You are on page 1of 58

CHAPTER 1

1.0 INTRODUCTION

Fundamentally beer is the product of the alcoholic fermentation by yeast of extracts of

malted barley. Whilst malt and yeast contribute substantially to the character of beers, the

quality of beer is at least as much a function of the water and, especially, of the hops used

in its production (http://www.ibdasiapac.com.au/brewing).

Barley starch supplies most of the sugars from which the alcohol is derived in the

majority of the world's beers.

The two major categories of beer are ales and lagers, and the yeast used in fermentation

determines the differences between the two.

Lagers are usually drier, crisper, and less fruity in taste than ales. Lagers are fermented

with Saccharomyces carlbengsis at cooler temperature for longer periods up to two

weeks, which encourages the flavour components in lagers to combine and mellow.

Lagers are aged for up to 12 weeks. Lagers are best served chilled.

1.1 HISTORY OF BEER

Archaeologists have turned up evidence that the Sumerian people in the Middle East were

brewing barley grain as long as 8,000 years ago. These early people may have discovered

the basic processes of brewing when they observed—and then tasted—what happened to

fruit juices or cereal extracts left exposed to the wild yeasts that naturally float in the

air(http://www.scienceclarified.com/Bi-Ca/Brewing).

1
Over the centuries, breweries sprang up throughout Europe where there was good water

for brewing. During the Middle Ages (400–1450), monasteries became the centres for

brewing, and the monks originated brewing techniques and created many of the beers still

popular today (http://www.scienceclarified.com/Bi-Ca/Brewing).

Bottled beer was introduced by the Joseph Schlitz Brewing Company in Milwaukee,

Wisconsin, in 1875. The Gottfried Krueger Brewing Company released the first canned

beer in America in 1935 (http://www.scienceclarified.com/Bi-Ca/Brewing).

2
CHAPTER 2

2.1 BREWING RAW MATERIALS

Quantitatively, the most important raw materials used for the production of beer are the

carbohydrate sources that is barley (usually malted), adjuncts such as maize, wheat and

sorghum. Less commonly oats, and, in some instances, flavoured sugars (e.g. chip sugar),

hops, water (used in production, making up over 90% of the final product) and yeast.

2.1.1 Barley

Barley, which is the main raw material in beer, does not give, as such, a fermentable

extract with the yeast. It must first be left to begin to germinate. The barley starch is

enclosed in the cell wall and proteins within the barley, and these wrappings are stripped

away in the malting process (essentially a limited germination of the barley grains),

leaving the starch preserved. The mature barley grain comprises the embryo and the

endosperm, which provides a store of carbohydrates (mainly starch) and protein to

support the initial growth of the germinating embryo. The endosperm comprises the

starchy endosperm, which is a non-living storage tissue, surrounded by a living non-

starch cell layer called aleurone. The grain is protected by an outer husk.

3
Figure 1: Transverse section of a raw barley grain

Source: Wikipedia, the free encyclopedia

Barley selected for use in the malting industry must meet special quality specifications

shown below. Accepted malting barley varieties have to modify evenly and produce

finished malt whose properties lie within the brewer's specifications. The malt quality of

a given barley variety is determined by its genetic background and the physical

conditions during growth, harvest and storage (http://www.crc.dk/flab/malting barley).

Malting quality has to be tested in micro-, pilot- and industrial malting trials, and brewing

trials also in pilot and production scale.

The physical condition of the barley must meet specifications concerning

• Germination % min. 97% after 3 days

• Germination index min. 6,0

• Water content 12.0 %, max. 13.0 %

• Protein content > 9.0 % and < 11.5 %

• Grading min. 90 % > 2.5 mm.

4
• β-glucan content max. 4 %

• Micro-organisms below a set level.

• Pesticide residues according to national law

• Ochratoxin according to national law

• Aflatoxin according to national law

• Variety purity min. 99 %

Source: http://www.crc.dk/flab/malting barley

5
Table 1: Raw barley Nutritional value per 100 g

Raw barley
Nutritional value per 100 g (3.5 oz)
Energy 350 kcal 1470 kJ

77.7
Carbohydrates
g
- Sugars 0.8 g
- Dietary fiber 15.6 g
Fat 1.2 g
Protein 9.9 g
Thiamin (Vit. B1) 0.2 mg 15%
Riboflavin (Vit. B2) 0.1 mg 7%
Niacin (Vit. B3) 4.6 mg 31%
Pantothenic acid (B5) 0.3 mg 6%
Vitamin B6 0.3 mg 23%
Folate (Vit. B9) 23 μg 6%
Vitamin C 0.0 mg 0%
Calcium 29.0 mg 3%
Iron 2.5 mg 20%
Magnesium 79.0 mg 21%
Phosphorus 221 mg 32%
Potassium 280 mg 6%
Zinc 2.1 mg 21%

6
Source: USDA Nutrient database

2.1.2 Water

Water is an important raw material in the brewing process, so the brewery is dependent

on a reliable supply of good quality water. Since breweries and good water have long had

a close association, water quality is generally taken for granted. Large quantities of water

are used in production (water making up over 90% of the final product) as well as for

cleaning, washing and sterilizing of equipment. Water used for brewing must be fit for

human consumption (potable). As such it must be free from contaminating organisms.

However, what is fit to drink is not necessarily fit for brewing use. Water for brewing is

of course boiled during the process (http://www.crc.dk/flab/water.htm).

From a chemical point of view, it has been noticed that pure water has a flat unpleasant

taste. A balanced content of inorganic salts is preferable. This is achieved at the water

treatment plants of the breweries and controlled by water analyses using sophisticated

instruments such as:

• Atomic Absorption Spectrophotometer for the determination of metal ions, such

as Na+, K+, Ca++, Mg++, Fe++(+) and Mn++.

• Dionex Ion Chromatograph for the determination of inorganic anions, such as

chloride, nitrate, phosphate and sulphate.

• Gas Chromatograph to ensure the absence of trihalomethanes, originating from

the chlorination of surface water (http://www.crc.dk/flab/water.htm).

7
From a microbiological point of view the main concern is the introduction of spoilage

organisms from water introduced after fermentation, for example during dilution of beer

following high gravity brewing or from vessels rinsed with contaminated water.

A variety of methods are available for water purification, and generally micro-organisms

may be removed very effectively.

Membrane filtration is generally used for complete removal of bacteria, viruses, proteins,

salts and ions. Chlorine dioxide can be applied to water systems to reduce or eliminate

brewery spoilage organisms. At levels of approximately 0.2 ppm chlorine dioxide

significantly reduces microbial count while causing no off-flavours or odours in the final

beer.

Bacteriological analysis of water is designed to detect recent faecal pollution, for example

by farm animals or by sewage effluent. Such analysis ensures absence of coliform

bacteria, which may be spread by contaminated water supplies. The most dangerous,

Salmonella Shigella spp. and Vibrio cholerae, occur irregularly and in small numbers in

contaminated water, and it is not normal practice to culture these pathogens directly. For

brewers, the most significant of the coliform bacteria are Aerobacter aerogenes which

may be the cause of biologically unstable wort. Algae and fungi from water supplies are

also able to create problems within a brewery causing undesirable odours and taints,

clogging filters and providing nutrients for bacterial growth. The wild yeast Pichia may

be found in some supplies; Pichia is quite tolerant to anaerobic conditions, is able to spoil

wort, and grows readily in unpasteurized finished beer.

2.1.3 Hops

Hops are the female flower cones of the hop plant (Humulus lupulus). They are used

primarily as a flavouring and stability agent in beer, and also in other beverages.

8
The hops have two principal components: resins and essential oils. The resins (so-called

α-acids) are changed ('isomerised') during boiling to yield iso-α-acids, which provide the

bitterness to beer. Hops are dried in an oat house before they are used in the brewing

process. Hop resins are composed of two main acids: alpha and beta acids

(http://en.wikipedia.org/wiki/Hops).

Alpha acids have a mild antibiotic/bacteriostatic effect against Gram-positive bacteria,

and favour the exclusive activity of brewing yeast in the fermentation of

beer(http://en.wikipedia.org/wiki/Hops).

Beta acids do not isomerise during the boil of wort, and have a negligible effect on beer

flavour. Instead they contribute to beer's bitter aroma, and high beta acid hop varieties are

often added at the end of the wort boil for aroma. Beta acids may oxidize into compounds

that can give beer off-flavours of rotten vegetables or cooked corn

(http://en.wikipedia.org/wiki/Hops).

The flavour imparted by hops varies by type and use: hops boiled with the wort (known

as "bittering hops") produce bitterness, while hops added to wort later impart some

degree of "hop flavour" (if during the final 10 minutes of boil) or "hop aroma" (if during

the final 3 minutes, or less, of boil) and a lesser degree of bitterness. Adding hops after

the wort has cooled and the beer has fermented is known as "dry hopping", and adds hop

aroma, but no bitterness. The degree of bitterness imparted by hops depends on the

degree to which otherwise insoluble alpha acids (AAs) are isomerized during the boil,

and the impact of a given amount of hops is specified in International Bitterness Units.

Unboiled hops are only mildly bitter(http://en.wikipedia.org/wiki/Hops).

9
This process is rather inefficient. Nowadays, hops oils are often extracted with liquefied

carbon dioxide and the extract is either added to the kettle or extensively isomerised

outside the brewery for addition to the finished beer (thereby avoiding losses due to the

tendency of the bitter substance to stick on to yeast(http://en.wikipedia.org/wiki/Hops)).

Flavours and aromas are described appreciatively using terms which include "grassy",

"floral", "citrus", "spicy", "piney" and "earthy". Most of the common commercial lagers

have fairly low hop influence, while true pilsners should have noticeable noble hop

aroma and certain ales (particularly the highly-hopped style known as India Pale Ale, or

IPA) can have high levels of bitterness (http://en.wikipedia.org/wiki/Hops).

2.1.4 Sugar

Free flowing sugar, syrups or honey are commonly used adjuncts, generally added during

wort boiling. Some are also added as non-fermentable sweeteners. Specially tailored

syrups allow the production of better quality low alcohol beers. The main concern in

brewing involves transfer of bacterial spores, principally from Bacillus sp., which can

withstand heat treatment, including boiling, and may persist into the finished beer

(although beer does not support the subsequent growth of these organisms).

2.1.5 Finings (Copper or kettle finings)

Copper finings are Carregeenan gels from the seaweed Euchema cottonii . They consist

of a linear sulphated polysaccharide containing galactose and anhydrogalactose.

Of the various forms, the kappa type is the most effective. In the presence of calcium or

potassium ions, copper finings will remove proteins, polypeptides and possibly

polyphenols. This is caused by binding to the sulphate groups on the galactose units.

They are added towards the end of the boil at 2–5 g/hl at a pH of above 5.2. Used at an

10
optimum dose, copper finings give good wort clarity with compact stable sediment (AB

Vickers Ltd., personal communication). They do not help the clarity of hot wort, where

they are solubilised but will improve the clarity of cold wort by helping to precipitate

trub. High gravity wort require more finings than normal gravity wort. Filtered beers that

have been previously fined with either carrageenan or isinglass have lower hazes than

unfined beers. This is due to the finings either increasing the size of haze particles and

removing them, or by removing the precursors.

2.2 ESSENTIAL PRE-BREWERY OPERATIONS

Certain operations are very important before brewing process can commence. These

essential processes determine overall success of brewing process.

2.2.1 Malting

In the malt house, barley grain germination is initiated by the uptake of water in a steeping

vessel (A). The grain imbibes water during controlled cycles of water spraying or water

immersion followed by aeration, until the water content of the grain reaches 42 to 48%. This

is checked with the aid of a moisture analyzer. Water enters the grain via the embryo, and

after approximately 24 hours, the first visible sign of germination is the appearance of the

root, as a white 'chit'. The grains are then transferred to malting beds where germination is

allowed to proceed over a period of about 5 days (B). The speed of germination is controlled

by temperature and aeration of the malt bed, while moisture content is maintained by

spraying. Further embryo growth, with the appearance of rootlets and acrospire, can lead to

root entangling. The grain bed is regularly turned with a rotating screw to prevent grains

matting together.

Green malt, produced after five days of germination, is kiln dried and partly cooked in a

forced flow of hot air (C). Hydrolases produced during malting are partially inactivated

11
during this process. Malt colour, enhanced by kilning at higher temperatures, may be

desirable for production of darker beer, but it leads to further heat-inactivation of hydrolases.

The brittle malt rootlets are separated from the malt and utilised in animal feeds.

The kilned malt is stable for storage and has a friable texture suitable for the milling process

which proceeds brewing.

Figure 2: A Schematic representation of malting process

Source: (http://www.crc.dk/flab/malting.htm)

In summary, malt is made by allowing a grain to germinate, after which it is then dried in a

kiln and sometimes roasted. The germination process creates a number of enzymes, notably

β-amylase and α-amylase, which will be used to convert the starch in the grain into sugar.

Depending on the amount of roasting, the malt will take on a dark colour and strongly

influence the colour and flavour of the beer.

12
2.2.2 Water treatment plant

Water is an important raw material in the brewing process, so the brewery is dependent on a

reliable supply of good quality water. This is achieved at the water treatment plants of the

brewery.

Treatment

The following processes are used in water purification plants.

Pre-treatment

1. Pumping and containment - The majority of water must be pumped from its source

or directed into pipes or holding tanks. To avoid adding contaminants to the water, this

physical infrastructure must be made from appropriate materials and constructed so that

accidental contamination does not occur.

2. Screening- The first step in purifying surface water is to remove large debris such as

sticks, leaves, trash and other large particles which may interfere with subsequent

purification steps. Most deep groundwater does not need screening before other purification

steps.

3. Storage - Water from rivers may also be stored in bankside reservoirs for periods

between a few days and many months to allow natural biological purification to take place.

This is especially important if treatment is by slow sand filters. Storage reservoirs also

provide a buffer against short periods of drought or to allow water supply to be maintained

during transitory pollution incidents in the source river.

4. Pre-conditioning – Hard water is treated with soda-ash (Sodium carbonate) to

precipitate calcium carbonate out utilising the common-ion effect.

13
5. Pre-chlorination - In many plants the incoming water is chlorinated to minimise the

growth of fouling organisms on the pipe-work and tanks. Because of the potential adverse

quality effects of chlorine, this has largely been discontinued.

Widely varied techniques are available to remove the fine solids, micro-organisms and

some dissolved inorganic and organic materials. The choice of method will depend on the

quality of the water being treated, the cost of the treatment process and the quality

standards expected of the processed water (http://en.wikipedia.org/wiki/water treatment).

pH Adjustment

Distilled water has a pH of 7 (neither alkaline nor acidic) and sea water has an average

pH of 8.3 (slightly alkaline). If the water is acidic (lower than 7), lime or soda ash is

added to raise the pH. Lime is the more common of the two additives because it is cheap,

but it also adds to the resulting water hardness. Making the water slightly alkaline ensures

that coagulation and flocculation processes work effectively and also helps to minimise

the risk of lead being dissolved from lead pipes and lead solder in pipe fittings. If the

water is alkaline, acid (HCl) or carbon dioxide (CO2) may be added in some

circumstances to lower the pH. Having an alkaline water does not necessarily mean that

lead or copper from the plumbing system will not be dissolved into the water but as a

generally, water with a pH above 7 is much less likely to dissolve heavy metals than a

water with a pH below 7 (http://en.wikipedia.org/wiki/water treatment).

Flocculation

Flocculation is a process which clarifies the water. Clarifying means removing any

turbidity or colour so that the water is clear and colourless. Clarification is done by

causing a precipitate to form in the water which can be removed using simple physical

14
methods. Initially the precipitate forms as very small particles but as the water is gently

stirred, these particles stick together to form bigger particles - this process is sometimes

called flocculation. Many of the small particles that were originally present in the raw

water absorb onto the surface of these small precipitate particles and so get incorporated

into the larger particles that coagulation produces. In this way the coagulated precipitate

takes most of the suspended matter out of the water and is then filtered off, generally by

passing the mixture through a coarse sand filter or sometimes through a mixture of sand

and granulated anthracite (high carbon and low volatiles coal). Coagulants / flocculating

agents that may be used include:

1. Iron (III) hydroxide. This is formed by adding a solution of an iron (III)

compound such as iron (III) chloride to pre-treated water with a pH of 7 or

greater. Iron (III) hydroxide is extremely insoluble and forms even at a pH as low

as 7. Commercial formulations of iron salts were traditionally marketed in the UK

under the name Cuprus.

2. Aluminium hydroxide is also widely used as the flocculating precipitate although

there have been concerns about possible health impacts and mishandling which

led to a severe poisoning incident in 1988 at Camelford in south-west UK when

the coagulant was introduced directly into the holding reservoir of final treated

water.

3. PolyDADMAC is an artificially produced polymer and is one of the class of

synthetic polymers that are now widely used. These polymers have a high

molecular weight and form very stable and readily removed flocs, but tend to be

more expensive in use compared to inorganic materials

(http://en.wikipedia.org/wiki/water treatment).

15
Sedimentation

Water exiting the flocculation basin may enter the sedimentation basin, also called a

clarifier or settling basin. It is a large tank with slow flow, allowing floc to settle to the

bottom. The sedimentation basin is best located close to the flocculation basin so the

transit between does not permit settlement or floc break up. Sedimentation basins can be

in the shape of a rectangle, where water flows from end to another, or circular where flow

is from the centre outward. Sedimentation basin outflow is typically over a weir so only a

thin top layer - furthest from the sediment - exits. The amount of floc that settles out of

the water is dependent on the time the water spends in the basin and the depth of the

basin. The retention time of the water must therefore be balanced against the cost of a

larger basin. The minimum clarifier retention time is normally 4 hours. A deep basin will

allow more floc to settle out than a shallow basin. This is because large particles settle

faster than smaller ones, so large particles bump into and integrate smaller particles as

they settle. In effect, large particles sweep vertically though the basin and clean out

smaller particles on their way to the bottom.

As particles settle to the bottom of the basin a layer of sludge is formed on the floor of the

tank. This layer of sludge must be removed and treated. The amount of sludge that is

generated is significant, often 3%-5% of the total volume of water that is treated. The cost

of treating and disposing of the sludge can be a significant part of the operating cost of a

water treatment plant. The tank may be equipped with mechanical cleaning devices that

continually clean the bottom of the tank or the tank can be taken out of service when the

bottom needs to be cleaned (http://en.wikipedia.org/wiki/water treatment).

16
Filtration

After separating most floc, the water is filtered as the final step to remove remaining

suspended particles and unsettled floc. The most common type of filter is a rapid sand

filter. Water moves vertically through sand which often has a layer of activated carbon or

anthracite coal above the sand. The top layer removes organic compounds, which

contribute to taste and odour. The space between sand particles is larger than the smallest

suspended particles, so simple filtration is not enough. Most particles pass through

surface layers but are trapped in pore spaces or adhere to sand particles. Effective

filtration extends into the depth of the filter. This property of the filter is key to its

operation: if the top layer of sand were to block all the particles, the filter would quickly

clog. To clean the filter, water is passed quickly upward through the filter, opposite the

normal direction (called backflushing or backwashing) to remove embedded particles.

Prior to this, compressed air may be blown up through the bottom of the filter to break up

the compacted filter media to aid the backwashing process; this is known as air scouring.

This contaminated water can be disposed of, along with the sludge from the

sedimentation basin, or it can be recycled by mixing with the raw water entering the

plant.

Some water treatment plants employ pressure filters. These work on the same principle as

rapid gravity filters, differing in that the filter medium is enclosed in a steel vessel and the

water is forced through it under pressure.

Advantages

• Filters out much smaller particles than paper and sand filters can.

• Filters out virtually all particles larger than their specified pore sizes.

17
• They are quite thin and so liquids flow through them fairly rapidly.

• They can be cleaned (back flushed) and reused.

Membrane filters are widely used for filtering both drinking water and sewage (for

reuse). For drinking water, membrane filters can remove virtually all particles larger than

0.2 um--including Giardia and cryptosporidium. Membrane filters are an effective form

of tertiary treatment when it is desired to reuse the water for industry, for limited

domestic purposes, or before discharging the water into a river that is used by towns

further downstream. They are widely used in industry, particularly for beverage

preparation (including bottled water). However no filtration can remove substances that

are actually dissolved in the water such as phosphorus, nitrates and heavy metal ions

(http://en.wikipedia.org/wiki/water treatment).

Disinfection

Disinfection is accomplished both by filtering out harmful microbes and also by adding

disinfectant chemicals in the last step in purifying drinking water. Water is disinfected to

kill any pathogens which pass through the filters. Possible pathogens include viruses,

bacteria, including Escherichia coli, Campylobacter and Shigella, and protozoans,

including Giardia lamblia and other cryptosporidia (http://en.wikipedia.org/wiki/water

treatment).

Chlorination: The most common disinfection method is some form of chlorine or its

compounds such as chloramine or chlorine dioxide. Chlorine is a strong oxidant that

rapidly kills many harmful micro-organisms. Because chlorine is a toxic gas, there is a

danger of a release associated with its use. This problem is avoided by the use of sodium

hypochlorite, which is a relatively inexpensive solution that releases free chlorine when

18
dissolved in water. Chlorine solutions can be generated on site by electrolyzing common

salt solutions. A solid form, calcium hypochlorite exists and releases chlorine on contact

with water. Handling the solid, however, requires greater routine human contact through

opening bags and pouring than the use of gas cylinders or bleach which are more easily

automated. The generation of liquid sodium hypochlorite is both inexpensive and safer

than the use of gas or solid chlorine. All forms of chlorine are widely used despite their

respective drawbacks. One drawback is that chlorine from any source reacts with natural

organic compounds in the water to form potentially harmful chemical by-products

trihalomethanes (THMs) and haloacetic acids (HAAs), both of which are carcinogenic in

large quantities and regulated by the United States Environmental Protection Agency

(EPA). The formation of THMs and haloacetic acids may be minimized by effective

removal of as many organics from the water as possible prior to chlorine addition.

Although chlorine is effective in killing bacteria, it has limited effectiveness against

protozoans that form cysts in water (Giardia lamblia and Cryptosporidium, both of which

are pathogenic), (http://en.wikipedia.org/wiki/water treatment).

19
Figure 3: Flow diagram of a conventional water treatment plant

Source: Hillis, P., ed., (2000). Membrane Technology in Water and Wastewater Treatment.

The Royal Society of Chemistry, Cambridge, UK. 269p.

20
CHAPTER 3

3.0 BREWING PROCESS

3.1 MILLING

The milling process involves the grinding of malted barley grain to produce relatively fine

particles, which are for the most part starch. This is done with the aid of a special German-

built Row mill. This grinds the malt to a fine, husky powder which is then termed grist.

Figure 4: A schematic representation of the milling process

Source: (http://www.crc.dk/flab/brewhous.htm)

The malt is milled into fine husky powder (grist) to increase their surface are and ensure

good access of water to grain particles in the subsequent phase of beer production. Milling

energy is a good indication of malt quality, where homogeneously modified malt has a lower

milling energy. Malt may be supplemented with solid adjunct, i.e. a sugar source such as

flaked or roasted barley (grain bill), in order to impart specific flavour or colour

characteristics to the finished beer (http://www.crc.dk/flab/brewhous.htm).

21
Figure 5: The brewery process chart

Source: (http://www.ibdasiapac.com.au/brewing)

22
3.2 MASHING

Mashing is the process of combining a mix of milled grain (typically malted barley with

supplementary grains such as corn, sorghum, rye or wheat), known as the "grain bill", and

water, known as "liquor", and heating this mixture up with rests at certain temperatures

(notably 45°C, 62°C and 73°C) to allow the enzymes in the malt to break down the starch in

the grain into sugars, typically maltose to create a malty liquid called wort.

There are basically two types of mashing:

1. Infusion mashing,

2. Decoction mashing.

Infusion mashing: Most breweries use infusion mashing, in which the mash is heated

directly to go from rest temperature to another. Some infusion mashes achieve temperature

changes by adding hot water, and there are also breweries that do single-step infusion,

performing only one rest before lautering.

Decoction mashing: Decoction mashing is where a proportion of the grains are boiled and

then returned to the mash, raising the temperature. This boiling process extracts more starch

from the grain by breaking down the cell walls of the grain.

This can be classified into one-, two-, and three-step decoctions, depending on how many

times part of the mash is drawn off to be boiled (http://en.wikipedia.org/wiki/mashing).

3.2.1 Mashing-in

Mixing of the strike water, water used for mashing in, and milled grist must be done in such

a way as to minimize clumping and oxygen uptake. Traditionally this was done by first

adding water to the mash vessel, and then introducing the grist from the top of the vessel in a

23
thin stream. This unfortunately led to a lot of oxygen absorption, and loss of flour dust to the

surrounding air. A premasher, which mixes the grist with mash-in temperature water while it

is still in the delivery tube, reduces oxygen uptake and prevents dust from being lost.

Mashing in is typically done between 35 °C and 45 °C (95 °F and 113 °F), but for single-

step infusion mashes mashing in must be done between 62 °C and 67 °C (143.6 °F and

152.6 °F) for amylases to break down the grain's starch into sugars. The weight-to-weight

ratio of strike water and grain varies from 1:2 for dark beers in single-step infusions to 1:4 or

even 1:5, ratios more suitable for light-coloured beers and decoction mashing, where much

mash water is boiled off (http://en.wikipedia.org/wiki/mashing).

3.2.2 Enzymatic rests

In step-infusion and decoction mashing, the mash is heated to different temperatures, at

which specific enzymes work optimally. Table 2 shows the optimal temperature enzymes

brewers most pay attention to, and what material those enzymes break down. There is some

contention in the brewing industry as to just what the optimal temperature is for these

enzymes, as it is often very dependent on the pH of the mash, and its thickness. A thicker

mash acts as a buffer for the enzymes. Once a step is passed, the enzymes active in that step

are denatured, and become permanently inactive. The time between rests is preferably as

short as possible, but if the temperature is raised more than 1 °C per minute, enzymes may

be prematurely denatured in the transition layer near heating elements.

β-glucanase rest: β-glucan is a chain of the beta isomer of glucose molecules, and found

mainly in the cell walls of plants, and in this context is also known as cellulose. A β-

glucanase rest done at 40 °C is practised in order to break down cell walls and make starches

more available, thus raising the extraction efficiency. Should the brewer let this rest go on

24
too long, it is possible that a large amount of β-glucan will dissolve into the mash, which can

lead to a stuck mash on brew day, and cause filtration problems later in beer production.

Table 2: Rest temperature for major mashing enzymes.

Optimal rest temperatures for major mashing enzymes


Temp °C Temp °F Enzyme Breaks down
40 °C 104.0 °F β-Glucanase β-Glucan
50 °C 122.0 °F Protease Protein
62 °C 143.6 °F β-Amylase Starch
72 °C 161.6 °F α-Amylase Starch

Source: (http://en.wikipedia.org/wiki/mashing)

Protease rest: Protein degradation via a proteolytic rest plays many roles: production of

free-amino nitrogen (FAN) for yeast nutrition, freeing of small proteins from larger proteins

for foam stability in the finished product, and reduction of haze-causing proteins for easier

filtration and increased beer clarity. In all-malt beers, the malt already provides enough

protein for good head retention, and the brewer needs to worry more about more FAN being

produced than the yeast can metabolize, leading to off flavours. The haze causing proteins

are also more prevalent in all-malt beers, and the brewer must strike a balance between

breaking down these proteins, and limiting FAN production.

Amylase rests: The amylase rests are responsible for the production of free fermentable and

non-fermentable sugar from starch in a mash.

25
Starch is an enormous molecule made up of branching chains of glucose molecules. β-

amylase breaks down these chains from the end molecules forming links of two glucose

molecules, i.e. maltose. β-amylase cannot break down the branch points, although some help

is found here through low α-amylase activity and enzymes such as limit dextrinase. The

maltose will be the yeast's main food source during fermentation. During this rest starches

also cluster together forming visible bodies in the mash. This clustering eases the lautering

process.

The α-amylase rest is also known as the saccharification rest, because during this rest the α-

amylase breaks down the starches from the inside, and starts cutting off links of glucose one

to four glucose molecules in length. The longer glucose chains, sometimes called dextrins or

maltodextrins, along with the remaining branched chains, give body and fullness to the beer.

Because of the closeness in temperatures of peak activity of α-amylase and β-amylase, the

two rests are often performed at once, with the exact temperature of the rest determining the

ratio of fermentable to non-fermentable sugars in the wort and hence the final sweetness of

the fermented drink; a hotter rest also a fuller-bodied, sweeter beer as α-amylase produces

more unfermentable sugars. 66 °C is a typical rest temperature for a pale ale or German

pilsner, while Bohemian pilsner and mild ale are rested more typically at 67-68 °C. This is

referred to as the sacchrification rest.

Decoction "rests": In decoction mashing, part of the mash is taken out of the mash tun and

placed in a cooker, where it is boiled for a period of time. This caramelizes some of the

sugars, giving the beer a deeper flavour and colour, and frees more starches from the grain,

making for a more efficient extraction from the grains. The portion drawn off for decoction

is calculated so that the next rest temperature is reached by simply putting the boiled portion

back into the mash tun. Before drawing off for decoction, the mash is allowed to settle a bit,

26
and the thicker part is typically taken out for decoction, as the enzymes have dissolved in the

liquid, and the starches to be freed are in the grains, not the liquid. This thick mash is then

boiled for around 15 minutes, and returned to the mash tun.

The mash cooker used in decoction should not be allowed to scorch the mash, but

maintaining a uniform temperature in the mash is not a priority. To prevent a scorching of

the grains, the brewer must continuously stir the decoction and apply a slow heating.

A Decoction mash brings out a higher malt profile from the grains and is typically used in

Bocks or Doppelbock style beers (http://en.wikipedia.org/wiki/mashing).

3.2.3 Mash-out

After the enzyme rests, the mash is raised to its mash out temperature. This frees up about

2% more starch, and makes the mash less viscous, allowing the lauter to process

faster. It would be nice to raise the mash to 100 °C for mash out and have a much

less viscous liquid, but α-Amylase quickly denatures above 78 °C and any

starches extracted above this temperature cannot be broken down and will cause a

starch haze in the finished product, or in larger quantities an unpleasantly harsh

flavour can evolve. Therefore the mash out temperature rarely exceeds 78 °C.

If the lauter tun is a separate vessel from the mash tun, the mash is transferred to the lauter

tun at this time. If the brewery has a combination mash-lauter tun, the agitator is stopped

after mash-out temperature is reached and the mash has mixed enough to ensure a uniform

temperature (http://en.wikipedia.org/wiki/mashing).

27
Figure 6: Mashing Graph

Source: http://www.crc.dk/flab/mashing.htm

1 Mashing-in: mixing of malt and water

2 Protein pause: release of peptides and amino acids

3 Sugar pause: release of maltose and dextrin

4 Mashing-off: degradation of residual starch, and inactivation of enzymes

Principal mashing enzymes include (1-3, 1-4)-β-glucanase and xylanase for cell wall

degradation, endo-peptidase and carboxypeptidase for protein degradation; and amylases,

limit dextrinase and α-glucosidase for starch degradation

(http://www.crc.dk/flab/mashing.htm).

3.3 LAUTERING

Lautering is a process in brewing beer in which the mash is separated into the clear liquid

wort and the residual grain or spent grain.

28
Lautering usually consists of 3 steps: mash-out, recirculation, and sparging.

1. Mash-out is the term for raising the temperature of the mash to 170°F (77°C). This

both stop the enzymatic conversion of starches to fermentable sugars, and makes the mash

and wort more fluid. Mash-out is considered especially necessary if there is less than 1.5

quarts of water per pound of grain (3 litres of water per kilogram of grain), or if the grain is

more than 25% wheat or oats. The mash-out step can be done by using external heat, or

simply by adding hot water.

2. Recirculation consists of drawing off wort from the bottom of the mash, and adding it

to the top. Lauter tubs typically have slotted bottoms to assist in the filtration process.

The mash itself functions much as a sand filter to capture mash debris and proteins.

This step is monitored by use of a turbidimeter to measure solids in the wort liquid by

their opacity.

3. Sparging is trickling water through the grain to extract sugars from the grain. This is a

delicate step, as the wrong temperature or pH will extract tannins from the chaff (grain

husks) as well, resulting in a bitter brew. Typically, 50% more water is used for

sparging than was originally used for mashing. Sparging is typically conducted in a

lauter-tun.

English sparging drains the wort completely from the mash, after which more water is

added, held for a while at 170°F and then drained again. The second draining can be used

in making a lighter-bodied low-alcohol beer known as Small Beer, or can be added to the

first draining. Some home brewers use English sparging, except that the second batch of

water is only held long enough for the grain bed to settle, after which recirculation and

draining occurs.

29
German sparging, which is used by commercial breweries and many home brewers,

uses continuous process sparging. When the wort reaches a desired level (typically about

an inch) above the grain-bed, water is added at the same slow rate that wort is being

drained. The wort gradually becomes weaker and weaker, and at a certain point, they stop

adding water. This results in greater yields (http://en.wikipedia.org/wiki/lautering).

3.4 BOILING

Boiling the malt extracts, called wort, ensures its sterility, and thus prevents a lot of

infections. During the boil hops are added, which contribute bitterness, flavour, and

aroma compounds to the beer, and, along with the heat of the boil, causes proteins in the

wort to coagulate and the pH of the wort to fall. Finally, the vapours produced during the

boil volatilise off flavours, including dimethyl sulphide precursors.

The boil must be conducted so that is it even and intense. The boil lasts between 50 and

120 minutes, depending on its intensity, the hop addition schedule, and volume of wort

the brewer expects to evaporate.

The simplest boil kettles are direct-fired, with a burner underneath. These can produce a

vigorous and favourable boil, but are also appropriate to scorch the wort where the flame

touches the kettle, causing caramelization and making clean up difficult.

Most breweries use a steam-fired kettle, which uses steam jackets in the kettle to boil the

wort. The steam is delivered under pressure by an external boiler.

State-of-the-art breweries today use many interesting boiling methods, all of which

achieve a more intense boiling and a more complete realisation of the goals of boiling.

30
Many breweries have a boiling unit outside of the kettle, sometimes called a calandria,

through which wort is pumped. The unit is usually a tall, thin cylinder, with many tubes

upwards through it. These tubes provide an enormous surface area on which vapour

bubbles can nucleate, and thus provides for excellent volitization. The total volume of

wort is circulated seven to twelve times an hour through this external boiler, ensuring that

the wort is evenly boiled by the end of the boil. The wort is then boiled in the kettle at

atmospheric pressure, and through careful control the inlets and outlets on the external

boiler, an overpressure can be achieved in the external boiler, raising the boiling point by

a few Celsius degrees. Upon return to the boil kettle, a vigorous vaporization occurs. The

higher temperature due to increased vaporization can reduce boil times up to 30%.

External boilers were originally designed to improve performance of kettles which did

not provide adequate boiling effect, but have since been adopted by the industry as a sole

means of boiling wort.

Modern brew houses can also be equipped with internal calandria, which requires no

pump. It works on basically the same principle as external units, but relies on convection

to move wort through the boiler. This can prevent over-boiling, as a deflector above the

boiler reduces foaming, and also reduces evaporation. Internal calandria are generally

difficult to clean (http://www.beerbrewing.info/beer+brewing/13).

3.5 WHIRLPOOLING

At the end of the boil, the wort is set into a whirlpool. The so-called teacup effect forces

the more dense solids (coagulated proteins, vegetable matter from hops) into a cone in the

centre of the whirlpool tank.

31
In most large breweries, there is a separate tank for whirlpooling. These tanks have a

large diameter to encourage settling, a flat bottom, a tangential inlet near the bottom of

the whirlpool, and an outlet on the bottom near the outer edge of the whirlpool. A

whirlpool should have no internal protrusions that might slow down the rotation of the

liquid. The bottom of the whirlpool is often slightly sloped towards the outlet. Newer

whirlpools often have "Denk rings" suspended in the middle of the whirlpool. These rings

are aligned horizontally and have about 75% of the diameter of the whirlpool. The Denk

rings prevent the formation of secondary eddies in the whirlpool, encouraging the

formation of a cohesive trub cone in the middle of the whirlpool. Smaller breweries often

use the brewkettle as a whirlpool. In the United Kingdom, it is common practice to use a

device known as a hopback to clear the green wort (green wort is wort to which yeast has

not yet been added). This device has the same effect as, but operates in a completely

different manner than, a whirlpool. The two devices are often confused but are in

function, quite different. While a whirlpool functions through the use of centrifugal

forces, a hopback uses a layer of fresh hop flowers in a confined space to act as a filter

bed to remove trub. Furthermore, while a whirlpool is only useful for the removal of

pelleted hops (as flowers don't tend to separate as easily), hopbacks are generally used

only for the removal of whole flower hops, as the particles left by pellets tend to make it

through the hopback (http://www.beerbrewing.info/beer+brewing/13).

3.6 WORT COOLING

After the whirlpool, the wort must be brought down to fermentation temperatures before

yeast is added. In modern breweries this is achieved through a plate heat exchanger. A

plate heat exchanger has many ridged plates, which form two separate paths. The wort is

pumped into the heat exchanger, and goes through every other gap between the plates.

32
The cooling medium, usually water, goes through the other gaps. The ridges in the plates

ensure turbulent flow. A good heat exchanger can drop 95 °C wort to 20 °C while

warming the cooling medium from about 10 °C to 80 °C. The last few plates often use a

cooling medium which can be cooled to below the freezing point, which allows a finer

control over the wort-out temperature, and also enables cooling to around 10 °C. After

cooling, oxygen is often dissolved into the wort to revitalize the yeast and aid its

reproduction (http://www.beerbrewing.info/beer+brewing/13).

3.7 FERMENTATION

Fermentation, as a step in the brewing process, starts as soon as yeast is added to the

cooled wort. This is also the point at which the product is first called beer. It is during this

stage that sugars won from the malt are metabolized into alcohol and carbon dioxide.

During the primary fermentation (H), the fermentable sugars, mainly maltose and glucose

are converted to ethanol and carbon dioxide. This action is performed by the brewing

yeast, which during the brewing process also produces many of the characteristic aroma

compounds found in beer. At the end of the primary fermentation, the yeast cells

flocculate and sediment at the bottom of the fermenter and can be cropped and used for a

new fermentation. Not all yeast cells sediment; some will remain in suspension, and these

cells are responsible for maturation of the beer. During this process the off-flavour,

diacetyl is degraded to below the taste threshold. The fermentation characteristics of

brewer's yeast are strain-dependent and are genetically inherited. Much of the genetics of

Saccharomyces yeasts has been elucidated, and the knowledge gained, forms the basis for

breeding of brewing yeast. Thus, new types of beer with altered aromas can be produced

with yeast strains selected through breeding (http://www.crc.dk/flab/fermentation.htm).

33
There are two main categories of beer here, “bottom-fermentation yeast” beers,

which are fermented at a low temperature (5 to 10°C) with a yeast that sinks to the

bottom of the beer, and “top-fermentation yeast” beers, fermented at 15 to 25°C, with a

yeast which rises to the surface of the beer after fermentation.

The former originated in central Europe and spread to the rest of the world. The second

are mainly brewed in England. Approximately half of the top yeast beer comes from

Belgium. In France, the top yeast method is rare. The difference in taste and flavour

between the two methods of production is very distinct.

The bottom yeast (Uvarum saccharomyces), having little flavour and a fairly neutral

taste, lets the flavour and taste of the hops circulate and gives fine beers. This is the yeast

used in classic pilsner beers.

Top yeast (Cerevisiae saccharomyces) is an energetic yeast that reproduces greatly and

only works well at a temperature of more than 15°C. It produces beers with much more

aroma and flavour, appearing lighter and more digestible even when their specific gravity

is very high. This yeast is ideal for specialty beers. It should be noted that top yeast beers

should not be drunk too cool. In both cases, there are many strains available to the

brewer. It is up to him to choose the one that best suits his particular needs.

In summary, fermentation is in two parts.

Part 1

Aerobic (Oxygen is present): This is the initial rapid process where the yeast is doubling

its colony size every 4 hours (Usually 24-48 hours).

34
Part 2

Anaerobic (No oxygen present): Slower activity and the yeast focuses on converting

sugar to alcohol rather that increasing the number of yeast cells (this process can take

from days to weeks depending on the yeast and the recipe).

The overall process of fermentation is to convert glucose sugar (C6H12O6) to alcohol

(CH3CH2OH) and carbon dioxide gas (CO2). The reactions within the yeast to make this

happen are very complex but the overall process is as follows:

C6H12O6 2CH3CH2OH + 2CO2

Sugar Alcohol + Carbondioxide gas

Source: (http://www.crc.dk/flab/fermentation.htm).

3.8 CONDITIONING

When the sugars in the fermenting beer have been almost completely digested, the

fermentation slows down and the yeast starts to settle to the bottom of the tank. At this

stage, the beer is cooled to around freezing, which encourages settling of the yeast, and

causes proteins to coagulate and settle out with the yeast. Unpleasant flavours such as

phenolic compounds become insoluble in the cold beer, and the beer's flavour becomes

smoother. During this time pressure is maintained on the tanks to prevent the beer from

going flat.

Often, the beer is then racked (siphoned) into another container, usually a carboy, for

aging or secondary fermentation. Fermentation is virtually complete, so the term

secondary fermentation actually refers to conditioning. Use of a hydrometer is

recommended to be absolutely sure all fermentation is finished; this is especially

35
important as a precaution when the beer is to be bottled. Racking is done to separate the

beer from the trub so that the remaining active yeasts do not consume it, as this can give

the beer an off-flavour. Racking also helps separate the beer from sediment, making it

less likely to find its way into the finished product. During secondary fermentation some

chemical by-products from the primary fermentation are digested, which considerably

improves the taste. Secondary fermentation can take from 2 to 4 weeks, sometimes

longer, depending on the type of beer. Additionally lagers, at this point, are aged at near

freezing temperatures for 1-6 months depending on style. This cold aging serves to

reduce sulphur compounds produced by the bottom-fermenting yeast and to produce a

cleaner tasting final product with fewer esters.

If the fermentation tanks have cooling jackets on them, as opposed to the whole

fermentation cellar being cooled, conditioning can take place in the same tank as

fermentation. Otherwise separate tanks (in a separate cellar) must be employed. This is

where aging occurs (http://www.beerbrewing.info/beer+brewing/13).

3.9 MATURATION

The fermented beer initially at a temperature of 130C is reset to 1500C and left for about

12-24hours before resetting again to 500C in preparation for cropping. During transfer

through chillers at a temperature of -1.500C, there carbonation and finings is injected to

act as clarifying agent. It is allowed to rest for about 3-5-days before filtration. The vessel

is normally purged of the excess carried over yeast.

36
3.10 FILTERING

Filtering the beer stabilizes the flavour, and gives beer its polished shine and brilliance.

Not all beer is filtered. When tax determination is required by local laws, it is typically

done at this stage in a calibrated tank.

Filters come in many types. Many use pre-made filtration media such as sheets or

candles, while others use a fine powder made of, for example, diatomaceous earth, also

called kieselguhr, which is introduced into the beer and recirculated past screens to form

a filtration bed.

Filters range from rough filters that remove much of the yeast and any solids (e.g. hops,

grain particles) left in the beer, to filters tight enough to strain colour from the beer.

Normally used filtration ratings are divided into rough, fine and sterile. Rough filtration

leaves some cloudiness in the beer, but it is noticeably clearer than unfiltered beer. Fine

filtration gives a glass of beer that you could read a newspaper through, with no

noticeable cloudiness. Finally, as its name implies, sterile filtration is fine enough that

almost all microorganisms in the beer are removed during the filtration process

(http://www.beerbrewing.info/beer+brewing/13).

3.10.1 Sheet (pad) filters

These filters use pre-made media and are relatively straightforward. The sheets are

manufactured to allow only particles smaller than a given size through, and the brewer is

free to choose how finely to filter the beer. The sheets are placed into the filtering frame,

sterilized (with hot water, for example) and then used to filter the beer. The sheets can be

flushed if the filter becomes blocked, and usually the sheets are disposable and are

37
replaced between filtration sessions. Often the sheets contain powdered filtration media

to aid in filtration.

It should be kept in mind that pre-made filters have two sides. One with loose holes, and

the other with tight holes. Flow goes from the side with loose holes to the side with the

tight holes, with the intent that large particles get stuck in the large holes while leaving

enough room around the particles and filter medium for smaller particles to go through

and get stuck in tighter holes.

Sheets are sold in nominal ratings, and typically 90% of particles larger than the nominal

rating are caught by the shee(http://www.beerbrewing.info/beer+brewing/13).

3.10.2 Kieselguhr filters

Filters that use a powder medium are considerably more complicated to operate, but can

filter much more beer before needing to be regenerated.

Common media include diatomaceous earth, or kieselguhr, and perlite

(http://www.beerbrewing.info/beer+brewing/13).

3.11 PACKAGING

Packaging is putting the beer into the containers in which it will leave the brewery.

Typically this means in bottles, aluminium cans and kegs, but it might include bulk tanks

for high-volume customers (http://www.beerbrewing.info/beer+brewing/13).

38
CHAPTER 4

4.1 FLAVOUR CHANGE IN BREWING

Many flavour active compounds present in uninfected beer are capable of changing their

levels during storage in the final package. Compounds may

(a) Decrease in level leading to flavour deterioration by loss of a desirable character

(b) Increase in level leading to flavour deterioration by an increase in an undesirable

character.

In turn, category “b” compounds may arise

(i) because they are produced de novo in a chemical reaction

(ii) by the release of pre-formed material that is bound up in the beer with a “ holding

agent ” that prevents their flavour from being expressed

(iii) because conditions have changed in a beer which makes the likelihood of either type

of change [(i) or (ii)] more likely, for example a change in redox conditions.

The relevant chemistry underpinning the changes described in (a), (b), (i), (ii) and (iii) is as

follows, remembering that, while all of these reactions are feasible and at some time or

another have been proposed as contributors to flavour change, they may have varying

degrees of actual relevance.

In passing we might note that there has been an over-emphasis on the compound E-2-

nonenal in the literature. To imply that a solitary compound is primarily responsible for

ageing is naive (Bamforth , 1985).

39
4.2 An evaluation of processes from barley to beer in the context of flavour

instability

Table 3: Process impacts on flavour instability (derived from Bamforth, 2004a)

40
Table 3

41
Table 3

42
4.3 Beer components that influence foam quality

Much research activity has been devoted to determining the key beer components that

influence beer foam quality. In the past many of these investigations have applied

reductionist scientific principles to identify these key foam determinants, in the hope of

identifying one component to manipulate to optimize foam quality. Combined, these

studies have established that foam quality, essentially its stability, is promoted by

interactions between proteins (5 kDa) contributed by malt and hop acids ( Asano and

Hashimoto, 1980 ; Bamforth, 1985 ). The proteins identified have included protein Z,

LTP1 and various hordein-derived species. In addition to these widely reported species,

proteomic techniques with mass spectrometric evaluation were recently applied to detect

a range of other proteins species in twice-foamed foam (2x foam, Hao et al., 2006 ),

although these analyses did not ascertain if these proteins were significant beer foam

components. Various divalent metal cations have also been shown to effectively cross-

link hop iso- α -acids to strengthen the bubble film. Components such as polyphenols and

non-starch polysaccharides i.e. β - glucan and arabinoxylan have been attributed potential

minor roles in promoting foam quality. Other components such as lipids, basic amino

acids and high levels of ethanol have been shown to destabilize beer foam. The attributes

and contributions of these “yin and yang” components have recently been

comprehensively reviewed by Evans and Sheehan (2002). On the whole it would appear

that the combined contributions of the promoting components such as proteins are

incremental but the destabilizing effects of lipids can be disastrous to foam quality.

To achieve these evaluations, physiochemical measurements of individual components

have been made to determine each constituent’s foam promoting ability. These

assessments include hydrophobicity ( Slack and Bamforth, 1983 ; Onishi and Proudlove,

1994 ; Yokoi et al., 1994 ), fluorescence recovery after photo bleaching ( Clark, 1991 ;

43
Hughes and Wilde, 1997 ), surface dilational rheology ( Douma et al., 1997 ; Hughes and

Wilde, 1997 ), and surface- viscometric activity ( Maeda et al., 1991 ; Yokoi et al., 1989).

Of these characteristics it is now considered that protein hydrophobicity is probably of

most importance (Bamforth, 1999). However, Evans and Sheehan (2002) concluded,

“these physical considerations are only, at best, a useful guide to the foam promoting

behaviour of beer components.” This is because the numbers of potential foam active

components are certainly numerous and their interactions are complex. Therefore the

practical value of each individual foam component can only be resolved in experiments

where their impact on foam quality is evaluated in the context of the whole beer system.

In the last decade there have been a number of investigations that have adopted this more

holistic approach (Evans et al., 1999c; Evans et al., 2003; van Nierop et al., 2004; Kapp

and Bamforth, 2002; Brey et al., 2002; Lewis and Lewis, 2003; Bamforth and Kanauchi,

2003; Bamforth, 2004a) to determine the interactions and interplay between components

that produce optimal foam quality.

For all these components to influence beer foam quality, they must be resilient and able

to survive the relatively challenging malting and brewing process conditions which

confronts them with heat and enzymatic denatureation/modification, substantial changes

in pH, stabilization treatments (haze, pasteurization) and the potential to be utilized or

adsorbed by the fermenting yeast.

4.3.1 Hop Acids and Foam Stability

Hop acids not only contribute to beer bitterness but are also essential partners with

proteins to achieve foam stability. To impart bitterness the hop acids require

isomerisation traditionally achieved during wort boiling or by catalytic isomerisation of

hop extracts. A further modification that may also be undertaken is to hydrogenate the

hop acid, primarily to achieve protection against the formation of light struck flavours.

44
This final modified form is often commonly referred to by its initial trademark name,

“tetra hop.”

The foam promoting properties of hop acids have long been known and that hop resins, in

particular isohumulone, promote foam stability (for review of in the bubble wall has been

attributed to the ionized form of iso- α -acids crosslinking foam proteins (Asano and

Hashimoto, 1976). It has been observed that isohumulone was more foam promoting than

isocohumulone (Diffor et al., 1978). Ono et al. (1983) observed that isohumulones were

foam concentrated to a greater extent than their less hydrophobic counterparts’

isocohumulones. Combined with the finding that more hydrophobic proteins are foam

promoting ( Slack and Bamforth, 1983 ; Onishi and Proudlove, 1994 ; Yokoi et al.,

1994), this supports Roberts ’ (1976) conclusion that the interaction between hops was

based on either hydrogen bonding or hydrophobic interactions. As such, the more

hydrophobic hydrogenated iso- α -acid hop products would be expected to be more foam

promoting. More recently Lusk et al., (2001a, b) found that tetrahydroisoalpha acids bind

with LTP1 in a molar ratio of 23:1. On the basis of protein structural analysis the binding

of hop acids was concluded to be the result of both ionic (interactions with basic amino

acids) and hydrophobic amino acids. Other natural but minor hop components such as

Structure and structural modification of humulone.

45
adprehumulone, dihydrohumulone and xanthohumol have been suggested to confer

increased benefits over isohumulone for promoting foam stability and lacing (Smith et

al., 1998; Wilson et al., 1999 ). The impact of hop addition is governed by the hop form

and the method by which foam quality is measured. Worts were made from two malt

samples (variety A and B) and fermented by the small scale brewing procedure ( _800

ml, 20 BU hop extract added, Stewart et al., 1998 ). As the hopping level increases, both

foam stability (Rudin) and lacing increases. Interestingly, the magnitude of the increase

across the range of hop concentrations is substantially greater for lacing ( _ 300%) than

for the stability tests (Rudin _ 150%). This observation demonstrates that hop iso- α

-acids are of particular importance in determining the ability of beer foams to lace.

4.4 Manipulating the brewing process to optimize foam quality

The preceding discussion and evaluation has summarized the current understanding of

what components in beer are important and could potentially be manipulated to optimize

foam quality. The emphasis in the statement is on the word “optimize.” In the beginning

of this chapter, customer expectations for foam quality were placed at the fore because

they are the brewers’ objective for brand placement. This can be a complicated

assignment because customer preferences differ not only on nationality but also on

gender and regional lines (Bamforth, 2000a; Smythe et al., 2002; Roza et al., 2006).

Perhaps this is a tall order in this current age of brewery consolidation and global beer

brands.

In reductionist science terms, one is often left with the impression from the literature that

the objective is to maximize (or minimize) the parameter in question, particularly with

foam quality. This somewhat black and white view is rather limiting. For instance if the

objective was to produce highly stable and foam-able beer, the blinkered choice would be

46
to use Fusarium infected malt. This would produce very foam-able, gushing beer that

would perhaps initially be exciting and entertaining but would soon be viewed as rather

pointless as half the package would reside on the ceiling in extreme cases, not to mention

the undesirable health side effects of the attendant mycotoxins. So generally rather than

wanting either of the extremes, that being more or less foam, some subtlety is required by

the brewer to meet their customers’ desires. This optimization of beer foam quality will

ideally require that a certain amount of foam to be produced, of a certain stability, with an

appealing colour (degree of whiteness), degree of creaminess, which during the course of

consumption will leave the desired amount of lacing so as to best satisfy the consumer.

Fortunately, there are a number of opportunities for the brewer to manipulate foam

quality by the selection of raw materials, modifying the brewing process, along with

palliative choices for using foam enhancing additives and devices (Bamforth , 1985 ).

4.5 Styling in Beer

Beer style is a term used to differentiate and categorize beers by various factors such as

colour, flavour, strength, ingredients, production method, recipe, history, or origin

(http://en.wikipedia.org/wiki/beer style). In 1989, Fred Eckhardt furthered Jackson's work

publishing The Essentials of Beer Style (Eckhardt, Fred (1993). The study of what

constitutes a beer's style can be broken down into various elements. These may include

the amount of bitterness imparted to a beer from bittering agents such as hops, roasted

barley, or herbs; the amount of sweetness from the sugar present in the beer; the strength

of the beer from the amount of fermentable material converted into alcohol; the

smoothness or viscosity of the beer in the mouth, commonly described as mouth-feel; and

the appearance of the beer, including the colour.

47
4.5.1 Elements of beer style

Appearance

The visual characteristics that may be observed in a beer are colour, clarity, and nature of

the head. Colour is usually imparted by the malts used, notably the adjunct malts added to

darker beers, though other ingredients may contribute to the colour of some styles such as

fruit beers. Colour intensity can be measured by systems such as EBC, SRM or

Lovibond, but this information is rarely given to the public.

Many beers are transparent, but some beers, such as hefeweizen, may be cloudy due to

the presence of yeast making them translucent. A third variety is the opaque or near-

opaque colour that exists with stouts, porters, schwarzbiers (black beer) and other deeply

coloured styles. Thickness and retention of the head and the lace it can leave on the glass,

are also factors in a beer's appearance.

Aroma

The aroma in a beer may be formed from the malt and other fermentables, the strength

and type of hops, the alcohol, esters, and various other aromatic components that can be

contributed by the yeast strain, and other elements that may derive from the water and the

brewing process.

Flavour

The taste characteristics of a beer may come from the type and amount of malt used,

flavours imparted by the yeast, and strength of bitterness. Bitterness can be measured on

an International Bitterness Units scale, and in North America a number of brewers record

the bitterness on this scale as IBUs.

48
Mouth-feel

The feel of a beer in the mouth, both from thickness of the liquid and from carbonation,

may also be considered as part of a beer's style. A more dextrinous beer feels thicker in

the mouth. The level of carbonation (or nitrogen, in "smooth" beers) varies from one beer

style to another. For some beers it may give the beer a thick and creamy feel, while for

others it contributes a prickly sensation.

Strength

The strength of beer is a general term for the amount of alcohol present. It can be

quantified either indirectly by measurement of specific gravity, or more directly by

determining the overall percentage of alcohol in the beer.

Gravity

Measurement of the specific gravity of the beer has traditionally been used to estimate the

strength of beer by measuring its density. Historically, several different scales have been

used for the measurement of gravity, including the Plato, Baumé, Balling, and Brix

scales, with the Plato scale being the most common contemporary measure.

This approach relies on the fact that dissolved sugars and alcohol each affect the density

of beer differently. Since sugars are converted to alcohol during the process of

fermentation, gravity can be used to estimate the final alcohol. In beer brewing, a

distinction is made between the original gravity, the gravity of the wort before

fermentation has begun, and the final gravity of the product when fermentation is

completed. Since the concentration of sugars is directly proportional to the gravity, the

original gravity gives a brewer an idea of the potential alcoholic strength of the final

49
product. After fermentation, the differences between the final and original gravities

indicates the amount of sugar converted into alcohol, allowing the concentration of

alcoholic strength to be calculated (http://en.wikipedia.org/wiki/beer style).

4.5.2 Beer Styles

Most beer styles fall into broad types roughly according to the time and temperature of

the primary fermentation and the variety of yeast used during fermentation.

There are four main families of beer styles determined by the variety of yeast used in

their brewing.

Ale (top-fermenting yeasts)

Ale is beer that is brewed using only top-fermenting yeasts, and is typically fermented at

higher temperatures than lager beer (15–23°C, 60–75°F). At these temperatures, ale

yeasts produce significant amounts of esters and other secondary flavours and aromas,

often resembling those of apple, pear, pineapple, grass, hay, banana, plum or prune.

Principal styles of ale include Barley Wine, Belgian Trippel, Belgian Dubbel, Altbier,

Bitter, Amber Ale, Brown Ale, Pale Ale, Kölsch, Porter, Stout, and Wheat

beer(http://en.wikipedia.org/wiki/beer style).

Lager (bottom-fermenting yeasts)

Pale lagers are the most commonly consumed type of beer in the world. Lagers are of

Central European origin, taking their name from the German lagern ("to store"). Lager

yeast is a bottom-fermenting yeast, and typically begins fermentation at 7-12°C (45-55°F)

(the "fermentation phase"), and then stored at 0-4°C (30-40°F) (the "lagering phase").

50
During the secondary stage, the lager clears and mellows. The cooler conditions also

inhibit the natural production of esters and other byproducts, resulting in a "crisper"

tasting beer.

With modern improved fermentation control, most lager breweries use only short periods

of cold storage, typically 1–3 weeks.

Most of today's lager is based on the original Pilsner style, pioneered in 1842 in the town

of Pilsen (Plzeň), in an area of the Austrian monarchy now located in the Czech Republic.

The modern pale lager that developed from Pilsner is light in colour and high in forced

carbonation, with an alcohol content of 3–6% by volume. Principal styles of lager include

American-style lager, Bock, Dunkel, Helles, Oktoberfestbier/Märzen, Pilsner,

Schwarzbier and Vienna lager (http://en.wikipedia.org/wiki/beer style).

Beers of Spontaneous Fermentation (wild yeasts)

These beers are nowadays primarily only brewed around Brussels, Belgium. They are

fermented by means of wild yeast strains that live in a part of the Zenne river which flows

through Brussels. These beers are also called Lambic beers. However with the advent of

yeast banks and the National Collection of Yeast Cultures, brewing these beers, although

not through spontaneous fermentation, is possible anywhere

(http://www.beerbrewing.info/beer+brewing/13).

Beers of mixed origin

These beers are blends of spontaneous fermentation beers and ales or lagers or they are

ales/lagers which are also fermented by wild yeasts

(http://www.beerbrewing.info/beer+brewing/13).

51
REFERENCE

Asano , K. and Hashimoto , N. ( 1980 ) Isolation and characterization of foaming proteins

of beer . Journal of the American Society of Brewing Chemists, 38, 129 – 137.

Bamforth , C. W. ( 1985 ) The foaming properties of beer . Journal of the Institute of

Brewing, 91, 370 – 383.

Bamforth , C. W. ( 1999 ) Bringing matters to a head: The status of research on beer foam .

European Brewing Convention Monograph, XXVII, Amsterdam , 10 – 23 .

Bamforth , C. W. ( 2000 a ) Perceptions of beer foam . Journal of the Institute of Brewing ,

106 , 229 – 238.

Bamforth , C. W. ( 2000 b ) Beer quality series: Foam . Brewers ’ Guardian , 129 ( 3 ) , 40 –

43 .

Bamforth , C. W. ( 2004 a ) Fresh controversy: Conflicting opinions on beer staling .

Proceedings of the 28th Convention of the Institute and Guild of Brewing, Asia-Pacific

Section, 63 – 73.

Diffor, D. W. , Lickens , S. T. , Rehberger , A. J. and Burkhardt , R. J. ( 1978 ) T he effect of

isohumulone/isocohumulone ratio on beer head retention . Journal of the American Society

of Brewing Chemists , 36 , 63 – 65 .

Eckhardt, Fred (1993). The Essentials of Beer Style: A Catalog of Classic Beer Styles for

Brewers and Beer Enthusiasts. Fred Eckhardt Communications. ISBN 978-0960630271.

52
Evans , D. E. , MacLeod , L. C. and Lance , R. C. M. ( 1995 ) The importance of protein Z

to the quality of barley and malt for brewing . Proceedings of the European Brewery

Convention Congress, Brussels , 25 , 225 – 232 .

Evans , D. E. and Hejgaard , J . ( 1999 ) The impact of malt derived proteins on beer foam

quality. Part I. The effect of germination and kilning on the level of protein Z4, protein Z7

and LTP1 . J ournal of the Institute of Brewing , 105 , 159 – 169 .

Evans, D. E., Nischwitz, R., Stewart, D. C., Cole, N. and MacLeod, L. C. (1999a) The

influence of malt foam-positive proteins and non-starch polysaccharides on beer foam

quality. European Brewing Convention Monograph , XXVII , Amsterdam, 114–128.

Evans , D. E. and Sheehan , M. C. ( 2002 ) Do not be fobbed off, the substance of beer foam,

a review . Journal of the American Society of Brewing Chemists, 60, 47 – 57.

Hillis, P., ed., (2000). Membrane Technology in Water and Wastewater Treatment. The

Royal Society of Chemistry, Cambridge, UK. 269p.

http://beerbrewing.info/beer+brewing/13

http://en.wikipedia.org/wiki/beer style

http://en.wikipedia.org/wiki/Hops

http://en.wikipedia.org/wiki/lautering

http://en.wikipedia.org/wiki/mashing

http://www.crc.dk/flab/brewhous.htm

http://www.crc.dk/flab/fermentation.htm

53
http://www.crc.dk/flab/malting

http://www.crc.dk/flab/malting barley

http://www.crc.dk/flab/mashing.htm

http://www.crc.dk/flab/water.htm

http://www.ibdasiapac.com.au/brewing/

http://www.scienceclarified.com/Bi-Ca/Brewing

Mallevialle, J. et al., eds. (1994). Water Treatment Membrane Processes. McGraw-Hill,

New York, NY.

Onishi , A. , Canterranne , E. , Clarke , D. J. and Proudlove , M. O. ( 1995 ) Barley lipid

binding proteins: Their role in beer stabilization . Proceedings of the European Brewery

Convention Congress, Brussels , 25 , 553 – 560 .

Ono , M. , Hashimoto , S. , Kakudo , Y. , Nagami , K. and Kumada , J. ( 1983 ) Foaming

and beer flavour . Journal of the American Society of Brewing Chemists , 41 , 19 – 23 .

Roza , J. R. , Wallin , C. E. and Bamforth , C. W. ( 2006 ) A comparison between

instrumental measurement of head retention/lacing and perceived foam quality . Master

Brewers Association of the Americas Technical Quarterly , 43 , 173 – 176 .

Slack, P. T. and Bamforth, C. W. ( 1983) The fractionation of polypeptides from barley and

beer by hydrophobic interaction chromatography: The infl uence of their hydrophobicity on

foam stability . Journal of the Institute of Brewing , 89 , 397 – 401 .

54
Smythe , J. E. , O’Mahony , M. A. and Bamforth , C. W. ( 2002 ) The impact of appearance

of beer on its perception . Journal of the Institute of Brewing , 108 , 37 – 42 .

Stewart , D. C. , Hawthorne , D. and Evans , D. E. ( 1998 ) Cold sterile fi ltration: A small

scale filtration test and investigation of membrane plugging . Journal of the Institute of

Brewing , 104 , 321 – 326.

www.yobrew.co.uk 1999-2008

55
TABLE OF CONTENTS

CHAPTER 1....................................................................................................... 1

1.0 INTRODUCTION.......................................................................................1

1.1 HISTORY OF BEER..................................................................................1

CHAPTER 2....................................................................................................... 3

2.1 BREWING RAW MATERIALS ...................................................................3

2.1.1 Barley..............................................................................................3

2.1.2 Water..............................................................................................7

2.1.3 Hops................................................................................................8

2.1.4 Sugar.............................................................................................10

2.1.5 Finings (Copper or kettle finings)..................................................10

2.2 ESSENTIAL PRE-BREWERY OPERATIONS..............................................11

2.2.1 Malting..........................................................................................11

2.2.2 Water treatment plant..................................................................13

Flocculation..............................................................................................14

Sedimentation..........................................................................................16

Filtration...................................................................................................17

Disinfection..............................................................................................18

CHAPTER 3..................................................................................................... 21

3.0 BREWING PROCESS...............................................................................21

3.1 MILLING ................................................................................................21

3.2 MASHING...............................................................................................23

3.2.1 Mashing-in ................................................................................23

3.2.2 Enzymatic rests.............................................................................24

3.2.3 Mash-out.......................................................................................27

After the enzyme rests, the mash is raised to its mash out temperature.
This frees up about 2% more starch, and makes the mash less viscous,
allowing the lauter to process faster. It would be nice to raise the mash to
100 °C for mash out and have a much less viscous liquid, but α-Amylase
quickly denatures above 78 °C and any starches extracted above this

56
temperature cannot be broken down and will cause a starch haze in the
finished product, or in larger quantities an unpleasantly harsh flavour can
evolve. Therefore the mash out temperature rarely exceeds 78 °C........27

3.3 LAUTERING..........................................................................................28

3.4 BOILING.................................................................................................30

3.5 WHIRLPOOLING...................................................................................31

3.6 WORT COOLING...................................................................................32

3.7 FERMENTATION...................................................................................33

3.8 CONDITIONING....................................................................................35

3.9 MATURATION.......................................................................................36

3.10 FILTERING..........................................................................................37

3.10.1 Sheet (pad) filters.......................................................................37

3.10.2 Kieselguhr filters.........................................................................38

3.11 PACKAGING.......................................................................................38

CHAPTER 4..................................................................................................... 39

4.1 FLAVOUR CHANGE IN BREWING..........................................................39

4.2 An evaluation of processes from barley to beer in the context of


flavour instability.........................................................................................40

4.3 Beer components that influence foam quality.....................................43

4.3.1 Hop Acids and Foam Stability........................................................44

4.4 Manipulating the brewing process to optimize foam quality...............46

4.5 Styling in Beer.....................................................................................47

4.5.1 Elements of beer style..................................................................48

Appearance..............................................................................................48

Aroma...................................................................................................... 48

Flavour..................................................................................................... 48

Mouth-feel................................................................................................49

Strength...................................................................................................49

4.5.2 Beer Styles....................................................................................50

57
REFERENCE.................................................................................................... 52

TABLE OF CONTENTS......................................................................................56

58

You might also like