v) and finally the tails with an average
alcoholic strength of 3 per cent v/v (range from 5-0 per cent v/v). The heart 2 is redistilled and mixed with
the brouillis from the first distillation in subsequent batches. Likewise, the heads and tails of this second
distillation, with the heads and tails of the first distillation, are usually reintroduced in the first successive
distillations, mixing with the initial wine. With this strategy at the end of a distillation campaign, almost
all the alcohol from the wine distilled initially is recovered (Léauté, 1990).
Another strategy is to use a still with a column, as is the case of the Armagnac (Fig. 6), which involves
introducing the wine continuously into the upper part of the column and this is stripped of alcohol as it
goes down the column to the boiler. The vapour generated in the boiler produces the bubbling in the trays
Brandy Production: Fundamentals and Recent Developments 615

of the column so that it is rectified as it rises through it. The vapour, once outside the column, condenses
in the condenser, using the wine that is continuously fed as a cooling fluid, which in turn warms up before
entering the column. Operating in this way, the energy consumption in the boiler is reduced. The obtained
distillate has an alcoholic strength of 50-54 per cent v/v. At the same time, exhausted wine (wine residue)
must be extracted from the bottom intermittently to operate in a semi-continuous manner (Fig. 7).

Figure 7. Armagnac column distillation operating in continuous (Adapted from Bureau National Interprofessionel
de l’Armagnac, http://www.armagnac.fr/distillation-et-vieillissement)

On the other hand, with similar equipment in the elaboration of Chilean Pisco (Fig. 8), the distillation
is usually performed discontinuously, working in devices with larger boilers, in which the wine is loaded
and distilled in a discontinuous manner, separating the head, heart and tail fractions. With this strategy,
the product obtained is a mixture of distillate in a range of concentrations in ethanol from 70-40 per cent
v/v, being therefore, the characteristic of the final product more complex aromatically than Armagnac.

3.4. Aging Systems


The aging of brandy is one of the most important operations to be carried out in wooden barrels of oak.
This process may last from several months to several years. Relatively new distillates are placed in
barrels with a maximum capacity of 225-1000 L, while more aged, distillates can be stored in 5000 L
barrels. During this period, there is a loss of 2-3.5 per cent of the distillate per year due to evaporation
(Tsakiris et al., 2014). The sensory attributes of the distillates, such as colour, flavour and taste are greatly
influenced by the botanical species of the wood, the different heat treatments applied to wooden barrels,
the times that the wooden barrel has been used and the ageing time. This treatment will largely define
the quality of the final product, due to the physico-chemical processes that involve the distilled alcohol
and the wood of the barrels. During aging, the brandy is oxidised slowly. Its acidity grows by oxidation
616 Applied Aspects of Winemaking: Production of Wine

of alcohol into volatile acids and by dissolution of the acid


substances in the wood. Moreover, acetals are formed
and their odours are softer than those of aldehydes. The
odour characteristic of young brandy becomes blurred and
eventually disappears, to be replaced by a vanilla odour
induced by vanillic aldehyde together with other phenolic
aldehydes and acids arising from the alcoholysis of lignin
in oak wood (Bertrand, 2003).
An oak is a tree or shrub in the genus Quercus. There
are around 600 species of Quercus worldwide, of which
about 20 are economically important. Historically, the
species most used in cooperage are Q. alba (American white
oak) and two European species, Q. robur (pedunculate oak)
and Q. petraea (sessile oak). In practice, they can often
serve the same purpose but there are differences in wood
extractives; most notably the fact that American oak wood
contains lower ellagitannin content than the European
species, and its wood is denser and more coarse-grained
than European oak (Hornsey, 2016). Figure 9 presents the
composition of oak heartwood.
Cernîsev (2017) suggested a simplified scheme
for lignin degradation during spirit maturation. Lignin
consists mainly of coniferyl, p-coumaryl and sinapyl
alcohols. These alcohols, relatively easy, quickly convert to
coniferaldehyde, p-coumaryl aldehyde and sinapaldehyde,
which then are oxidised to respective acids. Theoretically, Figure 8. Pisco distillation column
a part of the formed phenolic acids can be esterified in the
presence of alcohol. Ferulic and protocatechuic acids are also present in aged spirits: the first one is
formed as result of gentle oxidation of coniferaldehyde to vanillin. Formation of protocatechuic acid can
be explained by conversion via demethylation of vanillic acid or oxidation of p-hydroxybenzoic acid.
Aromatic acids can be also converted to volatile phenols, such as 4-methylguaiacol, 4-ethylguaiacol,
eugenol, guaiacol, syringol, 4-methylsyringol and 4-ethylphenol. Thus, the concentrations of ethyl
esters of fatty acids increase during ageing, but the concentrations of esters of other alcohols, such as
3-methylbutyl acetate, decrease by transesterification (Christoph and Bauer-Christoph, 2007). The colour
becomes brown by dissolution of tannin. Moreover, the taste softens with the appearance of sugars arising
from the hydrolysis of wood hemicelluloses. There is also the appearance of a rancid taste by oxidation
of fatty acids (Bertrand, 2003).
Among the many species of the genus Quercus, only a few are of major technological interest
for ageing: American white oak (Quercus alba), sessile oak (Quercus petraea) and pedunculate oak
(Quercus robur). While European pedunculate oak has high quantities of extractable ellagic tannins,
sessile oak releases much smaller quantities of polyphenols and white oak even less. The American
species have a greater aromatic potential than European oak due to their high content of cis/trans isomers
of β-methyl-γ-octalactone. American white oak is easily identified by the low quantity of extractable
polyphenols, the high methyl-octalactone content and the presence of two isomers of 3-oxo-retro-α-
ionol. European sessile oak and American white oak are ideal for aging fine wine. Pedunculate oak, with
its low aromatic potential and high ellagitannin content, is best suited to aging spirits. Proper control of
toasting operations in barrelmaking could facilitate the use of this type of oak, by modelling the release of
volatile and odourous substances from wood (Chatonnet and Dubourdieu, 1998; Prida and Puech, 2006).
The extraction of tannins will depend on the size of the grains with lesser for fine grains than for coarse
grains. One of the most important treatments is toasting, since it will affect both the aroma and the colour
of the final product. For example, the hydrolysable tannins extracted in the brandy will cause different
reactions, such as the formation of esters from alcohols and acids. Lignin is a wood material that releases
aromatic aldehydes, such as vanillin, coniferaldehyde and syringaldehyde. Vanillin is a characteristic
Brandy Production: Fundamentals and Recent Developments 617

Figure 9. Composition of oak heartwood, where the percentages indicate the percentage of wood mass
made up by the constituent (Adapted from Hornsey, 2016; Mosedale and Puech, 2003)
aroma of cognacs and other alcoholic beverages aged in oak, as it is found in large concentrations after
toasting. The sweetness of the brandies is attributed to glycerol, xylose, arabinose and glucose extracted
from wood. Other typical aromas due to the aging of brandies in oak are the oak lactones, since at very
low concentrations they contribute aromas of coconut – an aroma associated with alcoholic beverages
aged in oak. Semi-volatile and non-volatile compounds of wood change the colour of the distillate and
contribute to an up-rounded flavour. The wooden barrels that are permeable allow air to pass in and cause
ethanol to evaporate; thus, the ethanol content decreases and the aroma gets more intense, complex and
concentrated. It has been proved that a decrease in the oxygen transfer rate is governed mainly by the
advance of the moisture front in the wood (wood impregnation), in contrast to the hypothesis which
attributes it to the oxygen consumption by the soluble ellagitannins of the wood (del Alamo-Sanza and
Nevares, 2017). In addition, harsher aroma constituents are removed and the spirit changes to mellow.
The period of maturation depends on the size of the casks used, the alcoholic strength, as well as the
temperature and humidity in the warehouse, which leads to a smoother flavour (Christoph and Bauer-
Christoph, 2007). Another factor to consider is the aging temperature, since it affects the aging rate as
well as a faster loss of ethanol by evaporation.
Globally, the aging process is characterised by changes in flavour, aroma and colour of the brandies,
as well as a decrease in the amount of product and the alcoholic strength. The aging time in oak barrels
varies according to the country (Louw and Lambrechts, 2012). The European regulations fix that brandies
must be aged for at least six months in oak barrels, while in United States brandies must be labelled with
the word ‘immature’ if the period of aging is less than two years and in South Africa, the minimum time
is three years.
Regarding the aging time, it will obviously affect the amount of compounds extracted from the wood,
as well as the reactions that occur. Therefore, longer aging time means higher compounds extraction.
According to Bertrand (2003), a barrel can yield substances to the cognac for about 40 years though
several stages in aging may be distinguished:
• By 1.5-5 years, the main process is dissolution of substances in the wood.
• By 5-10 years, astringency decreases and the brandy becomes rounder.
• By 10-35 years, a rancid taste appears.
• After 40 years, one should no longer keep brandies out of the barrel.
618 Applied Aspects of Winemaking: Production of Wine

Caldeira et al. (2016) studied the kinetics of the odourant compounds of a wine brandy during two years
of ageing in two ageing systems on an industrial scale. The odourant compounds in the analysed brandies
changed significantly over the time, but with different evolution patterns. With regards to the odourant
compounds proceeding from the distillate, namely alcohols (isobutanol, 2+3-methyl-1-butanol,trans-2-
hexen-1-ol, linalool, 2-phenylethanol) and acids (butanoic, isovaleric, hexanoic and dodecanoic acid) the
tendency was to decrease over two years of ageing. Regarding the esters, also derived from the distillate,
an inverse tendency for two esters was noticed (diethyl malate and ethyl octanoate), which increased over
the time, while the other four esters (isobutyl acetate, ethyl butyrate, isoamyl acetate and ethyl hexanoate)
were not affected by the ageing time. The kinetics of the majority of wood-related odourant compounds
(acetic acid, furanic aldehydes, volatile phenols, vanillin, acetovanillone and cis, trans-b-methyl-c-
octalactone) followed a hyperbolic pattern with a major increment at the beginning of the ageing period,
along with the diffusion of the compounds from the wood into brandies. However, for some of the wood–
related compounds, such as vanillin, acetovanillone, guaiacol, eugenol, 4-methylguaiacol and trans-b-
methyl-c-octalactone, the initial hyperbolic increase was followed by a linear enrichment, suggesting
their formation during ageing.
Cernîsev (2017), in his study, performed on 24 wine distillates with ageing time from one to 50
years and containing 69.1-43.4 per cent v/v alcohol observed that concentration of unsaturated aromatic
compounds (sinapaldehyde, coniferaldehyde) decreases with increasing maturing time of the distillates
(Fig. 10), probably due to their oxidation during long maturation, while the concentration of other
substances, such as syringaldehyde, vanillin, syringic and vanillic acids, increases. It is also interesting to
note that ratio syringaldehyde/vanillin lies between 1.80 and 2.21 almost for all distillates. The fact that
the ratio of syringaldehyde/vanillin is constant (2.0 ± 0.2) for most of aged distillates indicates a possible
correlation between the transformations of these aldehydes.
The use of oakwood fragments, as an alternative to the traditional barrels, is a rapid and economical
method of ageing treatment. Nowadays, there is no legislation applied to the ageing of spirits in
contact with oakwood fragments, and as a result, research on the accelerated ageing of distillates is
scarce (van Jaarsveld and Hattingh, 2012; Rodriguez-Solana et al., 2017). Schwarz et al. (2014) carried
out accelerated aging on a laboratory scale of Brandy de Jerez, employing oak chips and ultrasound
as extraction method and achieved in one-month sensorial characteristics and acceptability which are
similar to those of brandies whose average aging time was between six and 18 months.
Delgado-González et al. (2017) developed an accelerated aging process consisting of circulation
of the distillate through packed oak chips and the application of ultrasound. The best conditions were
obtained with wine distillate of 65 per cent (v/v), operating at room temperature (25ºC) and equivalent
doses of 5 g/L oak chips, using longer than fragmented ultrasound pulses, since the brandy presents

Figure 10. Mean concentrations of ageing markers in wine distillates from different age groups
(Adapted from Cernîsev, 2017)
Brandy Production: Fundamentals and Recent Developments 619

higher intensity of colour, TPI and extraction of furfuraldehyde, syringic acid, vanillin, syringaldehyde
and aromatic intensities and visual impressions. On the other hand, the aeration is positive, since the
TPI and the extraction of tartaric acid, syringaldehyde and syringic acid increases. The sensory analysis
showed that the distillates aged by five distillates of different grape varieties used in the Jerez area allow
brandies with characteristics of spirits aged by traditional methods in only three days.
Nevertheless, Rodriguez-Solana et al. (2017) for a pomace brandy conclude that in general, high
toast-oak fragments provide greater colour intensity in the accelerated aged spirits but reduce the
antioxidant capacity of the corresponding beverage. The best results are obtained with smaller fragment
size (granular) from Quercus petraea with medium toast level. The contact time did not significantly
influence the parameters evaluated. Caldeira et al. (2017) compared the influence of brandy aging in
650-L wooden barrels and in 3000-L stainless steel tanks. Both had Limousin oak (Quercus robur) and
Portuguese chestnut staves (Castaneas ativa). The brandy samples were profiled by descriptive sensory
analysis during the ageing period. The brandies aged in stainless steel tanks with staves presented higher
intensity of attributes, such as topaz, coffee, caramel and unctuous and lower intensity of golden, woody
and green attributes, and higher overall quality than the same wine distillate aged in wooden barrels.
Regardless of these differences, it is not possible to clearly distinguish the brandy samples proceeding
from different ageing systems by multidimensional analysis (MDS) of data. This effect seems to
contain the influence of a cooperage heating process, which imparts a different volatile wood
composition to the staves and wood barrels. Consequently, the brandies produced in wooden barrels are
associated with higher amounts of acetic acid, 5-methylfurfural, eugenol, acetic acid, cis-b-methyl-c-
octalactone and HMF, while brandies from stainless steel tanks with staves are linked to higher amounts
of volatile phenols (4-methylsyringol, syringol, guaiacol, 4-methylguaiacol) and acetovanillone (Caldeira
et al., 2016).
Zang et al. (2013a) studied the effects of applying an electric field (EF) treatment with 1 kV/cm
(50 Hz) on brandy stored in 5-L and 2-L oak barrels, respectively, for over 14 months to simulate the
natural aging process and it was compared with brandy sample naturally aged in 225-L oak barrels.
Results demonstrated that the content of phenol compounds in brandy, such as tannins, total phenols as
well as volatile phenols, significantly increased after treatment by EF, with tannin concentration of the
brandy treated with EF being 54.4 per cent and 43.9 per cent higher than those of the control brandy after
14 months of maturation in 5-L and 2-L barrels, respectively. It was also demonstrated that the EF-treated
samples in 5-L barrels for seven months and in 2-L barrels for five months exceeded the content of tannins
of those naturally aged for 12 months in 225-L oak barrels used in the brandy industry. The kinetic model
of oak phenol compounds extracted by brandy demonstrated that EF treatment played a positive role in
accelerating the extraction of phenol compounds and its effect was more significant than the size of the
barrel. Zang et al. (2013b) demonstrated that the application of an EF, the ester concentration, in the
brandy was higher than in the untreated samples. On the other hand, the concentration of some unpleasant
compounds apparently decreased; thus, the aroma of the brandy improved.

4. Brandy Styles in the World


The elaboration of brandies varied considerably in the world and in many places, it has a protected
geographical indication (GI) under the corresponding regulations. In some zones, practices, such as
the addition of wine, must, caramel and flavouring extracts based on dried plums, raisins, walnuts and
almond shells, are carried out. These spirits are normally aged in wooden casks (usually oak), a process
by which colour, mouthfeel and flavour are significantly changed (Christoph and Bauer-Christoph, 2007).
The characteristics of different types of brandies are described in the following section.

4.1. France
According to European regulations (European Union, 2008) in France, the geographical indications of
Cognac and Armagnac are recognised in the category of wine spirits, and the brandy Français in the
category of brandies. Cognac and Armagnac are made from wines from a limited geographical area and
are not characterised by a single specific flavour, but their quality depends on factors, such as the grape
varieties as well as the distillation system, aging and blending (Christoph and Bauer-Christoph, 2007).
620 Applied Aspects of Winemaking: Production of Wine

Cognac is produced in the region of Cognac (France), located north of Bordeaux. In 1909, the
French government established by means of a law that Cognac can be denominated only in a well-defined
zone that surrounds the city of Cognac. The Cognac production area is divided into six zones: Grande
Champagne, Petite Champagne, Borderies, Fins Bois, Bons Bois and Bois à Terroirs.
The grape varieties used are Ugni Blanc, Colombard and Folle Blanche, which have a minimum of
90 per cent of the wine destined for distillation. Although other varieties, such as Semillon, Folignan,
Juraçon Blanc, Montils, Sélect and Meslier St-François are authorised but with a proportion of less than
10 per cent in the production of the wine destined for distillation.
The grape harvest takes place in the second half of October and is pressed in horizontal presses
(continuous presses are prohibited), with the use of sulphur dioxide. The distillation of wine is with lees
and chaptalisation is not allowed. The alcoholic strength is generally relatively low (8-10 per cent v/v).
The wines are too acidic for direct consumption at pH 3 or even less. This acidity makes it possible to
some extent to compensate for the absence of sulphur dioxide. However, according to regulations, the
wines must be distilled before the end of March (Bertrand, 2003; Owens and Dikty, 2009).
Distillation is usually done with a copper Charantais traditional alembic (Fig. 1) and the procedure
used is double distillation (Fig. 11). The volume of the boiler is 30 hL maximum. The first distillation
is called ‘chauffe du brouillis’ and the second distillation ‘bonne chauffe’. The distillation strategy
is described in Section 1.3.2. Usually the second distillation is performed at a lower temperature for
obtaining better rectification of the spirit.

Figure 11. Cognac distillation process (Adapted from Leauté, 1989)

There are also variants of these two methods. Varying the intensity of heating is important according
to the alcoholic strength of brouillis required. Slow distillation gives good rectification. An odourous, fine
brandy is obtained but with dryness that may be detected on tasting due to the lack of certain products of
tail distillation (e.g. ethyl lactate, diethyl succinate). In contrast, fast heating involves the formation of a
marrowy brandy with little bouquet. Excessive heating results in a heavy taste (Bertrand, 2003).
The cognac is aged in oak barrels (200-600 L) from the forests of Tronçais, Allier, Limousin and the
Vosges in France. During aging, there is a loss of volume (around 3 per cent on an average per year) and
a 1 per cent v/v reduction in the alcoholic strength. During conservation, there is little evolution of the
volatile substances. The alcohol concentrate, the esters, are slightly hydrolysed and the unsaturated fatty
acids oxidise, giving a rancid taste. The Cognac ages by slow oxidation in barrels. In bottles, there is no
further evolution (Bertrand, 2003).
During aging, once a year, the cognac is racked and all the barrels of the same production are mixed
together. The alcoholic strength is gradually diminished by adding demineralised water to finally obtain
an alcoholic strength of 40.0 per cent (v/v) in commercialised bottles (never less). During aging, it is
Brandy Production: Fundamentals and Recent Developments 621

common to mix spirits of different origins, quality and age in order to obtain the same odour and taste for
a given brand name.
Accordingly the ageing period can be defined for different Cognac categories, as:
• Very special (V.S.) or three stars (***) where the youngest brandy is at least two years old
• Very Superior Old Pale (V.S.O.P) or Réserve where the youngest brandy is at least four years old
• Napoléon, X.O or Hors d’âge where the youngest brandy is at least six years old.
The colour can be adjusted with the addition of caramel. The taste can be adjusted with the addition
of woody water extract obtained from small pieces of oakwood to give more body to the spirit, more
astringency and a little bitterness. On the other hand, excessive hardness can be diminished by adding
sucrose syrup, generally less than 8 g/L. Generally, all these procedures should be carried out at least two
or three months before bottling.
Armagnac is produced in the region of Gascoigne (France), located east of Bordeaux. The Armagnac
production area is divided into three zones: Haute Armagnac, Bas Armagnac and Tenareze. The main
grape varieties used are Ugni Blanc, Colombard, Folle Blanche and Baco 22 (Owens and Dikty, 2009).
The distillation system used is described in Section 1.3.2 (Fig. 6). The wines to be distilled can be made
from musts obtained with continuous press and distilled with their lees. It is forbidden to add sulphur
dioxide. Wines are required to be distilled between the end of the harvest and 31 March of the following
year. According to the regulation, the maximum and minimum alcoholic strength allowed are 72 per cent
(v/v) and 52 per cent (v/v), respectively.
The distillation by means of stills with columns allows regulation of the alcoholic strength of the
distillate varying in wine flow and heating power. Therefore, lowering the heating or increasing the wine
flow brings down the temperature at the head of the column and results in a higher alcoholic strength.
In this case, higher alcohol and ester concentrations remain constant; on the other hand, the amount of
substances called tailings, of which there is usually a surplus in Armagnac, decreases exponentially when
the percentage of alcohol increases. For prolonged aging, a large quantity of tailings is an advantage
because of the ‘winey’ character of their molecules; but if the Armagnac is to be marketed soon, it is
preferable to make a high-proof distillate to limit the amount of such substances (Bertrand, 2003).
Armagnac is aged in oak casks and the coarse-grained wood is preferred (Gascony or Limousin)
to fine-grained wood, as it is slightly more permeable to oxygen and yields more tannin. Although
there are a variety of aging methods, spirits are usually kept in new casks for six months to one year
before being transferred to old casks. Armagnac contains vanillin, syringaldehyde, coniferaldehyde and
sinapaldehyde, but only vanillin is detectable on tasting. Prior to being marketed, several wine spirits
are blended and the alcoholic strength of the blend is reduced to a minimum of 40 per cent (v/v) with
distilled water. The naturally golden-yellow colour can be enhanced with caramel. Sometimes, infusions
or decoctions made from oak shavings are added to make the Armagnac more astringent and to give it
more body; however, these preparations must be at least of the same age as the youngest spirit used for
the commercial designation of the final product. Sugar solutions are sometimes added to attenuate the
‘burn’ of the alcohol (about 6 g/L).
Finally, before being bottled, the spirits are cold processed (usually one week at 5ºC) and passed
through a cellulose filter to eliminate any possible cloudiness due to excess of calcium or fatty acids.
Armagnac comprises several hundreds of substances, but the main features of old Armagnac are its
aroma of prunes, its rancid taste, its complexity; it is vigorous and even rough, with a long-lasting palate.
The blend of Armagnacs from different years of aging is authorised to obtain a more homogeneous
product in quality, but the age of the blend is that of the youngest Armagnac. The Bureau National
Interprofessionnel de l’Armagnac (BNIA) has harmonised the Armagnac since 2010 into four categories:
• Three Stars (***) or Very Special (V.S.) where the youngest brandy is at least one year old
• Very Superior Old Pale (V.S.O.P) or Réserve, where the youngest brandy is at least four years old
Napoléon or Extra Old (X.O.) where the youngest brandy is at least six years old
• Hors d’Âge, where the youngest brandy is at least 10 years old
Lastly, the Millésimes (Vintage), 10 years of minimum of ageing in wood, specific to Armagnac,
corresponds solely to the year of harvest declared on the label.
622 Applied Aspects of Winemaking: Production of Wine

It has been found (Bertrand, 2003) that there are about four times less ethyl esters of fatty acids
with 8, 10 and 12 carbon atoms in Armagnac than in Cognac. Some brandies contain noticeably lower
quantities of esters than others. This can be attributed to the utilisation of wine distillate, resulting in
poorer distillates containing volatile substances (Tsakiris et al., 2014). Cognac can then be differentiated
from Armagnac because this spirit contains highest contents of furan derivatives, such as furfural,
5-methylfurfural, furfuryl ethyl ether, and 2-acetylfuran due to the effect of double distillation (Ledauphin
et al., 2010).

4.2. Spain
In Spain, there are two brandies with geographical indication in accordance with the European regulation
of spirits (European Union, 2008): Sherry brandies (Brandy de Jerez) and Penedès brandy (Brandy del
Penedès).
Sherry brandy is produced in the area between the municipalities of Jerez de la Frontera, Sanlúcar
de Barrameda and El Puerto de Santamaría, located in the province of Cádiz (Spain). The regulation on
the preparation of these brandies (RDE Brandy de Jerez, 2005) allows the use of wine alcohol without
specifying its origin; however, most of the alcohol used in its preparation is distilled in the region of La
Mancha, located in the centre of Spain, using wines from the Airén variety.
The traditional distillation system is by means of alquitaras, a system similar to the charantais
alembic, in which distillates obtained are denominated holandas with a graduation between 40-70 per cent
(v/v) and content in volatile substances of 200-600 g/hL p.a. This distillation system is used to elaborate
the highest quality Sherry brandies. It is also used in more modern and efficient distillation columns in
which the wine is fed continuously. There are two types of columns – the so-called low-grade ones in
which spirits are obtained between 70-86 per cent (v/v) and a content in volatile substances between
130-400 g/hL p.a., and the high-grade ones in which the distillates are with a graduation between 86-94.8
per cent (v/v) and a content of volatile substances less than 50 g/hL p.a. In traditional distillation with
alquitara, the holandas are obtained in a single distillation unlike the Cognac, since the wines distilled
have a greater alcoholic strength, conserving better the characteristics of the wines.
According to its specific regulation, sherry brandy is matured through the traditional criaderas
and soleras system (Fig. 12), during which the spirit extracts its principal components from the interior
surface of the oak casks, although the vintage system is also used. The barrels used for aging usually are
of American oak (Quercus alba) with a capacity less than 1000 L, which have previously been used in
the production of Sherry wines.

Figure 12. Sherry brandy traditional aging system of criaderas and soleras
Brandy Production: Fundamentals and Recent Developments 623

According to the characteristics of the spirits and distillates of wine, and the process of elaboration
and aging, three types of ‘Brandy de Jerez’ are distinguished: Solera, Solera Reserva and Solera Gran
Reserva. Table 2 indicates the main characteristics of ‘Brandy de Jerez’.

Table 2. Characteristics of ‘Brandy de Jerez’

Brandy de Jerez type Aging system Aging time Barrel size Volatile substances
(L) (g/hL p.a.)
Solera criaderas and solera > 6 mon 1000 150
Solera Reserva criaderas and solera > 1 yr 1000 200
Solera Gran Reserva criaderas and solera > 3 yr 1000 250

The Solera sherry brandy is the youngest and fruity, with an average aging of one year, while the
Reserva usually has an average aging time of three years and finally the brandy Gran Reserva has an
average of 10 years. However, in practice, the best reserves and large reserves are aged 12-15 years. With
this procedure, it is possible to maintain the characteristics (flavour, aroma and colour), identical quality
and peculiarities of each brand. It is usually aged at a graduation between 50-60 per cent (v/v), that allows
the processes of maturation or extraction of characteristics of the wood to be more intense. In prolonged
aging, the graduation is reduced to a certain extent, mainly due to the evaporation that in Jerez is intense
– 7 per cent annually, due to its climate.
The taste of sherry brandies is sweet of ripe fruit, not only due to aging, but also due to the elaboration
process that authorises the addition of different substances, called Cabeceo, in which small amounts
of extracts and fruit macerations are used. According to the legislation, they can also use concentrates
of grapes and plums, oak essences, almond pericarp, vanilla pods and green walnuts and can be also
sweetened with natural sugars and caramel to adjust the colour. The alcohol content must be between 36-
45 per cent (v/v), although the usual commercial value is 38 per cent v/v and the highest quality products
are 39 per cent (v/v).
Penedès brandies are made in Catalonia, in the northeast of Spain, near Barcelona. The grape varieties
used are usually Macabeo and Parellada, as well as the Ugni Blanc variety used in the elaboration of
Cognac. Penedès brandy production is currently limited to two producers (Owens and Dikty, 2009). In
accordance with Penedès brandy regulation (Reglament IGP Brandy del Penedès, 2017), they can be
obtained by continuous (tray column) or discontinuous distillation (double distillation as in Cognac). The
spirits suitable for the production of ‘Brandy del Penedès’, obtained by distilling wines, must preserve the
organoleptic characteristics and volatile components of the raw material.
The types of distillates obtained are wine ‘holandas’ with alcohol content less than 70 per cent (v/v),
wine spirits with an alcohol content between 70-86 per cent (v/v) and wine distillates wines with an
alcohol content between 86-94.8 per cent v/v. In the final product, the sum of holandas and wine spirits
must represent more than 50 per cent alcoholic strength. In the case of double distillation, the maximum
degree of the second distillation will be 70 per cent (v/v).
Aging is performed both in static and dynamic systems in the area of Protected Geographical
Indication Brandy del Penedès. After distillation, the holandas are introduced in oak containers for aging
for a minimum of six months. In the case of dynamic aging, the process can be started in containers of
more than 1000 L, but it must be finished in barrels less than 500 L. For static aging, it must be in barrels
of less than 500 L.
For dynamic aging, there are the same categories as for Jerez brandies: Solera, Solera Reserva and
Solera Gran Reserva; for static aging or vintages, they are distinguished: a) ‘Reserva’, when aging is
higher than one year, b) ‘Gran Reserva’, when aging is greater than three years and c) ‘Extra Or’, when
aging is greater than five years with a minimum alcoholic strength of 40 per cent (v/v). ‘Extra Or’ category
must be elaborated by double distillation with a final product lower than 65 per cent (v/v).
The Penedès brandies must have a graduation higher than 36 per cent (v/v) and a low sugar content
(less than 20 g/L), being a dry product. The colour varies from amber to topaz; aromatically they are
intense with traces of nuts and notes of species, such as cinnamon, vanilla and nutmeg. The oak manifests
itself in the mouth with caramel and roasted notes.
624 Applied Aspects of Winemaking: Production of Wine

4.3. Other European Countries


In Italy, there is a tradition to make brandies from grapes, but they do not have defined geographical
indications. Normally, they are elaborated from native grapes and the distillation system used is distillation
columns and subsequent minimum aging of one to two years in oak barrels, up to ages of eight years.
They are characterised as being light and delicate with a soft touch of sweetness. In Germany, brandies
have been distilled since the middle ages. They are called weinbrand and have traditionally been made
from imported wines. Most of them are made in alembics and are aged in oak barrels for a minimum of
six months. Those aged for a minimum of one year are called old (in German uralt or alter). Brandies
tend to be softer and lighter than Cognacs and have a touch of sweetness. In Greece, distillates are made
from grapes using traditional alembics. Greek brandy is distilled from Muscat wine. Mature distillates are
made from sun-dried Savatiano, Sultana and Black Corinth grape varieties blended with an aged Muscat
wine. One of the best known is called Metaxa, which is flavoured with muscatel wine, anise and other
species. There are many varieties: centennial, golden, great reserve, Rhodes, etc. The number of stars
indicates the years of aging: three means a minimum of three years of aging; five stars, five years; seven
stars, seven. ‘Private reserve’ means at least 20 years (Dhiman and Attri, 2011; Owens and Dikty, 2009).

4.4. America
In the United States, the production of grape brandy was confined to California due to the tradition
of the old Spanish missions. At present, after the vicissitudes due to phylloxera, the dry law, after the
Second World War in California, the so-called California-style was developed under the influence of the
University of California at Davis (Owens and Dikty, 2009; Amerine et al., 1989). The brandies made
in this style are clean on the palate, light and with aromas that make them suitable for consumption.
Brandies are made by distillation columns from table grapes. The most commonly used varieties of
grapes are the Thompson seedless and the Flame Tokay, although grapes with little varietal character
are currently used, such as Chenin blanc, Folle blanche, French Colombard and Palomino (Owens and
Dikty, 2009). The California brandies use the grape variety which is of lesser importance than clean low-
temperature fermentations; also use fresh wine of immature grapes, good distillation practice and proper
aging (Amerine et al., 1989). These brandies are aged from two to 12 years in used American oak barrels.
Some producers use the aging method of sherry brandy. The more expensive brandies usually contain a
certain percentage of very aged brandies distilled in alembic.
In Latin America, in countries like Mexico, elaborated wine is used practically to obtain brandy. The
distillation methods include both traditional alembics and distillation columns. The grape varieties most
used are Thompson Seedless, Palomino and Ugni Blanc. The method of aging most used is that of soleras.
Brandy is also distilled from grapes in Chile and Peru to make a drink called Pisco. The most
commonly used grape variety is muscatel in both countries and the distillation procedures are traditional
alembic mainly in Peru and still with column in Chile. While in Peru the product does not age, in Chile
the trend is also to make Pisco age in oak barrels. Pisco is a very aromatic drink with a remarkable
presence of terpenic aromas.
According to Chilean legislation, Pisco is defined as a brandy produced and bottled in units of
consumption in the Atacama and Coquimbo regions through the distillation of genuine wine, from
Vitis vinifera varieties specified by law and planted in these regions (Bordeu et al., 2012). The Pisco
is a non-aged or slightly aged brandy. The distillation is mainly done by using batch column alembics,
separating head, heart and tail fractions. The average alcoholic strength of heart fractions is around 60
per cent (v/v). After the maturation or aging, Pisco is adjusted to the final consumption alcohol content
(35-50 per cent v/v), clarified and bottled. The product ranges from completely colourless to light amber
and is characterised by fruit Muscat aromas combined in different degrees with aromas coming from
aging and oak.
The commercial categories are:
• Artisanal Pisco, produced by small distillers with wines elaborated with skin contact during
winemaking. The distillate is more complex and needs some maturation or aging.
• White Pisco is young, distillates with less than six months aging, in general with no contact with
wood, frequently come from single varietal wines and can be triple distilled.
Brandy Production: Fundamentals and Recent Developments 625

• Mature Pisco, aged between six and 12 months, usually comes from a mixture of grape varieties.
Aromatically it is a mixture of fruity aromas and wooden notes.
• Aged Pisco, aged around one year, comes in young American oak barrels and aromatically have an
important contribution of wood-derived aromas combined with fruity aromas. It is elaborated from
multi-varietal wines and have an amber colour.
In general, Chilean Pisco is characterised by Muscat aroma, like geraniol, linalool, raisin and honey.
Even if Pisco is not aged for long periods in wood, descriptors like vanilla, oak and toasted are important
in describing products with some aging (Bordeu et al., 2004).
Peruvian Pisco is a grape brandy distilled from must of fresh grape in alembics, which do not rectify
the final product. Thus, it is transparent, slightly yellowish with an alcohol content of about 42 per cent
v/v. According to the Peruvian legislation (Indecopi, 2006), Pisco is the product obtained exclusively by
distillation of fresh musts of freshly fermented Pisco grapes, using methods that maintain the principles
of quality. It is produced on the coast of the specified departments. The Pisco varieties can be aromatics
(Italy, Muscat, Albilla and Torontel) or non-aromatics (Quebranta, Negracriolla, Mollar and Uvina). The
distillation system could be a traditional alembic called ‘falca’ and the Charantais alembic with and
without wine pre-heater. Rectification is not authorised and the addition of water is not allowed to adjust
the alcohol content of the final product. Pisco must mature for three months in containers that do not give
aromas or flavour. The different types of Pisco recognised are:
• Pure Pisco (Pisco puro), obtained from a single grape variety
• Pisco green must (Pisco mosto verde), obtained from the distillation of fresh musts with incomplete
fermentation
• Pisco acholado, which is obtained from the mixture of aromatic and non-aromatic grapes, musts,
wines or Piscos
The main aromatic descriptors associated with Peruvian Pisco for the Quebranta variety are fresh
fruit, dried fruit, citrus, chocolate, herbaceous, syrup, alcohol, chemical, acetic and empyreumatic; for the
variety Italia. they are the same in addition to aromatic grass, floral and sulphurous (Cacho et al., 2012).

4.5. Other Regions


In South Africa, there is also a tradition in the production of brandies since the 17th century with the arrival
of the Dutch. The introduction of new regulations and modern processing techniques at the beginning
of the 20th century allowed the development and improvement of traditional products. The brandies are
usually made with grape varieties – Ugni Blanc, Colombard, Chenin Blanc and Palomino in pot stills and
in column stills. (Owens and Dikty, 2009; Bougass, 2014; Bougass et al., 2014).
According to South African Liquor Products Act No. 60 of 1989, Sections 10, 12, 13 and 14 define
the requirements for grape spirit, pot still brandy, brandy and vintage brandy, respectively. If the grape
spirit is distilled in a continuous still, the alcohol content will be higher than 75 per cent (v/v), while if
it is distilled by pot still, the alcohol content will be lower than 75 per cent (v/v). The minimum alcohol
content of commercial product will be 43 per cent (v/v). The pot still brandy is distilled below 75 per
cent (v/v) and matured by storage for a period of at least three years in oak casks, with a capacity of not
more than 340 L. The alcohol content of the commercial product is a minimum of 38 per cent (v/v) and it
accepts a maximum of 10 per cent of wine spirits. Brandy consists of a mixture of not less than 30 per cent
of pot still brandy without grape spirit, wine spirit, spirit or a mixture thereof added in terms of regulation.
Brandy shall have an alcohol content of at least 43 per cent (v/v). Vintage brandy shall be produced in
such a manner that at least 90 per cent of the volume thereof is brandy according to the regulation. The
pot still portion requires an additional period of maturation of at least five years in oaken casks with a
capacity of not more than 1000 L; and the other portion must be matured for at least eight years in oaken
casks. The alcohol content of the commercial product is at least 38 per cent (v/v).
Pot still brandy is by far the more complex brandy of the three styles and is considered the richest,
fruitiest and most layered in style (Bougass et al., 2014). South African brandies are distinguished by
their overt fragrances of stone fruit, like apricots, peaches and pears (Owens and Dikty, 2009).
626 Applied Aspects of Winemaking: Production of Wine

4.6. Fruit Brandies


Fruit distillates, normally called brandies, are produced exclusively through alcoholic fermentation and
distillation of fleshy fruit or the must of berry, or vegetable, with or without stones and distilled at less
than 86 per cent (v/v). According to the raw material, three main types of brandies can be distinguished:
distillates obtained with pome fruits, of which apples and pears are the most common; those obtained with
stone fruits, mainly sweet cherries, sour cherries, plums, apricots, and peaches; and finally the distillates
obtained from berries. While the distillates with pome fruits are those traditionally obtained from pears
and apples, new brandies with other fruits are also being developed to valorise surpluses (Lopez et al.,
2017).

5. Brandy Composition and Sensorial Characteristics


Brandy is a wine distillate, so its initial chemical composition will depend on the volatile compounds
in the wine. Although ethyl alcohol and water are the two main components of any distilled spirit, the
character of aroma and flavour depend on a multitude of minor compounds, which are generally referred
to as congeners or congenerics. After distillation, brandy does not contain the non-distillable wine’s
organic and inorganic compounds, and this absence of non-volatile organic acids in the distillate affects
the balance of the taste. The brandy’s high alcohol content, about 40 per cent v/v, ‘burns’ the mouth and
simultaneously augments the sweet taste (Tsakiris et al., 2014). Therefore, the aging process produces
a complexity of minor components that generally improve palatability and improves aroma and flavour.
More than 500 substances have been detected in spirits (Cantagrel, 2003); they belong to a large
number of chemical classes: alcohols (ethanol, methanol, higher alcohols, etc.), aldehydes, esters,
volatile acids (acetic acid, fatty acids, etc.), ketones, acetals, nitrogen-, oxygen- and sulphur-containing
heterocyclic compounds, phenolic acids, and aldehydes. These different aromatic constituents are
contributed by:
• the grape (primary aromas)
• winemaking (fermentation aromas)
• distillation (specific aromas produced by the heating process)
• aging (aromas imparted by the oak wood)
The volatile composition of brandies is also characterised by high contents of low molecular weight
ethyl esters and some phenols, aldehydes, and acetals. However, these major components, due to their
high concentrations in all brandies, may not be helpful in the discrimination of samples. The volatile
compositions of brandies, such as Mirabelle, Calvados, Cognac and Armagnac are qualitatively rather
similar. However, their organoleptic characteristics are really different, due to slight differences in the
concentrations of volatiles (Ledauphin et al., 2010). Wine distillates’ authentication, mainly in terms of
varieties and regions of geographical origin, is rather difficult. However, differentiation can be realised by
the utilisation of statistical methods, such as principal component analysis (Tsakiris et al., 2014).

5.1. Alcohols
Alcohols possessing more than two carbon atoms are known as higher alcohols or fusel alcohols and are
by-products of yeast fermentation during the elaboration of base wine and form an important part of the
volatile composition of brandy (Louw and Lambrechts, 2012), being quantitatively the largest group of
volatile flavour compounds. Table 3 summarises the odour quality, odour threshold value in water and/
or ethanol solution, and concentration range of single volatile compounds in distilled spirits produced
during alcoholic fermentation from carbohydrates by yeasts and other microorganisms.
The term ‘fusel alcohols’ refers to their malty and burnt flavour, with the exception of 2-phenethyl
alcohol, which has a rose-like odour. The higher alcohol content remains almost unaffected before
distillation. The most important alcohols are 1-propanol, 2-methyl-1-propanol, 2-methylbutanol,
3-methylbutanol, and the aromatic alcohol 2-phenylethanol. Most straight-chain alcohols and their esters
have a strong pungent smell. At high concentration levels, higher alcohols are characterised by pungent
odours, which mask the aromatic finesse. They reach concentration in the order of 2.5-5.0 g/Lp.a and
their recovery is about 90 per cent with the exception of 2-phenyl-ethanol which is recovered 10 per
Brandy Production: Fundamentals and Recent Developments 627

Table 3. Odor Quality, Odour Threshold Value in Water and/or Ethanol Solution and Concentration
Range of Main Alcohols in Distilled Spirits

Compound Flavour quality Threshold Threshold Typical


(mg/L water)* (mg/ L ethanol concentration
solution)* (g/L p.a.)
Methanol Alcoholic - 668 0.30–0.70
1-butanol Alcoholic 0.5; 1.3 820 0.0025-0.200
2-butanol Alcoholic - 1,000 0.001-0.8
1-propanol Stupefying 500 830 0.1-2
2-Methyl-1-propanol Alcoholic - 40; 75 0.1-1
2-Methyl-1-butanol Malty 0.32 7; 30 0.02-1.8
3-Methyl-1-butanol Malty 1 7; 30 0.01-3
Allyl alcohol Unpleasant 19 - 0.01-0.13
Phenethyl alcohol Rose-like 1 7.5; 10 0.01-0.08
*Minimum and maximum threshold values cited from the literature

cent (Christoph and Bauer-Christoph, 2007; Tsakiris et al., 2014). At lower concentration levels, higher
alcohols may add complexity to brandy due to synergistic interactions with other compounds. In addition,
higher alcohols may be esterified during the aging process, resulting in esters that impart more pleasant
aromas (Louw and Lambrechts, 2012).
Methanol, 1-butanol and 2-butanol are not compounds of alcoholic fermentation and their threshold
values are rather high (668, 820 and 1000 mg/L p.a., respectively) and therefore, they do not contribute
significantly to the flavour (Christoph and Bauer-Christoph, 2007). Methanol is always present in
very small quantities of 40-60 mg/L in wine. However, in distillates and brandy, it is found in higher
concentrations of 0.30–0.70 g/L p.a. with a recovery of 90 per cent. Its taste is similar to ethanol and it
does not affect the organoleptic quality of the spirits. However, it affects the safety of brandy because its
toxicity is well known. European Union legislation requires a limit lower than 2.0 g L−1 of pure alcohol.
Methanol is reduced during ageing in barrels (Tsakiris et al., 2014); 2-Butanol levels higher than 0.5 g/L
p.a. indicate a bacterial spoilage of raw materials or mash (Christoph and Bauer-Christoph, 2007).
The herbaceous odour of brandies is due to grape-derived carbonyl compounds with six carbon
atoms. Unripe grapes and continuous presses may induce herbaceous tastes by liberating compounds,
such as hexanols (hexanol-1 and hexanol-2) and hexenols (cis-3-hexene-1-ol, trans-2-hexen-1-ol, cis-2-
hexen-1-ol). 1-Octen-3-ol is characterised by a mushroom odour and is produced in grapes infected by
Botrytis cinerea (Tsakiris et al., 2014).

5.2. Terpenes
The Muscat flavor is produced by terpenes. The chemical compounds responsible for this specific aroma
are mainly geraniol, nerol, α-terpineol, linalool and ß-citronellol. These monoterpenes are part of a large
family of molecules obtained by association of isoprene units. The total free terpene concentration varies
between 0.6-1.5 mg/L, depending on the grape variety and cultural parameters, with important variations
in the respective proportions of terpenes. The olfactory perception thresholds of these compounds are
rather low (a few hundred micrograms per litre). In Muscat wines, terpenes are found either free or bound
to sugars as glycosides; the latter are also called aroma precursors, as they are unable to express their
aromatic character. The sensorial characteristics are described as rose flower (geraniol, nerol), rosewood
(linalool) and geranium flower (geraniol, citronellol, terpineol) (Tsakiris et al., 2014; Colonna-Ceccaldi,
2010).

5.3. Volatile Acids


Acetic acid can be produced during and/or after fermentation by oxidation of ethanol under aerobic
conditions by the acetic acid bacteria Acetobacter; acetic acid levels should not be higher than 1 g/L p.a.
628 Applied Aspects of Winemaking: Production of Wine

in distilled spirits, since higher levels may contribute to a typical vinegar-like off-flavor (Christoph and
Bauer-Christoph, 2007).
In wine, it is found in concentrations ranging between 300-700 mg/L, while in distillates the
concentration ranges from 0.20-1.0 g/L p.a. Recovery is as low as 2-5 per cent due to the removal of the
distillation ‘heads’. However, distillates produced by continuous distillation may contain higher amounts
of acetic acid due to the absence of the removal of the distillation heads. During wine maturation in
barrels, a small quantity of acetic acid can be produced from ethanol oxidation or it can be extracted from
wood hemicelluloses (Tsakiris et al., 2014).
Other carboxylic acids, such as propionic acid and butyric acids, may also be present and they are
associated with bacterial activity. Butyric acid is characterised by an unpleasant, buttery and cheesy
aroma and its concentration increases during ageing. Hexanoic, octanoic, decanoic, dodecanoic, myristic
(14 carbon atoms), palmitic (16 carbon atoms) and stearic (18 carbon atoms) acids are formed by
yeasts (Tsakiris et al., 2014). In distilled spirits the concentration of free fatty acids is low owing to the
esterification and separation by distillation; thus, the concentration in wine distillates, like Cognac, is in
the range of 50 mg 0.1 L-1 p.a. (Christoph and Bauer-Christoph, 2007).

5.4. Ethyl Esters


Esters are condensation products of the carboxyl group of an organic acid and the hydroxyl group of an
alcohol or a phenol. Yeasts produce esters after cell division has ceased. Table 4 summarises the odour
quality, odour threshold value in water and/or ethanol solution, and concentration range of main esters
in distilled spirits.

Table 4. Odour Quality, Odor Threshold Value in Water and/or Ethanol Solution and Concentration
Range of Main Alcohols in Distilled Spirits

Compound Flavour quality Threshold Threshold Typical


(mg/L water)* (mg/ L ethanol concentration
solution)* (g/L p.a.)
Ethyl acetate Solvent-like, nail polish 17.6 7.5 0.01–2
Ethyl butanoate Fruity, floral 0.001 0.02 <0.00025–0.008
Methyl butyl acetate Fruity, banana, pear 0.3 0.03 0.003–0.03
2-Phenethyl acetate Rose, honey, fruity 0.02 0.25 0.01–0.03
Ethyl hexanoate Apple, banana, violet 0.005 0.005 0.001–0.008
Ethyl octanoate Pineapple, pear 0.07 0.26; 0.002 0.01–0.05
Ethyl decanoate Floral, fatty 0.5 - 0.01–0.09
Ethyl dodecanoate Floral - - 0.004–0.08
Diethyl succinate - - 100 0.005–0.03
Ethyl lactate - - 100 <0.025-1
*Minimum and maximum threshold values cited from the literature

Ethyl acetate, mainly produced as a result of esterification of acetic acid, is the main ester to occur
in fermented products and their distillates; it contributes significantly to a solvent-like nail polish off-
flavour at levels higher than 400 mg /100 mL p.a. (Christoph and Bauer-Christoph, 2007) and the range
is about 0.4 to 0.8 g L-1 of pure alcohol (Tsakiris et al., 2014). Ethyl acetate has a recovery of 100 per
cent in continuous distillation and 60 per cent in batch distillation (Tsakiris et al., 2014). Ethyl acetate
is mostly found in the head fraction, and the ethyl lactate in the tail fraction; therefore, the distillation
cuts can influence the presence of these compounds in the final brandy. Ethyl lactate has a negative
influence on brandy quality, and usually is accompanied by diethyl succinate, which is regarded as
a spoilage compound that can be removed along with the tails (Louw and Lambrechts, 2012). Ethyl
butyrate concentration increases with ageing, with the absence of antioxidant and antimicrobial agents,
and mainly with increased temperature (Tsakiris et al., 2014).
Brandy Production: Fundamentals and Recent Developments 629

Longer-chain ethyl esters also contribute to the total ester content of brandy. These esters are
generally associated with fruity aromas and contribute to the overall brandy quality. Their quantities
and mutual proportions are of great importance for the perceived flavour of a spirit drink since their
concentrations are generally above the sensory threshold values. Especially the low-boiling ethyl esters,
like ethyl 2-methylbutanoate, ethyl hexanoate, and ethyl octanoate, and the acetates like ethyl acetate,
isoamyl acetate, isobutyl acetate, hexyl acetate, and 2-phenethyl acetate are of great importance for the
flavour of distilled spirits (Christoph and Bauer-Christoph, 2007). Over 160 esters have been identified in
wines and most of them are also present in brandies. Almost all esters are low-boiling-point compounds
and are distilled in the beginning. Thus, the point at which the heads are cut to remove undesirable
compounds must be controlled in such a way no to remove too many desirable esters. Similarly, the tails
must not be cut too early in the distillation to preserve more esters. Recovery varies between 40-60 per
cent, depending on the distillation technique (Louw and Lambrechts, 2012; Tsakiris et al., 2014).
The flavour fraction with the lowest volatility is composed of C14-C18 fatty acid esters; these esters
as well as the long-chain fatty alcohols may contribute to the stearine-like smell that is characteristic of
Scotch malt whisky, in particular. Malo-lactic fermentation also has an influence on the concentration of
these compounds; distillates show a loss of fruitiness and aroma intensity with decreasing levels of ethyl
hexanoate, hexyl acetate, 2-phenethyl acetate, and with increasing levels of ethyl lactate, acetic acid,
and diethyl succinate (Christoph and Bauer-Christoph, 2007). In general, wines pressed by continuous
presses contain higher amounts of ethyl esters of long-chain fatty acids (14-18) because they contain
relatively lower oxygen, which affects their synthesis by yeasts. Distilling wine in the presence of yeasts
contained in fermentation lees leads to enrichment in fatty acid esters, such as ethyl octanoate (fruity,
floral, pineapple, apple and pear) and ethyl decanoate (fruity, pear, wine, etc.). Aromatic synergy between
those esters strengthens their olfactory impact. Fermentation lees also supply the wine spirit with fatty
acids (C8-C18, saturated and unsaturated) which are at the origin of aromatic derivatives formed by their
oxidation into aldehydes or ketones during the ageing process (Lurton et al., 2012). Recent research
has shown that some brandies contain naturally rare ethyl esters, which may have some impact on their
aroma. Such compounds are ethyl 2-, 3- and 4-methylpentanoate and ethyl cyclohexanoate, which exhibit
pleasant strawberry–liquorices-like odours (Tsakiris et al., 2014). The concentration of ethyl pentanoate
is higher in German brandies than in Cognac and similar to some Spanish and French brandies (Uselmann
and Schieberle, 2015).
The ethyl ester content of brandies increases during ageing because of the slow esterification of
different organic acids with ethanol. As brandy matures, ethyl esters become less flavour-active due
to an increase in their solubility in aqueous ethanol by the wood-extracted materials. Preserving wines
before distillation in the presence of lees has been connected with increased content of ethyl esters, since
they are generally retained within yeast cells rather than being released into the fermenting must during
fermentation (Tsakiris et al., 2014). Duran-Guerrero et al. (2011) found that ethyl esters were the main
family of volatile compounds responsible for the differentiation among the three different categories of
sherry brandies. The concentration of ethyl esters increased during the aging process, appearing in higher
amounts in those sherry brandies submitted to longer aging in wood.

5.5. Higher Alcohol Acetates


The formation of esters between acetic acid and higher alcohols is also important since they may
provide a fruity character, for example, isoamyl acetate, which has a characteristic banana odour,
positively influences brandy’s aroma. Low fermentation temperatures favour synthesis of fruity esters,
such as isoamyl, isobutyl and hexyl acetates, while higher temperatures favour the production of higher
molecular-weight esters. Both low SO2 levels and juice clarification favour ester synthesis and retention.
The absence of oxygen during yeast fermentation enhances further ester formation (Tsakiris et al., 2014).
Malo-lactic fermentation also has an influence on the concentration of these compounds; distillates
show a loss of fruitiness and aroma intensity with decreasing levels of ethyl hexanoate, hexyl acetate,
2-phenethyl acetate and with increasing levels of ethyl lactate, acetic acid, and diethyl succinate (Christoph
and Bauer-Christoph, 2007).
630 Applied Aspects of Winemaking: Production of Wine

Higher alcohol acetate synthesis as well hydrolytic breakdown continue enzymatically during ageing
based on the chemical composition and storage conditions of the brandy, due to formation of acetic acid
from the xylans extracted from the wood (Louw and Lambrechts, 2012; Tsakiris et al., 2014)

5.6. Aldehydes and Ketones


Acetaldehyde (ethanal) is the major important carbonyl compound of alcoholic fermentation and is
formed as an intermediate compound by degradation of pyruvate; its production by yeasts depends on
the pyruvate decarboxylase activity of the yeast (Christoph and Bauer-Christoph, 2007). The toxicity
associated with acetaldehyde is well known and its presence in the alcoholic beverages is quite often
related to nausea and vomiting. In distillates and brandies, it is found in concentrations ranging between
0.20-0.25 g L−1 p.a. (Tsakiris et al., 2014) and in wine distillates, higher concentrations do not affect
the quality owing to an odour threshold of about 100 mg/L (Christoph and Bauer-Christoph, 2007).
Since acetaldehyde is one of the most volatile compounds, the highest levels are in the ‘head cut’ of
the distillation and thus can be separated from the ‘heart cut’ (Christoph and Bauer-Christoph, 2007).
Acrolein (2-propenal) has a peppery, horseradish-like smell and is formed either by dehydration of
glycerol during distillation in the presence of acids on hot metallic surfaces or especially by bacteria
during fermentation of spoiled raw materials. The biochemical pathway of the formation of acrolein
is initiated by a bacterial dehydratase enzyme, which converts glycerol to 3-hydroxypropionaldehyde
(Christoph and Bauer-Christoph, 2007). Acetaldehyde as well as acrolein react with ethanol to form the
acetals 1,1-diethoxyethane and 1,1,3-triethoxypropane, respectively, leading to a reduction in the pungent
odour of the aldehydes; the equilibrium concentration of the 1,1-diethoxyethane formed is close to 10 per
cent in relation to the amount of acetaldehyde present. Higher aldehydes and their acetals can be found at
concentrations less than 0.1 mg/L (Christoph and Bauer-Christoph, 2007).
Isobutanal, at concentrations higher than 25 mg L−l, could give an herbaceous character to the
brandy. However, during ageing, its content declines due to acetylysation and selective evaporation.
Trans-Nonenal is characterised by a paper-like sense, while octanal contributes to the aroma complexity
by adding an orange flavour. Although concentrations of both the above aldehydes is increased during
ageing, it generally remains below the perception threshold. However, significant statistical correlations
have been obtained among the herbaceous odour and aldehyde concentration. β-Damascenone is an
isoprenoid ketone, which is present in grapes. Since it is a highly odoriferous compound with a powerful
and pleasant fragrance, it is an important compound in the perfume and flavouring industries. As it has
a very low sensory threshold, β-damascenone is considered a key odour compound in brandy, imparting
a stewed apple, fruity–flowery and honey-like character (Tsakiris et al., 2014). The concentration of
β-damascenone is quite high in Cognac than in German brandies, as well as in Spanish and some French
brandies. The ratio of the β-damascenone concentration to that of ethyl pentanoate allows differentiating
the German and Spanish brandies from the Cognacs (Uselmann and Schieberle, 2015).
Diacetyl (2,3-dioxobutane) is a ketone produced during wine fermentation through oxidation of
acetoin, a degradation product of citric acid. It has an important sensory influence on brandy since its odour
is characterised as sweet, buttery or butterscotch-like (Tsakiris et al., 2014; Christoph and Bauer-Christoph,
2007). Other aldehydes, which may be found in brandies, are formaldehyde, 5-hydroxymethylfyrfural,
propionaldehyde, butyraldeyde, benzaldehyde, isovaleraldeyde and n-valeraldeyde (Tsakiris et al., 2014).

5.7. Furfural and Furanic Compounds


Another aldehyde having a sensory impact like that of ‘baked’ in brandy, is furfural (0.5-82.5 mg L−l of
pure alcohol). Its synthesis involves sugar oxidation and is activated by heat. It is mainly produced during
distillation from the remaining pentose content of the lees and consequently is highly influenced by the
distillation system employed and double distillation enhances the amount of all furanic species. For this
reason, the concentration of furfural in brandy shows high fluctuation. Furfural and its derivatives may
also be derived from both the wooden cask and the possible addition of caramel. Concentration of furanic
compounds, such as furfuryl ethyl ether, furfural, 2-acetylfuran and 5-methylfurfural, varies according
to the type of cask and ageing time. Furanic aldehydes are derived from thermal degradation of wood
polysaccharides of wooden casks (Tsakiris et al., 2014).
Brandy Production: Fundamentals and Recent Developments 631

5.8. Volatile Compounds from Casks


It was found that significant differences in vanilla, woody, caramel, burned/toasted, green, tails and
rubbery aromas existed between brandies aged in barrels from different wood origins. By storing distillates
in wooden casks, volatile aroma compounds, like cis-β-methyl-γ-octalactone and trans-β-methyl-γ-
octalactone (oak lactone), furfural, 4-hydroxy-2-butenoic acid lactone, hexanoic acid, vanillin, guaiacol,
eugenol, cresols and other phenolic compounds migrate from the toasted wood into the distillate. These
compounds are responsible for the characteristic oak wood and vanilla-like flavour (Tsakiris et al., 2014;
Christoph and Bauer-Christoph, 2007; Louw and Lambrechts, 2012).
The toasting process can modify strongly the volatile composition of the different types of wood,
particularly the levels of furanic aldehydes (furfural, 5-methylfurfural with toasted almond aromas,
5-hydroxylmethyl furfural), volatile phenols (syringol and 4-allyl-syringol), propanoic acid, 4-hydroxy-
2-butenoic acid lactone and vanillin (Tsakiris et al., 2014). In contrast, non-toasted oak wood had a
small quantity of volatile phenols, mainly eugenol and traces of phenolic aldehydes (Zhang et al., 2015).
After three consecutive uses (three years), the wood will no longer liberate fatty acids, coniferaldehyde,
furfural, 5-methylfurfural and 5-hydroxymethyl-furfural into the wine spirits (Tsakiris et al., 2014).

5.9. Non-volatile Compounds from Casks


The ageing of spirits in oak barrels is a complex process. Direct extraction of wood components or
degradation products of macromolecules of the wood may occur, as well as reactions between the
components of the distillate itself and/or those originating from the oak wood (polymerisations,
esterifications, acetylysations, and hydrolysis), in addition to major oxidation processes. Apart from
ellagitannins, oak releases a certain number of other compounds into brandies, mainly lignins. Depending
on conditions, oak may also release polysaccharides, mostly consisting of hemicelluloses that contribute
to spirit flavour (Tsakiris et al., 2014).

5.10. Other Compounds


Brandy may contain extremely low concentrations of different unpleasant (rotten eggs, garlic) volatile
sulphur compounds, such as hydrogen sulphide, carbonyl sulphide, sulphur dioxide, thiols, sulphides,
polysulfides and thiosterols. The determination of these compounds in wines and spirits is difficult
because of their volatility and their very low concentrations, which require the use of highly sensitive
detectors (Tsakiris et al., 2014). Ethyl carbamate has been detected in several types of fermented foods
and beverages, such as alcoholic beverages, particularly stone fruit spirits. Ethyl carbamate is typically
formed from various precursors and one of the most important precursors is urea, which might be formed
during the degradation of arginine by yeast; another source is cyanate from the oxidation of cyanide.
The precursors react with ethanol to form ethyl carbamate and this occurs even during storage after
distillation, which explains why the concentration of the chemical in spirits is high (Tsakiris et al., 2014;
Pang et al., 2017). Ethyl carbamate is a potential carcinogenic compound and its presence is strictly
monitored in wines and spirits. Ethyl carbamate cannot be completely avoided because it occurs naturally
in fermented foods and beverages and the maximum levels for ethyl carbamate must be established based
on risk assessment and the levels of this chemical in spirits must be periodically monitored by producers
and supervisors (Tsakiris et al., 2014; Pang et al., 2017).

5.11. Caramel
According to the European legislation, wine brandies may only contain added caramel as a means to acquire
colour. Nevertheless, there is no legal limit for the concentration of this additive, neither specifications
for the product and the technological procedure have been prescribed. The addition of caramel is quite
common in the production of aged spirit beverages since it gives them an amber colouration, that is
attractive to the consumer. The chemical composition of caramel is complex due to the large number of
substances produced because of pyrolysis of carbohydrates, such as sucrose, glucose or starch. However,
furanic compounds are also present, such as furfural or 5-hydroxylmethyl furfural, of which the latter
is found in much higher concentrations (Tsakiris et al., 2014; Canas and Belchior, 2013). For quality
632 Applied Aspects of Winemaking: Production of Wine

control purposes, the correlation analysis between the caramel concentration and the characteristics of
the brandies reveals that caramel mainly influences HMF content, total phenolic index and coordinate a*,
since the correlations are positive and significant for all brandies as well as for the lightness which are
negative and are very significant for all of them. It is also shown that the ratio of furfural/HMF is a useful
tool to detect the addition of caramel in aged wine brandies (Canas and Belchior, 2013).

References
Amerine, M.A., Berg, H.W., Kunkee, R.E., Ough, C.S., Singleton, V.L. and Webb, A.D. (1989). The
Technology of Wine Making. AVI Publishing, Westport, CT, USA, pp. 582-642.
Bertrand, A. (2003). Brandy and cognac: Armagnac, brandy, and cognac and their manufacture. pp. 584-
601. In: Caballero, B., Trugo, I.C. and Finglas, P.M. (Eds.). Encyclopedia of Food Sciences and
Nutrition, second ed. Academic Press, Oxford, UK.
Bordeu, E., Agosin, E. and Casaubon, G. (2012). Pisco: Production, flavour chemistry, sensory analysis
and product development. pp. 331-347. In: Piggott, J. (Ed.). Sensory Evaluation and Consumer
Research. Woodhead Publishing Ltd., Oxford (UK).
Bordeu, E., Formas, G. and Agosin, E. (2004). Proposal for standardised set of sensory terms for Pisco, a
young Muscat wine distillate. American Journal of Enology and Viticulture 55: 104-107.
Bougass, N.V., Van Rensburg, P., Snyman, C.L.C. and Lambrechts, M.G. (2014). Brandy, cognac and
armagnac. pp. 248-276. In: Bamforth, C.W. and Ward, R.E. (Eds.). Food Fermentations. Oxford
University Press, New York, USA.
Bougass, N.V. (2014). Factors influencing the style of brandy. Ph.D. thesis. Stellenbosch University.
Cacho, J., Moncayo, L., Palma, J., Ferreira, V. and Culleré, L. (2012). Characterisation of the aromatic
profile of the Italian variety of Peruvian Pisco by gas chromatography – olfactometry and gas
chromatography coupled with flame ionisation and mass spectrometry detection systems. Food
Research International 49: 117-125.
Caldeira, I., Santos, R., Ricardo-da-Silva, J.M., Anjos, O., Mira, H., Belchior, A.P. and Canas, S. (2016).
Kinetics of odourant compounds in wine brandies aged in different systems. Food Chemistry 211:
937-946.
Caldeira, I., Anjos, O., Belchior, A.P. and Canas, S. (2017). Sensory impact of alternative ageing
technology for the production of wine brandies. Ciência e Técnica Vitivinícola 32: 12-22.
Canas, S. and Belchior, A.P. (2013). Effects of caramel addition on the characteristics of wine brandies.
Ciência e Técnica Vitivinícola 28: 51-58.
Canas, S., Caldeira, I., Anjos, O., Lino, J., Soares, A. and Belchior, A.P. (2016). Physicochemical and
sensory evaluation of wine brandies aged using oak and chestnut wood simultaneously in wooden
barrels and in stainless steel tanks with staves. International Journal of Food Science and Technology
51: 2537-2545.
Cantagrel, R. (2003). Chemical composition and analysis of cognac. pp. 601-606. In: Caballero, B.,
Trugo, I.C. and Fingla, P.M. (Eds.). Encyclopedia of Food Science and Nutrition, second ed.
Academic Press, Oxford (UK).
Cernîsev, S. (2017). Analysis of lignin-derived phenolic compounds and their transformations in aged
wine distillates. Food Control 73: 281-290.
Chatonnet, P. and Dubourdieu, D. (1998). Comparative study of the characteristics of American white
oak (Quercus alba) and European oak (Quercus petraea and Q. robur) for production of barrels used
in barrel aging of wines. American Journal of Enology and Viticulture 49: 79-85.
Christoph, N. and Bauer-Christoph, C. (2007). Flavour of spirit drinks: Raw materials, fermentation,
distillation, and ageing. pp. 219-239. In: Berger, R.G. (Ed.). Flavours and Fragrances. Springer,
Berlin, Heidelberg, Germany.
Colonna-Ceccaldi, B. (2010). Use of terpene-producing yeasts in brandy production. pp. 63-67. In:
Walker, G.M. and Hughes, P.S. (Eds.). Distilled Spirits, New Horizons: Energy, Environmental and
Enlightenment. Nottingham University Press, Nottingham, United Kingdom.
Brandy Production: Fundamentals and Recent Developments 633

del Alamo-Sanza, M. and Nevares, I. (2017). Oak wine barrel as an active vessel: A critical review
of past and current knowledge. Critical Reviews in Food Science and Nutrition, DOI:
10.1080/10408398.2017.1330250
Delgado-González, M.J., Sánchez-Guillén, M.M. García-Moreno, M.V., Rodríguez-Dodero, M.C.,
García-Barroso, C. and Guillén-Sánchez, D.A. (2017). Study of a laboratory-scaled new method
for the accelerated continuous ageing of wine spirits by applying ultrasound energy. Ultrasonics
Sonochemistry 36: 226-235.
Dhiman, A.K. and Attri, S. (2011). Production of Brandy (Chapter 35). pp. 1-60. In: Joshi, V.K. (Ed.).
Handbook of Enology: Principles, Practices and Recent Innovations, vol. III. Asiatech Publisher,
INC, New Delhi.
Duran-Guerrero, E., Cejudo-Bastante, M.J., Castro-Mejías, R., Natera, R. and García-Barroso, C. (2011).
Differences in the volatile compositions of French labelled brandies (Armagnac Calvados Cognac
and Mirabelle) using GC-MS and PLS-DA. Journal of Agricultural and Food Chemistry 59: 2410-
2415.
European Union (2008). Regulation (EC) No. 110/2008 of the European Parliament and of the Council
of 15 January 2008. Official Journal of the European Union L 39: 30-31.
Guymon, J.F. (1974). Chemical aspects of distilling wines into brandy. pp. 232-253. In: Webb, A.D.
(Ed.). Advances in Chemistry, vol. 137. Chemistry of Winemaking. American Chemical Society,
Washington, DC, USA.
Hornsey, I. (2016). Way through the wood. Barrel Biochemistry for Brewers and Distillers, Brewer and
Distiller International, April: 48-55.
Indecopi (2006). Reglamento de la Denominación de Origen Pisco. INDECOPI. NTP 211.001.2006,
Bebidas Alcohólicas, Pisco, Requisitos, Norma Técnica, Peru, 2006: 1-11.
Kohlmann, H. (2016). Future of the US spirit is market in light of accelerated fragmentation. Global
Drinks Forum, October 10, Berlin, Germany.
Lambrecthts, M., van Velden, D., Louw, L. and van Rensburg (2016). Brandy and Cognac: Consumption,
Sensory and Health Effects. pp. 456-461. In: Caballero, B., Finglas, P.M. and Toldrá, F. (Eds.).
Encyclopedia in Health. Academic Press, Oxford, UK.
Léauté, R. (1990). Distillation in Alembic. American Journal of Enology and Viticulture 41: 90-103.
Ledauphin, J., Milbeau, C., Barillier, D. and Hennequin, D. (2010). Differences in the volatile compositions
of French labelled brandies (Armagnac Calvados Cognac and Mirabelle) using GC-MS and PLS-
DA. Journal of Agricultural and Food Chemistry 58: 7782-7793.
López, F., Rodríguez-Bencomo, J.J., Orriols, I. and Pérez-Correa, J.R. (2017). Fruit brandies. pp. 531-
556. In: Kosseva, M.R., Joshi, V. and Panesar, P. (Eds.). Science and Technology of Fruit Wine
Production. Academic Press.
Louw, L. and Lambrechts, M.G. (2012). Grape-based brandies: Production, sensory properties and
sensory evaluation. pp. 281-298. In: Piggot, J. (Ed.). Alcoholic Beverages: Sensory Evaluation and
Consumer Research. Woodhead Publishing Ltd., Cambridge, UK.
Lurton, L., Ferrari, G. and Snakkers, G. (2012). Cognac: Production and aromatic characteristics. pp.
242-266. In: Piggott, J. (Ed.). Sensory Evaluation and Consumer Research. Woodhead Publishing
Ltd., Oxford (UK).
Mosedale, J.R. and Puech, J.L. (2003). Barrels, wines, spirits and other beverages. pp.393-403. In:
Caballero, B., Trugo, I.C. and Finglas, P.M. (Ed.). Encyclopedia of Food Sciences and Nutrition,
second ed. Academic Press, Oxford, UK.
Owens, B. and Dikty, A. (2009). The art of distilling whiskey and other spirits. Query Books, Beverley,
Massachusetts (USA).
Pang, X.N., Li, Z.J., Chen, J.Y., Gao, L.J. and Han, B.Z. (2017). Comprehensive review of spirit drink
safety standards and regulations from an international perspective. Journal of Food Protection 80:
431-442.
Prida, A. and Puech, J.L. (2006). Influence of geographical origin and botanical species on the content of
extractives in American, French and East European oak woods. Journal of Agricultural and Food
Chemistry 54: 8115-8126.
634 Applied Aspects of Winemaking: Production of Wine

RDE Brandy de Jerez (2005). Reglamento de la Denominación Específica ‘Brandy de Jerez’ y su Consejo
Regulador. Boletín Oficial de la Junta de Andalucía No. 122, Spain, 2005.
Reglament IGP Brandy del Penedès (2017). ORDRE ARP/246/2017, de 27 d’octubre, per la qual s’aprova
el reconeixement transitori i el Reglament de la Indicació Geogràfica Protegida Brandy del Penedès,
i es reconeix el seu Consell Regulador provisional. Diari Oficial de la Generalitat de Catalunya
Núm, 7488, Spain, 2017.
Rodriguez-Solana, R., Rodriguez-Feijedo, S., Salgado, J.M., Dominguez, J.M. and Cortes-Dieguez, S.
(2017). Optimisation of accelerated ageing of grape marc distillate on a micro-scale process using a
Box–Benhken design: Influence of oak origin, fragment size and toast level on the composition of
the final product. Australian Journal of Grape and Wine Research 23: 5-14.
Statista (2018). https://www.statista.com/statistics/693822/global-market-share-spirits/
Schwarz, M., Rodríguez, M.C., Sánchez, M., Guillén, D.A. and Barroso, C.G. (2014). Development of an
accelerated aging method for Brandy. LWT – Food Science and Technology 59: 108-114.
Tsakiris, A., Kallithraka, S. and Kourkoutas, Y. (2014). Grape brandy production, composition and
sensory evaluation. Journal of the Science of Food and Agriculture 94: 404-414.
U.S. Code of Federal Regulations, Title 27: Alcohol, Tobacco and Firearms, PART 5, Subpart C. Standards
of Identity for Distilled Spirits (2013).
Uselmann, V. and Schieberle, O. (2015). Decoding the combinatorial aroma code of a commercial cognac
by application of the sensomics concept and first insights into differences from a German brandy.
Journal of Agricultural and Food Chemistry 63: 1948-1956.
van Jaarsveld, F.P. and Hattingh, S. (2012). Rapid induction of ageing character in brandy products:
Ageing and general overview. South African Journal for Enology and Viticulture 33: 225-252.
Wiseguy (2018). Global Spirits Market Report (2018). WISEGUY RESEARCH CONSULTANTS PVT
LTD. https://www.wiseguyreports.com/reports/2941237-global-spirits-market-report-2018
Zhang, B., Cai, J., Duan, C.Q., Reeves, M.J. and He, F. (2015). A review of polyphenolics in oak woods.
International Journal of Molecular Sciences 16: 6978-7014.
Zhang, B., Zeng, X.A., Lin, W.T., Sun, D.W. and Cai, J.L. (2013). Effects of electric field treatments
on phenol compounds of brandy aging in oak barrels. Innovative Food Science and Emerging
Technologies 20: 106-114.
Zhang, B., Zeng, X.A., Sun, D.W., Yu, S.J., Yang, M.F. and Ma, S. (2013). Effect of electric field
treatments on brandy aging in oak barrels. Food Bioprocess Technology 6: 1635-1643.

You might also like