article

The many styles of wine fermentation

Posted: 1 May 2014 | Linda F. Bisson, Professor & Geneticist, Department of Viticulture and Enology, University of California, Davis | No comments yet

Wine production is inherently a microbial process wherein components of the grape are transformed into flavour and aroma characteristics of the wine. If wine is a symphony of sensorial experiences then the grape precursors are the instruments and the microbes are the members of the orchestra. Wine fermentations may be inoculated with specific preparations of organisms or may be left to ferment by the indigenous populations of the vineyard and winery.

The many styles of wine fermentation

Microbial population dynamics can be influenced by seemingly inconsequential processing decisions and some microbial characters may be desired in some wines but considered spoilage in others. Fermentation management strategies must take into consideration the style of wine being made, the inherent risks of different processing decisions and seasonal or vintage variation in composition of both fruit and microbial biota.

A brief overview of the wine production process

The production of wine is a multistep yet simple process (Figure 1). The composition of the wine is influenced by the processing decisions made along the way to the bottle. The first important decision is the time of harvest. The timing of harvest not only impacts the composition of the grape influencing relative levels of sugars, amino acid content and types and amounts of flavour precursors, but also affects the relative numbers and species of microbiota that will be present on the grape and therefore present in the juice and ultimately in the wine itself. Different vineyards and sites within a vineyard will support differing populations of microbiota1 as will the presence of damage to the clusters or berries, the presence and types of insect vectors and physical factors such as temperature and exposure of the grape berries to UV light. These factors can be modified to an extent by vineyard management practices but some will be inherent to the site and type of grape varieties planted2.

Following harvest, grapes are crushed and the resulting must (juice, seeds and skins) can be used directly for fermentation as in red wine processing or the must pressed to release the juice from the skins and seeds which is used in white and blush wine production3. Grape skins contain the red pigments or anthocyanins that give red wines their characteristic colour and the fermentation on the skins releases these important pigmented compounds into the juice so that they will be present in the wine. Grapes are not washed post-harvest nor is there any heat or pasteurisation treatment except in special cases so materials other than grapes, such as soil, leaves and insects may also be a source of microorganisms3.

In red wines the must, and in white or blush wines the juice, is then transferred to a tank for fermentation. Fermentations may be inoculated with commercial preparations of microorganisms or left to ferment by the organisms coming from the vineyard or present in the winery. Both vineyards and wineries develop distinctive populations of microorganisms as has been recently revealed by next generation sequencing of vineyard, grape and winery surfaces4,5 and these differences can persist throughout the fermentation3. The primary or alcoholic fermentation is considered to be complete once sugar levels drop to a low to undetectable level, generally less than 2g/L from a starting concentration of 240g/L, an equimolar mixture of glucose and fructose3. The juice must be fermented in wood, stainless steel, plastic bins or concrete vats with or without temperature control. In red wine production, the carbon dioxide produced as a consequence of sugar catabolism via glycolysis becomes entrapped in the skins and as a consequence, the skins form a floating cap on the surface of the wine. The cap must be periodically bathed with fermenting wine or submerged and broken up into the wine to enable efficient extraction and to avoid the build-up of microbial populations on the upper surface of the cap. Cap management becomes important not only for colour extraction but for management of microbial populations3. Thus multiple phases present in red wine fermentations: the liquid phase of the juice, the solid phase of the grape skins and seeds, the surfaces of the containers that can house development of biofilms, and the exposure to air on the surface of the tank. Fermentations of red wines are conducted at temperatures ranging from 18 to 32°C, depending upon the variety of grape and style of wine desired3. Once fermentation is complete, the red wine is pressed off of the skins and seeds and subjected to the aging program of the winery.

White wine fermentations are restricted to the juice extracted from the grapes immediately upon crushing and pressing. White juices often undergo the process of ‘cold settling’ in which the juice is held at a low temperature (4 to 10°C) for 24 to 48 hours to allow the solids from the grape berries to settle to the bottom of the tank3. The juice is then separated from the solids and fermented at a warmer temperature, typically 12 to 16°C. The white fermentations are conducted at temperatures lower than those of red fermentations in order to retain volatile flavour components. In some situations, red musts are also held at low temperatures in a process referred to as a ‘cold soak’ to extract the readily soluble anthocyanins and other components6. This low temperature incubation of juices and musts allow the development of cyrotolerant microbial species early in the production of wine7-9. Changes in processing temperature impose a strong selection on resident microbial populations.

Following fermentation, wines are then aged prior to bottling. This aging period allows for important chemical transformations of the wine to occur that leads to flavour modification and integration. A second microbial fermentation may occur at this time, characterised by the conversion of the grape acid malate to lactate10. Members of the lactic acid bacteria conduct this conversion that is colloquially called the malolactic fermentation. As with the primary or alcoholic fermentation, the malolactic fermentation may be induced by addition of commercial preparations of bacteria, by using wine already undergoing active malolactic fermentation as an inoculum, or by simply relying on the lactic acid bacteria indigenous to the winery to conduct the fermentation. In some cases, if conditions are favourable, the malolactic fermentation may initiate prior to the initiation of the yeast fermentation.

It is important to emphasise that even today wine production is not a sterile process. It takes advantage of the ability of Saccharomyces to dominate mixed-culture fermentation and produce sufficient ethanol to inhibit other microbes. Similarly, the malolactic fermentation generally relies on the activity of a single lactic acid bacterium, Oenococcus oeni, for the evolution of desired bacterial flavours and for the consumption of nutrients that would otherwise be available for spoilage organisms10,11.

Fermentation processes

With respect to the primary fermentation, there are two classes of processes: inoculated and autochthonous. Autochthonous or native fermentations are conducted by the microbiota present on the grapes and in the winery. In either case, the fermentation is eventually dominated by the yeast Saccharomyces. Two species of Saccharomyces, S. cerevisiae and S. bayanus, are commonly found in wine fermentations. Many commercial strains and native isolates are hybrids of these and other species of Saccharomyces12-18.

Inoculated fermentations

Fermentations may be inoculated in one of two ways. The most common is to use a commercial active dry preparation of a yeast strain. There are over 100 commercial yeast strains available that differ in key oenological properties or that are targeted for specific fermentation conditions such as high ethanol tolerance or low nitrogen or micronutrient requirements. These strains also differ in aromatic profiles, fermentation efficiency, tolerances to temperature, and in ability to dominate the microbial profile of the wine. Some of the commercial yeasts have been selected for specific varieties of grapes or of style and produce more or less of the enzymes that impact the conversion of grape precursors to wine aromatic compounds.

Fermentations may also be inoculated using an existing fermentation. This process is also referred to as pitching and levels of inoculation are generally one to two per cent of the tank volume. In both types of inoculated fermentations, sulphur dioxide may be used to inhibit more sensitive microbes enabling a swifter domination of the fermentation by the inoculant. The nitrogen levels of the must are routinely monitored and adjusted to assure a complete fermentation and to avoid a slow or arrested fermentation19.

Autochthonous fermentations

In contrast, autochthonous or native fermentations are conducted by the organisms originally present in the juice or that gain access to the juice by passage through winery surfaces. In some cases, these yeasts may originate in the vineyard but in other cases the fermentations are conducted by winery residents2. There are different production styles for native fermentations. A true autochthonous fermentation receives no addition of microbes and no other adjustments (nutrients, pH or sulphur dioxide) so that there is minimal manipulation of the existing population dynamics. These types of fermentations can be at high risk for the development of off-odours or for slow or incomplete fermentations depending upon the microbes present and the innate nutrient composition of the fruit3. Manipulated autochthonous fermentations are those in which the conditions are adjusted to favour desired populations and to assure fermentation completion. The pH may be adjusted to reduce the activity of the bacterial species present. Most of the lactic acid bacteria are unable to grow below pH 3.5 and wine pH can vary from as low as pH 3.2 to as high as pH 4.2. Adjusting a juice to a pH below 3.5 greatly restricts the numbers of bacteria able to proliferate3. Similarly bacteria and non-Saccharomyces yeasts are more intolerant of sulphur dioxide than Saccharomyces is with the bacteria being most sensitive. Additions of sulphur dioxide can therefore also be used to adjust the microbial populations of the must. Nutrient additions can encourage the growth of the major populations present so can be timed to benefit a desired population by holding off the addition until Saccharomyces has become dominant.

Some autochthonous fermentations are more correctly termed partial native fermentations. In this case there is an inoculation with Saccharomyces at some point in the fermentation after the non-Saccharomyces populations have been able to impact wine aroma. Partial native fermentations are generally employed when there is a greater risk of arrest of fermentation or of the production of off-characters. The microbial populations present during autochthonous fermentations will be influenced by the vintage, winery sanitation practices, the variety of grape, and the adjacencies of the vineyard to other types of commercial operations.

Microbial contribution to wine aroma

The flavour and aroma compounds of a wine can derive from the grape itself, from microbial activity, from oxygen-dependent and independent chemical reactions that occur during aging or from extraction of cooperage such as oak during fermentation or aging. There are several classes of volatile compounds found in wine: alcohols, aliphatics, benzene derivatives, carbonyls, esters, organic acids, S-containing compounds, shikimic acid derivatives, polyols, isoprenoid-derived terpenes and C13 nor-isoprenoids, methoxypyrazines, vinyl and volatile phenols20-35. There are multiple ways in which the biological activity of Saccharomyces can impact wine flavour and aroma. Yeast may synthesise aroma-active compounds de novo or convert grape components into volatile odour-active forms. Yeast cellular components such as breakdown products of cell wall mannoproteins can modify wine flavour36,37. Metabolites produced by yeast can partake in the chemical reactions that occur in wine during aging38-41. Such modification may either create novel flavour-active components or lead to the loss of flavours derived from other sources47.

Saccharomyces is capable of producing a wide array of aromatic compounds that have been detected in wine22,32-34,37,42-76.The major classes of compounds formed are acids, alcohols, carbonyl compounds, esters and S-containing volatiles (Table 1). These compounds are often found in levels above the threshold of detection47. Table 1 also includes the common descriptive terms given to these compounds. Many are classified simply as ‘fruity’, meaning there is a perception of fruit character but not of a distinctive type of fruit. The fruity yeast esters serve an important role in amplifying the perception of the fruit identifier compounds of the specific wine variety. Non-Saccharomyces yeast can also contribute aroma compounds to the wine either via direct synthesis or due to modification of grape components.

Conclusion

The production of wine with a desired aroma and flavour profile hinges on the ability of the winemaker to manage the presence and metabolic activities of the organisms present during fermentation and aging of the wine. Yeast metabolites can enhance, modify or diminish wine aroma. Although Saccharomyces is the principle architect of wine production, numerous other microorganisms contribute to the final aromatic and flavour profile of the wine27. The goal of fermentation management strategies is to guide the process of microbial activity towards that of flawlessness, in short to act as maestro to achieve the desired masterpiece of flavour and aroma.

References

  1. Schuller D, Cardoso F, Sousa S, Gomes P, Gomes AC, Santos MAS, Casal M. Genetic diversity and population structure of Saccharomyces cerevisiae strains isolated from different grape varieties and winemaking regions PLOSone. 2012; 7(2):e32507
  2. Tello J, Cordero-Bueso G, Aporta I, Cabellos JM. Arroyo T. Genetic diversity in commercial wineries: effect of the farming system and vinification management on wine yeasts. J Appl Microbiol. 2011; 112:302-15
  3. Boulton RB, Singleton VL, Bisson LF, Kunkee RE. Principles and Practices of Winemaking. New York: Chapman & Hall; 1996
  4. Bokulich N, Ohta M, Richardson PM, Mills DA. Monitoring seasonal changes in winery resident microbiota. PLOSone 2013; 8(6) e66437
  5. Bokulich NA, Thorngate JH, Richardson PM, Mills DA. Microbial biogeography of wine grapes is conditioned by cultivar, vintage and climate. Proc Natl Acad Sci USA. 2014; 111:e139-e148
  6. Sacchi K L, Bisson L F. Adams DO. A review of the effect of winemaking techniques on phenolic extraction in red wines. Am J Enol Vitic. 2005; 56:197-206
  7. Fleet GH. Wine yeasts for the future. FEMS Yeast Res. 2008; 8:979-95
  8. Hierro N, Gonzalez A, Mas A, Guillamon JM. Diversity and evolution of non-Saccharomyces yeast populations during wine fermentation: effect of grape ripeness and cold maceration. FEMS Yeast Res. 2006; 6;102-11
  9. Zott K, Miot-Sertier C, Claisse O, Lonvaud-Funel A. Masneuf-Pomarede I. Dynamics and diversity of non-Saccharomyces yeasts during the early stages in winemaking. Int J Food Microbiol. 2008; 125:197-203
  10. Osborne JP, Edwards CG. Bacteria in winemaking. Adv Food Nutr Res. 2005; 50:139-77
  11. Bartowsky E, Henschke PA. Acetic acid bacteria spoilage of bottled wine – a review. Int J Food Microbiol. 2008; 125:60-70.
  12. Belloch C, Perez-Torrado R, Gonzalez SS, Perez-Ortin JE, Garcia-Martinez J, Querol A, Barrio E. Chimeric genomes of natural hybrids of Saccharomyces cerevisiae and Saccharomyces kudriavzevii, Appl Environ Microbiol. 2009; 75: 2534-2544
  13. Bisson LF. Geographic origin and diversity of wine strains of Am J Enol Vitic. 2012; 63:165-176
  14. Dequin S, Casaregola S. The genomes of fermentative Saccharomyces. C R Biologies. 2011; doi: 10.1016/j.crvi.2011.05.019
  15. González SS, Barrio E, Gafner J, Querol A. Natural hybrids from Saccharomyces cerevisiae, Saccharomyces bayanus and Saccharomyces kudriavzevii in wine fermentations. FEMS Yeast Res. 2006; 6:1221-34
  16. Gonzalez SS, Barrio E, Querol A. Molecular characterization of new natural hybrids of Saccharomyces cerevisiae and kudriavzevii in brewing. Appl Envion Microbiol. 2008; 74;2314-20
  17. Masneuf I, Hansen J, Groth C, Piskur J, Dubourdieu D. New hybrids between Saccharomyces sensu stricto yeast species found among wine and cider production strains. Appl Environ Microbiol. 1998; 64:3887-92
  18. Naumova, E. S., Naumov, G. I., Masneuf-Pomarede, I., Aigle, M. and Dubourdieu, D. Molecular genetic study of introgression between Saccharomyces bayanus and S. cerevisiae. Yeast. 2005; 22:1099-115
  19. Bisson LF. Stuck and sluggish fermentations. Am J Enol Vitic. 1999; 50:107-119
  20. Baumes R. Wine aroma precursors. In: Moreno-Arribas MV, Polo MC, editors. Wine Chemistry and Biochemistry. New York. Springer Science + Business Media; 2009
  21. Ferreira V, Lopez R, Escudero A, Cacho JF. The aroma of Grenache red wine: hierarchy and nature of its main odorants. J Sci Food Agric. 1998; 77:259-67
  22. Ferreira V, Cacho J. Identification of impact odorants of wines. In: Moreno-Arribas MV, Polo MC, editors. Wine Chemistry and Biochemistry. New York. Springer Science + Business Media; 2009
  23. Guth H. Identification of character impact odorants of different white wine varieties. J Agric Food Chem. 1997; 45:3022-26
  24. Guth H. Quantitation and sensory studies of character impact odorants of different white wine varieties. J Agric Food Chem. 1997; 45:3027-32
  25. Lee SH, Seo M, Riu M, Cotta JP, Block DE, et al. Vine microclimate and norisoprenoid concentration in Cabernet Sauvignon grapes and wines. Am J Enol Vitic. 2007; 58:291-301
  26. Mendes-Pinto MM. Carotenoid breakdown products the – norisoprenoids – in wine aroma. Arch Biochem Biophys. 2009; 483:236-45
  27. Heard GM, Fleet GH. Growth of natural yeast flora during the fermentation of inoculated wines. Appl Environ Microbiol. 1985; 50:727-28
  28. Polaskova P, Herszag J, Ebeler E. Wine flavour: chemistry in a glass. Chem Soc Rev. 2008; 37:2478- 89
  29. Preston LD, Block DE, Heymann H, Soleas G, Noble AC, et al. Defining vegetal aromas in Cabernet Sauvignon using sensory and chemical evaluations. Am J Enol Vitic. 2008; 59:137-45
  30. Rapp A, Mandery H. Wine aroma. Experientia 1986; 42:873-84
  31. Tominaga T, Peyrot des Gachons C, Dubourdieu D. A new type of flavour precursors in Vitis vinifera L. cv. Sauvignon Blanc: S-cysteine conjugates. J Agric Food Chem. 1998; 46:5215-19
  32. Tominaga T, Baltenweck Guyot R, Peyrot des Gachons C, Dubourdieu D. Contribution of volatile thiols to the aromas of white wines made from several Vitis vinifera grape varieties. Am J Enol Vitic. 2000; 51:178-81
  33. Tominaga T, Blanchard L, Darriet P and Dubourdieu D. A powerful aromatic volatile thiol, 2-furanmethanethiol, exhibiting roast coffee aroma in wines made from several Vitis vinifera grape varieties. J Agric Food Chem. 2009; 48:1799-802
  34. Vilanova M, Cortes S, Santiago JL, Martinez C, Fernandez E. Aromatic compounds in wines produced during fermentation: effect of three red cultivars. Int J Food Prop. 2007; 10:867-75
  35. Zhang M, Xu Q, Duan C, Qu W, Wu Y. Comparative study of aromatic compounds in young red wines from Cabernet Sauvignon, Cabernet Franc and Cabernet Gernischet varieties in China. J Food Sci. 2007; 72:248-52
  36. Pozo-Bayon MA, Reineccius G. Interactions between wine matrix macro-components and aroma compounds. In: Moreno-Arribas MV, Polo MC, editors. Wine Chemistry and Biochemistry. New York. Springer Science + Business Media; 2009
  37. Ugliano M, Henschke P. Yeasts and Wine Flavour. In: Moreno-Arribas MV, Polo MC. Editors. Wine Chemistry and Biochemistry. New York. Springer Science + Business Media; 2009
  38. Cullere L, Cacho J, Ferreira V. An assessment of the role played by some oxidation-related aldehydes in wine aroma. J Agric Food Chem. 2007; 55:876-81
  39. Cutzach I, Chatonnet P, Dubourdieu D. Influence of storage conditions on the formation of some volatile compounds in white fortified wines (Vins doux naturels) during the aging process. J Agric Food Chem. 2009; 48:2340-45
  40. Jarauta I, Cacho J, Ferreira V. Concurrent phenomena contributing to the formation of the aroma of wine during aging in oak wood: an analytical study. J Agric Food Chem. 2005; 53:4166-77
  41. Spillman PJ, Pollnitz AP, Liacopoulos D, Pardon KH, Sefton MA. Formation and degradation of furfuryl alcohol, 5-methylfurfurly alcohol, vanillyl alcohol and their ethyl esters in barrel-aged wines. J Agric Food Chem. 1998; 46:657-63
  42. Bardi L, Crivelli C, Marzona M. Esterase activity and release of ethyl esters of medium-chain fatty acids by Saccharomyces cerevisiae during anaerobic growth. Can J Microbiol. 1998; 44:1171-76
  43. Bartowsky EJ, Pretorius IS. Microbial formation and modification of flavour and off-flavour compounds in wine. In: Konig H, Unden G, Frohlich J, editors. Biology of Microorganisms on Grapes, in Musts and in Wine. Berlin Heidelberg. Springer-Verlag; 2009
  44. Dickinson JR, Eshantha L, Salgado J, Hewlins MJE. The catabolism of amino acids to long chain and complex alcohols in Saccharomyces cerevisiae. J Biol Chem. 2003; 278:8028-34
  45. Dubourdieu D, Tominaga T, Masneuf I, Peyrot des Gachons C, Murat ML. The role of yeasts in grape flavour development during fermentation: the example of Sauvignon blanc. Am J Enol Vitic. 2006; 57:81-8
  46. Fernandez-Gonzalez M, Di Stefano R. Fractionation of glycoside aroma precursors in neutral grapes. Hydrolysis and conversion by Saccharomyces cerevisiae. Lebensm-Wiss. U. Technol. 2003; 37:467-73
  47. Ferreira V, Lopez R, Aznar M. Olfactometry and aroma extract dilution analysis of wines. In: Jackson JF, Linskens HF, editors. Analysis of Taste and Aroma. Berlin Heidelberg. Springer-Verlag; 2009
  48. Hazelwood LA, Daran JM, van Maris AJA, Pronk JT, Dickinson JR. The Ehrlich pathway for fusel alcohol production: a century of research on Saccharomyces cerevisiae Appl Environ Microbiol. 2008; 74:2259-66
  49. Howell KS, Klein M, Swiegers JH, Hayasaki Y, Elsey GM, et al. Genetic determinants of volatile thiol release by Saccharomyces cerevisiae during wine fermentation. Appl Environ Microbiol. 2005; 71:5420-26
  50. Howell KS, Cozzolino, Bartowsky EJ, Fleet GH, Henschke PA. Metabolic profiling as a tool for revealing Saccharomyces interactions during wine fermentation. FEMS Yeast Res. 2006; 6:91-101
  51. Kotseridis Y, Baumes R. Identification of impact odorants in Bordeaux red grape juice, in the commercial yeast used for its fermentation, and in the produced wine. J Agric Food Chem. 2000; 48:400-6
  52. Landaud S, Helinck S, Bonnarme P. Formation of volatile sulfur compounds and metabolism of methionine and other sulfur compounds in fermented foods. Appl Microbiol Biotechnol. 2008; 77:1191-205
  53. Marchand S, de Revel G, Bertrand A. Approaches to wine aroma: Release of aroma compounds from reactions between cysteine and carbonyl compounds in wine. J Agric Food Chem. 2000; 48:4890-95
  54. Mason AB, Dufour JP. Alcohol acetyltransferases and the significance of ester synthesis in yeast. Yeast. 2000; 16:1287-98
  55. Miller AC, Wolff SR, Bisson LF, Ebeler SE. Yeast strain and nitrogen supplementation: dynamics of volatile ester formation in Chardonnay juice fermentations. Am J Enol Vitic. 2007; 58:470-83
  56. Molina AM, Swiegers JH, Varela C, Pretorius IS, Agosin E. Influence of wine fermentation temperature on the synthesis of yeast-derived volatile aroma compounds. Appl Microbiol Biotechnol. 2007; 77:675-87
  57. Moreira N, Mendes F, Pereira O, Guedes de Pinho P, Hogg T, et al. Volatile sulphur compounds in wines related to yeast metabolism and nitrogen composition of grape musts. Anal Chim Acta 2002; 458:157-67
  58. Murat ML, Masneuf I, Darriet P, Lavigne V, Tominaga T, et al. Effect of Saccharomyces cerevisiae yeast strains on the liberation of volatile thiols in Sauvignon blanc wine. Am J Enol Vitic. 2001; 52:136-9
  59. Nykanen L. Formation and occurrence of flavour compounds in wine and distilled beverages. Am J Enol Vitic. 1986; 37:84-96
  60. Parkkinen E, Suomalainen H. Esterases of Baker’s yeast. II. Substrate specificities towards esters formed during sugar fermentations. J Inst Brew. 1982; 88:34-8
  61. Rauhut D. Yeasts – Production of sulfur compounds. In: Fleet GH, editor. Wine Microbiology and Biotechnology. Chur. Harwood Academic Publishsers GmbH; 1993
  62. Rauhut D. Usage and formation of sulphur compounds. In: Konig H, Unden G, Frohlich J, editors. Biology of Microorganisms on Grapes, in Musts and in Wine. Berlin Heidelberg. Springer-Verlag; 2009
  63. Romano P, Fiore C, Paraggio M, Caruso M, Capece A. Function of yeast species and strains in wine flavour. Int J Food Microbiol. 2003; 86:169-80
  64. Rossouw D, Naes T, Bauer F. Linking gene regulation and the exo-metabolome: A comparative transcriptomics approach to identify genes that impact the production of volatile aroma compounds in yeast. BMC Genomics. 2008; doi:10.1186/1471-2164-9-530
  65. Saerens SM, Delvaux FR, Verstrepen KJ, Thevelein JM. Production and biological function of volatile esters in Saccharomyces cerevisiae. Microbial Biotech. 2009; doi1111/j.1751-7915.2009.00106
  66. Schoondermark-Stolk SA, Jansen M, Verkleij AJ, Verrips CT, Euvernink GJW, et al. Genome-wide transcription survey on flavour production in Saccharomyces cerevisiae. World J Microbiol Biotechnol. 2006; 22:1347-56
  67. Schoondermark-Stolk SA, Jansen M, Veurink JH, Verkleij AJ, Verrips CT, et al. Rapid identification of target genes for 3-methyl-1-butanol production in Saccharomyces cerevisiae. Appl Microbiol Biotechnol. 2006; 70:237-46.
  68. Spiropoulos A, Tanaka J, Flerianos I, Bisson LF. Characterization of hydrogen sulfide formation in commercial and natural wine isolates of Am J Enol Vitic. 2000; 51:233-48
  69. Suomalainen H. Yeast esterases and aroma esters in alcoholic beverages. J Inst Brew. 1981; 87:296-300
  70. Swiegers JH, Bartowsky EJ, Henschke PA, Pretorius IS. Yeast and bacteria modulation of wine aroma and flavour. Aus J Grape Wine Res. 2005; 11:139-73
  71. Swiegers JH, Pretorius IS. Modulation of volatile sulfur compounds by wine yeast. Appl Microbiol Biotechnol. 2007; 74:954-60
  72. Thorngate JH. Yeast strain and wine flavour: Nature or nurture? In: Waterhouse AL, Ebeler SE, editors. Chemistry of Wine Flavour. Washington DC. American Chemical Society; 1998
  73. Thurston PA, Quain DE, Tibb RS. Lipid metabolism and the regulation of volatile ester synthesis in Saccharomyces cerevisiae. J Inst Brew 1982; 88:90-4
  74. Ugliano M, Bartowsky EJ, McCarthy J, Moio L, Henschke PA. Hydrolysis and transformation of grape glycosidically bound volatile compounds during fermentation with three Saccharomyces yeast strains. J Agric Food Chem. 2006; 54:6322-31
  75. Verstrepen KJ, Van Laere SDM, Vanderhaegen BMP, Derdelinckx G, Dufour JP, et al. Expression levels of the yeast alcohol acetyltransferase genes ATF1, LgATF1, and ATF2 control the formation of a broad range of volatile esters. Appl Environ Microbiol. 2003; 69:5228-37
  76. Yoshimoto H, Fukushige T, Yonezawa T, Sone H. Genetical and physiological analysis of branched-chain alcohols and isoamyl acetate production in Saccharomyces cerevisiae. Appl Microbiol Biotechnol. 2002; 59:501-8

About the author

Linda Bisson received her BA and MA degrees in Biology and Microbiology respectively from San Francisco State University. She obtained her PhD from UC Berkeley and conducted postdoctoral research at Harvard Medical School. In 1985 she joined the faculty in the Department of Viticulture and Enology at UC Davis.

Related organisations