You are on page 1of 20

International Journal of Food Microbiology 89 (2003) 105 124

www.elsevier.com/locate/ijfoodmicro

Review article

Beer spoilage bacteria and hop resistance


Kanta Sakamoto a,b,*, Wil N. Konings b
b

a
Fundamental Research Laboratory, Asahi Breweries, Ltd. 1-21, Midori 1-chome, Moriya-shi, Ibaraki 302-0106, Japan
Department of Microbiology, Groningen Biomolecular Sciences and Biotechnology Institute, University of Groningen, Kerklaan 30,
9751NN Haren, The Netherlands

Received 4 September 2002; accepted 15 March 2003

Abstract
For brewing industry, beer spoilage bacteria have been problematic for centuries. They include some lactic acid bacteria such
as Lactobacillus brevis, Lactobacillus lindneri and Pediococcus damnosus, and some Gram-negative bacteria such as
Pectinatus cerevisiiphilus, Pectinatus frisingensis and Megasphaera cerevisiae. They can spoil beer by turbidity, acidity and the
production of unfavorable smell such as diacetyl or hydrogen sulfide. For the microbiological control, many advanced
biotechnological techniques such as immunoassay and polymerase chain reaction (PCR) have been applied in place of the
conventional and time-consuming method of incubation on culture media. Subsequently, a method is needed to determine
whether the detected bacterium is capable of growing in beer or not. In lactic acid bacteria, hop resistance is crucial for their
ability to grow in beer. Hop compounds, mainly iso-a-acids in beer, have antibacterial activity against Gram-positive bacteria.
They act as ionophores which dissipate the pH gradient across the cytoplasmic membrane and reduce the proton motive force
(pmf). Consequently, the pmf-dependent nutrient uptake is hampered, resulting in cell death. The hop-resistance mechanisms in
lactic acid bacteria have been investigated. HorA was found to excrete hop compounds in an ATP-dependent manner from the
cell membrane to outer medium. Additionally, increased proton pumping by the membrane bound H+-ATPase contributes to
hop resistance. To energize such ATP-dependent transporters hop-resistant cells contain larger ATP pools than hop-sensitive
cells. Furthermore, a pmf-dependent hop transporter was recently presented. Understanding the hop-resistance mechanisms has
enabled the development of rapid methods to discriminate beer spoilage strains from nonspoilers. The horA-PCR method has
been applied for bacterial control in breweries. Also, a discrimination method was developed based on ATP pool measurement
in lactobacillus cells. However, some potential hop-resistant strains cannot grow in beer unless they have first been exposed to
subinhibitory concentration of hop compounds. The beer spoilage ability of Pectinatus spp. and M. cerevisiae has been poorly
studied. Since all the strains have been reported to be capable of beer spoiling, species identification is sufficient for the
breweries. However, with the current trend of beer flavor (lower alcohol and bitterness), there is the potential risk that not yet
reported bacteria will contribute to beer spoilage. Investigation of the beer spoilage ability of especially Gram-negative bacteria
may be useful to reduce this risk.
D 2003 Elsevier B.V. All rights reserved.
Keywords: Beer; Spoilage; Bacteria; Detection; Hop; Antibacterial; Resistance

* Corresponding author. Fundamental Research Laboratory, Asahi Breweries, Ltd. 1-21, Midori 1-chome, Moriya-shi, Ibaraki 302-0106,
Japan. Tel.: +81-297-46-1504; fax: +81-297-46-1506.
E-mail address: kanta.sakamoto@asahibeer.co.jp (K. Sakamoto).
0168-1605/$ - see front matter D 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0168-1605(03)00153-3

106

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

1. Introduction
Beer has been recognized for hundreds of years as
a safe beverage. It is hard to spoil and has a remarkable microbiological stability. The reason is that beer
is an unfavorable medium for many microorganisms
due to the presence of ethanol (0.5 10% w/w), hop
bitter compounds (approximately 17 55 ppm of isoa-acids), the high content of carbondioxide (approximately 0.5% w/w), the low pH (3.8 4.7), the extremely reduced content of oxygen ( < 0.1 ppm) and
the presence of only traces of nutritive substances
such as glucose, maltose and maltotriose. These latter
carbon sources have been substrates for brewing yeast
during fermentation. As a result, pathogens such as
Salmonellae typhimurium and Staphylococcus aureus
do not grow or survive in beer (Bunker, 1955).
However, in spite of these unfavorable features, a
few microorganisms still manage to grow in beer.
These, so-called beer spoilage microorganisms, can
cause an increase of turbidity and unpleasant sensory
changes of beer. Needless to say that these changes
can affect negatively not only the quality of final
product but also the financial gain of the brewing
companies.
A number of microorganisms have been reported
to be beer spoilage microorganisms, among which
both Gram-positive and Gram-negative bacteria, as
well as so-called wild yeasts. Gram-positive beer
spoilage bacteria include lactic acid bacteria belonging to the genera Lactobacillus and Pediococcus.
They are recognized as the most hazardous bacteria
for breweries since these organisms are responsible
for approximately 70% of the microbial beer-spoilage
incidents (Back, 1994). The second group of beer
spoilage bacteria is Gram-negative bacteria of the
genera Pectinatus and Megasphaera. The roles of
these strictly anaerobic bacteria in beer spoilage have
increased since the improved technology in modern
breweries has resulted in significant reduction of
oxygen content in the final products. Wild yeasts do
cause less serious spoilage problem than bacteria but
are considered a serious nuisance to brewers because
of the difficulty to discriminate them from brewing
yeasts.
Considerable effort has been made by many microbiologist to control microbial contamination in
beer. The most commonly used method today for

detecting beer spoilage bacteria in breweries is still


traditional incubation on culture media. A number of
selective media have been developed since Louis
Pasteur published in 1876 Etudes sur La Biere
(Studies on beer) (Pasteur, 1876). It usually takes a
week or even longer for bacteria to form visible
colonies on plates or to increase the turbidity in
nutrients broths. Consequently, the products are often
already released for sale before the microbiological
results become available. If a beer spoilage bacterium
is then detected and identified in the beer product it
needs to be recalled from the market. This will cause
serious commercial damages to the brewery. Most
microbiologists have focused on developing more
specific and rapid methods for the detection of beer
spoilage bacteria than using traditional culture methods. However, a method is needed to determine
whether the bacteria detected are capable of growing
in beer or not. Up to date, the only available method is
the so-called forcing test in which the detected
bacterium is re-inoculated and incubated in beer. It
usually takes 1 month or even longer to detect visible
turbidity in the inoculated beer, meaning that this
method is not very practical. A rapid method to
predict beer-spoiling ability is therefore urgently
needed and the development of such a method will
be essential for understanding the nature of the beerspoiling ability.
Among the components of beer, hop compounds
have received a lot of attention for reason of their
preservative values and their bitterness. For centuries
it was generally believed that hops protect beer from
infection by most organisms, including pathogens, but
it was only in the 20th century that Shimwell
(1937a,b) showed that hop compounds only inhibit
growth of Gram-positive bacteria and not of Gramnegative bacteria. His findings had a great impact
because many pathogens such as Salmonella species
are Gram-negative bacteria. Feature(s) such as low pH
and alcohol content undoubtedly have a negative
effect on growth of these pathogens in beer. Among
Gram-positive bacteria, some species of lactic acid
bacteria are less sensitive to hop compounds and are
able to grow in beer. Insight in the mechanism of
resistance of lactic acid bacteria to hop compounds is
important for understanding their beer-spoiling ability.
The antibacterial activity of hop compounds and the
hop resistance of lactic acid bacteria have been

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

investigated. On the other hand, little studies have


been done on the beer-spoiling ability of Gramnegative bacteria.
This report reviews the currently available information about beer spoilage bacteria, their growth and
spoilage activity in beer.

Table 1
Beer spoilage bacteria

Gram-positive
bacteria

2. Beer spoilage bacteria


Beer is a poor and rather hostile environment for
most microorganisms. Its ethanol concentration ranges
from 0.5% to 10% (w/w) and is usually around 4 5%.
Beer is usually acid with pHs ranging from pH 3.8 to
4.7, which is lower than most bacteria can tolerate for
growth. Furthermore, the high carbondioxide concentration (approximately 0.5% w/w) and extremely low
oxygen content ( < 0.1 ppm) makes beer a near to
anaerobic medium.
Beer also contains bitter hop compounds (approximately 17 55 ppm of iso-a-acids), which are toxic,
especially for Gram-positive bacteria. The concentrations of nutritive substances, such as carbohydrates and amino acids, are very low since most
have been consumed by brewing yeasts during
fermentation.
Only a few bacteria are able to grow under such
inhospitable conditions and are able to spoil beer (see
Table 1). These bacteria include both Gram-positive
and Gram-negative species. Gram-positive beer spoilage bacteria belong almost always to the lactic acid
bacteria. They are regarded as most harmful for
brewing industry and are the cause of most of bacterial spoilage incidents. Only a few Gram-negative
bacteria are known to cause beer spoilage. Some of
these belong to the acetic acid bacteria and had
received most attention. Today, these aerobic bacteria
do not present a serious problem in beer spoilage
anymore since improved brewing technology has led
to a drastic reduction of the oxygen content in beer.
Instead, strictly anaerobic bacteria, typically Pectinatus spp. and Megasphaera cerevisiae, have become
serious beer spoilage bacteria.
2.1. Gram-positive bacteria
Almost all the beer spoilage Gram-positive bacteria
belong to lactic acid bacteria. They are a large group

107

Gram-negative
bacteria

Rod-shaped

Cocci

Lactobacillus spp.
Lb. brevis
Lb. brevisimilis
Lb. buchneri
Lb. casei
Lb. coryneformis
Lb. curvatus
Lb. lindneri
Lb. malefermentans
Lb. parabuchneri
Lb. plantarum
Pectinatus spp.
P. cerevisiiphilus
P. frisingensis
P. sp. DSM20764
Selenomonas sp.
S. lacticifex
Zymophilus sp.
Z. raffinosivorans

Pediococcus spp.
P. damnosus
P. dextrinicus
P. inopinatus
Micrococcus sp.
M. kristinae

Megasphaera sp.
M. cerevisiae

Zymomonas sp.
Z. mobilis

of species and genera of Gram-positive bacteria. Only


a few of these lactic acid bacteria are beer-spoiling
organisms. Most hazardous for the brewing industry
are those belonging to the genera Lactobacillus and
Pediococcus. In the period 1980 1990, 58 88% of
the microbial beer-spoilage incidents in Germany
were caused by lactobacilli and pediococci (Back et
al., 1988; Back, 1994). Also in Czech all beer spoilage
bacteria detected in the breweries belonged to lactic
acid bacteria (Hollerova and Kubizniakova, 2001).
The situation in other countries seems to be similar
although for commercial reasons little statistical information has been supplied. These lactic acid bacteria spoil beer by producing haze or rope and cause
unpleasant flavor changes such as sourness and atypical odor.
2.1.1. Lactobacilli
The genus Lactobacillus is the largest genus of
lactic acid bacteria and includes numerous species.
They are widely used in various fermentation processes, including food products such as beer, wine,
yoghurt and pickles. In contrast to the general believe
that all lactobacilli can grow in beer, only a few
species have been reported to be capable of beer
spoilage (Rainbow, 1981; Priest, 1987, 1996; Jespersen and Jakobsen, 1996). Lactobacillus brevis

108

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

appears to be the most important beer spoiling Lactobacillus species and is detected at high frequency in
beer and breweries. More than half of the bacterial
incidents were caused by this species (Back et al.,
1988; Back, 1994; Hollerova and Kubizniakova,
2001). Lb. brevis is an obligate heterofermentative
bacterium. It is one of the best-studied beer spoilage
bacteria and grows optimally at 30 jC and pH 4 6
and is generally resistant to hop compounds. It is
physiologically versatile and can cause various problems in beer such as super-attenuation, due to the
ability to ferment dextrins and starch (Lawrence,
1988). The antibacterial effects of hop compounds
and the mechanism(s) responsible for hop resistance
have been studied in detail in this species (Simpson,
1991, 1993a,b; Simpson and Smith, 1992; Fernandez
and Simpson, 1993; Simpson and Fernandez, 1994;
Sami et al., 1997a,b, 1998; Sakamoto et al., 2001,
2002; Suzuki et al., 2002).
The second most important beer spoiling lactobacillus, Lactobacillus lindneri is responsible for 15
25% of beer-spoilage incidents (Back et al., 1988;
Back, 1994). The physiology of Lb. lindneri is very
similar to that of Lb. brevis and only recently has Lb.
lindneri been recognized on the basis of its 16S rRNA
gene sequence as a phylogenetically separate species
in the genus Lactobacillus (Back et al., 1996; Anon.,
1997). Lb. lindneri is highly resistant to hop compounds (Back, 1981) and grows optimally at 19 23
jC (Priest, 1987; Back et al., 1996) but survives
higher thermal treatments than other lactic acid bacteria (Back et al., 1992). All Lb. lindneri strains, tested
so far, are capable of beer spoilage, while other
lactobacilli comprise both beer spoiling and nonspoiling strains (Rinck and Warkerbauer, 1987a,b; Storgards et al., 1998).
Lactobacillus buchneri, Lb. casei, Lb. coryneformis, Lb. curvatus and Lb. plantarum are less common
beer-spoiling bacteria than the two species described
above (Priest, 1996). Lb. buchneri resembles closely
Lb. brevis but differs in its ability to ferment melezitose. In addition, some Lb. buchneri strains require
riboflavin in their growth medium (Sharpe, 1979;
Back, 1981). Lb. casei can produce diacetyl, which
gives beer an unacceptable buttery flavor. Diacetyl
appears to be more potent in that respect than lactic
acid, the major end-product of lactic acid bacteria.
The threshold value of diacetyl in beer is much lower

(0.15 ppm) than of lactic acid (300 ppm) (Hough et


al., 1982). Diacetyl is also produced during normal
fermentation by yeast and too high levels of diacetyl
are produced when yeast is not properly removed
from young beer (Inoue, 1981). The pathway of
diacetyl formation by lactic acid bacteria has been
studied in detail (Speckman and Collins, 1968, 1973;
Jonsson and Pettersson, 1977).
Lactobacillus brevisimilis (Back, 1987), Lb. malefermentans, Lb. parabuchneri (Farrow et al., 1988),
Lb. collinoides and Lb. paracasei subsp. paracasei
(Hallerova and Kubizniakova, 2001) have also been
reported to be beer spoilage species.
Recently, few additional newly discovered species
of lactobacilli have been added to the list of beer
spoilers. According to its 16S rRNA gene sequence, a
strain called BS-1 is related to Lb. coryneformis, but
its narrow fermentation pattern (only limited to glucose, mannose and fructose) differs significantly not
only from that of Lb. coryneformis but also from other
Lactobacillus spp. (Nakakita et al., 1998). Two other
novel lactobacillus species were found in spoiled beer
with significantly different taxonomical properties
from those of other Lactobacillus spp. (Funahashi et
al., 1998). One species, LA-2, with a 16S rRNA gene
sequence 99.5% similar to that of Lb. collinoides, has
a strong beer soiling ability. The other species, LA-6,
has a weak beer-spoiling ability an did not show any
significant homology to the other Lactobacillus spp.
(Funahashi et al., 1998).
2.1.2. Pediococci
Pediococci are homofermentative bacteria which
grow in pairs and tetrads. They were originally known
as sarcinae because their cell organization was
confused with that of true sarcinae. Beer spoilage,
caused by cocci and characterized by acid formation
and buttery aroma of diacetyl, was therefore called
sarcinae sickness. Pediococcus spp. produce rope
and extensive amounts of diacetyl like Lb. casei. They
are found at many stages in the brewing process from
wort till finished beer. Several Pediococcus spp. have
been found in breweries: Pediococcus acidilactici,
Pediococcus damnosus P. dextrinicus, P. halophilus
which recently is classified as Tetragenococcus halophilus (Collins et al., 1990), P. inopinatus, P. parvulus and P. pentosaceus (Back, 1978; Back and
Stackebrandt, 1978). Among them is P. damnosus,

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

109

2.2. Gram-negative bacteria

stain, DSM20764, isolated from spoiled beer that


differs considerably in genotype from the two other
species (Weiss, personal communication, 1987). Its
16S rRNA gene sequence is distinctly different from
that of the other two species (Sakamoto, 1997).
Pectinatus spp. are nonspore-forming motile rods with
lateral flagella attached to the concave side of the cell
body. They swim actively and appear X-shaped when
cells are young and snake-like and longer when cells
are older. They possess some features that are characteristic for Gram-positive bacteria (Haikara et al.,
1981) and are regarded as being intermediate between
Gram-positive and Gram-negative bacteria. Growth
takes place between 15 and 40 jC with an optimum
around 32 jC (Chelak and Ingledew, 1987), between
pH 3.5 and 6 with an optimum at 4.5 and in media
containing up to 4.5% (w/w) of ethanol (Chelak and
Ingledew, 1987; Watier et al., 1993). During growth,
considerable amounts of propionic and acetic acids
are produced as well as succinic and lactic acids and
acetoin. Pectinatus spp. can also ferment lactic acid.
Since lactic acid is the sole source of carbon in SMMP
medium (see Section 3.1), this medium is used for the
selective isolation of Pectinatus spp. and Megaspaera
spp. (Lee, 1994). The most characteristic feature of
spoilage caused by Pectinatus spp. is extensive turbidity and an offensive rotten egg smell brought by
the combination of various fatty acids, hydrogen
sulfide and methyl mercaptan (Lee et al., 1978,
1980; Haikara et al., 1981). This spoilage activity
can cause serious damages for breweries.

2.2.1. Pectinatus
Pectinatus spp. are now recognized as one of the
most detestable beer spoilage bacteria. They play a
major role in 20 30% of bacterial incidents, mainly in
nonpasteurized beer rather than in pasteurized beer
(Back, 1994). Pectinatus species were long thought to
be Zymomonas spp. because of their phenotypical
similarities. The first isolate was obtained from breweries in 1971 (Lee et al., 1978) and so were all
subsequent isolates (Back et al., 1979; Haukeli,
1980; Kirchner et al., 1980; Haikara et al., 1981;
Takahashi, 1983, Soberka et al., 1988, 1989). The
natural habitat of the Pectinatus species are still
unknown (Haikara, 1991). Two species are found in
this genus: Pectinatus cerevisiiphilus and Pectinatus
frisingensis (Schleifer et al., 1990). There is also one

2.2.2. Megasphaera
Megasphaera has emerged in breweries along with
Pectinatus and is responsible for 3 7% of bacterial
beer incidents (Back et al., 1988; Back, 1994). They
are nonspore-forming, nonmotile, mesophilic cocci
that occur singly or in pairs and occasionally as short
chains. This genus includes two species, M. elsdeni
and M. cerevisiae. Since the first isolations in 1976,
only M. cerevisiae has been blamed to be responsible
for beer spoilage (Weiss et al., 1979; Haikara and
Lounatmaa, 1987; Lee, 1994). M. cerevisiae grows
between 15 and 37 jC with an optimum at 28 jC and
at pH values above 4.1. The growth is inhibited at
ethanol concentrations above 2.8 (w/v) but is still
possible up to 5.5 (w/v) (Haikara and Lounatmaa,
1987; Lawrence, 1988). It is the most anaerobic

the most common beer spoiler. It was responsible for


more than 20% of all bacterial incidents in the period
1980 1987 (Back, 1980; Back et al., 1988) but only
for 3 4% in 1992 1993 (Back, 1994). The incidence
of beer spoilage by pediococci has decreased most
likely as a result of improved sanitation conditions in
breweries. P. damnosus is generally resistant to hop
compounds. It is interesting that P. damnosus is
commonly found in beer and late fermentations, but
seldom in pitching yeast. In contrast, P. inopinatus is
frequently detected in pitching yeast but rarely in the
other stages of beer fermentations (McCaig and
Weaver, 1983; Priest, 1987). P. inopinatus and P.
dextrinicus can grow in beer at pH values above 4.2
and with low concentrations of ethanol and hop
compounds (Lawrence, 1988). P. pentsaceus and P.
acidilactici have never been reported to cause any
defect in finished beer (Simpson and Taguchi, 1995).
2.1.3. Other Gram-positive bacteria
In additional to Lactobacillus and Pediococcus
species, also, a species from the genus Microcococcus
has been reported to be occasionally responsible for
beer spoilage. M. kristinae can grow in beer with low
ethanol and hop compounds at pH values above 4.5
(Back, 1981). Micrococci are usually strictly aerobic
but M. kristinae can grow also under anaerobic
condition (Lawrence and Priest, 1981). It produces a
fruity atypical aroma in beer (Back, 1981).

110

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

species known to exist in the brewing environment


(Seidel et al., 1979). Beer spoilage caused by this
organism results in a similar extreme turbidity as
Pectinatus and the production of considerable quantities of butyric acid together with smaller amounts of
acetic, isovaleric, valeric and caproic acids as well as
acetoin (Seidel et al., 1979). Like Pectinatus, the
production of hydrogen sulfide causes a fecal odor
in beer (Lee, 1994), which makes this bacterium one
of the most feared organisms for brewers.
2.2.3. Other Gram-negative bacteria
In addition to the two genera described above,
some other Gram-negative bacteria have been found
to cause problems in the brewing industry. Anaerobic
Zymomonas spp. have been found in primed beer to
which sugar was added and in ale beer. Zymomonas
mobilis is an aerotolerant anaerobe and grows above
pH 3.4 and at ethanol concentrations below 10% (w/v)
(van Vuuren, 1996). There is no report of Z. mobilis
spoilage in lager beer, probably because of its selective fermentation character (nonfermentative of maltose, maltotriose but fermentative of glucose, fructose
and sucrose). It produces high levels of acetaldehydes
and hydrogen sulfide. Also another Zymophilus spp.,
Z. raffinosivorans, has been reported as a beer spoiler
(Schleifer et al., 1990; Seidel-Rufer, 1990). The genus
Zymophilus is phylogenetically close to the genus
Pectinatus. Zymophilus spp. can grow in beer, like
Pectinatus spp., at pH values above 4.3 4.6 and
ethanol concentrations below 5% (w/v). Also, their
beer spoilage activity is similar to that of Pectinatus
spp. (Jespersen and Jakobsen, 1996).
Another Gram-negative bacterium Selenomonas
lacitifex has also been reported to play a role in
certain beer spoilage incidents but this species has
hardly been studied (Schleifer et al., 1990). Historically, a lot of attention of the brewing industry has
been given to aerobic Gram-negative bacteria. Acetic
acid bacteria, i.e., Gluconobacter and Acetobacter
used to be well known to breweries. They convert
ethanol into acetate, which results in vinegary offflavor of beer. For reasons explained above, such
aerobes are no longer important in modern breweries.
Hafnia protea, formerly Obesumbacterium proteus, and Rahnella aquatilis, formerly Enterobacter
agglomerans, have been detected in pitching yeasts
but never in finished beer. They can retard the

fermentation process. Beer produced with yeasts contaminated with H. protea has a parsnip-like or fruity
odor and flavor (van Vuuren, 1996). Abnormally high
levels of diacetyl and dimethly sulfide were detected
in beer produced from wort contaminated by R.
aquatilis (van Vuuren, 1996).
Recently, a novel strictly anaerobic Gram-negative
bacterium was isolated from a brewery (Nakakita et
al., 1998). It is a rod-shape bacterium with no flagella
that can grow in beer at pH values above 4.3 and does
not produce propionic acid. Genetic and phenotipical
studies indicated that this bacterium is different from
Pectinatus, Zymomonas and Selenomonas spp.

3. Detection of beer spoilage bacteria


3.1. Culture media
It has been seen above that only a limited number
of bacterial species are responsible for beer spoilage
and that only a few species are major beer spoilage
bacteria. For quality assurance of finished beer it is
usually practically sufficient to control potential contaminations by Lactobacillus brevis, Lb. lindneri,
Pediococcus damnosus, and Pectinatus spp. Most
studies on beer spoilage bacteria have focused on
the taxonomical classification of these bacteria. These
studies have made it possible to identify the bacteria
detected in beer and breweries and to take the proper
measures to control them.
Along with the taxonomical studies, a number of
selective culture media for beer spoilage bacteria have
been developed and recommended by various brewery organizations as shown in Table 2. European
Brewery Convention recommends three media for
the detection of lactobacilli and pediococci: MRS
(de Man, Rogosa and Sharpe) agar supplemented with
cycloheximide to prevent growth of aerobes such as
yeasts and moulds, Raka-Ray medium supplemented
with cycloheximide and VLB S7-S (Versuchs- und
Lehranstalt fur Brauerei in Berlin). Other optional
media are UBA (Universal Beer Agar) supplemented
with cycloheximide, HLP (Hsus Lactobacillus and
Pediococcus medium), NBB (Nachweismedium fur
bierschadliche Bakterien), WLD (Wallerstein Laboratory Differential Medium), Nakagawa medium, SDA
(Schwarz Differential Agar) and MRS modified by

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124


Table 2
Examples of culture media for the detection of beer spoilage
bacteria
Media
MRS (de Man,
Rogosa and Sharpe)
Raka-Ray
VLB S7-S (Versuchsund Lehranstalt fur
Brauerei in Berlin)
HLP (Hsus Lactobacillus
and Pediococcus
medium)
WLD (Wallerstein
Differential)
Nakagawa
SDA (Schwarz
Differential Agar)
Concentrated MRS
PYF (Peptone, Yeasts
extract and Fructose)
Thioglycolate Medium
LL-Agar
UBA (Universal Beer Agar)
NBB (Nachweismedium fur
bierschadliche Bakterien)
Brewers Tomato
Juice Medium
LMDA (Lees MultiDifferential Agar)
BMB (Barney Miller
Brewery Medium)
SMMP (Selective Medium
for Megasphaera and
Pectinatus)

Recommended bya

Bacteria
b

LAB

LAB, G(
LAB

EBC, ASBC, BCOJ


)c

EBC, ASBC, BCOJ


EBC, BCOJ

LAB

EBC, BCOJ

LAB

EBC, BCOJ

LAB
LAB

EBC, BCOJ
EBC, BCOJ

G(
G(

)
)

EBC, BCOJ
EBC, BCOJ

G( )
G( )
LAB, G(
LAB, G(

)
)

EBC
EBC, BCOJ
EBC, ASBC, BCOJ
EBC, BCOJ

LAB, G(

ASBC

LAB

ASBC

LAB

ASBC

G(

ASBC, BCOJ

EBC, European Brewery Convention; ASBC, American


Society of Brewing Chemists; BCOJ, Brewery Convention of
Japan.
b
LAB, lactic acid bacteria.
c
G( ), Gram-negative bacteria.

addition of maltose and yeast extract at pH 4.7. None


of these media are suitable for detecting all strains of
lactobacilli and pediococci but a combination of some
of these media yields the best results. For the detection of Pectinatus and Megasphaera, the following
media are recommended: Concentrated MRS broth,
PYF (Peptone, Yeast extract and Fructose) and Thioglycolate Medium for enrichment of beer, LL-Agar
for growth in Lee tube, and UBA, NBB and Raka
Ray for routine analysis at breweries. For Zymomonas
spp., Zymomonas Enrichment Medium is also recommended (European Brewery Convention, 1992).

111

American Society of Brewing Chemists recommends


UBA and Brewers Tomato Juice Medium for general
microbial detection and other media including LMDA
(Lees Multi-Differential Agar), Raka-Ray, BMB
(Barney Miller Brewery Medium) and MRS for the
detection of lactic acid bacteria and SMMP (Selective
Medium for Megasphera and Pectinatus) (The Technical Committee and the Editorial Committee of the
American Society of Brewing Chemists, 1992). Brewery Convention of Japan also recommends the use of
some of those media (Brewery Convention of Japan,
1999). These culture media are listed in Table 2.
The conventional detection method based on culturing of the organisms in these media has the
significant disadvantage that is very time-consuming.
One week or even longer is needed to obtain visible
colonies on plates or turbidity in broths. Consequently, the products are often already released for sale
before the microbiological results become available.
Another problem is that these media are not speciesspecific. Media for the detection of beer spoilage
lactic acid bacteria allow also growth of nonbeerspoilage species such as Lactobacillus delbrueckii
and P. acidilactici. If the selectivity is increased by
the addition of specific chemicals to these media, even
longer detection times might be required.
3.2. Identification methods
Following the bacterial detection in these media,
species identification is needed. Besides the basic tests
such as colony morphology, cell morphology, Gramstaining and catalase assays, also biochemical tests
such as sugar fermentation pattern and chromatographic analysis of organic acids can be performed.
Also, specific detection and identification methods are
used such as immunoassays with polyclonal or monoclonal antibodies (Claussen et al., 1975, 1981; Dolezil
and Kirsop, 1976; Haikara, 1983; Gares et al., 1993;
Sato et al., 1994; Whiting et al., 1992, 1999a,b; Ziola
et al., 1999, 2000a,b), DNA DNA hybridization,
DNA sequencing (Doyle et al., 1995) and polymerase
chain reaction (PCR) (Tsuchiya et al., 1992, 1993,
1994; DiMichele and Lewis, 1993; Thompson et al.,
1994; Vogesser et al., 1955a,b; Yasui, 1995; Stewart
and Dowhanick, 1996; Yasui et al., 1997; Sakamoto,
1997; Sakamoto et al., 1997; Satokari et al., 1997,
1998; Juvonen and Satokari, 1999; Motoyama and

112

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

Ogata, 2000; Bischoff et al., 2001). These modern


methods have been reviewed (Barney and Kot, 1992,
Schmidt, 1992; Dowhanick and Russell, 1993; Dowhanick, 1995; Schofield, 1995; Hammond, 1996;
Schmidt, 1999).
3.3. Discrimination of beer spoilage bacteria from
nonspoilers
After detection and identification it is for some
species necessary to identify the bacterium as an
actual beer spoiler. While all strains belonging to
Pectinatus spp. and M. cerevisiae have been reported
to be capable of spoiling beer (Haikara, 1991), lactic
acid bacteria include both beer spoilage and nonspoilage strains. Among Lb. brevis and P. damnosus,
most of the strains are capable of spoiling beer and
only a few strains are not. On the other hand, the
number of beer spoilage strains in Lb. casei, Lb.
coryneformis and Lb. plantarum is limited. Exceptionally, all strains of Lb. lindneri have been reported
to be capable of spoiling beer (Rinck and Wackerbauer, 1987a,b; Storgards et al., 1998).
Before the beginning of 1990s, the only method
available for judging the beer spoiling potential of a
bacterium was the so-called forcing test. In this test,
the bacterium was re-inoculated into beer or beer
enriched with concentrated nutrient medium. However, this test has proven to be far from practical for
quality assurance since a few months are needed to
obtain conclusive results.
More rapid procedures have been developed. The
identification at the strain level can now be done at the
genome level. Ribotyping, based on Southern hybridization with a ribosomal gene as a probe, has been
successfully introduced (Motoyama et al., 1998, 2000;
Satokari et al., 2000; Suihko and Haikara, 2000;
Barney et al., 2001). Fully automated ribotyping
machines now commercially available and only 8 h
are needed to obtain conclusive results. Amplified
Fragment Length Polymorphisms (AFLP) (Perpete et
al., 2001) and Random Amplified Polymorphic DNA
(RAPD-PCR) (Savard et al., 1994; Tompkins et al.,
1996) have also successfully been applied for bacterial strain identification as well as for identification
of brewing yeasts. In these methods, the genotype of
each beer spoilage strain is registered in a database.
A comparison of the genotype of a newly detected

strain with registered genotypes can provide important intervention.


Another approach is to determine the common
physiological properties responsible for beer-spoiling
ability. For beer spoilage lactic acid bacteria the
common physiological denominator is hop resistance,
which allows growth of these bacteria in beer. However, measuring hop resistance by culturing in hop
containing medium, is time-consuming.

4. Antibacterial activity of hop compounds


4.1. History of hop usage in beer
The comfortable bitterness experienced in beer
drinking is characteristic for and is mainly caused
by hop compounds. These hop compounds are present
in the flowers of the hop plant, Humulus lupulus, L.,
which are added to the wort. This plant has been
known for thousands of years. However, its use in
beer is not old as the history of beer itself (5000 7000
years). A few descriptions of hops as a beer additive
as well as a decoration for gardens were found in
documents of the 6th century BC. German monks in
the 12th century often used hops in beer making. In
those days, as in ancient times, it was popular to use a
variety of fruits, herbs and spices to flavor beer (socalled gruit beer). Initially, the bitter taste from hops
was not particularly appreciated. However, when in
the 14th century beer production increased and beer
was exported, the importance of hops in beer was
gradually more and more appreciated, not only for its
contribution to beer flavor but also for its contribution
to the stability. Hopped beer can be preserved significantly longer than gruit beer. In 1516, Wilhelm IV,
the lord of Bayern, enacted the Reinheitsgebot (Purity Law) which ordered that beer must be made from
barley, water and hops. Since then, the use of hops
became more popular and standard. Many aspects of
this law were adopted by other countries, which made
hops indispensable for the brewing industry.
4.2. Hop plant
The hop plant is a vine, belonging to the family of
hemp. It is dioecious and blooms yearly. Nowadays,
it is mainly cultured for the brewing industry. Only

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

the female flowers, the so-called cones, are used for


beer. The mature cones contain golden resinous
granules, the lupulin, which are the most important
part of the flower for the bitterness and preservation
of beer. Hop resins are extracted and fractionated as
shown in Fig. 1.
4.3. Antibacterial compounds in hops
Hop chemistry has been developed since 19th
century and has been extensively reviewed (Verzele,
1986; Moir, 2000). Research has especially been
focused on the antibacterial properties of hop compounds and the bitter substances derived from hops.
This research goes back to 1888 when Hayduck
showed first that antiseptic properties of the hops
are due to the soft resins (Hayduck, 1888). In the
Institute of Brewing of the United Kingdom, Walker
conducted from 1922 till 1941 a long-term investigation on the preservative principles of hops (Pyman et
al., 1922; Walker, 1923a,b, 1924a,b, 1925, 1938,
1941; Hasting et al., 1926; Hastings and Walker,
1928a,b, 1929; Walker and Hastings, 1931, 1933a,b;
Walker et al., 1931, 1932, 1935, 1940; Walker and
Parker, 1936, 1937, 1938, 1940a,b). The study focused on the antiseptic properties of a-acids fraction
(humulone) and h-acid fraction (lupulone).
The a-acid fraction is a mixture of homologous
compounds, the a-acids, which are not transferred as
such to beer. During the wort boiling stage in the
brewing process, a-acids are converted by a rear-

113

rangement or isomerization to iso-a-acids, which are


much more soluble and bitterer than the original
compounds. This conversion, which is very important
in hop chemistry, was advanced first in 1888 (Hayduck). Wieland et al. (1925) suggested that the hydrolysis of humulone to humulinic acid proceeded via
an intermediate. Windisch et al. (1927) investigated
the humulone boiling products under alkaline conditions and isolated a resinous and bitter oil termed
Resin A with chemical properties similar to the
intermediate and isomeric with humulone. Around
1950, Rigby and Bethune, (1952, 1953) showed that
a-acid fraction is a mixture of three major compounds: humulone, cohumulone and adhumulone
(Fig. 2). The bittering compounds of beer were found
to comprise three major analogues of these three aacids, which are now known as iso-a-acids: isohumulone, isocohumulone and isoadhumulone (Fig. 2).
Stereoisomers (cis- and trans-) exist for each iso-aacid. Finally, the chemical structure and configuration
of naturally occuring ( )-humulone (De Keukeleire
and Verzele, 1970) and isohumolones (De Keukeleire
and Verzele, 1971) were elucidated. The isomerization
yield of a-acids during wort boiling process is low
(typically of the order of 30%) (Hughes, 2000) due to
relatively acidic condition of wort (ca. pH 5.2) and the
adsorption to the wort coagulum during boiling and
fermentation.
h-acids or lupulones in hops are very poorly
soluble in wort and beer and cannot undergo the same
isomerization processes as a-acids. Consequently,

Fig. 1. Fractionation of hop resins (Hough et al., 1971).

114

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

Fig. 2. Chemical structures of hop compounds. The name of each a-acid (I) and h-acid (II) is dependent on its side chain. During wort boiling
process, a-acids (naturally R-body; III) are isomerized to result in stereoisomers of trans-isohumulones (IV) and cis-isohumulones (V).

they are not transferred to beer and have no direct


value in brewing.
4.4. Antibacterial mechanism of hop compounds
The antibacterial activities of a-acid (humulone)
and h-acid (lupulone) have been studied before 1950.
Their antibacterial activities are higher than of iso-aacids but they dissolve to less extent in beer and water.
Studies of the antiseptic properties of hopped wort and
hop boiling product showed that they inhibit growth
of Gram-positive bacteria but not of Gram-negative
bacteria (Shimwell, 1937a; Walker and Blakebrough,
1952). It was first reported by Shimwell (1937a) that
the antiseptic potency of hop increases at lower pH.
Interestingly, he predicted that the antiseptic potency
of hop is associated with permeability changes of the
bacterial cell wall Shimwell, (1937b). The bacteriostatic power was also studied of hop compounds,
including humulone and the humulone boiling product (Walker and Blakebrough, 1952). The humulone
boiling product had less bacteriostatic potency in malt
extract (pH 5.5) and wort (ph 5.2) than the original
humulone, while its potency was the same at pH 4.3,
the pH of beer. The hop constituents (lupulone,

humulone, isohumulone and humulinic acid) were


found to cause leakage of the cytoplasmic membrane
of Bacillus subtilis, resulting in the inhibition of active
transport of sugar and amino acids (Teuber and
Schmalreck, 1973). Subsequently, inhibition of respiration and synthesis of protein, RNA and DNA was
also observed. Since the iso-a-acids are mainly present in beer among the hop resins and their derivatives,
a precise investigation of the antibacterial activity of
iso-a-acids was needed for understanding the preservation or bacterial stability of beer. The molecular
mechanism of antibacterial activity of iso-a-acids and
the effects of pH of the growth medium and other
variables on the antibacterial activity of hop compounds were investigated after 1990 (Simpson and
Smith, 1992). Hop compounds are weak acids and the
undissociated forms are mainly responsible for inhibition of bacterial growth (Fig. 3).
In Lb. brevis (Simpson, 1993b), trans-isohumulone
reduces the uptake of leucine and causes slow leakage
of accumulated leucine. trans-Isohumulone dissipates
effectively the transmembrane pH gradient (DpH) of
the proton motive force but not the transmembrane
electrical potential (Dw). Inhibition of H+-ATPase
activity was not observed. Potentiometric studies

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

115

naturally occurring analogues (Price and Stapely,


2001).

5. Hop resistance in lactic acid bacteria

Fig. 3. Dissociation of trans-isohumulone (Fernandez and Simpson,


1995b).

revealed that undissociated trans-isohumulone acts


most likely as an ionophore, catalyzing electroneutral
influx of undissociated isohumulone, internal dissociation of (H+)-isohumulone and efflux of the complex
of isohumulone with divalent cations such as Mn2 +.
This cation is known to be present at high concentrations in lactobacillus cells (Archibald and Fridovich, 1981a,b; Archibald and Duong, 1984). The
results of this activity is a decrease of the pH gradient
across the membrane. It was reported that the antibacterial activity of trans-isohumulone can be influenced by the presence of cations in the medium.
Protonophoric activity of trans-isohumulone requires
the presence of monovalent cations such as K+, Na+ or
Rb+ and increases with the concentration of these
monovalent cations (Simpson and Smith, 1992).
trans-Isohumulone cannot bind K+ unless a divalent
cation, such as Mn2 +, Mg2 +, Ni2 + and Ca2 + or a
trivalent cation, such as Li3 + and Al3 +, is present in
the medium (Simpson et al., 1993; Simpson and
Hughes, 1993). Thus, the ability of hop compounds
to bind simultaneously two or more cations may be
crucial for their antibacterial action but the reason has
been still unclear.
The properties of other hop acids are similar to
those of trans-isohumulone and it is likely that the
mechanism of their antibacterial activities is also
similar. Some strains of lactic acid bacteria, which
are sensitive to trans-isohumulone, are also sensitive
to ( )-humulone and colupulone and other strains
resistant to trans-isohumulone are also resistant to the
related compounds (Fernandez and Simpson, 1993).
The antibacterial activities of six naturally occurring
iso-a-acids, five chemically reduced iso-a-acids and a
reduced iso-a-acids mixture were higher at lower pH
values while several hydrophobic reduced iso-a-acids
were found to be far more antibacterial than their

Beer spoiling lactic acid bacteria have to be hop


resistant in order to grow in beer. The understanding
and elucidation of the mechanism of hop resistance is
not only of scientific interest but is also important for
the microbial control in brewing industry to predict
the beer-spoiling ability of lactic acid bacteria.
5.1. Variation of hop resistance
The extent of hop resistance varies between bacteria. Among beer-spoiling lactic acid bacteria, Lb.
brevis is so far the most resistant to hop compounds.
The degree of hop resistance varies among the different strains of Lb. brevis (Hough et al., 1957; Harris
and Watson, 1960; Simpson and Fernandez, 1992;
Fernandez and Simpson, 1993). Hop resistance of
lactobacilli decreases upon prolonged serial subculturing in the absence of hop compounds (Shimwell,
1936; Yamamoto, 1958; Richards and Macrae,
1964). Hop resistance was thought to be caused by
immunity acquired by prolonged contact with hop
compounds under brewing conditions (Shimwell,
1937a). The necessity of beer spoiling lactic acid
bacteria to acclimatize to beer or hop compounds in
order to reproduce in beer (Yamamoto, 1958) was
solidly documented by Richards and Macrae in 1964.
Hop resistance increased 8- to 20-fold in strains of
lactobacilli upon serial subculturing in media containing increasing concentrations of hop compounds,
while subculturing of resistant populations in the
absence of hop compounds resulted gradually in
decreased hop resistance. It took about 1 year of
subculturing in unhopped beer to maximally reduce
hop resistance of lactobacilli, indicating that the
acquired hop resistance can be a very stable property
(Shimwell, 1936). However, organisms isolated from
spoiled beer frequently fail to grow upon reinoculation in beer. Preculturing in the presence of subinhibitory concentrations of isohumulone is needed in order
to make growth in beer possible (Simpson and Fernandez, 1992). The stability of hop resistance in Lb.
brevis appears to vary from strain to strain. The hop

116

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

resistance in Lb. brevis strain BSO310 could not be


altered by plasmid curing or mutation induced with
ultraviolet light (Fernandez and Simpson, 1993), suggesting that it may be generally a stable character,
both phenotypically and genetically. The Lb. brevis
strain ABBC45 can develop hop resistance in the
same way as observed by Richards and Macrea (Sami
et al., 1997a). Hop resistance increased with the copy
number of plasmid pRH45. When Lb. brevis ABBC45
was cured from this plasmid by serial subculturing in
the absence of hop compounds, the degree of hop
resistance decreased (Sami et al., 1998; Suzuki et al.,
2002). This plasmid contains the horA gene that codes
for a polypeptide that is 53% identical to LmrA, the
lactococcal ATP-binding cassette (ABC) multidrug
transporter (van Veen et al., 1996). HorA protein
was expressed heterologously in Lactococcus lactis
and found to function as an ABC-type multidrug
transporter and to excrete hop compounds (Sakamoto
et al., 2001).
5.2. Features of hop resistance
The pattern of resistance or sensitivity of several
species of lactobacilli to isohumulone is similar to that
of the related hop acids humulone and lupulone
(Richards and Macrae, 1964; Fernandez and Simpson,
1993). Fernandez and Simpson (1993) compared the
properties of hop-resistant strains of lactobacilli and
pediococci with those of sensitive strains. No obvious
correlation was found between hop resistance and cell
morphology, colony morphology, pH range for
growth, sugar utilization profile, products of metabolism, manganese requirement and sensitivity to superoxide radicals, expression of cellular proteins and
resistance to various antibacterial agents. However,
differences were found in the transmembrane pH
gradient (DpH) and the cellular ATP pool. Because
hop compounds act as protonophores that dissipate
the DpH across the cellular membrane (Simpson,
1993a,b), these differences are of great importance
for understanding the mechanism of hop resistance.
5.3. Mechanisms of hop resistance
The molecular structure of antibacterial agents
might supply an insight in the mechanism of resistance. Resistance to trans-isohumulone also results in

resistance to ( )-humulone and colupulone (Fernandez and Simpson, 1993), suggesting a common mechanism of resistance against a broad range of hop acids.
trans-Isohumulone and ( )-humulone have three
hydrophobic side chains while colupulone has four.
The side chains are attached to a five-membered ring
of trans-isohumulone, but to a six-membered ring of
( )-humulone and colupulone (Fig. 2). On the other
hand, resistance was not observed to other ionophores
(nigericin, A23187, CCCP, monesin), weak acid food
preservatives (sorbic acid, benzoic acid), solvents
(ethanol) or antibiotics (ampicillin, cefamandole, vancomycin) (Fernandez and Simpson, 1993). The resistance mechanism might be specific for the h-triketone
group of the hop acids, which plays an essential role
in the antibacterial action (Simpson, 1991).
Microorganisms have developed various ways to
resist the toxicity of antibacterial agents:
(i) enzymatic drug inactivation. A well known
example is beta-lactamase which hydrolyzes the beta-lactam ring into innocuous substrates. In hop-resistant strains of Lb. brevis, neither conversion nor
inactivation of trans-isohumulone was found (Simpson and Fernandez, 1994).
(ii) target alteration. Cellular targets can be altered
by mutation or enzymatic modification in such a way
that the affinity of the target for the antibiotics is
reduced. In the case of trans-isohumulone, the target
site is the cell membrane (Teuber and Schmalreck,
1973; Schmalreck et al., 1975; Simpson, 1993a,b). A
sake (Japanese rice wine) spoilage bacterium, Lb.
heterohiochii, contains extremely long-chain fatty
acids in its membrane (Uchida, 1974), which may
play a role in ethanol resistance (Ingram and Buttke,
1984). It is possible but not yet investigated that hopresistant Lb. brevis has also a changed lipid composition of its membrane to lower the permeability to
hop compounds.
(iii) inhibition of drug influx. The outer membrane
of Gram-negative bacteria restricts the permeation of
lipophilic drugs, while the cell wall of the Grampositive mycobacteria has been found to be an exceptionally efficient barrier. The permeation of hop compounds might also be affected by the presence of a
galactosylated glycerol teichoic acid in beer spoilage
lactic acid bacteria (Yasui and Yoda, 1997b).
(iv) active extrusion of drugs. The presence of
(multi) drug resistance pump in the cytoplasmic

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

membrane of many bacteria has been extensively


documented (Putman et al., 2000). A multidrug resistance pump HorA (Sakamoto et al., 2001) has been
found in Lb. brevis ABBC45, which is overexpressed
when exposed to hop compounds (Sami, 1999). In
addition, a proton motive force-dependent hop excretion transporter was suggested in this strain (Suzuki
et al., 2002).
(v) other mechanisms to tolerate the toxic effects of
drugs. Since hop compounds act as protonophores and
dissipate the transmembrane pH gradient (DpH), the
cells could respond by increasing the rate at which
protons are expelled. The hop-resistant strains maintain a larger DpH than hop-sensitive strains (Simpson
and Fernandez, 1993) and Lb. brevis ABBC45
increases its H+-ATPase activity upon acclimatization
to hop compounds (Sakamoto et al., 2002). The ATP
pool in hop-resistant strains was also found to be a
larger than in hop-sensitive strains (Simpson and
Fernandez, 1994; Okazaki et al., 1997). The ability

117

of hop-resistant strains to produce large amounts of


ATP in the cell is needed for the increased activity of
the H+-ATPase and for the hop extruding activity of
HorA. It is interesting that these responses occurred
also in hop-sensitive strains at subinhibitory concentration of trans-isohumulone (Simpson, 1993a; Simpson and Fernandez, 1994). However, at higher
concentrations, both the DpH and the pool decreased
in the hop-sensitive strains, but not in the hop-resistant strains. A role of HitA in hop resistance was
suggested (Hayashi et al., 2001). It is about 30%
identical to the natural resistance-associated macrophage proteins (Nramp), which function as divalentcation transporters in many prokaryotic and eukaryotic organisms. HitA could play a role in transport of
divalent cations, while isohumulone has been claimed
to exchange protons for cellar divalent cations such as
Mn2 +. Thus, a simple mechanism of hop resistance
does not appear to exist. The resistance mechanisms
found so far in Lb. brevis are illustrated in Fig. 4.

Fig. 4. Mechanisms of hop resistance. Hop compounds act as ionophores that exchange protons for cellular divalent cations. In a hop-sensitive
cell, hop compounds (Hop-H) invade the cell and dissociate into hop anions and protons due to the higher internal pH. Hop anions trap divalent
cations such as Mn2 + and diffuse out of the cell. The ionophoric action together with the diffusion of the hop metal complex results in an
electroneutral exchange of cations. Release of protons from hop compounds decreases the intracellular pH and results in a dissipation of the
transmembrane proton gradient (DpH) and the proton motive force (pmf). Consequently, pmf-driven uptake of nutrients will be decreased. In
hop-resistant cells, hop compounds can be expelled from the cytoplasmic membrane by HorA (a) (Sakamoto et al., 2001) and probably also by a
pmf-dependent transporter (b) (Suzuki et al., 2002). Furthermore, overexpressed H+-ATPase increases the pumping of protons released from the
hop compounds (c) (Sakamoto et al., 2002). More ATP is generated in hop-resistant cells than in hop-sensitive cells (Simpson and Fernandez,
1994). Galactosylated glycerol teichoic acid in the cell wall (Yasui et al., 1997) and a changed lipid composition of the cytoplasmic membrane
of beer spoilage lactic acid bacteria may increase the barrier to hop compounds.

118

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

6. Beer-spoiling ability in lactic acid bacteria


Hop resistance is crucial to beer-spoiling ability of
lactic acid bacteria. In many bacteria, hop-resistance
mechanisms need to be induced before growth in beer
is possible. For this induction, some bacteria need to
be exposed to subinhibitory concentrations of hop
compounds (Simpson, 1993a,b).
6.1. Factors affecting the growth in beer
Beer spoilage and bacterial growth depend on the
strain and the type of beer. The ability of 14 hopresistant lactic acid bacterial strains, including Lb.
brevis and P. damnosus strains, was investigated for
their capacity to grow in 17 different lager beers with
the biological challenge test (Fernandez and Simpson,
1995a). A statistical analysis of the relationship between spoilage potential and 56 parameters of beer
composition revealed a correlation with eight parameters: pH, beer color, the content of free amino nitrogen,
total soluble nitrogen content and the concentrations of
a range of individual amino acids, maltotriose, undissociated SO2 and hop compounds. The effects of
dissolved carbon dioxide (CO2) and phenolic compounds including catechin, gallic, phytic and ferulic
acids on beer spoilage were also investigated (Hammond et al., 1999). CO2 was found to inhibit the
growth of lactobacilli at the concentrations present in
typical beer but stimulate at the lower concentrations.
Among the phenolic compounds, ferulic acid, a component of barley cell wall and hence present in all
beers, exerted a stronger antibacterial activity after
enzymatic conversion into 4-vinyl guaiacol. Organic
acids in beer may also influence bacterial growth but
this aspect has hardly been studied.
6.2. Prediction of beer spoilage by lactic acid
bacteria
A number of attempts have been made to develop
methods to predict beer-spoiling ability. For lactic
acid bacteria, hop resistance is the key factor. As
described above, some factors have been identified
to cause hop resistance. Rapid procedures for detecting these factors would be very beneficial for microbial control in breweries. A set of PCR primers have
been made that can specifically detect the horA gene

or its homologues in a wide range of lactobacilli


(Sami et al., 1997b). Most horA-positive strains were
found to have beer-spoiling activity, indicating that
this is a very useful prediction method. Another
prediction method is based on ATP pool measurements in lactobacillus cells (Okazaki et al., 1997).
Polyclonal and monoclonal antibodies specific only for beer spoilage strains have been reported. A
series of antisera were made against the Lactobacillus
group E antigen, a cell wall glycerol teichoic acid
beneath the S-layer protein (Yasui et al., 1992, 1995;
Yasui and Yoda, 1997a) and known to be present in
Lb. brevis, Lb. buchneri, Lb. delbrueckii subsp. lactis
and subsp. bulgaricus (Sharpe, 1955; Sharpe et al.,
1964). When the beer spoilage strain Lb. brevis 578
was used as an antigen, the resulting antiserum
reacted specifically with other beer spoilage strains
of Lb. brevis. Surprisingly, this antiserum also reacted
with beer spoilage strains of P. damnosus, but not with
any strains of Lb. linderi (Yasui and Yoda, 1997a).
Galactosylated glycerol teichoic acid was found to be
the most likely epitope that presumably selectively
increases the cell barrier to hop compounds (Yasui
and Yoda, 1997b). Three monoclonal antibodies specific for beer spoilage ability of lactic acid bacteria
were obtained by immunizing mice with cells cultured
in beer (Tsuchiya et al., 2000). The monoclonal
antibody raised against Lb. brevis reacted with all
beer spoilage strains of Lb. brevis and several beer
spoilage strains of P. damnosus, but not with nonspoilage strains of Lb. brevis, P. damnosus and other
lactic acid bacteria. The monoclonal antibody raised
against P. damnosus reacted significantly with all beer
spoilage strains of P. damnosus and weakly with many
of the beer spoilage strains of Lb. brevis. On the other
hand, the monoclonal antibody raised against Lb.
lindneri reacted specifically only with Lb. lindneri.
The reactivity of the still unknown antigens did not
change regardless of the presence of hop compounds
in their culture media.
D-lactate dehydrogenase (LDH) of 60 strains of Lb.
brevis, including 44 beer spoiling strains and 16
nonspoiling strains, was also investigated. The strains
could be divided in five groups (A, B, C, D and E) on
the basis of the mobility of their D-LDH in native
polyacrylamide gels (Takahashi et al., 1999). Forty
out of forty-four beer spoilage strains were classified
to the group B, suggesting a relationship between the

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124


D-LDH

profile and beer-spoiling ability. The purified


D-LDH of those groups had different pH and temperature optima and isoelectric points. Especially the
temperature optimum of 50 jC of the Group B DLDH is significantly lower than that of the other DLDHs (60 jC).
Except for the forcing test, none of the other
methods so far available can predict beer-spoiling
ability of all strains. Since hop resistance in lactic
acid bacteria does not appear to be a general feature of
all beer-spoiling bacteria, combination of several
methods will be required to detect all potential beer
spoiling strains.

References
Anon, Wil N., 1997. Validation of the publication of new names
and new combinations previously effectively published outside
the IJSB. Int. J. Syst. Bacteriol. 47, 601 602.
Archibald, F.S., Duong, M.-N., 1984. Manganese acquisition by
Lactobacillus plantarum. J. Bacteriol. 158, 1 8.
Archibald, F.S., Fridovich, I., 1981a. Manganese and defenses
against oxygen toxicity in Lactobacillus plantarum. J. Bacteriol.
145, 442 451.
Archibald, F.S., Fridovich, I., 1981b. Manganese, superoxide
dismutase, and oxygen tolerance in some lactic acid bacteria.
J. Bacteriol. 146, 928 936.
Back, W., 1978. Zur Taxonomie der Gattung Pediococcus, Brauwissenschaft 31, 237 250, 312 320, 336.
Back, W., 1980. Bierschadlich Bakterien. Nachweis und Kultivierung bierschadlicher Bakterien im Betriebslabor. Brauwelt 120,
1562.
Back, W., 1981. Beer spoilage bacteria. Taxonomy of beer spoilage
bacteria. Gram positive species. Mon.schr. Brau. 34, 267 276.
Back, W., 1987. Neubeschreibung einer bierschaedlichen Laktobazillen-Art. Lactobacillus brevisimilis spec. nov. Mon.schr. Brauwiss. 40, 484 488.
Back, W., 1994. Secondary contamination in the filling area. Brauwelt Int. 4, 326 328.
Back, W., Stackebrandt, E., 1978. DNS/DNS-Homologie studien
innerhalb der Gattung Pediococcus. Arch. Mikrobiol. 118,
79 85.
Back, W., Weiss, N., Seidel, H., 1979. Isolierung und systematische
Zuordnung bierschadlicher gramnegativer Bakterien. Brauwissenschaft 32, 233 238.
Back, W., Breu, S., Weigand, C., 1988. Infektionsursachen im jahre
1987. Brauwelt 178, 1358 1362.
Back, W., Leibhard, M., Bohak, I., 1992. Flash pasteurization
membrane filtration. Comparative biological safety. Brauwelt
Int. 12, 42 49.
Back, W., Bohak, I., Ehrmann, M., Ludwig, W., Schleifer, K.H.,
1996. Revival of the species Lactobacillus lindneri and the

119

design of a species specific oligonucleotide probe. Syst. Appl.


Microbiol. 19, 322 325.
Barney, M., Kot, E., 1992. A comparison of rapid microbiological
methods for detecting beer spoilage microorganisms. MBAA
T.Q. 29, 91 95.
Barney, M., Volgyi, A., Navarro, A., Ryder, D., 2001. Riboprinting
and 16S rRNA gene sequencing for identification of brewery
Pediococcus isolates. Appl. Environ. Microbiol. 67, 553 560.
Bischoff, E., Bohak, I., Back, W., Leibhard, S., 2001. Schnellnachweis von bierschadlichen Bakterien mit PCR und universellen
Primern. Mon.schr. Brau. 54, 4 8.
Brewery Convention of Japan, S., 1999. BCOJ Biseibutu Bunsekihou, 8.2. Brewing Society of Japan, Tokyo. In Japanese.
Bunker, H.J., 1955. The survival of pathogenic bacteria in beer.
Proc. Eur. Brew. Conv. Baden-Baden. Elsevier Publishing Company, Houston, pp. 330 341.
Chelack, B.J., Ingledew, W.M., 1987. Anaerobic Gram-negative
bacteria in brewinga review. J. Am. Soc. Brew. Chem. 45,
123 127.
Claussen, O.G., Hegna, I.K., Solberg, O., 1975. Agglutination
tests as an aid to the classification of pediococci and aerococci.
J. Inst. Brew. 81, 440 443.
Claussen, O.G., Hegna, I.K., Kjeldaas, L., Palmer, L., Tjade, O.H.,
1981. Precipitation tests as an aid to the classification of pediococci and aerococci. J. Inst. Brew. 87, 141 146.
Collins, M.D., Williams, A.M., Wallbanks, S., 1990. The phylogeny of Aerococcus and Pediococcus as determined by 16S
rRNA sequence analysis: description of Tetragenococcus gen
nov. FEMS Microbiol. Lett. 70, 255 262.
de Keukeleire, D., Verzele, M., 1970. The structure and absolute
configuration of ( ) humulone. Tetrahedron 26, 385 393.
de Keukeleire, D., Verzele, M., 1971. The absolute configuration of
the isohumulones and the humulinic acids. Tetrahedron 27,
4939 4945.
DiMichele, L.J., Lewis, M.J., 1993. Rapid species-specific detection of lactic acid bacteria from beer using polymerase chain
reaction. J. Am. Soc. Brew. Chem. 51, 63 66.
Dolezil, L., Kirsop, B.H., 1976. The detection and identification of
Pediococcus and Micrococcus in breweries using a serological
method. J. Inst. Brew. 82, 93 95.
Dowhanick, T.M., 1995. Advances in yeast and contaminant determination. The future of the so-called rapid methods. Cerevisia
Belg. J. Brew. Biotechnol. 20, 40 50.
Dowhanick, T.M., Russell, I., 1993. Beer and wine production. In:
Gump, B.H. (Ed.), ACS Symp. Ser. vol. 536. American Chemical Society, Washington, D.C., pp. 13 30.
Doyle, L.M., McInerney, J.O., Mooney, J., Powell, R., Haikara, A.,
Moran, A.P., 1995. Sequence of the gene encoding 16S rRNA of
the spoilage organism Megasphaera cerevisiae. J. Ind. Microbiol. 15, 67 70.
European Brewery Convention, A.P., 1992. Detection of Contaminants. EBC Analytica Microbiologica II, Section 4 European
Brewery Convention, Fachverlag Hans Carl, Nurnberg.
Farrow, J.A.E., Phillips, B.A., Collins, M.D., 1988. Nucleic acid
studies on some heterofermentative lactobacilli, Description of
Lactobacillus malefermentans sp. nov. and Lactobacillus parabuchneri sp. nov. FEMS Microbiol. Lett. 55, 163 168.

120

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

Fernandez, J.L., Simpson, W.J., 1993. Aspects of the resistance of


lactic acid bacteria to hop bitter acids. J. Appl. Bacteriol. 75,
315 319.
Fernandez, J.L., Simpson, W.J., 1995a. Measurement and prediction of the susceptibility of lager beer to spoilage by lactic acid
bacteria. J. Appl. Bacteriol. 78, 419 425.
Fernandez, J.L., Simpson, W.J., 1995b. A cost-effective solution to
control of microbiological stability? Proc. Eur. Conv. Brussels.
Oxford University Press, New York, pp. 713 722.
Funahashi, W., Suzuki, K., Ohtake, Y., Yamashita, H., 1998. Two
novel beer-spoilage Lactobacillus species isolated from breweries. J. Am. Soc. Brew. Chem. 56, 64 69.
Gares, S.L., Whiting, M.S., Ingledew, W.M., Ziola, B., 1993.
Detection and identification of Pectinatus cerevisiiphilus using surface-reactive monoclonal antibodies in a membrane
filter-based fluoroimmunoassay. J. Am. Soc. Brew. Chem. 51,
158 163.
Haikara, A., 1983. Immunological characterization of Pectinatus cerevisiiphilus strains. Appl. Environ. Microbiol. 46, 1054 1058.
Haikara, A., 1991. The genera Pectinatus and Megasphaera. In:
Balows, A., Truper, H.G., Dworkin, M., Harder, W., Schleifer,
K.-H. (Eds.), The Procaryotes. Springer-Verlag, New York, pp.
1993 2004.
Haikara, A., Lounatmaa, K., 1987. Characterization of Megasphaera sp. A new anaerobic beer spoilage coccus. Proc. Eur.
Brew. Conv., Madrid. IRL Press Ltd., Oxford, pp. 473 480.
Haikara, A., Penttila, L., Enari, T.M., Lounatmaa, K., 1981. Microbiological, biochemical, and electron microscopic characterization of a Pectinatus strain. Appl. Environ. Microbiol. 41,
511 517.
Hammond, J.R.M., 1996. Brewing microbiology 100 years after
Pasteur. Cerevisia Belg. J. Brew. Biotechnol. 21, 59 62.
Hammond, J., Brennan, M., Price, A., 1999. The control of microbial spoilage of beer. J. Inst. Brew. 105, 113 120.
Harris, J.O., Watson, W., 1960. Observation on some beer spoilage
lactobacilli. J. Inst. Brew. 66, 151 154.
Hastings, J.H.H., Walker, T.K., 1928a. Report on the preservative
principles of hops: Part VIII. Some suggested modifications of
the gravimetric method for the evaluation of hops. J. Inst. Brew.
34, 9 13.
Hastings, J.H.H., Walker, T.K., 1928b. Report on the preservative
principles of hops: Part IX. The influence of special methods of
drying at low temperatures upon the antiseptic properties of
hops. J. Inst. Brew. 34, 556 565.
Hastings, J.H.H., Walker, T.K., 1929. Report on the preservative
principles of hops: Part X. A simplifying modification of Ford
and Taits gravimetric process for the evaluation of hops. J. Inst.
Brew. 35, 223 229.
Hastings, J.H.H., Pyman, F.L., Walker, T.K., 1926. Report on the
preservative principles of hops: Part VII. A comparison of methods of determining the preservative properties of hops. J. Inst.
Brew. 32, 484 502.
Haukeli, A.D., 1980. En ny lskadelig bakterie i tappet 1. Referat
fran 18. Det Skan Dinaviska. Bryggeritekniska Motet, Stockholm, pp. 111 122.
Hayashi, N., Ito, M., Horiike, S., Taguchi, H., 2001. Molecular
cloning of a putative divalent-cation transporter gene as a new

genetic marker for the identification of Lactobacillus brevis


strain capable of growing in beer. Appl. Microbiol. Biotechnol.
55, 596 603.
Hayduck, F., 1888. Making tetrahydroisoalpha acids and hexahydroisoalpha acids by catalytic reduction and isomerization. Wochenschr. Brau. 5, 937.
Hollerova, I., Kubizniakova, P., 2001. Monitoring Gram positive
bacterial contamination in Czech breweries. J. Inst. Brew. 107,
355 358.
Hough, J.S., Howard, G.A., Slater, C.A., 1957. Bacteriostatic activities of hop resin materials. J. Inst. Brew. 63, 331 333.
Hough, J.S., Briggs, D.E., Stevens, R., 1971. Malting and Brewing
Science. Chapman & Hall, London, p. 324.
Hough, J.S., Briggs, D.E., Stevens, R., Young, T.W., 1982. Malting
and Brewing Science, vol. 2, 2nd ed. Chapman & Hall, London.
Hughes, P., 2000. The significance of iso-a-acids for beer quality,
Cambridge prize paper. J. Inst. Brew. 106, 271 276.
Ingram, L.O., Buttke, T.M., 1984. Effects of alcohols on microorganisms. Adv. Microb. Physiol. 25, 253 300.
Inoue, T., 1981. The relationship between the performance of yeast
and acetohydroxy acid formation during wort fermentation.
MBAA T.Q. 18, 62 65.
Jespersen, L., Jakobsen, M., 1996. Specific spoilage organisms in
breweries and laboratory media for their detection. Int. J. Food
Microbiol. 33, 139 155.
Jonsson, H., Petterson, H.-E., 1977. Studies on the citric acid fermentation in lactic starter cultures with special interest in a-acetolactic: 2. Metabolic studies. Milchwissenschaft 32, 587 594.
Juvonen, R., Satokari, R., 1999. Detection of spoilage bacteria in
beer by polymerase chain reaction. J. Am. Soc. Brew. Chem. 57,
99 103.
Kirchner, G., Lurz, R., Matsuzawa, K., 1980. Biertrubungen
durch Bakterien der Gattung Bacteroides. Mon.schr. Brau.
33, 461 467.
Lawrence, D.R., 1988. Spoilage organisms in beer. In: Robinson,
R.K. (Ed.), Developments in Food Microbiology. Elsevier, London, pp. 1 48.
Lawrence, D.R., Priest, F.G., 1981. Identification of brewery cocci.
Proc. Eur. Brew. Conv., Copenhagen. IRL Press Ltd., Oxford,
pp. 217 227.
Lee, S.Y., 1994. SMMPa medium for selective isolation for Megasphaera and Pectinatus from the brewery. J. Am. Soc. Brew.
Chem. 52, 115 119.
Lee, S.Y., Mabee, M.S., Jangaard, N.O., 1978. Pectinatus, a new
genus of the family bacteriodaceae. Int. J. Syst. Bacteriol. 28,
582 594.
Lee, S.Y., Mabee, M.S., Jangaard, N.O., Horiuchi, E.K., 1980.
Pectinatus, a new genus of bacteria capable of growth in hopped
beer. J. Inst. Brew. 86, 28 30.
McCaig, R., Weaver, R.L., 1983. Physiological studies on pediococci. MBAA T.Q. 20, 31 38.
Moir, M., 2000. Hopsa millenium review. J. Am. Soc. Brew.
Chem. 58, 131 146.
Motoyama, Y., Ogata, T., 2000. Detection of Pectinatus spp. by
PCR using 16S-23S rDNA spacer regions. J. Am. Soc. Brew.
Chem. 58, 4 7.
Motoyoma, Y., Ogata, T., Sakai, K., 1998. Characterization of Pec-

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124


tinatus cerevisiiphilus and P. frisingensis by ribotyping. J. Am.
Soc. Brew. Chem. 56, 19 23.
Motoyama, Y., Funahashi, W., Ogata, T., 2000. Characterization of
Lactobacillus spp. by ribotyping. J. Am. Soc. Brew. Chem. 58,
1 3.
Nakakita, Y., Takahashi, T., Sugiyama, H., Shigyo, T., Shinotsuka,
K., 1998. Isolation of novel beer-spoilage bacteria from brewery
environment. J. Am. Soc. Brew. Chem. 56, 114 117.
Okazaki, Z., Usui, Y., Suzuki, K., 1997. Early evaluation of lactobacilli for production of beer. Japanese Patent, JP2869860.
Pasteur, M.L., 1876. Etudes sur La Biere. Gauthier-Villars, Imprimeur-Libraire, Paris.
Perpete, P., van Cutsem, P., Boutte, C., Colson-Corbisier, A.-M.,
Collin, S., 2001. Amplified fragment-length polymorphism, a
new method for the analysis of brewers yeast DNA polymorphism. J. Am. Soc. Brew. Chem. 59, 195 200.
Price, A., Stapely, S.J., 2001. The antibacterial activity of hop isoa-acids. BRi Q. 1, 20 21.
Priest, F.G., 1987. Gram-positive brewery bacteria. In: Priest, F.G.,
Cambell, I. (Eds.), Brewing Microbiology. Elsevier, London,
pp. 121 154.
Priest, F.G., 1996. Gram-positive brewery bacteria. In: Priest, F.G.,
Cambell, I. (Eds.), Brewing Microbiology, 2nd ed. Chapman &
Hall, London, pp. 127 161.
Putman, M., van Veen, H.W., Konings, W.N., 2000. Molecular
properties of bacterial multidrug transporters. Microbiol. Mol.
Biol. Rev. 64, 672 693.
Pyman, F.L., Rogerson, H., Walker, T.K., 1922. Report on the preservative principles of hops. Part I. J. Inst. Brew. 28, 929 934.
Rainbow, C., 1981. Beer spoilage microorganisms. In: Rainbow, C.
(Ed.), Brewing Science, vol. 2. Academic Press, London, pp.
491 550.
Richards, M., Macrae, R.M., 1964. The significance of the use of
hops in regard to the biological stability of beer: II. The development of resistance to hop resins by strains of lactobacilli. J. Inst.
Brew. 70, 484 488.
Rigby, F.L., Bethune, J.L., 1952. Countercurrent distribution of hop
constituents. Proc. - Am. Soc. Brew. Chem., 98 105.
Rigby, F.L., Bethune, J.L., 1953. Countercurrent distribution of
hop constituents. The isolation and properties of cohumulone.
Proc. - Am. Soc. Brew. Chem., 119 129.
Rinck, M., Wackerbauer, K., 1987a. Detection and serological classification of lactic acid bacteria, especially L. lindneri. Mon.schr.
Brauwiss. 40, 324 327.
Rinck, M., Wackerbauer, K., 1987b. Detection and serological classification of lactic acid bacteria, especially L. lindneri. Brauwelt
Int. 2, 181 184.
Sakamoto, K., 1997. Detection of the genus Pectinatus. Japanese
Patent Application JP09-520359, World Patent Application
WO97/20071, US Patent 586942.
Sakamoto, K., Funahashi, W., Yamashita, H., Eto, M., 1997. A
reliable method for detection and identification of beer-spoilage
bacteria with internal positive control PCR (IPC-PCR). Proc.
Eur. Brew. Conv. Maastricht. Oxford University Press, New
York, pp. 631 638.
Sakamoto, K., Margolles, A., van Veen, H.W., Konings, W.N.,
2001. Hop resistance in the beer-spoilage bacterium Lactobacil-

121

lus brevis is mediated by the ATP-binding cassette multidrug


transporter HorA. J. Bacteriol. 183, 5371 5375.
Sakamoto, K., van Veen, H.W., Saito, H., Kobayashi, H., Konings,
W.N., 2002. Membrane bound ATPase contributes to hop resistance of Lactobacillus brevis. Appl. Environ. Microbiol. 68,
5374 5378.
Sami, M., 1999. PhD thesis. The university of Tokyo.
Sami, M., Yamashita, H., Hirono, T., Kadokura, H., Kitamoto, K.,
Yoda, K., Yamasaki, M., 1997a. Hop-resistant Lactobacillus
brevis contains a novel plasmid harboring a multidrug resistance-like gene. J. Ferment. Bioeng. 84, 1 6.
Sami, M., Yamashita, H., Kadokura, H., Kitamoto, K., Yoda, K.,
Yamasaki, M., 1997b. A new and rapid method for determination of beer-spoilage ability of lactobacilli. J. Am. Soc. Brew.
Chem. 55, 137 140.
Sami, M., Suzuki, K., Sakamoto, K., Kadokura, H., Kitamoto, K.,
Yoda, K., 1998. A plasmid pRH45 of Lactobacillus brevis confers hop resistance. J. Gen. Appl. Microbiol. 44, 361 363.
Sato, A., Ohtake, Y., Mori, T., Miyamoto, Y., 1994. A monoclonal
antibody specific for lactic acid bacteria. Japanese Patent Application JP06-046881.
Satokari, R., Juvonen, R., von Wright, A., Haikara, A., 1997. Detection of Pectinatus beer spoilage bacteria by using the polymerase chain reaction. J. Food Prot. 60, 1571 1573.
Satokari, R., Juvonen, R., Mallison, K., Wright, A.V., Haikara, A.,
1998. Detection of beer spoilage bacteria Megasphaera and
Pectinatus by polymerase chain reaction and colorimetric microplate hybridization. Int. J. Food Microbiol. 45, 119 127.
Satokari, R., Mattia-Sandholm, T., Suihko, M.-L., 2000. Identification of pediococci by ribotyping. J. Appl. Microbiol. 88,
260 265.
Savard, L., Hutchinson, J.N.G., Dowhanick, T.M., 1994. Characterization of different isolates of Obesumbacterium proteus using
randam amplified polymorophic DNA. J. Am. Soc. Brew.
Chem. 52, 62 65.
Schleifer, K.H., Leuteritz, M., Weiss, N., Ludwig, W., Kirchhof, G.,
Seidel-Rufer, H., 1990. Taxonomic study of anaerobic, Gramnegative, rod-shaped bacteria from breweries, Emended description of Pectinatus cerevisiiphilus and description of Pectinatus
frisingensis sp. nov., Selenomonas lacticifex. sp. nov., Zymophilus raffinosivorans gen. nov., sp. nov., and Zymophilus paucivorans sp. nov. Int. J. Syst. Bacteriol. 40, 19 27.
Schmalreck, A.F.M., Teuber, W., Reininger, W., Hartl, A., 1975.
Structural features determining the antibiotic potencies of natural and synthetic hop bitter resins, their precursors and derivatives. Can. J. Microbiol. 21, 205 212.
Schmidt, H.-J., 1992. Mikrobiologische SchnellnachweisverfahrenEine kritische Betrachtung. Brauwelt 132, 402 409.
Schmidt, H.-J., 1999. Neue und alte Wege in der mikrobiologischen
Absicherung. Brauwelt 139, 1166 1171.
Schofield, M., 1995. The use of PCR technology. BRFI Q. Jan,
20 21.
Seidel, H., Back, W., Weiss, N., 1979. Isolation and systemic
classification of beer spoilage, Gram-negative bacteria: Part
III. What dangers do the Gram-negative cocci and rods described in parts I and II represent for beer? Brauwissenschaft
32, 262 270.

122

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

Seidel-Rufer, H., 1990. Pectinatus and other morphologically similar Gram negative anaerobic bacilli from brewing. Mon.schr.
Brauwiss. 43, 101 105.
Sharpe, M.E., 1955. A serological classification of lactobacilli.
J. Gen. Microbiol. 12, 107 122.
Sharpe, M.E., 1979. In: Skinner, F.A., Lovelock, D.W. (Eds.), Identification Methods for Microbiologists, 2nd ed. Academic Press,
London, p. 233.
Sharpe, M.E., Davison, A.L., Baddiley, J., 1964. Teichoic acids and
group antigens in lactobacilli. J. Gen. Microbiol. 34, 333 340.
Shimwell, J.L., 1936. Saccharrobacillus infection. J. Inst. Brew. 42,
452 460.
Shimwell, J.L., 1937a. On the relation between the staining properties of bacteria and their reaction towards hop antiseptic. Part I,
II. J. Inst. Brew. 43, 111 118.
Shimwell, J.L., 1937b. On the relation between the staining properties of bacteria and their reaction towards hop antiseptic. Part III.
J. Inst. Brew. 43, 191 195.
Simpson, W.J., 1991. Molecular structure and antibacterial function
of hop resin materials. PhD thesis, Council for National Academic Awards, UK.
Simpson, W.J., 1993a. Cambridge prize lecture, Studies on the
sensitivity of lactic acid bacteria to hop bitter acids. Inst. Brew.
99, 405 411.
Simpson, W.J., 1993b. Ionophoric action of trans-isohumulone on
Lactobacillus brevis. J. Gen. Microbiol. 139, 1041 1045.
Simpson, W.J., Fernandez, J.L., 1992. Selection of beer-spoilage
lactic acid bacteria and induction of their ability to grow in beer.
Lett. Appl. Microbiol. 14, 13 16.
Simpson, W.J., Fernandez, J.L., 1994. Mechanism of resistance of
lactic acid bacteria to trans-isohumulone. J. Am. Soc. Brew.
Chem. 52, 9 11.
Simpson, W.J., Hughes, P.S., 1993. Cooperative binding of potassium ions to trans-isohumulone in the presence of divalent and
trivalent cations. Bioorg. Med. Chem. Lett. 3, 79 772.
Simpson, W.J., Smith, A.R.W., 1992. Factors affecting antibacterial
activity of hop compounds and their derivatives. J. Appl. Bacteriol. 72, 327 334.
Simpson, W.J., Taguchi, H., 1995. The genus Pediococcus, with
notes on the genera Tetragenococcus and Aerococcus. In: Wood,
B.J.B., Holzapfel, W.H. (Eds.), The Genera of Lactic Acid Bacteria. The Lactic Acid Bacteria, vol. 2. Blackie Academic &
Professional, London, pp. 125 172.
Simpson, W.J., Fernandez, J.L., Hughes, P.S., Parker, D.K., Price,
A.C., 1993. The chemistry of iso-a-acids, an explanation of
their mode of action. Proc. Eur. Brew. Conv. Oslo. Oxford University Press Inc., New York, pp. 183 192.
Soberka, R., Sciazko, D., Warzecha, A., 1988. Pectinatus, a new
bacterium potentially able to affect the biological stability of
wort and beer. Part 1. Bios 19, 31 37.
Soberka, R., Sciazko, D., Warzecha, A., 1989. Pectinatus, a new
bacterium potentially able to affect the biological stability or
wort and beer. Part 2. Bios 20, 51 59.
Speckman, R.A., Collins, E.B., 1968. Diacetyl biosynthesis in
Streptococcus diacetilactis and Leuconostoc citrovorum. J. Bacteriol. 95, 174 180.
Speckman, R.A., Collins, E.B., 1973. Incorporation of radioactive

acetate into diacetyl by Streptococcus diacetilactis. Appl. Microbiol. 26, 744 746.
Stewart, R.J., Dowhanick, T.M., 1996. Rapid detection of lactic
acid bacteria in fermenter samples using a nested polymerase
chain reaction. J. Am. Soc. Brew. Chem. 54, 78 84.
Storgards, E., Suihko, M.-L., Pot, B., Vanhonacker, K., Janssens,
D., Broomfield, P.L.E., Banks, J.G., 1998. Detection and identification of Lactobacillus lindneri from brewery environments.
J. Inst. Brew. 104, 47 54.
Suihko, M.-L., Haikara, A., 2000. Characterization of Pectinatus
and Megasphaera strains by automated ribotyping. J. Inst. Brew.
107, 175 184.
Suzuki, K., Sami, M., Kadokura, H., Nakajima, H., Kitamoto, K.,
2002. Biochemical characterization of horA-independent hop
resistance mechanism in Lactobacillus brevis. Int. J. Food Microbiol. 76, 223 230.
Takahashi, N., 1983. Presumed Pectinatus strain isolated from Japanese beer. Bull. Brew. Sci. 28, 11 14.
Takahashi, T., Nakakita, Y., Sugiyama, H., Shigyo, T., Shinotsuka,
K., 1999. Classification and identification of strains of Lactobacillus brevis based on electrophoretic characterization of Dlactate dehydrogenase, Relationship between D-lactate dehydrogenase and beer-spoilage ability. J. Biosci. Bioeng. 88,
500 506.
Teuber, M., Schmalreck, A.F., 1973. Membrane leakage in Bacillus subtilis 168 induced by the hop constituents lupulone,
humulone, isohumulone and humulic acid. Arch. Mikrobiol.
94, 159 171.
The Technical Committee and the Editorial Committee of the
American Society of Brewing Chemists, A.F., 1992. Microbiology. Microbiological control. Methods of Analysis of the American Society of Brewing Chemists, 8th Revised ed. with 1996
and 1999 Supplements. American Society of Brewing Chemists,
Minnesota, USA.
Thompson, A.N., Wright, D.M., Pawson, E.C., Meaden, P.G., 1994.
Polymerase chain reaction tests for lactic acid bacteria. Proc.
Aviemore Conf. Malt Brew. Distill. 4, 213 216.
Tompkins, T.A., Stewart, R., Savard, L., Russell, I., Dowhanick,
T.M., 1996. RAPD-PCR characterization of brewery yeast and
beer spoilage bacteria. J. Am. Soc. Brew. Chem. 54, 91 96.
Tsuchiya, Y., Kaneda, H., Kano, Y., Koshino, S., 1992. Detection of
beer spoilage organisms by polymerase chain reaction technology. J. Am. Soc. Brew. Chem. 50, 64 67.
Tsuchiya, Y., Kano, Y., Koshino, S., 1993. Detection of Lactobacillus brevis in beer using polymerase chain reaction technology.
J. Am. Soc. Brew. Chem. 51, 40 41.
Tsuchiya, Y., Kano, Y., Koshino, S., 1994. Identification of lactic
acid bacteria using temperature gradient gel electrophoresis for
DNA fragments amplified by polymerase chain reaction. J. Am.
Soc. Brew. Chem. 52, 95 99.
Tsuchiya, Y., Nakakita, Y., Watari, J., Shinotsuka, K., 2000. Monoclonal antibodies specific for the beer-spoilage ability of lactic
acid bacteria. J. Am. Soc. Brew. Chem. 58, 89 93.
Uchida, K., 1974. Occurrence of saturated and mono-unsaturated
fatty acids with unusually-long-chains (C20 C30) in Lactobacillus heterohiochii, an alcoholophilic bacterium. Biochim. Biophys. Acta 348, 86 93.

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124


van Veen, H.W., Venema, K., Bolhuis, H., Oussenko, I., Kok, J.,
Poolman, B., Driessen, A.J.M., Konings, W.N., 1996. Multidrug
resistance mediated by a bacterial homologue of the human
multidrug transporter MDR1. Proc. Natl. Acad. Sci. U. S. A.
93, 10668 10672.
van Vuuren, H.J.J., 1996. Gram-negative spoilage bacteria. In:
Priest, F.G., Campbell, I. (Eds.), Brewing Microbiology, 2nd
ed. Elsevier, London, pp. 163 191.
Verzele, M., 1986. Centenary review, 100 years of hop chemistry
and its relevance to brewing. J. Inst. Brew. 92, 32 48.
Vogesser, G., Donhauser, S., Winnewisser, W., 1995a. Rapid detection of beer spoilage Lactobacillus species with the polymerase
chain reaction (PCR). Proc. Eur. Brew. Conv. Brussels. Oxford
University Press Inc., New York, pp. 629 636.
Vogesser, G., Winnewisser, W., Geiger, E., 1995b. Anwendungen
der PCR im Brauerebereich. Schnellnachweis von bierschaedlichen Lactobacillen (Teil 1). Brauwelt 135, 1635 1637.
Walker, T.K., 1923a. Report on the preservative principles of hop:
Part II. The nature of the preservative principles of hops. J. Inst.
Brew. 29, 373 378.
Walker, T.K., 1923b. Report on the preservative principles of hops:
Part III. The extraction of humulon and lupulon from hop-cones
and a new method for the estimation of the resin content of the
cones. J. Inst. Brew. 29, 379 399.
Walker, T.K., 1924a. Report on the preservative principles of hops:
Part IV. Improvements in the method of isolation of lupulon and
a preliminary examination of the other constituents of the soft
resins. J. Inst. Brew. 30, 570 578.
Walker, T.K., 1924b. Report on the preservative principles of hops:
Part V. The chemical constitution of lupulon. J. Inst. Brew. 30,
712 724.
Walker, T.K., 1925. Report on the preservative principles of hops:
Part VI. Determination of the relative antiseptic values of the
soft resin constituents. J. Inst. Brew. 31, 562 577.
Walker, T.K., 1938. Report on the preservative principles of hops:
Part XIX. Quantitative studies of the changes in preservative
value during the boiling and fermentation of hopped worts.
Section I. and II. J. Inst. Brew. 44, 11 28.
Walker, T.K., 1941. Report on the preservative principles of hops:
Part XX. A summary of the investigations described in Parts I to
XIX and a discussion of the significance of the principal results.
J. Inst. Brew. 47, 362 370.
Walker, T.K., Blakebrough, N., 1952. Bacteriostatic power of humulone boiling-product. J. Inst. Brew. 58, 13 24.
Walker, T.K., Hastings, J.H.H., 1931. Report on the preservative
principles of hops: Part XI. Some preliminary observations of
the effects produced by certain constituents of hop resin during
the boiling of hopped wort. J. Inst. Brew. 37, 509 512.
Walker, T.K., Hastings, J.H.H., 1933a. Report on the preservative
principles of hops: Part XIV. Observations on the development
of the antiseptic constituents of hops and of the tannin during
ripening. J. Inst. Brew. 39, 15 27.
Walker, T.K., Hastings, J.H.H., 1933b. Report on the preservative
principles of hops: Part XV. The gravimetric estimation of the
antiseptic constituents of hops. J. Inst. Brew. 39, 509 512.
Walker, T.K., Parker, A., 1936. Report on the preservative principles of hops: Part XVII. Colour-standards for use in the gravi-

123

metric determination of the antiseptic constituents of hops.


J. Inst. Brew. 42, 267 272.
Walker, T.K., Parker, A., 1937. Report on the preservative principles of hops: Part XVIII. The theoretical basis of the log phase
method for the evaluation of bacteriostatic power, and the procedure in using phenol as a standard of value. J. Inst. Brew. 43,
17 30.
Walker, T.K., Parker, A., 1938. Report on the preservative principles of hops: Part XIX. Quantitative studies of the changes in
preservative value during the boiling and fermentation of hopped worts. Section III. The adsorption of hop antiseptics on wort
colloids. J. Inst. Brew. 44, 140 163.
Walker, T.K., Parker, A., 1940a. Report on the preservative principles of hops: Part XIX. Quantitative studies of the changes in
preservative value during the boiling and fermentation of hopped worts. Section IV. Deductions from experimental data obtained by application of the log phase method for estimation of
the bacteriostatic powers of hop resin constituents. The expression of the bacteriostatic powers of worts and beers in terms of
pure humulon. J. Inst. Brew. 46, 235 242.
Walker, T.K., Parker, A., 1940b. Report on the preservative principles of hops: Part XIX. Quantitative studies of the changes in
preservative value during the boiling and fermentation of hopped worts. Section VI. Changes in preservative value during
brewery processes. J. Inst. Brew. 46, 337 361.
Walker, T.K., Hastings, J.H.H., Farrar, E.J., 1931. Report on the
preservative principles of hops: Part XII. A method for the
quantitative comparison of very small degrees of antiseptic activity and of very small differences between more pronounced
degrees of activity. J. Inst. Brew. 37, 512 533.
Walker, T.K., Hastings, J.H.H., Farrar, E.J., Day, F.E., 1932. Report on the preservative principles of hops: Part XIII. A method for the quantitative comparison of the relative stabilities of
hopped worts before and after fermentations. J. Inst. Brew. 38,
303 312.
Walker, T.K., Hastings, J.H.H., Vero, E., 1935. Report on the preservative principles of hops: Part XVI. A study of the action of
antiseptics on the logarithmic phase of the growth of B. bulgaricus and the selection of a standard of antiseptic value for use in
the log-phase method for the determination and comparison of
antiseptic powers. J. Inst. Brew. 41, 92 102.
Walker, T.K., Hastings, J.H.H., Parker, A., 1940. Report on the
preservative principles of hops: Part XIX. Quantitative studies
of the changes in preservative value during the boiling and
fermentation of hopped worts. Section V. The extent to which
the preservative power of humulon, of lupulon and of h-soft
resin, respectively, is lost during boiling in wort and subsequently during fermentation of the boiled wort. J. Inst. Brew.
46, 304 320.
Watier, D., Lenguerinel, I., Hornez, J.-P., Dubourguier, H.-C., 1993.
Influence de facteurs physico-chmiques sur les cinetiques de
croissance de Pectinatus cerevisiiphilus et de Megasphaera cerevisiae. Sci. Aliments 13, 297 304.
Weiss, N., Seidel, H., Back, W., 1979. Isolation and systematic
classification of beer spoilage, Gram-negative bacteria: Part I.
Gram-negative strictly-anaerobic cocci. Brauwissenschaft 32,
189 194.

124

K. Sakamoto, W.N. Konings / International Journal of Food Microbiology 89 (2003) 105124

Whiting, M., Crichlow, M., Ingledew, W.M., Ziola, B., 1992. Detection of Pediococci spp. in brewing yeast by a rapid immunoassay. Appl. Environ. Microbiol. 58, 713 716.
Whiting, M.S., Gares, S.L., Ingledew, W.M., Ziola, B., 1999a.
Brewing spoilage lactobacilli detected using monoclonal antibodies to bacterial surface antigen. Can. J. Microbiol. 45,
51 58.
Whiting, M.S., Ingledew, W.M., Lee, S.Y., Ziola, B., 1999b.
Bacterial surface antigen-specific monoclonal antibodies used
to detect beer spoilage pediococci. Can. J. Microbiol. 45,
670 677.
Wieland, H., Martz, E., Hoek, H., 1925. Chemical nature of the hop
resin acids. Chem. Ber. 58B, 2012 2017.
Windisch, W., Kolbach, P., Schleicher, R., 1927. Transformation of
the alpha-bitter acid of hops in boiling aqueous solutions of
various reactions and the nature of the products formed. Wochenschr. Brau. 44, 453 459, 473 478, 485 490, 497 502.
Yamamoto, Y., 1958. Bakushu Kougyou Yuugaikinrui. Jyouzou
Kagaku Kenkyuusho Suita Kenkyuusho (in Japanese).
Yasui, T., 1995. Detection and identification of lactic acid bacteria.
Japanese Patent Application JP07-289295.
Yasui, T., Yoda, K., 1997a. The group E antigen is masked by the
paracrystalline surface layer in Lactobacillus brevis. J. Ferment.
Bioeng. 84, 35 40.
Yasui, T., Yoda, K., 1997b. Purification and partial characterization

of antigen specific to Lactobacillus brevis strains with beer


spoilage activity. FEMS Microbiol. Lett. 151, 169 176.
Yasui, T., Taguchi, H., Kamiya, T., 1992. Method of group antigen exposure to the cell surface for rapid detection of Lactobacillus brevis using enzyme immunoassay. Proc. Inst. Brew.
22, 190 192.
Yasui, T., Yoda, K., Kamiya, T., 1995. Analysis of S-layer proteins
of Lactobacillus brevis. FEMS Microbiol. Lett. 131, 181 186.
Yasui, T., Okamoto, T., Taguchi, H., 1997. A specific oligonucleotide primer for the rapid detection of Lactobacillus lindneri by
polymerase chain reaction. Can. J. Microbiol. 43, 157 163.
Ziola, B., Gares, S.L., Lorrain, B., Ingledew, W.M., Lee, S.Y., 1999.
Epitope mapping of monoclonal antibodies specific for directly
cross-linked mesodiaminopimelic acid peptidoglycan found in
the anaerobic beer spoilage bacterium Pectinatus cerevisiiphilus.
Can. J. Microbiol. 45, 779 785.
Ziola, B., Gee, L., Berg, N.N., Lee, S.Y., 2000a. Serogroup of the
beer spoilage bacterium Megasphaera cerevisiae correlate with
the molecular weight of the major EDTA-extractable surface
protein. Can. J. Microbiol. 46, 95 100.
Ziola, B., Ulmer, M., Bueckert, J., Giesbrecht, D., 2000b. Monoclonal antibodies showing surface reactivity with Lactobacillus
and Pediococcus beer spoilage bacteria. J. Am. Soc. Brew.
Chem. 58, 63 68.

You might also like