Você está na página 1de 16

CH08CH08-Bamforth ARI 8 March 2017 10:41

V I E W Review in Advance first posted online


E on March 12, 2017. (Changes may
R

still occur before final publication

S
online and in print.)

C E
I N

N
A
D V A

Progress in Brewing Science


and Beer Production
C.W. Bamforth
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

Department of Food Science and Technology, University of California, Davis, California 95616;
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

email: cwbamforth@ucdavis.edu

Annu. Rev. Chem. Biomol. Eng. 2017. 8:8.1–8.16 Keywords


The Annual Review of Chemical and Biomolecular beer, brewing, tradition, alternative paradigms, quality
Engineering is online at chembioeng.annualreviews.org

https://doi.org/10.1146/annurev-chembioeng- Abstract
060816-101450
The brewing of beer is an ancient biotechnology, the unit processes of which
Copyright  c 2017 by Annual Reviews. have not changed in hundreds of years. Equally, scientific study within the
All rights reserved
brewing industry not only has ensured that modern beer making is highly
controlled, leading to highly consistent, high-quality, healthful beverages,
but also has informed many other fermentation-based industries.

8.1

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

HISTORY AND OVERVIEW


The brewing of beer is an ancient practice of somewhere between 6,000 and 8,000 years (1). Still
considered by some to owe more to art than science, it is undeniably science driven, relying on
fundamentals such as pH, kinetic analysis of enzymes, use of pure strains in fermentation, and
statistical tools emerging from research conducted within brewing laboratories (2). Particularly
since the latter decades of the nineteenth century, brewing has advanced to become a finely
controlled technology in which consistent beer excellence can be ensured, thereby rendering
products that are reassuringly predictable, unlike, perhaps, the vagaries of wine quality (3).
Despite the enormous depths of understanding of the brewing process, it is nevertheless a fact
that the fundamental shape of the brewing process has not changed in millennia. Were one to be
time-transported to the Middle Ages, one would find fundamentally the same unit operations that
continue to be employed in the brewing of beer. There is an inherent reluctance to fundamentally
change the status quo in this respect, in large part because the very nature of beer is because of
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

the classic raw materials and processes and not despite them (4, 5).
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

Briefly—and see Reference 2 for fuller detail, also Figure 1—most beers are produced from
grain that has been malted, in which process the cereal is steeped in water to trigger metabolism,
then germinated over a limited period to synthesize enzymes that will degrade cell wall and protein
components in the sprouting, thereby softening the grain and making it more readily millable.
Also synthesized are the enzymes that will degrade the main food reserve, starch, in the brewhouse.
At the end of the germination phase, there is a kilning stage to drive off moisture, thereby halting

Barley
Water Steeping (1–2 days; 14–18°C)
Heat Germination (3–7 days; 16–20°C)
Kilning (24 h; 50–110°C)

Malt

Storage (3–4 weeks, ambient)

Stored malt
Water
Milling (ambient)
Heat Mashing (ca. 1 h; 45–72°C)
Starch adjuncts Wort separation (1–3 h; 73–80°C)

Sweet wort
Hops
Sugar adjuncts Boiling (1–3 h; 100°C)
Clarification (10 min–1 h; 90–100°C)
Heat

Pitching wort
Oxygen
Yeast Fermentation and maturation
Refrigeration (3–14 days; 6–25°C)
Cold conditioning
Stabilizers (3–21 days; –2–0°C)
Filter aids Stabilization, filtration, packaging
(<24 h; ca. 0°C)
Packaging materials
Heat (if pasteurizing) Beer
Figure 1
An overview of malting and brewing.

8.2 Bamforth

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

the growth of the grain while at the same time modifying the color and flavor of the grain to
the advantage of the beer that will be made from the malt. After several weeks of storage, the
malt is milled to produce more readily extractable particles, and the milled grist is soaked in hot
water to allow the starch granules to gelatinize and become amenable to digestion by the amylases.
Then the liquid (wort) is drained from the spent grains with the aid of sparging water, with the
collected sweet liquid being boiled, traditionally with hops to extract bitterness and aroma. After
clarification and cooling, the wort is pitched with yeast and allowed to ferment for several days,
during which process the sugars are converted into ethanol and carbon dioxide, alongside various
other flavor changes. Thereafter, there is a removal of yeast, subjection of the green beer to cold
and other stabilization treatments, clarification, and packaging.
Viewed dispassionately, this time-honored process lacks logic. With its multiple inputs of water
and demand for this often precious commodity for heating, cooling, and cleaning, as well as other
demands on energy and the need to deal with coproducts (spent grains, surplus yeast, carbon
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

dioxide), there must be a simpler way. Russell et al. (6) calculated the theoretical environmental
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

and cost impacts of processes of increasing radicalization and confirmed the savings potential;
meanwhile, Heymann et al. (7) examined the potential for an ersatz beer synthesized from an
alcoholic base and flavorings and determined that at least one population of consumers placed
little store on whether a product is made in traditional ways. Most brewers would be unconvinced
by these approaches, at least until such time as flavor technology has advanced to allow the
production of tried and trusted beer styles, even individual brands (8, 9). Furthermore, the simple
expedient of creating product by mixing quality-determining ingredients with a bland alcoholic
base would be prohibitively expensive in most locations, as the products would be taxed as spirits
rather than the much cheaper beer. Indeed, this is the very reason why certain long-flavored
alcoholic beverages (formerly called alcopops) are made from a beer-type grist involving brewing
technology.
Tax is a major driving force for technological change in the brewing industry. Another example
would be the production of beverages (Happoshu, third category) in Japan from grists of <25%
and zero malt, respectively, on the basis that they are taxed at drastically lower rates than those
beers in which malted barley is the main grist component (10). For mainstream beer brands, the
major cost components apart from taxation are personnel costs, packaging and distribution, and
sales and marketing (11). In some ways, this reinforces the loyalty of brewers to long-standing
processes, as it does not make sense to jeopardize products by trying to take shortcuts on raw
materials and time-honored strategies.
One approach that has been periodically revisited for many years is the replacement of malt
with raw barley, with the use of microbial enzymes in place of the ones that would otherwise be
made in malting (12). Although the logic is clear, in terms of process savings and environmental
friendliness, few extant beer brands could be made with such a switch, as the quality of the malt
makes an incontrovertible contribution to the nature of the beer.

BARLEY
Breeding programs, monitored closely by the brewing industry through collaborative organiza-
tions such as the American Malting Barley Association, continue to develop new malting-quality
barleys, with ongoing targets including increased yields of extractable material; enhanced agro-
nomic characteristics, including straw strength; and reduced susceptibility to pests and diseases
(http://ambainc.org/). Of particular concern to brewers is the mold Fusarium, which produces
undesirable health risks, such as deoxynivalenol as well as hydrophobin, a small protein that induces
the spontaneous foaming of beer, a phenomenon known as gushing (13).

www.annualreviews.org • Brewing Science and Beer Production 8.3

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

In recent years, mutant barleys have been developed, including low-proanthocyanidin types
deficient in polyphenols that contribute to haze formation in beer (14), as well as barleys unable
to synthesize the enzyme lipoxygenase, which may contribute to staling in beer (15). There is
inherent resistance within the brewing industry to the use of genetically modified materials, and
although certain desirable traits might conceivably be introduced into barley that would benefit
the brewer and the beer, in the current climate it is unlikely that they would be used (16).
Fundamental studies on the composition of barley in relation to changes occurring during
malting and brewing have reached an advanced stage. Compositionally, much work has been de-
voted to the cell walls of the starchy endosperm, which are primarily composed of a mixed-linkage
β-glucan and arabinoxylan, with small amounts of ferulic acid and acetic acid. A model of the wall
structure has been proposed in which there is a preponderance of glucan in the inner regions and
of arabinoxylan on the periphery (17). This explains the relative solubility of these entities and also
the advantage of xylanases in allowing better access by both solvents and glucanases to the molecule
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

that can have several disadvantageous impacts on the brewing process and beer quality (18).
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

The primary protein reservoir in barley is in the form of hordeins of various types. Efficient
solubilization and degradation of these molecules are of great relevance for rendering the starch
that is embedded in them amenable to attack, as well as for lessening their tendency to cause
haze in beer and for the production of the amino acids required for yeast metabolism (19). These
proteins are also the source of the peptide sequences to which gluten-intolerant individuals are
sensitive (20). The balance of hordein degradation products and proteins from within the albumin
fraction, such as protein Z and lipid transfer protein, has a major impact on the stability of beer
foam (21).
Surprisingly, there has been relatively less work on starch in the context of brewing than has
been dedicated to the cell wall and proteins (22). It would be reasonable to expect that future
research might focus on the differences in gelatinization properties between varieties of barley
and the causes of these differences. In respect to the enzymology of starch degradation, most
effort has been devoted to understanding the properties of limit dextrinase and the reasons why it
acts to only an extremely limited extent during mashing, meaning that the alpha and beta amylases
alone function to render 80% of the starch in a fermentable form (23). It appears that the limit
dextrinase is synthesized during germination in a bound and inactive form. This can be released
in an active condition by lowering the pH (24).
Barley is 3% lipid, but most of this is lost with insoluble fractions generated during the brewing
process, notably the spent grains and the hot and cold breaks generated during wort boiling and
cooling, respectively (25). That which remains is generally considered to be undesirable from the
perspective of foam and flavor stability.
The husk fraction of the grain is of prime significance as a filter bed during most traditional
wort-collection processes. However, its high levels of polyphenols and silicate mean that it is a
source of materials that can decrease the stability of beer, although those high levels of silicate
that find their way into beer from the grain have potential health benefits (26).

WATER
Although vastly more water is used in growing crops (barley, wheat, adjuncts, and hops) and in
the malting process than is used in brewing, brewing is nonetheless perceived as being a water-
intensive process (27–29). Nowadays, it is entirely feasible to have a water-usage ratio (liters of
water per liters of packaged beer) of 3:1 or even higher in operations within the brewery walls.
Water entering the brewery is generally processed to render it free from taints and to ensure it
possesses the composition deemed appropriate for the beers that are being produced. This ability

8.4 Bamforth

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

to dial in the desired water composition is critical for the matching of a single brand produced
in different global locations, rendering rather foolish the concept of terroir. Water will also be
processed to remove hardness prior to use in steam-generation systems and will be deaerated
when used for various operations in the brewery, notably, the adjustment of alcohol content in
high-gravity brewing. Hydrophobic membrane–based systems are available for such operations,
as they are also for adjusting the gas content of finished beer (30). Wastewater is also increasingly
treated by brewers on site through aerobic and anaerobic digestion systems, the latter of which
offers opportunities for energy recovery (31). Indeed, diverse systems generally are being used in
breweries to maximize energy recovery, including wind and solar power.

HOPS
Hops remain the main spice for beer, although an ever-widening range of other materials is
apparently being used to add novelty to beer choice. Just as for barley, breeding of new hop
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

varieties is ongoing (32), with aroma contributions being a particular goal, as well as increased
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

resin levels in those varieties primarily intended to provide bitterness. Hops are quite an intensive
crop to grow, not least for their tendency to spoil. There is a particular interest in low-trellis hops,
which are easier to cultivate and harvest (33). As for cereals used in brewing, there is equal interest
in the possibilities of organically grown hops (34). Some brewers insist that it is only the whole hop
cone that provides character to the beer, preferring hops that have not even been dried (which is
essential for longer-term storage of the crop). But it is normal for hops to receive some degree of
processing to render them easier to handle, and in particular to enhance the recovery of bitterness,
which is lost throughout the process, especially from native cones (35). Most brewers use pelletized
hops of various types, though some use hops extracted with liquid carbon dioxide. These can be
subfractionated into resin and aroma extracts, allowing flexibility in the introduction of bitterness
separate from aroma. The resin fraction can be isomerized outside the brewery, adding bitterness
to the finished product; still further, such isomerized extracts can be reduced by using hydrogen
over a palladium catalyst to produce bitterness molecules that are no longer converted by light to
3-methyl-2-butene-1-thiol, which delivers a skunky aroma to beer exposed to sunlight. This is why
most brewers prefer to use brown glass bottles, though marketing pressure to use green or clear
glass packages means that some (though not all) brewers use these reduced side-chain preparations.

YEAST AND OTHER ORGANISMS


The two main brewing yeasts remain Saccharomyces cerevisiae (ale yeast) and Saccharomyces pastori-
anus (lager yeast) (36). The latter has undergone several name changes and is the more complicated
organism, seemingly having arisen by happenstance in the relatively recent past via a melding of
S. cerevisiae and Saccharomyces eubayanus (37). There are vastly more ale strains than lager strains,
largely because the former can be isolated from natural sources. There is much interest in selection
of yeasts that will afford different character to beers, although for most strains the nature of the
flavor-active substances produced is similar, albeit in different proportions. Relatively few brewing
strains produce distinct flavorsome entities; perhaps the best examples are the strains of S. cerevisiae
that are used in the production of hefeweizen products, which produce very high levels of isoamy-
lacetate, as well as the relatively unusual 4-vinylguiaicol, which affords the clove or medicinal
character associated with this beer style (38). This substance is produced through the decarboxy-
lation of ferulic acid by an enzyme that is coded for by a gene in this type of ale yeast that is not
found in other ale yeasts or lager yeasts, though it may be found in certain so-called wild yeasts.
The search for flavor diversity has led brewers to explore alternative organisms (39). There is
nothing new in this; for example, Brettanomyces was a yeast named in honor of the British, who

www.annualreviews.org • Brewing Science and Beer Production 8.5

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

employed it in the maturation of many of their beers, with the attendant production of aromas
generally referred to as barnyard. This organism and very many more yeasts and bacteria are
spontaneously involved in the production of wild beers, with the best known being Belgium’s
lambic and gueuze products (40).
The search for diversity or intended process goals has also led to the pursuit of mutants of
Saccharomyces deficient in certain undesirable traits, such as the increased ability to produce sulfur
dioxide (41) and decreased ability to produce dimethyl sulfide (42). Just as for barley, the ambitions
for genetically modified brewing yeasts have been thwarted by brewers’ inherent reluctance to use
such constructs, even though they have been produced successfully and even approved for use.

MALT HOUSE OPERATIONS


The malting process remains a somewhat traditional one of steeping, germination, and kilning.
Fundamental changes to the process, such as the use of abrasion and gibberellic acid to stimulate
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

and homogenize germination, were introduced many years ago, but even to this day they are
by no means universally applied (43, 44). Modifications to the kilning process to prevent the
development of nitrosamines are also a study and an application of many years’ duration (45). A
little more recently, changes have been proposed including the seeding of lactic acid bacteria to
outgrow undesirable microorganisms (46), as well as advances in the understanding of the impact
of malting conditions on malt flavor (47).

BREWHOUSE OPERATIONS
Technology for beer brewing evolves organically rather than radically (48). Malt must be stored
for several weeks prior to use; otherwise, solid-liquid-separation operations in the brewery are
jeopardized. It has been suggested that during this storage there is a decrease in the level in the
malt of an enzyme(s) promoting the bridging of high–molecular weight molecules that form a
framework that impedes liquid flow. Candidates for this enzyme are lipoxygenase (49) and thiol
oxidase (50).
Apart from brewers operating on a very large scale, most brewers grind the malt using two-
roll mills, sometimes with prior steam treatment of the malt to make the husk more pliable or
even to the extent of so-called wet milling, where the malt is soaked for several minutes prior
to grinding (48). To optimize the particle size distribution in the milled grist, it is customary
for larger production concerns to use six-roll mills, in which three pairs of rollers are set with
different gap settings. The aim in any roller milling operation is to keep the husk as intact as
possible, which forms a filter bed in the wort separation system, while ensuring that the starch
endosperm is crumbled to flour and finer grits to maximize hydration possibilities and extraction
of enzymes and substrates to allow enzymolysis.
There is increasing interest, however, in wort-separation systems that do not depend on the
husk, notably the mash filter. When such a device is installed, it makes sense to pulverize the malt
as much as possible to maximize extraction of materials, which is achieved using a hammer mill.
Of course, the increased extraction of materials from very small particles includes those that are
undesirable as well as those that are desirable.
Modern brewhouses are designed to minimize air ingress, as oxygen is of concern as a source
of enzymic browning but also because it may lessen the flavor stability of beer, a claim that is
refuted by some (51). Nonetheless, reduction in air ingress is generally achieved through the
use of intimate mixing stems; bottom entry into vessels; and, for some, the deaeration of milled
grist and/or brewing water. The advent of mash mixers with steam jackets and rousers (agitators)

8.6 Bamforth

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

rendered unnecessary the historic approaches to changing the temperature of mashes—with a


low-temperature stand to complete the degradation of cell wall materials and protein [although
there are those who argue that proteolysis is minimal in mashing (52)], followed by a higher
temperature sufficient to gelatinize starch while allowing enzyme survival. However, as in most
aspects of the malting and brewing process, there are adherents to older approaches, in this case,
the so-called decoction approach, whereby portions of the mash are removed and boiled before
reintroducing to raise the temperature.
Use of very well modified malt means it should indeed not be strictly necessary to have a low-
temperature commencement to mashing, other than for the production of beers such as hefeweis-
sens, in which the lower temperature is needed for survival of feruloyl esterase, the enzyme that
releases ferulic acid from the cell walls of the grain (38). One strategy to avoid a low-temperature
mash, even for less-well-modified grists, is to employ heat-tolerant enzymes of microbial origin.
But there are those who resist this, as they believe that enzymes endogenous to malt are preferable
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

(53).
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

Although traditional mash tuns, wherein mashing and wort collection are effected in a single
vessel, remain in use in some breweries, most brewers keep these stages separate, with a mash
mixer followed by a lauter tun or mash filter. In both instances, wort collection occurs through
Darcy’s law, with particular emphasis on large surface area (wide lauter tuns, many mash filter
plates) and small bed depth (shallow grain beds in lauter tuns, narrow chambers in mash filters).
Excessive pressure in a lauter tun can lead to blocking of the holes in the platers, whereas pressure
is fundamentally an advantage in the mash filter, allowing squeezing out of wort and thus shorter
turnarounds. Andrews (54) has succinctly compared the relative advantages and disadvantages of
these two systems.
In both instances, the coproduct is spent grains, whose high moisture content demands that
they be shipped as soon as possible. It is uneconomical to dry them, and they are most widely used
as cattle food, although a range of other possibilities have been proposed (55). A radical approach
would be to fractionate the grist premashing, rendering a dry spent grain product, though this has
not attracted commercial consideration (11). There have, however, been efforts through selective
milling approaches to minimize husk contribution to traits such as increased astringency, which
is due to their polyphenol content.
The most thermodynamically demanding stage in brewing is the wort boiling phase, which
is required to sterilize wort; precipitate proteinaceous complexes; drive off unwanted volatiles;
concentrate wort; and, traditionally, extract bitterness from hops. Many have attempted to improve
the efficiency of the boil in terms of energy reduction, without losing sight of its critical importance
for the reasons just stated, all of which requirements are dependent on turbulence (vigor) in the
boil. A range of novel boiling systems have been proposed, some of which replace thermal energy
with gas flow to purge off unwanted flavors (56).
The insoluble complex generated in boiling (hot break or trub) is removed usually using a
whirlpool prior to cooling the wort in a so-called paraflow heat exchanger, in which countercurrent
flow of hot wort and coolant (cold water, propylene glycol) separated by thin metal plates ensures
the rapid transfer of heat. Spontaneously fermented beers, such as lambic, are cooled in a shallow
vessel open to the elements known as a coolship.

FERMENTATION AND MATURATION


Controlled fermentation is the fulcrum about which beer consistency is achieved, depending as
it does on the production of a consistent growth medium for the yeast (e.g., wort), as well as the
employment of the necessary yeast in the correct amounts and in the correct condition. Thus,

www.annualreviews.org • Brewing Science and Beer Production 8.7

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

the main variables in fermentation are yeast quality and quantity; wort composition and strength,
including additions of oxygen and often zinc; temperature; and vessel geometry (36).
Wort composition is dictated by raw material selection and brewhouse operations. Its strength
is controlled by regulating parameters such as water–grist ratio; water sparging; boil length (e.g.,
percent evaporation); and the possible addition of extra sugars, such as high maltose corn syrup,
sucrose, or candi sugar, into the wort at the boiling stage. Wort clarity is also critical, as particles
in the wort form nucleation sites that cause carbon dioxide to exit solution in the fermenter as
bubbles and deliver a convective force in the vessel, which tends to lead to better mixing and thus
faster formation (57). This clarity must be consistent between batches to ensure a consistent rate
of fermentation and thus consistent beer flavor. Although certain yeast foods are available, in the
shape of proprietary mixtures of vitamins, amino acids, and minerals, in reality the only material
likely to be limiting is zinc, and this is frequently added by brewers. Yeast strains differ in their
requirement for oxygen for the synthesis of unsaturated fatty acids and sterols that are significant
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

components of the membranes of the cell. Sufficient oxygen must be provided to allow the yeast
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

to multiply to allow the desired fermentation profile and avoid so-called stuck fermentations.
However, excessive oxygen leads to overgrowth of yeast, with a shift in the balance or alcohol-
to-yeast yield [which is significant in large-scale production (58)], as well as a shift in the flavor
spectrum of the product, notably toward lower levels of esters.
On a larger production scale, many brewers use so-called high-gravity brewing (59). The
wort produced for pitching is of higher strength (original extract) than required for the brew in
question, but following fermentation and conditioning, the beer is brought to the target alcohol
concentration by adding deaerated water. The driving force is a maximization of the throughput in
fermenters. Excessive original extracts must be avoided (e.g., >16–18o Plato), as yeast encountering
very high sugar concentrations will produce disproportionate amounts of esters.
Yeast continues to be used for the most part in a liquid slurry form, although there have
been proposals to adopt dried yeast (60). Provided the concentration of alcohol developed in a
fermentation is not excessive [e.g., less than 6–7% alcohol by volume (ABV)], yeast produced
in fermentation can generally be reused. Some companies have pursued such a transfer over
many successive fermentations, but yeast properties unavoidably shift over time, and stressed cells
develop that perform poorly and display a tendency toward autolysis and the release of undesirable
materials into the beer, notably proteinase A, which hydrolyses foam-active polypeptides (61).
Many brewers will switch to newly propagated yeast after 5–10 successive fermentations.
Whether new or not, the yeast pitched into a new fermentation must be viable; this is rou-
tinely evaluated by staining with methylene blue, although there has been much interest in newer
approaches to assessing yeast viability (62). There is also interest in assessing the so-called vitality
of yeast, namely, the health and vigor of living yeast, though as yet none of the proposed meth-
ods is routinely applied (63). Traditional approaches to measuring yeast concentration, notably
the hemocytometer, are often replaced these days by instrumental approaches. Perhaps the best
known are the devices that measure yeast capacitance (64), which are now available for inline use
alongside other inline measuring probes employed through the brewery, such as those involved
in the assay of turbidity, pH, oxygen, carbon dioxide, ethanol, and specific gravity (65).
The physical behavior of yeast is a critical consideration in fermentation. The phenomenon
of flocculation, wherein yeast cells agglomerate to produce flocs that either rise or fall in the
fermenter, is highly significant (66). Yeast must not separate out from suspension prematurely;
if this occurs, fermentation will halt. Equally, the yeast should flocculate once fermentation is
complete, as this renders ensuing clarification processes easier. Flocculation is a characteristic of
yeast strain but can also be influenced by wort composition, most notably the level of calcium.
Certain malts can also deliver a premature yeast flocculation factor to wort (67).

8.8 Bamforth

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

To this day, fermentations are still monitored by the decrease in specific gravity of the wort,
although there have been advocates for other measurements (68). It is also customary to monitor
pH decline.
Much attention is paid to ensuring that the yeast also removes the vicinal diketones (VDKs) that
are spontaneously produced during fermentation and that afford the buttery/popcorn-like aroma
that is anathema to most beer brands (69). This VDK removal is at the heart of beer maturation,
but for many brewers it is actually an integral part of the fermentation process. These molecules,
diacetyl and pentanedione, are produced by the oxidative decarboxylation of acetolactate and
acetohydroxybutyrate, intermediates in the production of amino acids that leak out of the cell. A
sufficiently healthy yeast cell will scavenge the VDKs given enough time. Alternative strategies
(70) for accelerating the handling of the VDKs have included the use of a heating stage, in which
newly fermented beer is centrifuged to remove yeast then heated profoundly to accelerate the
conversion of the precursors into the free VDKs, followed by cooling and the passage of the
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

beer through a column of immobilized yeast, which scavenges the diacetyl and pentanedione. An
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

alternative is the addition of a bacterial enzyme, acetolactate decarboxylase, during fermentation,


which affects decarboxylation as opposed to oxidative decarboxylation (71). The product in the case
of acetolactate is acetoin, a molecule of much less flavor potency than diacetyl and the substance
that yeast first reduces the diacetyl to when that molecule is reabsorbed into the cell.
Nonetheless, some advocate for prolonged maturation of beer, although the evidence for much
worthwhile happening in this time is extremely sparse (72). Surplus yeast is used in some markets
to produce savory extracts, but it is also used as an animal feed (73). The 1960s and 1970s saw
considerable interest in the use of continuous fermentation in breweries, but this has been widely
discontinued, not least because it does not lend itself to operations wherein many different wort
and beer streams are in play (74).

DOWNSTREAM PROCESSING AND PACKAGING


Some traditional beers, such as the cask-conditioned ales in the United Kingdom, continue to
enjoy a relatively rapid route from fermenter to glass. They are essentially racked directly to
barrels with the addition of sugar, which the residual yeast will use to produce additional carbon
dioxide in the package, and isinglass finings, which are collagen digests from the swim bladders
of certain fishes that interact with yeast and other insoluble materials to allow them to settle
according to Stokes’ law (75).
Most other beers move from fermentation/maturation to cold conditioning, which serves the
dual purpose of chilling out cold-sensitive proteinaceous complexes and aiding their subsequent
settling. It is typical to take the beer to at least −1◦ C for this purpose, the alcohol and other
dissolved components lowering the freezing point. The lower the temperature, short of freezing,
the better in terms of material emerging from solution (76). Effective removal of cold-sensitive
materials (notably complexes of proteins linked through polymerized polyphenols) at this stage
reduces the tendency of such entities to throw a haze in the finished beer. However, it is also
possible to remove polyphenols using polyvinyl polypyrollidone and haze-causing proteins using
silica gels, tannic acid, and enzymes such as papain, although the latter lessens foam potential
(77). More recently, prolyl endopeptidase has been advocated for the removal of haze-causing
polypeptides, which are rich in the amino acid proline. This enzyme has the added advantage of
removing the last traces of the peptides that gluten-sensitive individuals are unable to tolerate
(78). Actually, there is a massive decline in the levels of problematic entities throughout the
brewing process (79). Most beers are clarified by filtration, and the most effective power-based
system remains that employing diatomaceous earth as the filter aid, although there has been a shift

www.annualreviews.org • Brewing Science and Beer Production 8.9

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

toward perlite but also toward non-powder-based systems, notably cross-flow, to avoid concerns
with waste material disposal (30).
Bright beer will be adjusted for carbon dioxide content, usually by pinpoint carbonation but
also sometimes via membranes or more traditionally by natural in-package carbonation. In terms
of packaging, which is the most expensive stage of the operation but which largely falls outside
the scope of this review, modern-day small-pack operations may reach rates of 2,500 cans/min or
1,200 bottles/min, with highly controlled fill heights and minimum oxygen levels (80).

QUALITY CONSIDERATIONS IN BEER


Raw material quality and process performance manifest themselves in the perceived quality of
beer.
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

Foam
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

Whereas the quantity of foam produced is a direct consequence of carbon dioxide level in the
beer and the temperature of dispense, the release of carbon dioxide from solution is dependent
on nucleation (81), which can be induced through agitation (e.g., a vigorous pour) or the use of
nucleation devices, e.g., widgets (82) and etched glasses. Stability of the ensuring foam is dependent
on the presence of foam-stabilizing surfactant molecules and the absence of foam-negative entities,
including those present at the point of dispense, e.g., in ill-washed glasses or those containing
residual detergent from the washing process (83, 84). Principle among the foam-positive substances
are amphipathic polypeptides (85), more specifically, the relative proportions of strongly foam-
stabilizing species, such as lipid-transfer protein and those with lesser stabilizing potential, notably
the hydrolysis products of hordeins produced during malting (21). The foaming capacities of
LTP1 appear to be enhanced in the boiling process, presumably through partial denaturation of
the molecules and the exposure of hydrophobic regions from the interior of the molecule (86). The
foaming polypeptides are linked through interactions with the iso-alpha-acid bittering molecules
that are produced through the isomerization of hop alpha acids (87). This interaction, which is
promoted by divalent ions such as zinc, occurs in the foam itself, leading to a rapid increase in
surface varicosity through a solidification of the foam, which in turn results in the foam lacing
the walls of the glass as the beer is consumed (88). Particularly efficacious foaming species may be
produced during malt kilning and roasting (89); such species may include melanoidins produced
in the Maillard reaction, but also complexes of polypeptides with polysaccharides (90). Materials
may be added to beer to further enhance foam stability, including propylene glycol alginate (91)
but also nitrogen gas (92), which greatly boosts foam stability by lowering the bubble collapse rate
through disproportionation (93).
The primary foam-negative material in beer is ethanol, although this is clearly not a variable
and is built into a product specification. At lower levels (e.g., less than 6% ABV), the ethanol is
insufficient to overcome the stronger positive impact of proteins and bitter substances, provided
these are in ample quantities. Thus, a beer made from a relatively low proportion of malt and
substantial amounts of adjunct that do not contribute protein, such beers low in bitterness (e.g.,
malt liquors), is more likely to display poor foam. At higher alcohol concentrations, including the
ludicrously high-alcohol products being produced by freeze-concentration processes, foam can
be severely debilitated (83).
Foam is also destabilized by lipid materials, including those that can emerge through the
brewing process (94). Certain specialty malts contain lipid materials, including oxidized lipids, that
have a pronounced foam-damaging effect (95, 96). It also appears that avoiding low-temperature

8.10 Bamforth

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

commencement of mashing and lowering mash pH to approximately 5.1 improves foam, most
likely through the lesser activity of lipoxygenase, an enzyme capable of catalyzing the oxidation
of unsaturated fatty acids (97).

Gushing
Fundamentally, gushing is caused by either agitation of beer or the presence of nucleation sites
in beer that promote the formation of bubbles, the latter including hydrophobin (98), hop resin
degradation products, and oxalate (84).

Color
Color is afforded to beer via Maillard reaction products generated during malt kilning, as well as
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

through oxidation of polyphenols in the brewhouse (99). Although still not unheard of, there is
relatively little use these days of caramels. The matter of color measurement, namely, the potential
merits of using tristimulus and chromaticity in preference to simple light absorption at a single
wavelength, has perhaps garnered the most attention in recent years (100).

Clarity
Most beers are intended to be served free from turbidity (haze), although there appears to be
a peculiar move to make distinct cloudiness in beer a desirable trait in so-called craft beers.
Fundamentally any macromolecule insufficiently degraded in the malting and brewing processes
can constitute a problem in this respect; thus, proteins complexed with polyphenols, β-glucans,
arabinoxylans, and starch have all been implicated in this context. Strategies for minimizing haze
risks have been reviewed (101).

Flavor
Myriad species contribute to the flavor of beer, contributed from grist materials, water, yeast,
and, for some beers, other microorganisms. They can be categorized as esters, alcohols, acids,
sulfur-containing substances, hydrocarbons, and a myriad of aromatic substances. Their levels are
regulated by careful control of raw materials and processing (102).

Flavor Instability
A perceptible deviation in quantity of any flavorsome component in beer is deemed flavor insta-
bility, although the primary focus of many (misguidedly) has been on avoiding the development
of papery/cardboard notes in beer that arise owing to the production of carbonyl substances such
as E-2-nonenal. The chemistry of the formation of this and other carbonyl-containing substances
has been studied extensively, with unsaturated fatty acids, alcohols, amino acids, and iso-a-acids
among the claimed precursors (51). Since the original proposal that the key triggers of staling are
oxygen radicals (103), much research has been devoted to unraveling the mechanisms involved,
employing tools such as electron spin resonance spectroscopy (104). However, the key strategy
for minimizing flavor change in beer remains the minimization of in-package oxygen levels during
filling and storage of beer at the lowest practical temperature short of freezing.

www.annualreviews.org • Brewing Science and Beer Production 8.11

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

Microbiological Instability
Beer is relatively resistant to spoilage because it represents an unbalanced feedstock for most
organisms, because it has low pH and low levels of oxygen but high levels of carbon dioxide, and
because of the presence of antimicrobial agents, notable ethanol and iso-α-acids. Nonetheless, a
range of microbes will spoil wort and beer (see 40 for review).

Safety and Wholesomeness


For the most part, brewers have always focused on minimizing the level of undesirable molecules
in beer, notably nitrosamines (105). More recently, there has been a willingness to consider the po-
tential health benefits of beer in the way that there has been less reluctance to champion red wine.
Thus, the clear benefit of moderate alcohol consumption in lessening the risk of atherosclerosis
is now linked to ethanol, per se, rather than resveratrol, a molecule that to have any impact would
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

demand vastly more daily consumption than even one bottle of red wine daily (106). Composi-
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

tionally, beer has demonstrably more nutritional value than wine, including silicate (26), which is
claimed to counter osteoporosis (107); soluble fiber (108); prebiotics (109); the assimilable phe-
nolic antioxidant ferulic acid (110); and B vitamins, notably folic acid (111). The myth of the beer
belly has been refuted (112), and any greater perceived healthfulness of wine drinkers over beer
drinkers has been linked to overall lifestyle (113). For a recent review, see Bamforth (114).

DISCLOSURE STATEMENT
The author is not aware of any affiliations, memberships, funding, or financial holdings that might
be perceived as affecting the objectivity of this review.

LITERATURE CITED
1. Hornsey IS. 2003. A History of Beer and Brewing. Cambridge, UK: R. Soc. Chem.
2. Bamforth C. 2009. Beer: Tap into the Art and Science of Brewing. New York: Oxford Univ. Press. 3rd ed.
3. Bamforth C. 2008. Grape Versus Grain. New York: Cambridge Univ. Press
4. Bamforth CW. 2000. Brewing and brewing research: past, present and future. J. Sci. Food Agric.
80:1371–78
5. Bamforth CW, Stewart GG. 2010. Brewing—the transformation of a craft into a technology. Biologist
57:139–47
6. Russell ST, Singh RP, Bamforth CW. 2008. Alternative paradigms for the production of beer. J. Inst.
Brew. 114:349–56
7. Heymann H, Goldberg JR, Wallin CE, Bamforth CW. 2010. A “beer” made from a bland alcohol base.
J. Am. Soc. Brew. Chem. 68:75–76
8. Bamforth CW. 2005. Intensifying malting and brewing—Does the customer care? Brew. Guard.
134(8):30–32
9. Bamforth CW. 2006. Don’t like it—but hard to deny it. Brew. Guard. 135(1):20–22
10. Nakatani K. 2007. Beer in Japan—present and future trends. The story of Happoshu and the “Third
Way.” Brew. Distill. 3(5):54–57
11. Bamforth CW. 2004. Opportunities for newer technologies in the oldest biotechnology, brewing. Appl.
Biotechnol. Food Sci. Policy 1:213–22
12. Goode DL, Wijngaard HH, Arendt EK. 2005. Mashing with unmalted barley—impact of malted barley
and commercial enzyme (Bacillus spp.) additions. Master Brewer Assoc. Am. Tech. Q. 42:184–98
13. Sarlin T, Kivioja T, Kalkkinen N, Linder MB, Nakari-Setälä T. 2012. Identification and characteriza-
tion of gushing-active hydrophobins from Fusarium graminearum and related species. J. Basic Microbiol.
52:184–94

8.12 Bamforth

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

14. Von Wettstein D, Nilan RA, Ahrenst-Larsen B, Erdal K, Ingversen J, et al. 1985. Proanthocyanidin-free
barley for brewing: progress in breeding for high yield and research tool in polyphenol chemistry. Master
Brewer Assoc. Am. Tech. Q. 22:41–52
15. Hirota N, Kaneko T, Kuroda H, Kaneda H, Takashio M, et al. 2005. Characterization of lipoxygenase-1
null mutants in barley. Theor. Appl. Genet. 111:1580–84
16. Hammond JRM. 2016. Biographical review: microscopes, microbes, and manipulation: 35 years in
brewing. J. Am. Soc. Brew. Chem. 74:157–72
17. Bamforth CW, Kanauchi M. 2001. A simple model for the cell wall of the starchy endosperm in barley.
J. Inst. Brew. 107:235–40
18. Scheffler A, Bamforth CW. 2005. Exogenous β-glucanases and pentosanases and their impact on mash-
ing. Enzym. Microb. Technol. 36:813–17
19. Wrobel R, Jones BL. 1992. Appearance of endoproteolytic enzymes during the germination of barley.
Plant Physiol. 100:1508–16
20. Bamforth CW. 2009. Producing gluten-free beer—an overview. In The Science of Gluten-Free Foods and
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

Beverages, ed. EK Arendt, F Dal Bello, pp. 113–17. St. Paul: AACC Int.
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

21. Bamforth CW, Milani C. 2004. The foaming of mixtures of albumin and hordein protein hydrolysates
in model systems. J. Sci. Food Agric. 84:1001–4
22. Bamforth CW. 2003. Barley and malt starch in brewing: a general review. Master Brew. Assoc. Am. Tech.
Q. 40:89–97
23. Stenholm K, Home S. 1999. A new approach to limit dextrinase and its role in mashing. J. Inst. Brew.
105:205–10
24. Heisner CB, Bamforth CW. 2008. Thioredoxin in barley: Could it have a role in releasing limit dextrinase
in brewery mashes? J. Inst. Brew. 114:122–26
25. Anness BJ, Reed RJR. 1985. Lipids in the brewery—a material balance. J. Inst. Brew. 91:82–87
26. Casey TR, Bamforth CW. 2010. Silicon in beer and brewing. J. Sci. Food. Agric. 90:784–88
27. Fillaudeau L, Blanpain-Avet P, Daufin G. 2006. Water, wastewater and waste management in brewing
industries. J. Clean. Prod. 14:463–71
28. Palmer J, Kaminski C. 2013. Water: A Comprehensive Guide for Brewers. Boulder: Brewers
29. Eumann M. 2006. Water in brewing. See Reference 115, pp. 183–207
30. Freeman G. 2006. Filtration and stabilisation of beer. See Reference 115, pp. 275–92
31. Reed R. 2006. Waste handling in the brewing process. See Reference 115, pp. 335–57
32. Henning J. 2006. The breeding of hop. See Reference 115, pp. 102–22
33. Darby P. 2004. Hop growing in England in the twenty first century. J. R. Agric. Soc. Eng. 165.
http://adha.us/sites/default/files/downloads/UK%20Report%20on%20the%20Future%20of%
20Low%20Trellis.pdf
34. Turner SF, Benedict CA, Darby H, Hoagland LA, Simonson P, et al. 2011. Challenges and opportunities
for organic hop production in the United States. Agron. J. 103:1645–54
35. Roberts TR. 2016. Hops. In Brewing Materials and Processes: A Practical Approach to Beer Excellence, ed.
CW Bamforth, pp. 47–75. San Diego, CA: Elsevier
36. Boulton C, Quain D. 2006. Brewing Yeast and Fermentation. Oxford: Wiley-Blackwell
37. Gallone B, Mertens S, Crauwelse S, Lievense B, Verstrepen1 KJ, Steensels J. 2017. Genomics and
evolution of beer yeasts. In Brewing Microbiology: Current Research, Omics and Microbial Ecology, ed. NA
Bokulich, CW Bamforth. Poole, UK: Caister Acad. In press
38. Coghe S, Benoot K, Delvaux F, Vanderhaegen B, Delvaux FR. 2004. Ferulic acid release and 4-
vinylguaiacol formation during brewing and fermentation: indications for feruloyl esterase activity in
Saccharomyces cerevisiae. J. Agric. Food Chem. 52:602–8
39. Bokulich N, Bamforth CW, Mills DA. 2012. Brewhouse resident microbiota are responsible for
multi-stage fermentation of American Coolship Ale. PLOS ONE 7(4):e35507. doi:10.1371/journal.
pone.0035507
40. Bokulich NA, Bamforth CW. 2013. The microbiology of malting and brewing. Microbiol. Mol. Biol. Rev.
77:157–72

www.annualreviews.org • Brewing Science and Beer Production 8.13

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

41. Chen YF, Yang X, Zhang SJ, Wang XQ, Guo CH, et al. 2012. Development of Saccharomyces cerevisiae
producing higher levels of sulfur dioxide and glutathione to improve beer flavor stability. Appl. Biochem.
Biotechnol. 166:402–13
42. Hansen J, Bruun SV, Bech LM, Gjermansen C. 2002. The level of MXR1 gene expression in brewing
yeast during beer fermentation is a major determinant for the concentration of dimethyl sulfide in beer.
FEMS Yeast Res. 2:137–49
43. Briggs DE. 1987. Accelerating malting: a review of some lessons of the past from the United Kingdom.
J. Am. Soc. Brew. Chem. 45:1–8
44. Davies N. 2016. Malts. In Brewing Materials and Processes: A Practical Approach to Beer Excellence, ed. CW
Bamforth, pp. 1–25. San Diego, CA: Elsevier
45. Wainwright T. 1986. Nitrosamines in malt and beer. J. Inst. Brew. 92:73–80
46. Rouse S, van Sinderen D. 2008. Bioprotective potential of lactic acid bacteria in malting and brewing.
J. Food Prot. 71:1724–33
47. Murray JP, Bennett SJE, Chandra GS, Davies NI, Pickles JL. 1999. Sensory analysis of malt. Master
Brew. Assoc. Am. Tech. Q. 36:15–19
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

48. Andrews JMH. 2006. The brewhouse. See Reference 115, pp. 208–27
49. Pöyri S, Mikola M, Sontag-Strohm T, Kaukovirta-Norja A, Home S. 2002. The formation and hydrolysis
of barley malt gel-protein under different mashing conditions. J. Inst. Brew. 108:261–67
50. Bamforth CW, Roza JR, Kanauchi M. 2009. Storage of malt, thiol oxidase and brewhouse performance.
J. Am. Soc. Brew. Chem. 67:89–94
51. Vanderhaegen B, Neven H, Verachtert H, Derdelinckx G. 2006. The chemistry of beer aging—a critical
review. Food Chem. 95:357–81
52. Lewis MJ, Robertson IC, Dankers SU. 1992. Proteolysis in the protein rest of mashing—An appraisal.
Master Brew. Assoc. Am. Tech. Q. 29:117–21
53. Bamforth CW. 2009. Current perspectives on the role of enzymes in brewing. J. Cereal Sci. 50:353–57
54. Andrews JA. 2004. Review of progress in mash separation technology. Master Brew. Assoc. Am. Tech. Q.
41:45–49
55. Aliyu S, Bala M. 2011. Brewer’s spent grain: a review of its potentials and applications. Afr. J. Biotechnol.
10:324–31
56. Bamforth C. 2009. Wizards, volcanoes and beautiful music: developments in wort boiling. Brew. Guard.
138(5):23–24
57. Siebert KJ, Blum PH, Wisk TJ, Stenroos LE, Anklam WJ. 1986. The effect of trub on fermentation.
Master Brew. Assoc. Am. Tech. Q. 23:37–43
58. Bamforth CW, Boulton CA, Clarkson SP, Large PJ. 1988. The effects of oxygen on brewery process
performance. Proc. 20th Conv. Inst. Brew. (Aust. NZ Sect.), Brisbane, pp. 211–19
59. Stewart GG. 2010. High gravity brewing and distilling—past experiences and future prospects. J. Am.
Soc. Brew. Chem. 68:1–9
60. Gosselin Y, Fels S. 1998. Fermentation characteristics from dried ale and lager yeasts. Master Brew. Assoc.
Am. Tech. Q. 35:129–32
61. Cooper DJ, Stewart GG, Bryce JH. 1998. Some reasons why high gravity brewing has a negative effect
on head retention. J. Inst. Brew. 104:83–87
62. Chan LL, Driscoll D, Kuksin D, Saldi S. 2016. Measuring lager and ale yeast viability and vitality using
fluorescence-based image cytometry. Master Brew. Assoc. Am. Tech. Q. 53:49–54
63. Heggart HM, Margaritis A, Pilkington H, Stewart RJ, Dowhanick TM, Russell I. 1999. Factors affecting
yeast viability and vitality characteristics: a review. Master Brew. Assoc. Am. Tech. Q. 36:383–406
64. Carvell JP, Turner K. 2003. New applications and methods utilizing radio-frequency impedance mea-
surements for improving yeast management. Master Brew. Assoc. Am. Tech. Q. 40:30–38
65. Boulton CA. 2012. 125th anniversary review: advances in analytical methodology in brewing. J. Inst.
Brew. 118:255–63
66. Verstrepen KJ, Derdelinckx G, Verachtert H, Delvaux FR. 2003. Yeast flocculation: what brewers should
know. Appl. Microbiol. Biotechnol. 61:197–205
67. Vidgren V, Londesborough J. 2011. 125th anniversary review: yeast flocculation and sedimentation in
brewing. J. Inst. Brew. 117:475–87

8.14 Bamforth

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

68. Daoud IS, Searle BA. 1990. On-line monitoring of brewery fermentation by measurement of CO2
evolution rate. J. Inst. Brew. 96:297–302
69. Krogerus K, Gibson BR. 2013. 125th anniversary review: diacetyl and its control during brewery fer-
mentation. J. Inst. Brew. 119:86–97
70. Inoue T. 2008. Diacetyl in Fermented Foods and Beverages. St. Paul: Am. Soc. Brew. Chem.
71. Rostgaard-Jensen B, Svendsen I, Ottesen M. 1987. Isolation and characterization of an α-acetolactate
decarboxylase useful for accelerated beer maturation. Proc. Eur. Brew. Conv. Cong. Madrid, pp. 393–400
72. Bamforth C. 2002. Great brewing debates: part 4. Does beer get better with ageing? Brew. Guard.
131(10):26–28
73. Tanguler H, Erten H. 2008. Utilisation of spent brewer’s yeast for yeast extract production by autolysis:
the effect of temperature. Food Bioprod. Process. 86:317–21
74. Brányik T, Vicente AA, Dostálek P, Teixeira JA. 2008. A review of flavour formation in continuous beer
fermentations. J. Inst. Brew. 114:3–13
75. Leather RV. 1998. From field to firkin: an integrated approach to beer clarification and quality. J. Inst.
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

Brew. 104:9–18
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

76. Miedl M, Bamforth CW. 2004. The relative importance of temperature and time in the cold conditioning
of beer. J. Am. Soc. Brew. Chem. 62:75–78
77. Rehmanji M, Gopal C, Mola A. 2005. Beer stabilization technology—clearly a matter of choice. Master
Brew. Assoc. Am. Tech. Q. 42:332–38
78. Akeroyd M, van Zandycke S, den Hartog J, Mutsaers J, Edens L, et al. 2016. AN-PEP, proline-specific
endopeptidase, degrades all known immunostimulatory gluten peptides in beer made from barley malt.
J. Am. Soc. Brew. Chem. 74:91–99
79. Guerdrum LJ, Bamforth CW. 2012. Prolamin levels through brewing and the impact of prolyl endopep-
tidase. J. Am. Soc. Brew. Chem. 70:35–38
80. Klimovitz R, Ockert K, eds. 2014. Beer Packaging. St Paul: Master Brew. Assoc. Am. 2nd ed.
81. Lynch DM, Bamforth CW. 2002. Measurement and characterization of bubble nucleation in beer.
J. Food Sci. 67:2696–701
82. Lee WT, McKechnie JS, Devereux MG. 2011. Bubble nucleation in stout beer. Phys. Rev. E 83:1609
83. Evans DE, Bamforth CW. 2009. Beer foam: achieving a suitable head. In Beer: A Quality Perspective, ed.
CW Bamforth, pp. 1–60. Burlington, MA: Academic
84. Lusk LT. 2016. Controlling beer foam and gushing. In Brewing Materials and Processes: A Practical
Approach to Beer Excellence, ed. CW Bamforth, pp. 175–98. San Diego, CA: Elsevier
85. Slack PT, Bamforth CW. 1983. The fractionation of polypeptides from barley and beer by hydrophobic
interaction chromatography: the influence of their hydrophobicity on foam stability. J. Inst. Brew. 89:397–
401
86. Bech LM, Vaag P, Heinemann B, Breddam K. 1995. Throughout the brewing process barley lipid
transfer protein 1 (LTP1) is transformed into a more foam promoting form. Proc. Eur. Brew. Conv.
Congr., Brussels, pp. 561–68. Oxford: IRL
87. Simpson WJ, Hughes PS. 1994. Stabilization of foams by hop-derived bitter acids. Chemical interactions
in beer foam. Cerevisiae Biotechnol. 19:39–44
88. Roza JR, Wallin CE, Bamforth CW. 2006. A comparison between the instrumental measurement of
head retention/lacing and perceived foam quality. Master Brew. Assoc. Am. Tech. Q. 43:173–76
89. Bishop LR. 1977. The nitrogenous complexes of haze and foam and their measurement. J. Inst. Brew.
83:350–55
90. Lusk LT, Goldstein H, Ryder D. 1995. Independent role of beer proteins, melanoidins and polysaccha-
rides in foam formation. J. Am. Soc. Brew. Chem. 53:93–103
91. Jackson G, Roberts RT, Wainwright T. 1980. Mechanism of beer foam stabilization by propylene glycol
alginate. J. Inst. Brew. 86:34–37
92. Carroll TCN. 1979. The effect of dissolved nitrogen gas on beer foam and palate. Master Brew. Assoc.
Am. Tech. Q. 16:116–19
93. Ronteltap AD, Hollemans M, Bisperink CGJ, Prins A. 1991. Beer foam physics. Master Brew. Assoc. Am.
Tech. Q. 28:25–32

www.annualreviews.org • Brewing Science and Beer Production 8.15

Changes may still occur before final publication online and in print
CH08CH08-Bamforth ARI 8 March 2017 10:41

94. Dickie KH, Cann C, Norman EC, Bamforth CW, Muller RE. 2001. Foam-negative materials. J. Am.
Soc. Brew. Chem. 59:17–23
95. Combe AL, Ang JK, Bamforth CW. 2013. Positive and negative impacts of specialty malts on beer foam:
a comparison of various cereal products for their foaming properties. J. Sci. Food Agric. 93:2094–101
96. Ang JK, Bamforth CW. 2014. Foam inhibitors from specialty malts. J. Inst. Brew. 120:193–200
97. Kobayashi N, Kaneda H, Kano Y, Koslimo S. 1994. Behavior of lipid hydroperoxides during mashing.
J. Am. Soc. Brew. Chem. 52:141–45
98. Wessels JGH. 1996. Fungal hydrophobins: proteins that function at an interface. Trends Plant Sci. 1:9–15
99. Shellhammer T, Bamforth CW. 2008. Assessing color quality of beer. In Color Quality of Fresh and
Processed Foods, ed. CA Culver, RE Wrolstad, pp. 192–202. Washington, DC: Am. Chem. Soc.
100. De Lange AJ. 2016. Color. In Brewing Materials and Processes: A Practical Approach to Beer Excellence, ed.
CW Bamforth, pp. 199–249. San Diego, CA: Elsevier
101. Bamforth CW. 1999. Beer haze. J. Am. Soc. Brew. Chem. 57:81–90
102. Bamforth CW. 2014. Practical Guides for Beer Quality: Flavor. St. Paul: Am. Soc. Brew. Chem.
103. Bamforth CW, Parsons R. 1985. New procedures to improve the flavor stability of beer. J. Am. Soc.
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Hacettepe Universitesi on 03/17/17. For personal use only.

Brew. Chem. 43:197–202


104. Uchida M, Suga S, Ono M. 1996. Improvement for oxidative flavor stability of beer—rapid prediction
method for beer flavor stability by electron spin resonance spectroscopy. J. Am. Soc. Brew. Chem. 54:205–
11
105. Tricker AR, Preussmann R. 1991. Volatile and non-volatile nitrosamines in beer. J. Cancer Res. Clin.
Oncol. 117:130–32
106. Gaziano JM, Hennekens CH, Godfried SL, Sesso HD, Glynn RJ, et al. 1999. Type of alcoholic beverage
and risk of myocardial infarction. Am. J. Cardiol. 83:52–57
107. Rico H, Gallego-Lago JL, Hernández ER, Villa LF, Sanchez-Atrio A, et al. 2000. Effect of silicon
supplement on osteopenia induced by ovariectomy in rats. Calcif. Tissue Int. 66:53–55
108. Bamforth CW, Gambill SC. 2007. Fiber and putative prebiotics in beer. J. Am. Soc. Brew. Chem. 65:67–69
109. Kanyer AJ. 2016. The digestive fate of beta-glucan oligosaccharides in beer. MS Thesis, Univ. Calif., Davis
110. Bourne L, Paganga G, Baxter D, Hughes P, Rice-Evans C. 2000. Absorption of ferulic acid from low-
alcohol beer. Free Radic. Res. 32:273–80
111. Owens JE, Clifford AJ, Bamforth CW. 2007. Folate in beer. J. Inst. Brew. 113:243–48
112. Wannamethee SG, Shaper AG, Whincup PH. 2005. Alcohol and adiposity: effects of quantity and type
of drink and time relation with meals. Int. J. Obes. 29:1436–44
113. Galobardes B, Morabia A, Bernstein MS. 2001. Diet and socioeconomic position: Does the use of
different indicators matter? Int. J. Epidemiol. 30:334–40
114. Bamforth C. 2015. “Beer is good for you” as a message in academia. In Ethanol and Education: Alcohol as
a Theme for Teaching Chemistry, ed. R Barth, pp. 113–18. Washington, DC: Am. Chem. Soc.
115. Bamforth CW. 2006. Brewing: New Technologies. Cambridge, UK: Woodhead

8.16 Bamforth

Changes may still occur before final publication online and in print

Você também pode gostar