effects of sugar concentration and yeast …

137
EFFECTS OF SUGAR CONCENTRATION AND YEAST INOCULATION STRATEGY ON MANGO WINE FERMENTATION CHAN LI JIE (B. Sc. National University of Singapore) A THESIS SUBMITTED FOR THE DEGREE OF MASTER OF SCIENCE DEPARTMENT OF CHEMISTRY NATIONAL UNIVERSITY OF SINGAPORE 2014 brought to you by CORE View metadata, citation and similar papers at core.ac.uk provided by ScholarBank@NUS

Upload: others

Post on 24-May-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

EFFECTS OF SUGAR CONCENTRATION AND

YEAST INOCULATION STRATEGY ON MANGO

WINE FERMENTATION

CHAN LI JIE

(B. Sc. National University of Singapore)

A THESIS SUBMITTED

FOR THE DEGREE OF MASTER OF SCIENCE

DEPARTMENT OF CHEMISTRY

NATIONAL UNIVERSITY OF SINGAPORE

2014

brought to you by COREView metadata, citation and similar papers at core.ac.uk

provided by ScholarBank@NUS

Page 2: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

i

DECLARATION

I hereby declare that the thesis is my original work and it has been written by me in its

entirety, under the supervision of Dr Liu Shao Quan, (in the Food Science and Technology

research laboratory, S13-05), Chemistry Department, National University of Singapore,

between July 2010 and June 2014.

I have duly acknowledged all the sources of information which have been used in the

thesis.

This thesis has not been submitted for any degree in any university previously.

The content of the thesis has been partly published in:

Chan, L.J., Lee, P.R., Li, X., Chen, D., Liu, S.Q., and Trinh, T.T.T. (2012). Tropical fruit

wine: an untapped opportunity. In: D. Cabel (Ed.), Food and Beverage Asia, Dec/Jan

2011/2012 (pp. 48–51). Singapore: Pablo Publishing Pte Ltd. ISSN: 2010-2364.

http://www.foodbeverageasia.com/ebook/FBA_DecJan2012/index.html.

Li, X., Chan, L.J., Yu, B., Curran, P., and Liu, S.-Q. (2014) Influence of Saccharomyces

cerevisiae and Williopsis saturnus var. mrakii on mango wine characteristics. Acta

Alimentaria. 43(3), 473-481.

Chan Li Jie 20 June 2014

Name Signature Date

Page 3: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

ii

ACKNOWLEDGEMENTS

I would like to express my heartfelt gratitude to my supervisor, Dr Liu Shao Quan for

his unrelenting patience, support and assistance in this project. I have been very, very

fortunate to have such an understanding supervisor.

Next, I would like to thank the Food Science and Technology program for all the

assistance they have rendered. I would also like to thank the flavorists at Firmenich Asia for

assisting with the sensory evaluation despite their busy schedule.

In addition, I would like to thank my research group members, Dr Cheong Mun Wai,

Dr Lee Pin Rou, Dr Li Xiao, Dr Sun Jingcan and Ms Chen Dai for their support, concern and

presence in my academic and personal life. They truly have made my days in the lab

brighter.

Last but not least, I would like to thank my family for their unwavering support and

faith in me. Without them, I wouldn’t have been the person I am.

Page 4: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

iii

Table of Contents

DECLARATION……………………………………………………………………….. i

ACKNOWLEDGEMENTS……………………………………………………………. ii

Table of Contents……………………………………………………………………….iii

List of Tables…………………………………………………………………...………vi

List of Figures…………………………………………………………………..………ix

Chapter 1 Introduction and Literature Review………………………………………....1

1.1 Mango fruit ......................................................................................................... 1

1.1.1 Nutritional content ........................................................................................ 1

1.1.2 Volatile compounds and mango flavour ....................................................... 2

1.2 Mango wine ........................................................................................................ 3

1.3 Biochemistry of fermentation, wine flavour and quality ................................... 6

1.4 Influence of fermentation conditions and yeast strains ...................................... 8

1.4.1 Amelioration of must .................................................................................... 8

1.4.2 Saccharomyces in wine production .............................................................. 9

1.4.3 Non-Saccharomyces species in wine production ....................................... 10

1.4.4 Inoculation strategies in wine fermentation ................................................ 10

1.5 Research aims and objectives ........................................................................... 13

Chapter 2 Materials and Methods…………………………………………………….15

2.1 Mango fruits and preparation of mango juice .................................................. 15

2.2 Yeast and culture media ................................................................................... 16

2.3 Preparation of yeast starter culture ................................................................... 16

Page 5: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

iv

2.4 Fermentation of mango juice with different initial sugar concentrations

— Chapter 3 .................................................................................................... 17

2.5 Fermentation of mango juice with co-inoculated S. cerevisiae and W. saturnus

— Chapter 4 .................................................................................................... 17

2.6 Fermentation of mango juice with different sequential inoculation strategies

— Chapter 5 .................................................................................................... 18

2.7 Analytical methods ........................................................................................... 19

2.7.1 pH, oBrix and yeast enumeration ................................................................ 19

2.7.2 Analysis of sugars and organic acids ......................................................... 19

2.7.3 Analysis of volatile compounds ................................................................. 20

2.8 Sensory analysis ............................................................................................... 22

2.9 Statistical analysis ............................................................................................ 23

Chapter 3 Effects of Sugar Concentration on Volatile Production by Saccharomyces

cerevisiae MERIT.ferm………………………………...…………….…….24

3.1 Introduction ........................................................................................................... 24

3. 2 Results and discussion ...................................................................................... 25

3.2.1 Mango juice volatile composition .............................................................. 25

3.2.2 Changes in pH and organic acids ............................................................... 31

3.2.3 Yeast growth, total soluble solids and sugar concentration ........................ 32

3.2.4 Glycerol production ................................................................................... 36

3.2.5 Effects of initial sugar concentration on volatile production ..................... 39

3.2.6 Effects of redox potential on overall wine quality ..................................... 52

3.3 Conclusion ............................................................................................................ 53

Chapter 4 Effects of Co-fermentation of Saccharomyces cerevisiae and Williopsis

saturnus Yeasts on Volatile Production…………………………………...55

4.1 Introduction ...................................................................................................... 55

4.2 Results and discussion ...................................................................................... 56

Page 6: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

v

4.2.1 Total soluble solids, sugar concentrations and yeast cell count ................. 56

4.2.2 pH and organic acids .................................................................................. 59

4.2.3 Volatile composition of mango wines ........................................................ 59

4.2.4 Miscellaneous compounds........................................................................ 77

4.3 Conclusion ........................................................................................................ 78

Chapter 5 Effects of Different Sequential Inoculation Strategies of Saccharomyces

cerevisiae and Williopsis saturnus on Volatile Production……………….79

5.1 Introduction ..................................................................................................... 79

5.2 Results and discussion ...................................................................................... 80

5.2.1 Physicochemical properties of mango wine ............................................... 80

5.2.2 Yeast biomass evolution ............................................................................ 82

5.2.3 Volatile composition................................................................................... 86

5.2.4 Sensory evaluation .................................................................................... 108

5.3 Conclusion ............................................................................................... 109

Chapter 6 General conclusions and recommendations……………………………...110

Bibliography………………………………………………………………………….115

Page 7: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

vi

SUMMARY

The project first investigated the effects of initial sugar concentration on volatile and

glycerol production by Saccharomyces cerevisiae MERIT.ferm in mango wine fermentation.

Generally, high sugar concentration had a negative impact on volatile production but

enhanced glycerol production. Significantly lower amounts of esters and higher alcohols and

more acetic acid and acetaldehyde were produced in high sugar fermentation.

The effects of a non-Saccharomyces yeast, Williopsis saturnus var. mrakii NCYC 500

and a mixed culture fermentation consisting of S. cerevisiae and W. mrakii in the ratio of

1:1000 were then studied in mango wine fermentation. The volatile profile of the mango

wine produced by the mixed culture fermentation resembled that of the one fermented with a

monoculture of S. cerevisiae, while the mango wine produced by fermentation with W. mrakii

differed significantly from the wine fermented with a monoculture of S. cerevisiae and the

mixed culture fermentation. W. mrakii produced higher amounts of acetate esters but

significantly lower amounts of ethyl esters and fusel alcohols. This is highly likely due to the

metabolic activities of the different yeast strains.

The effects of varying the sequence of inoculation of S. cerevisiae and W. mrakii were

subsequently investigated. Simultaneous mixed culture fermentation (MCF) was conducted

by co-inoculation of S. cerevisiae and W. mrakii in the ratio of 1:1000. Negative sequential

fermentation (NSF) was conducted by first inoculating S. cerevisiae and allowing the

fermentation to proceed for 7 days before inactivating the S. cerevisiae by ultrasonication;

then inoculating W. mrakii and allowing the second phase of the fermentation to proceed for

14 days. For positive sequential fermentation (PSF), the sequence of inoculation was

reversed.

Page 8: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

vii

The volatile profiles of the mango wines that resulted from the three different

sequential inoculation strategies varied significantly. NSF generally produced the least

amount of volatile, especially the higher alcohols and esters while PSF produced more of the

desirable volatile ester compounds. MCF typically produced levels of volatiles between NSF

and PSF.

The findings obtained in this study potentially could have an impact on fruit wine

production and allow wine makers to design an inoculation strategy that would cater to the

desired wine style.

Page 9: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

viii

List of Tables

Table 3.1 Volatile compounds in fresh Chok Anan mango juice analysed using HS-SPME-

GC-MS/FID .......................................................................................................... 28

Table 3.2 Physicochemical properties, organic acid and sugar concentrations of mango wine

before and after fermentation ................................................................................ 32

Table 3.3 Volatiles in mango juice catabolised during fermentation ...................................... 39

Table 3.4 Volatile composition of mango wines with different initial sugar concentrations . 40

Table 3.5 Selected volatiles quantified in mango wine and their odour activity values (OAV)

in low (unfortified mango juice 16.6oB), medium (initial TSS of 23°B) and high

(initial TSS of 30°B) sugar fermentation .............................................................. 44

Table 4.1 Physicochemical properties, reducing sugar and organic acid content before and

after fermentation .................................................................................................. 57

Table 4.2 Complete volatiles for mango wine fermented with yeasts S. cerevesiae

MERIT.ferm, W. mrakii NCYC 500 and mixed culture ....................................... 61

Table 4.3 Major volatile compounds of alcoholic fermentation by S. cerevisiae, W. mrakii

and a mixed culture of S. cerevisiae and W. mrakii .............................................. 65

Table 5.1 Physicochemical properties, yeast cell count, organic acid and sugar concentrations

of mango wines ..................................................................................................... 81

Table 5.2 Complete volatiles for mango wine produced from different inoculation strategies

............................................................................................................................... 87

Table 5.3 Major volatiles quantified for mango wines with different inoculation strategies..

............................................................................................................................... 92

Page 10: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

ix

List of Figures

Figure 3.1 Changes in yeast cell count and TSS (oBrix) for low ( ), medium ( ) and high ( )

sugar fermentation................................................................................................. 34

Figure 3.2 Fructose ( ), glucose ( ) and sucrose ( ) consumption kinetics of S.cerevisiae

during fermentation ............................................................................................... 36

Figure 3.3 Concentrations of glycerol in mango wines for low (16.6oBrix), medium (23

oBrix)

and high sugar (30oBrix) fermentation.................................................................. 38

Figure 3.4 Metabolic pathways governing the production of ethanol, glycerol and acetic

acid………………………………………………………………………………49

Figure 4.1 (a) Changes TSS (°B) during fermentation for S. cerevisiae MERIT.ferm

monoculture ( ), W. mrakii NCYC 500 monoculture ( )and mixed culture

fermentation ( ) (S. cerevisiae:W. mrakii at a ratio 1:1000) (b) Changes in yeast

population during fermentation for S. cerevisiae MERIT.ferm monoculture ( ),

W. mrakii NCYC 500 monoculture ( ), S. cerevisiae in mixed culture

fermentation ( )and W. mrakii in mixed culture fermentation ( ); S. cerevisiae:W.

mrakii at a ratio 1:1000 in mixed culture fermentation ........................................ 58

Figure.4.2 Changes in β-myrcene in the fermentation of mango juice with a monoculture of S.

cerevisiae MERIT.ferm ( ), monoculture of W. mrakii NCYC 500 ( ) and a

mixed culture of S. cerevisiae and W. mrakii ( ) ................................................. 67

Figure 4.3 Changes in ethanol content in the fermentation of mango juice with a monoculture

of S. cerevisiae MERIT.ferm ( ), monoculture of W. mrakii NCYC 500 ( ) and a

mixed culture of S. cerevisiae and W. mrakii ( ) ................................................. 68

Figure 4.4 Changes in isobutyl alcohol and 2-phenylethyl alcohol content in the fermentation

of mango juice with a monoculture of S. cerevisiae MERIT.ferm ( ),

monoculture of W. mrakii NCYC 500 ( ) and a mixed culture of S. cerevisiae and

W. mrakii ( ) ........................................................................................................ 70

Figure 4.5 Changes in ethyl octanoate in the fermentation of mango juice with a monoculture

of S. cerevisiae MERIT.ferm ( ), monoculture of W. mrakii NCYC 500 ( ) and a

mixed culture of S. cerevisiae and W. mrakii ( ) ................................................. 72

Figure 4.6 Changes in 2-phenylethyl acetate and ethyl acetate content in the fermentation of

mango juice with a monoculture of S. cerevisiae MERIT.ferm ( ), monoculture

of W. mrakii NCYC 500 ( ) and a mixed culture of S. cerevisiae and W. mrakii ( )

............................................................................................................................... 74

Figure 4.7 Changes in acetic acid content in the fermentation of mango juice with a

monoculture of S. cerevisiae MERIT.ferm ( ), monoculture of W. mrakii NCYC

500 ( ) and a mixed culture of S. cerevisiae and W. mrakii ( ) ........................... 76

Figure 5.1 Changes in oBrix values during fermentation for PSF

1 ( ), NSF

2 ( ) and MCF

3 ( )

............................................................................................................................... 82

Page 11: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

x

Figure 5.2 Changes in (a) S. cerevisiae MERIT.ferm for PSF1 ( ), NSF

2 ( ) and MCF

3 ( ), (b)

W. mrakii NCYC 500 cell population during fermentation for PSF1 ( ), NSF

2 ( )

and MCF3 ( ) (c) Changes in S. cerevisiae MERIT.ferm ( ) and W. mrakii

NCYC 500 ( ) in MCF......................................................................................... 84

Figure 5.3 Changes in ethanol concentration during PSF1

( ), NSF2 ( ) and MCF

3 ( ) ........ 95

Figure 5.4 Changes in isoamyl alcohol, 2-phenylethyl alcohol, isobutyl alcohol during PSF1

( ), NSF2 ( ) and MCF

3 ( ) .................................................................................. 97

Figure 5.5 Changes in linalool during PSF1 ( ), NSF

2 ( ) and MCF

3 ( ) .............................. 99

Figure 5.6 Changes in (a) ethyl octanoate, (b) ethyl decanoate, (c) ethyl acetate during PSF1

( ), NSF2 ( ) and MCF

3 ( ) ............................................................................... 102

Figure 5.7 Changes in 2-phenylethyl acetate and isoamyl acetate during during PSF1 ( ),

NSF2 ( ) and MCF

3 ( ) ....................................................................................... 103

Figure 5.8 Changes in acetic acid and hexanoic acid during PSF1 ( ), NSF

2 ( ) and MCF

3 ( )

............................................................................................................................. 106

Figure 5.9 Sensory profile of PSF1 ( ), NSF

2 ( ) and MCF

3 ( ), mango wine .................... 108

Page 12: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

1

Chapter 1 Introduction and Literature Review

1.1 Mango fruit

1.1.1 Nutritional content

The mango fruit (Mangifera indica L.) is one of the most popular and economically

important tropical fruits due to its exotic and appealing flavour and taste. Typically, the pulp

of the fruit is consumed, and the skin and seed discarded. The major constituents of the pulp

are water, carbohydrates, organic acids, fats, tannins, vitamins and flavor compounds (Sagar

et al. 1999); the chemical composition varies with location of cultivation, variety, and stage

of maturity (Chauhan et al. 2010).

In the Southeast Asian region, the Chok Anan mango is one of the most popular

cultivars consumed for its mild and pleasant flavour. Reported to have a total soluble solids

content of about 14 to16 °Brix (Vásquez-Caicedo et al. 2002), with sucrose as the

predominant sugar (approximately 7.5 g/100 g), followed by fructose and glucose at 5 g/100

g and 1.5 g/100 g respectively (Vásquez-Caicedo et al. 2002). Citric acid (0.32 g/100 g) and

malic acid (0.25 g/100 g) are the main organic acids; other organic acids include succinic,

oxalic, pyruvic, adipic, mucic, galacturonic and glucuronic acids (Tharanathan et al. 2006).

During the ripening of mango, both sugars and pH increase, due to glucogenesis and

hydrolysis of polysaccharides (starch), resulting in an increase in sweetness. This

accumulation of sugars and organic acids results in an excellent sugar/acid ratio that is

responsible for taste development (Chauhan et al. 2010).

The major amino acids reported in mangoes are alanine, arginine, glycine, serine,

leucine and isoleucine, with a protein content of approximately 0.8 g/100 g (Tandon and

Kalra 1984). Lipids constitute less than 1% (w/w), with triglyceride being the major

Page 13: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

2

component (Gholap and Bandyopadhyay 1975). Although mango is rich in vitamin C, the

maximum level occurs in the early stage of growth instead of ripening stage (Spencer et al.

1956). β-Carotene increases as the mango fruit matures and ripens (Jungalwala and Cama

1963); the increase in β-carotene correlates with a decrease in acids and an increase in sugar

content (Godoy and Rodriguez-Amaya 1987).

1.1.2 Volatile compounds and mango flavour

Although more than a hundred volatiles have been identified in the mango flavour

profile, terpene hydrocarbons (monoterpene and sesquiterpene) make up the predominant

class of volatiles in mango flavour across all varieties. Amongst the terpene hydrocarbons,

the dominant ones include δ-3-carene, α-pinene, α-phellandrene, α-terpinolene and β-

caryophyllene (Chauhan et al. 2010). In addition to these terpene hydrocarbons, many

varieties also have considerable amounts of oxygenated volatile compounds, including esters

(e.g. ethyl 2-methylpropanoate, ethyl butanoate, methyl benzoate), lactones (e.g. γ-

hexalactone), aldehydes (e.g. cis-2-nonenal, 2,6-nonadienal, decanal), furanones (e.g. 2,5-

dimethyl-4-methoxy-3(2H)-furanone), C13-norisoprenoids (e.g. β-ionone), etc (Pino et al.

2005; Pino and Mesa 2006). These compounds are produced through metabolic pathways

during ripening, harvesting and post-harvest storage, and depend on many factors related to

the species, variety and type of technological treatments (Léchaudel and Joas 2007). In

addition, the maturity of the fruit at harvest was also found to affect the overall flavour of

mangoes. Mangoes harvested at the green stage exhibited a higher amount of monoterpenes,

sesquiterpenes and aromatic compounds while fruits harvested at the fully ripe stage had

higher concentrations of esters, alkanes, and norisoprenoids (Lalele et al. 2003a).

Generally regarded as the terpinolene type, the Chok Anan mango is known for its pine

needle-like terpene note (Chauhan et al. 2010), likely due to the major volatiles present being

Page 14: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

3

δ-thujene, α-pinene, δ-3-carene, mycrene, α-phellandrene and α-terpinolene. α-Terpinolene,

the major compound, has been described as sweet, floral with pine-like aroma notes while δ-

3-carene, the second major monoterpene in Chok Anan mangoes has been described as sweet,

floral and mango leaf-like (Laohakunjit et al. 2005). These terpene notes are further

enhanced by the presence of compounds such as 3-hexanol, 2-hexanol, γ- and δ-lactones, and

furan compounds to give the overall flavour perception (Vásquez-Caicedo et al. 2002). In

addition, esters give rise to the overall fruity character while the overall floral top notes of the

mango flavour is derived from the alcoholic compounds such as linalool, 2-phenylethyl

alcohol, nerol and citronellol (Chauhan et al. 2010).

1.2 Mango wine

Despite its popularity, mango juice/pulp is not an optimum substrate for fruit wine

fermentation due to its relatively low sugar content, and organic acid and amino acid

composition. Wine grapes, the optimum substrate for wine fermentation, typically have a

sugar content of approximately 150 to 250 g/100 mL [17 to 26 % (w/v)], with tartaric and

malic acids as the main organic acids (Alexandre et al. 1994); glutamate, glutamine, alanine,

arginine and proline make up 90% of all amino acids present in wine grapes. Furthermore,

certain cultivars of mangoes also contain high levels of pectins which may lead to pectin haze,

resulting in undesirable cloudiness in wine (Vásquez-Caicedo et al. 2002). To overcome

these differences and limitations, various groups have conducted research to optimise the

production of a mango wine with quality comparable to conventional grape wine

After the first anonymous report in 1963, the technology for mango wine production

was tested on the Hilacha variety in 1966. The study concluded that mango was an excellent

raw material for production of good quality white, semi-sweet table wine (Czyhrinciwk 1966).

Page 15: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

4

Following which, several varieties of mangoes were screened for their suitability in wine

production with the conclusion that mango wine had rather high acceptability (Kulkarni,

Singh and Chadha 1980; Onkarayya and Singh 1984; Onkarayya 1986). Another study

utilised Saccharomyces cerevisiae (S. cerevisiae) and Schizosaccharomyces species isolated

from local palm wine and concluded that Schizosaccharomyces yeasts were suitable for the

production of sweet, table mango wine while S. cerevisiae yeasts were suitable for the

production of dry mango wine with a higher ethanol level (Obisanya et al. 1987). However,

detailed vinification techniques and the chemical composition of the wine produced were not

reported in these studies.

The gap in knowledge was addressed by Reddy and colleagues with extensive

investigations on mango wine with special focus on several aspects (Reddy et al. 2009;

Reddy and Reddy 2009; Sudheer et al. 2009; Reddy and Reddy, 2011; Varakumar et al.

2011). After an initial study that reported on the production and characterization of mango

wine from popular Indian cultivars (Reddy and Reddy 2005, 2007), studies on the

optimisation of fermentation conditions by employing response surface methodology

(Sudheer et al. 2009), the effects of enzymatic maceration on synthesis of higher alcohols

(Reddy and Reddy, 2009), the analysis of volatile composition of wine fermented from

Indian mango cultivars (Reddy et al. 2010), the effects of fermentation conditions on yeast

growth and volatile production (Reddy and Reddy, 2011), and the antioxidant potential of

mango wine (Sudheer et al. 2012) were reported subsequently. These studies concluded that

25°C, pH 5, with 100 ppm SO2 and initial aeration were optimum fermentation conditions for

the production of mango wine. In addition, Reddy et al. (2010) also concluded that the

aromatic compounds of mango wine were comparable in concentration to those of grape wine.

Pectinase enzyme treatment was also found to be effective in increasing the yield of juice, the

production of ethanol, increasing the yield of higher alcohols (Reddy and Reddy 2009) and

Page 16: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

5

improving mango wine (Li et al. 2013). Pulp maceration was also discovered to have a

positive effect on the chemical profile of mango wine (Li et al. 2013). In addition, it was

reported that the cultivar of mango used had an effect on the volatile composition of the

mango wine produced, indicating that it might be possible to produce mango wines with

‘varietal’ character, just like grape wines (Li et al. 2012).

The effects of mixed yeast culture fermentation (S. cerevisiae and Metschnikowia

pulcherrima or Torulaspora delbrueckii; and S. cerevisiae and Williopsis saturnus) on the

aroma and sensory properties of mango wine were also investigated (Li et al. 2012;

Varakumar et al. 2012). Differences in volatiles produced were observed in wines produced

by different yeast strains and/or mixed culture. The two non-Saccharomyces yeasts (M.

pulcherrima and W. saturnus ) were unable to complete the fermentation, but the mixed

cultures were able to produce similar levels of ethanol relative to the monoculture of S.

cerevisiae, coupled with a higher glycerol content but lower volatile and total acidity

(Varakumar et al. 2012).

Malolactic fermentation with lactic acid bacteria is often conducted to convert the sharp

and tart-tasting malic acid into the mellow and softer tasting lactic acid. Oenococus oeni (O.

oeni) Lactobacillus and Pediococcus are also some of the common microorganisms used.

The simultaneous inoculation of O. oeni and S. cerevisiae resulted in higher amounts of ethyl

acetate and a bigger decrease in acetaldehyde content than with S. cerevisiae alone

(Varakumar et al. 2013). In addition, mango wine inoculated with O. oeni gave higher

sensory impacts than the control wine (monoculture of S. cerevisiae), and simultaneous

inoculation showed better sensorial attributes in flavour, fruity aroma, and overall

acceptability than sequential inoculation (Varakumar et al. 2012).

Page 17: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

6

The analysis of volatile compounds in mango wine has been conducted by a few groups

(Reddy and Reddy 2010; Li et al. 2011; Pino and Queris 2011). Reddy and colleagues

reported on the concentration of major alcohols, esters, fatty acids and ketones of mango

wines made from the Banginapalli and Alphonso varieties. Pino and Queris (2011) and Li et

al. (2011) assessed the contribution of the identified volatile compounds to the aroma of

mango wine on the basis of their odour activity values (OAV). Pino and Queris (2011)

identified 40 esters, 15 alcohols, 12 terpenes, 8 acids, 6 aldehydes and ketones, 4 lactones, 2

phenols, 2 furans and 13 miscellaneous in the mango wine. Li et al (2012) identified 4 acids,

7 alcohols, 25 esters, 8 carbonyl compounds and 1 miscellaneous compound as being

important to the flavour of mango wine.

1.3 Biochemistry of fermentation, wine flavour and quality

Wine making is a complex process that involves interactions between the fermentative

yeasts and the numerous compounds present in the must through a series of biochemical

reactions. A wide range of compounds (fermentative products) are produced during

fermentation, affecting appearance, aroma, flavour and mouth-feel and overall organoleptic

properties of the wine (Plata et al. 2003; Francis and Newton 2005; Jones et al. 2008; Sáenz-

Navajas et al. 2010).

The two main biochemical reactions that affect wine quality are primary alcoholic and

glyceropyruvic fermentation. The former converts sugars into ethanol anaerobically while

the latter produces glycerol which affects the mouthfeel and texture of the wine (Nieuwoudt

et al. 2002; Jones et al. 2008). Pyruvate is generated as an intermediate product in both

pathways; the by-products of pyruvate metabolism are also precursors for other flavour

compounds (Swiegers et al. 2005; Ardö 2006). The major volatile products are typically

Page 18: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

7

ethanol, fusel alcohols, esters and other compounds which make up the fermentation bouquet

that impacts the overall flavour profile of the wine. In addition, fermentative flavour can also

be influenced by other compounds such as long-chain fatty acids, nitrogenous and sulphur

containing compounds. These compounds are not directly fermented, but can diffuse into the

yeast cells and undergo biochemical reactions producing numerous volatile by-products

(Salmon et al. 1993); the production levels and metabolism of these compounds are variable

and yeast strain specific (Pretorius et al. 1999).

Wine quality is heavily influenced by the chemical composition of the wine at the point

of consumption due to their interaction with the sense of taste and smell. Wine flavour is due

to the interaction of non-volatile chemicals (which lead to taste sensations) and volatile

compounds (which are responsible for the odour — one of the most important factors

affecting wine quality). Some major non-volatiles include glycerol, sugars, organic acids and

polyphenols which affect the mouthfeel, sweetness, sourness and astringency, respectively.

The concentration of non-volatiles needs to be at least approximately 1% to have an effect on

the wine taste (Rapp and Mandery 1986). The volatile composition is especially important as

many consumers associate complexity in aroma with a higher quality wine. These volatile

compounds may originate from the original fruit (primary aroma), or from oenological

processes such as fermentation (secondary aroma) or aging process (tertiary aroma), and

include alcohols, esters, aldehydes, ketones, lactones, fatty acids, benzene derivatives,

terpenes, C13-norisoprenoids, thiols, etc. (Rapp and Mandery 1986).

The primary aroma is dominated by the major volatiles found in the fruit such as

terpene hydrocarbons and norisoprenoids (El Hadri, et al. 2010), while fermentative products

such as alcohols and esters dominate the secondary aroma. During alcoholic fermentation,

each glucose molecule is broken down into two pyruvate molecules that are subsequently

Page 19: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

8

decarboxylated into acetaldehyde and carbon dioxide by pyruvate decarboxylase. This is

followed by the reduction of acetaldehyde to ethanol by alcohol dehydrogenase.

In addition, higher alcohols can be formed from amino acids through the Ehrlich

pathway. After the initial transamination reaction, the resultant α-keto acids can be further

converted into respective higher alcohols (Ardö 2006). Some higher alcohols can also be

formed through sugar metabolic pathways since α-keto acids can be formed in the

tricarboxylic acid cycle (TCA cycle) (Pietruszka et al. 2010). Furthermore, hydrolytic

enzymes produced by yeasts also release a wide range of secondary metabolites initially

present in the must as non-volatile flavour precursors, contributing to varietal character

(Lambrechts and Pretorius 2000; Swiegers et al. 2005; Dubourdieu et al. 2006; Jolly et al.

2006; Li et al. 2013).

Tertiary flavour development during the aging of wine is a slow and gradual process

(Sumby et al. 2010). Acetate esters usually decrease in the first few years of aging due to

acid-catalysed hydrolysis resulting in a loss of freshness and fruitiness, especially in white

wines. On the other hand, straight-chain fatty acid ethyl esters remain relatively constant

(Rapp and Mandery 1986) as hydrolysis of ethyl esters is slower than that of acetate esters

considering the high concentration of the hydrolytic product ethanol which may cause

inhibition of ethyl ester hydrolysis (Rapp and Mandery 1986).

1.4 Influence of fermentation conditions and yeast strains

1.4.1 Amelioration of must

Conventionally, wine grapes are carefully cultivated to contain the optimum amounts

of nutrients for the production of a balanced wine. However, many factors beyond the

Page 20: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

9

control of the viticulturist and winemaker (for instance, weather and climate change) may

affect the composition of the grape (Jackson and Lombard 1993), resulting in the need for

additions of adjuncts to produce a balanced wine. Some of the common adjuncts include

acids (for optimum pH), yeast nutrients (nitrogenous source) and sugars (This et al. 2006;

Mendoza et al. 2009).

Initial sugar content is one of the most important parameter monitored; insufficient

sugars would result in a wine with insufficient ethanol content. However, high sugar

concentration may have an effect on final wine quality due to the effects of hyperosmotic

stress on the yeasts which includes rapid reduction in internal cell volume, efflux of water

from the cell, lowering turgor pressure, reducing water availability and causing cell shrinkage

(Hohmann 1997). Yeasts accumulate compatible solutes and osmoprotectants under

hyperosmotic conditions (Thomas et al. 1994); and osmotolerant yeasts are able to retain

synthesized glycerol as osmoregulator, some species even have active glycerol uptake pumps

(van Zyl et al. 1990). However, osmosensitive species, such as S. cerevisiae, tend to leak

glycerol significantly, except under hyperosmotic conditions when retention is improved

(Bauer and Pretorius 2000). In addition, hyperosmotic conditions also impact volatile

production due to the imbalance of redox potential (Jain et al. 2011; Styger et al. 2013). The

most significant effects on wine quality are likely to be the consequence of excess acetic acid

and acetaldhyde production (Swiegers et al. 2005).

1.4.2 Saccharomyces in wine production

The Saccharomyces species, especially S. cerevisiae, is the dominant yeast utilised in

commercial wine production today. It is highly adapted to fermenting grape must in

monoculture and has the ability to modulate most of the major constituents of wine, including

residual sugars, ethanol, polyols, acids and phenolics (Swiegers et al. 2005; Pretorius 2007).

Page 21: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

10

Commercially, strain selection places an emphasis on the ability to produce wine with low

residual sugar and high ethanol (Jackson and Lombard 1993). However, different strains

exhibit different traits and have different nutrient requirements; hence, the suitability of each

strain is dependent on both the desired wine style as well as the initial physicochemical

properties of the must (Heard and Fleet 1986). S. cerevisiae typically produces fruity/estery

wines that contain higher concentrations of ethyl esters of fatty acids with lower

concentrations of higher alcohols, which could otherwise mask aroma intensity (van der

Merwe and van Wyk, 1981); or wines with enhanced varietal character from the release

and/or modification of native flavour compounds to yield varietal aroma compounds such as

the fruity, long-chain, polyfunctional thiols 4-mercapto-4-methylpentan-2-one, 3-

mercaptohexan-1-ol and 3- mercaptohexyl acetate (Dubourdieu et al. 2006; Swiegers et al.

2008).

1.4.3 Non-Saccharomyces species in wine production

The aroma profiles of non-Saccharomyces fermented wines are distinctly different from

S. cerevisiae wines. Some species such as Williopsis saturnus and Kloeckera apiculate

produce significantly higher amounts of desirable acetate esters which may impart floral and

fruity notes to the wine bouquet (Li et al. 2012); while other species such as Torulaspora

delbrueckii produce low amounts of acetic acid (Ciani and Maccarelli 1999; Rojas, et al.

2001; Jolly, et al. 2006; Fernández, et al. 2000). In addition, some other high ester producing

non-Saccharomyces species include the Hanseniaspora family, Pichia anomala, etc. These

non-conventional yeast strains yield novel aroma profiles, many of which are perceived as

positive, indicate potential application in wine production by providing aroma diversity

(Dubourdieu et al., 2006).

Page 22: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

11

Glycosidases can also potentially enhance the aroma and flavour properties of wine.

However, most Saccharomyces species have low glycosidase activities (Heard and Fleet 1986;

Li et al. 2013), therefore, non-Saccharomyces species with glycosidase activities have been

reported to contribute positively to enzymatic reactions during the early stages of vinification.

Some of these species include the Candida, Debaryomyces, Hanseniaspora, Kloeckera,

Kluyveromyces, Metschnikowia, Pichia, Saccharomycodes, Schizosaccharomyces, and

Zygosaccharomyces genera (Rosi, Vinella and Domizio 1994). The liberation of aromatic

terpenols from their odourless precursors has also been linked to the enzymatic activities of

non-Saccharomyces species (Lagace and Bisson 1990). In addition, some strains such as

Kloeckera apiculata has the potential to reduce protein haze due to their significant protease

content (Lagace and Bisson 1990).

1.4.4 Inoculation strategies in wine fermentation

Due to its weak weak fermentative capability and low ethanol tolerance, complete

fermentation with a non-Saccharomyces monoculture is not possible, therefore resulting in

the limited use of non-Saccharomyces species in commercial wine fermentation.

Consequently, multistarter cultures have been explored and utilised to harness the advantages

of non-Saccharomyces yeast species. Two strategies have evolved to harvest the desirable

aroma profile and to enable complete fermentation with non-Saccharomyces yeasts — co-

fermentation with a robust Saccharomyces strain; and sequential fermentation, in which the

non-Saccharomyces yeast and Saccharomyces strain are inoculated successively, in order to

complete fermentation. Wines produced by these methods have proven to have distinct and

desirable traits (Soden et al. 2000; Holzapfel 2002;, Jolly et al. 2006; Li et al. 2012; Maturano

et al. 2012)

Page 23: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

12

A successful co-fermentation depends on the physiological properties of the individual

yeasts – its compatibility with other yeasts and the effects on growth rate and biomass

development. Suppression of one yeast by the other can result in its reduced metabolic

activity and hence decreased impact on the wine characteristics. One strategy may be the co-

inoculation of a weakly fermentative yeast at high ratio to a strongly fermentative yeast to

also achieve a greater impact of the former yeast and produce a more balanced wine during

co-fermentation. Diffusion of various metabolites between yeasts with different ‘metabolic

tuning’ can result in metabolite concentrations different from those that would be achieved by

blending wines (Ciani and Maccarelli 1999; Fernández et al. 2000; Clemente-Jimenez et al.

2005; Jolly et al., 2005; Augustyn and Pretorius 2006; Lee et al. 2012a; Lee et al. 2012b; Li

et al. 2012).

A sequential fermentation of Pichia fermentans and S. cerevisiae conferred greater

complexity to wine through the enhancement of desirable flavour compounds production and

glycerol content (Clemente-Jimenez et al. 2005). In addition, the use of multistarter

fermentations to reduce the negative sensorial characteristics and for biological acidification

of wines has also been reported. Sequential fermentations can also be used to favour weak

fermentative strains by delaying inoculation of S. cerevisiae so that the desirable traits

conferred by the weak fermentative strain can be developed first, for instance, good acidity,

low volatile acidity, intense fruity ester production, and in some wines, a desirable ‘wild

yeast’ fermentation character (Moreno et al. 1991; Kapsopoulou et al. 2007; Bely et al. 2008).

Page 24: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

13

1.5 Research aims and objectives

The overall objective of this research project was to study the effects of sugar

concentration and inoculation strategies on mango wine fermentation.

Aim 1: Effect of sugar concentration on mango wine fermentation with S. cerevisiae

MERIT.ferm — Chapter 3

High sugar concentration or hyperosmotic pressure creates redox imbalance and the

efforts of the yeasts to combat this stress to ensure its survival have implications on wine

quality. The effects of initial sugar concentration on yeast biomass accumulation, glycerol

and volatile compounds production with special emphasis on acetaldehyde and acetic acid

were investigated. While it has been hypothesised that high initial sugar concentration will

impact wine quality negatively, the specific effects are unknown. Hence, the specific effects

of high initial sugar concentration on volatile production and its relation to redox equilibrium

are presented in Chapter 3.

Aim 2: Effect of co-inoculation of S. cerevisiae and W. saturnus on mango wine

fermentation — Chapter 4

As yeast strain is one of the most crucial factors determining the quality of wine, effects

were investigated by analysing the differences in volatile composition and growth kinetics

between the two different yeasts, S. cerevisiae MERIT.ferm and W. saturnus NCYC 500 and

simultaneous mixed culture fermentation with MERIT.ferm and W. saturnus NCYC 500 in

the ratio of 1:1000. The results are presented in Chapter 4.

Aim 3: Effect of different sequential inoculation strategies on mango wine fermentation

— Chapter 5

Due to the effects that different inoculation strategies have on the flavour and quality of

wine, and the lack of studies investigating the effects of sequential inoculation strategies on

the production of mango wine involving the use of S. cerevisiae MERIT.ferm and W.

Page 25: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

14

saturnus NCYC 500, Chapter 5 presents information on the differences in yeast growth

kinetics, volatile composition and sensory perception in three wines with different inoculation

strategies.

Page 26: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

15

Chapter 2 Materials and Methods

2.1 Mango fruits and preparation of mango juice

Mango fruits (Mangifera indica L.) of the Chok Anan variety imported from Malaysia

were purchased from a local fresh produce wholesale market. Whole, healthy looking fruits

were selected and stored at room temperature until fully ripe and soft. After washing with tap

water to remove dirt and being allowed to air dry naturally at room temperature, the skin and

flesh were removed manually and separated. The resulting flesh was then juiced in a

commercial juicer, Sona juice extractor (Cahaya Electronics, Singapore) with the resulting

puree being centrifuged at 4°C, 41 415 x g (Beckman Centrifuge, Brea, CA, USA) for 15 min

and the supernatant removed and stored at -50°C before use.

The mango juice (initial pH 4.5 to 4.6 and 15 – 18 °Brix) was then adjusted to a pH of

3.5 with 50% (w/v) DL-malic acid solution before the addition of 100 ppm potassium

metabisulfite (K2S2O5, Goodlife Homebrew center, Norfolk, England). The mixture was left

to stand for 24 h at 25°C for sterilisation. This sterilisation process aimed to kill any wild

yeast that may be present in the juice.

Handling of all materials was conducted in a bio-fumehood to maintain sterility. The

effectiveness of sterilisation by 100 ppm SO2 sterilisation was verified by by streak plating on

potato dextrose agar medium plate (39 g/L, Oxoid, Basingstoke, Hampshire, England).

Page 27: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

16

2.2 Yeast and culture media

Williopsis saturnus var. mrakii NCYC 500 from the National Collection of Yeast

Culture (Norwich, UK) and Saccharomyces cerevisiae MERIT.ferm from Chr.-Han.

(Denmark) were received in the active freeze dried form used in this study.

Active dried yeast was propagated in a sterilised nutrient broth consisting of (on a w/v

basis), 2% glucose, 0.25% bacteriological peptone, 0.25% yeast extract and 0.25% malt

extract in deionised water, pH 5.0. This solution was first autoclaved for 15 min at 121°C

before inoculation. These yeast cultures were then sub-cultured on potato dextrose agar

plates (0.4% w/v potato extract, 2% w/v dextrose and 1.5% w/v agar, pH 5.6 at 25°C) for 2

days at 25°C. An isolated single colony was suspended into 10 mL of above-mentioned

nutrient broth; following which, yeast strains were maintained in the nutrient broth and

incubated at 25°C for 48 – 72 h without aeration. Finally, 20% glycerol was added to the

culture before being stored at –80°C until further use.

All media and equipment were sterilised at 121°C for 15 min before use and purity

checks on stock cultures were conducted prior to all fermentations.

2.3 Preparation of yeast starter culture

Each pre-culture was prepared with sterilised mango juice inoculated with 10% (v/v) of

with S. cerevisiae MERIT.ferm or W. saturnus NCYC 500. The pre-cultures were then

incubated at 25°C for 48 to 72 h for the yeasts to reach a concentration of 107 CFU/mL.

Assessment of yeast cell growth was conducted via the spread plating method on PDA plates.

This method of pre culture fermentation was used for the single culture fermentations

(Chapter 3).

Page 28: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

17

2.4 Fermentation of mango juice with different initial sugar concentrations

— Chapter 3

Three different sets of mango juice (200 mL of sanitised mango juice in sterilised

conical flasks) were prepared for fermentation. Two sets were supplemented with glucose to

attain medium and high sugar contents with readings of 23°Brix and 30°Brix, respectively. A

set of control fermentation with no sugar supplementation (low sugar fermentation) was also

conducted. Each conical flask was inoculated with 105

CFU/mL of S. cerevisiae

MERIT.ferm then plugged with cotton wool and wrapped with aluminium foil. Static

fermentation was carried out for 35 days at 20°C. The fermentations were conducted in

triplicate.

2.5 Fermentation of mango juice with co-inoculated S. cerevisiae and W. saturnus

— Chapter 4

Two different starter cultures were prepared as describe in section 2.3 – one S.

cerevisiae and one W. saturnus NCYC 500. Triplicates of 200 mL of sanitised mango juice

in sterilised conical flasks plugged with cotton wool and wrapped with aluminium foil were

fermented for 21 days at 20°C. S. cerevisiae MERIT.ferm and W. saturnus NCYC 500 were

each inoculated into a flask of sterilised mango juice (200 mL) at a concentration 105

CFU/mL for the single culture fermentations. For the mixed culture fermentation, S.

cerevisiae MERIT.ferm and W. saturnus NCYC 500 from the two single cultures were

inoculated in a ratio of 1:1000 into the mango juice simultaneously.

Page 29: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

18

2.6 Fermentation of mango juice with different sequential inoculation strategies

— Chapter 5

Briefly, three different sequential inoculation strategies were studied. Triplicates of

200 mL of sanitised mango juice in sterilised conical flasks plugged with cotton wool and

wrapped with aluminium foil were fermented for 21 days at 20°C.

In the simultaneous mixed culture fermentation (MCF), 105

CFU/mL of W. saturnus

NCYC 500 and 102

CFU/mL of S. cerevisiae MERIT.ferm, in a ratio of 1000:1 were

simultaneously added. For the positive sequential fermentation (PSF), 105

CFU/mL W.

saturnus NCYC 500 was added and fermentation was carried out for 14 days before the yeast

cells were deactivated by being subjected to ultrasonication. S. cerevisiae MERIT.ferm was

then added at 102

CFU/mL and the fermentation continued for another 7 days. For the

negative sequential fermentation (NSF), 102

CFU/mL of S. cerevisiae MERIT.ferm was

inoculated and fermentation was carried out for 7 days before the yeast cells were deactivated

by ultrasonication. 105

CFU/mL of W. saturnus NCYC 500 was then added and fermentation

continued for further 14 days. S. cerevisiae MERIT.ferm fermentation was halted at Day 7

because it had been shown that S. cerevisiae was able to complete the fermentation within

that timeframe. Ultrasonication was conducted with the probe of the ultrasonicator

(Hielscher – Ultrasound Technology, UIP 1000, 1000W).being sterilised with 70% ethanol

solution prior to each run. Each conical flask was partially immersed in ice water to prevent

the sample from overheating and affecting the flavour profile. The sample was then

subjected to 15 min of treatment at 20 kHz. Plating on potato dextrose agar (PDA) was done

to check for sterility of mango juice after ultrasonication treatment by drawing samples and

doing streak plating.

Page 30: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

19

2.7 Analytical methods

2.7.1 pH, oBrix and yeast enumeration

Sampling was done at regular intervals throughout the fermentation process. Aliquots

of approximately 10 mL were drawn under aseptic conditions after swirling to obtain a

homogenous sample. The pH and total soluble solids were measured using a refractomer

(Atago, Tokyo, Japan) and pH meter (Metrohm, Herisau, Switzerland), respectively.

Cell counts for yeasts were carried out via the spread plating method on potato dextrose

agar (PDA). Suitable dilutions of up to 108

were carried out with 1% peptone water to obtain

suitable cell counts. The plates were then incubated at 25°C for 48 h before yeast colonies

were counted. Lysine agar is unable to support the growth of Saccharomyces yeast (Erten

and Tanguler 2010) and hence was used to differentiate Saccharomyces and non-

Saccharomyces yeast growth pattern for the mixed culture experiments. Other than the

selective media used to differentiate between the MERIT.ferm yeast and NCYC 500 yeast,

the appearance of the colonies was used as a means of identification. S. cerevisiae yeasts

appeared as small, off-white colonies with a glossy surface while NCYC 500 colonies had a

dull, white appearance with a slightly bigger size. All yeast enumeration analyses were done

in duplicate.

2.7.2 Analysis of sugars and organic acids

The instrumental analysis and quantification of sugars and organic acids were

conducted using a Shimadzu modular chromatographic system (LC solution software version

1.25) equipped with LC-20AD XR pumps and coupled to a SPD-M20A photodiode array

detector, a low temperature evaporative light, scattering detector (ELSD-LT), a SIL-20AC

XR autoinjector.

Page 31: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

20

Prior to HPLC analysis, samples were centrifuged at 4 248 g at 4°C for 25 min (Sigma

3-18K centrifuge, Osterode am Harz, Germany), filtered with a 0.20-μm RC membrane

(Sartorius, Gottingen, Germany) and stored at –50°C before analysis. Analysis was

conducted in triplicate. Compounds were identified by comparing retention time, spectrum

and concentration with external reference standards.

Analysis of organic acids was conducted with a Supelcogel C-610 H column (300 × 7.8

mm, Supelco, Bellefonte, PA, USA). The mobile phase was 0.1% (v/v) sulphuric acid at a

flow rate of 0.4 mL/min at 40°C and detection was done by photodiode array at 210 nm

wavelength.

Sugars were analysed by using the ELSD-LT (gain: 5; 40°C; 350 kPa) coupled with a

Zorbax carbohydrate column (150 x 4.6 mm, Agilent, Santa Clara, CA, USA) using a mixture

of acetonitrile and water (80:20 v/v) as the mobile phase with a flow rate of 1.4 mL/min at

40°C.

2.7.3 Analysis of volatile compounds

Analysis of volatiles in both mango juice and wine was carried out using optimised

headspace-solid phase microextraction combined with gas chromatography-flame ionization

detector/mass spectrometry (HS-SPME GCMS/FID). Although traditionally used as a

qualitative and semi-quantitative method (Sánchez-Palomo et al. 2005; Trinh et al. 2011), it

has been demonstrated that HS-SPME GCMS/FID can be applied quantitatively if extraction

and analytical conditions were optimised and consistently employed (Baptista et al. 2001;

Lee et al. 2010). The analytical method was optimised by varying desorption temperature

and time, flow rate, temperature profile gradient and evaluated based on the peak areas

obtained for the compounds of interest. The nature of the matrix, the amount of sample,

Page 32: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

21

desorbing conditions, fiber coating, extraction temperature and time can have an effect on the

analytical results.

A SPME fused silica fiber coated with 85 μm carboxen/polydimethylsiloxane

(CAR/PDMS) (Supelco, Sigma-Aldrich, Barcelona, Spain) was used for extraction. A 5-mL

sample was placed in a 20-mL glass vial tightly capped with a PTFE/silicone septum and

extracted by HS-SPME at 60°C for 40 min with 250 rpm agitation; after which, the fibre was

desorbed at 250°C for 3 min and injected into Agilent 7890A GC (Santa Clara, CA, USA),

coupled to FID and Agilent 5975C triple-axis MS. Chromatographic separation was

achieved via a capillary column (Agilent DB-FFAP) of 60 m × 0.25 mm I.D. coated with

0.25 µm film thickness of polyethylene glycol modified with nitroterephthalic acid.

Helium was used as the carrier gas with a liner velocity of 1.2 mL/min, transfer line

temperature 280°C. Mass detector conditions were set at 70eV electron impact (EI) mode,

230°C source temperature. The mass scanning parameters were: 3 min → 22 min: m/z 25–

280 (5.36 scan/s); 22 min → 71 min: m/z 25–550 (2.78 scan/s) under full-scan acquisition

mode.

Volatiles were identified by matching their mass spectra with the Wiley mass spectrum

library and confirmed with the linear retention index (LRI) values, which were determined on

the FFAP column against a series of alkanes (C5-C25) separated under identical operating

conditions. The linear retention index (LRI) was used to identify the compound, and the

calculation for the LRI is given as :

LRI=100×[(ti-tz)/(tz+1-tz)+z], where

ti = retention time of compound,

tz = retention time of preceding n-alkane,

tz +1 = retention time of subsequent n-alkane

Page 33: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

22

Several volatiles were selected to be quantified using external standards (supplied by

Firmenich Asia Pte Ltd, Tuas, Singapore) dissolved in 10% v/v micro-filtered (0.45 μm)

mango juice diluted with water, except for ethanol that was dissolved in 100% micro-filtered

mango juice. The analyses were carried out in triplicate. Following the analyses, a standard

curve was constructed for the linear range of each compound. The results shown represent

the means of three independent fermentations.

Concentrations of volatile compounds were determined by using the linear regression

equations of the corresponding standards. Odour activity values (OAVs) of quantified

volatiles were calculated according to their known thresholds from the literature (Bartowsky

and Pretorius, 2009; Ferreira, Lopez, and Cacho, 2000). The formula for the calculation of

OAV used was as follows:

OAV =

.

2.8 Sensory analysis

Sensory analysis was conducted for the mango wines produced by the simultaneous

mixed-culture and sequential fermentations by a panel of six experienced flavourists (females

and males) from Firmenich Asia Pte Ltd (Singapore) using quantitative descriptive analysis

(QDA) methodology. A constant volume of wine was presented in wine-testing glasses and

was arbitrarily coded. The samples were sniffed and the aroma intensity of each sensory

descriptor was rated on a 5-point hedonic scale, where 0 indicated that the descriptor was not

perceivable and 5 indicated that the descriptor had extremely high intensity. The eight

mutually agreed sensory descriptors to describe the wine aroma were acidic, alcoholic,

buttery, cocoa, fruity, fusel, sweet and yeasty notes.

Page 34: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

23

2.9 Statistical analysis

The ANOVA test using Microsoft Excel (ver 2007) was used to analyse the data

obtained. Results were considered statistically significant if the value of P was less than 0.05.

Each set of experiment was conducted in triplicate; hence, mean values and standard

deviations were calculated from the triplicate.

Page 35: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

24

Chapter 3 Effects of Sugar Concentration on Volatile Production by

Saccharomyces cerevisiae MERIT.ferm in Mango Juice Fermentation

3.1 Introduction

Oenological conditions could be stressful for most yeast strains, with many of these

stresses encountered simultaneously or rapidly one after another. For instance, the high sugar

content (hyperosmotic stress), nutrient deficiency (lack of nitrogen sources, oxygen, vitamins

and other minerals), low pH, extreme temperatures all present as stressful conditions for the

propagation and growth of most yeast strains during the onset of fermentation while ethanol

toxicity occurs later in the fermentation process. Although commercial wine yeasts are

selected for their inherent ability to cope with such environmental stresses, yeast metabolism

is still significantly affected by fermentation conditions.

High gravity fermentation or fermentation with high sugar content has been shown to

have an effect on volatile production. The level of sugar substrate directly affects yeast

metabolism which regulates the biochemical assimilation (energy consumption) and

dissimilation (energy generation) of nutrients by yeast cells (Walker 1998); this pathway is

intrinsically linked to the production of both volatile compounds and non-volatile compounds.

Most studies reported increases in acetaldehyde and acetic acid and glycerol production in

high gravity fermentation by yeasts (Bely et al. 2003; Chaney et al. 2006).

The plasma membrane and actin skeletion are damaged due to the sudden loss in turgor

pressure in response to hyperosmotic stress, potentially leading to the cessation of growth

(Arrizon and Gschaedler 2002). To compensate for the efflux of water, water from the

vacuole is released to provide a buffer for a short adjustment period (Bauer and Pretorius

2000). Concurrently, glycerol production is stimulated and the export channel closes;

Page 36: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

25

accumulation of glycerol occurs until the influx of water restores the critical cell size for cell

growth to occur (Chaney et al. 2006).

This chapter investigated the response of S. cerevisiae MERIT.ferm to hyperosmotic

pressure in mango juice fermentation due to high sugar content with a focus on yeast growth,

sugar consumption, glycerol production and key volatile production in low sugar

fermentation (unfortified mango juice with initial TSS of 16.6°B), medium sugar

concentration (fortified mango juice with initial TSS of 23°B) and high sugar fermentaiton

(fortified mango juice with initial TSS of 30°B). The sugar concentrations were selected for

this study based on the typical total soluble solid (TSS) in °Brix for wine production. For

still table wines, the recommended °Brix typically ranged from 20°B to 23°B (medium sugar

fermentation) while German regulations for the production of ice wine (or Eiswein) had to be

between 26 to 30°B (high sugar fermentation) (Boulton, et al. 1996).

3. 2 Results and discussion

3.2.1 Mango juice volatile composition

In order to ascertain the effects that fermentation has on the volatile compounds, it is

necessary to gain knowledge of the original volatile profile of the fresh Chok Anan mango

juice. From the literature research conducted, there is only one detailed analysis of the

flavour profile of Chok Anan mango (Li et al. 2011). Other information sources mentioned

major character impact volatile compounds but did not provide a detailed list. Most studies

focused on sensory evaluation (Vásquez-Caicedo et al., 2002) but did not report on the full

list of volatiles and only reported on the key compounds believed to have an effect on

consumers’ acceptability (Laohakunjit et al. 2005).

Page 37: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

26

In this study, more than a hundred volatiles were detected in the fresh mango juice; a

total of 59 major volatile compounds deemed to have an impact on the flavour profile were

identified in the fresh Chok Anan mango juice (Table 3.1). These major volatile compounds

identified (by relative peak area or RPA) were mainly monoterpenes (12), sesquiterpenes (6),

alcohols (11) and esters (12). Monoterpenes made up the largest group of volatile

compounds and constitute approximately 58% of the volatiles by RPA.

The major volatile compound by RPA was α-terpinolene, followed by p-cymenene, p-

cymene and δ-3-carene. α-Terpinolene has been described as sweet, floral with pine-like

aroma notes, while δ-3-carene has been described as sweet, floral and mango leaf-like. This

finding differs slightly from data reported by Laohakunjit et al. (2005) where δ-3-carene as

the second most abundant monoterpene. Some of the other terpene hydrocarbons identified

in the same study were δ-thujene, α-pinene, mycrene, and α-phellandrene. Most of these

volatile compounds were also identified in this current study. Terpenic notes were thought to

be enhanced by the presence of other volatile compounds such as 3-hexanol, 2-hexanol, γ and

δ-lactones, and furan compounds to give the overall flavour perception (Chauhan et al. 2010).

However, some of these compounds such as 2-hexanol, 3-hexanol and δ-lactone were not

found in this current study. This discrepancy could be due to the differences between the

mangoes used. It has been shown previously that different climatic conditions and maturity

of the fruit when harvested could lead to differences in the organoleptic quality (Lalel et al.

2003).

Many of the volatiles that are regarded as important to the aroma profile of the mango

flavour were identified in this study. Some of these include butyl butanoate, hexyl formate,

3-hexenyl acetate, cis-rose oxide, cis-3-hexenol, β-damascenone and trans-2-hexanal (Pino et

al. 2005). The esters give rise to the overall fruity character while the overall floral top note

Page 38: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

27

of the mango flavour is derived from the alcoholic compounds such as linalool, 2-

phenylethanol, nerol and citronellol (Chauhan et al. 2010).

Page 39: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

28

Table 3.1 Volatile compounds in fresh Chok Anan mango juice analysed using HS-SPME-GC-MS/FID

CAS No1 LRI

2 Compound Peak Area (×10

6) RPA (%) Odour descriptors

3

Terpenes 007785-70-8 1088 α-Pinene 3.78 ± 0.05 0.46 Fresh, sweet, pine

013466-78-9 1206 δ-3-Carene 52.8 ± 1.02 6.49 Sweet citrus

002867-05-2 1219 α-Thujene 11.98 ± 0.09 1.47 Woody, green, herb-like

018172-67-3 1226 β-Pinene 1.14 ± 0.08 0.14 Sharp, terpenic, conifer

000099-86-5 1235 α-Terpinene 35.2 ± 0.61 4.33 Sharp, terpenic, lemon

095327-98-3 1254 Limonene 35.6 ± 0.55 4.38 Citrius, terpenic, orange note

000555-10-2 1265 β- Phellandrene 3.04 ± 0.28 0.37 Mint terpentine

027400-71-1 1290 β-Ocimene 1.1 ± 0.25 0.13 Citrus, green, lime

000099-85-4 1305 γ-Terpinene 19.13 ± 0.41 2.35 Fatty, terpenic, lime

000099-87-6 1339 p-Cymene 12.7 ± 2.36 1.56 Citrus, terpenic, woody

000535-77-3 1343 m-Cymene 73.97 ± 5.06 9.09 Citrus, terpenic, woody

000586-62-9 1352 α-Terpinolene 348.97 ± 5.67 42.91 Citrus, lime, pine

000673-84-7 1450 allo-Ocimene 0.73 ± 0.47 0.09 Floral, nutty, peppery

001195-32-0 1529 p-Cymenene 60.7 ± 2.1 7.46 Citrus, pine-like

001879-84-6 1628 β-Farnesene 0.34 ± 0.03 0.04 Woody, citrus, sweet

000087-44-5 1695 trans-Caryophyllene 0.39 ± 0.02 0.05 Woody, clove note

000473-13-2 1696 α-Selinene 0.35 ± 0.03 0.04 Amber

004630-07-3 1826 Valencene 0.19 ± 0.02 0.02 Orange, citrus, woody

017066-67-0 1829 β- Selinene 1.12 ± 0.05 0.14 Herbal

000515-17-3

γ-Selinene 0.56 ± 0.06 0.07 Woody

Subtotal

663.77 81.61

Alcohol 000064-17-5 1032 Ethanol 7.42 ± 0.59 0.91 Alcoholic

000928-95-0 1445 trans-2-Hexen-1-ol 0.84 ± 0.06 0.1 Fruity, green, leafy

000111-27-3 1448 1-Hexanol 2.73 ± 0.3 0.34 Alcoholic

000928-96-1 1476 cis-3-Hexanol 65.45 ± 1.91 8.05 Green, leafy

000104-76-7 1527 2-Ethyl-1-hexanol 1.66 ± 0.08 0.2 Oily, rose, sweet

Page 40: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

29

Table 3.1 (Continued)

000470-08-6 1794 β-Fenchol 0.11 ± 0.01 0.01 Camphor-like, woody

000464-43-7 1808 Endo-borneol 0.21 ± 0.08 0.03 Camphor-like, woody

000106-22-9 1867 Citronellol 0.4 ± 0.02 0.05 Floral, rose, sweet, green, citrus

000078-70-6 1999 Linalool 0.2 ± 0 0.02 Fresh, floral, herbal, rosewood

000060-12-8 2035 2-Phenylethanol 0.24 ± 0.02 0.03 Rose, honey, floral

000128-37-0 2090 p−Cresol 0.25 ± 0.03 0.03 Cresol, medicinal, leather

Subtotal

79.51 9.78

Ester 000141-78-6 1009 Ethyl acetate 9.67 ± 0.85 1.19 Ethereal, fruity, sweet

000105-54-4 1095 Ethyl butanoate 0.76 ± 0.07 0.09 Sweet, fruity

000123-92-2 1112 Isoamyl acetate 11.7 ± 1.04 1.44 Sweet fruity, banana-like

000109-21-7 1284 Butyl butanoate 0.03 ± 0 0 Fruity, pineapple, sweet

003681-71-8 1396 3-Hexenyl acetate 25.07 ± 4.3 3.08 Sharp, fruity-green, sweet

002497-18-9 1410 trans-2-Hexenyl acetate 3.82 ± 0.3 0.47 Fruity, green, leafy

000629-33-4 1440 Hexyl formate 0.82 ± 0.02 0.1 Green, ethereal, fruity

033467-74-2 1466 cis-3-Hexenyl propionate 0.11 ± 0.01 0.01 Fresh, fruity, green

016491-36-4 1546 cis-3-Hexenyl isobutyrate 0.19 ± 0.01 0.02 Apple, fruity, green

065405-80-3 1700 cis-3-Hexenyl trans-3-butenoate 0.07 ± 0 0.01 Green, sweet, fruity

000103-45-7 1862 2-Phenylethyl acetate 1.19 ± 0.09 0.15 Sweet, honey, floral, rosy

000110-38-3 1948 Ethyl dodecanote 0.62 ± 0.05 0.08 Sweet, wine, brandy

Subtotal

54.05 6.65

Acid 000064-19-7 1549 Acetic acid 0.23 ± 0.02 0.03 Vinegar-like

000067-43-6 1728 Butanoic acid 0.72 ± 0.07 0.09 Cheesy, rancid butter

000124-07-2 2171 Octanoic acid 0.6 ± 0.04 0.07 Acidic, fatty, soapy

Subtotal

1.55 0.19

Aldehyde 000066-25-1 1152 Hexanal 0.17 ± 0.01 0.02 Fatty, green, grassy

006728-26-3 1310 trans-2-Hexenal 4.59 ± 0.28 0.56 Green, leafy, fruity

Page 41: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

30

Table 3.1 (Continued)

000100-52-7 1633 Benzaldehyde 0.4 ± 0.03 0.05 Bitter almond

000620-23-5 1731 3-Methyl-benzaldehyde 0.96 ± 0.08 0.12 Cherry-like

000104-87-0 1771 p-Toloualdehyde 2.76 ± 0.23 0.34 Sweet aromatic, bitter almond

Subtotal

8.88 1.09

Ketone 000096-48-0 1758 5-Methyl dihydro-2(3H)-furanone 0.83 ± 0.08 0.1 Herbaceous, waxy, creamy note

023696-85-7 1938 β-Damascenone 2.6 ± 0.23 0.32 Sweet, floral, fruity, coconut

000104-50-7 2051 γ—octalactone 0.53 ± 0.05 0.06 Coconut, lactonic

Subtotal

3.95 0.49

Furan 003208-16-0 1044 2-Ethylfuran 2.52 ± 0.09 0.31 Ethereal run, cocoa note

Subtotal

2.52 0.31

Ether 068780-91-6 1537 trans-Linalool oxide 0.58 ± 0.03 0.15 Sweet, lemon

016409-43-1 1434 cis-Linalool oxide 0.58 ± 0.04 0.15 Earthy, floral, sweet, woody

001786-08-9 1563 Nerol oxide 0.31 ± 0.01 0.08 Floral, orange blossom, green, sweet

016409-43-1 1434 cis-Rose oxide 0.17 ± 0.01 0.04 Rose, geranium

Total

813.35 100 1

CAS numbers were obtained from Wiley database library

2 LRI of all the tables was determined on the DB-FFAP column, relative to C5-C40 alkanes.

3 Descriptors were retrieved from http://www.thegoodscentscompany.com

Page 42: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

31

3.2.2 Changes in pH and organic acids

There were no significant changes in pH throughout fermentation with pH values for all

three treatments maintained at about 3.50 to 3.52. This is likely due to the insignificant

changes in most organic acids before and after fermentation (Table 3.2). Fermentation had

no significant effects on citric, succinic and tartaric acids. Citric acid remained at

approximately 0.25 g/100 mL, succinic acid stayed at around 0.08 g/100 mL and tartaric acid

varied in the range of 0.12 to 0.15 g/100 mL before and after fermentation. Malic acid

decreased by approximately 50% after fermentation from 0.79 to 0.9 g/100 mL before

fermentation to 0.33 to 0.41 g/100 mL after fermentation. The relatively high malic acid

content before fermentation was due to the addition of D/L-malic acid to a attain a pH of 3.5

for the sanistised mango juice before fermentation. The decrease in malic acid after

fermentation was likely due to the passive diffusion of D-malic acid into the yeast cells as

reported previously (Coloretti et al. 2002). It was unlikely that malolactic fermentation

occurred due to the absence of lactic acid in the mango wine and the absence of lactic acid

bacteria in the microbial analyses. Furthermore, the addition of 100 ppm of potassium

metabisulfite would have inhibited the growth of wild yeasts and bacteria.

Page 43: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

32

Table 3.2 Physicochemical properties, organic acid and sugar concentrations of mango wine

before and after fermentation Day 0 Day 35

16.6°B 23°B 30°B 16.6°B 23°B 30°B

pH 3.51± 0.01 a 3.51± 0.01

a 3.52± 0.01

a 3.51± 0.01

a 3.50± 0.01

a 3.50± 0.01

a

oBrix (°B) 16.6 ± 0.01

a 23.1± 0.02

a 30.1± 0.04

a 5.6 ± 0.06

b 7.2 ± 0.01

c 10.6 ± 0.17

d

Plate count

(x105 CFU/mL)

5.65±0.58 a 5.85±0.58

a 6.35±0.58

a 8567±750

c 923±60

d 935±58

d

Sugar (g/100 mL)

Fructose 4.96±0.08 a 5.58±0.05

b 5.96±0.09

c N.D.* 1.11±0.02

d 2.46±0.15

e

Glucose 0.63±0.02 a 7.46±0.03

b 12.39±0.06

c N.D.* 0.38±0.05

d 0.22±0.06

e

Sucrose 12.25±0.31 a 13.86±0.12

b 14.19±0.36

c 0.013±0.00

d 2.53±0.12

e 3.84±0.36

f

Organic acid (g/100 mL)

Citric acid 0.26±0.04a 0.26±0.03

a 0.24±0.04

a 0.23±0.03

a 0.24±0.03

a 0.25±0.03

a

Malic acid 0.79±0.05 a 0.90±0.09

b 0.88±0.09

b 0.36±0.05

c 0.33±0.02

c 0.41±0.05

d

Succinic acid 0.083±0.012 a 0.079±0.012

a 0.081±0.012

a 0.086±0.011

a 0.079±0.012

a 0.081±0.012

a

Tartaric acid 0.12±0.03 a 0.14±0.03

a 0.14±0.03

a 0.14±0.02

a 0.11±0.02

a 0.14±0.02

a

abcd ANOVA (n=4) at 95% confidence level with same letters in the same row indicating no

significant difference

* N.D. Not Detected

3.2.3 Yeast growth, total soluble solids and sugar concentration

The total soluble solids (TSS) declined throughout fermentation with similar trends.

The oBrix value decreased from 16.6°B to 5.6°B for the low sugar fermentation, from 23.1°B

to 7.2°B for the medium sugar fermentation and 30.1°B to 10.6°B for the high sugar

fermentation. The decrease in oBrix value was more rapid initially before the rate of decline

decreased (Figure 3.1). The decrease in °Brix also correlated to the increase in yeast growth

(Figure 3.1) and overall sugar concentrations (Figure 3.2).

Different trends for yeast biomass were observed for the three treatments (Figure 3.1).

From Day 0 to around Day 10, the increases in yeast biomass and maximum cell populations

were inversely correlated with initial sugar concentrations. Similar trends were observed for

the low and medium sugar fermentation where viable cell population peaked between Days 9

and 11 before decreasing slowly until Day 16 when viable cell population decreased sharply.

Page 44: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

33

However, the maximum cell population for the medium sugar fermentation was about 0.5 log

lower than that in the low sugar fermentation (16.6°B). For the high sugar fermentation

(30°B), the increase in yeast biomass was slower than the other two fermentations. The

maximum viable cell population for the high sugar fermentation peaked between Days 9 to

11 but only attained a level of 7.7×107

CFU/mL, lower than that in the other two

fermentations. This could be attributed to the osmotic stress from the high sugar

concentration (Hohmann et al. 2003; Devantier, Pedersen and Olsson 2005). However,

surprisingly, there was no significant decline in cell numbers after Day 11. Viable cell

population was maintained at about 7×107

CFU/mL until the fermentation was terminated at

Day 35. This could be due to the presence of sugars and other nutrients, while nutrients were

depleted in the other two treatments resulting in the starvation of the yeast cells which

ultimately led to the decline in the number of viable cells. Furthermore, S. cerevisiae

MERIT.ferm is known to have a high ethanol tolerance according to the manufacturer’s

information sheet, hence, in the presence of sufficient nutrients, ethanol toxicity is not

expected to have a detrimental effect on biomass increase with increasing sugar

concentrations.

Page 45: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

34

Figure 3.1 Changes in yeast cell count and TSS (oBrix) for low ( ),

medium ( ) and high ( ) sugar fermentation

Fermentation was effectively completed at Day 9 and Day 17 for the low and medium

sugar fermentation respectively, when the oBrix value remained constant (Figure 3.1) and the

concentration of sugars no longer declined (Figure 3.2). Fermentation proceeded to

completion for both low and medium sugar fermentation successfully. On the other hand, fa

sluggish fermentation was obseved for high sugar fermentation. The rate of decrease reduced

dramatically after the first 11 days of fermentation. This is consistent with many studies that

have indicated the likelihood of sluggish/stuck fermentations occurring in high density musts

(juices). This has implications for the wine quality as the high residual sugar content could

1

2

3

4

5

6

7

8

9

10

0 10 20 30

Log

CF

U/m

L

Time (Day)

0

5

10

15

20

25

30

0 10 20 30

TS

S (°B

)

Time (Days)

Page 46: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

35

suppport the growth of undesirable spoilage microorganisms that may result in detrimental

wine quality or result in a sweet wine.

From Day 0 to Day 3, an increase in fructose was observed in the low and high sugar

fermentations, with an increase in glucose also observed in the low sugar fermentation. After

which, all sugars decreased throughout the fermentation. On the other hand, there was no

increase in any of the sugars quantified for the medium sugar fermentation. The rate of

decline in sugars was generally inversely proportional to initial sugar concentration. It was

also observed that glucose was utilised more rapidly than fructose and sucrose. This is

consistent with the available data claiming that S. cerevisiae strains are mostly glucophillic in

nature (Berthels et al. 2004). Available literature data also suggests that fructose utilisation

was inhibited more than glucose in the presence of high ethanol content (Berthels et al. 2004).

This was also observed in this study where there was lesser glucose and more fructose

present after the high sugar fermentation than the medium sugar fermentation (Table 3.2).

The increase in fructose and glucose, together with the decrease in sucrose, is likely due

to the enzymatic activity of S. cerevisiae which produces invertase that converts sucrose into

glucose and fructose (Berthels et al. 2004). It might be possible that similar invertase activity

occurred in the medium sugar fermentation but the higher sugar concentration (relative to low

sugar fermentation) could have led to invertase suppression (Myers et al. 1997), leading to

lesser amounts of fructose and glucose production. Although this appears to contradict the

data for high sugar fermentation, it can be easily explained that the high sugar concentration

affected the normal metabolic functioning of the yeast cells, leading to less efficient sugar

consumption (Arrizon and Gschaedler 2002).

Page 47: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

36

Figure 3.2 Fructose ( ), glucose ( ) and sucrose ( ) consumption

kinetics of S.cerevisiae during fermentation

3.2.4 Glycerol production

The amount of glycerol produced at the end of each fermentation increased with

increasing sugar content (Figure 3.3). The final glycerol concentrations in the low, medium

and high sugar fermentations were 1.34 g/100 mL, 1.45 g/100 mL and 1.59 g/100 mL

0

4

8

12

0 10 20 30

Su

gar

con

ten

t (g

/10

0 m

L)

Time (Days)

Low sugar fermentation

0

5

10

15

20

0 10 20 30

Su

gar

con

ten

t (g

/10

0 m

L)

Time (Days)

Medium sugar fermentation

0

5

10

15

20

25

0 10 20 30

Su

gar

con

ten

t (g

/10

0 m

L)

Time (Days)

High sugar fermentation

Page 48: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

37

respectively. Glycerol can affect perceived wine quality positively by improving the

mouthfeel of the wine and may have minor contributions to the sweetness of the wine, this

effect is concentration dependent.

The usual amount of glycerol in wines ranges from 4 to 9 g/L (or 0.4 to 0.9 g/100 mL)

(Remize et al. 1999). Since yeast strain is the strongest influencing factor on the amount of

glycerol produced, it is likely that MERIT.ferm is a high glycerol producing strain and has a

potential to improve wine quality. Other factors that influence glycerol production include

sulfite concentration, pH, fermentation temperature, aeration, inoculation level, sugar

concentration and nitrogen content (Radle and Schülz 1982).

Glycerol is an osmoprotectant naturally synthesized by yeasts in response to the high

concentration of osmotically active substances (especially glucose and fructose) present in

the must or juice. The hypertonic conditions experienced by the yeast cells lead to an efflux

of water from the cell, diminished turgor pressure and reduced water availabilty (Scanes et al.

1998). The induction of glycerol biosynthesis results in the accumulation of glycerol inside

the cell for the equilibration of the osmotic pressure between the intracellular and

extracellular environment o re-establish cell turgor pressure (Hohmann, 1997).

Page 49: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

38

Figure 3.3 Concentrations of glycerol in mango wines for low (16.6oBrix),

medium (23oBrix) and high sugar (30

oBrix) fermentation

ab ANOVA (n=4) at 95% confidence level with same letters in the same row

indicating no significant difference

a, b

b

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

1.60

1.80

Low Medium High

Gly

cero

l (g

/10

0m

L)

a

Page 50: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

39

3.2.5 Effects of initial sugar concentration on volatile production

The volatile profile of mango wine was significantly different from the fresh mango

juice. A number of volatiles detected in the fresh mango juice were not detected in the final

mango wine. These catabolised compounds include a number of esters, aldehydes, ketones,

furans and terpene hydrocarbons. The complete list of catabolised compounds can be found

in Table 3.3 below. The absence of some significant character impact odourants suggests that

the mango wine produced had a different flavour profile from the mango juice.

Table 3.3 Volatiles in mango juice catabolised during fermentation

Class Compounds

Alcohol trans-2-Hexenol, 2-ethylhexanol, para−cresol, β-

fenchol, endo-borneol

Ester Hexyl formate, cis-3-hexenyl isobutyrate, cis-3-

hexenyl propionate, cis-3-hexenyl-3-butenoate,

butyl butanoate

Aldehyde trans-2-Hexenal, 3-methylbenzaldehyde, hexanal

Ketone 5-Methyldihydro-2(3H)-furanone, γ-hexalactone,

4-methyl-2-heptanone, mesifurane

Furan 2-ethylfuran, 2,3-dihydro-2-methyl-benzofuan

Terpene

hydrocarbon

α-Terpinene, cis-ocimene, β-phellandrene, α-

thujene, α-pinene, allo-ocimene, γ-selinene, α-

selinene, trans-farnesene, valencene

A total of 11 alcohols, 35 esters, 3 aldehydes, 3 ketones, 8 terpenes and 6 organic acids

were detected in the final mango wines (Table 3.4). The FID peak area was used to compare

volatile compounds that were not quantified, since FID peak area may semi-quantitatively

represent the amount of different volatiles (Alves et al. 2010; Lee et al. 2010; Trinh et al.

2010). The relative peak area (RPA) is the area under each peak expressed as a percentage of

the total peak area of all the compounds detected

Page 51: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

40

Table 3.4 Volatile composition of mango wines with different initial sugar concentrations CAS

1 LRI

2 Compound Low Medium High Odour descriptors

3

Peak Area RPA (%) Peak Area RPA Peak Area RPA

000064-17-5 1028 Ethanol 820.9 ± 11.55a 67.89 1115.94 ± 81.93

b 87.43 1515.94 ± 81.93

c 91.24 Alcoholic

Subtotal 820.9 67.89 1115.94 87.43 1515.94 91.24

000078-83-1 1172 Isobutyl alcohol 10.63 ± 0.69 a 0.8792 5.03 ± 0.08

b 0.416 6.11 ± 0.12

c 0.3677 Citrus, orange, floral, rose

Alcohol 000123-51-3 1237 Isoamyl alcohol 48.55 ± 2.75 a 4.0154 46.79 ± 0.25

a 3.8699 51.52 ± 0.57

b 3.1009 Ethereal, winey

000111-27-3 1448 Hexanol 0.65 ± 0.03 a 0.0538 0.31 ± 0.01

b 0.0243 0.34 ± 0.02

b 0.0205 Pungent, ethereal, fruity

and alcoholic

000928-96-1 1475 cis-3-Hexenol 2.76 ± 0.13 a 0.2283 0.16 ± 0.01

b 0.0132 0.09 ± 0.01

c 0.0054 Green, grassy

000505-10-2 1706 Methionol 0.78 ± 0.03 a 0.0645 0.29 ± 0.01

b 0.024 0.2 ± 0.01

c 0.0125 Cooked potato-like

000470-08-6 1794 β-Fenchyl alcohol 0.55 ± 0.01 a 0.0455 0.04 ± 0.003

b 0.0033 0.036 ± 0.006

b 0.0024 Camphor, pine, woody

000106-24-1 1861 Geraniol 0.23 ± 0.002 a 0.019 0.03 ± 0.002

b 0.0025 0.02 ± 0002

b 0.0012 Fusel, etherial, cognac,

fruity

000106-22-9 1867 Citronellol 1.1 ± 3.32 a 0.091 0.57 ± 0.03

b 0.0471 0.56 ± 0.01

b 0.0337 Floral, rosy, sweet, citrus

000060-12-8 1964 2-Phenylethyl alcohol 54.82 ± 3.52 a 4.534 0.66 ± 0.04

b 0.0546 0.78 ± 0.02

b 0.0482 Sweet, floral, rose-like,

honey

000078-70-6 1999 Linalool 0.71 ± 0.01 a 0.0587 0.21 ± 0.02

b 0.0174 0.2 ± 0.01

b 0.0126 Mild floral

Subtotal 120 9.99 53.5 4.47 59.36 3.59

Ethyl ester 000141-78-6 1009 Ethyl acetate 1.23 ± 0.09 a 0.1017 0.91 ± 0.02

b 0.0713 0.15 ± 0.01

c 0.009 Ethereal, fruity, sweet

000105-54-4 1034 Ethyl butyrate 15.04 ± 0.56 a 1.2439 1.43 ± 0.06

b 0.112 1.02 ± 0.08

c 0.0614 Sweet, fruity, lifting

000123-66-0 1297 Ethyl hexanoate 2.52 ± 0.17 a 0.2084 7.46 ± 0.18

b 0.5845 3.23 ± 0.13

c 0.1944 Sweet, pineapple, fruity,

waxy

64187-83-3 1358 Ethyl cis-3-hexenoate 0.13 ± 0.02 a 0.0108 0.15 ± 0.01

a 0.0118 0.12 ± 0.01

a 0.0072 Fruity, sweet, apple,

ethereal

000106-30-9 1369 Ethyl heptanoate 0.09 ± 0.001 a 0.0074 0.11 ± 0.01

a 0.0086 0.09 ± 0.01

a 0.0054 Sweet, waxy, soapy

054653-25-7 1392 Ethyl 5-hexenoate 3.28 ± 0.12 a 0.2713 0.48 ± 0.02

b 0.0376 0.44 ± 0.02

b 0.0265 Fruity, pineapple

000106-32-1 1453 Ethyl octanoate 22.31 ± 1.97 a 1.8452 1.59 ± 0.11

b 0.1246 0.93 ± 0.04

c 0.056 Waxy, sweet, musty, fruity

35194-38-8 1486 Ethyl 7-octenoate 8.74 ± 0.48 a 0.7229 1.66 ± 2.16

b 0.1301 0.32 ± 0.01

c 0.0193 Tropical, fruity

000123-29-5 1624 Ethyl nonanoate 1.79 ± 0.1 a 0.148 0.15 ± 0.01

b 0.0118 0.09 ± 0.01

c 0.0054 Waxy, soapy, cognac

000110-38-3 1746 Ethyl decanoate 132.33 ± 2.82 a 10.9447 0.31 ± 0.02

b 0.0243 0.23 ± 0.01

c 0.0138 Sweet, fruity, pineapple

067233-91-4 1795 Ethyl 9-decenoate 8.83 ± 0.22 a 0.7303 0.74 ± 0.04

b 0.058 0.61 ± 0.02

c 0.0367 Fruity, fatty

Page 52: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

41

Table 3.4 (Continued)

000106-33-2 1887 Ethyl dodecanoate 16.12 ± 1.57 a 1.3233 1.02 ± 0.05

b 0.0799 0.84 ± 0.04

c 0.0506 Sweet, waxy, fruity

000124-06-1 2201 Ethyl tertradecanoate 0.51 ± 0.01 a 0.0422 0.15 ± 0.02

b 0.0118 0.1 ± 0.01

c 0.006 Sweet, waxy, creamy

000628-97-7 2373 Ethyl hexadecanoate 2.51 ± 0.2 a 0.2076 0.28 ± 0.01

b 0.0219 0.19 ± 0.03

c 0.0114 Waxy, fruity, creamy

054546-22-4 2402 Ethyl 9-hexadecenoate 3.52 ± 0.08 a 0.2911 13.25 ± 1.09

b 1.0381 10.26 ± 0.65

c 0.6175 -

Subtotal 218.83 18.1 29.69 2.33 18.62 1.12

Acetate ester 000109-60-4 1002 Propyl acetate 0.75 ± 0.05 a 0.062 0.61 ± 0.04

b 0.0478 0.44 ± 0.03

c 0.0265 Fusel, sweet, fruity

000110-19-0 1020 Isobutyl acetate 1.18 ± 0.09 a 0.0976 0.84 ± 0.06

b 0.0658 0.6 ± 0.04

c 0.0361 Sweet, fruity, banana,

tropical

000123-86-4 1061 Butyl acetate 0.65 ± 0.04 a 0.0538 0.52 ± 0.04

b 0.0407 0.42 ± 0.01

c 0.0253 Ethereal, solvent, fruity,

banana

000628-63-7 1097 Active amyl acetate 0.18 ± 0.02 a 0.0149 0.13 ± 0.01

b 0.0102 0.1 ± 0.02

c 0.006 Ethereal, fruity, banana

000123-92-2 1112 Isoamyl acetate 0.48 ± 0.03 a 0.0397 1.25 ± 0.06

b 0.0979 0.91 ± 0.09

c 0.0548 Sweet, fruity, banana,

solvent

003681-71-8 1324 cis-3-Hexenyl acetate 0.07 ± 0.01 a 0.0058 0.29 ± 0.02

b 0.0227 0.26 ± 0.03

b 0.0156 Fresh, green, sweet, fruity

2497-18-9 1367 2-Hexenyl acetate 0.11 ± 0.01 a 0.0091 0.08 ± 0.01

b 0.0063 0.06 ± 0.01

b 0.0036 Sweet, leafy green

000142-92-7 1411 n-Hexyl acetate 0.33 ± 0.02 a 0.0273 0.24 ± 0.02

b 0.0188 0.19 ± 0.01

c 0.0114 Fruity, green, apple,

banana, sweet

000150-84-5 1659 Citronellyl acetate 0.69 ± 0.07 a 0.0571 0.52 ± 0.04

b 0.0407 0.39 ± 0.03

c 0.0235 Floral, green rose, fruity

16630-55-0 1676 Methionyl acetate 0.21 ± 0.01 a 0.0174 0.17 ± 0.01

b 0.0133 0.13 ± 0.01

c 0.0078 Fatty, ester

000141-12-8 1753 Neryl acetate 3.69 ± 0.09 a 0.3052 2.81 ± 0.07

b 0.2202 2.23 ± 0.06

c 0.1342 Floral, rose, soapy, citrus

000105-87-3 1790 Geranyl acetate 2.56 ± 0.12 a 0.2117 2.07 ± 0.1

b 0.1622 1.68 ± 0.08

c 0.1011 -

000103-45-7 1862 2-Phenylethyl acetate 2.33 ± 0.15 a 0.1927 1.42 ± 0.08

b 0.1113 1.03 ± 0.04

c 0.062 Sweet, honey, floral, rosy

Subtotal 13.23 1.0942 10.95 0.8579 8.44 0.508

Higher ester 005461-06-3 1642 Isobutyl octanoate 2.23 ± 0.28 a 0.1844 1.68 ± 0.21

b 0.1316 1.27 ± 0.16

c 0.0764 Fruity, green, oily, floral

002035-99-6 1762 Isoamyl octanoate 1.42 ± 0.05 a 0.1174 2.05 ± 0.06

b 0.1606 1.45 ± 0.26

c 0.0873 Sweet, fruity, waxy, green

30673-60-0 1803 Propyl decanoate 0.73 ± 0.03 a 0.0604 0.58 ± 0.02

b 0.0454 0.46 ± 0.02

c 0.0277 Waxy, fruity, fatty, green

030673-38-2 1859 Isobutyl decanoate 0.53 ± 0.02 a 0.0438 0.45 ± 0.02

b 0.0353 0.38 ± 0.02

c 0.0229 Oily, sweet, brandy,

apricot, cognac

000103-52-6 1907 2-Phenylethyl

butanoate

1.64 ± 0.12 a 0.1356 1.33 ± 0.1

b 0.1042 1.08 ± 0.08

c 0.065 Fruity, floral, green, winey

Page 53: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

42

Table 3.4 (Continued)

002306-91-4 1973 Isoamyl decanoate 2.44 ± 0.06 a 0.2018 1.25 ± 0.17

b 0.0979 0.95 ± 0.06

c 0.0572 Waxy, fruity, sweet,

cognac

006309-51-9 2180 Isoamyl dodecanoate 0.36 ± 0.02 a 0.0298 3.98 ± 0.12

b 0.3118 2.62 ± 0.08

c 0.1577 Cognac, green, waxy

Subtotal 9.35 0.7733 11.32 0.8869 8.21 0.4941

Aldehyde 000075-07-0 939 Acetaldehyde 1.33 ± 0.0012a 0.11 3.29 ± 0.17

b 0.2578 4.18 ± 0.25

c 0.2516 Green, cherry

000100-52-7 1637 Benzaldehyde 0.11 ± 0.01 a 0.0091 0.14 ± 0.01

b 0.011 0.12 ± 0.01

a 0.0072 Almond, fruity, nutty

000104-87-0 1773 p-Tolualdehyde 0.26 ± 0.0011a 0.0215 0.22 ± 0.036

b 0.0172 0.19 ± 0.03

b 0.0114 Pungent, fresh, green

Subtotal 1.7 0.1406 3.65 0.286 4.49 0.2702

Ketone 000821-55-6 1398 2-Nonanone 0.23 ± 0.02 a 0.019 0.05 ± 0.006

b 0.0039 0.05 ± 0.003

b 0.003 Fruity, sweet, waxy,

soapy, cheese

000513-86-0 1401 Acetoin 0.01 ± 0.001 a 0.001 0.01 ± 0.001

a 0.001 0.01 ± 0.001

a 0.001 Sweet, buttery, creamy

023696-85-7 1938 β-Damascenone 0.25 ± 0.02 a 0.0207 33.88 ± 3.16

b 2.6544 28.57 ± 0.88

c 1.7196 Woody, sweet, fruity,

earthy, floral

Subtotal 0.48 0.0397 33.93 2.6583 28.62 1.7226

Terpene 095327-98-3 1254 Limonene 0.01 ± 0.003 a 0 0.22 ± 0.04

b 0.0172 0.16 ± 0.01

c 0.0096 Citrus, terpenic, orange

note

000099-85-4 1305 γ-Terpinene 0.1 ± 0.02 a 0.0083 0.15 ± 0.11

b 0.0118 0.19 ± 0.09

b 0.0114 Sweet, citrus

000099-87-6 1339 p-Cymene 1.37 ± 0.31 a 0.1133 1.02 ± 0.03

b 0.0799 1.01 ± 0.03

b 0.0608 Citrus, terpenic, woody

000586-62-9 1352 α-Terpinolene 0.31 ± 0.02 a 0.0256 0.09 ± 0.001

b 0.0071 0.05 ± 0.01

c 0.003 Citrus, lime, pine

000087-44-5 1695 β-Caryophyllene 0.4 ± 0.01 a 0.0331 0.11 ± 0.02

b 0.0086 0.12 ± 0.01

b 0.0072 Sweet, woody, spice,

clove

006753-98-6 1725 α-Caryophyllene 3.83 ± 4.91 a 0.3168 1.83 ± 0.18

b 0.1434 1.7 ± 0.1

c 0.1023 Woody

017066-67-0 1829 β-Selinene 0.49 ± 0.01 a 0.0405 0.98 ± 0.05

b 0.0768 0.92 ± 0.03

b 0.0554 Mint, turpentine

Subtotal 6.5 0.5376 4.4 0.3447 4.15 0.2498

Organic acid 000064-19-7 1549 Acetic acid 6.54 ± 0.3 a 0.5409 7.53 ± 0.01

b 0.5899 8.26 ± 0.03

c 0.4972 Sharp, pungent, sour,

vinegar 000142-62-1 1890 Hexanoic acid 0.73 ± 0.06

a 0.0604 0.69 ± 0.11

b 0.0541 0.66 ± 0.11

b 0.0397 Sour, fatty, sweat, cheese

000124-07-2 2170 Octanoic acid 4.92 ± 0.15 a 0.4069 0.71 ± 0.02

b 0.0556 0.6 ± 0.01

b 0.0361 Fatty, waxy, rancid

Page 54: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

43

Table 3.4 (Continued)

000334-48-5 2390 Decanoic acid 1.63 ± 0.7 a 0.1348 0.19 ± 0.01

b 0.0149 0.18 ± 0.01

b 0.0108 Rancid, sour, fatty

004436-32-9 2474 9-Decenoic acid 1.59 ± 0.06 a 0.1315 1.08 ± 0.08

b 0.0846 1.03 ± 0.06

c 0.062 Waxy, green, fatty, soapy

000143-07-7 2544 Dodecanoic acid 2.68 ± 0.19 a 0.2217 2.8 ± 0.17

a 0.2194 2.91 ± 0.12

a 0.1751 Fatty, waxy

Subtotal 18.09 1.4962 13 1.0185 13.64 0.821

Grand total 1209.08 100 1276.38 100 1661.47 100

1 CAS numbers were obtained from Wiley library

2 LRI of all the relative tables was determined on the DB-FFAP column, relative to C5-C40 hydrocarbons.

3 Descriptors were retrieved from http://www.thegoodscentscompany.com

abc ANOVA at 95% confidence level, same letters in the same row indicate no significant difference

Page 55: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

44

Table 3.5 Selected volatiles quantified in mango wine and their odour activity values (OAV) in low (unfortified mango juice 16.6oB), medium (initial TSS of

23°B) and high (initial TSS of 30°B) sugar fermentation

CAS1 LRI

2 Compounds Low

(mg /L)

OAV Medium

(mg /L)

OAV High

(mg /L)

OAV Odour

threshold

Organoleptics3

Alcohol 000078-83-1 1172 Isobutyl alcohol 62.3±8.3a 1.56 31.9±6.2

b 0.80 39.1±7.3

c 0.98 40 Fruity, wine-like

000123-51-3 1237 Isoamyl alcohol 107.2±10.3 a 3.57 387.1±12.2

b 12.90 414.4±15.6

c 13.81 30 Alcoholic, fruity, banana-like

000060-12-8 1964 2-Phenylethyl alcohol 56.0±7.2 a 5.60 2.26±0.15

b 0.23 2.49±0.18

c 0.25 10 Rose, floral

Ester 000141-78-6 1009 Ethyl acetate 3.93±1.19 a 0.52 2.34±0.82

b 0.31 0.56±0.15

c 0.07 7.5 Pineapple, fruity, varnish

000110-19-0 1020 Isobutyl acetate 0.042±0.012 a 0.03 0.025±0.092

b 0.02 0.013±0.006

c 0.01 1.6 Fruity, sweet, apple

000123-92-2 1112 Isoamyl acetate 0.258±0.083 a 8.60 0.7847±0.11

b 26.16 0.5529±0.12

c 18.43 0.03 Fruity, banana, sweet

000103-45-7 1862 2-Phenylethyl acetate 0.764±0.16a 3.06 0.423±0.051

b 1.69 0.341±0.043

c 1.36 0.25 Floral, rose, sweet

000123-66-0 1297 Ethyl hexanoate 0.301±0.025 a 21.50 0.792±0.062

b 56.57 0.375±0.028

c 26.79 0.014 Sweet, pineapple, waxy

000106-32-1 1453 Ethyl octanoate 1.79±0.25a 895.00 0.4481±0.052

b 224.05 0.3865±0.042

b 183.25 0.002 Floral, fruity, brandy

000110-38-3 1746 Ethyl decanoate 1.09±0.23 a 5.45 0.0615±0.008

b 0.31 0.0605±0.005

b 0.30 0.2 Waxy, sweet, apple

000106-33-2 1887 Ethyl dodecanoate 1.75±0.45 a 0.30 0.706±0.056

b 0.12 0.692±0.052

b 0.12 5.9 Soapy, waxy, floral

Acid 000064-19-7 1549 Acetic acid 442.25±59.1 a 2.21 487.69±56.2

b 2.44 538.63±66.5

c 2.69 200 Vinegar, pungent

000142-62-1 1890 Hexanoic acid 1.81±0.21 a 0.60 1.67±0.21

b 0.56 1.57±0.16

b 0.52 3 Fatty, soapy, sour

000124-07-2 2170 Octanoic acid 4.64±0.76 a 0.53 1.18±0.14

b 0.13 1.08±0.09

b 0.12 8.8 Fatty, soapy, sour, fruity

000334-48-5 2390 Decanoic acid 2.16±0.25 a 0.22 1.81±0.14

b 0.18 1.79±0.12

b 0.18 10 Fatty, rancid, sour

000143-07-7 2544 Dodecanoic acid 0.6339±0.05 a 0.06 0.6402±0.081

b 0.06 0.6453±0.093

b 0.06 10 Fatty, waxy

1 CAS numbers were obtained from Wiley library

2 LRI of all the relative tables was determined on the DB-FFAP column, relative to C5-C40 hydrocarbons.

3 Descriptors were retrieved from http://www.thegoodscentscompany.com

abc ANOVA at 95% confidence level, same letters in the same row indicate no significant difference

Page 56: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

45

3.2.5.1 Ethanol

Ethanol accounted for 67.89%, 87.43% and 91.24% (by RPA) of all volatiles produced

for the low (unfortified, 16.6oB), medium (23°B) and high (30°B) sugar fermentations

respectively (Table 3.4). At low levels, ethanol may enhance the sensory perception of

aroma compounds. However, in excess, ethanol has a masking effect and can directly lead to

a burning sensation (Swiegers et al., 2005). This may be especially prominent in the high

sugar fermentation resulting in a detrimental effect on wine quality by masking the other

volatiles present with ethanol accounting for more than 90% of all volatiles present.

At a cellular level, ethanol toxicity can lead to the inhibiton of volatile production by

affecting the transport of sugar and nitrogen into the cell due to its effects on membrane

permeability and may reduce proton motive force (Bisson and Karpel 2010; Fleet and Heard,

1993). This appears to be the case where volatile production was significantly lower in high

sugar fermentation. It has been suggested that oxygenation and increasing assimilable

nitrogen content can be used to reduce the effects of ethanol toxicity for the production of a

high ethanol wine (Casey and Ingledew 1986) since oxygenation allows for the formation of

unsaturated fatty acids which decreases membrane fluidity caused by high ethanol content

(Kyung et al. 2003).

3.2.5.2 Fusel alcohols

Fusel alcohols are important components of the wine bouquet due to their organoleptic

properties. The major fusel alcohols produced were 2-phenylethyl alcohol, isoamyl alcohol

and isobutyl alcohol. Most fusel alcohols detected were present at the highest concentrations

in the low sugar fermentation (Table 3.4).

2-Phenylethyl alcohol and isobutyl alcohol were found at the highest concentrations in

the low sugar fermentation. Both higher alcohols were present at levels above their odour

Page 57: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

46

thresholds in the low sugar fermentation; therefore, they are likely to have a positive impact

on the wine flavour. However, these two alcohols were present at concentrations below their

odour thresholds in the medium and high sugar fermentations.

Isoamyl alcohol was found at levels above its odour threshold in all three mango wines,

with the highest concentration in the high sugar fermentation and the lowest in the low sugar

fermentation. Although isoamyl alcohol has a pleasant fruity and sweet odour, the excessive

amounts (387.1 and 414.1 mg/L respectively) detected in the medium and high sugar

fermentations may have an undesirable effect on wine aroma. It has been reported that fusel

alcohols can be detrimental to wine aroma at levels above 350 mg/L (Swiegers et al. 2005).

The reason for this abnormality is not clear, but it may be due to the presence of leucine (one

of the major amino acids in mango juice) and the redox imbalance presented due to the

hypersomotic stress. Furthermore, isoamyl alcohol can also be synthesized from sugar

metabolism via pyruvate (which can be converted to leucine ) (Kohlhaw 2003). The higher

amounts of isoamyl alcohol present in the high sugar fermentation is highly likely due to the

increased sugar metabolism producing large amounts of pyruvate which were consequently

converted to isoamyl alcohol. Previous studies have shown that high sugar concentrations

favour the formation of fusel alcohols rather than the corresponding carboxylic acid (Jain et

al. 2011; Styger et al. 2013) due to the regeneration of NADH from NAD+. The production

of leucine from pyruvate also regenerates NADH from NAD+ (Kohlhaw 2003)

Other than their contribution to the wine bouquet, higher alcohols also serve as

important precursors for the formation of branch-chained esters which are beneficial to wine

flavour. In this study, the amounts of higher alcohols were within the desirable range and

likely to make positive contributions to the overall mango wine flavour profile.

Page 58: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

47

3.2.5.3 Terpene alcohols

A number of terpene alcohols were also detected in all three mango wines (Table 3.4).

Citronellol, geraniol and linalool were detected but at levels below their quantification limits.

Hence, their contribution to the overall wine flavour profile would be likely due to their

synergistic effects. Citronellol also increased after fermentation, indicating its biosynthesis

of S. cerevisiae (Carrau et al. 2005) or release from the bound terpenols (glycosides) due to

enzymatic activity (Gunata et al. 1986). S. cerevisiae is able to convert geraniol to citronellol

(Carrau, et al. 2005). This may explain the low amounts of geraniol present in the mango

wines.

3.2.5.4 Esters

Ethyl esters were the major esters present in mango wine. In low sugar fermentation, ,

ethyl esters made up 18.8% of all volatiles (by RPA). The amount of ethyl esters detected in

the medium sugar and high sugar fermentation was significantly lower, only 2.23% and 1.13%

(by RPA) of total volatile production respectively.

In terms of potential impact on the wine flavour profile, ethyl octanoate is likely to

have the most significant influence on wine flavour due to its low odour threshold, despite

not being produced at an exceptionally high level (Table 3.5). Ethyl hexanoate was also

present at levels above its odour threshold in all three mango wines and detected in the

highest concentration in the medium sugar fermentation at 0.792 mg/L (Table 3.5).

Ethyl esters are likely to be produced enzymatically during the synthesis or degradation

of fatty acids (Alves et al. 2010) and/or chemical reactions that occur in the presence of

alcohol with their prevalence attributed to the high production of medium-chain fatty acids in

S. cerevisiae yeasts (Saerens et al. 2008). Ethyl esters are produced by transferase reactions

in which alcohols react with fatty acyl-CoAs derived from the metabolism of fatty acids. It

Page 59: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

48

has been shown that S. cerevisiae exhibited enzymatic activity required for the synthesis of

medium-chain fatty acid ethyl esters (Saerens et al. 2008).

The major acetate esters produced were 2-phenylethyl acetate, isoamyl acetate, geranyl

acetate and neryl acetate (Table 3.4). Acetate esters confer fruity and floral notes and are

important to wine flavour. Acetate esters are produced from the reaction of acetyl-CoA with

alcohols (Perestrelo et al. 2006) and thus, the higher production of acetates in the low sugar

fermentation may be due to the higher quantities of branched-chain fusel alcohols present.

With the exception of isoamyl acetate, all major acetate esters were found in the highest

concentration in the low sugar fermentation. Isoamyl acetate and 2-phenylethyl acetate were

both detected at levels above their odour thresholds and may impart fruity and floral notes to

the wine flavour. Isoamyl acetate was found in the highest amount in the high sugar

fermentation, similar to isoamyl alcohol. This is likely due to the higher amounts of isoamyl

alcohol as a precursor present in the high sugar fermentation.

3.2.5.5 Organic acids

The main acid produced in all three treatments was acetic acid which made up more

than 90% of all acids formed, making it main determining factor of volatile acidity. Acetic

acid exerts a detrimental effect on wine quality if present at excessive levels. The amount of

acetic acid produced in the low, medium and high sugar fermentations were 442, 487 and 538

mg/L respectively; this translates to OAV of 2.21, 2.44 and 2.69 respectively. Studies have

shown that S. cerevisiae produces relatively small quantities of acetic acid in must with a

moderate sugar concentration (less than 220 g/L) (Bely et al. 2003), therefore, the relatively

high levels of acetic acid produced was likely due to the high sugar concentrations.

Page 60: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

49

Other acids that were detected include hexanoic acid, octanoic acid, decanoic acid, 9-

decenoic acid and dodecanoic acid. These short- and medium-chain fatty acids may

contribute to undesirable fatty, rancid and soapy off-odours. The amounts of short- and

medium-chain fatty acids produced were all well below their odour thresholds (Table 3.5)

and were unlikely to impart negative flavour notes but may impart some desired complexity

to the wine flavour. With the exception of acetic acid, most of the other organic acids were

found in the highest amount in the low sugar fermentation.

During normal alcoholic fermentation, the pyruvate produced during glycolysis is

decarboxylated to form acetaldehyde, which is then reduced to ethanol, with the production

of NAD+. However, NAD

+ is generated when glycerol is produced by yeast cells under

hyperosmotic conditions, resulting in an excess of NAD+ (with a deficit of NADH). To

rectify this redox imbalance, acetaldehyde is converted into acetic acid and this biochemical

reaction reduces NAD+ to NADH instead (Figure 3.4). Consequently, more acetic acid was

produced in the mango wine with a higher starting sugar content. This is supported by the

fact that glycerol was produced in higher amounts in the medium and high sugar

fermentations.

Page 61: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

50

Figure 3.4 Metabolic pathways governing the production of ethanol, glycerol and acetic

acid

3.2.5.5 Carbonyl compounds

The major aldehydes detected in mango wine were acetaldehyde, p-tolualdehyde and

benzaldehyde with acetaldehyde being the main aldehyde. Acetaldehyde production was

directly correlated to the initial sugar concentration. This can be explained by the redox

imbalance due to the high sugar concentration. As mentioned previously, acetaldehyde is

oxidised to acetic acid (instead of reduced to ethanol) for NADH regeneration when excess

glycerol is produced under hyperosmotic conditions (Figure 3.4). However, due to the

excessive amounts of glycerol produced, a shortage of NADH occurs, causing part of the

acetaldehyde to be excreted by S.cerevisiae into the wine. This is further supported by the

fact that acetaldehyde and glycerol concentents were directly correlated.

Glyceraldehyde-3-phosphate

Pyruvate

Acetaldehyde

Ethanol

Glucose

Fructose-1,6-Bisphosphate

ATP

ADP

NAD+

NADH

Acetic acid

NAD+ NADH

NADH

NAD+

Dihydroxyacetone phosphate

Glycerol-3-phosphate

Glycerol-3-phosphate

Glycerol

NADH

NAD+

Page 62: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

51

Acetaldehyde imparts a fresh, fruity aroma to wines at low levels. However, at higher

concentrations, acetaldehyde results in green apple like, even a pungent irritating odour

(Miyake and Shibamoto 1993). Other than its impact on wine quality due to its organoleptic

properties, acetaldehyde is a precursor for acetoin (Romano et al. 1994).

The ketones detected in mango wine were β-damascenone, 2-nonanone and acetoin,

with β-damascenone. β-Damascenone was the major ketone and one of the few compounds

detected in both fresh mango juice and mango wine; and it decreased during fermentation.

Although these ketones had low concentrations, they may contribute to “floral” or “fruity”

aroma to mango wines synergistically. In addition, it was found at greater concentrations in

the medium and high sugar fermentation than in low sugar fermentation. This appears to

suggest that the increased hyperosmotic stress may have disrupted the normal metabolic

activities of the yeast, resulting in decreased degradation of certain compounds.

3.2.5.6 Terpene hydrocarbons

Despite being the predominant class of volatiles in fresh mango juice, the terpenes

made up less than 1% (by RPA) of all volatiles present in mango wine. Most of these

compounds were catabolised during the fermentation process (Table 3.1). This suggests that

the characteristic flavour of the fresh mango juice may not be present in the mango wine.

The major terpene hydrocarbon detected in the mango wine was α-caryophyllene, an

isomer of β-caryophyllene (El Hadri et al. 2010), a terpene found in fresh mango juice;

therefore, this suggests that some form of biotransformation occurred during the fermentation

process. In addition, although not quantified, the peak area of α-caryophyllene was about 10

times that of β-caryophyllene. This suggests that some form of biosynthesis of α-

caryophyllene occurred during the fermentation process. Although this compound is a

quantitatively minor compound, it is of interest to note that studies have shown that many of

Page 63: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

52

the monoterpenes and sequisterpenes, i.e. the hydrocarbon terpenes, have been linked to

potential health benefits (El Hadri et al. 2010; Buchbauer and Ilic 2013). Some of the other

terpene hydrocarbons linked to possible health benefits include limonene, β-caryophyllene

and myrcene, all of which are detected in mango wine (Buchbauer and Ilic 2013). Therefore,

reducing the degradation of these terpene hydrocarbons may not only improve the flavour

profile of the mango wine but also potentially harness some of the health benefits of mango

wine if consumed in adequate amounts.

3.2.6 Effects of redox potential on overall wine quality

Mouthfeel and flavour complexity are two of the most important factors affecting

perceived wine quality, and they are influenced by glycerol and volatile composition

respectively. In turn, hyperosmotic stress influences wine quality due to the corrective

mechanisms undertaken by the yeast cells for survival by producing glycerol. Therefore, the

balance of NAD+ and NADH can influence the volatile and glycerol production during

fermentation and consequently affect wine quality. However, as each yeast strain may have

different metabolic capabilities, it may be necessary for more in-depth studies to be

conducted on S. cerevisiae MERIT.ferm to fully maximise its potential for the fermentation

of a mango wine.

Metabolically, glycerol is primarily produced for the maintenance of intracellular redox

balance, by converting the excess NADH generated during biomass formation to NAD+

(Hohmann 1997; Remize et al. 1999). Hence, when excess glycerol is produced under

hyperosmotic conditions to prevent the exodus of water from the cell, an excess of NAD+

arises, leading to imbalance. Acetaldehyde is very toxic to cells and excess of it must be

excreted. While an increase in glycerol concentration may improve wine quality by

enhancing the mouthfeel of the wine, this is coupled to an increase in acetic acid and

Page 64: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

53

acetaldehyde production and hence may result in an undesirable vinegar-like character if

acetic acid is produced at excessive levels (Swiegers and Pretorius, 2005).

On the other hand, excess NADH promotes production of fusel alcohol and hinders the

production of esters due to the need for precursor fusel acids which requires NAD+

Therefore, it has been hypothesized that the imbalance in NAD

+/NADH levels, due to

hyperosmotic conditions generally drove the production of higher alcohols in an attempt to

reduce NADH to NAD+ (Jain et al. 2011; Styger et al. 2013). Furthermore, the production of

a fusel alcohol versus fusel acids is therefore also dependent on the redox requirements of the

yeast (Bisson and Karpel, 2010).

In addition, despite the stuck/laggish fermentation in the high sugar juice samples, the

final ethanol content obtained was still higher than that of the other two fermentations. This

indicates that, with the right conditions, the production of mango wines with high alcohol

content is possible. Several studies have suggested that supplementing a high sugar

fermention with amino acids and nitrogenous compounds reduces the risks of a stuck

fermentation (Arrizon and Gschaedler, 2002; Alexandre, Rousseauz and Charpentier, 1994;

Chaney et al., 2006).

3.3 Conclusion

In conclusion, the initial sugar content in the mango juice had a dramatic effect on the

volatile production and fermentation kinetics: Generally, the higher the initial sugar

concentration, the more significant the hyperosmotic stress the starting material presented to

the yeast. This hyperosmotic stress affected the normal metabolism of the cell and

consequently also has a negative effect on the production of volatile compounds. It could be

partly due to the redox imbalance hyperosmotic stress presented to the yeast and/or the

Page 65: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

54

effects of ethanol toxicity affecting the functioning of the cellular membrane and resulting in

a disruption to normal cellular mechanisms.

Although it was expected that initial sugar concentration would have a negative impact

on wine quality, the results obtained in this chapter demonstrated the specific effects of

hyperosmotic stress on the production of volatiles, and how the balance of NAD+ and NADH

affects the metabolic pathways and consequently affects the flavour profile of the wine. for

instance, higher amounts of ketones and terpene hydrocarbons were detected in wines with

higher sugar concentrations. Potentially, the data obtained from this study can be further

investigated to use initial sugar concentration as a mean of modulating wine flavour.

Furthermore, the data obtained from this study can be used as an indicator of the flavour

profile of the wine that will be produced from batches of mangoes with different maturities,

and consequently different sugar content.

Page 66: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

55

Chapter 4 Effects of Co-fermentation of Saccharomyces cerevisiae and Williopsis

saturnus Yeasts on Volatile Production

4.1 Introduction

With its ability to modulate most major constituents of wine (including residual sugar,

ethanol, polyols, acids and phenolics) (Pretorius, 2000), S. cerevisiae is widely utilised

commercially, specifically for producing wines with high ethanol and low residual sugar

content. However, non-Saccharomyces yeasts are able to produce a diversity of novel

aromatic compounds (such as acetate esters and other branch-chained esters) different from S.

cerevisiae wines and has potential application in wine production to increase aroma diversity

(Dubourdieu et al. 2006) due to the host of enzymes such as esterases, glycosidases, amongst

others. In addition, Williopsis saturnus var. saturnus strains has a high acetate ester producing

ability and are known to convert higher alcohols into the corresponding acetate esters

(Yilmaztekin et al. 2009) and was chosen for its potential for flavour enhancement by contributing

to the fruity and floral note in wine (Iwase et al. 1995; Yilmaztekin et al. 2009). S. cerevisiae

MERIT.ferrm was selected due to its vigorous fermenting ability and tolerance toward high sugar and

ethanol concentrations.

To harness the benefits of both Saccharomyces and non-Saccharomyces yeasts, mixed

culture (also known as co-fermentation or co-inoculation) is a potential fermentation strategy

that can be utilised. Mixed culture fermentations involving Saccharomyces and non-

Saccharomyces yeasts for grape wines have been extensively studied and shown to allow for

differentiation of wines due to increased flavour complexity and mouthfeel enhancement

(Fleet 2008; Viana et al. 2009). Some commercial non-Saccharomyces yeasts in use now

include Kluyveromyces thermotolerans and Torulapora delbrueckii (Pretorius et al. 1999).

Page 67: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

56

This chapter studied the chemical composition of three different mango wines – one

produced from a S. cerevisiae MERIT.ferm, one produced from W. saturnus NCYC 500, and

a mango wine produced from a simultaneous fermentation of yeasts S. cerevisiae

MERIT.ferm and W. mrakii NCYC 500.

4.2 Results and discussion

4.2.1 Total soluble solids, sugar concentrations and yeast cell count

The growth kinetics, changes in TSS and sugar concentration is illustrated in Table 4.1.

The yeast cell populations increased from the initial 5.50×105 CFU/mL for S. cerevisiae,

6.33×105 CFU/mL for W. mrakii to 9.01×10

8 CFU/mL and 9.65×10

7 CFU/mL respectively

after a 21-day fermentation period. For the mixed culture fermentation, S. cerevisiae and W.

mrakii yeasts were inoculated in a 1:1000 ratio. S. cerevisiae was initially present at a level

of 2×102 CFU/mL and W. mrakii was present at a level of 1.67 ×10

5 CFU/mL. At the end of

21 days, no W. mrakii was detected and S. cerevisiae reached a population size of 7.05×108

CFU/mL (Table 4.1). Lysine agar was used to enumerate W. mrakii towards the end of the

fermentation when numbers were too low to count on PDA.

In the mixed culture fermentation, where S. cerevisiae and W. mrakii were

simultaneously inoculated, S. cerevisiae dominated the fermentation. This was expected as S.

cerevisiae is known to be a vigorous fermenter with the ability to rapidly adapt to the harsh

oenological conditions. S. cerevisiae grew to about 108 CFU/mL and remained stationary

until the end of the fermentation. W. mrakii, on the other hand, declined rapidly and steadily

(Figure 4.1) These could either be due to the inhibitory effects from the toxic metabolites of

S. cerevisiae fermentation (such as ethanol), and/or the lack of nutrients for growth due to its

weaker capability to compete for nutrients with the more robust Saccharomyces species. The

Page 68: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

57

toxic effects of ethanol coupled with the lack of nutrients might have a detrimental effect on

the survival of the W. mrakii. Other potential factors for the early death of W. mrakii yeasts

include oxygen unavailability (Holm Hansen, et al. 2001), formation of other toxic

compounds, quorum sensing and cell-to-cell contact mechanism (Fleet and Heard 1993;

Nissen and Arneborg 2003; Nissen et al. 2003; Farkas et al. 2005).

Table 4.1 Physicochemical properties, reducing sugar and organic acid content before and after fermentation Day 0 Day 21

Yeast strains MERIT NCYC 500 Mixed Culture MERIT NCYC 500 Mixed Culture

Physiochemical properties pH 3.51± 0.01

a 3.50± 0.01

a 3.51± 0.01

a 3.52± 0.02

a 3.50± 0.01

a 3.51± 0.01

a

oBrix 16.7 ± 0.01

a 16.7 ± 0.01

a 16.7 ± 0.01

a 5.3 ± 0.01

b 11.46 ± 0.5

c 5.5 ± 0.01

b

Plate count (105 CFU/mL)

S. cerevisiae MERIT.ferm 6.33 ± 1.10a - 0.002 ±0.001

b 9012 ± 1291

c - 7051 ± 993

d

W. mrakii NCYC 500** - 6.33 ± 1.10a 6.67±1.10

a - 965 ± 85

b N.D*.

Organic acid (g/100 mL) Citric acid 0.26 ± 0.03

a 0.25 ± 0.02

a 0.26 ± 0.02

a 0.25 ± 0.03

a 0.24 ± 0.02

a 0.25 ± 0.02

a

Malic acid 0.90 ± 0.09a 0.91 ± 0.08

a 0.90 ± 0.07

a 0.46 ± 0.05

b 0.43 ± 0.02

b 0.48 ± 0.05

b

Succinic acid 0.079 ± 0.012a 0.079 ± 0.011

a 0.081 ± 0.012

a 0.085 ± 0.011

b 0.079 ± 0.012

a 0.084 ± 0.014

b

Tartaric acid 0.15 ± 0.03a 0.14 ± 0.03

a 0.14 ± 0.04

a 0.14 ± 0.02

a 0.13 ± 0.02

a 0.13 ± 0.01

a

Reducing sugars (g/100 mL) Fructose 4.99 ± 0.05

a 4.97 ± 0.04

a 4.98 ± 0.06

a N.D.* 1.11 ± 0.02

b N.D.*

Glucose 0.68 ± 0.03a 0.66 ± 0.05

a 0.68 ± 0.02

a N.D.* 0.38 ± 0.05

b N.D.*

Sucrose 12.45 ± 0.12a 12.43 ± 0.16

a 12.47± 0.17

a 0.013 ± 0.00

b 2.53 ± 0.12

c 0.013 ± 0.00

b

*N.D.: not detected

** W. mrakii was enumerated on both lysine and potato dextrose agar abcd

ANOVA (n=4) confidence level at 95%, same letters in the same row indicate no significant

differences

The °Brix values decreased from 16.7 °B (in fresh mango juice) to 5.3 ° B for s

monoculture and to 5.5 ° B for the mixed culture after 21 days. The sugar contents of the

mango wine fermented with S. cerevisiae monoculture and mixed culture showed no

significant differences. Only trace levels of sucrose detected at 0.013 and 0.015g/100 mL

Page 69: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

58

respectively (Table 4.1) while both fructose and glucose were completely metabolised. On

the other hand, significant amounts of all three sugars were detected in the W. mrakii wine,

with 1.11 g/100 mL of fructose, 0.38 g/100 mL of glucose and 2.53 g/100 mL of sucrose; the

total soluble solids content was about 11.46 °B in the NCYC 500-fermented wine. Sugar

consumption was much lower by Williopsis yeast than Saccharomyces. Interestingly, in the

NCYC 500-fermented wine, the largest decrease amongst the sugars was for sucrose. This

indicates that W. mrakii utilised sugars in a different manner from the glucophillic S.

cerevisiae (Gonzalez et al. 2013).

Figure 4.1 (a) Changes TSS (°B) during fermentation for S. cerevisiae

MERIT.ferm monoculture ( ), W. mrakii NCYC 500 monoculture ( )and

mixed culture fermentation ( ) (S. cerevisiae:W. mrakii at a ratio 1:1000)

(b) Changes in yeast population during fermentation for S. cerevisiae

MERIT.ferm monoculture ( ), W. mrakii NCYC 500 monoculture ( ), S.

cerevisiae in mixed culture fermentation ( )and W. mrakii in mixed culture

0

5

10

15

20

0 5 10 15 20

TS

S (°B

rix

)

Time (Days)

(a)

-2

0

2

4

6

8

10

0 5 10 15 20

Log

CF

U /

mL

Time (Days)

(b)

Page 70: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

59

fermentation ( ); S. cerevisiae:W. mrakii at a ratio 1:1000 in mixed culture

fermentation

4.2.2 pH and organic acids

The changes in pH and organic acid contents before and after fermentation can be

found in Table 4.1. The pH remained relatively constant throughout the fermentation period

between 3.4 to 3.6. With the exception of malic acid, there were no significant changes in the

organic acids (Table 4.1). Malic acid decreased by approximately 50% after fermentation. As

D,L-malic acid was added to the fresh mango juice to achieve a pH of 3.5, the total malic

acid increased from initial 0.3 g/100 mL to about 0.88 g/100 mL after pH adjustment (data

not shown). The absence of malolactic fermentation (as indicated by the absence of lactic

acid bacteria and lactic acid), and the inability of the selected yeast strains in this study to

metabolise malic acid suggest that the observed decrease in malic acid content is most likely

due to the passive diffusion of D-malic acid into the cells (Coloretti et al. 2002). Furthermore,

the addition of 100 ppm of potassium metabisulphite to the juice at the beginning of the

fermentation would inhibit the growth of unwanted microorganisms. There was a minor

increase in the amount of succinic acid after fermentation which could be formed from the

reduction of oxaloacetate and malate from the Krebs cycle. Succinic acid is commonly

present in alcoholic fermentation (Swiegers et al. 2005) and undesirable at high

concentrations due to the “salty” or “bitter” taste imparted (Whiting 1976).

4.2.3 Volatile composition of mango wines

A total of 10 alcohols, 49 esters, 3 aldehydes, 1 furan, 4 ketones, 2 lactones, 20

hydrocarbon terpenes, 2 oxides of terpene alcohols, 10 organic acids and 5 other

miscellaneous compounds (such as furans and lactones) were identified. The complete list

of volatiles produced can be found in Table 4.2 while some of the potentially impactful

odourants which were quantified can be found in Table 4.3.

Page 71: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

60

While a large number of volatiles found in mango juice were metabolised, a number of

volatile compounds were also produced from both chemical and enzymatic reactions during

fermentation. Consequently, there were significant differences between the S. cerevisiae, W.

mrakii and mixed culture fermented mango wines. The volatile profile for the mixed culture

wine was similar to that of the S. cerevisae wine, while the W. mrakii wine exhibited a

significantly different volatile profile. This was expected as it was apparent that S. cerevisiae

dominated the fermentation in the mixed culture fermentation.

Page 72: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

61

Table 4.2 Complete volatiles for mango wine fermented with yeasts S. cerevisiae MERIT.ferm, W. mrakii NCYC 500 and mixed culture

CAS1 LRI

2 Compounds Merit 500 Mixed Odour descriptor

3

Peak Area RPA (%) Peak Area RPA (%) Peak Area RPA (%)

000064-17-5 1028 Ethanol 2025 ± 50.33a 71.21 575.5 ± 26.48

b 53.69 2188 ± 63.57

a 70.25 Strong, alcoholic

Subtotal 2025 71.21 575.5 53.69 2188 70.25 s

Alcohol 000078-83-1 1172 Isobutyl alcohol 14.19 ± 0.57 a 0.499 9.06 ± 0.21

b 0.845 15.5 ± 0.440

c 0.498 Ethereal, winey

000137-32-6 1220 Active amyl alcohol 6.23 ± 0.18 a 0.219 5.67 ± 0.19

b 0.529 6.57 ± 0.18

a 0.211 Fusel, winey and solvent-like

000123-51-3 1237 Isoamyl alcohol 28.28 ± 1.36 a 0.995 15.63 ± 1.22

b 1.458 43.29 ± 0.650

c 1.39 Fusel, alcoholic, cognac, fruity,

banana

000111-27-3 1448 Hexanol 0.63 ± 0.0288a 0.022 1.03 ± 0.11 b 0.096 0.67 ± 0.031

a 0.021 Ethereal, fruity and alcoholic,

fusel oil

000928-96-1 1475 cis-3-Hexenol 1.04 ± 0.05 a 0.037 1.71 ± 0.08

b 0.16 0.38 ± 0.01

c 0.012 Green, grassy

000505-10-2 1706 Methionol 1.24 ± 0.029 a 0.044 0.94 ± 0.033

b 0.088 1.37 ± 0.010

c 0.044 Cooked vegetable

000470-08-6 1794 β-Fenchol 0.63 ± 0.03 a 0.022 0.12 ± 0.01

b 0.011 0.15 ± 0.01

c 0.005 Camphor, pine , sweet

000106-22-9 1867 Citronellol 1.3 ± 0.07 a 0.046 0.64 ± 0.02

b 0.06 1.31 ± 0.07

a 0.042 Floral, rosy, sweet, citrus

000060-12-8 1964 2-Phenylethyl alcohol 56.97 ± 1.60 a 2.004 11.84 ± 1.49

b 1.105 87.52 ± 1.560

c 2.811 Floral, rose and honey-like

000078-70-6 1999 Linalool 0.49 ± 0.01 a 0.017 0.43 ± 0.03

b 0.04 1.43 ± 0.070

c 0.046 Citrus, orange, floral and rose

000128-37-0 2090 p−Cresol 0.25 ± 0.02 a 0.009 0.13 ± 0.01

b 0.012 0.22 ± 0.02

a 0.007 Mild floral

Subtotal 111.3 3.994 47.2 4.405 158.5 5.087

Ethyl ester 000105-54-4 1098 Ethyl butanoate 2.3 ± 0.10 a 0.081 1.21 ± 0.03

b 0.113 2.54 ± 0.14

a 0.082 Fruity, sweet, apple, fresh

000123-66-0 1297 Ethyl hexanoate 0.28 ± 0.01 a 0.01 0.71 ± 0.52

b 0.066 6.28 ± 0.220

c 0.202 Sweet, pineapple, fruity

64187-83-3 1358 Ethyl cis-3-hexenoate 0.1 ± 0.003 a 0.004 0.03 ± 0.001

b 0.003 0.1 ± 0.005

a 0.003 Sweet, fruity, pineapple, green

054653-25-7 1392 Ethyl 5-hexenoate 3.01 ± 0.030 a 0.106 0.02 ± 0.001

b 0.001 2.51 ± 0.03

c 0.081 Fruity, pineapple

000106-32-1 1453 Ethyl octanoate 83.84 ± 2.02 a 2.949 3.14 ± 0.06

b 0.293 138.8 ± 2.140

c 4.456 Waxy, sweet, pineapple and

fruity

35194-38-8 1486 Ethyl 7-octenoate 0.41 ± 0.005 a 0.014 0.01 ± 0.001

b 0.001 0.36 ± 0.04

c 0.012 Tropical, fruity

000123-29-5 1624 Ethyl nonanoate 1.81 ± 2.32 a 0.064 1.04 ± 0.062

b 0.097 1.37 ± 0.110

c 0.044 Waxy, soapy, cognac

000110-38-3 1746 Ethyl decanoate 305.3 ± 5.12 a 10.738 1.05 ± 0.05

b 0.098 246.2 ± 5.470

c 7.907 Sweet, waxy, fruity

067233-91-4 1795 Ethyl 9-decenoate 87.45 ± 1.39 a 3.076 11.3 ± 0.565

b 1.054 63.07 ± 0.600

c 2.025 Fruity, fatty

000106-33-2 1887 Ethyl dodecanoate 71.41 ± 0.53 a 2.512 0.82 ± 0.70

b 0.077 36.94 ± 0.260

c 1.186 Sweet, waxy, soapy

Page 73: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

62

Table 4.2 (Continued)

000124-06-1 2201 Ethyl tertradecanoate 2.42 ± 0.0434a 0.131 0.54 ± 0.031

b 0.05 1.36 ± 0.066

c 0.044 Sweet, waxy, creamy

000628-97-7 2373 Ethyl hexadecanoate 24.88 ± 0.62 a 0.875 1.77 ± 0.014

b 0.166 1.97 ± 0.041

c 0.063 Waxy, fruity, creamy

054546-22-4 2402 Ethyl 9-hexadecenoate 1.22 ± 0.0427

a

0.043 0.05 ± 0.001 b 0.005 0.8 ± 0.01

c 0.026

Subtotal 598.3

21.044 34.32

3.202 516.6

16.588

Acetate ester 000109-60-4 1002 Propyl acetate 0.75 ± 0.01 a 0.026 3.98 ± 5.18

b 0.371 1.27 ± 0.03

c 0.041 Fusel, sweet, fruity

000141-78-6 1009 Ethyl acetate 4.41 ± 0.40 a 0.155 148.6 ± 1.45

b 13.869 94.35 ± 0.38

c 3.03 Ethereal, fruity, sweet, green

000110-19-0 1020 Isobutyl acetate 1.17 ± 0.08 a 0.041 2.18 ± 0.02

b 0.203 2.07 ± 0.08

b 0.066 Sweet, fruity, banana, tropical

000123-86-4 1061 Butyl acetate 0.65 ± 0.036 a 0.023 0.88 ± 0.00

b 0.082 0.65 ± 0.03

a 0.021 Fruity, banana

000628-63-7 1097 Active amyl acetate 0.21 ± 0.01 a 0.007 9.63 ± 0.21

b 0.899 3.83 ± 0.16

c 0.123 Ethereal, fruity, banana

000123-92-2 1112 Isoamyl acetate 1.97 ± 0.08 a 0.069 86.67 ± 1.87

b 8.087 0.42 ± 0.01

c 0.013 Sweet, fruity, banana

003681-71-8 1324 cis-3-Hexenyl acetate 0.66 ± 0.04 a 0.023 13.85 ± 0.30

b 1.292 2.91 ± 0.14

c 0.093 Fresh, green, sweet, fruity,

apple

002497-18-9 1367 2-Hexenyl acetate 0.11 ± 0.01 a 0.004 0.45 ± 0.01

b 0.042 0.03 ± 0.001

c 0.001 Sweet, leafy green, fruity

000142-92-7 1413 n-Hexyl acetate 0.33 ± 0.02 a 0.012 2.22 ± 0.138

b 0.207 0.43 ± 0.01

c 0.014 Fruity, green apple banana

sweet

000150-84-5 1659 Citronellyl acetate 0.69 ± 0.01 a 0.024 2.82 ± 0.04

b 0.263 4.15 ± 0.10

c 0.133 Floral, green, rose, fruit

000141-12-8 1753 Neryl acetate 3.69 ± 0.09 a 0.13 0.49 ± 0.06

b 0.046 0.11 ± 0.01

c 0.004 Floral, rose, citrus, pear

16630-55-0 1764 Methionyl acetate 0.21 ± 0.01 a 0.007 0.34 ± 0.01

b 0.032 0.33 ± 0.03

b 0.011 Sulphurous

000105-87-3 1790 Geranyl acetate 2.56 ± 0.12 a 0.09 3.56 ± 0.151

b 0.332 2.64 ± 0.04

a 0.085 Floral, green, herbal

000103-45-7 1862 2-Phenylethyl acetate 10.56 ± 0.52 a 0.371 41.93 ± 0.20

b 3.913 40.99 ± 1.10

c 1.316 Sweet, honey, floral rosy

Subtotal 27.97 0.984 317.6 29.638 154.2 4.951

Higher

esters

000624-13-5 1506 Propyl octanoate 0.09 ± 0.0059

a

0.003 0.12 ± 0.003 b 0.011 0.16 ± 0.001

c 0.005 Waxy, fruity, fatty, green, fruity

005461-06-3 1642 Isobutyl octanoate 2.22 ± 0.03 a 0.078 3.21 ± 0.10

b 0.3 2.92 ± 0.03

c 0.094 Fruity, green, oily, floral

030673-60-0 1721 Propyl decanoate 0.73 ± 0.02 a 0.026 0.38 ± 0.01

b 0.035 0.8 ± 0.01

c 0.026 Waxy, fruity, fatty, fruity

002035-99-6 1762 Isoamyl octanoate 6.76 ± 0.04 a 0.238 7 ± 0.320

a 0.653 9.54 ± 0.110

c 0.306 Sweet, fruity, waxy

030673-38-2 1859 Isobutyl decanoate 2.25 ± 0.094 a 0.079 0.37 ± 0.018

b 0.034 1 ± 0.045

c 0.032 Oily, sweet, brandy, apricot

000103-52-6 1914 2-Phenylethyl butanoate 1.64 ± 0.12 a 0.058 0.02 ± 0.001

b 0.002 0.2 ± 0.001

c 0.006 Fruity, floral, green, winey

002306-91-4 1973 Isoamyl decanoate 5.45 ± 0.28 a 0.192 0.11 ± 0.004

b 0.01 1.96 ± 0.120

c 0.063 Waxy, banana, cognac, green

Page 74: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

63

Table 4.2 (Continued)

006309-51-9 2180 Isoamyl dodecanoate 0.55 ± 0.01 a 0.019 0.07 ± 0.002

b 0.007 0.15 ± 0.000

c 0.005 Cognac, green, waxy

005457-70-5 2397 2-Phenylethyl octanoate 0.36 ± 0.01 a 0.013 0.07 ± 0.000

b 0.007 0.02 ± 0.001

c 0.001 Sweet, fruity, creamy, floral

Subtotal 20.05 0.705 11.35 1.059 16.75 0.538

Aldehydes 000075-07-0 939 Acetaldehyde 5.27 ± 0.03 a 0.185 1.1 ± 0.11

b 0.103 10.16 ± 0.530

c 0.326 Pungent, fresh, green

000100-52-7 1637 Benzaldehyde 0.19 ± 0.0065

a

0.007 0.07 ± 0.003 b 0.007 2.11 ± 0.050

c 0.068 Almond, fruity, nutty

000104-87-0 1773 p-Tolualdehyde 3.05 ± 0.05 a 0.107 1.78 ± 0.03

b 0.166 1.79 ± 0.01

b 0.057 Cherry

Subtotal 8.51 0.299 2.95 0.276 14.06 0.452

Ketones

000821-55-6 1331 2-Nonanone 1.03 ± 0.0366

a

0.036 1.85 ± 0.006 b 0.172 1.32 ± 0.020

c 0.042 Fruity, sweet, waxy

000513-86-0 1401 Acetoin 0.03 ± 0.0012

a

0.001 0.75 ± 0.027 b 0.07 0.25 ± 0.330

c 0.008 Sweet, buttery creamy, dairy

002305-05-7 1418 γ-Dodecalactone 0.06 ± 0.0030

a

0.002 0.12 ± 0.007 b 0.011 0.07 ± 0.0035

a

0.002 Fruity, peach, creamy

023696-85-7 1938 β-Damascenone 0.57 ± 0.01 a 0.02 1.86 ± 0.044

b 0.174 1.05 ± 0.05

c 0.034 Woody, sweet, fruity, floral

Subtotal 1.69 0.059 4.58 0.428 2.69 0.086

Furan

000539-52-6 1331 3-(4-Methyl-3-

pentenyl)-furan

0.04 ± 0.002 a 0.001 0.13 ± 0.010

b 0.012 0.04 ± 0.001

a 0.001 Woody

Subtotal 0.04 0.001 0.13 0.012 0.04 0.001

Terpenes 003387-41-5 1173 Sabinene 0.46 ± 0.01 a 0.016 3.97 ± 0.07

b 0.37 0.39 ± 0.014

c 0.013 Woody, spicy, citrus

013466-78-9 1206 δ-3-Carene 0.09 ± 0.001 a 0.003 1.55 ± 0.06

b 0.145 0.19 ± 0.01

c 0.006 Sweet, citrus

000123-35-3 1246 β-Myrcene 0.68 ± 0.008 a 0.024 4.43 ± 0.09

b 0.404 0.38 ± 0.031

c 0.012 Woody, citrus, fruity, mango,

minty

095327-98-3 1254 Limonene 0.07 ± 0.00 a 0.002 0.22 ± 0.01

b 0.021 0.22 ± 0.009

b 0.007 Sweet, citrus, peely

000555-10-2 1265 β-Phellandrene 0.11 ± 0.01 a 0.004 0.6 ± 0.02

b 0.056 0.18 ± 0.001

c 0.006 Mint, turpentine

027400-71-1 1290 β-Ocimene 0.49 ± 0.0191

a

0.017 1.95 ± 0.05 b 0.182 0.1 ± 0.001

c 0.003 Green, tropical, woody, floral

000099-85-4 1305 γ-Terpinene 0.71 ± 0.03 a 0.025 1.74 ± 0.01

b 0.162 0.69 ± 0.032

a 0.022 Citrus, lime-Iike, oily

000099-87-6 1339 p-Cymene 1.59 ± 0.25 a 0.056 6.1 ± 0.09

b 0.569 1.99 ± 0.010

c 0.064 Woody, lemon-like

000586-62-9 1352 α-Terpinolene 1.72 ± 0.12 a 0.061 20.97 ± 1.01

b 1.957 1.86 ± 0.020

c 0.06 Sweet, fresh, piney citrus

Page 75: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

64

043124-56-7 1382 Isoterpinolene 0.13 ± 0.00 a 0.005 0.57 ± 0.03

b 0.053 0.09 ± 0.004

c 0.003 -

016409-43-1 1434 cis-Rose oxide 0.32 ± 0.02 a 0.011 3.62 ± 2.69

b 0.338 0.31 ± 0.0101

a

0.01 Green, vegetative, floral, herbal

001195-32-0 1529 p-Cymenene 9.76 ± 0.23 a 0.343 19.26 ± 1.14

b 1.797 9.66 ± 0.451

a 0.31 Phenolic, spicy, musty

000087-44-5 1695 β-Caryophyllene 0.22 ± 0.0075

a

0.008 0.64 ± 0.023 b 0.06 0.2 ± 0.006

a 0.006 Sweet, woody, spice, clove

006753-98-6 1725 α- Humulene 0.57 ± 0.03 a 0.02 1.1 ± 0.038

b 0.103 0.92 ± 0.010

b 0.03 Woody

017066-67-0 1829 β-Selinene 0.24 ± 0.0083

a

0.009 0.44 ± 0.02 b 0.041 0.21 ± 0.003

a 0.007 Herbal

Subtotal 17.17 0.604 67.07

6.258 17.4

0.559

Misc

compounds

016409-43-1 1434 cis-Linalool oxide 0.023 ± 0.0005a 0.001 0.29 ± 0.22

b 0.027 0.024 ± 0.001

a 0.001 Woody, floral, slightly green

001786-08-9 1563 Nerol oxide 0.71 ± 0.01 a 0.025 0.95 ± 0.02

b 0.089 1.23 ± 0.020

c 0.039 Green, vegetative, floral, mint

Subtotal 0.73 0.026 1.24 0.116 1.25 0.04

Acids 000064-19-7 1549 Acetic acid 2.96 ± 0.134 a 0.104 1.03 ± 0.047

b 0.097 0.72 ± 0.040

c 0.023 Sharp, pungent, sour, vinegar

000079-31-2 1575 Isobutyric acid 0.51 ± 0.23 a 0.018 0.35 ± 0.012

b 0.033 0.55 ± 0.025

c 0.018 Acidic note

000067-43-6 1728 Butanoic acid 0.24 ± 0.011a 0.008 0.53 ± 0.02

b 0.049 0.23 ± 0.01

a 0.007 Sharp, dairy-like, cheesy

000142-62-1 1890 Hexanoic acid 0.8 ± 0.036 a 0.028 1.09 ± 0.11

b 0.102 0.72 ± 0.002

c 0.023 Sour, fatty, sweat, cheese

000124-07-2 2170 Octanoaic acid 19.41 ± 0.64 a 0.683 1.74 ± 0.08

b 0.162 19.8 ± 0.360

a 0.636 Fatty, waxy, rancid

000334-48-5 2390 Decanoic acid 0.53 ± 0.03 a 0.019 0.06 ± 0.002

b 0.006 16.33 ± 0.060

c 0.524 Rancid, sour, fatty

000065-85-0 2455 Benzoic acid 0.54 ± 0.02 a 0.019 0.55 ± 0.040

a 0.051 0.87 ± 0.020

b 0.028 Faint, balsam

14436-32-9 2474 9-Decenoic acid 4.46 ± 0.09 a 0.157 1.25 ± 0.067

b 0.116 2.36 ± 0.050

c 0.076 Waxy, green, fatty, soapy

000143-07-7 2544 Dodecanoic acid 1.75 ± 0.084 a 0.061 0.95 ± 0.052

b 0.088 1.92 ± 0.070

c 0.062 Mild fatty, coconut, bay oil

Subtotal 31.2 1.097 7.55 0.704 43.5 1.397

Grand Total 2843 100 1072 100 3114 100

1 CAS numbers were obtained from Wiley database library

2 LRI of all the relative tables was determined on the DB-FFAP column, relative to C5-C40 hydrocarbons.

3 Descriptors were retrieved from http://www.thegoodscentscompany.com

abc ANOVA (n=4)at 95% confidence level with the same letters in the same row indicating no significant difference

Page 76: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

65

Table 4.3 Major volatile compounds of alcoholic fermentation by S. cerevisiae, W. mrakii and a mixed culture of S. cerevisiae and W. mrakii

CAS N1o LRI

2 Compounds S. cerevisiae MERIT

(mg /L)

OAV W. mrakii NCYC 500

(mg /L)

OAV Mixed-culture

(mg /L)

OAV Odour

threshold

Organoleptics3

Alcohol 000078-83-1 1172 Isobutyl alcohol 89.1±11.9a 2.2 20.3±0.98

b 0.5 78.8±13.6

a 2 40 Fruity, wine-like

000123-51-3 1237 Isoamyl alcohol 101.5±5.7 a 3.4 21.7±5.0

b 0.7 93.8±3.4

c 3.1 30 Alcoholic, fruity,

banana,

000060-12-8 1964 2-Phenylethyl alcohol 54.0±5.9 a 5.4 6.7±0.6

b 0.7 58.2±10.0

a 5.8 10 Sweet, rose, floral

Terpenol 000098-55-5 1352 α-Terpinolene 0.01±0.00 a 0.05 0.1±0.02

b 0.5 0.02±0.00

a 0.1 N.A. Woody, sweet, pine

000106-22-9 1867 β-Citronellol 0.013±0.00 a 0.13 0.006±0.000

b 0.06 0.017±0.003

c 0.17 0.1 Rose, citrus, green

Ester 000141-78-6 1009 Ethyl acetate 3.93±1.19 a 0.5 349.5±33.7

b 46.6 30.3±4.3

c 4 7.5 Pineapple, fruity,

varnish

000110-19-0 1020 Isobutyl acetate 0.026±0.012 a 0.016 0.22±0.04

b 0.14 0.027±0.019

a 0.017 1.6 Fruity, sweet, apple

000123-92-2 1112 Isoamyl acetate 0.27±0.07 a 9 3.54±0.19

b 118 0.51±0.05

c 17 0.03 Fruity, banana, sweet

000103-45-7 1862 2-Phenylethyl acetate 0.59±0.15 a 2.4 3.30±0.18

b 13.2 0.51±0.10

a 2 0.25 Floral, rose, sweet

000123-66-0 1297 Ethyl hexanoate 0.27±0.05 a 19.3 0.006±0.001

b 0.4 0.25±0.08

a 17.9 0.014 Sweet, pineapple,

waxy

000106-32-1 1453 Ethyl octanoate 1.13±0.30 a 565 0.053±0.002

b 26.5 0.95±0.39

a 475 0.002 Floral, fruity, brandy

000110-38-3 1746 Ethyl decanoate 1.26±0.25 a 6.3 0.059±0.002

b 0.3 0.84±0.23

c 4.2 0.2 Waxy, sweet, apple

000106-33-2 1887 Ethyl dodecanoate 1.21±0.45 a 0.2 0.083±0.005

b 0.01 0.89±0.05

c 0.15 5.9 Soapy, waxy, floral

Acid 000064-19-7 1549 Acetic acid 452±49 a 2.3 680±14

b 3.4 647±105

c 3.2 200 Vinegar, pungent

000142-62-1 1890 Hexanoic acid 1.39±0.15 a 0.46 0.32±0.02

b 0.11 2.17±0.56

c 0.72 3 Fatty, soapy, sour

000124-07-2 2170 Octanoic acid 4.64±0.76 a 0.53 0.24±0.03

b 0.03 3.87±0.59

c 0.44 8.8 Fatty, soapy, sour

000334-48-5 2390 Decanoic acid 1.57±0.27 a 0.16 0.35±0.01

b 0.04 1.45±0.03

a 0.15 10 Fatty, rancid, sour

000143-07-7 2544 Dodecanoic acid 0.65±0.05 a 0.065 0.48±0.02

b 0.048 0.66±0.04

a 0.066 10 Coconut, fatty

1 CAS numbers were obtained from Wiley database library

2 LRI of all the relative tables was determined on the DB-FFAP column, relative to C5-C40 hydrocarbons.

3 Descriptors were retrieved from http://www.thegoodscentscompany.com

abc ANOVA (n=4)at 95% confidence level with the same letters in the same row indicating no significant difference

Page 77: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

66

4.2.3.1 Terpenes

The major terpenes in mango wines were α-caryophyllene, β-caryophyllene, β-myrcene,

β-ocimene and β-phellandrene (Table 4.3). Terpenes constituted the largest group of

volatiles in fresh mango juice (Table 3.2) but most of the terpene hydrocarbons in fresh

mango juice decreased substantially after fermentation. The decrease in terpenes was

significantly more pronounced in the S. cerevisiae predominant fermentations. This is in

contrast with some previous reports which claimed that fermentation would not affect the

concentration of terpenes (Rapp 1998; Ong and Acree 1999; Alves et al. 2010). However, it

has also been reported that some Saccharomyces strains would cause the decrease of terpenes

(Zoecklein et al. 1997). This could be due to the inherent variation in the genome makeup for

the different strains and should be further investigated for the production of wine with more

varietal character, since studies have indicated that varietal character of mango is often

distinguished by the presence of terpene hydrocarbons (Chauhan et al. 2010). This could be

an area of W. mrakii fermentation that can be further investigated and utilised to produce

wine where the original mango aroma is better preserved, because W. mrakii can better retain

the characteristic mango aroma.

Some new terpene hydrocarbons were formed. For instance, β-myrcene was not

detected in fresh mango juice but detected after fermentation. This could be due to the

metabolic activities of the yeasts. Similar to the general trend observed for terpene

hydrocarbons, β- myrcene was found in higher amounts in the W. mrakii –fermented wine.

The concentration of β- myrcene decreased throughout the S. cerevisiae fermentation with the

largest decrease observed from Day 2 to Day 4, before the decrease became more gradual.

On the other hand, β- myrcene increased from Day 2 to Day 4 for the W. mrakii and mixed

culture fermentation before a significant decrease was obsesrved in mixed culture

fermentation and a less drastic decrease was observed in W. mrakii fermentation. β- myrcene

Page 78: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

67

steadily increased again from Day 10 in W. mrakii fermentation while β- myrcene decreased

gradually in the mixed culture fermentation. The trend in evolution for the mixed culture

fermentation could be largely due to the domination of S. cerevisiae after the initial

domination of W. mrakii resulting in similar trends as the S. cereisiae monoculture

fermentation and final content of β- myrcene. β-myrcene is known to be the product of

linalool degradation, hence it is likely that the linalool produced during fermentation

simultaneously undergoes degradation due to other metabolic reactions, resulting in the

formation of β- myrcene (Förster-Fromme and Jendrossek 2010).

Figure.4.2 Changes in β-myrcene in the fermentation of mango juice with a

monoculture of S. cerevisiae MERIT.ferm ( ), monoculture of W. mrakii

NCYC 500 ( ) and a mixed culture of S. cerevisiae and W. mrakii ( )

4.2.3.2 Alcohols

Alcohols make up the largest group of volatiles (by RPA) in all three mango wines with

ethanol being the most dominant alcohol formed with an RPA of 71.2% (v/v) for strain S.

cerevisiae, 26.4% v/v for strains NCYC 500 and 70.2% (v/v) for mixed culture fermentation

(Table 4.3,). The major alcohols in fresh mango juice were cis-3-hexanol, ethanol, hexanol

and trans-2-hexenol (Table 3.1) while the major alcohols in mango wines were isobutyl

alcohol, isoamyl alcohol and 2-phenylethyl alcohol (Table 4.3). The amounts of all higher

0.0E+00

1.0E+06

2.0E+06

3.0E+06

4.0E+06

5.0E+06

6.0E+06

0 5 10 15 20

Pea

k a

rea

Time (Days)

β-Myrcene

Page 79: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

68

alcohols found in mango juice decreased after fermentation and significant amounts of new

higher alcohols were formed. Generally, the S.cerevisiae-dominated fermentations produced

higher amounts of alcohols relative to the W. mrakii-fermented wine.

There were significant differences in the amounts of ethanol produced from the three

different treatments. The monoculture of S.cerevisiae produced the most ethanol, followed

by the mixed culture and lastly the monoculture of W.mrakii. Ethanol production peaked at

Day 7 and 10 for S.cerevisiae and mixed culture fermentation respectively, before a gradual

decline (Figure 4.3). The reason for this decline is unclear as there were no corresponding

increases in the ethyl esters (Figure 4.5) which might have utilised ethanol as a precursor.

Figure 4.3 Changes in ethanol content in the fermentation of mango juice with

a monoculture of S. cerevisiae MERIT.ferm ( ), monoculture of W. mrakii

NCYC 500 ( ) and a mixed culture of S. cerevisiae and W. mrakii ( )

The major fusel alcohols produced were isoamyl alcohol, isobutyl alcohol and 2-

phenylethyl alcohol. All three alcohols were found at levels higher than their odour

thresholds (OAV>1) in the mango wine fermented by the S. cerevisiae yeasts and mixed

culture (Table 4.2). The mixed culture fermentation produced slightly lesser amounts of

isoamyl and isobutyl alcohols than the S. cerevisiae monoculture. 2-Phenylethyl alcohol was

0.0E+00

5.0E+08

1.0E+09

1.5E+09

2.0E+09

2.5E+09

3.0E+09

3.5E+09

4.0E+09

0 5 10 15 20

Pea

k a

rea

Time (Days)

Ethanol

Page 80: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

69

found in the highest concentration in the mixed culture fermentation. W. mrakii fermentation

generally produced the lowest amounts of fusel alcohols.

Isoamyl and isobutyl alcohol exhibited similar changes during the course of

fermentation (data for isoamyl alcohol not shown). Production increased drastically from

Day 2 to 4 for S. cerevisiae fermentation and from Day 4 to 7 for mixed culture fermentation

(Figure 4.4). After which, both alcohols generally decreased as the fermentation proceeded.

On the other hand, W. mrakii, produced isoamyl and isobutyl alcohol a lot more slowly

(Figure 4.4). However, the final concentration of isobutyl alcohol did not differ significantly

between the mixed culture and W. mrakii fermentation due to the decrease in isobutyl alcohol

during the later part of the mixed culture fermentation. The production of 2-phenylethyl

alcohol was more gradual for all three treatments. Maximum concentration was attained at

Day 10 and Day 14 for mixed culture and S. cerevisiae fermentation respectively while

production increased steadily throughout the fermentation period for W. mrakii. Slight

decreases were observed for both S. cerevisiae and mixed culture fermentations.

Fusel alcohols have a positive effect on wine flavour at concentrations below 350 mg/L;

typically, wines have been reported to contain 80–540 mg/L (Rapp and Mandery, 1986). In

this study, the major higher alcohol content falls within the desirable range. The production

of such moderate levels of alcohols may be beneficial to the flavour profile. S. cerevisiae

yeast may be harnessed for its production of desirable higher alcohols. On the other hand, the

W. mrakii yeast produced significantly lower amounts of these higher alcohols. It is likely

that the individual higher alcohol was not able to contribute much to the flavour profile, but

there may be a synergistic effect and the total sum of these alcohols may contribute positively

to the overall flavour profile.

Page 81: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

70

Figure 4.4 Changes in isobutyl alcohol and 2-phenylethyl alcohol content in

the fermentation of mango juice with a monoculture of S. cerevisiae

MERIT.ferm ( ), monoculture of W. mrakii NCYC 500 ( ) and a mixed

culture of S. cerevisiae and W. mrakii ( )

Terpene alcohols or terpenols such as citronellol and linalool were also detected in the

mango wine. These compounds may impart attributes such as “fruity” or “floral” flavour

(Iwase et al. 1995). However, both terpenols were produced at levels lower than their odour

thresholds for all three fermentations. Although these compounds may have synergistic

effects with other higher alcohols, they may contribute little to the overall flavour profile of

the mango wine individually. Increases in linalool were also after fermentation were

observed and were inversely proportional to the myrcene content. As β-myrcene is the

product of linalool degradation (Förster-Fromme and Jendrossek 2010), moderation of this

mechanism might have a positive effect on the flavour profile of mango wine

0.0E+00

5.0E+06

1.0E+07

1.5E+07

2.0E+07

2.5E+07

0 5 10 15 20

Pea

k a

rea

Time (Days)

Isobutyl alcohol

0.0E+00

1.0E+07

2.0E+07

3.0E+07

4.0E+07

5.0E+07

6.0E+07

0 5 10 15 20

Pea

k a

rea

Time (Days)

2-Phenylethyl alcohol

Page 82: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

71

4.2.3.3 Esters

Esters made up the second largest group (over 25% RPA). Ethyl esters were the major

esters detected in the S. cerevisiae-dominated fermented wines while acetate esters dominated

the W. mrakii-fermented wine. The ethyl esters that were quantified include ethyl hexanoate,

ethyl octanoate, ethyl decanoate, ethyl dodecanoate while the acetate esters that were

quantified were ethyl acetate, isobutyl acetate, isoamyl acetate and 2-phenylethyl acetate.

While only two ethyl esters of significance were detected in fresh mango juice, both the

number and amount of ethyl esters increased after fermentation. The major ethyl esters

present in the S. cerevisiae and mixed culture fermentations were ethyl octanoate, ethyl

decanoate and ethyl dodecanoate. In terms of potential contribution to wine flavour profile,

ethyl octanoate (described as floral, fruity, brandy) may have a more significant impact.

Production of ethyl octanoate peaked at Day 4 and Day 7 for S.cerevisiae and mixed culture

fermentation respectively (Figure 4.5). After which, ethyl octanoate decreased steadily for

the S.cerevisiae fermentation while some a small increase was observed from Day 10 to 14 in

the mixed culture fermentation. On the other hand, ethyl octanoate production was

significantly lower in W.mrakii fermentation. A noticeable increase was observed from Day

10 to Day 14 before ethyl octanoate content remained relatively constant.

Ethyl hexanoate and ethyl decanoate were also found at levels above their odour

thresholds, while ethyl dodecanoate was only present at levels above detection thresholds in

the mixed culture fermentation. With the exception of ethyl octanoate, all the other ethyl

esters quantified were below their odour thresholds.

Page 83: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

72

Figure 4.5 Changes in ethyl octanoate in the fermentation of mango juice

with a monoculture of S. cerevisiae MERIT.ferm ( ), monoculture of W.

mrakii NCYC 500 ( ) and a mixed culture of S. cerevisiae and W. mrakii ( )

At appropriate levels, these medium-chain esters add moderate notes of ripe fruits to

fermented wine (Alves et al. 2010). In this study, ethyl hexanoate, ethyl octanoate and ethyl

decanoate were all found at levels above their odour thresholds in the Saccharomyces-

fermented wine. In the mixed culture fermentation, ethyl dodecanoate was produced at high

levels and exceeded its odour threshold with an OAV of 5.9. The presence of much higher

ethyl esters content in the S. cerevisiae fermentation may imply that the efficiency of fatty

acid metabolism is much lower in W. mrakii than S. cerevisiae. However, these esters may

impart rancid and soapy flavours at high concentrations; as such, careful moderation of these

esters may make a positive contribution to the wine aroma. In this respect, mixed culture

fermentation resulted in a lower concentration of ethyl decanoate and ethyl dodecanoate.

Hence, such a fermentation strategy may prove to be beneficial in moderating wine flavour

profile.

W. mrakii fermentation produced higher amounts of acetate esters. Acetate esters may

have positive contributions to the overall quality of the wine and most impart moderate

“floral” or “fruity” flavour notes. All the acetate esters quantified in this study were present

0.0E+00

2.0E+07

4.0E+07

6.0E+07

8.0E+07

1.0E+08

1.2E+08

1.4E+08

0 5 10 15 20

Pea

k a

rea

Time (Days)

Ethyl octanoate

Page 84: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

73

in amounts higher than their threshold values (Table 4.3), indicating the likelihood of their

contribution to fruity, floral and sweet flavours to the final wine bouquet. The timepoint

evolution of ethyl acetate and 2-phenylethyl acetate can be found in Figure 4.6. There was no

significant differences between 2-phenylethyl acetate between W. mrakii and mixed culture

fermentations after 21 days. 2-Phenylethyl acetate was initially produced at a much faster

rate in the mixed culture fermentation and exceeded that in the W. mrakii fermentation until a

significant decline occurred during Day 7. The reason for this decline would be due to the

increased acetate ester hydrolysing activity of the S. cerevisiae being more dominant (Fukuda

et al. 1998).

Furthermore, the presence of rather significant amounts of citronellyl acetate and neryl

acetate further demonstrated the diverse range of volatiles in the W. markii mango wine.

However, one of the possible areas of concern is the high ethyl acetate levels in the NCYC

500-fermented wine. Ethyl acetate increased consistently in the W. mrakii fermentation

(Figure 4.6) and is considered beneficial at levels of 150–200 mg/L (Jackson 2008) and

possibly detrimental to wine flavour by introducing “nail varnish”-like flavours. Therefore,

the high ethyl acetate content in the W. mrakii wine may prove detrimental to wine quality.

Acetate esters are produced from the reaction of acetyl CoA with alcohols (Perestrelo et

al. 2006). The significantly higher acetate ester contents in W. mrakii fermentations could be

due to the higher stability of the alcohol-acetyl transferases (AATase) in W. mrakii (Inoue et

al. 1997) relative to labile nature of the same enzymes in S. cerevisiae (Minetoki 1992). The

high concentrations of acetate esters could enhance the activity of hydrolyzing esterase of S.

cerevisiae (Kurita 2008). This could explain why acetate esters increased and then decreased

in mixed-culture fermentation in similar studies.

Page 85: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

74

S. cerevisiae consistently produced higher amounts of other higher branched-chain

esters than W. mrakii. This could be due to the higher amounts of fusel alcohols present in

the S. cerevisiae fermentation than in the W. mrakii fermentation. However, the significantly

lower amounts of these esters in mixed culture fermentation was unexpected. A possible

explanation would include the potential killer toxins secreted by the W. mrakii yeasts

(Magliani et al. 2008) inhibiting the normal functioning of the cellular metabolism of the S.

cerevisiae yeast.

Figure 4.6 Changes in 2-phenylethyl acetate and ethyl acetate content in the

fermentation of mango juice with a monoculture of S. cerevisiae MERIT.ferm

( ), monoculture of W. mrakii NCYC 500 ( ) and a mixed culture of S.

cerevisiae and W. mrakii ( )

0.0E+00

1.0E+07

2.0E+07

3.0E+07

4.0E+07

5.0E+07

0 5 10 15 20

Pea

k a

rea

Time (Days)

2-Phenylethyl acetate

0.0E+00 2.0E+07 4.0E+07 6.0E+07 8.0E+07 1.0E+08 1.2E+08 1.4E+08 1.6E+08 1.8E+08

0 5 10 15 20

Pea

k a

rea

Time (Days)

Ethyl acetate

Page 86: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

75

4.2.3.4 Acids

Acetic, octanoic, decanoic and dodecanoic acids were the major fatty acids detected in

all three mango wines and quantified in this study. Acetic acid was the only acid detected at

levels above its odour threshold. As discussed in the Chapter 3, acetic acid is produced when

yeasts are exposed to hyperosmotic stress and may impart an unpleasant, vinegar off-odour

when present at excessive levels. Similar trends in acetic acid production were observed Day

0 to Day 7 for both monoculture fermentations while the mixed culture produced

significantly more acetic acid. From Day 10 onwards, acetic acid production for W. mrakii

fermentation increased significantly until the end of the fermentation. On the other hand,

acetic acid production for S. cerevisiae fermentation remained relatively constant while a

decrease in acetic acid was observed from Day 14 until the end of the fermentation in the

mixed culture fermentation.

It is likely that S. cerevisiae is better adapted to such osmotic stress and therefore,

produced lesser amounts of acetic acid relative to the W. mrakii yeast. However, it was

interesting to note that acetic acid was present at significantly higher levels in the mixed

culture–fermented wine compared to the S. cerevisiae fermentation despite S. cerevisiae

dominating the fermentation. This could be due to the inhibitory effects that W. mrakii may

have on S. cerevisiae. Previous studies have shown that W. mrakii produce killer toxins that S.

cerevisiae are sensitive to. This will be discussed in greater detail in the next chapter.

Although acetic acid concentration was higher than the threshold level in all three mango

wines, it may not be detrimental to wine quality as a concentration between 0.02 to 0.07

g/100 mL was considered optimal depending on the style of wine (Lambrechts and Pretorius,

2000); the amount of acetic acid produced in this study ranged from 0.045 to 0.068 g/100 mL.

Page 87: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

76

Figure 4.7 Changes in acetic acid content in the fermentation of mango juice

with a monoculture of S. cerevisiae MERIT.ferm ( ), monoculture of W.

mrakii NCYC 500 ( ) and a mixed culture of S. cerevisiae and W. mrakii ( )

The C6 to C10 fatty acids impart mild and pleasant aroma to wine at concentrations of 4

to 10 mg/L; however, at levels beyond 20 mg/L, their impact on wine becomes negative with

fatty, rancid and soapy off-odours. Hence, they must be controlled at low levels or at least

not higher than their threshold levels. These fatty acids may arise from the auto-oxidation of

saturated lipids present as by-products of yeast fatty acid metabolism; and are extremely

important to wine flavour not only for their own odour characters but also for their

contribution to the synthesis of volatile esters (Saerens et al. 2008; Pino and Queris 2011). In

this study, the amounts of these fatty acids produced were well below 20 mg/L and may

therefore contribute to some complexity in the mango wine flavour profile.

4.2.3.5 Aldehydes and Ketones

Acetaldehyde, benzaldehyde, o-tolualdehyde, and p-tolualdehyde were identified in all

three mango wines. Acetaldehyde was the major aldehyde in the Saccharomyces dominated

wines while o-tolualdehyde was the major aldehyde in NCYC 500-fermented wines (Table

4.2). Compared with other volatiles, aldehydes were only a minor group with less than 1%

RPA.

0.0E+00

2.0E+06

4.0E+06

6.0E+06

8.0E+06

1.0E+07

1.2E+07

0 5 10 15 20

Pea

k a

rea

Time (Days)

Acetic acid

Page 88: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

77

Acetaldehyde is an important compound in terms of influence on wine quality. At low

levels, acetaldehyde imparts a fresh, fruity note to wine. However, at excessive levels, it can

impart an objectionable, green odour. Acetaldehyde was not quantified in this study due to

its instability. However, by comparing RPA, it was present in the highest amount in the

mixed culture fermentation, followed by the S.cerevisiae fermentation then W. mrakii

fermentation. As mentioned in Chapter 3, acetaldehyde is very toxic to the yeast cells and

has to be excreted, the data obtained in this chapter appears to suggest that W. mrakii is more

prone to converting acetaldehyde to acetic acid when compared to S. cerevisiae which

appears to secrete the acetaldehyde. Acetaldehyde and acetic acid are both undesirable at

excessive levels, hence the effects of this discrepancy in acetaldehyde utilisation and

secretion on wine flavour have to be further evaluated.

4.2.3.6 Miscellaneous compounds

Nerol oxide and cis-linalool oxide were still detected after fermentation. However,

trans-linalool oxide (which was present in fresh mango juice) was not detected in any of the

mango wines. Generally, S. cerevisiae fermentation produced the least amounts of these

compounds. Although cis-linalool oxide decreased after fermentation, it was found in

significantly higher concentration in W. mrakii fermentation. On the other hand, nerol oxide

was found in the highest concentration in the mixed culture fermentation, followed by the W.

mrakii fermentation. Nerol oxide can be degraded to form nerol before being converted into

neryl acetate. Although the S. cerevisiae fermentation produced the lowest amount of nerol

oxide, it contained the highest amount of neryl acetate. This is interesting because it would

be expected that W. mrakii with its high acetate synthesising activity would result in the W.

mrakii producing the highest amounts of neryl acetate. One plausible reason for this

Page 89: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

78

phenomenon could be due to the possible de novo production of nerol by S. cerevisiae yeast

(Herrero et al. 2008).

Methionol was identified in and it increased after the fermentation which was probably

produced by yeasts through L-methionine metabolism (Seow et al. 2010; Tan et al. 2012).

The amount of methionol produced was higher in the S. cerevisiae-dominated fermentations

than the W. mrakii fermentation.

4.3 Conclusion

The final flavour profile of the wine and consequently wine quality is very heavily

influenced by the choice of starter cultures, which includes choice of yeast strains and/or

monoculture or mixed culture. This chapter has illustrated the differences in the metabolism

and modulation of volatile compounds by the different starter cultures. Although the volatile

profile of the mixed culture fermentation is similar to that of the Saccharomyces monoculture,

there were still significant differences. The possibility of utilising a mixed culture

fermentation to improve the flavour profile of mango wine with similar ethanol content to a S.

cerevisiae–fermented is worth investigating. In addition, it points to the possibility of

harnessing the benefits of aroma enhancement by W. mrakii yeast and the ability for a

complete fermentation by S. cerevisiae. The possibility of using sequential inoculation can

be further explored to maximise the benefits of both strains and for the production of a mango

wine with superior wine quality.

Page 90: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

79

Chapter 5 Effects of Different Sequential Inoculation Strategies of Saccharomyces

cerevisiae and Williopsis saturnus on Volatile Production

5.1 Introduction

Mixed starter cultures have been demonstrated to enhance flavour complexities in wine

production in some studies (Ciani et al. 2006; Moreira et al. 2008; Lee et al. 2010), while

others have reported that non-Saccharomyces yeasts in mixed starter cultures haves limited

influence due to their early growth arrest (Clemente-Jimenez et al. 2005; Farkas, et al. 2005;

Bely, Stoeckle, et al. 2008; Varakumar et al. 2012). The early demise of the non-

Saccharomyces yeasts could be attributed to their lower resistance to stresses under

oenological conditions.

Sequential fermentation has been suggested to be the most adequate strategy in strain

combination; the kinetic behaviour is similar to a successful spontaneous fermentation with

he wine produced different from that of a simultaneous fermentation (Clemente-Jimenez et al.

2005; Rodríguez et al. 2010). This is likely due to the prolonged action of non-

Saccharomyces species in controlled sequential fermentation due to the lack of inhibition by

the strongly fermentative Saccharomyces species. To date, there has been no published study

on the effects of a controlled sequential fermentation on the quality of mango wine.

This chapter investigated the effects of simultaneous mixed culture fermentation (MCF)

where S. cerevisiae MERIT.ferm and W. mrakii NCYC 500 were co-inoculated against a

positive sequential fermentation (PSF) and negative sequential fermentation (NSF). In PSF,

W. mrakii NCYC 500 was first inoculated and fermentation was allowed to proceed for 14

days before killing the W. mrakii NCYC 500 yeast and inoculating the S. cerevisiae

MERIT.ferm yeast and fermentation was conducted for further 7 days. For the NSF, the

Page 91: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

80

sequence of yeast inoculation was reversed. S. cerevisiae MERIT.ferm was inoculated and

fermentation carried out for 7 days before its forced demise and then, W. mrakii NCYC 500

was inoculated and fermentation conducted for further 14 days.

5.2 Results and discussion

PSF was meant to stimulate the conditions of natural spontaneous fermentation under

controlled conditions. The absence of strongly competitive S. cerevisiae MERIT.ferm yeast

and other competitive microorganisms in the sterile mango juice was meant to maximise the

action of W. mrakii NCYC 500 yeasts on the mango juice while the strongly fermentative S.

cerevisiae MERIT.ferm yeast was inoculated later to produce the necessary ethanol content.

On the other hand, NSF reversed the natural sequence of microbial growth. Previous studies

have shown that some odour active volatile compounds increased briefly after fermentation,

then decreased (Li et al. 2012). Reversing the order of propagation of yeasts would

theoretically allow the important volatile compounds to be retained in the final mango wine

with the high ethanol content associated with Saccharomyces fermented wines, imparting

positive aroma attributes of Williopsis fermented wines (Ciani et al. 1999).

5.2.1 Physicochemical properties of mango wine

There were no significant changes in the pH and most organic acids between the three

mango wines and fresh mango juice (Table 5.1). The pH of all three mango wines produced

ranged from 3.50 to 3.52 while only malic acid showed significant decreases as discussed in

Chapter 4.

Page 92: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

81

Table 5.1 Physicochemical properties, yeast cell count, organic acid and sugar

concentrations of mango wines

Day 0 Day 21

Yeast strains PSF1 NSF

2 MCF PSF NSF MCF

Physicochemical properties

pH 3.50±0.01 a 3.50±0.01

a 3.50±0.01

a 3.51±0.02

a 3.52±0.01

a 3.50±0.01

a

Brix (°B) 17.2±0.01a 17.3±0.01

a 17.3±0.01

a 5. 61±0.01

b 7.8±0.11

c 5.58±0.01

b

Plate count (105 CFU/mL)

MERIT.ferm - 6.67±1.20 a 0.002 ±0.001

b 3.50±0.64

c - 7051 ± 993

d

NCYC 500 5.00±0.95a - 6.67±1.10

b - 53.4±7.85

c N.D*.

Organic acid (g/100 mL)

Citric acid 0.30±0.05 a 0.29±0.06

a 0.29±0.06

a 0.27±0.03

a 0.28±0.03

a 0.27±0.03

a

Malic acid 0.90±0.08 a 0.92±0.09

a 0.92±0.09

a 0.36±0.05

b 0.41±0.05

c 0.36±0.05

b

Succinic acid 0.083±0.018 a 0.088±0.011

a 0.088±0.011

a 0.086±0.011

a 0.081±0.012 0.086±0.011

a

Tartaric acid 0.15±0.04 a 0.16±0.04

a 0.16±0.04

a 0.14±0.02

a 0.14±0.02

a 0.14±0.02

a

Sugars (g/100 mL)

Fructose 4.98±0.08 a 5.03±0.09

a 4.98 ± 0.06

a N.D.* 2.46±0.15

b N.D.*

Glucose 0.63±0.02 a 0.62±0.06

a 0.68 ± 0.02

a N.D.* 0.22±0.06

b N.D.*

Sucrose 12.25±0.31 a 12.56±0.36

a 12.47± 0.17

a 0.013±0.00

b 3.84±0.36

c 0.013±0.00

b

*N.D.: not detected

abcd ANOVA (n=4) confidence level at 95%, same letters in the same row indicate no

significant difference

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S.

cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii was then

inoculated at D14. Fermentation was ceased at Day 21.

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in the ratio of

1:1000. Fermentation was ceased at Day 21.

The evolution of oBrix values for NSF and MCF was almost identical for the first 7

days of the fermentation (Figure 5.1). This is likely due to the fact that S. cerevisiae was the

dominant yeast species in MCF. Although W. mrakii was present in MCF, due to its low

consumption of sugar, the main determining factor of sugar consumption was still due to

Saccharomyces growth and cell population. The sudden increase in °Brix at Day 7 for NSF

was due to the supplementation of glucose for the next phase of fermentation; glucose was

added to ensure that there were sufficient nutrients for the subsequently inoculated NCYC

500. The trend observed for the decrease in oBrix values correlated with the trend for yeast

Page 93: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

82

cell growth, especially the growth of S. cerevisiae (Figure 5.2). This is easily explained by

the fact that S. cerevisiae consumed sugars at a much faster rate than W. MRAKII.

Figure 5.1 Changes in oBrix values during fermentation for PSF

1 ( ), NSF

2 ( )

and MCF3 ( )

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication.

S. cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii

was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0

in the ratio of 1:1000. Fermentation was ceased at Day 21

5.2.2 Yeast biomass evolution

The evolution of S. cerevisiae and W. mrakii for the three types of fermentations is

shown in Figure 5.2. In PSF, W. mrakii reached a level of 108

CFU/mL in the sterile mango

juice before being killed off via ultrasonication. The absence of the Saccharomyces yeast

allowed proliferation of W. mrakii, giving W. mrakii longer persistence to produce esters.

Interestingly, the growth kinetics of S. cerevisiae when inoculated after the demise of W.

mrakii was different from that of both NSF and MCF. Despite its suitability for alcoholic

0

2

4

6

8

10

12

14

16

18

20

0 5 10 15 20

TS

S (°B

)

Time (Days)

Addition of glucose

Page 94: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

83

fermentation and the presence of sufficient nutrients in the juice, S. cerevisiae cell population

declined almost immediately after inoculation. It is unlikely that the primary cause of the

death of S. cerevisiae in sequential culture was ethanol because ethanol production and sugar

consumption by W. mrakii were low. Nutrient depletion could be ruled out because similar

studies demonstrated that S. cerevisiae with higher inoculum levels could still survive (Li

2013). Consequently, it is likely that the typically robust S. cerevisiae may have failed to

thrive due to the presence of mycosins or killer toxins produced by W. mrakii (Liu and Tsao

2010), to which S. cerevisiae was sensitive (Yap et al., 2000). Previous studies have also

shown that W. mrakii could produce mycotoxins that affects the synthesis of yeast cell walls

by inhibiting the β-1,3-glucan synthesis occurring at a budding site or a conjugating tube,

resulting in cell death (Yamamoto et al. 1986; Inoue et al. 1995; Kimura et al. 1999; Guyard

et al. 2002). Therefore, despite the completion of fermentation, it is likely that W. mrakii

might have inhibited the growth of S. cerevisiae due to the accumulation of mycosin.

Therefore, it might be necesssary for a S. cerevisiae to be inoculated at a higher dosage.

Page 95: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

84

Figure 5.2 Changes in (a) S. cerevisiae MERIT.ferm for PSF1 ( ), NSF

2 ( ) and MCF

3 ( ),

(b) W. mrakii NCYC 500 cell population during fermentation for PSF1 ( ), NSF

2 ( ) and

MCF3 ( ), (c) Changes in S. cerevisiae MERIT.ferm ( ) and W. mrakii NCYC 500 ( ) in

MCF

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S. cerevisiae was

then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii was then inoculated

at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in the ratio of 1:1000. Fermentation was ceased at Day 21

0

1

2

3

4

5

6

7

8

9

0 5 10 15 20

Log

CF

U/m

L

Time (Days)

(a)

0

1

2

3

4

5

6

7

8

9

10

0 5 10 15 20

Log

CF

U/m

L

Time (Days)

(b)

0

1

2

3

4

5

6

7

8

0 5 10 15 20

Log

CF

U/m

L

Time (Days)

(c)

Page 96: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

85

On the other hand, it is likely that S. cerevisiae may inhibit W. mrakii due to high

ethanol production. In addition, studies have indicated that W. mrakii produced and

accumulated toxins in the stationary phase (Li 2013). Therefore, it is likely that W. mrakii

was unable to kill S. cerevisiae in MCF because S. cerevisiae already reached a high cell

density and produced sufficient ethanol to inhibit W. mrakii before the amount of toxin

compounds produced by W. mrakii rose above its lethal concentration. This is supported by a

study that showed that more than 6% (v/v) ethanol significantly inhibited the growth of W.

mrakii (Li 2013). This result was consistent with another report on the behaviour of

Williopsis in Japanese sake, demonstrating the inhibitory effect of 6 to 7% ethanol on

Williopsis (Inoue et al. 1994).

For NSF, S. cerevisiae reached a similar maximal population to those in the single

culture fermentations, but NCYC 500 yeast inoculated after the forced demise of the S.

cerevisiae yeast was unable to attain the maximal cell population obtained in the single

culture fermentations in Chapter 4. Maximal W. mrakii population only reached 106

CFU/mL, as opposed to 108 CFU/mL in the single culture fermentation and PSF. This could

be due to the toxic effects of ethanol as discussed previously.

Page 97: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

86

5.2.3 Volatile composition

A wide variety of volatile compounds could be produced from different inoculation

strategies due to the metabolic interactions between the different yeast species (Ciani et al.

2010), and potentially the chemical reactions of the different metabolites and aroma

compounds that are produced during the different stages of the fermentation. These volatile

compounds include fatty acids, alcohols, aldehydes, esters, ketones, volatile phenols and

terpenoids. Although similar compounds were detected in all three fermentations, the

amounts of these volatile compounds differed rather significantly. Generally, PSF appeared

to produce higher amounts of desirable volatile compounds while NSF produced lower

amounts of most volatiles. This may be due to the effects of ethanol toxicity as the amount of

ethanol produced in NSF appears to be significantly higher than PSF. The complete list of

volatiles and their RPA can be found in Table 5.2 and the impact odourants quantified can be

found in Table 5.3.

Page 98: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

87

Table 5.2 Complete volatiles for mango wine produced from different inoculation strategies

CASA LRI

B Compound PSF

1 MCF

2 NSF

3 Odour descriptor

C

Peak area (×106) RPA (%) Peak area (×10

6) RPA (%) Peak area

(×106)

RPA (%)

000064-17-5 1028 Ethanol 1783 ± 137 a 78.219 2187 ± 63

b 73.05 2775 ± 68

c 86.2 Alcoholic

Subtotal 1783 78.22 2187 73.05 2775 86.2

Alcohols 000071-23-8 1039 Propanol 1.24 ± 0.15 a 0.054 1.37 ± 0.01

a 0.046 2.79 ± 0.03

c 0.087 Alcoholic, fusel

000078-83-1 1172 Isobutyl alcohol 14.68 ± 1.49 a 0.644 15.5 ± 0.44

b 0.518 21.01 ± 1.08

c 0.653 Fruity, wine-like

000123-51-3 1237 Isoamyl alcohol 48.48 ± 0.58 a 2.126 43.29 ± 0.65

b 1.445 65.97 ± 0.17

c 2.049 Alcoholic, fruity, banana

000928-96-1 1475 cis-3-hexenol 0.25 ± 0.01 a 0.011 0.38 ± 0.02

b 0.013 0.3 ± 0.01

c 0.009 Green, grassy

000505-10-2 1706 Methionol 0.69 ± 0.02 a 0.03 1.37 ± 0.01

b 0.046 0.82 ± 0.01

c 0.026 Cooked cabbage

000078-70-6 1999 Linalool 3.16 ± 0.06 a 0.139 1.43 ± 0.07

b 0.048 4.8 ± 0.08

c 0.149 Mild floral

000562-74-3 1691 4-Terpineol 0.3 ± 0.01 a 0.013 0.11 ± 0.03

b 0.003 0.24 ± 0.05

c 0.008 Woody

000098-55-5 1716 α-Terpineol 0.66 ± 0.04 a 0.029 1.23 ± 0.02

b 0.041 0.84 ± 0.01

c 0.026 Floral, lilac

000470-08-6 1794 β-Fenchol 0.12 ± 0.009 a 0.005 0.15 ± 0.01

b 0.005 0.13 ± 0.01

a 0.004 Camphor, pine, woody

000106-22-9 1867 β-Citronellol 0.11 ± 0.008 a 0.004 N.D - 0.27 ± 0.01

b 0.008 Floral, rosy, citrus

000106-25-2 1833 Nerol 0.25 ± 0.01 a 0.011 0.59 ± 0.02

b 0.02 0.89 ± 0.01

c 0.028 Ethereal, cognac, fruity

000106-24-1 1834 Geraniol 1.99 ± 0.1 a 0.087 0.73 ± 0.01

b 0.024 1.75 ± 0.06

c 0.054 Sweet, floral, rose-like

000060-12-8 2035 2-Phenylethyl alcohol 42.53 ± 0.71 a 1.865 87.52 ± 1.56

b 2.922 95.01 ± 4.43

c 2.951 Rose, honey, floral

00499-75-2 2236 Carvacrol 0.29 ± 0.04 a 0.013 0.68 ± 0.05

b 0.023 0.21 ± 0.01

c 0.006 Spicy, woody, herbal

Subtotal 114.74 5.03 154.34 5.15 195 6.06

Ethyl esters 000141-78-6 1009 Ethyl acetate 59.13 ± 0.15 a 2.593 14.24 ± 0.28

b 0.476 1.92 ± 0.11

c 0.06 Ethereal, fruity, sweet

000105-54-4 1034 Ethyl butyrate 0.41 ± 0.02 a 0.018 0.32 ± 0.01

b 0.011 N.D - Sweet, fruity

000123-66-0 1297 Ethyl hexanoate 22.33 ± 1.07 a 0.979 6.28 ± 0.22

b 0.21 13.25 ± 0.36

c 0.412 Sweet, pineapple, fruity

054653-25-7 1392 Ethyl 5-hexenoate N.D - 2.51 ± 0.03 a 0.084 0.05 ± 0

b 0.002 Fruity, pineapple

064187-83-3 1358 Ethyl cis-3-hexenoate 0.16 ± 0.007 a 0.007 0.12 ± 0.002

b 0.003 0.06 ± 0.005

c 0.002 Fruity, sweet, apple

000106-30-9 1369 Ethyl heptanoate 0.2 ± 0.005 a 0.009 0.16 ± 0.008

b 0.005 0.19 ± 0.01

a 0.006 Sweet,

Page 99: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

88

Table 5.2 (Continued)

000106-32-1 1453 Ethyl octanoate 78.71 ± 1.53 a 3.451 138.77 ± 2.14

b 4.634 58.52 ± 0.66

c 1.817 Sweet, fruity

035194-38-8 1486 Ethyl 7-octenoate 0.14 ± 0.02 a 0.006 0.36 ± 0.03

b 0.012 0.29 ± 0.04

c 0.009 Tropical, fruity

000123-29-5 1624 Ethyl nonanoate 0.07 ± 0.003 a 0.003 0.37 ± 0.04

b 0.012 0.09 ± 0.01

a 0.003 Cognac

000110-38-3 1746 Ethyl decanoate 72.97 ± 0.66 a 3.2 246.23 ± 5.47

b 8.222 36.77 ± 0.82

c 1.142 Sweet, fruity, pineapple

067233-91-4 1795 Ethyl 9-decenoate 5.49 ± 4.13 a 0.241 63.07 ± 0.6

b 2.106 16.47 ± 0.47

c 0.512 -

000106-33-2 1887 Ethyl dodecanoate 19.66 ± 0.68 a 0.862 36.94 ± 0.26

b 1.233 0.84 ± 0.01

c 0.026 Sweet, waxy, fruity

000692-86-4 1986 Ethyl undecenoate 0.66 ± 0.05 a 0.029 0.52 ± 0.01

b 0.017 0.71 ± 0.01

c 0.022 Waxy, fruity, creamy, cognac

000124-06-1 2201 Ethyl tertradecanoate 0.64 ± 0.05 a 0.028 0.68 ± 0.04

a 0.023 0.5 ± 0.01

c 0.015 Sweet, waxy, creamy

000628-97-7 2373 Ethyl hexadecanoate 0.42 ± 0.01 a 0.019 0.43 ± 0.03

a 0.014 0.25 ± 0.01

c 0.008 Waxy, fruity, creamy,

balsamic

054546-22-4 2402 Ethyl 9-hexadecenoate 0.87 ± 0.03 a 0.038 0.8 ± 0.01

a 0.027 0.26 ± 0.01

c 0.008 -

Subtotal 261.86 11.483 511.78 17.09 130.17 4.051

Acetate esters 000110-19-0 1020 Isobutyl acetate 0.66 ± 0.04 a 0.029 0.51 ± 0.04

b 0.015 0.06 ± 0.01

c 0.002 Sweet, fruity, banana

000123-92-2 1112 Isoamyl acetate 23.71 ± 1.6 a 1.04 4.25 ± 0.16

b 0.128 1.52 ± 0.02

c 0.047 Sweet, fruity, banana

000142-92-7 1206 n-Hexyl acetate 1.36 ± 0.06 a 0.06 0.43 ± 0.01

b 0.142 0.13 ± 0.01

c 0.004 Fruity, green

003681-71-8 1324 cis-3-Hexenyl acetate 5.49 ± 0.27 a 0.241 2.93 ± 0.13

b 0.098 1.25 ± 0.08

c 0.038 Fresh, green, fruity

000112-06-1 1546 Heptyl acetate 0.12 ± 0.011 a 0.005 N.D - N.D - Fresh, green, rum

000150-84-5 1659 Citronellyl acetate 1.05 ± 0.11 a 0.046 4.15 ± 0.15

b 0.139 N.D - Floral, green rose, fruity

000112-17-4 1726 Decyl acetate 0.21 ± 0.01 a 0.009 0.11 ± 0.009

b 0.003 0.16 ± 0.02

c 0.005 Waxy, soapy, fatty

000105-87-3 1790 Geranyl acetate 0.05 ± 0.006 a 0.002 2.64 ± 0.04

b 0.088 N.D - Floral, rose, lavender

000103-45-7 1862 2-Phenylethyl acetate 25.15 ± 2.73 a 1.102 41.23 ± 1.15

b 1.377 31.1 ± 0.73

c 0.969 Sweet, honey, floral, rosy

Subtotal 57.81 2.535 55.76 1.862 34.24 1.063

Higher esters 000624-13-5 1523 Propyl octanoate N.D - 0.16 ± 0.009 a 0.005 0.13 ± 0.0081

b 0.004 Coconut, cocoa, gin

002035-99-6 1762 Isoamyl octanoate 2.34 ± 0.11 a 0.103 9.5 ± 0.13

b 0.319 2.77 ± 0.07

c 0.087 Sweet, fruity,

000112-32-3 1803 Octyl formate 1.53 ± 0.21 a 0.067 N.D - 1.59 ± 0.04

a 0.049 Fruity, rose, orange

030673-38-2 1859 Isobutyl decanoate 1.11 ± 0.013 a 0.011 1.03 ± 0.069

b 0.033 0.29 ± 0.015

c 0.009 Oily, sweet, brandy, apricot,

cognac

Page 100: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

89

Table 5.2 (Continued)

002306-91-4 1973 Isoamyl decanoate 1.18 ± 0.16 a 0.052 1.96 ± 0.12

b 0.065 1.57 ± 0.03

c 0.049 Waxy, fruity, sweet, cognac

006309-51-9 2180 Isoamyl dodecanoate 0.09 ± 0.01 a 0.004 0.15 ± 0.009

b 0.005 0.14 ± 0.01

b 0.004 Cognac, green, waxy

006290-37-5 2261 2-Phenylethyl hexanoate 0.32 ± 0.01 a 0.014 N.D - 0.33 ± 0.01

a 0.01 Sweet, honey, floral

005457-70-5 2419 2-Phenylethyl octanoate 0.36 ± 0.02 a 0.016 N.D - 0.46 ± 0.01

b 0.014 Sweet, waxy, green, cocoa

Subtotal 6.93 0.305 12.81 0.427 7.28 0.225

Organic acids 000064-19-7 1549 Acetic acid 4.39 ± 0.2 a 0.193 0.72 ± 0.043

b 0.024 5.24 ± 0.04

c 0.163 Sharp, pungent, sour, vinegar

000079-31-2 1575 Isobutyric acid 0.6 ± 0.04 a 0.026 0.55 ± 0.024

b 0.018 0.55 ± 0.02

b 0.017 Acidic

000067-43-6 1728 Butanoic acid 0.46 ± 0.03 a 0.02 0.23 ± 0.01

b 0.008 0.43 ± 0.01

a 0.013 Sharp, dairy-like, cheesy,

000142-62-1 1890 Hexanoic acid 0.49 ± 0.36 a 0.021 0.72 ± 0.027

b 0.024 0.72 ± 0.45

b 0.022 Sour, fatty, sweat, cheesy

000124-07-2 2170 Octanoic acid 13.26 ± 1.25 a 0.581 19.8 ± 0.36

b 0.661 20.43 ± 0.57

c 0.635 Fatty, waxy, rancid

000373-49-9 2853 9-Hexadecenoic acid 0.27 ± 0.01 a 0.012 0.29 ± 0.016

a 0.01 0.09 ± 0.003

b 0.003 Waxy

000334-48-5 2390 Decanoic acid 11.12 ± 1.07 a 0.487 16.33 ± 0.06

b 0.545 7.59 ± 0.01

c 0.236 Rancid, sour, fatty

014436-32-9 2474 9-Decenoic acid 0.46 ± 0.03 a 0.02 2.36 ± 0.05

b 0.079 2.21 ± 0.1

c 0.069 Waxy, green, fatty, soapy

000065-85-0 2455 Benzoic acid 0.5 ± 0.01 a 0.022 0.87 ± 0.02

b 0.029 0.47 ± 0.06

a 0.015 Faint balsam, urine

000143-07-7 2544 Dodecanoic acid 1.04 ± 0.06 a 0.046 1.92 ± 0.07

b 0.064 1.63 ± 0.15

c 0.051 Fatty, coconut, bay oil

Subtotal 32.59 1.428 43.79 1.462 39.36 1.224

Terpenes 013466-78-9 1206 δ-3-Carene N.D - 0.19 ± 0.01 b 0.006 N.D - Sweet citrus

000123-35-3 1246 β-Myrcene 0.43 ± 0.02a 0.019 0.17 ± 0.02

b 0.006 0.29 ± 0.06

c 0.009 Terpenic, herbaceous, woody,

citrus

000099-86-5 1235 α-Terpinene 0.32 ± 0.02 a 0.014 0.12 ± 0.01

b 0.004 0.14 ± 0.03

b 0.004 Sharp, terpenic, lemon

095327-98-3 1254 Limonene 0.12 ± 0.01 a 0.005 0.22 ± 0.03

b 0.007 0.19 ± 0.01

b 0.006 Citrius, terpenic, orange note

027400-71-1 1290 β-Ocimene N.D - 0.101 ± 0.02 a 0.003 0.13 ± 0.02

b 0.004 Citrus, green, lime

000100-42-5 1403 p-α-Dimethyl styrene 0.23 ± 0.01 a 0.01 0.18 ± 0.03

a 0.006 N.D - Sweet, balsam, floral

000099-87-6 1339 p-Cymene 1.47 ± 0.16 a 0.064 2.01 ± 0.02

b 0.066 4.92 ± 0.10

c 0.041 Citrus, terpenic, woody

001195-32-0 1529 p-Cymenene 9.19 ± 0.24 a 0.403 4.25 ± 0.2

b 0.142 7.64 ± 0.05

c 0.237 Spicy, balsamic, musty

000535-77-3 1343 m-Cymene 0.06 ± 0.001 a 0.003 N.D - N.D - Citrus, terpenic, woody

Page 101: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

90

Table 5.2 (Continued)

000087-44-5 1695 β-Caryophyllene 0.13 ± 0.01 a 0.006 0.2 ± 0.005

b 0.007 0.07 ± 0.009

b 0.002 Woody, clove note

006753-98-6 1725 α-caryophyllene N.D - 0.92 ± 0.01 0.031 N.D - Woody

017066-67-0 1829 β-Selinene 0.12 ± 0.01 a 0.005 0.21 ± 0.03

b 0.007 0.15 ± 0.01

c 0.005 Herbal

000079-92-5 1117 Camphene 0.32 ± 0.03 a 0.014 0.34 ± 0.01 0.012 9.3 ± 0.38 0.289 Camphoraceous, cooling,

minty

000586-62-9 1352 α-Terpinolene 0.75 ± 0.05 a 0.033 1.86 ± 0.02 0.062 0.64 ± 0.02 0.02 Citrus, lime, pine

Subtotal 14.34 0.628 10.93 0.365 23.47 0.729

Aldehydes 000075-07-0 939 Acetaldehyde 2.76 ± 0.06 a 0.121 10.16 ± 0.53

b 0.339 3.46 ± 0.07

c 0.108 Green apple, fresh

000104-87-0 1773 p-Tolualdehyde 0.57 ± 0.04 a 0.025 1.39 ± 0.03

b 0.046 5.23 ± 0.03

c 0.162 Pungent, fresh, green

015764-16-6 1836 2, 4-Dimethyl

benzaldehyde

0.49 ± 0.03 a 0.022 0.55 ± 0.01

b 0.018 1.3 ± 0.05

c 0.04 Sweet, almond, cherry

000100-52-7 1637 Benzaldehyde 1.03 ± 0.09 a 0.045 2.11 ± 0.05

b 0.07 1.07 ± 0.08

a 0.033 Almond, fruity, nutty

Subtotal 4.85 0.213 14.21 0.473 11.06 0.343

Ketone 023696-85-7 1938 β-Damascenone 0.94 ± 0.05 a 0.041 N.D - N.D - Sweet, floral, fruity, coconut

000821-55-6 2-Nonanone 1.52 ± 0.13 a 0.067 1.32 ± 0.02

b 0.044 2.21 ± 0.1

c 0.069 Fruity, sweet, cheese

000513-86-0 1401 Acetoin 0.17 ± 0.02 a 0.007 0.25 ± 0.33

b 0.008 0.16 ± 0.02

a 0.005 Sweet, buttery, creamy

Subtotal 1.69 0.074 1.57 0.052 2.37 0.074

Others 016409-43-1 1434 cis- Rose oxide 0.61 ± 0.02 a 0.027 0.38 ± 0.01

b 0.013 0.39 ± 0.02

c 0.012 Rose, geranium-like

trans-Rose oxide 0.37 ± 0.01 a 0.016 0.19 ± 0.01

b 0.006 0.11 ± 0.01

c 0.004 Floral, rose-like

001786-08-9 1563 Nerol oxide 0.75 ± 0.05 a 0.033 1.23 ± 0.02

b 0.041 0.51 ± 0.01

c 0.016 Floral, orange blossom sweet

Subtotal 1.73 0.076 1.8 0.06 1 0.032

Grand total 2280.46 100 2994.94 100 3219.6 100

N.D. Not detected A

CAS numbers were obtained from Wiley database library

B LRI of all the relative tables was determined on the DB-FFAP column, relative to C5-C40 hydrocarbons.

Page 102: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

91

C Descriptors were retrieved from http://www.thegoodscentscompany.com

abc ANOVA (n=4)at 95% confidence level with the same letters in the same row indicating no significant difference

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S. cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in the ratio of 1:1000. Fermentation was ceased at Day 21

Page 103: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

92

Table 5.3 Major volatiles quantified for mango wines with different inoculation strategies CAS

A LRI

B PSF

1 MCF

2 NSF

3 Odour

threshold

OrganolepticsC

mg/L OAV mg/L OAV mg/L OAV (mg/L)

Alcohol 000078-83-1 1172 Isobutyl Alcohol 8.21 ± 0.61

b 0.21 10.1 ± 0.75

a 0.25 15.32 ± 1.22

a 0.38 40 Fruity, wine-like

000123-51-3 1237 Isoamyl Alcohol 150.28 ± 8.98

b 5.01 83.87 ± 5.01

a 2.8 142.09 ± 17.60

b 4.74 30 Banana-like, sweet, fruity

000078-70-6 1699 Linalool Trace - 0.012 ± 0.001

a 0.46 0.028 ± 0.001

b 1.12 0.025 Fresh floral, herbal, rosewood

000060-12-8 2035 2-Phenylethyl Alcohol 14.82 ± 1.42

b 1.48 42.46 ± 4.07

a 4.25 44.51 ± 0.69

a 4.45 10 Floral, rose-like

Esters 000141-78-6 1009 Ethyl acetate 73.22 ± 7.04

b 9.76 34.51 ± 5.35

a 4.6 26.51 ± 2.74

c 3.53 7.5 Ethereal, fruity, sweet,

000110-19-0 1020 Isobutyl acetate 0.02 ± 0.01

a 0.01 0.02 ± 0.01

a 0.01 0.01 ± 0.00

a 0.01 1.6 Sweet, fruity, tropical

000123-92-2 1112 Isoamyl acetate 0.68 ± 0.14

b 2.28 0.86 ± 0.18

a 2.85 0.09 ± 0.03

c 0.3 0.3 Sweet, fruity, banana

000123-66-0 1297 Ethyl hexanoate 0.61 ± 0.20

b 43.43 0.36 ± 0.12

a 25.48 0.25 ± 0.17

a 17.58 0.014 Sweet, pineapple

000106-32-1 1453 Ethyl octanoate 0.73 ± 0.30

b 366.27 1.09 ± 0.44

a 546.75 0.55 ± 0.24

c 274 0.002 Sweet, fruity

000110-38-3 1746 Ethyl decanoate 1.01 ± 0.28

a 5.05 0.96 ± 0.27

a 4.81 0.84 ± 0.07

c 4.18 0.2 Sweet, fruity, apple

000103-45-7 1862 2-Phenylethyl acetate 0.77 ± 0.06

b 3.06 0.88 ± 0.07

a 3.52 0.61 ± 0.05

c 2.44 0.25 Sweet, honey, floral, rosy

000106-33-2 1887 Ethyl dodecanoate 0.91 ± 0.09

a 0.16 1.05 ± 0.10

a 0.18 0.95 ± 0.04

a 0.16 5.9 Sweet, waxy, soapy

Acid 000064-19-7 1549 Acetic Acid 765.91 ± 33.28

b 3.83 597.6 ± 33.77

a 2.99 777.95 ± 31.29

b 3.89 200 Vinegar-like, pungent

000142-62-1 1890 Hexanoic acid 4.45 ± 0.04

b 1.48 3.13 ± 0.03

a 1.04 4.02 ± 0.29

c 1.34 3 Sour, fatty, cheese

000124-07-2 2170 Octanoic Acid 5.4 ± 1.00

b 0.61 4.16 ± 0.77

a 0.47 6.69 ± 0.51

c 0.76 8.8 Fatty, rancid

000334-48-5 2390 Decanoic Acid 4.43 ± 0.22

b 0.44 4.07 ± 0.20

a 0.41 4.3 ± 0.46

b 0.43 10 Rancid, sour, fatty

000143-07-7 2544 Dodecanoic Acid 5.56 ± 0.11

b 0.56 5.44 ± 0.11

a 0.54 5.53 ± 0.48

b 0.55 10 Mild fatty, coconut

Trace: Compound detected but below linear range. A

CAS numbers were obtained from Wiley database library

B LRI of all the relative tables was determined on the DB-FFAP column, relative to C5-C40 hydrocarbons.

Page 104: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

93

C Descriptors were retrieved from http://www.thegoodscentscompany.com

abc ANOVA (n=4)at 95% confidence level with the same letters indicating no significant difference

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S. cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in the ratio of 1:1000. Fermentation was ceased at Day 21

Page 105: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

94

5.2.3.1 Alcohols

The main product of alcoholic fermentation is ethanol. By RPA, PSF produced the

least amount of ethanol while NSF produced the most ethanol. Generally, NSF produced the

highest amounts of other alcohols while PSF produced the lowest amounts of other alcohols.

This may have an effect on wine quality since excessive amounts of alcohols could mask the

other volatile aromatic compounds (Swiegers et al. 2005).

Ethanol production rate was significantly higher in NSF than in MCF despite both

fermentations were dominated by the high ethanol producing S. cerevisiae (Figure 5.3). As

discussed previously, this is likely due to the toxic effect of the mycosins secreted by W.

mrakii yeast affecting the metabolism of the S. cerevisiae. It is likely that while mycosin

accumulation was not rapid enough to kill the S. cerevisiae yeast, mycosins affected S.

cerevisiae metabolism, resulting in the slower growth rate and production of ethanol. In

MCF, the ethanol content remained unchanged after Day 11, indicating the end of alcoholic

fermentation. This is in general agreement with the trend observed for °Brix values (Figure

5.1), which indicated no decrease in TSS after Day 7.

There was a slight decline in the amount of ethanol after the inoculation of W. mrakii at

Day 7 for NSF (Figure 5.3). This decline could be due to the effects of ultrasonication as the

vibrations could have caused the loss of volatile compounds via evaporation. It is unlikely

that the ethanol was converted into ethyl acetate by W. mrakii as the amount of ethyl acetate

did not increase after Day 7 (Figure 5.6). Although there were residual sugars at the end of

the fermentation, this is likely due to the addition of glucose just before the inoculation W.

mrakii, which consumed sugars very slowly, hence resulting in the residual sugars.

For PSF, rate of ethanol production was very slow for the W. mrakii yeast for the first

14 days of fermentation (Figure 5.3). This is expected as W. mrakii is not a vigorous

Page 106: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

95

fermenter. Ethanol production increased rapidly after Day 14, upon the inoculation of the

high ethanol producing S. cerevisiae. At the point of cessation, there were only trace

amounts of sucrose detected (Table 5.1), indicating a complete fermentation.

Figure 5.3 Changes in ethanol concentration during PSF1 ( ), NSF

2( ),

MCF3 ( )

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S.

cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii

was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in

the ratio of 1:1000. Fermentation was ceased at Day 21

After ethanol, the major alcohols found in all three mango wines were isoamyl alcohol,

2-phenylethyl alcohol and isobutyl alcohol (Figure 5.4). NSF produced the most fusel

alcohol at the fastest rate for the first seven days of the fermentation. MCF also produced

significant amounts of fusel alcohols, but at slower rate and lesser amounts than NSF. PSF

produced the lowest amounts of fusel alcohols consistently.

The trend observed for isoamyl alcohol, 2-phenylethyl alcohol and isobutyl alcohol for

NSF was slightly different. Isoamyl alcohol increased gradually throughout the fermentation.

0.E+00

5.E+08

1.E+09

2.E+09

2.E+09

3.E+09

3.E+09

4.E+09

4.E+09

0 5 10 15 20

Pea

k A

rea

Time (Days)

Ethanol

Page 107: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

96

Isobutyl alcohol increased rapidly from Day 0 to Day 3 before the rate of increase became

more gradual. On the other hand, 2-phenylethyl alcohol showed a modest increase from Day

0 to Day 3 before a drastic increase from Day 3 to Day 7. The initial increase is expected as S.

cerevisiae are known to be efficient fusel alcohol producers. After the demise of S.

cerevisiae at Day 7, the rate of increase in the fusel alcohols generally decreased. W.mrakii

generally do not produce much fusel alcohol. this is also observed in the initial stages of PSF

where only W. mrakii was present.

Isobutyl alcohol was somewhat different from both isoamyl alcohol and 2-phenylethyl

alcohol. Isoamyl alcohol and 2-phenylethyl alcohol only showed substantial increases after

the demise of W. mrakii and inoculation of S. cerevisiae. However, isobutyl alcohol showed

a steady increase from the beginning of the fermentation. At the end of fermentation, the

amount of isobutyl alcohol was similar in NSF and MCF. This was not a trend observed for

either isoamyl alcohol or 2-phenylethyl alcohol.

These higher alcohols are known for the floral and fruity notes that they impart and

their positive contribution to wine flavour if present at desirable levels (below 350 mg/L)

(Swiegers et al. 2005). Isoamyl alcohol and 2-phenylethyl alcohol were both detected at

levels above their threshold values (Table 5.3) and would likely contribute fruity and floral

notes to the overall mango wine flavour profile. On the other hand, isobutyl alcohol was

produced at levels below its odour threshold value; hence its influence may be negligible.

Page 108: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

97

Figure 5.4 Changes in isoamyl alcohol, 2-phenylethyl alcohol, isobutyl

alcohol during PSF1 ( ), NSF

2( ), MCF

3 ( )

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S.

cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii

was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in

the ratio of 1:1000. Fermentation was ceased at Day 21

0.E+00

2.E+07

4.E+07

6.E+07

8.E+07

1.E+08

0 5 10 15 20

Pea

k A

rea

Time (Days)

Isoamyl alcohol

0.E+00

2.E+07

4.E+07

6.E+07

8.E+07

1.E+08

1.E+08

1.E+08

0 5 10 15 20

Pea

k A

rea

Time (Days)

2-Phenylethyl alcohol

0.E+00

5.E+06

1.E+07

2.E+07

2.E+07

3.E+07

3.E+07

0 5 10 15 20

Pea

k A

rea

Time (Days)

Isobutyl alcohol

Page 109: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

98

Some of the terpene alcohols, or terpenols, detected were gerianol, α-terpineol, nerol, 4-

terpineol and citronellol (Table 5.3). These terpenols have been identified to be important

compounds in the aroma of wine from aromatic grape varieties including Muscat, Riesling,

and other aromatic grape varieties (Marais 1983; Swiegers et al. 2005). Amongst these

terpenols, linalool was the only one produced at levels above its threshold odour in NSF

(Table 5.2); the delayed inoculation of W. mrakii appeared to have a positive effect on the

production of terpene alcohols. For all three fermentations, linalool increased from Day 0 to

Day 3 before displaying different trends (Figure 5.5). In MCF, linalool decreased until Day

11 before showing a slight increase from Day 11 to Day 21 (Figure 5.5). On the other hand,

for PSF, linalool content did not change significantly after Day 3 (Figure 5.5). In NSF,

linalool decreased from Day 3 to Day 7 before increasing from Day 7 to Day 21 (Figure 5.5).

This appears to imply that W. mrakii has the ability to either release bound linalool glycosides

or transform other terpenols such as geraniol and/or nerol into linalool (King and Dickinson

2000). It is unlikely that the increase in linalool was due to de novo synthesis by W. mrakii

since the strain has not been reported to possess such a capability.

Page 110: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

99

Figure 5.5 Changes in linalool during PSF1 ( ), NSF

2( ), MCF

3 ( )

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S.

cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii

was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in

the ratio of 1:1000. Fermentation was ceased at Day 21

Although the rest of the terpenols were at low levels, it is probable that they may

interact in a synergistic manner to contribute positively to the overall flavour profile due to

the fruity and floral notes associated with these aromatic compounds. The low levels of these

terpenols could also be due to their degradation during the fermentation period (King and

Dickinson 2000), such as the trend observed for MCF in this case. Furthermore, the

differences between the concentrations of the various alcohols could be due to the different

metabolic activities of the yeasts.

The amounts of geraniol and nerol were also significantly different amongst the three

treatments and this is likely due to geraniol being converted into citronellol and other

terpenols while nerol can be converted into α-terpineol, terpinolene and limonene, de novo by

S. cerevisiae through an alternative pathway involving leucine (Carrau et al. 2005).

0.E+00

1.E+06

2.E+06

3.E+06

4.E+06

5.E+06

6.E+06

7.E+06

0 5 10 15 20

Pea

k A

rea

Time (Days)

Linalool

Page 111: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

100

S. cerevisiae do not efficiently excrete monoterpenes but synthesize the phosphorylated

form of geraniol, geranyl diphosphate (GDP), as intermediate of farnesyl diphosphate

synthesis, a key molecule in the isoprenoid pathway that leads to the synthesis of dolicols,

ubiquionones, and sterols. It has also been shown that yeasts use terpenes as biosynthetic

intermediates for sterol synthesis. Therefore, differences in the terpene profile of wines

probably depend on beta-glucosidase activity, terpene bioconversion rate and the percentage

of terpenes accumulated by yeasts (Sadoudi, et al. 2012).

5.2.3.2 Esters

Esters are important contributors to the fruity flavours of alcoholic beverages (Russell,

2003). The major esters produced in all three fermentations were mainly ethyl octanoate,

ethyl decanoate, ethyl dodecanoate, 2-phenylethyl acetate and isoamyl acetate (Table 5.2).

The ethyl esters were generally found in the highest concentration in MCF and could be

attributed to efficient fatty acid metabolism of S. cerevisiae and the longer exposure time for

S. cerevisiae during fermentation as opposed to the shorter time for both NSF and PSF.

Amongst these esters, ethyl octanoate could have the most influence on the wine profile due

to its very low odour threshold value and the significant amounts produced. Ethyl acetate and

ethyl decanoate could also have significant influences on the overall flavour profile as their

OAVs all exceeded one (

Page 112: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

101

Table 5.3).

The time-course formation of three ethyl esters is presented in Figure 5.6. In PSF, ethyl

octanoate and ethyl decanoate both increased from Day 15 to the end of the fermentation with

minimal changes during the first 14 days (Figure 5.6). This is likely due to the fatty acid

metabolism of the S. cerevisiae yeast inoculated at Day 14. The trend indicated that W.

mrakii is a low ethyl ester producer. In MCF, ethyl octanoate and ethyl decanoate increased

slowly from Day 0 to Day 3 before showing significant increases from Day 3 onwards

(Figure 5.6). This can be explained by the dominant yeast strain present. W. mrakii was the

dominant strain in the beginning of the fermentation, hence, resulting in the lacklustre

production of ethyl esters. However, as the S. cerevisiae population increased and dominated

the fermentation, S. cerevisiae yeast became prominent.

Low amounts of all three ethyl esters were produced in NSF throughout the

fermentation (Figure 5.6). Although it was hypothesised that the higher S. cerevisiae

population in NSF would give rise to higher concentrations of ethyl esters as compared to

MCF (Rojas et al. 2003), this was not observed in this study; most esters were found in lower

concentrations in NSF (Table 5.2). The likely cause could be the demise of S. cerevisiae at

Day 7. The lack of time for S. cerevisiae metabolism could be the likely reason for the low

amounts of ethyl ester formation. Furthermore, the subsequently inoculated W. mrakii could

possibly have caused the hydrolysis of ethyl esters, as illustrated by the decrease in all three

ethyl esters after Day 7.

Page 113: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

102

Figure 5.6 Changes in (a) ethyl octanoate, (b) ethyl decanoate, (c) ethyl

acetate during PSF1 ( ), NSF

2( ), MCF

3 ( )

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S.

cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii

was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in

the ratio of 1:1000. Fermentation was ceased at Day 21

0.E+00

2.E+07

4.E+07

6.E+07

8.E+07

1.E+08

1.E+08

1.E+08

2.E+08

0 5 10 15 20

Pea

k A

rea

Time (Days)

Ethyl octanoate

0.E+00 5.E+07 1.E+08 2.E+08 2.E+08 3.E+08 3.E+08 4.E+08 4.E+08

0 5 10 15 20

Pea

k A

rea

Time (Days)

Ethyl decanoate

0.E+00 2.E+07 4.E+07 6.E+07 8.E+07 1.E+08 1.E+08 1.E+08 2.E+08

0 5 10 15 20 25

Pea

k A

rea

Time (Days)

Ethyl acetate

Page 114: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

103

The major acetate esters produced during fermentation were 2-phenylethyl acetate and

isoamyl acetate; these compounds were found in the lowest amounts in NSF, with PSF and

MCF displaying mixed trends as illustrated by 2-phenylethyl acetate and isomayl acetate in

Figure 5.7.

Figure 5.7 Changes in 2-phenylethyl acetate and isoamyl acetate during

during PSF1 ( ), NSF

2( ), MCF

3 ( )

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S.

cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii

was then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in

the ratio of 1:1000. Fermentation was ceased at Day 21

W. mrakii was dominant in the initial stages of MCF before the more rigorous S.

cerevisiae became dominant. The rapid increase of 2-phenylethyl acetate and isoamyl acetate

0.E+00

1.E+07

2.E+07

3.E+07

4.E+07

5.E+07

6.E+07

7.E+07

0 5 10 15 20

Pea

k A

rea

Time (Days)

2-Phenylethyl acetate

0.E+00

2.E+07

4.E+07

6.E+07

8.E+07

1.E+08

1.E+08

1.E+08

2.E+08

0 5 10 15 20

Pea

k A

rea

Time (Days)

Isoamyl acetate

Page 115: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

104

in MCF from Day 0 to Day 7 was due to the acetate ester synthesising activity of W. mrakii.

This increase in 2-phenylethyl acetate is more significant in MCF than PSF due to the

production of 2-phenylethyl alcohol by S. cerevisiae concurrently. On the other hand, in the

absence of W. mrakii in NSF, despite the high amounts of 2-phenylethyl alcohol and isoamyl

alcohol present (Figure 5.4), significantly lower amounts of the corresponding acetate esters

were produced, likely due to the low acetate ester synthesising activities of S. cerevisiae. The

decline in both acetate esters after Day 7 in MCF could be due to esterase activities in S.

cerevisiae which dominated the fermentation. In PSF, there were no significant changes in

both esters from Day 7 to Day 14. A sharp decline was observed in isoamyl acetate after the

inoculation of S. cerevisiae while the decline in 2-phenylethyl acetate was not significant.

The discrepancy in the acetate hydrolysing enzymatic activity of S. cerevisiae could be

ascribed to the balance between the degradation and synthesis of esters governed by esterase

and alcohol acetyltransferase (Fukuda et al. 1998).

5.2.3.3 Acids

The main organic acids identified were acetic, octanoic and decanoic acids (Table 5.2).

The production of acetic acid was lower in MCF while no significant differences were found

between PSF and NSF. It is likely that the fatty acids are derived from yeast anabolic

pathways or β-oxidation of higher fatty acids (Tehlivets et al. 2007). Only acetic and

hexanoic acids were detected in the mango wine at levels above their odour threshold limits

(Table 5.2).

Acetic acid increased for the first three days for all three fermentations (Figure 5.8).

After which, NSF and MCF both showed significant decreases from Day 3 to Day 7. On the

other hand, acetic acid continued increasing steadily until Day 11 when a slight decrease

occurred from Day 11 to Day 14. From Day 14 to Day 21, a drastic increase in acetic acid

occurred. Although W. mrakii has been reported to produce higher amounts of acetic acid,

Page 116: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

105

this was not observed in this study. PSF, in which W. mrakii was present only in the first 14

days, did not show significantly more acetic acid production. The amount of acetic acid also

decreased upon the inoculation of S. cerevisiae (Figure 5.8).

On the other hand, MCF produced the most acetic acid during the first 14 days of

fermentation. This is likely due to the stress that the mycosin produced by the W. mrakii on

the S. cerevisiae since acetic acid is a sign of stress and redox imbalance as presented in

Chapter 3. However, the amount of acetic acid decreased drastically from Day 15 to Day 21.

The reason for this sharp decline is not clear, and further work should be conducted to verify

this trend.

For NSF, there was a rapid increase in acetic acid from Day 0 to 3 before a significant

decrease occurred (Figure 5.8). This could likely be due to the intial adaptation period

required for S. cerevisiae; the relatively high sugar concentration resulting in an increase in

acetic acid production (Devantier et al. 2005). However, it could be seen that W. mrakii once

inoculated, produced large amounts of acetic acid (Day 7 to 11). This is consistent with data

presented previously in Chapter 4 stating the W. mrakii produced higher amounts of acetic

acid relative to S. cerevisiae.

Hexanoic acid showed a different trend from acetic acid (Figure 5.8). Large increases

were observed from Day 0 to Day 5 for NSF and MCF. This is likely due to the action of S.

cerevisiae; the strain has efficient fatty acid metabolism and produced large amounts of

hexanoic acid rapidly once inoculated. A sharp decrease in hexanoic acid content was

observed after Day 7 after the inoculation of W. mrakii. In MCF, the amount of hexanoic

acid increased in a more gradual manner from Day 7 to Day 11 before decreasing gradually

from Day 11 to Day 15 and then decreasing drastically from Day 15 to Day 21. For PSF,

there were very little changes in the hexanoic acid content, even after the inoculation of S.

Page 117: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

106

cerevisiae at Day 14. Again, the lack of expected metabolic activity from the S. cerevisiae is

likely due to the effects of the mycosin produced by the W. mrakii yeasts.

Figure 5.8 Changes in acetic acid and hexanoic acid during PSF1 ( ), NSF

2( ),

MCF3 ( )

1 PSF W.mrakii was inoculated on Day 0, inactivated at Day 14 by ultrasonication. S.

cerevisiae was then inoculated at D14. Fermentation was ceased at Day 21.

2 NSF S. cerevisiae on on Day 0, inactivated at Day 7 by ultrasonication. W.mrakii was

then inoculated at D14. Fermentation was ceased at Day 21

3 MCF S. cerevisiae and W. mrakii were both simultaneously inoculated on Day 0 in the

ratio of 1:1000. Fermentation was ceased at Day 21

0.E+00

1.E+06

2.E+06

3.E+06

4.E+06

5.E+06

6.E+06

7.E+06

8.E+06

9.E+06

0 5 10 15 20

Pea

k A

rea

Time (Days)

Acetic acid

0.E+00

5.E+06

1.E+07

2.E+07

2.E+07

3.E+07

3.E+07

4.E+07

4.E+07

5.E+07

0 5 10 15 20

Pea

k A

rea

Time (Days)

Hexanoic acid

Page 118: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

107

5.2.3.4 Terpenes

Similar to data presented in previous chapters, most of the terpenes found in fresh

mango juice were metabolized to trace levels (Table 5.2). Of interest is the significant

increase in the concentration of para-α-dimethyl styrene (Table 5.3). This volatile compound

is found naturally in bay leave oil and imparts nutty nuances and hence may provide an

interesting flavour note to the mango wine profile. Camphene was also found in significant

amounts for NSF despite being found in negligible amounts in PSF and MCF (Table 5.2).

The difference in the amounts of terpenes detected after the three different fermentations may

prove to be useful in distinguishing between the wines, since terpene profile has been shown

to play an important role in differentiating wine varietal character. However, due to their low

levels, it is more likely their impact, if any, on the overall aroma flavor profile would be a

result of their synergistic effects.

5.2.3.5 Carbonyl compounds

Acetaldehyde was produced in significantly higher amounts in MCF relative to NSF

and PSF, which had the lowest concentration (Table 5.2). As mentioned previously,

acetaldehyde can impart a ‘fresh’ note at low levels but becomes objectionable at higher

concentrations. Hence, PSF may prove to be a fermentation strategy to modulate the amount

of acetaldehyde produced to suit the desired wine style.

Ketones make up a very small proportion of the volatiles identified (Table 5.2). 2-

Nonanone was the major ketone detected, with NSF producing the most 2-nonanone,

followed by PSF and MCF. Varying amounts of acetoin were also detected in the three

mango wines, with MCF produced the highest amount, followed by NSF and PSF. Acetoin

is responsible for the creamy, buttery note and is considered to be beneficial to wine flavour

profile at desired levels. The low level of acetoin is likely due to the lack of malolactic

fermentation in this study. However, with the low amounts of these compounds detected, it is

Page 119: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

108

unlikely that carbonyl compounds could have much potential impact on the overall wine

flavor.

5.2.4 Sensory evaluation

The sensory profile of the mango wines is presented in Figure 5.9 and scores are

compared by one-way ANOVA. In terms of ranking, NSF wine was found to be more green,

winey, sweet and yeasty; PSF wine was considered to be more fruity, waxy and creamy; and

MCF wine was deemed to be more terpenic.

Figure 5.9 Sensory profile of PSF ( ), MCF ( ) and NSF ( ), mango wine (refer to

legend of Fig 1 for definition of PSF, MCF and NSF)

For instance, NSF wine contained higher amounts of acetic acid and was deemed to be

more slightly more acidic. This is in agreement with most literature stating that acetic acid is

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

green

fruity

waxy

terpenic

rosy

winey

sweet

yeasty

acidic

creamy

Page 120: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

109

largely responsible for the volatile acidity in wine. In addition, NSF did produce significantly

more hexanoic and octanoic acids as well (Table 5.3). However, it is interesting to note that

NSF was also deemed to be sweet and winey; this could be due to the synergistic effects of

the terpenols and terpene hydrocarbons that were retained in higher amounts in NSF.

In PSF wine, the significantly larger amounts of ethyl and acetate esters could account

for its higher rating for the ‘fruity’ attribute. It was also deemed to be creamy and this could

be due to the higher amount of acetoin. However, it was considered waxy, which is generally

deemed to be a negative attribute and could be caused by the longer-chain ethyl esters. In

addition, the higher rating scored for ‘winey’ in PSF could be due to the higher ethanol

content as well.

The MCF wine was deemed to be more terpenic (least green). Although the amounts of

terpene hydrocarbons were not excessively high in the MCF mango wine, it is likely that the

overall synergistic effects of the various terpene hydrocarbons resulted in the terpenic notes.

5.3 Conclusion

Different inoculation strategies resulted in different flavour profiles of mango wines.

Although most volatile compounds produced were similar, the amounts produced by each

inoculation strategy differed. Generally, PSF produced more esters than NSF and MCF.

This could be due to the effects of ethanol toxicity that the high ethanol content had on the

Williopsis yeast, resulting in the significantly lower concentrations of volatiles produced by

NSF. However, there were also certain volatile compounds that were not detected in some of

the mango wine. This is likely due to the effects of sequential inoculation on the metabolic

process of the yeast, resulting in different biochemical reactions and metabolism.

Page 121: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

110

Chapter 6 General conclusions and recommendations

This research project provided an insight into how the various fermentation strategies

could be employed to optimise the production of mango wine from fresh mango juice. This

thesis provides an exposition of techniques and analysis of chemical composition (especially

the volatile composition) of mango wine. Different species of yeasts, S. cerevisiae and W.

mrakii, were tested for optimising the flavour profile. Several inoculation strategies were

also studied for their effects on volatile production and sensory profile.

S. cerevisiae was the major yeast in mango wine fermentation and it was not only

responsible for converting the sugars into ethanol, but also for producing a wide range of

volatiles. S. cerevisiae was efficient to produce alcohols, ethyl esters, and medium-chain fatty

acids. It also led to significant decomposition of terpenes which were the signature mango

varietal aroma compounds. W. mrakii, on the other hand, was not effective in converting

sugars into alcohols; however, it was a prolific producer of acetate esters and branched-

chained esters. Some of these compounds include ethyl acetate, isobutyl acetate, isoamyl

acetate and 2-phenylethyl acetate. In addition, W. mrakii retained more terpenes in mango

wine. However, W. mrakii should be well controlled in case of overproduction of ethyl

acetate and acetic acid. This study not only provided details on vinification techniques with

using S. cerevisiae and W. mrakii but also on chemical composition of mango wine fermented

by both yeast species. In the wine industry, screening and selection of suitable strains is

critical to make attractive mango wine. The correct strain should be confirmed before scale-

up in the industry.

Initial sugar concentration also had a significant effect on volatile and glycerol

production. High sugar concentration generally inhibits volatile production due to the toxic

Page 122: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

111

effects of ethanol on metabolic activities. However, excessive levels of acetaldehyde and

acetic acid can be produced by S. cerevisiae under hyperosmotic conditions which may have

a detrimental effect on wine flavour. On the other hand, glycerol production is directly

correlated to initial sugar concentrations. Glycerol has a positive effect on the mouthfeel of

the wine, and may impact consumers’ perception of wine quality. However, with the data

gathered from this study, the use of high sugar concentration to enhance glycerol production

may not be an appropriate technique due to the effects on volatile production. Hence, more

studies may be required to enhance glycerol production in S. cerevisiae MERIT.ferm.

Mixed culture fermentation between S. cerevisiae and W. mrakii produced a wider

range of volatile compounds and more balanced aroma. It also provided an opportunity for

winemakers to modulate volatile profile by varying inoculation strategy. In addition, an

antagonistic effect between S. cerevisiae and W. mrakii was observed by this study. The

antagonism could be attributed to ethanol and mycosin, with ethanol produced by S.

cerevisiae inhibiting W. mrakii, and mycosin secreted by W. mrakii suppressing S. cerevisiae.

These findings will be helpful to optimise the vinification techniques and inoculation strategy.

The volatile profiles of the mango wines from these different inoculation strategies were

evaluated by a trained panel of flavourists. Different inoculation strategies produced wines

with different flavour profile.

The sequence of inoculation of S .cerevisiae and W. mrakii also had a significant

impact on volatile production. The persistence of W. mrakii without the highly competitive S.

cerevisiae allowed the W. mrakii to produce and accumulate sufficient amounts of mycosin

which had an inhibitory effect on the growth and metabolic activities of S. cerevisiae in PSF.

Consequently, volatile production was affected. However, PSF proved to be a generally

better inoculation strategy than NSF. When S. cerevisiae was inoculated first, ethanol

Page 123: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

112

production was extremely rapid and the high ethanol content led to severe detrimental effects

on the subsequently inoculated W. mrakii. This could be seen by the lower amounts of

volatile production for amongst all classes of compounds. Especially prominent was the

drastic reduction in the amount of acetate esters. It was originally hypothesized that the

delayed inoculation of W. mrakii might allow the final wine to contain higher amounts of

acetate esters due to the reduced exposure of acetate ester hydrolysing S. cerevisiae.

However, this study proved that this was not the case. A better way to obtain a fruity and

floral mango wine might be by blending a W. mrakii fermented wine with a S. cerevisiae

fermented wine.

All in all, S. cerevisiae and W. mrakii are deemed to have potential in commercial

winemaking applications due to their ability to produce desirable volatile and non-volatiles

that are deemed to be beneficial to wine quality. However, further studies on how to better

regulate the metabolic pathways of these 2 yeast species, either in monocultures, mixed

cultures, or sequential fermentation is necessary to reap the most benefits that either or both

can offer.

Recommendations

The mango juice used in this study was not supplemented with additional nitrogenous

sources; therefore, the relatively low nitrogenous content may contribute to another additional

stress factor, resulting in a stuck fermenation for the high sugar fermentations. In addition, an

excessively high initial sugar content could result in excessive amounts of acetaldehyde and

acetic acid being produced, which could be detrimental to the wine quality. Studies have

shown that adding a nitrogenous source either at the beginning of the fermentation (Bely et al.

2003) or during the exponential phase (Arrizon and Gschaedler 2002) of the fermenation

increased fermentation efficiency and reduced volatile acidity. These supplementations

Page 124: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

113

includes soluble nutrients such as tryptone, yeast extracts, and a mixture of purine and

pyrimidine bases (Thomas, Hynes and Ingledew 1994). Such supplementation could

improve the fermentation efficiency of a high sugar medium for a high ethanol content fruit

wine.

Previous studies have shown that Saccharomyces has the ability to convert geraniol into

β-citronellol and other terpenols (Carrau et al. 2005), while nerol and linalool can also be

converted into α-terpineol and terpinolene (Carrau et al. 2005). In addition, production of

terpenols can occur and has previously been hypothesized to be a result of de novo synthesis

by S. cerevisiae through an alternative pathway involving metabolism of L-leucine (Carrau et

al. 2005).

Fermentation conditions have been known to affect fermentation kinetics and volatile

production (Reddy and Reddy 2011). The effects of nitrogen supplementation on volatile

production especially its effects on terpene alcohols have not been studied in detail. Further

studies in this aspect and its application in the production of mango wine could be

investigated to fully develop the ‘varietal’ character of mango wine.

While the use of mixed starter cultures in different inoculation sequences has been

studied in this research, the duration of the fermentation period for each stage can be

investigated to produce the most desirable flavour profile, since it is possible that degradation

of some volatile compounds could occur during fermentation. Furthermore, the possibility of

the aging process in improving the mango wine profile could be investigated. This would not

only optimise the volatile composition and wine flavour profile, it could also lead to process

optimisation and economic benefits for wine makers.

In addition, another aspect of mango wine that can be studied in greater detail would be

the blending of a high ethanol S. cerevisiae mango wine and a low ethanol W. mrakii wine

Page 125: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

114

with more floral and fruity attributes. A good blend between the two different styles could

potentially lead to synergistic effects, producing a wine that harvests the best of both styles.

Page 126: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

115

Bibliography

Alexandre, H., I. Rousseauz, and C. Charpentier. “Ethanol adaptation mechanisms in

Saccharomyces cerevisiae.” Biotechnology and Applied Biochemistry 20, no. 2 (1994):

173-183.

Alves, J. A., L. C. D. O. Lima, D. R. Dias, C. A. Nunes, and R. F. Schwan. "Effects of

spontaneous and inoculated fermentation on the volatile profile of lychee (Litchi

chinensis Sonn.) fermented beverages." International Journal of Food Science &

Technology 45, no. 11 (2010): 1365-2621.

Ambesi, A., M. Miranda, V.V. Petrov, and C.W. Slayman. “Biogenesis and function of the

yeast plasma membrance H(+)-ATPase.” Journal of Experimental Biology 203, no. 1

(2000): 155-160.

Ardö, Y. "Flavour formation by amino acid catabolism." Biotechnology Advances 24, no. 2

(2006): 238-242.

Arrizon, L., and A. Gschaedler. “Increasing fermentation efficiency at high sugar

concentrations by supplementing an additional source of nitrogen during the

exponential phase of the tequila fermenation process.” Canadian Journal of

Microbiology 48, no. 11 (2002): 965-970.

Baptista, J. A., J. F. D. P. Tavares, and R. C. Carvalho. " Comparison of polyphenols and

aroma in red wines from Portuguese mainland versus Azores Islands." Food Research

International, 34, no. 4 (2001): 345-355.

Bauer, F.F., and I.S. Pretorius. “Yeast stress response and fermenation efficiency: how to

survive the making of wine - a review.” South African Journal of Enology and

Viticulture 21 , Special issue (2000): 27-51.

Bely, M., A. Rinaldi, and D Dubourdieu. “Influence of assimiable nitrogen on volatile acidity

production by Saccharomyces cerevisiae during high sugar fermentation.” Journal of

Bioscience and Bioengineering 96, no. 6 (2003): 507-512.

Bely, M., P. Stoeckle, I. Masneuf-Pomarède, and D. Dubourdieu. "Impact of mixed

Torulaspora delbrueckii–Saccharomyces cerevisiae culture on high-sugar

fermentation." International Journal of Food Microbiology 122, no. 3 (2008): 312-320.

Berthels, N.J., R.R. Cordero Otero, F.F, Bauer, J.M. Thevelein, and I.S. Pretorius.

“Discrepancy in glucose and fructose utilisation during fermentatation by

Saccharomyces cerevisiae wine yeast strains.” FEMS Yeast Research 4 (2004): 683-

689.

Bisson, L. F., and J. E. Karpel. "Genetics of yeast impacting wine quality." Annual Review of

Food Science and Technology, no. 1 (2010): 139-162.

Boulton, R.B., V.L. Singleton, L.F. Bisson, and R. E. Kunkee. “Viticulture for Winemakers.”

In Principles and Practices of Winemaking, by R.B. Boulton, V.L. Singleton, L.F.

Bisson and R. E. Kunkee, 13-64. New York: Springer US, 1996.

Page 127: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

116

Buchbauer, G., and A. Ilic. “Biological activities of selected mono- and sesquiterpenes:

Possible uses in medicine.” In Natural Products: Phytochemistry, Botany and

Metabolism of Alkaloids, Phenolics and Terpenes , 4109-4159. Berlin: Springer Berlin

Heidelberg, 2013.

Carlson, M., B.C., Osmond, and D. Botstein. “Mutants of yeast defective in sucrose

utilization” Genetics 98, no. 1 (1981): 25-40.

Carrau, F, M., K. Medina, E. Boido, L. Farina, C. Gaggero, E. Dellacassa, G. Versini, and

P.A. Henschke. "De novo synthesis of monoterpenes by Saccharomyces cerevisiae

wine yeasts. "FEMS Microbiology Letters 243, no. 1 (2005): 107-115.

Cartwright, C., J.-R. Juroszek, M. Beavan, and F. Ruby. “Ethanol dissipates the proton-

motive force across the plasma membrane of Saccharomyces cerevisiae.” Journal of

General Microbiology 132 , no. 132 (1986): 369-377.

Casey, G.P., and W.M. Ingledew. “Ethanol tolerance in yeasts.” Critial Review in

Microbiology 13, no. 3 (1986): 219-280.

Chauhan, O.P., P.S. Raju, and A.S. Bawa. "Mango Flavor." In Handbook of Fruit and

Vegetable Flavors, by Y.H. Hui, edited by Y.H. Hui, F. Chen and L.M.L. Nollet, 319-

344. New Jersey: John Wiley & Sons, Inc., 2010.

Ciani, M., and F. Maccarelli. "Oenological properties of non-Saccharomyces yeasts

associated with wine-making." World Journal of Microbiology & Biotechnology 14, no.

2 (1999): 199-203.

Clemente-Jimenez, J. M., L. Mingorance-Cazorla, S. Martínez-Rodríguez, F. J. Las Heras-

Vázquez, and F. Rodríguez-Vico. "Influence of sequential yeast mixtures on wine

fermentation." International Journal of Food Microbiology 98, no. 3 (2005): 301-308.

Coloretti, F., C. Zambonelli, L. Castellari, V. Tini, and S. Rainieri. "The effect of DL-malic

acid on the metabolism of L-malic acid during wine alcoholic fermentation." Food

Technology and Biotechnology 40, no. 4 (2002): 317-320.

Devantier, R., S. Pedersen, and L. Olsson. ". Characterization of very high gravity ethanol

fermentation of corn mash. Effect of glucoamylase dosage, pre-saccharification and

yeast strain." Applied Microbiology and Biotechnology 68, no. 5 (2005): 622-629.

Dubourdieu, D., T. Tominaga, I. Masneuf, C. P. des Gachons, and M. L Murat. "The role of

yeasts in grape flavor development during fermentation: the example of Sauvignon

blanc." American Journal of Enology and Viticulture 57, no. 1 (2006): 81-88.

El Hadi, M. A. M., Zhang, Wu F. J., F. F., H., C. Zhou, and J. Tao. "Advances in fruit aroma

volatile research." Molecules 18, no. 7 (2013): 8200-8229.

Erten, H., and H. Tanguler. "Influence of Williopsis saturnus yeasts in combination with

Saccharomyces cerevisiae on wine fermentation." Letters in Applied Microbiology 50,

no. 5 (2010): 474-479.

Page 128: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

117

Farkas, G., J. M. Rezessy-Szabo, F. Zakany, and A. Hoschke. "Interaction of Saccharomyces

and non-Saccharomyces yeast strains in an alcoholic fermentation process." Acta

Alimentaria 34, no. 1 (2005): 81-90.

Fernández, M., J. F. Ubeda, and A. I. Briones. "Typing of non-Saccharomyces yeasts with

enzymatic activities of interest in wine-making." International Journal of Food

Microbiology 59, no. 1 (2000): 29-36.

Fleet, G.H., and G.M. Heard. "Yeasts: growth during fermentation." In Wine Microbiology

and Biotechnology, by G.H Fleet, 27-54. London: Harwood Academic Press GmbH,

1993.

Förster-Fromme, K., and D. Jendrossek. "Catabolism of citronellol and related acyclic

terpenoids in Pseudomonads." Applied Microbiology and Biotechnology 87, no. 3

(2010): 859-869.

Francis, I. L., and J. L. Newton. "Determining wine aroma from compositional data."

Australian Journal of Grape and Wine Research 11, no. 2 (2005): 114-126.

Fukuda, K., et al. “Balance of activities of alcohol acetyltransferase and esterase in

Saccharomyces cerevisiae is important for production of isoamyl acetate.” Applied and

Environmental Microbiology 80, no. 12 (1998): 4076-4078.

Gholap, A., and C. Bandyopadhyay. "Comparative assessment of aromatic principles of ripe

Alphonso and Langra mango." Journal of Food Science and Technology 49, no. 12

(1975): 262-263.

Godoy, H. T., and D. B. Rodriguez-Amaya. "Changes in individual carotenoids on processing

and storage of mango (Mangifera indica L.) slices and puree." International Journal of

Food Science and Technology 22, no. 5 (1987): 451-460.

Gonzalez, R., M. Quirós, and P. Morales. "Yeast respiration of sugars by non-Saccharomyces

yeast species: A promising and barely explored approach to lowering alcohol content of

wines." Trends in Food Science & Technology 29, no. 1 (2013): 55-61.

Gunata, Y. Z., C. L. Bayonove, R. L. Baumes, and R. E. Cordonnier. “Stability of free and

bound fractions of some aroma components of grapes cv. Muscat during the wine

processing: preliminary results.” American Journal of Enology and Viticulture 37, no. 2

(1986): 112-114.

Guyard, C., et al. "Involvement of β-glucans in the wide-spectrum antimicrobial activity of

Williopsis saturnus var. mrakii MUCL 41968 killer toxin." Molecular Medicine 8, no.

11 (2002): 686-694.

Hallsworth, J.E. “Ethanol induced water stress in yeast.” Journal of Fermentation

Bioengineering 85, no.2 (1998): 125-137.

Heard, GM, and GH Fleet. "Occurrence and growth of yeast species during the fermentation

of some Australian wines." Food Australia 38, no. 1 (1986): 22-25.

Page 129: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

118

Herrero, Ó., D. Ramón, and M. Orejas. "Engineering the Saccharomyces cerevisiae

isoprenoid pathway for de novo production of aromatic monoterpenes in wine."

Metabolic Engineering 10, no. 2 (2008): 78-86.

Hohmann, S. “Shaping up: the response of yeast to osmotic stress.” In Yeast Stress

Responses, by S. Hohmann and W.H. Mager, 101-145. Austin : R.G. Landes Company,

1997.

Hohmann, S. “Shaping up: The response of yeast to osmotic stress.” In Yeast Stress

Responses, by S. Hohmann and W.H. Mager, 101-145. Berlin: Springer-Verlag Berlin

Heidelberg New York, 2003.

Holm Hansen, E., P. Nissen, P. Sommer, J. C. Nielsen, and N. Arneborg. "The effect of

oxygen on the survival of non‐Saccharomyces yeasts during mixed culture

fermentations of grape juice with Saccharomyces cerevisiae." Journal of Applied

Microbiology 9, no. 13 (2001): 541-547.

Holzapfel, W. H. "Appropriate starter culture technologies for small-scale fermentation in

developing countries." . International Journal of Food Microbiology 75, no. 3 (2002):

197-212.

Inoue, S.B., N. Takewakt, T. Takasuka, T. Mio, M. Adachi, Y. Fujii, C. Miyamoto, M.

Arisawa, Y. Furuichi, and T. Watanabe. "Characterization and gene cloning of 1,

3‐β‐D‐glucan synthase from Saccharomyces cerevisiae." European Journal of

Biochemistry 231, no. 3 (1995): 845-854.

Iwase, T., T. Morikawa, H. Fukuda, K. Sasaki, and M. Yoshitake. "Production of fruity odor

by genus Williopsis." Journal of the Brewing Society of Japan 90, no. 5 (1995): 394-

396.

Jackson, D. I., and P. B. Lombard. "Environmental and management practices affecting grape

composition and wine quality-a review." American Journal of Enology and Viticulture

44, no. 4 (1993): 409-430.

Jackson, R. S. Wine Science: Principles and Applications. Oxford: Academic Press, 2008.

Jain, V. K., B. Divol, B. A. Prior, and F. F. Bauer. "Elimination of glycerol and replacement

with alternative products in ethanol fermentation by Saccharomyces cerevisiae."

Journal of Industrial Microbiology & Biotechnology 38, no. 9 (2011): 1427-1435.

Jolly, N.P., O.P.H. Augustyn, and I.S. Pretorius. "The role and use of non-Saccharomyces

yeasts in wine production." South African Journal of Viticulure 27, no. 1 (2006): 15-39.

Jones, P. R., R. Gawel, I. L. Francis, and E. J. Waters. "The influence of interactions between

major white wine components on the aroma, flavour and texture of model white wine."

Food Quality and Preference 19, no. 6 (2008): 596-607.

Jungalwala, F., and H. Cama. "Carotenoids in mango (Mangifera indica L.) fruit." Indian

Journal of Chemistry 62, no. 1 (1963): 36-40.

Kapsopoulou, K., A. Mourtzini, M. Anthoulas, and E. Nerantzis. "Biological acidification

during grape must fermentation using mixed cultures of Kluyveromyces thermotolerans

Page 130: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

119

and Saccharomyces cerevisiae." World Journal of Microbiology and Biotechnology 23,

no. 5 (2007): 735-739.

Kimura, T., T. Komiyama, Y. Furuichi, Y. Iimura, S. Karita, K. Sakka, and K. Ohmiya.. "N-

glycosylation is involved in the sensitivity of Saccharomyces cerevisiae to HM-1 killer

toxin secreted from Hansenula mrakii IFO 0895." Applied microbiology and

biotechnology 51, no. 2 (1999): 176-184.

King, A., and J.R Dickinson. ""Biotransformation of monoterpene alcohols by

Saccharomyces cerevisiae, Torulaspora delbrueckii and Kluyveromyces lactis." Yeast

16, no. 6 (2000): 499-506.

Kulkarni, J. H., H. Singh, and K.L. Chadha. "Preliminary screening of mango varieties for

wine making." Journal of Food Science and Technology 17, no. 5 (1980): 218-220.

Kurita, O. "Increase of acetate ester‐hydrolysing esterase activity in mixed cultures of

Saccharomyces cerevisiae and Pichia anomala." Journal of Applied Microbiology 104,

no. 4 (2008): 1051-1058.

Kyung, M.Y., C.-L. Rosenfield, and D.C. Knipple. “Ethanol tolerance in the yeast

Saccharomyces cerevisiae is dependent on cellular oleic acid content.” Applied and

Environmental Microbiology 69, no. 3 (2003): 1499-1503.

Lagace, LS, and LF Bisson. "Survey of yeasts acid proteases for effectiveness of wine haze

reduction." American Journal of Viticulture and Enology 41, no. 2 (1990): 147–155.

Lalel, H. J., Z. Singh, and S. C. Tan. "Glycosidically-bound aroma volatile compounds in the

skin and pulp of 'Kensington Pride' mango fruit at different stages of maturity."

Postharvest Biology and Technology 29, no. 2 (2003a): 205-218.

Lalel, H., Z. Singh, and S. Tan. "Maturity stage at harvest affects fruit ripening, quality and

biosynthesis of aroma volatile compounds in 'Kensington Pride' mango." Journal of

Horticultural Science and Biotechnology 78, no. 2 (2003b): 225-233.

Lambrechts, M. G., and I. S. Pretorius. "Yeast and its importance to wine aroma-a review."

South African Journal of Enology and Viticulture, 21., no. 1 (2000): 97-129.

Laokhakunjit, N., A. Uthairatakji, O. Kerdchoechuen, A. Chatpaisarn, and S. Photchanachai.

Identification of changes in volatile compound in gamma-irradiated mango during

storage. Technical report, Bangkok: International Symposium "New Frontier of

Irradiated Food and Non-food Products", 2005.

Léchaudel, M., and J. Joas. "An overview of preharvest factors influencing mango fruit

growth, quality and postharvest behaviour." Brazilian Journal of Plant Physiology 19,

no. 4 (2007): 287-298.

Lee, P. R., Y. L. Ong, B. Yu, P. Curran, and S. Q. Liu. "Evolution of volatile compounds in

papaya wine fermented with three Williopsis saturnus yeasts." International Journal of

Food science & Technology 45, no. 10 (2010a): 2032-2041.

Page 131: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

120

Lee, P.-R., I. S.-M. Chong, B. Yu, P. Curran, and S.-Q. Liu. "Effects of sequentially

inoculated Williopsis saturnus and Saccharomyces cerevisiae on volatile profiles of

papaya wine." Food Research International 45, no. 1 (2012b): 177-183.

Lee, P.-R., S. H. C. Kho, B. Yu, Curran, P., and S.-Q Liu. "Yeast ratio is a critical factor for

sequential fermentation of papaya wine by Williopsis saturnus and Saccharomyces

cerevisiae." Microbial Biotechnology 6, no. 4 (2012c): 385-393.

Li, X. Impact of Selected Yeasts and Glycosidases on Chemical Composition of Mango Wine.

PhD Thesis, Singapore: National University of Singapore, 2013.

Li, X., B. Yu, P. Curran, and S. Q Liu. "Chemical and volatile composition of mango wines

fermented with different Saccharomyces cerevisiae yeast strains." South African

Journal of Enology & Viticulture 32, no. 1 (2011): 117-128.

Li, X., B. Yu, P. Curran, and S.-Q. Liu. "Impact of two Williopsis yeast strains on the volatile

composition of mango wine." International Journal of Food Science and Technology

47, no. 4 (2012): 808-815.

Li, X., L.J. Chan, B. Yu, P. Curran, and S.-Q. Liu. "Fermentation of three varieties of mango

juices with a mixture of Saccharomyces cerevisiae and Williopsis saturnus var. mrakii."

International Journal of Food Microbiology 158, no. 1 (2012): 28-35.

Li, X., Lim, S.L., B. Yu, P. Curran, and S.-Q. Liu. "Mango wine aroma enhancement by pulp

contact and β-glucosidase." International Journal of Food Science and Technology 48,

no. 11 (2013a): 2258-2266.

Li, X., S.L. Lim, B. Yu, P. Curran, and S.-Q. Liu. "Impact of pulp on the chemical profile of

mango wine." South African Journal of Enology and Viticulture 34, no. 1 (2013b): 119-

128.

Lilly, M., F. F. Bauer, M. G. Lambrechts, J. H. Swiegers, D. Cozzolino, and I. S. Pretorius.

"The effect of increased yeast alcohol acetyltransferase and esterase activity on the

flavour profiles of wine and distillates." Yeast 23, no. 9 (2006): 641-659.

Liu, S. Q., and M. Tsao. "Biocontrol of spoilage yeasts and moulds by Williopsis saturnus

var. saturnus in yoghurt." Nutrition & Food Science 40, no. 2 (2010): 166-175.

Magliani, W., S. Conti, L. R. Travassos, and L. Polonelli. "From yeast killer toxins to

antibiobodies and beyond." FEMS Microbiology Letters 288, no. 1 (2008): 1-8.

Marais, J. "Terpenes in the aroma of grapes and wines: a review." South African Journal of

Enology and Viticulture 4, no. 2 (1983): 49-60.

Maturano, Y. P., et al. "Multi-enzyme production by pure and mixed cultures of

Saccharomyces and non-Saccharomyces yeasts during wine fermentation."

International Journal of Food Microbiology 155, no. 1 (2012): 43-50.

Mendoza, L. M., M. C. M. de Nadra, E. Bru, and M. E. Farías. "Influence of wine-related

physicochemical factors on the growth and metabolism of non-Saccharomyces and

Saccharomyces yeasts in mixed culture." Journal of Industrial Microbiology and

Biotechnology 36, no. 2 (2009): 229-237.

Page 132: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

121

Minetoki, T. "Alcohol acetyl transferase of sake yeast." Journal of the Brewing Society of

Japan 87, no. 5 (1992): 334-340.

Miyake, T., and T. Shibamoto. "Quantitative analysis of acetaldehyde in foods and

beverages." Journal of Agricultural and Food Chemistry 41, no. 11 (1993): 1968-1970.

Moreno, J. J., C. Millán, J. M. Ortega, and M. Medina. "Analytical differentiation of wine

fermentations using pure and mixed yeast cultures." Journal of Industrial Microbiology

and Biotechnology 7, no. 3 (1991): 181-189.

Myers, D.K., D.T.M. Lawlor, and P.V. Attfield. "Influence of invertase acitivity and glycerol

synthesis and retention on fermentation of media with a high sugar concnetration by

Saccharomyces cerevisiae." Applied and Environmental Microbiology 63, no. 1 (1997):

145-150.

Nieuwoudt, H. H., B. A. Prior, I. S. Pretorius, and F. F Bauer. "Glycerol in South African

table wines: an assessment of its relationship to wine quality." South African Journal

for Enology and Viticulture 23, no. 1 (2002): 22-30.

Nissen, P., and N. Arneborg. "Characterization of early deaths of non-Saccharomyces yeasts

in mixed cultures with Saccharomyces cerevisiae." Archives of Microbiology 180, no. 4

(2003): 257-263.

Nissen, P., D. Nielsen, and N Arneborg. "Viable Saccharomyces cerevisiae cells at high

concentrations cause early growth arrest of non‐Saccharomyces yeasts in mixed

cultures by a cell–cell contact‐mediated mechanism." Yeast 20, no. 4 (2003): 331-341.

Obisanya, M. O., J.O. Aina, and G.B Oguntimein. "Production of wine from mango

(Magnifera indica L.) using Saccharomyces and Schizosaccharomyces species isolated

from palm wine." Journal of Applied Microbiology 63, no. 3 (1987): 191-196.

Ong, P. K., and T. E. Acree. "Similarities in the aroma chemistry of Gewürztraminer variety

wines and lychee (Litchi chinesis Sonn.) fruit." Journal of Agricultural and Food

Chemistry 47, no. 2 (1999): 665-670.

Onkarayya, H, and H, Singh. "Screening of mango varieties for Dessert and Madeira-style

wines." American Journal of Enology and Viticulture 35, no. 2 (1984): 63-65.

Onkarayya, H. "A rapid madeirization process to improve mango dessert wines." Journal of

Food Science and Technology 23, no. 3 (1986): 175-176.

Perestrelo, R., A. Fernandes, F. F. Albuquerque, J. C. Marques, and J. S. Camara. "Analytical

characterization of the aroma of Tinta Negra Mole red wine: Identification of the main

odorants compounds." Analytica Chimica Acta 563, no. 1 (2006): 154-164.

Pietruszka, M., K Pielech-Przybylska, and J. S Szopa. "Synthesis of higher alcohols during

alcoholic fermentation of rye mashes." Zeszyty Naukowe. Chemia Spożywcza i

Biotechnologia/Politechnika Łódzka 74, no. 108 (2010): 51-64.

Pino, J. A., and J. Mesa. "Contribution of volatile compounds to mango (Mangifera indica

L.) aroma." Flavour and Fragrance Journal 21, no. 2 (2006): 207-213.

Page 133: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

122

Pino, J. A., and O Queris. "Analysis of volatile compounds of mango wine." Food Chemistry

125, no. 4 (2011): 1141-1146.

Pino, J. A., J. Mesa, Y. Muñoz, M. P. Martí, and R. Marbot. "Volatile components from

mango (Mangifera indica L.) cultivars." Journal of Agricultural and Food Chemistry

53, no. 6 (2005): 2213-2223.

Plata, C., C. Millan, J. C. Mauricio, and J. M. Ortega. "Formation of ethyl acetate and

isoamyl acetate by various species of wine yeasts." Food Microbiology 20, no. 2

(2003): 217-224.

Pretorius, I. S. "Tailoring wine yeast for the new millennium: novel approaches to the ancient

art of winemaking." Yeast 16, no. 8 (2007): 675-729.

Pretorius, I., T. Van der Westhuizen, and O. Augustyn. "Yeast biodiversity in vineyards and

wineries and its importance to the South African wine industry: A review." South

African Journal of Enology and Viticulture 20, no. 2 (1999): 61-70.

Quijano, C. E., G. Salamanca, and J. A. Pino. "Aroma volatile constituents of Colombian

varieties of mango (Mangifera indica L.)." Flavour and Fragrance Journal 22, no. 5

(2007): 401-406.

Radle, F., and H. Schülz. "Glycerol production of various strains of Saccharomyces."

American Journal of Enology and Viticulture no. 1 (1982): 36–40.

Rapp, A. "Volatile flavour of wine: correlation between instrumental analysis and sensory

perception." Food/Nahrung 42, no. 6 (1998): 351-363.

Rapp, A., and H. Mandery. "Wine aroma." Experientia 42, no. 8 (1986): 873-884.

Reddy, L. V. A., and O. V. S. Reddy. ""Production and characterization of wine from mango

fruit (Mangifera indica L)."." World Journal of Microbiology and Biotechnology 21,

no. 8-9 (2005): 1345-1350.

Reddy, L. V. A., and O. V. S. Reddy. "Effect of fermentation conditions on yeast growth and

volatile composition of wine produced from mango Mangifera indica L.) fruit juice."

Food and Bioproducts Processing 89, no. 4 (2011): 487-49.

Reddy, L.V., and O Reddy. "Production of ethanol from mango (Mangifera indica L.) fruit

juice fermentation." Research Journal of Microbiology 2, no. 10 (2007): 763-769.

Reddy, L.V., O.V.S. Reddy, and V.K. Joshi. "Production, optimization and characterization

of wine from mango (Mangifera indica Linn.)." Natural Product Radiance 8, no. 4

(2009): 426-435.

Reddy, L.V.A., and O. V. S. Reddy. "Effect of enzymatic maceration on synthesis of higher

alcohols during mango wine fermentation." Journal of Food Quality 32, no. 1 (2009):

34-47.

Remize, F., J.L. Roustan, J.M. Sablayrolles, P. Barre, and S. Dequin. "Glycerol

overproduction by engineered Saccharomyces cerevisiae wine yeast strains leads to

Page 134: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

123

substantial changes in by-product formation and to a stimulation of fermentation rate in

stationary phase." Applied and Environmental Microbiology 65, no. 1 (1999): 143-149.

Rojas, V., J. V. Gil, F. Piñaga, and P. Manzanares. "Studies on acetate ester production by

non-Saccharomyce wine yeasts." International Journal of Food Microbiology 70, no. 3

(2001): 283-289.

Romano, P., G. Suzzi, L. Turbanti, and M. Polsinelli. "Acetaldehyde production in

Saccharomyces cerevisiae wine yeasts." FEMS Microbiology Letters, 118, no. 3

(1994): 213-218.

Rosi, I, M Vinella, and P Domizio. "Characterization of β-glucosidase activity in yeast of

oenological origin." Journal of Applied Bacteriology 77, no. 5 (1994): 519-527.

Sadoudi, M., R. Tourdot-Maréchal, S. Rousseaux, D. Steyer, J-J Gallardo-Chacón, J.

Ballester, S. Vichi, R. Guérin-Schneider, J. Caixach, and H. Alexandre. "Yeast–yeast

interactions revealed by aromatic profile analysis of Sauvignon Blanc wine fermented

by single or co-culture of non-Saccharomyces and Saccharomyces yeasts."Food

microbiology 32, no. 2 (2012): 243-253.

Sáenz-Navajas, M. P., E. Campo, P. Fernández-Zurbano, D. Valentin, and V. Ferreira. "An

assessment of the effects of wine volatiles on the perception of taste and astringency in

wine." Food Chemistry 121, no. 4 (2010): 1139-1149.

Saerens, S.M.G., F. Delvaus, K.J, Verstrepen, P. Van Dijck, M. Thevelein, and F.R. Delvaux.

“Parameters affecting ethyl ester production by Saccharomyces cerevisiae during

fermentation.” Applied and Environmental Microbiology 74, no. 2 (2008): 454-461.

Sagar, V., D. Khurdyia, and K. Balakrishnan. "Quality of dehydrated ripe mango slices as

affected by packaging material and mode of packaging." Journal of Food Science and

Technology 36, no. 1 (1999): 67-70.

Sales, K., W. Brandt, E. Rumbak, and G. Lindsey. “The LEA-like protein HSP12 in

Saccharomyces cerevisiae has a plasma membrane location and protects membrane

against desiccation and ethanol-induced stress.” Biochimica et Biophysica Acta (BBA) -

Biomembranes 1463, no. 2 (2000): 267-278.

Salmon, J.-M., O. Vincent, J. Mauricio, C. Bely, and P. Barre. “Sugar transport inhibition and

apparent loss of activity in Saccharomyces cerevisiae as a major limiting factor of

enological fermentations.” American Journal of Enology and Viticulture 44, no. 1

(1993): 56-64.

Sánchez-Palomo, E., M. C. Diaz-Maroto, and M. S. Pérez-Coello. "Rapid determination of

volatile compounds in grapes by HS-SPME coupled with GC–MS." Talanta 66, no. 5

(2005): 1152-1157.

Scanes, K.T., S. Hohmann, and B.A. Prior. “Glycerol production by the yeast Saccharomyces

cerevisiae and its relevance to wine: a review.” South African Journal of Enology and

Viticulture 19 no. 1 (1998): 17-24.

Page 135: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

124

Seow, Y. X., P. K. Ong, and S. Q. Liu. " Production of flavour-active methionol from

methionine metabolism by yeasts in coconut cream." International Journal of Food

Microbiology 143, no. 3 (2010): 235-240.

Slaninova, I., S. Sestak, A. Svoboda, and V. Farkas. “Cell wall and cytoskeleton

reorganization as a response to hyperosmotic shock in Saccharomyces cerevisiae.”

Archives of Microbiology 173, no. 4 (2000): 245-252.

Soden, A., I. L. Francis, H. Oakey, and P. A Henschke. "Effects of co‐fermentation with

Candida stellata and Saccharomyces cerevisiae on the aroma and composition of

Chardonnay wine." Australian Journal of Grape and Wine Research 6, no.1 (2000): 21-

30.

Spencer, J., M. Morris, and W Kennard. "Vitamin C concentration in developing and mature

fruits of mango (Mangifera indica L.)." Plant Physiology 31, no. 1 (1956): 79-80.

Styger, G., D. Jacobson, B. A. Prior, and F. F. Bauer. "Genetic analysis of the metabolic

pathways responsible for aroma metabolite production by Saccharomyces cerevisiae."

Applied Microbiology and Biotechnology 97, no. 10 (2013): 4429-4442.

Sudheer, K., Yannam, S.P. Reddy, and O.V.S Reddy. "Optimisation of fermentation

conditions for mango (Mangifera indica L.) wine production by employing response

surface methodology." International Journal of Food Science & Technology 44, no. 11

(2009): 2320-2327.

Sumby, K. M., P. R. Grbin, and V. Jiranek. " Microbial modulation of aromatic esters in

wine: current knowledge and future prospects." Food Chemistry 121, no. 1 (2010): 1-

16.

Swiegers, J. H., E. J. Bartowsky, P. A. Henschke, and I. S. Pretorius. " Yeast and bacterial

modulation of wine aroma and flavour." Australian Journal of Grape and Wine

Research 11, no. 2 (2005): 139-173.

Tan, A. W., P. R. Lee, Y. X. Seow, P. K. Ong, and S. Q. Liu. "Volatile sulphur compounds

and pathways of L-methionine catabolism in Williopsis yeasts." Applied Microbiology

and Biotechnology 95, no. 4 (2012): 1011-1020.

Tandon, D. K., and S. K. Kalra. "Pectin changes during the develoment of mango fruit cv.

Dashehari." Journal of Horticultural Science 59 (1984): 283-286.

Tehlivets, O., K. Scheuringer, and S. D. Kohlwein. "Fatty acid synthesis and elongation in

yeast." Biochimica et Biophysica Acta (BBA)-Molecular and Cell Biology of Lipids

1771, no. 3 (2007): 255-270.

Tharanathan, R., H. Yashode, and T. Prabha. "Mango (Mangifera inica L), "The king of

fruits"- An overview." Food Reviews International 22, no. 2 (2006): 95-123.

This, P., T. Lacombe, and M. R. Thomas. "Historical origins and genetic diversity of wine

grapes." TRENDS in Genetics 22, no. 9 (2006): 511-519.

Thomas, K. C., S. H. Hynes, and W. M. Ingledew. "Effects of particulate materials and

osmoprotectants on very-high-gravity ethanolic fermentation by Saccharomyces

Page 136: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

125

cerevisiae." Applied and Environmental Microbiology 60(5), 60, no. 5 (1994): 1519-

1524.

Thomas, KC, S H Hynes, and W M Ingledew. “Effects of particulate materials and

osmoprotectants on very high gravity ethanolic fermentation by Saccharomyces

cerevisiae.” Applied and Environmental Microbiology 60, no. 5 (1994): 1519-1524.

Trinh, T. T. T., B. Yu, P. Curran, and S. Q. Liu. "Dynamics of volatile compounds during

longan juice fermentation by three yeasts from the genus Williopsis." Acta Alimentaria

40, no. 3 (2011): 385-395.

Ugliano, M., E. J. Bartowsky, J. McCarthy, L. Moio, and P. A. Henschke. " Hydrolysis and

transformation of grape glycosidically bound volatile compounds during fermentation

with three Saccharomyces yeast strains." Journal of Agricultural and Food Chemistry

54, no. 17 (2006): 6322-6331.

van Zyl, P. J., S. G. Kilian, and B. A. Prior. "The role of an active transport mechanism in

glycerol accumulation during osmoregulation by Zygosaccharomyces rouxii." Applied

Microbiology and Biotechnology 34, no. 2 (1990): 231-235.

Varakumar, S., K. Naresh, P. S. Variyar, A. Sharma, and O. V. S. Reddy. "Role of malolactic

fermentation on the quality of mango (Mangifera indica L.) Wine." Food

Biotechnology 27, no. 2 (2013): 119-136.

Varakumar, S., N. Kondapalli, and O.V.S. Reddy. "Effect of co-fermentation with

Saccharomyces cerevisiae and Torulaspora delbrueckii or Metschnikowia pulcherrima

on the aroma and sensory properties of mango wine." Annals of Microbiology 62, no. 4

(2012): 1353-1360.

Varakumar, S., Y.S. Kumar, and O.V.S. Reddy. "Carotenoid composition of mango

(Mangifera indica L.) wine and its antioxidant activity." Journal of Food Biochemistry

35, no. 5 (2011): 1538-1547.

Vásquez-Caicedo, A. L., S. Neidhart, P. Pathomrungsiyounggul, P. Wiriyacharee, A.

Chattrakul, P. Sruamsiri, P. Manochai, F. Bangerth, and R. Carle. "Physical, Chemical,

and Sensory Properties of Nine Thai Mango Cultivars and Evaluation of their

Technological and Nutritional Potential." In International Symposium: Sustaining Food

Security and Managing Natural Resources in Southeast Asia Challenges for the 21st

Century, pp. 8-11. 2002

Viana, F., J. V. Gil, S. Vallés, and P. Manzanares. " Increasing the levels of 2-phenylethyl

acetate in wine through the use of a mixed culture of Hanseniaspora osmophila and

Saccharomyces cerevisiae." International Journal of Food Microbiology 135, no. 1

(2009): 68-74.

Walker, G. M. Yeast Physiology and Biotechnology. West Sussex: John Wiley & Sons., 1998.

Wang, D., Y. Su, J. Hu, and G. Zhao. “Fermentation kinetics of different sugars by apple

wine yeast Saccharomyces cerevisiae.” Journal of the Institute of Brewing 110, no. 4

(2004): 340-346.

Page 137: EFFECTS OF SUGAR CONCENTRATION AND YEAST …

126

Whiting, G. C. "Organic acid metabolism of yeasts during fermentation of alcoholic

beverages—a review." Journal of the Institute of Brewing 82, no. 2 (1976): 84-92.

Yamamoto, T., M. Imai, K. Tachibana, and M. Mayumi. "Application of monoclonal

antibodies to the isolation and characterization of a killer toxin secreted byHansenula

mrakii." FEBS Letters 195, no. 1 (1986): 253-257.

Yilmaztekin, M., H. Erten, and T. Cabaroglu. "Enhanced production of isoamyl acetate from

beet molasses with addition of fusel oil by Williopsis saturnus var. saturnus." Food

Chemistry 112, no. 2 (2009): 290-294.

Zoecklein, B. W., J. E. Marcy, J. M. Williams, and Y. Jasinski. "Effect of native yeasts and

selected strains of Saccharomyces cerevisiae on glycosyl glucose, potential volatile

terpenes, and selected aglycones of White Riesling Vitis vinifera L.) wines." Journal of

Food Composition and Analysis 10, no. 1 (1997): 55-65.