27
33 Biochemistry of Beer Fermentation Ronnie Willaert Introduction The Beer Brewing Process Carbohydrate Metabolism—Ethanol Production Wort Carbohydrates Uptake and Metabolism Maltose and Maltotriose Metabolism Glycogen and Trehalose Metabolism Wort Fermentation Metabolism of Bioflavoring by-Products Biosynthesis of Higher Alcohols Biosynthesis of Esters Biosynthesis of Organic Acids Biosynthesis of Vicinal Diketones Secondary Fermentation Vicinal Diketones Hydrogen Sulfide Acetaldehyde Development of Flavor Fullness Beer Fermentation Using Immobilized Cell Technology Carrier Materials Applications of ICT in the Brewing Industry Flavor Maturation of Green Beer Production of Alcohol-Free or Low-Alcohol Beer Production of Acidified Wort Using Immobilized Lactic Acid Bacteria Continuous Main Fermentation Acknowledgments References Abstract: The carbohydrate metabolism and flavor formation dur- ing yeast primary and secondary fermentation (maturation) is re- viewed. Carbohydrate metabolism and ethanol production during the primary fermentation is discussed firstly. Next, the metabolism of the bioflavoring by-product formation, that is, higher alcohols, esters, organic acids, and vicinal diketones, is elaborated. The next step of the fermentation process is the maturation process where the vicinal diketones, acetaldehyde, and hydrogen sulfide concen- tration needs to be reduced to acceptable levels. The chapter is concluded with a discussion about the use of immobilized cell tech- nology to intensify the fermentation process and its impact on flavor production. INTRODUCTION The production of alcoholic beverages is as old as history. Wine may have an archeological record going back more than 7500 years, with the early suspected wine residues dating from early to mid-fifth millennium bc (McGovern et al. 1996). Clear evidence of intentional winemaking first appears in the repre- sentations of wine presses that date back to the reign of Udimu in Egypt, some 5000 years ago. The direct fermentation of fruit juices, such as that of grape, had doubtlessly taken place for many thousands of years before early thinking man devel- oped beer brewing and, probably coincidentally, bread baking (Hardwick 1995). The oldest historical evidence of formal brew- ing dates back to about 6000 bc in ancient Babylonia is a piece of pottery found there, which shows workers either stirring or skimming a brewing vat. Nowadays, alcoholic beverage production represents a sig- nificant contribution to the economies of many countries. The most important beverages today are beer, wine, distilled spir- its, cider, sake, and liqueurs (Lea and Piggott 1995). In Bel- gium (“the beer paradise”), beer is the most important alcoholic beverage, although the beer consumption declined in the last 40 years: from 11,096,717 hL in 1965 to 9,703,000 hL in 2004 (NN 2005). In this time frame, wine consumption doubled from 1,059,964 to 2,215,579 hL. Another trend is the spectacular in- crease in waters and soft drinks consumption (from 5,215,056 to 26,395,000 hL). In this chapter, the biochemistry and fermentation of beer is reviewed. First, the carbohydrate metabolism in brewer’s yeast is discussed. The maltose metabolism is of major importance in beer brewing, since this sugar is in a high concentration present in wort. For the production of a high-quality beer, a well-controlled Food Biochemistry and Food Processing, Second Edition. Edited by Benjamin K. Simpson, Leo M.L. Nollet, Fidel Toldr´ a, Soottawat Benjakul, Gopinadhan Paliyath and Y.H. Hui. C 2012 John Wiley & Sons, Inc. Published 2012 by John Wiley & Sons, Inc. 627

Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

Embed Size (px)

Citation preview

Page 1: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33Biochemistry of Beer Fermentation

Ronnie Willaert

IntroductionThe Beer Brewing ProcessCarbohydrate Metabolism—Ethanol Production

Wort Carbohydrates Uptake and MetabolismMaltose and Maltotriose MetabolismGlycogen and Trehalose MetabolismWort Fermentation

Metabolism of Bioflavoring by-ProductsBiosynthesis of Higher AlcoholsBiosynthesis of EstersBiosynthesis of Organic AcidsBiosynthesis of Vicinal Diketones

Secondary FermentationVicinal DiketonesHydrogen SulfideAcetaldehydeDevelopment of Flavor Fullness

Beer Fermentation Using Immobilized Cell TechnologyCarrier MaterialsApplications of ICT in the Brewing Industry

Flavor Maturation of Green BeerProduction of Alcohol-Free or Low-Alcohol BeerProduction of Acidified Wort Using Immobilized

Lactic Acid BacteriaContinuous Main Fermentation

AcknowledgmentsReferences

Abstract: The carbohydrate metabolism and flavor formation dur-ing yeast primary and secondary fermentation (maturation) is re-viewed. Carbohydrate metabolism and ethanol production duringthe primary fermentation is discussed firstly. Next, the metabolismof the bioflavoring by-product formation, that is, higher alcohols,esters, organic acids, and vicinal diketones, is elaborated. The nextstep of the fermentation process is the maturation process wherethe vicinal diketones, acetaldehyde, and hydrogen sulfide concen-tration needs to be reduced to acceptable levels. The chapter isconcluded with a discussion about the use of immobilized cell tech-

nology to intensify the fermentation process and its impact on flavorproduction.

INTRODUCTIONThe production of alcoholic beverages is as old as history.Wine may have an archeological record going back more than7500 years, with the early suspected wine residues dating fromearly to mid-fifth millennium bc (McGovern et al. 1996). Clearevidence of intentional winemaking first appears in the repre-sentations of wine presses that date back to the reign of Udimuin Egypt, some 5000 years ago. The direct fermentation offruit juices, such as that of grape, had doubtlessly taken placefor many thousands of years before early thinking man devel-oped beer brewing and, probably coincidentally, bread baking(Hardwick 1995). The oldest historical evidence of formal brew-ing dates back to about 6000 bc in ancient Babylonia is a pieceof pottery found there, which shows workers either stirring orskimming a brewing vat.

Nowadays, alcoholic beverage production represents a sig-nificant contribution to the economies of many countries. Themost important beverages today are beer, wine, distilled spir-its, cider, sake, and liqueurs (Lea and Piggott 1995). In Bel-gium (“the beer paradise”), beer is the most important alcoholicbeverage, although the beer consumption declined in the last40 years: from 11,096,717 hL in 1965 to 9,703,000 hL in 2004(NN 2005). In this time frame, wine consumption doubled from1,059,964 to 2,215,579 hL. Another trend is the spectacular in-crease in waters and soft drinks consumption (from 5,215,056 to26,395,000 hL).

In this chapter, the biochemistry and fermentation of beer isreviewed. First, the carbohydrate metabolism in brewer’s yeastis discussed. The maltose metabolism is of major importance inbeer brewing, since this sugar is in a high concentration present inwort. For the production of a high-quality beer, a well-controlled

Food Biochemistry and Food Processing, Second Edition. Edited by Benjamin K. Simpson, Leo M.L. Nollet, Fidel Toldra, Soottawat Benjakul, Gopinadhan Paliyath and Y.H. Hui.C© 2012 John Wiley & Sons, Inc. Published 2012 by John Wiley & Sons, Inc.

627

Page 2: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

628 Part 5: Fruits, Vegetables, and Cereals

fermentation needs to be performed. During this fermentation,major flavor-active compounds are produced (and some of themare again metabolized) by the yeast cells. The metabolism ofthe most important fermentation by-products during main andsecondary fermentation is discussed in detail. The latest trendin beer fermentation technology is the process intensificationusing immobilized cell technology (ICT). This new technologyis explained and some illustrative applications—on small andlarge scale—are discussed.

THE BEER BREWING PROCESSThe principal raw materials used to brew beer are water, maltedbarley, hops and yeast. The brewing process involves extractingand breaking down the carbohydrate from the malted barley tomake a sugar solution (called “wort”), which also contains es-sential nutrients for yeast growth, and using this as a source ofnutrients for “anaerobic” yeast growth. During yeast fermenta-tion, simple sugars are consumed, releasing heat and producingethanol and other flavoring metabolic by-products. The major

biological changes, which occur in the brewing process, are cat-alyzed by naturally produced enzymes from barley (during malt-ing) and yeast. The rest of the brewing process largely involvesheat exchange, separation, and clarification, which only pro-duces minor changes in chemical composition when comparedto the enzyme catalyzed reactions. Barley is able to produce allthe enzymes that are needed to degrade starch, β-glucan, pen-tosans, lipids, and proteins, which are the major compounds ofinterest to the brewer. An overview of the brewing process isshown in Figure 33.1, where also the input and output flows areindicated. Table 33.1 gives a more detailed explanation of eachstep in the process.

CARBOHYDRATEMETABOLISM—ETHANOL PRODUCTIONWort Carbohydrates Uptake and Metabolism

Carbohydrates in wort make up 90–92% of wort solids. Wortfrom barley malt contains the fermentable sugars sucrose, fruc-tose, glucose, maltose, and maltotriose together with some

Malt

Wort cooling and aeration

Primary fermentation

Beer filtration

Maturation and conditioning

Beer stabilization

Beer packaging

Brewing water

Unmalted cereals

Hops/hopproducts

Syrups

Yeast

Spent grains

Hot trub

Spent hops

Yeast

Cold trub

Yeast

Milling

Mashing

Wort separation

Wort boiling

Wort clarification

Figure 33.1. Schematic overview of the brewing process (input flows are indicated on the left side and output flows on the right side).

Page 3: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 629

Table 33.1. Overview of the Brewing Processing Steps: From Barley to Beer

Process Action Objectives Time Temperature (◦C)

MaltingSteepingGerminationKilning

Moistening and aeration ofbarley

Barley germinationKilning of the green malt

Preparation for the germinationprocess

Enzyme production, chemicalstructure modification

Ending of germination andmodification, production offlavoring and coloringsubstances

48 h3–5 d24–48 h

12–222222–110

Milling Grain crushing withoutdisintegrating the husks

Enzyme release and increase ofsurface area

1–2 h 22

Mashing + wortseparation

Addition of warm/hot water Stimulation of enzyme action,extraction and dissolution ofcompounds, wort filtration,to obtain the desiredfermentable extract as quickas possible

1–2 h 30–72

Wort boiling Boiling of wort and hops Extraction and isomerization ofhop components, hot breakformation, wort sterilization,enzyme inactivation,formation of reducing,aromatic and coloringcompounds, removal ofundesired volatile aromacompounds, wortacidification, evaporationof water

0.5–1.5 h >98

Wort clarification Sedimentation orcentrifugation

Removal of spent hops,clarification (whirlpool,centrifuge, settling tank)

<1 h 100–80

Wort cooling andaeration

Use of heat exchanger,injection of air bubbles

Preparing the wort for yeastgrowth

<1 h 12–18

Fermentation Adding yeast, controlling thespecific gravity, removal ofyeast

Production of green beer, toobtain yeast for subsequentfermentations, carbondioxide recovery

2—7 d 12–22 (ale)4–15 (lager)

Maturation andconditioning

Beer storage in oxygen freetank, beer cooling, addingprocessing aids

Beer maturation, adjustment ofthe taste, adjustment of CO2

content, sedimentation ofyeast and cold trub, beerstabilization

7–21 h −1–0

Beer clarification Centrifugation, filtration Removal of yeast and cold trub 1–2 h −1–0Biological

stabilizationPasteurization of sterile

filtrationKilling or removing of

microorganisms1–2 h 62–72 (past.)

−1–0 (filtr.)Packaging Filling of bottles, cans, casks,

and kegs; pasteurization ofsmall volumes in packings

Production of packaged beeraccording to specifications

0.5–1.5 h −1–roomtemperature

dextrin material (Table 33.2). The fermentable sugars typicallymake up 70–80% of the total carbohydrate (MacWilliam 1968).The three major fermentable sugars are glucose, α-glucosidesmaltose, and maltotriose. Maltose is by far the most abundantof these sugars, typically accounting for 50–70% of the total

fermentable sugars in an all-malt wort. Sucrose and fructose arepresent in a low concentration. The unfermentable dextrins playlittle part in brewing. Wort fermentability may be reduced orincreased by using solid (e.g., corn grits, flakes, or rice) or liquidadjuncts (e.g., sugar syrups).

Page 4: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

630 Part 5: Fruits, Vegetables, and Cereals

Table 33.2. Carbohydrate Composition of Worts

Origin Type of WortDanish Lager

11◦PCanadian

Lager 13◦PBritish Pale

Ale 10◦PCanadian CornAdjunct Wort

All-Malt Wort18◦P

Fructose (g/L) (%)a 2.12.7

1.51.6

3.34.8

1.31.3

3.0

Glucose (g/L) (%)a 9.111.6

10.310.9

10.014.5

14.715.7

13.0

Sucrose (g/L) (%)a 2.32.9

4.24.5

5.37.7

1.81.9

Maltose (g/L) (%)a 52.466.6

60.464.2

38.956.5

62.867.0

80.0

Maltotriose (g/L) (%)a 12.816.3

17.718.8

11.416.5

13.214.1

24.0

Total ferm. sugars (g/L) 78.7 94.1 68.9 93.8 121.0Maltotetraose (g/L) 2.6 7.2 2.0Higher sugars (g/L) 21.3 26.8 25.2Total dextrins (g/L) 23.9 34.0 25.2Total sugars (g/L) 102.6 128.1 94.1 117.5

Source: Patel and Ingledew (1973), Huuskonen et al. (2010), and Hough et al. (1982).aPercent of the total fermentable sugars.

Brewing strains consume the wort sugars in a specific se-quence: glucose is consumed first, followed by fructose, maltose,and finally maltotriose. The uptake and consumption of maltoseand maltotriose is repressed or inactivated at elevated glucoseconcentrations. Only when 60% of the wort glucose has beentaken up by the yeast, the uptake and consumption of maltosewill start. Maltotriose uptake is inhibited by high glucose andmaltose concentrations. When high amounts of carbohydrateadjuncts (e.g., glucose) or high-gravity wort are employed, theglucose repression is even more pronounced, resulting in fer-mentation delays (Stewart and Russell 1993).

The efficiency of brewer’s yeast strains to effect alcoholic fer-mentation is dependent upon their ability to utilize the sugarspresent in wort. This ability very largely determines the fer-mentation rate as well as the final quality of the beer produced.In order to optimize the fermentation efficiency of the primaryfermentation, a detailed knowledge of the sugar consumptionkinetics, which is linked to the yeast growth kinetics, is required(Willaert 2001).

Maltose and Maltotriose Metabolism

The yeast Saccharomyces cerevisiae transports the monosaccha-rides across the cell membrane by the hexose transporters. Thereare 20 genes encoding hexose transporters (Dickinson 1999,Rintala et al. 2008): 18 genes encoding transporters (HXT1 toHXT17, GAL2) and two genes encoding (SNF3, RGT2). Thedisaccharide maltose and the trisaccharide maltotriose are trans-ported by specific transporters into the cytoplasm, where thesemolecules are hydrolyzed by the same α-glucosidase yieldingtwo or three molecules of glucose, respectively (Panchal andStewart 1979, Zheng et al. 1994a).

Maltose utilization in yeast is conferred by any one of fiveMAL loci: MAL1 to MAL4 and MAL6 (Bisson et al. 1993,Dickinson 1999). Each locus consists of three genes (MALx1,where x stands for one of the five loci): gene 1 encodes amaltose transporter (permease), gene 2 encodes a maltase (α-glucosidase), and gene 3 encodes a transcriptional activator ofthe other two genes. Thus, for example, the maltose transportergene at the MAL1 locus is designated MAL61. The three genesof a MAL locus are all required to allow fermentation. Alterna-tively, some authors use gene designations such as for the MAL1locus: MAL1T (transporter = permease), MAL1R (regulator),and MAL1S (maltase). The genetic and biochemical analysis ofmaltose fermentation by yeast cells revealed a series of five un-linked telomere-associated multigene MAL loci: MAL1 (chro-mosome VII), MAL2 (chromosome III), MAL4 (chromosomeII), and MAL6 (chromosome VIII). The MAL loci exhibit a veryhigh degree of homology and are telomere linked, suggestingthat they evolved by translocation from telomeric regions ofdifferent chromosomes (Michels et al. 1992). Since a fully func-tional or partial allele of the MAL1 locus is found in all strainsof S. cerevisiae, this locus is proposed as the progenitor of theother MAL loci (Chow et al. 1983), as all S. cerevisiae strains,and even its closest related yeast species S. paradoxus, containMAL1 sequences near the right telomere of chromosome VII.The genes in the MAL loci show a high degree of sequenceand functional similarity, but there can be extensive variability,and several different alleles that determine distinct phenotypes(i.e., MAL-inducible and MAL-constitutive strains) have beendescribed (Novak et al. 2004). Gene dosage studies performedwith laboratory strains of yeast have shown that the transport ofmaltose in the cell may be the rate-limiting step in the utilizationof this sugar (Goldenthal et al. 1987). Constitutive expression

Page 5: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 631

of the maltose transporter gene (MALT) with high-copy-numberplasmids in a lager yeast strain has been found to acceleratethe fermentation of maltose during high-gravity (24◦P) brewing(Kodama et al. 1995). The constitutive expression of MALS andMALR had no effect on maltose fermentability.

The control over MAL gene expression is exerted at threelevels. The presence of maltose induces, whereas glucose re-presses, the transcription of MALS and MALT genes (Federoffet al. 1983a, 1983b, Needleman et al. 1984). The constitutivelyexpressed regulatory protein (MALR) binds near the MALS andMALT promotors and mediates the induction of MALS andMALT transcription (Cohen et al. 1984, Chang et al. 1988,Ni and Needleman 1990). Experiments with MALR-disruptedstrains led to the conclusion that MalRp is involved in glucoserepression (Goldenthal and Vanoni 1990, Yao et al. 1994). Rela-tively little attention has been paid to posttranscriptional control,that is, the control of translational efficiency, or mRNA turnover,as mechanisms complementing glucose repression (Soler et al.1987). The addition of glucose to induced cells has been re-ported to cause a 70% increase in the liability of an mRNA pop-ulation containing a fragment of MALS (Federoff et al. 1983a).The third level of control is posttranslational modification. Inthe presence of glucose, maltose permease is either reversiblyconverted to a conformational variant with decreased affinity(Siro and Lovgren 1979, Peinado and Loureiro-Dias 1986) orirreversibly proteolytically degraded depending on the physio-logical conditions (Lucero et al. 1993, Riballo et al. 1995). Thelatter phenomenon is called catabolite inactivation. Glucose re-pression is accomplished by the Mig1p repressor protein, whichis encoded by the MIG1 gene (Nehlin and Ronne 1990). It hasbeen shown that Mig1p represses the transcription of all threeMAL genes by binding upstream of them (Hu et al. 1995). TheMIG1 gene has been disrupted in a haploid laboratory strainand in an industrial polyploid strain of S. cerevisiae (Klein et al.1996). In the MIG1-disrupted haploid strain, glucose repressionwas partly alleviated; that is, maltose metabolism was initiatedat higher glucose concentrations than in the corresponding wild-type strain. In contrast, the polyploid �mig1 strain exhibited aneven more stringent glucose control of maltose metabolism thanthe corresponding wild-type strain, which could be explainedby a more rigid catabolite inactivation of maltose permease,affecting the uptake of maltose.

All α-glucoside transport systems so far characterized in yeastare H+-transporters that use the electrochemical proton gradientto actively transport these sugars into the cell (Crumplen et al.1996, Stambuk and de Araujo 2001). It has been shown thatmaltose uptake is the rate-limiting step of fermentation (Kodamaet al. 1995, Wang et al. 2002, Rautio and Londesborough 2003).At least three different maltose transporters have been identi-fied in S. cerevisiae, and while the MALx1 transporters (andprobably the two MPH2 and MPH3 alleles) encode high affinity(Km = 2–4 mmol l−1) maltose permeases, the AGT1 permease (agene present in partially functional mal1g loci) transports mal-tose with lower (Km ≈ 20 mmol l−1) affinity (Han et al. 1995,Stambuk and de Araujo 2001, Day et al. 2002a, Alves et al.2007, 2008). AGT1 is found in many S. cerevisiae laboratorystrains and maps to a naturally occurring, partially functional

allele of the MAL1 locus (Han et al. 1995). Agt1p is a highlyhydrophobic, postulated integral membrane protein. It is 57%identical to Mal61p (the maltose permease encoded at MAL6)and is also a member of the 12 transmembrane domain super-family of sugar transporters (Nelissen et al. 1995). Like Mal61p,Agt1p is a high-affinity, maltose/proton symporter, but Mal61p iscapable of transporting only maltose and turanose, while Agt1ptransports these two α-glucosides as well as several others in-cluding isomaltose, α-methylglucoside, maltotriose, palatinose,trehalose and melezitose. AGT1 expression is maltose inducibleand induction is mediated by the Mal-activator.

Brewing strains of yeast are polyploid, aneuploid, or, in thecase of lager strains, alloploid. Recently, Jespersen et al. (1999)examined 30 brewing strains of yeast (5 ale strains and 25 lagerstrains) with the aim of examining the alleles of maltose andmaltotriose transporter genes contained by them. All the strainsof brewer’s yeast examined, except two, were found to containMAL11 and MAL31 sequences, and only one of these strainslacked MAL41. MAL21 was not present in the 5 ale strainsand 12 of the lager strains. MAL61 was not found in any ofthe yeast chromosomes other than those known to carry MALloci. Sequences corresponding to the AGT1 gene (transport ofmaltose and maltotriose) were detected in all but one of theyeast strains.

Although maltose is easily fermented by the majority of yeaststrains after glucose exhaustion, maltotriose is not only the leastpreferred sugar for uptake by these Saccharomyces cells, butmany yeasts may not use this α-glucosidase at all (Zheng et al.1994a, Yoon et al. 2003). Incomplete maltotriose uptake dur-ing brewing fermentations results in yeast fermentable extractin beer, material loss, greater potential for microbiological sta-bility and sometimes atypical beer flavor profiles (Stewart andRussell 1993). Maltotriose uptake from wort is always slowerwith ale strains than with lager strains under similar fermenta-tion conditions. However, the initial transport rates are similarto those of maltose in a number of ale and lager strains. El-evated osmotic pressure inhibits the transport and uptake ofglucose, maltose and maltotriose with maltose and maltotriosebeing more sensitive to osmotic pressure than glucose in bothlager and ale strains. Ethanol (5% w/v) stimulated the transportof maltose and maltotriose, due in all probability to an ethanol-induced change in the plasma membrane configuration, but hadno effect on glucose transport. Higher ethanol concentrationsinhibited the transport of all three sugars.

Maltotriose uptake shows complex kinetics indicating thepresence of high- and low-affinity transport activities, and stud-ies on sugar utilization revealed that maltose and maltotrioseare apparently transported by different permeases (Zheng et al.1994b, Zastrow et al. 2001). Two genes are recognized as per-mease genes for transporting maltotriose in yeasts: AGT1 trans-porter from S. cerevisiae and MTY1 (also called MTT1) perme-ase from S. pastorianus. They are characterized as low-affinity(Km ≈ 20 mmol l−1) maltotriose transporters (Stambuk and deAraujo 2001, Dietvorst et al. 2005, Salema-Oom et al. 2005,Alves et al. 2007, 2008).

Recent reports regarding the observed patterns of maltose andmaltotriose utilization by yeast cells show contradicting results.

Page 6: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

632 Part 5: Fruits, Vegetables, and Cereals

Data indicated that all known α-glucoside transporters present inS. cerevisiae, including the maltose permeases MAL31, MAL61,MPH2, and MHP3, allowed growth of the yeast cells on bothmaltose and maltotriose (Day et al. 2002a, 2002b). Their kineticanalysis of maltose and maltotriose uptake indicated that allthese transporters, including AGT1 permease, could transportboth sugars with practically the same affinities and capacity.Recently, to better understand maltotriose utilization by yeaststrains, an analysis of maltotriose and maltotriose utilization by52 laboratory and industrial Saccharomyces yeast strains wasperformed (Duval et al. 2009). Microarray comparative genomehybridization (aCGH) was used to correlate the observed phe-notypes with copy number variations in genes known to be in-volved in maltose and maltotriose utilization by yeasts. Theresults showed that S. pastorianus strains utilized maltotriosemore efficiently than S. cerevisiae strains and highlighted theimportance of the AGT1 gene for efficient maltotriose utiliza-tion by S. cerevisiae yeasts. The fermentation performance ofa lager (S. pastorianus) strain was improved when its AGT1gene was replaced with the AGT1 gene of an ale (S. cerevisiae)strain, since the ATG1 gene of the lager strains studied con-tained a premature stop codon and did not encode functionaltransporters (Vidgren et al. 2005, 2009). The transformantswith repaired AGT1 had higher maltose transport activity, es-pecially after growth on glucose, which represses endogeneousα-glucoside transporter genes. The sequences of two AGT1-encoded α-glucosidase transporters with different efficiencies ofmaltotriose transport in two Saccharomyces strains were com-pared (Smit et al. 2008). The amino acids Thr505 and Ser557,which are respectively located in the transmembrane (TM) seg-ment TM11 and on the intracellular segment after TM12 of theAGT1-encoded α-glucosidase transporters, are critical for effi-cient transport of maltotriose in S. cerevisiae. It was also shownthat maltotriose utilization could be improved by attaching amaltase encoded by MAL32 to the yeast cell surface (Dietvorstet al. 2007).

Glycogen and Trehalose Metabolism

Glycogen and trehalose are the main storage carbohydrates inyeast cells (Panek 1991). The synthesis of both compoundscommences with the formation of uridine diphosphate (UDP)-glucose catalyzed by UDP-glucose pyrophosphorylase.

Glycogen is a reserve carbohydrate that is metabolized duringperiods of starvation. It is a branched polysaccharide composedof linear α-(1,4)-glycosyl chains with α-(1,6)-linkages (similarto starch but with a higher degree of branching). It is synthesizedstarting from glucose via glucose-6-phosphate and glucose-1-phosphate. Glycogen is synthesized from glucose, via glucose-6-phosphate and glucose-1-phosphate (Francois and Parrou 2001).UDP serves as a carrier of glucose units and is formed by atwo-step reaction, catalyzed by phosphoglucomutase and UDP-glucose phosphorylase. Glycogen synthesis is initiated by glyco-genin that produces a short α-(1,4)-glucosyl chain, which iselongated by glycogen synthase. The α-(1,6)-glucosidic bondsare formed by the branching enzyme Glc3. The dissimilationof glycogen occurs through the action of glycogen phosphory-

lase, which releases glucose-1-phosphate from the nonreducingends of the glycogen chains, and the debranching enzyme Gdb1,which transfers a maltosyl unit to the end of an adjacent linearα-(1,4) chain and releases glucose by cleaving the remainingα-(1,6)-linkage.

When yeast cells are pitched in aerated wort, an immediateglycogen mobilization is observed (Quain 1988, Boulton 2000).Glycogen accumulates during the exponential growth phase, af-ter oxygen has been consumed. When the yeast growth is ceasedtoward the end of the primary fermentation, are maximum glyco-gen levels obtained. In the stationary phase, the glycogen levelsdecline slowly. Glycogen provides energy for the synthesis ofsterols and unsaturated fatty acids during the aerobic phase of thebeer fermentation, and energy for the cellular maintanance func-tions during the stationary phase in the storage phase betweencropping and pitching (Quain and Tubb 1982, Boulton et al.1991). The glycogen content is directly related to subsequentfermentation performance. Therefore, yeast storage occurs bestat a low temperature, without agitation and under an atmosphereof nitrogen or carbon dioxide to minimize glycogen breakdown(Murray et al. 1984).

The regulation of the glycogen content is complex and occursin part by the cAMP/PKA pathway (Smith et al. 1998). Glycogenaccumulation is repressed by a high PKA-activity (Francois andParrou 2001). The glycogen content increases and the PKA-activity is reduced, when an essential nutrient is progressivelyconsumed from the growth medium.

Trehalose is a disaccharide (α-d-glucopyranosyl-1,1- α-d-glucopyranoside) which contains two molecules of d-glucose(Boulton and Quain 2006). Trehalose biosynthesis is catalyzedby the trehalose synthase complex, which forms trehalose-6-phosphate from UDP-glucose and glucose-6-phosphate, andnext dephosphorylates it to trehalose. Trehalose is degradaded bythe neutral (Nth1) or the acid (Ath1) trehalase (Panek and Panek1990, Francois and Parrou 2001). Like glycogen, trehalose alsoaccumulates in yeast under conditions of nutrient limitation.It has been observed that trehalose rapidly accumulates in re-sponse to environmental stress, such as dehydration, heatingand osmotic stress (during high gravity brewing) (Majara et al.1996, Hounsa et al. 1998). Under stress conditions, the higherlevels of trehalose protect the cells by binding to membranes andproteins (Francois and Parrou 2001). Trehalose accumulation inresponse to stress is, however, a transient phenomenon (Parrouet al. 1997).

In the beginning of the primary fermentation, high glucoselevels activate the camp/PKA pathway (Zahringer et al. 2000).This results in posttranslational activation of neutral treha-lase and in induction of its NTH1 gene, and in repression ofthe trehalose synthase genes, which results in reduced levelsof trehalose.

Wort Fermentation

Before the fermentation process starts, wort is aerated. This isa necessary step since oxygen is required for the synthesis ofsterols and unsaturated fatty acids, which are incorporated inthe yeast cell membrane (Rogers and Stewart 1973). It has been

Page 7: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 633

shown that ergosterol and unsaturated fatty acids increase bothin concentration as long as oxygen is present in the wort (e.g.,Haukeli and Lie 1979). A maximum concentration is obtained in5–6 hours after pitching, but the formation rate is dependent uponthe pitching rate and the temperature. Unsaturated fatty acids canalso be taken up from the wort, but all malt wort does not containsufficient unsaturated lipids to support a normal growth rate ofyeast. Adding lipids to the wort, especially unsaturated fattyacids might be an interesting alternative (Moonjai et al. 2000a,2000b).

The oxygen required for lipid biosynthesis can also be in-troduced by oxygenation of the separated yeast cells. The useof the preoxygenation technique resulted in more controllableand consistent fermentations, and as a consequence, in a morebalanced beer flavor profile (Jakobsen 1982, Ohno and Taka-hashi 1986a, 1986b, Boulton et al. 1991, Devuyst et al. 1991,Masschelein et al. 1995, Depraetere et al. 2003, Depraetere2007). Recently, the impact of yeast preoxygenation on yeastmetabolism has been assessed (Verbelen et al. 2009). Therefore,expression analysis was performed of genes that are of impor-tance in oxygen-dependent pathways, oxidative stress responseand general stress response during 8 hours of preoxygenation.The gene expressions of both the important transcription fac-tors Hap1 and Rox1, involved in oxygen sensing, were mainlyincreased in the first 3 hours, while YAP1 expression, which isinvolved in the oxidative stress response, increased drasticallyonly in the first 45 minutes. The results also show that stress-responsive genes (HSP12, SSA3, PAU5, SOD1, SOD2, CTA1,and CTT1) were induced during the process, together with theaccumulation of trehalose. The accumulation of ergosterol andunsaturated fatty acids was accompanied by the expression ofERG1, ERG11, and OLE1. Genes involved in respiration (QCR9,COX15, CYC1, and CYC7) also increased during preoxygena-tion. Yeast viability did not decrease during the process, andthe fermentation performance of the yeast reached a maximumafter 5 hours of preoxygenation. These results suggest that yeastcells acquire a stress response along the preoxygenation period,which makes them more resistant against the stressful conditionsof the preoxygenation process and the subsequent fermentation.

Different devices are used to aerate the cold wort: ceramic orsintered metal candles, aeration plants employing venturi pipes,two component jets, static mixers, or centrifugal mixers (Kunze1999). The principle of these devices is that very small air (oxy-gen) bubbles are produced and quickly dissolve during turbulentmixing.

As a result of this aeration step, carbohydrates are degradedaerobically during the first few hours of the “fermentation”process. The aerobic carbohydrate catabolism takes typically12 hours for a lager fermentation.

During the first hours of the fermentation process, oxidativedegradation of carbohydrates occurs through the glycolysis andKrebs (TCA) cycle. The energy efficiency of glucose oxidationis derived from the large number of NADH2

+ produced for eachmole of glucose oxidized to CO2. The actual wort fermentationgives alcohol and carbon dioxide via the Embden-Meyerhof-Parnas (glycolytic) pathway. The reductive pathway from pyru-vate to ethanol is important since it regenerates NAD+. Energy

is obtained solely from ATP-producing steps of the Embden-Meyerhof-Parnas pathway. During fermentation, the activityof the TCA cycle is greatly reduced, although it still servesas a source of intermediates for biosynthesis (Lievense andLim 1982).

Lagunas (1979) observed that during aerobic growth of S.cerevisiae, respiration accounts for less than 10% of glucosecatabolism, the remainder being fermented. Increasing sugarconcentrations resulting in a decreased oxidative metabolism isknown as the Crabtree effect. This was traditionally explained asan inhibition of the oxidative system by high concentrations ofglucose. Nowadays, it is generally accepted that the formationof ethanol at aerobic conditions is a consequence of a bottleneckin the oxidation of pyruvate, for example, in the respiratorysystem (Petrik et al. 1983, Rieger et al. 1983, Kappeli et al.1985, Fraleigh et al. 1989, Alexander and Jeffries 1990).

A reduction of ethanol production can be achieved bymetabolic engineering of the carbon flux in yeast resulting in anincreased formation of other fermentation product. A shift of thecarbon flux towards glycerol at the expense of ethanol formationin yeast was achieved by simply increasing the level of glycerol-3-phosphate dehydrogenase (Michnick et al. 1997, Nevoigt andStahl 1997, Remize et al. 1999, Dequin 2001). The GDP1 gene,which encodes glycerol-3-phosphate dehydrogenase, has beenoverexpressed in an industrial lager brewing yeast to reducethe ethanol content in beer (Nevoigt et al. 2002). The amountof glycerol produced by the GDP1-overexpressing yeast in fer-mentation experiments—simulating brewing conditions—wasincreased 5.6 times and ethanol was decreased by 18% com-pared to the wild-type strain. Overexpression did not affect theconsumption of wort sugars and only minor changes in theconcentration of higher alcohols, esters and fatty acids couldbe observed. However, the concentrations of several other by-products, particularly acetoin, diacetyl, and acetaldehyde, wereconsiderably increased.

Mutants of S. cerevisiae strains deficient in tricarboxylic acidcycle genes have been reported as suitable strains for the produc-tion of nonalcoholic beer (Navratil et al. 2002). Strains deficientin fumarase and α-ketoglutarate dehydrogenase made nonal-coholic beers with an alcohol content lower than 0.5% (v/v)(Selecky et al. 2008). The low ethanol content was compensatedby the considerable increase of organic acids (citrate, succinate,fumarate, and malate). Some of the mutants released high levelsof lactic acid, which protects beers against contamination andmasks an unacceptable worty off-flavor.

METABOLISM OF BIOFLAVORINGBY-PRODUCTSYeast is an important contributor to flavor development in fer-mented beverages. The compounds, which are produced dur-ing fermentation, are many and varied, depending on both theraw materials and the microorganisms used. The interrelationbetween yeast metabolism and the production of bioflavoringby-products is illustrated in Figure 33.2. The important flavorcompounds produced by yeast can be classified into five cat-egories: alcohols, esters, organic acids, carbonyl compounds

Page 8: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

634 Part 5: Fruits, Vegetables, and Cereals

Sulphite

Sulphate

Sulphite

Glucose

Glucose

Maltose Maltotriose

Sucrose homocysteine

H2S H2S

Amino acids Fructose-6-PFructose

Pyruvate Keto acids

Acetaldehyde

Vicinal diketones

Amino acids

Fatty acyl CoA

Acetyl CoA Organic acids

Fatty acids Esters

Ethanol

Higheralcohols

Lipids

Figure 33.2. Interrelation between yeast metabolism and theproduction of bioflavoring by-products.

(aldehydes and vicinal diketones), and sulfur-containing com-pounds (Hammond 1993). In addition, some speciality beers cancontain important concentrations of volatile phenol compounds(Vanbeneden et al. 2007).

Biosynthesis of Higher Alcohols

During beer fermentation, higher alcohols (also called “fuselalcohols”) are produced by yeast cells as by-products and repre-sent the major fraction of the volatile compounds. More than35 higher alcohols in beer have been described. Table 33.3gives the most important compounds, which can be classified

into aliphatic (n-propanol, isobutanol, 2-methylbutanol (or ac-tive amyl alcohol), and 3-methylbutanol (or isoamyl alcohol)),and aromatic (2-phenylethanol, tyrosol, tryptophol) higher alco-hols. Aliphatic higher alcohols contribute to the “alcoholic” or“solvent” aroma of beer, and produce a warm mouthfeel. Thearomatic alcohol 2-phenylethanol has a sweet rose-like aromaand has a positive contribution to the beer aroma. It is believedthat this compound masks the dimethyl sulfide (DMS) percep-tion (Hegarty et al. 1995). The aroma of tyrosol and tryptopholare undesirable, but they are only present above their thresholdsin some top-fermented beers.

Higher alcohols are synthesized by yeast during fermenta-tion via the catabolic (Ehrlich) and anabolic pathway (aminoacid metabolism) (Ehrlich 1904, Chen 1978, Oshita et al. 1995,Hazelwood et al. 2008). In the catabolic pathway, the yeastuses the amino acids of the wort to produce the correspond-ing α-keto acid via a transamination reaction. Isoamyl alcohol,isobutanol, and phenylethanol are produced via this route fromleucine, valine, and phenylalanine, respectively. An outsider inthis pathway is propanol, which is derived from threonine viaan oxidative deamination. The excess oxoacids are subsequentlydecarboxylated into aldehydes and further reduced (alcohol de-hydrogenase) to higher alcohols. This last reduction step alsoregenerates NAD+.

Dickinson and coworkers looked at the genes and enzymes,which are used by S. cerevisiae in the catabolism of leucineto isoamyl alcohol (Dickinson et al. 1997), valine to isobu-tanol (Dickinson et al. 1998), and isoleucine to active amylalcohol (Dickinson et al. 2000). In all cases, the general se-quence of biochemical reactions is similar, but the details forthe formation of the individual alcohols are surprisingly differ-ent. The branched-chain amino acids are first deaminated to thecorresponding α-ketoacids (α-ketoisocapric acid from leucine,α-ketoisovaleric acid from valine, and α-keto-β-methylvaleric

Table 33.3. Major Higher Alcohols in Beer (Partly Adapted From NN 2000.)

CompoundFlavor Threshold

(mg/L) Aroma or Taste(b)

Concentration Range(mg/L) BottomFermentation

Concentration Range(mg/L) Top Fermentation

n-Propanol 600(c), 800(b) Alcohol 7–19 (12)(a),(f), 6–30(j) 20–45(i)

Isobutanol 100(c), 80–100(g),200(b)

Alcohol 4–20 (12) (f), 6–32(j) 10–24(i)

2-Methylbutanol 50(c), 50–60(g), 70(b) Alcohol 9–25 (15) (a), 12–16(j) 80–140(i)

3-Methylbutanol 50(c), 50–60(g), 65(b) Fusely, pungent 25–75 (46) (a), 30–60(j) 80–140(i)

2-Phenylethanol 5(a), 40(c), 45–50(g),75(d), 125(b)

Roses, sweetish 11–51 (28) (f), 4–22(g),16–42(h), 4–10(j)

35–50(g), 8–25(a), 18–45(i)

Tyrosol 10(a), 10–20(e), 20(c),100(d), (g), 200(b)

Bitter chemical 6–9(a), 6–15(a) 8–12((g), 7–22(g)

Tryptophol 10(a), 10–20(e), 200(d) Almonds, solvent 0.5–14(a) 2–12(g)

Sources: (a) Szlavko (1973), (b) Meilgaard (1975a), (c) Engan (1972), (d) Rosculet (1971), (e) Charalambous et al. (1972), (f) Values in48 European lagers (Dufour unpublished date), (g) Reed and Nogodawithana (1991), (h) Iverson (1994), (i) Derdelinckx (unpublished data),(j) immobilised cells (Willaert and Nedovic 2006).aMean value.

Page 9: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 635

acid from leucine). There are significant differences in the wayeach α-ketoacid is subsequently decarboxylated. The catabolismof phenylalanine to 2-phenylethanol and of tryptophan were alsostudied (Dickinson et al. 2003). Phenylalanine and tryptophanare first deaminated to 3-phenylpyruvate and 3-indolepyruvate,respectively, and then decarboxylated. These studies revealedthat all amino acid catabolic pathways studied to date use a sub-tle different spectrum of decarboxylases from the five-memberedfamily that comprises Pdc1p, Pdc5p, Pdc6p, Ydl080cp, andYdr380wp. Using strains containing all possible combinations ofmutations affecting the seven AAD genes (putative aryl alcoholdehydrogenases), five ADH and SFA1 (other alcohol dehydroge-nase genes), showed that the final step of amino acid catabolismcan be accomplished by any one of the ethanol dehydrogenases(Ahd1p, Ahd2p, Ahd3p, Ahd4p, Ahd5p) or Sfa1p (formalde-hyde dehydrogenase).

In the anabolic pathway, the higher alcohols are synthesizedfrom α-keto acids during the synthesis of amino acids from thecarbohydrate source. The pathway choice depends on the indi-vidual higher alcohol and on the level of available amino acidsavailable. The importance of the anabolic pathway decreasesas the number of carbon atoms in the alcohol increases (Chen1978) and increases in the later stage of fermentation as wortamino acids are depleted (MacDonald et al. 1984). Yeast strain,fermentation conditions, and wort composition all have signif-icant effects on the combination and levels of higher alcoholsthat are formed (MacDonald et al. 1984).

Conditions that promote yeast cell growth such as high lev-els of nutrients (amino acids, oxygen, lipids, zinc, . . .) and in-creased temperature and agitation stimulate the production ofhigher alcohols (Engan 1969, Engan and Aubert 1977, Landaudet al. 2001, Boswell et al. 2002). The synthesis of aromatic al-cohols is especially sensitive to temperature changes. On theother hand, conditions that restrict yeast growth, such as lowertemperature and higher pressure, reduce the extent of higheralcohol production. Higher pressures can reduce the extent ofcell growth and, therefore, the production of higher alcohols(Landaud et al. 2001). The yeast strain, fermentation conditions,and wort composition have all significant effects on the patternand concentrations of synthesized higher alcohols. Supplemen-

tation of wort with valine, isoleucine, and leucine induces theformation of isobutanol, amyl alcohol, and isoamyl alcohol, re-spectively (Ayrapaa 1971, Kodama et al. 2001). The overexpres-sion of the branched chain amino acid transferases genes BAT1and BAT2 result in an increased production of isoamyl alcoholand isobutanol (Lilly et al. 2006).

Biosynthesis of Esters

Esters are very important flavor compounds in beer. They havean effect on the fruity/flowery aromas. Table 33.4 shows themost important esters with their threshold values, which areconsiderably lower than those for higher alcohols. The majoresters can be subdivided into acetate esters and C6–C10 medium-chain fatty acid ethyl esters. They are desirable components ofbeer when present in appropriate quantities and proportions butcan become unpleasant when in excess.

Esters are produced by yeast both during the growth phase(60%) and also during the stationary phase (40%) (NN 2000).They are formed by the intracellular reaction between a fattyacyl-coenzyme A and an alcohol:

R′OH + RCO − ScoA → RCOOR′ + CoASH (1)

This reaction is catalyzed by alcohol acyltransferases(AATases; or ester synthethases) of the yeast. Since acetyl CoAis also a central molecule in the synthesis of lipids and sterols,ester synthesis is linked to the fatty acid metabolism (see alsoFig. 33.2). The majority of acetyl-CoA is formed by oxidativedecarboxylation of pyruvate, while most of the other acyl-CoAsare derived from the acylation of free CoA, catalyzed by acyl-CoA synthase.

Several enzymes are involved in the synthesis, the best char-acterized are AATase I and II that are encoded by ATF1 andATF2. Alcohol acetyltransferase (AAT) has been localized inthe plasma membrane (Malcorps and Dufour 1987) and foundto be strongly inhibited by unsaturated fatty acids, ergosterol,heavy metal ions, and sulfydryl reagents (Minetoki et al. 1993).Subcellular fractionation studies conducted during the batchfermentation cycle demonstrated the existence of both cy-tosolic and membrane-bound AAT (Ramos-Jeunehomme et al.1989, Ramos-Jeunehomme et al. 1991). In terms of controlling

Table 33.4. Major Esters in Beer (Adapted From Dufour and Malcorps 1994.)

Compound Flavor Threshold (mg/L) AromaConcentration Range (mg/L) in

48 Lagers

Ethyl actetate 20–30, 30(a) Fruity, solvent-like 8–32 (18.4)(a), 5–40(b)

Isoamyl acetate 0.6–1.2, 1.2(a) Banana, peardrop 0.3–3.8 (1.72), <0.01–2.8(b)

Ethyl caproate (ethylhexanoate)

0.17–0.21, O.21(a) Apple-like with noteof aniseed

0.05–0.3 (0.14), 0.01–0.54(b)

Ethyl caprylate(ethyl octanoate)

0.3–0.9, 0.9(a) Apple-like 0.04–0.53 (0.17), 0.01–1.2(b)

2-Phenylethylacetate

3.8(a) Roses, honey, apple,sweetish

0.10–0.73 (0.54)

Source: (a) Meilgaard (1975b) and (b) immobilised cells (Willaert and Nedovic 2006).aMean value.

Page 10: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

636 Part 5: Fruits, Vegetables, and Cereals

ester formation on a metabolic basis, it has further been shownthat ester synthesizing activity of AAT is dependent on its po-sitioning within the yeast cell. An interesting feature of thisdistribution pattern is that specific rates of acetate ester for-mation varied directly with the level of cytosolic AAT activity(Masschelein 1997).

The ATF1 gene, which encodes AAT, has been cloned from S.cerevisiae and brewery lager yeast (S. cerevisiae uvarum) (Fujiiet al. 1994). An hydrophobicity analysis suggested that AATdoes not have a membrane-spanning region that is significantlyhydrophobic, which contradicts the membrane-bound assump-tion. A Southern analysis of the yeast genomes in which theATF1 gene was used as a probe revealed that S. cerevisiae hasone ATF1 gene, while brewery lager yeast has one ATF1 geneand another homologous gene (Lg-ATF1). The AAT activitieshave been compared in vivo and in vitro under different fermen-tation conditions (Malcorps et al. 1991). This study suggestedthat ester synthesis is modulated by a repression–induction ofenzyme synthesis or processing the regulation of which is pre-sumably linked to lipid metabolism. Other enzymes are Eht1(ethanolhexanoyl transferase) and Eeb1, which are responsiblefor the formation of ethyl esters (Verstrepen et al. 2003, Saerenset al. 2006).

The ester production can be altered by changing the synthesisrate of certain fusel alcohols. Hirata et al. (1992) increased theisoamyl acetate levels by introducing extra copies of the LEU4gene in the S. cerevisiae genome. A comparable S. cerevisiaeuvarum mutant has been isolated (Lee et al. 1995). The mutantshave an altered regulation pattern of amino acid metabolism andproduce more isoamyl acetate and phenylethyl acetate.

Isoamyl acetate is synthesized from isoamyl alcohol andacetyl coenzyme A by AAT and is hydrolyzed by esterases atthe same time in S. cerevisiae. To study the effect of balancingboth enzyme activities, yeast strains with different numbers ofcopies of ATF1 gene and isoamyl acetate-hydrolyzing esterasegene (IAH1) have been constructed and used in small-scale sakebrewing (Fukuda et al. 1998). Fermentation profiles as well ascomponents of the resulting sake were largely alike. However,the amount of isoamyl acetate in the sake increased with in-creasing ratio of AAT/Iah1p esterase activity. Therefore, it wasconcluded that the balance of these two enzyme activities isimportant for isoamyl acetate accumulation in sake mash.

The synthesis of acetate esters by S. cerevisiae during fer-mentation is ascribed to at least three acetyltransferase activi-ties, namely, AAT, ethanol acetyltransferase, and isoamyl AAT(Lilly et al. 2000). To investigate the effect of increased AATactivity on the sensory quality of Chenin blanc wines and distil-lates from Colombar base wines, the ATF1 gene of S. cerevisiaewas overexpressed. Northern blot analysis indicated constitu-tive expression of ATF1 at high levels in these transformants.The levels of ethyl acetate, isoamyl acetate and 2-phenylethylacetate increased 3- to 10-fold, 3.8- to 12-fold, and 2- to 10-fold, respectively, depending on the fermentation temperature,cultivar, and yeast used. The concentrations of ethyl caprate,ethyl caprylate, and hexyl acetate only showed minor changes,whereas the acetic acid concentration decreased by more thanhalf. This study established the concept that the overexpression

of acetyltransferase genes such as ATF1 could profoundly affectthe flavor profiles of wines and distillates deficient in aroma.

In order to investigate and compare the roles of the known S.cerevisiae AATs, Atf1p, Atf2p, and Lg-Atf1p, in volatile esterproduction, the respective genes were either deleted or overex-pressed in a laboratory strain and a commercial brewing strain(Verstrepen et al. 2003). Analysis of the fermentation productsconfirmed that the expression levels of ATF1 and ATF2 greatlyaffect the production of ethyl acetate and isoamyl acetate. GC-MS analysis revealed that Atf1p and Atf2p are also responsiblefor the formation of a broad range of less volatile esters, suchas propyl acetate, isobutyl acetate, pentyl acetate, hexyl acetate,heptyl acetate, octyl acetate, and phenyl ethyl acetate. With re-spect to the esters analyzed in this study, Atf2p seemed to playonly a minor role compared to Atf1p. The atf1�atf2� doubledeletion strain did not form any isoamyl acetate, showing thattogether, Atf1p and Atf2p are responsible for the total cellu-lar isoamyl AAT activity. However, the double deletion strainstill produced considerable amounts of certain other esters, suchas ethyl acetate (50% of the wild-type strain), propyl acetate(50%), and isobutyl acetate (40%), which provides evidence forthe existence of additional, as-yet-unknown ester synthases inthe yeast proteome. Interestingly, overexpression of different al-leles of ATF1 and ATF2 led to different ester production rates,indicating that differences in the aroma profiles of yeast strainsmay be partially due to mutations in their ATF genes.

Recently, it has been discovered that the Atf1 enzyme is lo-calized inside lipid vesicles in the cytoplasm of the yeast cell(Verstrepen 2003). Lipid vesicles are small organelles in whichcertain neutral lipids are metabolized or stored. This indicatesthat fruity esters are possibly by-products of these processes.

Ester formation is highly dependent on the yeast strain used(Nykanen and Nykanen 1977, Peddie 1990, Verstrepen et al.2003b) and on certain fermentation parameters such as tem-perature (Engan and Aubert 1977, Gee and Ramirez 1994,Sablayrolles and Ball 1995), specific growth rate (Gee andRamirez 1994), pitching rate (Maule 1967, D’Amore et al. 1991,Gee and Ramirez 1994), and top pressure (NN 2000, Verstrepenet al. 2003b). Additionally, the concentrations of assimilable ni-trogen compounds (Hammond 1993, Calderbank and Hammond1994, Sablayrolles and Ball 1995), carbon sources (Pfisterer andStewart 1975, White and Portno 1979, Younis and Stewart 1998,2000), dissolved oxygen (Anderson and Kirsop 1975a, 1975b,Avhenainen and Makinen 1989, Sablayrolles and Ball 1995),and fatty acids (Thurston et al. 1981, 1982) can influence theester production rate.

Acetate ester formation in brewer’s yeast is controlled mainlyby the expression level of the AATase-encoding genes (Ver-strepen et al. 2003b). Additionally, changes in the availabilityof the two substrates for ester production, higher alcohols andacyl-CoA, also influences ester synthesis rates. Any factor thatinfluences the expression of the ester synthase genes and/orthe concentrations of substrates will affect ester production ac-cordingly. Perhaps the most convenient and selective way toreduce ester production is applying tank overpressure, if neces-sary in combination with (slightly) lower fermentation temper-atures, low wort free amino nitrogen (FAN) and glucose levels

Page 11: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 637

and elevated wort aeration or wort lipid concentration (Ver-strepen et al. 2003b). Enhancing ester production is slightly morecomplicated.

If it is possible, overpressure or wort aeration can be reduced.Otherwise, worts rich in glucose and nitrogen combined withhigher fermentation temperatures and lower pitching rates orapplication of the drauflassen technique may prove helpful.

Biosynthesis of Organic Acids

Over hundred different organic acids have been reported inbeer (Meilgaard 1975b). Important organic acids detected inbeer include pyruvate, acetate, lactate, succinate, pyroglutamate,malate, citrate, α-ketoglutarate, and α-hydroxyglutarate; and themedium-chain length fatty acids, caproic (C6), caprylic (C8),and capric (C10) acid (Coote and Kirsop 1974, Meilgaard 1975b,Klopper et al. 1986). They influence flavor directly when presentabove their taste threshold and by their influence on beer pH.These components have their origin in raw materials (malt, hops)and are produced during the beer fermentation. Organic acids,which are excreted by yeast cells, are synthesized via aminoacid biosynthesis pathways and carbohydrate metabolism. Es-pecially, they are overflow products of the incomplete Krebscycle during beer fermentation. Excretion of organic acids isinfluenced by yeast strain and fermentation vigor. Sluggish fer-mentations lead to lower levels of excretion. Pyruvate excretionfollows the yeast growth: maximal concentration is reached justbefore the maximal yeast growth and is next taken up by theyeast and converted to acetate. Acetate is synthesized quicklyduring early fermentation and is later partially reused by theyeast during yeast growth. At the end of the fermentation, ac-etate is accumulated. The reduction of pyruvate results in theproduction of d-lactate of l-lactate (most yeast strains producepreferentially d-lactate). The highest amount of lactate is pro-duced during the most active fermentation period.

The change in organic acid productivity by disruption of thegene encoding fumarase (FUM1) has been investigated and ithas been suggested that malate and succinate are produced viathe oxidative pathway of the TCA cycle under static and sakebrewing conditions (Magarifuchi et al. 1995). Using a NAD+-dependent isocitrate dehydrogenase gene (IDH1, IDH2) disrup-tant, approximately half of the succinate in sake mash was foundto be synthesized via the oxidative pathway of the TCA cycle insake yeast (Asano et al. 1999).

Sake yeast strains possessing various organic acid productivi-ties were isolated by gene disruption (Arikawa et al. 1999). Sakefermented using the aconitase gene (ACO1) disruptant containeda twofold higher concentration of malate and a twofold lowerconcentration of succinate than that made using the wild-typestrain. The fumarate reductase gene (OSM1) disruptant producedsake containing a 1.5-fold higher concentration of succinate,whereas the α-ketoglutarate dehydrogenase gene (KGD1) andfumarase gene (FUM1) disruptants gave lower succinate con-centrations. In S. cerevisiae, there are two isoenzymes of fu-marate reductase (FRDS1 and FDRS2), encoded by the FRDSand OSM1 genes, respectively (Arikawa et al. 1998). Recentresults suggest that these isoenzymes are required for the reox-

idation of intracellular NADH under anaerobic conditions, butnot under aerobic conditions (Enomoto et al. 2002).

Succinate dehydrogenase is an enzyme of the TCA cycleand thus essential for respiration. In S. cerevisiae, this enzymeis composed of four nonidentical subunits, that is, the flavo-protein, the iron–sulfur protein, the cytochrome b560, and theubiquinone reduction protein encoded by the SDH1, SDH2,SDH3, and SDH4 genes, respectively (Lombardo et al. 1990,Chapman et al. 1992, Bullis and Lamire 1994, Daignan-Fournieret al. 1994). Sdh1p and Sdh2p comprise the catalytic domain in-volved in succinate oxidation. These proteins are anchored tothe inner mitochondrial membrane by Sdh3p and Sdh4p, whichare necessary for electron transfer and ubiquinone reduction,and constitute the succinate:ubiquinone oxidoreductase (com-plex II) of the electron transport chain. Single or double disrup-tants of the SDH1, SDH1b (which is a homologue of the SDH1gene), SDH2, SDH3, and SDH4 genes have been constructed andshown that the succinate dehydrogenase activity was retained inthe SDH2 disruptant and that double disruption of SDH1 andSDH2 or SDH1b genes is necessary to cause deficiency of suc-cinate dehydrogenase activity in sake yeast (Kubo et al. 2000).The role of each subunit in succinate dehydrogenase activity andthe effect of succinate dehydrogenase on succinate productionusing strains that were deficient in succinate dehydrogenase,have also been determined. The results suggested that succinatedehydrogenase activity contributes to succinate production un-der shaking conditions, but not under static and sake brewingconditions.

The medium-chain fatty acids account for 85–90% of thefatty acids in beer and impart an undesirable goaty, sweaty, andyeasty flavor (Chen 1980). These fatty acids are produced denovo by yeast during anaerobic fermentation and are not theresult of β-oxydation of wort or yeast long-chain fatty acids. Asa result, any change in fermentation conditions that promote theextent of yeast growth also favor increased levels of medium-chain fatty acids in beer. Higher temperature, increased wortoxygenation, and possibly elevated pitching rates are all effectivein this respect (Boulton and Quain 2006). The presence of thesemedium-chain fatty acids in beer is also related to yeast autolysis(Masschelein 1981). Yeast autolytic off-flavors are stimulatedat high temperatures, high yeast concentrations, and prolongedcontact times at the end of primary fermentation and duringsecondary fermentation.

Biosynthesis of Vicinal Diketones

Vicinal diketones are ketones with two adjacent carbonyl groups.During fermentations, these flavor-active compounds are pro-duced as by-products of the synthesis pathway of isoleucine,leucine, and valine (ILV pathway) (see Fig. 33.3) and thusalso linked to amino acid metabolism (Nakatani et al. 1984)and the synthesis of higher alcohols. They impart a “buttery”,“butterscotch” aroma to alcoholic drinks. Two of these com-pounds are important in beer, that is, diacetyl (2,3-butanedione)and 2,3-pentanedione. Diacetyl is quantitatively more impor-tant than 2,3-pentanedione. It has a taste threshold in lagerbeer from 17 µg/L (Saison et al. 2009) to 150 µg/L (Meilgaard

Page 12: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

638 Part 5: Fruits, Vegetables, and Cereals

Threonine

Threonine

α-Ketobutyrate

α-Acetohydroxybutyrate

Isoleucine

Pyruvate

α-Acetolactate α-Acetolactate

LeucineValine

α-Acetohydroxy- butyrate

Diacetyl

Acetoin

2,3-Butanediol

2,3-Pentane-dione

2,3-Pentanediol

Acetylethyl- carbinol

CO2CO2

Carbohydratemetabolism

Bacterialacetolactatedecarboxylase

CO2ILV5

Figure 33.3. The synthesis and reduction of vicinal diketones in Saccharomyces cerevisiae.

1975b), approximately 10 times lower than that of pentanedione(Wainwright 1973).

The excreted α-acetohydroxy acids are overflow products ofthe ILV pathway that are nonenzymatically degraded to thecorresponding vicinal diketones (Inoue et al. 1968). Tetraploidgene dosage series for various ILV genes have been constructedand the obtained yeast strains were used to study the influenceof the copy number of ILV genes on the production of vicinaldiketones (Debourg et al. 1990, Debourg 2002). It was shownthat the ILV5 activity is the rate-limiting step in the ILV path-way and responsible for the overflow (Fig. 33.3). The nonen-zymatic oxidative decarboxylation step is the rate-limiting stepand proceeds faster at a higher temperature and a lower pH (In-oue and Yamamoto 1970, Haukeli and Lie 1978). The producedamount of α-acetolactate is very dependent on the used yeaststrain. The production increases with increasing yeast growth.For a classical fermentation, 0.6 ppm α-acetolactate is formed(Delvaux 1998). At high aeration, this value can be increasedto 0.9 ppm and in cylindro-conical fermentations tanks evento 1.2–1.5 ppm.

It has been shown that valine inhibits the synthesis ofα-acetolactate through feedback inhibition (Magee and deRobinson-Szulmajster 1968). This inhibition is directed to theprotein Ilv6p, which is the regulatory subunit of acetohydroxyacid synthase (the catalytic subunit encoded by ILV2)(Pang andDuggleby 1999, 2001). Because the uptake of valine is delayedin a normal beer fermentation, the suppressive effect of va-line accounts for the postponed onset of total diacetyl (sumof actual diacetyl and α-acetolactate), and this effect persistslonger in worts with high levels of FAN content. In contrast,low FAN levels give two diacetyl peaks as a result of the re-quirement for valine biosynthesis. Therefore, a minimum FANlevel above the critical value of 50 ppm (Nakatani et al. 1984)

or 140 ppm (Pugh et al. 1997) should be maintained dur-ing the fermentation to ensure the presence of valine in thefermenting wort.

Yeast cells posses the necessary enzymes (reductases) to re-duce diacetyl to acetoin and further to 2,3-butanediol, and 2,3-pentanedione to 2,3-pentanediol (Bamforth and Kanauchi 2004).These reduced compounds have much higher taste thresholdsand have no impact on the beer flavor (Van Den Berg et al.1983). The reduction reactions are yeast strain dependent. Thereduction occurs at the end of the main fermentation and duringthe maturation. Sufficient yeast cells in suspension are necessaryto obtain an efficient reduction. Yeast strains that flocculate earlyduring the main fermentation needs a long maturation time toreduce the vicinal diketones. Diacetyl can be complexed usingSO2. These complexes cannot be reduced, but diacetyl can againbe liberated at a later stage by aldehydes. This situation is es-pecially applicable to yeast strains, which produce a lot of SO2.Worts, which are produced using a high content of adjuncts, canbe low in free amino acid content. These worts can give rise toa high diacetyl peak at the end of the fermentation.

There are several strategies, which can be chosen to reducethe vicinal diketones amount during fermentation:

1. Since the temperature has a positive effect on the reductionefficiency of the α-acetohydroxy acids, a warm rest periodat the end of the main fermentation and a warm maturationare applied in many breweries. In this case, temperatureshould be well controlled to avoid yeast autolysis.

2. Since the rapid removal of vicinal diketones requires yeastcells in an active metabolic condition, the addition of5–10% Krausen (containing active, growing yeast) is aprocedure, which gives enhanced transformation of vic-inal diketones (NN 2000). This procedure can lead to

Page 13: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 639

overproduction of hydrogen sulfide, depending upon theproportions of threonine and methionine carried forwardfrom primary fermentation.

3. Heating up the green beer to a high temperature (90◦C)and hold it there for a short period (ca. 7–10 min) to de-carboxylate all excreted α-acetohydroxy acids. To avoidcell autolysis, yeast cells are removed by centrifugationprior to heating up. The vicinal diketones can be further re-duced by immobilized yeast cells in a few hours (typicallyat 4◦C) (see further).

4. Adding the enzyme α-acetolactate decarboxylase (Godt-fredsen et al. 1984, Rostgaard-Jensen et al. 1987): Thisenzyme decarboxylates α-acetolactate directly into ace-toin (see Fig. 33.3). It is not present in S. cerevisiae, buthas been isolated from various bacteria such as Enter-obacter aerogenes, Aerobacter aerogenes, Streptococcoslactis, Lactobacillus casei, Acetobacter aceti, and Aceto-bacter pasteurianus. It has been shown that the additionof α-acetolactate decarboxylase from L. casei can reducethe maturation time to 22 hours (Godtfredsen et al. 1983,1984). An example of a commercial product is Maturex Lfrom Novo Nordisk (Denmark) (Jensen 1993). Maturex Lis a purified α-acetolactate decarboxylase produced by agenetically modified strain of Bacillus subtilis, which hasreceived the gene from Bacillus brevis. The recommendeddosage is 1 a 2 kg per 1000 hL wort, to be added to thecold wort at the beginning of fermentation.

5. Using genetic modified yeast strains:a. Introducing the bacterial α-acetolactate decarboxylase

gene into yeast chromosomes (Fujii et al. 1990, Suihkoet al. 1990, Blomqvist et al. 1991, Enari et al. 1992,Linko et al. 1993, Yamano et al. 1994, Tada et al. 1995,Onnela et al. 1996). Transformants possessed a veryhigh α-acetolactate decarboxylase activity that reducedthe diacetyl concentration considerably during beerfermentations.

b. Modifying the biosynthetic flux through the ILV path-way by partially deactivation of ILV2. Spontaneousmutants resistant to the herbicide sulfometuron methylhave been selected. These strains showed a partial in-activation of the α-acetolactate synthase activity andsome mutants produced 50% less diacetyl compared tothe parental strain (Gjermansen et al. 1988).

c. Increasing the flux of α-acetolactate acid isomerore-cuctase activity encoded by the ILV5 gene (Dillemanset al. 1987). Since α-acetolactate acid isomerore-cuctase activity is responsible for the rate-limitingstep, increasing its activity reduces the overflow of α-acetolactate. A multicopy transformant resulted in a70% decreased production of vicinal diketones (Villa-neuba et al. 1990), whereas an integrative transformantgave a 50% reduction (Goossens et al. 1993). A tan-dem integration of multiple ILV5 copies resulted alsoin elevated transciption in a polyploidy industrial yeaststrain (Mithieux and Weiss 1995). Vicinal diketonesproduction could be reduced by targeting the mito-chondrial Ilv5p to the cytosol (Omura 2008).

SECONDARY FERMENTATIONDuring the secondary fermentation or maturation of beer, severalobjectives should be realized:

� Sedimentation of yeast cells� Improvement of the colloidal stability by sedimentation of

the tannin–protein complexes� Beer saturation with carbon dioxide� Removal of unwanted aroma compounds� Excretion of flavor-active compounds from yeast to give

body and depth to the beer� Fermentation of the remaining extract� Improvement of the foam stability of the beer� Adjustment of the beer color (if necessary) by adding color-

ing substances (e.g., caramel)� Adjustment of the bitterness of beer (if necessary) by

adding hop products

In the presence of yeast, the principal changes that occur arethe elimination of undesirable flavor compounds, such as vicinaldiketones, hydrogen sulfide, and acetaldehyde, and the excretionof compounds enhancing the flavor fullness (body) of beer.

Vicinal Diketones

In traditional fermentation lagering processes, the eliminationof vicinal diketones required several weeks and determined thelength of the maturation process. Nowadays, the maturationphase is much shorter since strategies are used to acceleratethe vicinal diketones removal (see the preceding text). Diacetylis used as a marker molecule. The objective during lagering isto reduce the diacetyl concentration below its taste threshold(<0.10 mg/mL).

Hydrogen Sulfide

Sulfite and hydrogen sulfide are intermediates in the biosynthe-sis of the sulfur-containing amino acids methionine and cysteine(Van Haecht and Dufour 1995, Duan et al. 2004). Hydrogen sul-fide plays an important role during maturation. Inorganic sulfateis taken up by the yeast cells via a permease. Subsequently,it is reduced to sulfide via the intermediates adenylyl sulfate,phosphoadenylyl sulfate and sulfite (see Fig. 33.4). H2S andSO2, which are not incorporated in S-containing amino acids,are excreted by the yeast cell during the growth phase (Ryderand Masschelin 1983, Thomas and Surdin-Kerjan 1997). Theexcreted amount depends on the used yeast strain, the sulfatecontent of the wort and the growth conditions (Romano andSuzzi 1992). Methionine causes decreased production of sulfurcompounds by feedback inhibition, while threonine increasesthe production (Thomas and Surdin-Kerjan 1997). H2S and SO2

can also be formed from the catabolism of S-containing aminoacids (Dual et al. 2004). The production of H2S and SO2 de-pends on the yeast strain, sulfate content in the wort, and growthconditions. The production of H2S could be reduced by the ex-pression of cystathione synthase genes from S. cerevisiae in abrewing yeast strain (Tezuka et al. 1992).

Page 14: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

640 Part 5: Fruits, Vegetables, and Cereals

Hydrogen Sulphide

Sulphate Sulphate

Adenylyl sulphate

Phospoadenylyl sulphate

Sulphite Sulphite

Hydrogen sulphide

Homocysteine

Methionine

S-adenosylmethionine

Cystathionine

Cysteine

o-Acetylhomoserine

1

Gluthatione

S-adenosylhomocysteine

CH3-THF

CH2-THF

THF

Serine

Glycine CH3-X

X

2

3

4

5

6

7

8

910

11

12

13

14

Figure 33.4. The remethylation, transulfuration, and sulfur assimilation pathways. Genes and enzymes catalyzing individual reactions are asfollows: 1, sulfate permease; 2, ATP sulfurylase; 3, MET14: adenylylsulfate kinase (EC 2.7.1.25); 5, MET10: sulfite reductase (EC 1.8.1.2); 6,sulfite permease; 7, MET17: O-acetylhomoserine (thiol)-lyase (EC 2.5.1.49); 8, CYS4: cystathionine β-synthase (CBS; EC 4.2.1.22); 9,CYS3: cystathionine γ-lyase (EC 4.4.1.1), 10, MET6: methionine synthase (EC 2.1.1.14); 11, SAM1 and SAM2: S-adenosylmethioninesynthetase (EC 2.5.1.6); 12, SAH1: S-adenosylhomocysteine hydrolase (EC 3.3.1.1); 13, SHM1 and SHM2: serine hydroxymethyltransferase(SHMT; EC 2.1.2.1); 14, MET12 and MET13: methylenetetrahydrofolate reductase (MTHFR; EC 1.5.1.20). “X” represents any methyl groupacceptor; THF, tetrahydrofolate; CH2-THF, 5,10-methylenetetrahydrofolate; CH3-THF, 5-methyltetrahydrofolate. (Partly adapted from Chanand Appling 2003.)

At the end of the primary fermentation and during the matu-ration, the excess H2S is reutilized by the yeast. A warm con-ditioning period at 10 a 12◦C may be used to remove excessivelevels of H2S.

Brewing yeasts produce H2S when they are deficient in thevitamin pantothenate (Walker 1998). This vitamin is a precursorof coenzyme A, which is required for metabolism of sulfate intomethionine. Therefore, panthothenate deficiency may result inan imbalance in sulfur amino acid biosynthesis, leading to excesssulfate uptake and excretion of H2S (Slaughter and Jordan 1986).

Sulfite is a versatile food additive used to preserve a largerange of beverages and foodstuffs. In beer, sulfite has a dualpurpose, acting both as an antioxidant and an agent for maskingof certain off-flavors. Some of the flavor stabilizing propertiesof sulfite is suggested to be due to complex formation of bisul-fate with varying carbonyl compounds, of which some wouldgive rise to off-flavors in bottled beer (Dufour 1991). Especially,

the unwanted carbonyl trans-2-nonenal has received particularattention, since it is responsible for the “cardboard” flavor ofsome types of stale beer. It has been suggested that it would bebetter to use a yeast strain with reduced sulfite excretion duringfermentation and to add sulfite at the point of bottling to ensuregood flavor stability (Francke Johannesen et al. 1999). There-fore, a brewer’s yeast disabled in the production of sulfite hasbeen constructed by inactivating both copies of the two allelesof the MET14 gene (which encodes for adenylylsulfate kinase).Fermentation experiments showed that there was no qualitativedifference between yeast-derived and artificially added sulfite,with respect to trans-2-nonenal content and flavor stability ofthe final beer.

The elimination of the gene encoding sulfite reductase(MET10) in brewing strains of Saccharomyces results in in-creased accumulation of SO2 in beer (Hansen and Kielbrandt1996a). The inactivation of MET2 resulted into elevated sulfite

Page 15: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 641

concentrations in beer (Hansen and Kielbrandt 1996b). Beersproduced with increased levels of sulfite showed an improvedflavor stability.

Acetaldehyde

Aldehydes—in particular, acetaldehyde (green apple-likeflavor)—have an impact on the flavor of green beer. The accu-mulation of acetaldehyde is dependent on the kinetic propertiesof enzymes responsible for its formation (pyruvate decarboxy-lase and acetaldehyde dehydrogenase) and dissimilation (alco-hol dehydrogenase) (Boulton and Quain 2006). Acetaldehydesynthesis is linked to yeast growth (Geiger and Piendl 1976).Its concentration is maximal at the end of the growth phase andis reduced at the end of the primary fermentation and duringmaturation by the yeast cells. As with diacetyl, levels may beenhanced if yeast metabolism is stimulated during transfer, es-pecially by oxygen ingress. Removal also requires the presenceof enough active yeast. Fermentations with early flocculatingyeast cells can result in too high acetaldehyde concentrations atthe end.

Development of Flavor Fullness

During maturation, the residual yeast will excrete compounds(i.e., amino acids, phosphates, peptides, nucleic acids, . . .) intothe beer. The amount and “quality” of these excreted materialsdepend on the yeast concentration, yeast strain, its metabolicstate and the temperature (NN 2000). Rapid excretion of materialis best achieved at a temperature of 5–7◦C during 10 days (Vande Meersche et al. 1977).

When the conditioning period is too long or when the temper-ature is too high, yeast cell autolysis will occur. Some enzymesare liberated (e.g., α-glucosidase), which will produce glucosefrom traces of residual maltose (NN 2000). At the bottom ofa fermentation tank, the amount of α-amino-nitrogen can riseto 40–10,000 mg/L, which account for an increase of 30 mg/Lfor the total beer volume. The increase in amino acid concen-tration in the beer has a positive effect on the flavorfullness ofthe beer. Undesirable medium-chain fatty acids can also be pro-duced in significant amounts if the maturation temperature istoo high (Masschelein 1981). Measurement of these compoundsindicates the level of autolysis and permits the determination ofthe most appropriate conditioning period and temperature.

BEER FERMENTATION USINGIMMOBILIZED CELL TECHNOLOGYThe advantages of continuous fermentation—such as greaterefficiency in utilization of carbohydrates and better use ofequipment—led also to the development of continuous beerfermentation processes. Since the beginning of the twentiethcentury, many different systems using suspended yeast cellshave been developed. The excitement for continuous beer fer-mentation led—especially during the 1950 and 1960s—to thedevelopment of various interesting systems. These systems canbe classified as (i) stirred versus unstirred tank reactors, (ii)

single-vessel systems versus a number of vessels connected inseries, (iii) vessels that allow yeast to overflow freely with thebeer (“open system”) versus vessels that have abnormally highyeast concentrations (“closed” or “semiclosed system”) (Well-hoener 1954, Coutts 1957, Bishop 1970, Hough et al. 1982,).However, these continuous beer fermentation processes werenot commercially successful due to many practical problems,such as the increased danger of contamination (not only duringfermentation but also during storage of wort in supplementaryholdings tanks that are required since the upstream and down-stream brewing processes are usually not continuous), changesin beer flavor (Thorne 1968) and a poor understanding of thebeer fermentation kinetics under continuous conditions. One ofthe well-known exceptions is the successful implementation of acontinuous beer production process in New Zealand by MortonCoutts (Dominion Breweries), which is still in use today (Coutts1957, Hough et al. 1982).

In the 1970, there was a revival in developing continuousbeer fermentation systems due to the progress in research onimmobilization bioprocesses using living cells. Immobilizationgives fermentation processes with high cell densities, resultingin a drastic increase in fermentation productivities compared tothe traditional time-consuming batch fermentation processes.

The last 30 years, ICT has been extensively examined andsome designs have reached already commercial exploitation.Immobilized cell systems are heterogeneous systems in whichconsiderable mass transfer limitations can occur, resulting in achanged cell yeast metabolism. Therefore, successful exploita-tion of ICT needs a thorough understanding of mass transfer andintrinsic yeast kinetic behavior of these systems.

Carrier Materials

Cell immobilization can be classified into four categories basedon the mechanism of cell localization and the nature of sup-port material: (i) attachment to the support surface, which canbe spontaneous or induced by linking agents; (ii) entrapmentwithin a porous matrix; (iii) containment behind or within a bar-rier; and (iv) self-aggregation, naturally or artificially induced(Karel et al. 1985, Willaert and Baron 1996). Various cell im-mobilization carrier materials have been tested and used forbeer production/bioflavoring. Selection criteria are summarizedin Table 33.5. Depending on the particular application, reac-tor type and operational conditions, some selection criteria willbe more appropriate. Examples of selected carrier materials forparticular applications are tabulated in Table 33.6.

Cell immobilization by self-aggregation is based on the for-mation of cell clumps or flocs. Actually, flocs are formed at theend of the primary fermentation when flocculent strains are used.Flocculent strains can also be used in continuous fermentationsystems (see “Coutts continuous beer fermentation system” de-scribed in the preceding text). Very high yeast concentrationsmay be achieved by the use of inclined tubes, still zones aroundoutlet pipes and by holding the yeast in a filter (Hough et al.1982). Growth of the sedimentary yeast may be controlled bythe amount of air injected, while carbon dioxide is used to causesome mixing. A flocculent strain has also been used in the tower

Page 16: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

642 Part 5: Fruits, Vegetables, and Cereals

Table 33.5. Selection Criteria for Yeast CellImmobilization Carrier Materials

� High cell mass loading capacity� Easy access to nutrient media� Simple and gentle immobilization procedure� Immobilization compounds approved for food applications� High surface area-to-volume ratio� Optimum mass transfer distance from flowing media to

centre of support� Mechanical stability (compression, abrasion)� Chemical stability� Highly flexible: rapid start-up after shut-down� Sterilizable and reusable� Suitable for conventional reactor systems� Low shear experienced by cells� Easy separation of cells and carrier from media� Readily up-scalable� Economically feasible (low capital and operating costs)� Desired flavor profile and consistent product� Complete attenuation� Controlled oxygenation� Control of contamination� Controlled yeast growth� Wide choice of yeast

Source: Nedovic et al. (2005) and Verbelen et al. (2010).

fermenter system, which comprises a vertical tube into the baseof which wort is pumped (Royston 1966). The sedimentary yeastforms a solid plug at the base of the vessel and through it thewort permeates. Fermentation proceeds as the wort rises, withthe rate of wort injection being so adjusted that at the top of thetower the wort is completely fermented.

Yeast flocculation is a reversible, asexual, and calcium-dependent process in which cells adhere to form flocs consistingof thousands of cells (Stratford 1989, Bony et al. 1997, Jinand Speers 1999). Many fungi contain a family of cell wall“adhesines” (which are glycoproteins) that confer unique ad-hesion properties (Teunissen and Steensma 1995, Guo et al.2000, Hoyer 2001, Sheppard et al. 2004). These molecules arerequired for the interactions of fungal cells with each other (floc-culation and filamentation) (Teunissen and Steensma 1995, Loand Dranginis 1998, Guo et al. 2000, Viyas et al. 2003), withinert surfaces such as agar and plastic (Gaur and Klotz 1997, Loand Dranginis 1998, Reynolds and Fink 2001, Li and Palecek2003) and with mammalian tissues/cells (Cormack et al. 1999,Staab et al. 1999, Fu et al. 2002, Li and Palecek 2003). Theyare also crucial for the formation of fungal biofilms (Baillie andDouglas 1999, Reynolds and Fink 2001, Green et al. 2004). Theadhesin proteins in S. cerevisiae are encoded by FLO genes,including FLO1, FLO5, FLO9, FLO10, and FLO11 (Verstrepenet al. 2004). These proteins are called flocculins (Caro et al.1997) because these proteins promote cell–cell adhesion to formmulticellular clumps that sediment out of solution.

The flocculation phenomenon is genetically controlled by33 genes (Teunissen and Steensma 1995). Recently, the fivemembers of the FLO-adhesine family were studied by selectively

overexpressing each FLO gene in the laboratory strain S288C(Van Mulders et al. 2009). As all FLO genes are transcriptionallysilent in the S288C-strain background, each FLO gene can beactivated one by one and each resultant phenotype can be investi-gated. The FLO1, FLO5, FLO9, and FLO10 genes share consid-erable sequence homology. The member proteins of the adhesinfamily have a modular configuration that consists of three do-mains (A, B, and C) and an N-terminal secretory sequence thatmust be removed as the protein moves through the secretorypathway to the plasma membrane (Hoyer et al. 1998). The N-terminal domain (A) is involved in sugar recognition (Kobayashiet al. 1998). The adhesins undergo several posttranslational mod-ifications, that is, N- and O-glycosylations. They move from theendoplasmic reticulum (ER) through the Golgi and pass throughthe plasma membrane and find their final destination in the cellwall, where they are anchored by a glycosyl phosphatidylinositol(GPI) (Teunissen et al. 1993a, 1993b, Bidard et al. 1994, Bonyet al. 1997, Hoyer et al. 1998). The GPI anchor is added tothe C-terminus in the ER, and mannose residues are added tothe many serine and threonine residues in domain B in the Golgi(Udenfriend and Kodukula 1995, Bony et al. 1997, Frieman et al.2002, De Groot et al. 2003). The FLO1 gene product (Flo1p)has been localized at the cell surface by immunofluorescent mi-croscopy (Bidard et al. 1995). The amount of Flo proteins inflocculent strains increased during batch yeast growth and theFlo1p availability at the cell surface determined the flocculationdegree of the yeast. Flo proteins are polarly incorporated into thecell wall at the bud tip and the mother–daughter neck junction(Bony et al. 1997). The transcriptional activity of the flocculationgenes is influenced by the nutritional status of the yeast cells aswell as other stress factors (Verstrepen et al. 2003a). This impliesthat during beer fermentation, flocculation is affected by numer-ous parameters such as nutrient conditions, dissolved oxygen,pH, fermentation temperature, and yeast handling and storageconditions.

Applications of ICT in the Brewing Industry

Beer production with immobilized yeast has been the subject ofresearch for approximately 40 years, but has so far found lim-ited application in the brewing industry, because of engineeringproblems, unrealised cost advantages, microbial contaminations,and an unbalanced beer flavor (Linko et al. 1998, Branyik et al.2005, Nedovic et al. 2005, Willaert and Nedovic 2006, Verbelenet al. 2010). The ultimate aim of this research is the productionof beer, of desired quality, within 1–3 days. Traditional beer fer-mentation systems use freely suspended yeast cells to fermentwort in an unstirred batch reactor. The primary fermentationtakes approximately 7 days with a subsequent secondary fer-mentation (maturation) of several weeks. A batch culture systememploying immobilization could benefit from an increased rateof fermentation. However, it appears that in terms of increasingproductivity, a continuous fermentation system with immobi-lization would be the best method (Verbelen et al. 2006). Animportant issue of the research area is to whether beer can beproduced by immobilized yeast in continuous culture with thesame characteristic as the traditional method.

Page 17: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 643

Table 33.6. Some Selected Applications of Cell Immobilization Systems Used for Beer Production

Carrier Material Reactor Type Reference

Flavor maturationCalcium alginate beads Fixed bed Shindo et al. (1994)DEAE-cellulose Fixed bed Pajunen and Gronqvist (1994)Polyvinyl alcohol beads Fixed bed Smogrovicova et al. (2001)Porous glass beads Fixed bed Linko et al. (1993), Aivasidis (1996)Alcohol-free beerDEAE-cellulose beads Fixed bed Collin et al. 1991, Lomni et al. 1990Porous glass beads Fixed bed Aivasidis et al. (1991)Silicon carbide rods Monolith reactor Van De Winkel et al. (1991)

Acidified wortDEAE-cellulose beads Fixed bed Pittner et al. (1993)

Main fermentationCalcium alginate beads Gas lift White and Portno (1979), Nedovic et al. (1997)Calcium pectate beads Gas lift Smogrovicova et al. (1997)κ-Carrageenan beads Gas lift Mensour et al. (1996), Decamps et al. (2004)Ceramic beads Fixed bed Inoue (1995)Corncobs Gas lift Branyik et al. (2006a)Gluten pellets Fixed bed Bardi et al. 1997Polyvinyl alcohol beads Gas lift Smogrovicova et al. (2001)Polyvinyl alcohol Lentikats R© Gas lift Smogrovicova et al. (2001), Bezbradica et al. (2007)Polyvinyl chloride granules Gas lift Moll et al. (1973)Porous glass beads Fixed bed Virkajarvi and Kronlof (1998)Porous chitosan beads Fluidized bed Unemoto et al. (1998), Maeba et al. (2000)Silicon carbide rods Monolith reactor Andries et al. (1996)Spent grains Gas lift Branyik et al. (2002, 2004)Stainless steel fibre cloths Gas lift Verbelen et al. (2006)Wood (aspen beech) chips Fixed bed Linko et al. (1997), Kronlof and Virkajarvi (1999),

Pajunen et al. (2001)

Started in 1971, one of the first ICT systems for the fer-mentation and maturation of beer was developed by TREPAL(the R&D Centre of the Konenbourg Brewery and the Euro-pean Brewery) with INSA (Institut National des Sciences Ap-pliquees, Toulouse, France) and INRA (Institus National desRecherche Agronomique, Dijon, France) (Moll et al. 1973,Moll and Dueurtre 1996, Moll 2006). A continuous pilot in-stallation for fermentation and maturation of beer with a flowrate of 5 L/h was developed and worked for 9 months withoutany microbial contamination. Around the same time, Narzissand Hellich (1971) developed an ICT process where yeast cellswere immobilized in kieselguhr (which is widely used in thebrewing industry as a filter aid) and a kieselguhr filter was em-ployed as bioreactor (called the “bio-brew bioreactor”). Theiryeast cell immobilization method was based on a method forthe immobilization of enzymes (Berdelle-Hilge 1966). Fromthe beginning of the nineteen seventies, various systems havebeen developed and some have been implemented on anindustrial scale.

ICT processes have been developed for (i) the flavor mat-uration of green beer, (ii) the production of alcohol-free orlow-alcohol beer, (iii) the production of acidified wort usingimmobilized lactic acid bacteria, and (iv) the continuous beer

fermentation (Branyik et al. 2005, Verbelen et al. 2006, Willaertand Nedovic 2006, Willaert 2007, Verbelen et al. 2010).

Flavor Maturation of Green Beer

The objective of flavor maturation is the removal of di-acetyl and 2,3-pentanedione, and their precursors α-acetolactateand α-acetohydroxybutyrate, which are produced during themain fermentation (see the preceding text). The conversionof α-acetohydroxy acids to the vicinal diketones is the rate-limiting step. This reaction step can be accelerated by heat-ing the beer—after yeast removal—to 80–90◦C during a cou-ple of minutes. The resulting vicinal diketones are subse-quently reduced by immobilized cells into their less-flavor-activecompounds.

The traditional maturation process is characterized by a near-zero temperature, low pH and low yeast concentration, resultingin a very long maturation period of 3–4 weeks. Different strate-gies have been developed to accelerate diacetyl removal (seeSection “Vicinal diketones”). One of the techniques is the useof an ICT process to reduce the maturation period to about2 hours.

Page 18: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

644 Part 5: Fruits, Vegetables, and Cereals

Two continuous maturation systems have been implementedindustrially so far: one at Sinebrychoff Brewery (Finland, capac-ity: 1 million hL per year) (Pajunen 1995) and another system,developed by Alfa Laval and Schott Engineering (Mensour et al.1997). They are both composed of a separator (to prevent grow-ing yeast cells in the next stages), an anaerobic heat treatmentunit (to accelerate the chemical conversion of α-acetolactateto diacetyl, but also the partial directly conversion to acetoin),and a packed bed reactor with yeast immobilized on DEAE-cellulose granules or porous glass beads (to reduce the remain-ing diacetyl), respectively (Yamauchi et al. 1994). Later on, theDEAE-cellulose carriers were replaced by cheaper wood chips(Virkajarvi 2002). The heat treatment has been replaced by anenzymatic transformation in a fixed bed reactor in which the α-acetolactate decarboxylase is immobilized in special multilayercapsules, followed by the reduction of diacetyl by yeast in asecond packed bed reactor (Nitzsche et al. 2001).

Production of Alcohol-Free or Low-Alcohol Beer

The classical technology to produce alcohol-free or low-alcoholbeer is based on the suppression of alcohol formation by ar-rested batch fermentation (Narziss et al. 1992). However, theresulting beers are characterized by an undesirable wort aroma,since the wort aldehydes have only been reduced to a limiteddegree (Collin et al. 1991, Debourg et al. 1994, van Iersel et al.1998). The reduction of these wort aldehydes can be quicklyachieved by a short contact time with immobilized yeast cells ata low temperature without undesirable cell growth and ethanolproduction. A disadvantage of this short contact process is theproduction of only a small amount of desirable esters.

Controlled ethanol production for low-alcohol and alcohol-free beers have been successfully achieved by partial fermen-tation using DEAE-cellulose as carrier material, which waspacked in a column reactor (Collin et al. 1991, Van Dieren1995). This technology has been successfully implemented byBavaria Brewery (The Netherlands) to produce malt beer on anindustrial scale (150,000 hL/year) (Pittner et al. 1993). Severalother companies—that is, Faxe (Denmark), Ottakringer (Aus-tria), and a Spanish brewery—have also implemented this tech-nology (Mensour et al. 1997). In Brewery Beck (Germany), afluidized-bed pilot scale reactor (8 hL/day) filled with porousglass beads was used for the continuous production of nonal-coholic beer (Aivasidis et al. 1991, Breitenbucher and Mistler1995, Aivasidis 1996). Yeast cells immobilized in silicon carbiderods and arranged in a multichannel loop reactor (Meura, Bel-gium) have been used to produce alcohol-free beer at pilot scaleby Grolsch Brewery (The Netherlands) and Guinness Brewery(Ireland) (Van De Winkel 1995).

Nuclear mutants of S. cerevisiae that are defective in thesynthesis of tricarboxylic acid cycle enzymes; that is, fu-marase (Kaclıkova et al. 1992) or 2-oxoglutarate dehydrogenase(Mockovciakova et al. 1993) have been immobilized in calciumpectate gel beads and used in a continuous process for the pro-duction of nonalcoholic beer (Navratil et al. 2000). These strainsproduced minimal amounts of ethanol and they were also ableto produce much lactic acid (up to 0.64 g/dm3).

Production of Acidified Wort Using Immobilized LacticAcid Bacteria

The objective of this technology is the acidification of the wortaccording to the “Reinheidsgebot”, before the start of the boil-ing process in the brewhouse. An increased productivity ofacidified wort has been obtained using immobilized Lactobacil-lus amylovorus on DEAE-cellulose beads (Pittner et al. 1993,Meersman 1994). The pH of wort was reduced below a valueof 4.0 after contact times of 7–12 minutes using a packed-bedreactor in downflow mode. The produced acidified wort wasstored in a holding tank and used during wort production toadjust the pH.

Continuous Main Fermentation

During the main fermentation of beer, not only ethanol is beingproduced but also a complex mixture of flavor-active secondarymetabolites, of which the higher (or fusel) alcohols and estersare the most important (Verbelen et al. 2010). In addition, di-acetyl and some sulfury compounds can cause off-flavors. Sincethis complex flavor profile is closely related to the amino acidmetabolism and consequently to the growth of the yeast cells,differences in the growth metabolic state between freely sus-pended and immobilized yeast cell systems are most probablyresponsible for the majority of alterations in the beer flavor. Forthat reason, it is important that the physiological and metabolicstate of the yeast in conventional batch systems is mimicked asmuch as possible during the continuous fermentation with im-mobilized yeast. In the continuous mode of operation, cells arenot exposed to significant alterations of the environment, influ-encing the metabolism of the cells and consequently the flavor.Hence, the microbial population of continuous systems lacks thedifferent growth phases of a batch culture. To imitate the batchprocess as much as possible, plug-flow reactors or a series ofreactors can be used. As can be assumed, both the continuousmode of operation and the immobilization of yeast cells caninfluence the beer flavor.

The Japanese brewery Kirin developed a multistage contin-uous fermentation process (Inoue 1995, Yamauchi et al. 1994,Yamauchi and Kasahira 1995). The first stage is a stirred tankreactor for yeast growth, followed by packed-bed fermenters,and the final step is a packed-bed maturation column. The firststage ensures adequate yeast cell growth with the desirable FANconsumption. Ca-alginate was initially selected as carrier ma-terial to immobilize the yeast cells. These alginate beads werelater replaced by ceramic beads (“Bioceramic R©”). This systemallowed to produce beer within 3–5 days.

The engineering company Meura (Belgium) developed a reac-tor configuration with a first stage with immobilized yeast cellswhere partially attenuation and yeast growth occurs, followedby a stirred tank reactor (with free yeast cells) for complete at-tenuation, ester formation, and flavor maturation (Andries et al.1996, Masschelein and Andries 1995). Silicon carbide rods areused in the first reactor as immobilization carrier material. Thestirred tank (second reactor) is continuously inoculated by freecells that escape from the first immobilized yeast cell reactor.

Page 19: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 645

Labatt Breweries (Interbrew, Canada) in collaboration withthe Department of Chemical and Biochemical Engineering atthe University of Western Ontario (Canada) developed a contin-uous system using κ-carrageenan immobilized yeast cells in anairlift reactor (Mensour et al. 1995, 1996, 1997). Pilot scale re-search showed that full attenuation was reached in 20–24 hourswith this system compared to 5–7 days for the traditional batchfermentation. The flavor profile of the beer produced using ICTwas similar to the batch fermented beer.

Hartwell Lahti and VTT Research Institute (Finland) devel-oped a primary fermentation system using ICT on a pilot scaleof 600 L/day (Kronlof and Virkajarvi 1999). Woodchips wereused as the carrier material that reduced the total investment costby one-third compared to more expensive carriers. The resultsshowed that fermentation and flavor formation were very simi-lar compared to a traditional batch process, although the processtime was reduced to 40 hours.

Andersen et al. (1999) developed a new ICT process in whichthe concentration of carbon dioxide is controlled in a fixed-bedreactor in such a way that the CO2 formed is kept dissolved andis removed from the beer without foaming problems. DEAE-cellulose was used as carrier material. High-gravity beer of ac-ceptable quality has been fermented in 20 hours at a capacityof 50 L/h.

A well-explored concept for main beer fermentation is the useof a gas-lift bioreactor system (see Table 33.6). In this system,mixing established by the circulation of liquid and solid phases,provided high liquid circulation rates, low shear environment,and good mass transfer properties (Siegel and Robinson 1992,Vunjak-Novakovic et al. 1992, Chisti and Moo-Young 1993,Baron et al. 1996). Additionally, they possess the following pos-itive characteristics: high loadings of solids, simple construction,low risk of contamination, easy adjustment and control of op-erational parameters and simple capacity enlargement (Vunjak-Novakovic et al. 1998). Various carrier materials for gas-liftbioreactors have been studied to perform the main fermentation(see Table 33.6).

The optimization of temperature, wort gravity, feed volume,and wort composition seems to be an important tool for thecontrol of the flavor-active compounds formation in immobi-lized beer fermentation systems (Verbelen et al. 2006, 2010,Willaert and Nedovic 2006, Branyik et al. 2008). Many re-searchers have concluded that the optimization of aeration duringcontinuous fermentation is essential for the quality of the finalbeer (Virkajarvi et al. 1999). Oxygen is needed for the forma-tion of unsaturated fatty acids and sterols that needed for growth(Depraetere et al. 2003). However, excess oxygen will lead tolow ester production but to excessive diacetyl, acetaldehyde, andfusel alcohol formation (Okabe et al. 1992, Wackerbauer et al.2003, Branyik et al. 2004). It is possible to adjust the flavor ofthe produced beer by ensuring the adequate amount of dissolvedoxygen by sparging with a mixture of air, nitrogen, or carbondioxide (Kronlof and Linko 1992, Branyik et al. 2004). How-ever, it remains difficult to predict the right amount of oxygenbecause the oxygen availability to the immobilized yeast cellsis dependent of external and internal mass transfer limitations(Willaert and Baron 1996, Willaert et al. 2004).

Also, the reactor design, the carrier, and yeast strain can havea dominant effect on flavor formation (Cop et al. 1989, Linko,et al. 1997, Smogrovicova and Domeny 1999, Tata et al. 1999,Virkajarvi and Pohjala 2000).

ACKNOWLEDGMENTSThis work was supported by the Belgian Federal Science Pol-icy Office and European Space Agency PRODEX program, theInstitute for the Promotion of Innovation by Science and Tech-nology in Flanders (IWT), and the Research Council of theVUB.

REFERENCES

Aivasidis A. 1996. Another look at immobilized yeast systems.Cerevisia 21(1): 27–32.

Aivasidis A et al. 1991. Continuous fermentation of alcohol-freebeer with immobilized yeast cells in fluidized bed reactors. Pro-ceedings of the European Brewery Convention Congress, Lisbon,pp. 569–576.

Alexander MA, Jeffries TW. 1990. Respiratory efficiency andmetabolite partitioning as regulatory phenomena in yeast. En-zyme Microb Technol 12: 2–19.

Alves SL Jr et al. 2007. Maltose and maltotriose active transportand fermentation by Saccharomyces cerevisiae. J Am Soc BrewChem 65: 99–104.

Alves SL Jr et al. 2008. Molecular analysis of maltotriose activetransport and fermentation by Saccharomyces cerevisiae reveals adeterminant role for the AGT1 permease. Appl Environ Microbiol74: 1494–1501.

Andersen K et al. 1999. New process for the continuous fermenta-tion of beer. Proceedings of the European Brewery ConventionCongress, Cannes, pp. 771–778.

Anderson RG, Kirsop BH. 1975a. Oxygen as regulator of esteraccumulation during the fermentation of worts of high gravity.J Inst Brew 80: 111–115.

Anderson RG, Kirsop BH. 1975b. Quantitative aspects of the controlby oxygenation of acetate ester formation of worts of high specificgravity. J Inst Brews 81: 296–301.

Andries M et al. 1996. Design and application of an immobilizedloop bioreactor for continuous beer fermentation. In: RH Wijffelset al. (eds.) Immobilized Cells: Basics and Applications. Elsevier,Amsterdam, pp. 672–678.

Arikawa Y et al. 1998. Soluble fumarate reductase isoenzymes fromSaccharomyces cerevisiae are required for anaerobic growth.FEMS Microbiol Lett 165: 111–116.

Arikawa Y et al. 1999. Isolation of sake yeast strains possessingvarious levels of succinate- and/or malate-producing abilities bygene disruption or mutation. J Biosci Bioeng 87: 333–339.

Asano T et al. 1999. Effect of NAD+-dependent isocitrate dehydro-genase gene (IDH1, IDH2) disruption of sake yeast on organicacid composition in sake mash. J Biosci Bioeng 88: 258–263.

Avhenainen J, Makinen V. 1989. The effect of pitching yeast aera-tion on fermentation and beer flavor. Proceedings of the EuropeanBrewery Convention Congress, pp. 517–519.

Ayrapaa T. 1971. Biosynthetic formation of higher alcohols byyeast. Dependence on the nitrogenous nutrient level of themedium. J Inst Brew 77: 266–275.

Page 20: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

646 Part 5: Fruits, Vegetables, and Cereals

Baillie GS and Douglas LJ. 1999. Role of dimorphism in the de-velopment of Candida albicans biofilms. J Med Microbiol 48:671–679.

Bamforth CW and Kanauchi M. 2004. Enzymology of vicinal dike-tone reduction in brewer’s yeast. J Inst Brew 110: 83–93.

Bardi E et al. 1997. Room and low temperature brewing with yeastimmobilized on gluten pellets. Process Biochem 32: 691–696.

Baron GV et al. 1996. Immobilised cell reactors. In: RG Willaertet al. (eds.) Immobilised Living Cell Systems: Modellingand Experimental Methods. John Wiley & Sons, Chichester,pp. 67–95

Berdelle-Hilge P. 1966. Verfahren und Vorrichtung zur kontinuer-lichen Behandlung von Flussigkeiten mit Einzymtragern. Ger-man patent 1517814.

Bezbradica D et al. 2007. Immobilization of yeast cells in PVAparticles for beer fermentation. Process Biochem 42: 1348–1351.

Bidard F et al. 1994. Cloning and analysis of a FLO5 floccula-tion gene from Saccharomyces cerevisiae yeast. Curr Genet 25:196–201.

Bidard F et al. 1995. The Saccharomyces cerevisiae FLO1 floccu-lation gene encodes for a cell surface protein. Yeast 11: 809–822.

Bishop LR. 1970. A system of continuous fermentations. J InstBrew 76: 172–181.

Bisson LF et al. 1993. Yeast sugar transporters. Crit Rev BiochemMol Biol 28: 259–308.

Blomqvist K et al. 1991. Chromosomal integration and expressionof two bacterial α-acetolactate decarboxylase genes in brewer’syeast. Appl Environ Microbiol 57: 2796–2803.

Bony M et al. 1997. Localisation and cell surface anchoring of theSaccharomyces cerevisiae flocculation protein Flo1p. J Bacteriol179: 4929–4936.

Boswell CD et al. 2002. Studies on the effect of mechanical agitationon the performance of brewing fermentations: fermentation rate,yeast physiology, and development of flavor compounds. J AmSoc Brew Chem 60: 101–106.

Boulton CA. 2000. Trehalose, glycogen and sterol. In: KA Smart(ed.) Brewing Yeast Fermentation Performance. Blackwell Sci-ence, Oxford, pp. 10–19.

Boulton CA et al. 1991. Yeast physiological condition and fermen-tation performance. Proceedings of the 21st European BreweryConvention, pp. 385–392.

Boulton C, Quain D. 2006. Brewing Yeast and Fermentation. Black-well Science, Oxford.

Branyik T et al. 2002. Continuous primary beer fermentation withbrewing yeast immobilized on spent grains. J Inst Brew 108:410–415.

Branyik T et al. 2004. Continuous primary fermentation of beerwith yeast immobilized on spent grains—the effect of operationalconditions. J Am Soc Brew Chem 62: 29–34.

Branyik T et al. 2005. Continuous beer fermentation using immobi-lized yeast cell bioreactor systems. Biotechnol Prog 21: 653–663.

Branyik T et al. 2006a. Continuous immobilized yeast reactorsystem for complete beer fermentation using spent grains andcorncobs as carrier materials. J Ind Microbiol Biotechnol 33:1010–1018.

Branyik T et al. 2008. A review of flavour formation in continuousbeer fermentations. J InstBrew 114: 3–13.

Breitenbucher K, Mistler M. 1995. Fluidized-bed fermenters forthe continuous production of non-alcoholic beer with open-pore

sintered glass carriers. EBC Monograph XXIV, EBC Sympo-sium on Immobilized Yeast Applications in the Brewing Industry,pp. 77–89.

Bullis BL, Lamire BD. 1994. Isolation and characterization of theSaccharomyces cerevisiae SDH4 gene encoding a membraneanchor subunit of succinate dehydrogenase. J Biol Chem 269:6543–6549.

Calderbank J, Hammond JRM. 1994. Influence of higher alcoholavailability on ester formation by yeast. J Am Soc Brew Chem 52:84–90.

Caro LH et al. 1997. In silicio identification of glycosyl-phosphatidylinositol-anchored plasma-membrane and cell wallproteins of Saccharomyces cerevisiae. Yeast 13: 1477–1489.

Chan SY, Appling DR. 2003. Regulation of S-adenosylmethioninelevels in Saccharomyces cerevisiae. J Biol Chem 278:43051–43059.

Chang YS et al. 1988. MAL63 codes for a positive regulator ofmaltose fermentation in Saccharomyces cerevisiae. Curr Genet14: 201–209.

Chapman KB et al. 1992. SDH1, the gene encoding the succinatedehydrogenase flavoprotein subunit from Saccharomyces cere-visiae. Gene 118: 131–136.

Charalambous G et al. 1972. Effect of structurally modified polyphe-nols on beer flavour and flavour stability, part II. Tech QuartMaster Brew Assoc Am 9: 131–135.

Chen EC-H. 1978. Relative contribution of Ehrlich and biosyntheticpathways to the formation of fusel alcohols. J Am Soc Brew Chem36: 39–43.

Chen EC-H. 1980. Utilisation of fatty acids by yeast during fermen-tation. J Am Soc Brew Chem 38: 148–153.

Chisti Y, Moo-Young M. 1993. Improve the performance of airliftreactors. Chem Eng Prog 89: 33–45.

Chow T et al. 1983. Identification and physical characterization ofthe maltose of yeast maltase structural genes. Mol Gen Genet191: 366–371.

Cohen JD et al. 1984. Mutational analysis of MAL1 locus of Saccha-romyces: identification and functional characterization of threegenes. Mol Gen Genet 196: 208–216.

Collin S et al. 1991. Yeast dehydrogenase activities in relation to car-bonyl compounds removal from wort and beer. Proceedings of theEuropean Brewery Convention Congress, Lisbon, pp. 409–416.

Coote N, Kirsop BH. 1974. The content of some organic acids inbeer and in other fermented media. J Inst Brew 80: 474–483.

Cop J et al. 1989. Reactor design optimization with a view onthe improvement of amino acid utilization and flavour devel-opment of calcium-alginate entrapped brewing yeast fermenta-tions. Proceedings of the European Brewery Convention, Zurich,pp. 315–322.

Cormack BP et al. 1999. An adhesin of the yeast pathogen Candidaglabrata mediating adherence to human epithelial cells. Science285: 578–582.

Coutts MW. 1957. A continuous process for the production of beer.UK Patents 872, 391–400.

Crumplen RM et al. 1996. Characteristics of maltose transporteractivity in an ale and large strain of the yeast Saccharomycescerevisiae. Lett Appl Microbiol 23: 448–452.

Daignan-Fournier B et al. 1994. Structure and regulation ofSDH3, the yeast gene encoding the cytochrome b560 subunit byrespiratory complex II. J Biol Chem 269: 15469–15472.

Page 21: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 647

D’Amore T et al. 1991. Advances in the fermentation of highgravity wort. Proceedings of the European Brewery ConventionCongress, pp. 337–344.

Day RE et al. 2002a. Characterization of the putative maltosetransporters encoded by YDL247w and YJR160c. Yeast 19:1015–1027.

Day RE et al. 2002b. Molecular analysis of maltotriose transport andutilization by Saccharomyces cerevisiae. Appl Environ Microbiol68: 5326–5335.

Debourg A. 2002. Yeast in action: from wort to beer. Cerevisia27(3): 144–154.

Debourg A et al. 1990. Effets of ILV genes doses on the flux throughthe isoleucine-valine pathway in Saccharomyces cerevisiae. Ab-stract C31, 6th International Symposium on Genetics of IndustrialMicroorganisms.

Debourg A et al. 1994. Wort aldehyde reduction potential in free andimmobilized yeast systems. J Amer Soc Brew Chem 52: 100–106.

Decamps C et al. 2004. Continuous pilot plant-scale immobilizationof yeast in κ-carrageenan gel beads. AIChE J 50: 1599–1605.

De Groot PWJ et al. 2003. Genome-wide identification of fungalGPI proteins. Yeast 20: 781–796.

Delvaux F. 1998. Beheersing van gistingsproducten met belangrijkesensorische eigenschappen. Cerevisia 23(4): 36–45.

Depraetere SA. 2007. The potential of yeast preoxygenation forapplication in the brewing industry. Doctoral dissertation, no.751, Faculty of Bioscience Engineering, Katholieke UniversiteitLeuven, Belgium.

Depraetere SA et al. 2003. Evaluation of the oxygen requirement oflager and ale yeast strains by preoxygenation. MBAA Tech Quart40: 283–289.

Dequin S. 2001. The potential of genetic engineering for improvedbrewing wine-making and baking yeasts. Appl Microbiol Biotech-nol 56: 577–588.

Devuyst R et al. 1991. Oxygen transfer efficiency with a view toimprovement of lager yeast performance and eer quality. Pro-ceedings of the European Brewery Convention Congress, Lisbon,pp. 377–384.

Dickinson JR. 1999. Carbon metabolism. In: JR Dickinson,M Schweizer (eds.) The Metabolism and Molecular Physiol-ogy of Saccharomyces cerevisiae. Taylor & Francis, London,pp. 23–55.

Dickinson JR et al. 1997. A 13C nuclear magnetic resonanceinvestigation of the metabolism of leucine to isoamyl alco-hol in Saccharomyces cerevisiae. J Biol Chem 272: 26871–26878.

Dickinson JR et al. 1998. An investigation of the metabolism ofvaline to isobutyl alcohol in Saccharomyces cerevisiae. J BiolChem 273: 25751–25756.

Dickinson JR et al. 2000. An investigation of the metabolism ofisoleucine to active amyl alcohol in Saccharomyces cerevisiae.J Biol Chem 275: 10937–10942.

Dickinson JR et al. 2003. The catabolism of amino acids to longchain and complex alcohols in Saccharomyces cerevisiae. J BiolChem 278: 8028–8034.

Dietvorst J et al. 2005. Maltotriose utilization in lager yeast strains:MTT1 encodes a maltotriose transporter. Yeast 22: 775–788.

Dietvorst J et al. 2007. Attachment of MAL32-encoded maltase onthe outside of yeast cells improves maltotriose utilization. Yeast24: 27–38.

Dillemans M et al. 1987. The amplification effect of the ILV5 gene onthe production of vicinal diketones in Saccharomyces cerevisiae.J Am Soc Brew Chem 45: 81–85.

Duan W et al. 2004. A parallel analysis of H2S and SO2 formationby brewing yeast in response to sulphur-containing amino acidsand ammonium ions. J Am Soc Brew Chem 62: 35–41.

Dufour J-P. 1991. Influence of industrial brewing and fermentationworking conditions on beer SO2 levels and flavour stability. Pro-ceedings of the European Brewery Convention Congress, Lisbon,pp. 209–216.

Dufour J-P, Malcorps P. 1994. Ester synthesis during fermenta-tion: enzyme characterization and modulation mechanism. Pro-ceedings of the 4th Institute of Brewing Aviemore Conference,Aviemore, pp. 137–151.

Duval EH et al. 2009. Microarray karyotyping of maltose-fermenting Saccharomyces yeasts with differing maltotriose uti-lization profiles reveals copy number variation in genes involvedin maltose and maltotriose utilization. J Appl Microbiol 109:248–259.

Ehrlich F. 1904. Uber das naturliche isomere des leucins. Berichteder Deutschen Chemisten Gesellschaft 37: 1809–1840.

Enari T-M et al. 1992. Process for accelerated beer production byintegrative expression in the PGK1 or ADC1 genes. US Patent5,108,925.

Engan S. 1969. Wort composition and beer flavour. I. The influenceof some amino acids on the formation of higher aliphatic alcoholsand esters. J Inst Brew 76: 254–261.

Engan S. 1972. Organoleptic threshold values of some alcohols andesters in beer. J Inst Brew 78: 33–36.

Engan S, Aubert O. 1977. Relations between fermentation temper-ature and the formation of some flavour components. Proceed-ings of the European Brewery Convention Congress, Amsterdam,pp. 591–607.

Enomoto K et al. 2002. Physiological role of soluble fumarate re-ductase in redox balancing during anaerobiosis in Saccharomycescerevisiae. FEMS Microbiol Lett 24: 103–108.

Federoff HJ et al. 1983a. Carbon catabolite repression of mal-tase synthesis in Saccharomyces cerevisiae. J Bacteriol 156:301–307.

Federoff HJ et al. 1983b. Regulation of maltase synthesis in Sac-charomyces carlsbergensis. J Bacteriol 154: 1301–1308.

Fraleigh S et al. 1989. Regulation of oxidoreductive yeastmetabolism by extracellular factors. J Biotechnol 12: 185–198.

Francke Johannesen P et al. 1999. Construction of S. carlsbergensisbrewer’s yeast without production of sulfite. Proceedings of theEuropean Brewery Convention Congress, Cannes, pp. 655–662.

Francois J, Parrou JL. 2001. Reserve carbohydrates metabolism inthe yeast Saccharomyces cerevisiae. FEMS Microbiol Rev 25:125–145.

Frieman MB et al. 2002. Modular domain structure in the Candidaglabrata adhesin Epa1p, a β-1,6 glucan-cross-linked cell wallprotein. Mol Microbiol 46: 479–492.

Fu Y et al. 2002. Candida albicans Als1p: an adhesin that is adownstream effector of the EFG1 filamentation pathway. MolMicrobiol 44: 61–72.

Fujii T et al. 1990. Application of a ribosomial DNA integration vec-tor in the construction of a brewers’ yeast having α-acetolactatedecarboxylase activity. Appl Environ Microbiol 56: 997–1003.

Page 22: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

648 Part 5: Fruits, Vegetables, and Cereals

Fujii T et al. 1994. Molecular cloning, sequence analysis, and ex-pression of the yeast alcohol acetyltransferase gene. Appl EnvironMicrobiol 60: 2786–2792.

Fukuda K et al. 1998. Balance of activities of alcohol acetyltrans-ferase and esterase in Saccharomyces cerevisiae is importantfor production of isoamyl acetate. Appl Environ Microbiol 64:4076–4078.

Gaur NK, Klotz SA. 1997. Expression, cloning, and characteriza-tion of a Candida albicans gene, ALA1, that confers adherenceproperties upon Saccharomyces cerevisiae for extracellular ma-trix proteins. Infect Immun 65: 5289–5294.

Gee DA, Ramirez WF. 1994. A flavour model for beer fermentation.J Inst Brew 100: 321–329.

Geiger E, Piendl A. 1976. Technological factors in the formation ofacetaldehyde during fermentation. MBAA Tech Quart 13: 51–61.

Gjermansen C et al. 1988. Towards diacetyl-less brewers’ yeast. In-fluence of ilv2 and ilv5 mutations. J Basic Microbiol 28: 175–183.

Godtfredsen SE et al. 1983. Application of the acetolactate de-carboxylase from Lactobacillus casei for accelerated matura-tion of beer. Proceedings of the European Brewery ConventionCongress, London, pp. 161–168.

Godtfredsen SE et al. 1984. Occurrence of α-acetolactate decar-boxylases among lactic acid bacteria and their utilization formaturation of beer. Appl Microbiol Biotechnol 20: 23–28.

Goldenthal MJ, Vanoni M. 1990. Genetic mapping and biochem-ical analysis of mutants in the maltose regulatory gene of theMAL1 locus of Saccharomyces cerevisiae. Arch Microbiol 154:544–549.

Goldenthal MJ et al. 1987. Regulation of MAL gene expression inyeast: gene dosage effects. Mol Gen Gen 209: 508–517.

Goossens E et al. 1993. Decreased diacetyl production in lagerbrewing yeast by integration of the ILV5 gene. Proceedings of theEuropean Brewery Convention Congress, Oslo, pp. 251–258.

Green CB et al. 2004. T-PCR detection of Candida albicans ALSgene expression in the reconstituted human epithelium (RHE)model of oral candidiasis and in model biofilms. Microbiology150: 267–275.

Guo B et al. 2000. A Saccharomyces gene family involved in inva-sive growth, cell-cell adhesion, and mating. Proc Natl Acad SciUSA 97: 12158–12163.

Hammond JRM. 1993. Brewer’s yeast. In: AH Rose, JS Harrison(eds.) The Yests. Academic Press, London, pp. 7–67.

Han E-K et al. 1995. Characterization of AGT1 encoding a generalα-glucoside transporter from Saccharomyces. Mol Microbiol 17:1093–1107.

Hansen J, Kielbrandt MC. 1996a. Inactivation of MET10 in brewers’yeast specifically increases SO2 formation during beer produc-tion. Nat Biotechnol 14: 1587–1591.

Hansen J, Kielbrandt MC. 1996b. Inactivation of MET2 in brewers’yeast increases the level of sulphite in beer. J Biotechnol 50:75–87.

Hardwick WA. 1995. History and antecedents of brewing. In: WAHardwick (ed.) Handbook of Brewing. Marcel Dekker, New York,pp. 37–51.

Haukeli AD, Lie S. 1978. Conversion of α-acetolactate and removalof diacetyl: a kinetic study. J Inst Brew 84: 85–89.

Haukeli AD, Lie S. 1979. Yeast growth and metabolic changes dur-ing brewery fermentation. Proceedings of the European BreweryConvention Congress, Berlin, pp. 461–473.

Hazelwood LA et al. 2008. The Ehrlich pathway for fusel alcoholproduction: a century of research on Saccharomyces cerevisiaemetabolism. Appl Environ Microbiol 74: 2259–2266.

Hegarty PK et al. 1995. Phenyl ethanol—a factor determining lagercharacter. Proceedings of the European Brewery ConventionCongress, pp. 515–522.

Hirata D et al. 1992. Stable overproduction of isoamylalcohol byS. cerevisiae with chromosome-integrated copies of the LEU4genes. Biosci Biotechnol Biochem 56: 1682–1683.

Hough JS et al. 1982. Malting and Brewing Science—Volume 2:Hopped Wort and Beer. Chapman and Hall, London.

Hounsa CG et al. 1998. Role of trehalose in survival of Sac-charomyces cerevisiae under osmotic stress. Microbiology 144:671–680.

Hoyer LL. 2001. The ALS gene family of Candida albicans. TrendsMicrobiol 9: 176–180.

Hoyer LL et al. 1998. Identification of Candida albicans ALS2 andALS4 and localization of ALS proteins to the fungal cell surface.J Bacteriol 180: 5334–5343.

Huuskonen A et al. 2010. Selection from industrial lager yeaststrains of variants with improved fermentation performance invery-high-gravity worts. Appl Environ Microbiol 76: 1563–1573.

Hu Z et al. 1995. MIG1-dependent and MIG1-independent glucoseregulation of MAL gene expression in Saccharomyces cerevisiae.Curr Genet 28: 258–266.

Inoue T. 1995. Development of a two-stage immobilized yeastfermentation system for continuous beer brewing. Proceed-ings of the European Brewery Convention Congress, Brussels,pp. 25–36.

Inoue T, Yamamoto Y. 1970. Diacetyl and beer fermentation. Pro-ceedings of the Annual Meeting of the American Society ofBrewing Chemists, pp. 198–208.

Inoue T et al. 1968. Mechanism of diacetyl formation in beer. Pro-ceedings of the Annual Meeting of the American Society ofBrewing Chemists, MN, pp. 158–165.

Iverson WG. 1994. Ethyl acetate extraction of beer: quantitativedetermination of additional fermentation by-products. J Am SocBrew Chem 52: 91–95.

Jakobsen M. 1982. The effect of yeast handling procedures on yeastoxygen requirement and fermentation. Proc. Inst. Brew. (Aust.N.Z. Section) Conv, Adelaide, pp. 132–137.

Jensen S. 1993. Using ALDC to speed up fermentation. Brewers’Gardian, September: 55–56.

Jespersen L et al. 1999. Multiple α-glucoside transporter genes inbrewer’s yeast. Appl Environ Microbiol 65: 450–456.

Jin YL, Speers RA. 1999. Flocculation of Saccharomyces cere-visiae. Food Res Int 31: 421–440.

Kaclıkova E et al. 1992. Fumaric acid overproduction in yeast mu-tants deficient in fumarase. FEMS Microbiol Lett 91: 101–106.

Kappeli O et al. 1985. Transient responses of Saccharomycesuvarum to a change of growth-limiting nutrient in continuousculture. J Gen Microbiol 131: 47–52.

Karel SF et al. 1985. The immobilization of whole cells: engineeringprinciples. Chem Eng Sci 40: 1321–1353.

Klein CJ.L et al. 1996. Alleviation of glucose repression of maltosemetabolism by MIG1 disruption in Saccharomyces cerevisiae.Appl Environ Microbiol 62: 4441–4449.

Klopper WJ et al. 1986. Organic acids and glycerol in beer. J InstBrew 92: 311–318.

Page 23: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 649

Kobayashi O et al. 1998. Region of Flo1 proteins responsible forsugar recognition. J Bacteriol 180: 6503–6510.

Kodama Y et al. 1995. Improvement of maltose fermentation effi-ciency: constitutive expression of MAL genes in brewing yeasts.J Am Soc Brew Chem 53: 24–29.

Kodama Y et al. 2001. Control of higher alcohol production bymanipulation of the BAP2 gene in brewing yeast. J Am Soc BrewChem 59: 57–62.

Kronlof J, Linko M. 1992. Production of beer using immobilizedyeast encoding α-acetolactate decarboxylase. J Inst Brew 98:479–491.

Kronlof J, Virkajarvi I. 1999. Primary fermentation with immobi-lized yeast. Proceedings of the European Brewery ConventionCongress, Cannes, pp. 761–770.

Kubo Y et al. 2000. Effect of gene disruption of succinate dehy-drogenase on succinate production in sake yeast strain. J BiosciBioeng 90: 619–624.

Kunze W. 1999. Technology of Brewing and Malting. VLB, Berlin.Landaud S et al. 2001. Top pressure and temperature control of the

fusel alcohol/ester ratio through yeast growth in beer fermenta-tion. J Inst Brew 10: 107–117.

Lagunas R. 1979. Energetic irrelevance of aerobiosis for S. cere-visiae growing on sugars. Mol Cell Biochem 27: 139–146.

Lea AH, Piggott JRP. 1995. Fermented Beverage Production.Blackie Academic and Professional, Glasgow.

Lee S et al. 1995. Yeast strain development for enhanced productionof desirable alcohols/esters in beer. J Am Soc Brew Chem 39:153–156.

Li F, Palecek SP. 2003. EAP1, a Candida albicans gene in-volved in binding human epithelial cells. Eukaryot Cell 2:1266–1273.

Lievense JC, Lim HC. 1982. The growth and dynamics of Saccha-romyces cerevisiae. In: GT Tsao et al. (eds.) Annual Reports onFermentation Processes. Academic Press, New York.

Lilly M et al. 2000. Effect of increased yeast alcohol acetyltrans-ferase activity on flavor profiles of wine and distillates. ApplEnviron Microbiol 66: 744–753.

Lilly M et al. 2006. The effect of increased branched-chain aminoacid transaminase activity in yeast on the production of higheralcohols and on the flavour profiles of wine and distillates. FEMSYeast Res 6: 726–743.

Linko M et al. 1993. Use of brewer’s yeast expressing α-acetolactatedecarboxylase in conventional and immobilized fermentations.MBAA Tech Quart 30: 93–97.

Linko M et al. 1997. Main fermentation with immobilized yeast—abreaktrough? Proceedings of the European Brewery ConventionCongress, Maastricht, pp. 385–394.

Linko M et al. 1998. Recent advances in the malting and brewingindustry. J Biotechnol 65: 85–98.

Lo WS, Dranginis AM. 1998. The cell surface flocculin Flo11 isrequired for pseudohyphae formation and invasion by Saccha-romyces cerevisiae. Mol Biol Cell 9: 161–171.

Lombardo A et al. 1990. Cloning and characterization of theiron-sulfur subunit gene succinate dehydrogenase from Saccha-romyces cerevisiae. J Biol Chem 265: 10419–10423.

Lomni H et al. 1990. Immobilized yeast reactor speeds beer pro-duction. Food Technol 5: 128–133.

Lucero P et al. 1993. Catabolite inactivation of the yeast maltosetransporter is due to proteolysis. FEBS Lett 333: 165–168.

MacDonald J et al. 1984. Current approaches to brewery fermenta-tions. Prog Ind Microbiol 19: 47–198.

MacWilliam I.C. 1968. Wort composition. J Inst Brew 74:38–54.

Maeba H et al. 2000. Primary fermentation with immobilized yeastin porous chitosan beads. Pilot scale trial. Proceedings of the 26thConvention of the Institute Brewing, Australia and New ZealandSection, Singapore, pp. 82–86.

Magarifuchi T et al. 1995. Effect of yeast fumarase gene (FUM1)disruption on the production of malic, fumaric and succinic acidsin sake mash. J Ferment Bioeng 80: 355–361.

Magee PT, de Robinson-Szulmajster H. 1968. The regulationof isoleucine-valine biosynthesis in Saccharomyces cerevisiae.Properties and regulation of the activity of acetohydroxyacid syn-thase. Eur J Biochem 3: 507–511.

Majara N et al. 1996. Trehalose: a stress protectant and stress in-dicator compound for yeast exposed to adverse conditions. J AmSoc Brew Chem 54: 221–227.

Malcorps P, Dufour J-P. 1987. Proceedings of the European Brew-ery Convention Congress, Madrid, p. 377.

Malcorps P et al. 1991. A new model for the regulation of ester syn-thesis by alcohol acetyltransferase in Saccharomyces cerevisiae.J Am Soc Brew Chem 49: 47–53.

Masschelein CA. 1981. Role of fatty acids in beer flavour. EBCFlavour Symposium, Copenhagen, Denmark. EBC MonographVII, pp. 211–221, 237–238.

Masschelein CA. 1997. A realistic view on the role of research inthe brewing industry today. J Inst Brew 103: 103–113.

Masschelein CA, Andries M. 1995. Future scenario of immobilizedsystems: promises and limitations. EBC Monograph XXIV, EBCSymposium on immobilized yeast applications in the brewingindustry, Espoo, pp. 223–241.

Masschelein CA et al. 1995. The membrane loop concept: a new ap-proach for optimal oxygen transfer into high cell density pitchingyeast suspensions. Proceedings of the European Brewery Con-vention Congress, Brussels, pp. 377–386.

Maule DRJ. 1967. Rapid gas chromatographic examination of beerflavour. J Inst Brew 73: 351–361.

McGovern PE et al. 1996. Neolithic resinated wine. Nature 381:480–481.

Meersman E. 1994. Biologische aanzuring met geımmobiliseerdemelkzuurbacterien. Cerevisia Biotechnol 19(4): 42–46.

Meilgaard MC. 1975a. Flavour chemistry of beer. Part 1: flavourinteractions between principle volatiles. Tech Quart Master BrewAssoc Am 12: 107–117.

Meilgaard MC. 1975b. Flavour chemistry of beer. Part 2: flavourand threshold of 239 aroma volatiles. Tech Quart Master BrewAssoc Am 12: 151–173.

Mensour N et al. 1995. Gas lift systems for immobilized cell sys-tems. EBC Monograph XXIV, EBC Symposium on ImmobilizedYeast Applications in the Brewing Industry, Espoo, pp. 125–133.

Mensour N et al. 1996. Applications of immobilized yeast cellsin the brewing industry. In: RH Wijffels et al. (eds.) Immo-bilized Cells: Basics and Applications. Elsevier, Amsterdam,pp. 661–671.

Mensour N et al. 1997. New developments in the brewing industryusing immobilized yeast cell bioreactor systems. J Inst Brew 103:363–370.

Page 24: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

650 Part 5: Fruits, Vegetables, and Cereals

Michels CA et al. 1992. The telomere-associated MAL3 locus ofSaccharomyces is a tandem array of repeated genes. Yeast 8:655–665.

Michnick S et al. 1997. Modulation of glycerol and ethanolyields during alcoholic fermentation in Saccharomyces cerevisiaestrains overexpressed or disrupted for GPD1 encoding glycerol-3-phosphate dehydrogenase. Yeast 13: 783–793.

Minetoki T et al. 1993. The purification, properties and internalpeptide sequences of alcohol acetyltransferase isolated from Sac-charomyces cerevisiae. Biosci Biotechnol Biochem 57: 2094–2098.

Mithieux SM, Weiss AS. 1995. Tandem integration of multipleILV5 copies and elevated transcription in polyploid yeast. Yeast11: 311–316.

Mockovciakova D et al. 1993. The ogd1 and kgd1 mutants lacking2-oxoglutarate dehydrogenase activity in yeast are allelic and canbe differentiated by the cloned amber suppressor. Curr Genet 24:377–381.

Moll M. 2006. Fermentation and maturation of beer with immo-bilised yeasts. J Inst Brew 112: 346.

Moll M, Duteurtre B. 1996. Fermentation and maturation of beerwith immobilised microorganisms. Brauwelt Int 14: 248–252.

Moll M et al. 1973. Continuous production of fermented liquids.French Patent 73/23397; US Patent 4009286.

Moonjai N et al. 2000a. Unsaturated fatty acid supplementation ofstationary phase brewing yeast. Cerevisia 25(3): 37–50.

Moonjai N et al. 2000b. The effects of linoleic acid supplementationof cropped yeast on its performance and acetate ester synthesis.J Inst Brew 108: 227–235.

Murray CR et al. 1984. The effect of yeast storage conditions onsubsequent fermentations. MBBA Tech Quart 21: 189–194.

Nakatani K et al. 1984. Kinetic studies of vicinal diketones in brew-ing. 2. Theoretical aspects for the formation of total vicinal dike-tones. Tech Quart Master Brew Assoc Am 21: 175–183.

Narziss L, Hellich P. 1971. Ein Beitrag zur wesentlichen Beschle-unigung der Garung und Reifung des Bieres. Brauwelt 111:1491–1500.

Narziss L et al. 1992. Technology and composition of non-alcoholicbeers. Brauwelt Int 4: 396.

Navratil M et al. 2000. Fermented beverages produced by yeastcells entrapped in ionotropic hydrogels of polysaccharide nature.Minerva Biotechnol 12: 337–344.

Navratil M et al. 2002. Production of non-alcoholic beer using freeand immobilized cells of Sccharomyces cerevisiae deficient intricarboxylic acid cycle. Biotechnol Appl Biochem 35: 133–141.

Nedovic VA et al. 1997. Analysis of liquid axial dispersion in aninternal loop gas-lift bioreactor for beer fermentation with immo-bilized yeast cells. Proceedings of the 2nd European Conferenceon Fluidization, Bilbao, pp. 627–635.

Nedovic V et al. 2005. Beer production using immobilised cells.In: V Nedovic, R Willaert (eds.) Applications of Cell Immobili-sation Biotechnology. Kluwer Academic Publishers, Dordrecht,pp. 259–273.

Needleman RB et al. 1984. MAL6 of Saccharomyces: a complexgenetic locus containing three genes required for maltose fer-mentation. Proc Natl Acad Sci USA. 81: 2811–2815.

Nehlin JO, Ronne H. 1990. Yeast MIG1 repressor is related to themammalian early growth response and Wilms’ tumour fingerproteins. EMBO J 9: 2891–2898.

Nelissen B et al. 1995. Phylogenetic classification of the ma-jor superfamily of membrane transport facilitators, as de-duced from yeast genome sequencing. FEBS Lett 377:232–236.

Nevoigt E, Stahl U. 1997. Reduced pyruvate decarboxylase andincreased glycerol-3-phosphate degydrogenase [NAD+] levelsenhance glycerol production in Saccharomyces cerevisiae. Yeast12: 1331–1337.

Nevoigt E et al. 2002. Genetic engineering of brewing yeast toreduce the content of ethanol in beer. FEMS Yeast Res 2: 225–232.

Ni B, Needleman RB. 1990. Identification of the upstream activatingsequence of MAL and the binding sites for MAL63 activator ofSaccharomyces cerevisiae. Mol Cell Biol 10: 3797–3800.

Nitzsche F et al. 2001. A new way for immobilized yeast sys-tems: secondary fermentation without heat treatment. Proceed-ingsof the European Brewery Convention, Budapest, pp. 486–494.

NN. 2000. Fermentation and Maturation. In: European BreweryConvention Manual of Good Practice. Getranke-Fachverlag HansCarl, Nurnberg.

NN. 2005. Dossier sectoroverzicht van de Belgische brouwerijni-jverheid. Het Brouwersblad. juni: 6–20.

Novak S et al. 2004. Regulation of maltose transport and metabolismin Saccharomyces cerevisiae. Food Technol Biotechnol 42:213–218.

Nykanen L, Nykanen I. 1977. Production of esters by different yeaststrains in sugar fermentations. J Inst Brew 83: 30–31.

Ohno T, Takahashi R. 1986a. Role of wort aeration in the brewingprocess. Part 1: oxygen uptake and biosynthesis of lipids by thefinal yeast. J Inst Brew 92: 84–87.

Ohno T, Takahashi R. 1986b. Role of wort aeration in the brewingprocess. Part 2: the optimal aeration conditions for the brewingprocess. J Inst Brew 92: 88–92.

Okabe M et al. 1992. Growth and fermentation characteristics ofbottom brewer’s yeast under mechanical stirring. J Ferment Bio-eng 73: 148–152.

Omura F. 2008. Targeting of mitochondrial Saccharomyces cere-visiae Ilv5p to the cytosol and its effect on vicinal diketone for-mation in brewing. Appl Microbiol Biotechnol 78: 503–513.

Onnela M-L et al. 1996. Use of a modified alcohol dehydroge-nase, ADH1, promotor in construction of diacetyl non-producingbrewer’s yeast. J Biotechnol 49: 101–109.

Oshita K et al. 1995. Clarification of the relationship between fuselalcohol formation and amino acid assimilation by brewing yeastusing 13C-labeled amino acid. Proceedings of the European Brew-ery Convention Congress, Brussels, pp. 387–402.

Pajunen E. 1995. Immobilized yeast lager beer maturation: DEAE-cellulose at Synebrychoff. EBC Monograph XXIV, EBC Sympo-sium on Immobilized Yeast Applications in the Brewing Industry,Espoo, pp. 24–40.

Pajunen E, Gronqvist A. 1994. Immobilized yeast fermenters forcontinuous lager beer maturation. Proceedings of the 23rd Con-vention Institute of Brewing Australia and New Zealand Section,Sydney, pp. 101–103.

Panchal CJ, Stewart GG. 1979. Utilization of wort carbohydrates.Brew Digest June: 36–46.

Pajunen E et al. 2001. Controlled beer fermentation with contin-uous one-stage immobilized yeast reactor. Proceedings of theEuropean Brewery Convention, Budapest, pp. 465–476.

Page 25: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 651

Panek AD. 1991. Storage carbohydrates. In: AH Rose, JS Harrison(eds.) The Yeasts, Volume 4, Yeast Organelles, 2nd edn. AcademicPress, London, pp. 655–678.

Panek AD, Panek AC. 1990. Metabolism and thermotolerance func-tion of trehalose in Saccharomyces cerevisiae: a current perspec-tive. J Biotechnol 14: 229–238.

Pang SS, Duggleby RG. 1999. Expression, purification, charac-terization, and reconstitution of the large and small subunitsof yeast acetohydroxyacid synthase. Biochemistry 38: 5222–5231.

Pang SS, Duggleby RG. 2001. Regulation of yeast acetohydroxy-acid synthase by valine and ATP. Biochem J 357: 749–757.

Patel GB, Ingledew WM. 1973. Trends in wort carbohydrate uti-lization. Appl Microbiol 26: 349–353.

Parrou JL et al. 1997. Effects of various types of stress on themetabolism of reserve carbohydrates in Saccharomyces cere-visiae: genetic evidence for a stress-induced recycling of glyco-gen and trehalose. Microbiology 143: 1891–900.

Peddie HAB. 1990. Ester formation in brewery fermentations, J InstBrew 96: 327–331.

Peinado JM, Loureiro-Dias MC. 1986. Reversible loss of affinityinduced by glucose in the maltose-H+ symport of Saccharomycescerevisiae. Biochim Biophys Acta 856: 189–192.

Petrik M et al. 1983. An expanded concept for the glucose effectin the yeast Saccharomyces uvarum: involvement of short- andlong-term regulation. J Gen Microbiol 129: 43–49.

Pfisterer E, Stewart GG. 1975. Some aspects on the fermentation ofhigh gravity worts. Proceedings of the European Brewery Con-vention Congress, Nice, pp. 255–267.

Pittner H et al. 1993. Continuous production of acidified wortfor alcohol-free-beer with immobilized lactic acid bacteria. Pro-ceedings of the European Brewery Convention Congress, Oslo,pp. 323–329.

Pugh TA et al. 1997. The impact of wort nitrogen limitation onyeast fermentation performance and diacetyl. MBAA Tech Quart34: 185–189.

Quain DE. 1988. Studies on yeast physiology: impact of fermenta-tion performance and product quality. J Inst Brew 95: 315–323.

Quain DE, Tubb RS. 1982. The importance of glycogen in brewingyeast. MBAA Tech Quart 19: 19–23.

Ramos-Jeunehomme C et al. 1989. Proceedings of the EuropeanBrewery Convention Congress, pp. 513–519.

Ramos-Jeunehomme C et al. 1991. Why is ester formation inbrewery fermentations yeast strain dependent? Proceedings ofthe European Brewery Convention Congress, Lisbon, pp. 257–264.

Rautio J, Londesborough J. 2003. Maltose transport by brewer’syeast in brewer’s wort. J Inst Brew 109: 251–261.

Reed G, Nogodawithana TW. 1991. Yeast Technology. Chapter 3.Van Nostrand Reinhold, New York.

Remize F et al. 1999. Glycerol overproduction by engineered Sac-charomyces cerevisiae wine yeast strains leads to substantialchanges in by-product formation and to stimulation of fermenta-tion rate in stationary phase. Appl Environ Microbiol 65: 143–149.

Reynolds TB, Fink GR. 2001. Baker’s yeast, a model for fungalbiofilm formation. Science 291: 878–881.

Riballo E et al. 1995. Catabolite inactivation of the yeast maltosetransporter occurs in the vacuole after internalization by endocy-tosis. J Bacteriol 177: 5622–5627.

Rieger M et al. 1983. The role of limited respiration in the incom-plete oxidation of glucose by Saccharomyces cerevisiae. J GenMicrobiol 129: 653–661.

Rintala E et al. 2008. Transcription of hexose transporters of Sac-charomyces cerevisiae is affected by change in oxygen provision.BMC Microbiol 8: 53.

Rogers PJ, Stewart PR. 1973. Mitochondrial and peroxisomal con-tributions to the energy metabolism of Saccharomyces cerevisiaein continuous culture. J Gen Microbiol 79: 205–217.

Romano P, Suzzi G. 1992. Production of H2S by different yeaststrains during fermentation. Proceedings of the 22nd Conventionof the Institute of Brewing, Melbourne, Institute of Brewing,Adelaide, Australia, pp. 96–98.

Rosculet G. 1971. Aroma and flavour of beer. Part II: the origin andnature of less-volatile and non-volatile components of beer. BrewDigest 46(6): 96–98.

Rostgaard-Jensen B et al. 1987. Isolation and characterization of anα-acetolactate decarboxylase useful for accelerated beer matura-tion. Proceedings of the European Brewery Convention Congress,pp. 393–400.

Royston MG. 1966. Tower fermentation of beer. Process Biochem1: 215–221.

Ryder D, Masschelein CA. 1983. Aspects of metabolic regulatorysystems and physiological limitations with a view to the improve-ment of brewing performance. EBC Symposium on Biotechnol-ogy, Monograph IX, Nutfiels, pp. 2–29.

Sablayrolles JM, Ball CB. 1995. Fermentation kinetics and the pro-duction of volatiles during alcoholic fermentation. J Am Soc BrewChem 53: 71–78.

Saerens SM et al. 2006. The Saccharomyces cerevisiae EHT1 andEEB1 genes encode novel enzymes with medium-chain fatty acidethyl ester synthesis and hydrolysis capacity. J Biol Chem 281:4446–4456.

Saison D et al. 2009. Contribution of staling compounds to the agedflavour of lager beer by studying their flavour thresholds. FoodChem 114: 1206–1215.

Salema-Oom M et al. 2005. Maltotriose utilization by industrialSaccharomyces strains: characterization of a new member of thealpha-glucoside transporter family. Appl Environ Microbiol 71:5044–5049.

Selecky R et al. 2008. Beer with reduced ethanol content producedusing Saccharomyces cerevisiae yeasts deficient in various tricar-boxylic acid cycle enzymes. J Inst Brew 114: 97–101.

Sheppard DC et al. 2004. Functional and structural diversity inthe Als protein family of Candida albicans. J Biol Chem 279:30480–30489.

Shindo S et al. 1994. Suppression of α-acetolactate formation inbrewing with immobilized yeast. J Inst Brew 100: 69–72.

Siegel MH, Robinson CW. 1992. Applications of airlift gas-liquid-solid reactors in biotechnology. Chem Eng Sci 47: 3215–3229.

Siro M-R, Lovgren T. 1979. Influence of glucose on the α-glucosidepermease activity of yeast. Eur J Appl Microbiol 7: 59–66.

Slaughter JC, Jordan B. 1986. The production of hydrogen sulphideby yeast. In: I Campbell, FG Priest (eds.) Proceedings of theSecond Aviemore Conference on Malting, Brewing and Distilling.Institute of Brewing, London, pp. 308–310.

Smit A et al. 2008. The Thr505 and Ser557 residues of the AGT1-encoded alpha-glucoside transporter are critical for maltotriose

Page 26: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

652 Part 5: Fruits, Vegetables, and Cereals

transport in Saccharomyces cerevisiae. J Appl Microbiol 104:1103–1111.

Smith AU et al. 1998. Yeast PKA represses Msn2p/Msn4p-dependent gene expression to regulate growth, stress responseand glycogen accumulation. EMBO J 17: 3556–3564.

Smogrovicova D, Domeny Z. 1999. Beer volatile by-product for-mation at different fermentation temperature using immobilizedyeasts. Process Biochem 34: 785–794.

Smogrovicova D et al. 1997. Reactors for the continuous primarybeer fermentation using immobilised yeast. Biotechnol Tech 11:261–264.

Smogrovicova D et al. 2001. Continuous beer fermentation usingpolyvinyl alcohol entrapped yeast. Proceedings of the EuropeanBrewery Convention Congress, Budapest, pp. 1–9.

Soler AP et al. 1987. Differential translational efficiency of themRNA isolated from derepressed and glucose repressed Saccha-romyces cerevisiae. J Gen Microbiol 133: 1471–1480.

Staab JF et al. 1999. Adhesive and mammalian transglutaminasesubstrate properties of Candida albicans Hwp1. Science 283:1535–1583.

Stambuk BU, de Araujo PS. 2001. Kinetics of active alpha-glucosidetransport in Saccharomyces cerevisiae. FEMS Yeast Res 1: 73–78.

Stewart GG, Russell I. 1993. Fermentation—the “black box” of thebrewing process. MBAA Tech Quart 30: 159–168.

Stratford M. 1989. Yeast flocculation: calcium specificity. Yeast 5:487–496.

Suihko M et al. 1990. Recombinant brewer’s yeast strains suitablefor accelerated brewing. J Biotechnol 14: 285–300.

Szlavko CM. 1973. Tryptophol, tyrosol and phenylethanol—thearomatic higher alcohols in beer. J Inst Brew 83: 283–243.

Tada S et al. 1995. Pilot scale brewing with industrial yeasts whichproduce the α-acetolactate decarboxylase of Acetobacter acetissp. xylium. Proceedings of the European Brewery ConventionCongress, Brussels, pp. 369–376.

Tata M et al. 1999. Immobilized yeast bioreactor systems for con-tinuous beer fermentation. Biotechnol Prog 15: 105–113.

Teunissen AW et al. 1993a. Sequence of the open reading frameof the FLO1 gene from Saccharomyces cerevisiae. Yeast 9: 423–427.

Teunissen AW et al. 1993b. Physical localization of the flocculationgene FLO1 on chromosome I of Saccharomyces cerevisiae. Yeast9: 1–10.

Teunissen AW, Steensma HY. 1995. Review: the dominant floccu-lation genes of Saccharomyces cerevisiae constitute a new sub-telomeric gene family. Yeast 11: 1001–1013.

Tezuka H et al. 1992. Cloning of a gene suppressing hydrogen sul-phide production by Saccharomyces cerevisiae and its expressionin a brewing yeast. J Am Soc Brew Chem 50: 130–133.

Thomas D, Surdin-Kerjan Y. 1997. Metabolism of sulfur aminoacids in Saccharomyces cerevisiae. Microbiol Mol Biol Rev 61:503–532.

Thorne RSW. 1968. Continuous fermentation in retrospect. BrewDigest 43(2): 50–55.

Thurston PA et al. 1981. Effects of linoleic acid supplements onthe synthesis by yeast of lipid and acetate esters. J Inst Brew 87:92–95.

Thurston PA et al. 1982. Lipid metabolism and the regulation ofvolatile synthesis in Saccharomyces cerevisiae. J Inst Brew 88:90–94.

Udenfriend S, Kodukula K. 1995. How glyco-sylphosphatidylinositol-anchored membrane proteins aremade. Annu Rev Biochem 64: 563–591.

Unemoto S et al. 1998. Primary fermentation with immobilizedyeast in a fluidized bed reactor. MBAA Tech Quart 35: 58–61.

Vanbeneden N et al. 2007. Formation of 4-vinyl and 4-ethyl deriva-tives from hydroxycinnamic acids: occurrence of volatile pheno-lic flavour compounds in beer and distribution of Pad1-activityamong brewing yeasts. Food Chem 107: 221–230.

Van de Meersche J et al. 1977. The role of yeasts in the maturation ofbeer flavour. Proceedings of the European Brewery ConventionCongress, Amsterdam, pp. 561–575.

Van Den Berg R et al. 1983. Diacetyl reducing activity inbrewer’s yeast. Proceedings of the European Brewery ConventionCongress, London, pp. 497–504.

Van De Winkel L. 1995. Design and optimization of a multipurposeimmobilized yeast bioreactor system for brewery fermentations.Cerevisia 20(1): 77–80.

Van De Winkel L et al. 1991. The application of an immobi-lized yeast loop reactor to the continuous production of alcohol-free beer. Proceedings of the European Brewery ConventionCongress, Lisbon, pp. 307–314.

Van Dieren D. 1995. Yeast metabolism and the production ofalcohol-free beer. EBC Monograph XXIV, EBC Symposium onImmobilized Yeast Applications in the Brewing Industry, Espoo,pp. 66–76.

Van Haecht JL, Dufour JP. 1995. The production of sulfur com-pounds by brewing yeast: a review. Cerevisiae 20: 51–64.

van Iersel MFM et al. 1998. Effect of environmental conditions onflocculation and immobilization of brewer’s yeast during produc-tion of alcohol-free beer. J Inst Brew 104: 131–136.

Van Mulders SE et al. 2009. Phenotypic diversity of Flo proteinfamily-mediated adhesion in Saccharomyces cerevisiae. FEMSYeast Res 9: 178–190.

Verbelen PJ et al. 2006. Immobilized yeast cell systems for contin-uous fermentation applications. Biotechnol Lett 28: 1515–1525.

Verbelen PJ et al. 2009. The influence of yeast oxygenation priorto brewery fermentation on yeast metabolism and the oxidativestress response. FEMS Yeast Res 9: 226–239.

Verbelen PJ et al. 2010. Bioprocess intensification of beer fer-mentation using immobilised cells. In: NJ Zuidam, VA Nedovic(eds.) Encapsulation Technologies for Active Food Ingredientsand Food Processing. Springer, New York, pp. 303–325.

Verstrepen KJ. 2003. Flavour-active ester synthesis in Saccha-romyces cerevisiae. PhD thesis, Katholieke Universiteit Leuven,Belgium.

Verstrepen KJ et al. 2003a. Yeast flocculation: what brewers shouldknow. Appl Microbiol Biotechnol 61: 197–205.

Verstrepen KJ et al. 2003b. Flavor-active esters: adding fruitiness tobeer. J Biosci Bioeng 96: 110–118.

Verstrepen KJ et al. 2003. Expression levels of the yeast alcoholacetyltransferase genes ATF1, Lg-ATF1, and ATF2 control theformation of a broad range of volatile esters. Appl Environ Mi-crobiol 69: 5228–5237.

Verstrepen KJ et al. 2004. Origins of variation in the fungal cellsurface. Nat Rev Microbiol 2: 533–540.

Vidgren V et al. 2005. Characterization and functional analysis ofthe MAL and MPH Loci for maltose utilization in some ale andlager yeast strains. Appl Environ Microbiol 71: 7846–7857.

Page 27: Food Biochemistry and Food Processing (Simpson/Food Biochemistry and Food Processing) || Biochemistry of Beer Fermentation

P1: SFK/UKS P2: SFK

BLBS102-c33 BLBS102-Simpson March 21, 2012 14:5 Trim: 276mm X 219mm Printer Name: Yet to Come

33 Biochemistry of Beer Fermentation 653

Vidgren V et al. 2009. Improved fermentation performance of a lageryeast after repair of its AGT1 maltose and maltotriose transportergenes. Appl Environ Microbiol 75: 2333–2345.

Villaneuba KD et al. 1990. Subthreshold vicinal diketone levelsin lager brewing yeast fermentations by means of ILV5 geneamplification. J Am Soc Brew Chem 48: 111–114.

Virkajarvi I, Kronlof J. 1998. Long-term stability of immobilizedyeast columns in primary fermentation. J Am Soc Brew Chem 56:70–75.

Virkajarvi I, Pohjala N. 2000. Primary fermentation with immobi-lized yeast: effects of carrier materials on the flavour of the beer.J Inst Brew 106: 311–318.

Virkajarvi I et al. 1999. Effects of aeration on flavor compoundsin immobilized primary fermentation. Monatschrift fur Brauwis-senschaft 52: 9–12, 25–28.

Virkajarvi I et al. 2002. Productivity of immobilized yeast reactorswith very-high-gravity worts. J Am Soc Brew Chem 60: 188–197.

Viyas VK et al. 2003. Snf1 kinases with different β-subunit isoformsplay distinct roles in regulating haploid invasive growth. Mol BiolCell 23: 1341–1348.

Vunjak-Novakovic G et al. 1992. Flow regimes and liquid mixingin a draft tube gas-liquid-solid fluidized bed. Chem Eng Sci 47:3451–3458.

Vunjak-Novakovic G et al. 1998. Air-lift reactors: research andpotential applications. Trends Chem Eng 5: 159–172.

Wackerbauer K et al. 2003. Measures to improve long term stabilityof main fermentation with immobilized yeast. Proceedings of theEuropean Brewery Convention, Dublin, pp. 445–457.

Wainwright T. 1973. Diacetyl—a review. J Inst Brew 79: 451–470.Walker GM. 1998. Yeast—Physiology and Biotechnology. John Wi-

ley & Sons, Chichester.Wang X et al. 2002. Intracellular maltose is sufficient to induce MAL

gene expression in Saccharomyces cerevisiae. Eukaryot Cell 1:696–703.

Wellhoener HJ. 1954. Ein Kontinuierliches Gar- und Reifungsver-fahren fur Bier. Brauwelt 94(1): 624–626.

White FH, Portno AD. 1979. The influence of wort composition onbeer ester levels. Proceedings of the European Brewery Conven-tion Congress, Berlin, pp. 447–460.

Willaert R. 2001. Sugar consumption kinetics by brewer’s yeastduring the primary beer fermentation. Cerevisia 26(1): 43–49.

Willaert R. 2007. Cell immobilization and its applications inbiotechnology: current trends and future prospects. In: EMT El-Mansi et al. (eds.) Fermentation Microbiology and Biotechnol-ogy. CRC Press, Boca Raton, FL, pp. 287–361.

Willaert R, Baron GV. 1996. Gel entrapment and micro-encapsulation: methods, applications and engineering principles.Rev Chem Eng 12: 1–205.

Willaert R et al. 2004. Diffusive mass transfer in immobilised cellsystems. In: VA Nedovic, R Willaert (eds.) Fundamentals ofCell Immobilisation Biotechnology. Kluwer Academic Publish-ers, Dordrecht, pp. 359–387.

Willaert R, Nedovic VA. 2006. Primary beer fermentation by immo-bilised yeast—a review on flavour formation and control strate-gies. J Chem Technol Biotechnol 81: 1353–1367.

Yamano S et al. 1994. Construction of a brewer’s yeast hav-ing α-acetolactate decarboxylase gene from Acetobacter acetiissp. xylinum integrated in the genome. J Biotechnol 32: 173–178.

Yamauchi Y, Kashihara T. 1995. Kirin immobilized system. EBCMonograph XXIV, EBC Symposium on Immobilized Yeast Ap-plications in the Brewing Industry, pp. 99–117.

Yamauchi Y et al. 1994. Beer brewing using an immobilized yeastbioreactor design of an immobilized yeast bioreactor for rapidbeer brewing system. J Ferment Bioeng 78: 443–449.

Yao B et al. 1994. Shared control of maltose induction and cataboliterepression of the MAL structural genes in Saccharomyces. MolGen Genet 243: 622–630.

Yoon SH et al. 2003. Specificity of yeast (Saccharomyces cerevisiae)in removing carbohydrates by fermentation. Carbohydr Res 338:1127–1132.

Younis OS, Stewart GG. 1998. Sugar uptake and subsequent es-ter and higher alcohol production by Saccharomyces cerevisiae.J Inst Brew 104: 255–264.

Younis OS, Stewart GG. 2000. The effect of wort maltose con-tent on volatile production and fermentation performance inbrewing yeast. In: K Smart (ed.) Brewing Yeast Fermentationand Performance, vol 1. Blackwell Science, Oxford, pp. 170–176.

Zahringer H et al. 2000. Induction of neutral trehalase Nth1 byheat and osmotic stress is controlled by STRE elements andMsn2/Msn4 transcription factors: variations of PKA effect duringstress and growth. Mol Microbiol 35: 397–406.

Zastrow CR et al. 2001. Maltotriose fermentations by Saccha-romyces cerevisiae. J Ind Microbiol Biotechnol 27: 34–38.

Zheng X et al. 1994a. Transport kinetics of maltotriose in strains ofSaccharomyces. J Ind Microbiol 13: 159–166.

Zheng X et al. 1994b. Factors influencing maltotriose utilizationduring brewery wort fermentations. J Am Soc Brew Chem 52:41–47.