You are on page 1of 175

AN ABSTRACT OF THE THESIS OF

Ann Dumont for the degree of Master of Science in Food Science and
Technology presented on November 23. 1994.

Title: Sensory and Chemical Evaluation of Riesling. Chardonnay and Pinot

Noir Fermented by Different Strains of Saccharomyces cerevisiae .

Abstract approved: .
Barney T. Watson

Fermentation of the grape must into wine is one of the most important steps

in winemaking. Selected yeast strains of Saccharomyces cerevisiae have

been used for many years to ensure complete and even fermentations. The

formation of volatile compounds also occurs during fermentation and will

influence the sensory perception of the wine. The main objective of the

research was to study the sensory and chemical composition of 1992 Oregon

Riesling, Chardonnay and Pinot Noir wines fermented with different

commercial S.cerevisiae strains.

In the first study, Free-Choice Profiling was used to study the sensory

profiles of all three varieties after 7 and 20 months of aging. This method was

used in order to utilize a panel of expert winemakers for the tastings. The

sensory data analyzed through Generalized Procrustes Analysis showed that

some strains were similar while others were different in terms of aroma and
flavor at 7 months of age. After 20 months of aging, differences and

similarities were still present although the sensory profiles were different from
the young wines. This last finding showed that differences are still present
after a period of aging.
In the second study, the chemical composition of all three varieties and the
volatile composition of selected wines of Riesling and Chardonnay were

studied. In both white varieties, statistical differences in titratable acidity,

residual sugar, volatile acidity and malate content resulted from fermentation
with different yeast strains. The volatile composition was qualitatively similar,

but some quantitative differences, as relative concentration, were found.

Whether or not those differences had a sensory impact was not investigated.
Results of the present study showed the need for further studies in order to

understand the role of yeast in flavor development. Relationships between

sensory profiles and volatile composition could help winemakers to

understand the influence of a selected strain on a particular variety and to

select yeast strains to optimize wine quality.


©Copyright by Ann Dumont
November 23, 1994
All rights reserved
Sensory and Chemical Evaluation of Riesling,

Chardonnay and Pinot Noir Fermented by


Different Strains of Saccharomyces cerevisiae

by

Ann Dumont

A THESIS

submitted to
Oregon State University

in partial fulfillment of

the requirements for

the degree of

Master of Science

Completed November 23, 1994

Commencement June 1995


Master of Science thesis of Ann Dumont presented on November 23. 1994.

APPROVED:

Senior Instructor of Food Science and Technology in charge of major

■<J
QAji\ Department oJ^Food Science anoTe*

Dean of t^e^raduate
tne Graduate Scaool
Sbttc

I understand that my thesis will become part of the permanent collection of


Oregon State University libraries. My signature below authorizes release of
my thesis to any reader upon request.

Ann Dumont, Author


ACKNOWLEDGMENTS

WINE IS A SIMPLE PROOF THAT GOD LOVES US, AND A CONSTANT

REMINDER THAT HE INTENDS FOR US TO BE HAPPY.. Benjamin Franklin


I would like to thank Barney Watson and Mina McDaniel for their support,

encouragement and expertise throughout this work. I would also like to thank

Dr. D. Selivonchick, Dr. M. Deinzer and Dr. C. Pereira for accepting to be on


my committee and giving me a strong background.

I would also like to thank Lallemand Inc. for allowing me to come to OSU to

do my Masters degree. I would also like to thank Thomas Henick-Kling and

Terry Acree at Cornell University for their valuable advice.


This work was supported by other scientists that I would like to thank. Dr. L.
Libbey for valuable information on GC/MS, B. Yorgey for the help with the

grape crushing, S. Price for teaching me about grapes, Hsiao Ping Chen for

her help in the lab, D. Griffin for the GC-MS help, J. Fleurent for the DNA

fingerprints, S. Rubico and C. Lederer, whose friendship and advice in

sensory science were equally important. Very importantly, all of you

wonderful ladies in the office who knew everything, and were always so nice

to me (Tracy, Rosanna, Anne, Cheryl, Reitha). This work would also not have
been possible without the time and expertise of several Oregon winemakers

who came to taste the wines.

I would like to thank mostly my "support staff", a bit far but always with me:

J.-C. Frigon who believed so strongly, M. Bombardier, S. Comer, B. Sander.


You were all the things friends should be and there is a little bit of you in this.
My friends in Oregon: JoLynne, Anne, Paulo, Alice, Sonia, Cindy B., Scott,
Bob, Dan, Dave, Wolfgang and Joanna, Ean-Tun (Alex), Sudarshan and Pam.

Lastly, I would like to thank my family: Lorraine, Nathalie, Jean-Pierre,


Richard, Nicolas, Marc, Lucie and S. and T. Dumont. Their thoughts were
always with me. I also have to thank two important people in my life; my

grandfather who is no longer with me, and Renee Lapointe who is an example
of pride and courage.
I would like to thank Lallemand Inc. and The Oregon Wine Advisory Board

for the financial support of this work.


CONTRIBUTION OF AUTHORS

Dr. Mina R. McDaniel was involved in the design, analysis and writing of the
sensory evaluation. Dr. Max Deinzer was involved in the data collection for

the chemical analysis for the study.


TABLE OF CONTENTS

Page

CHAPTER 1. INTRODUCTION 1

CHAPTER 2. YEASTS AND WINES: AN OVERVIEW 4

Flavor Compounds Produced By Yeast Fermentation 5

Alcohols 6

Aldehydes 7

Ketones 8

Acids 9

Esters 9

Terpenes 10

Amines 12

Sulfur Compounds 13

Lactones 14

Phenols 14

Volatile Compounds Analysis 15

Wine Extraction 15

Gas Chromatography 16

Gas Chromatography-Mass Spectrometry 17

Sensory Evaluation 18

Statistical Analysis 20

Literature Cited 20
TABLE OF CONTENTS (Continued)

Page

CHAPTER 3. SENSORY EVALUATION OF RIESLING,


CHARDONNAY AND PINOT NOIR FERMENTED WITH
DIFFERENT COMMERCIAL SACCHAROMYCES CEREVISIAE
STRAINS BY FREE-CHOICE PROFILING AT SEVEN AND
TWENTY MONTHS OF AGE 27

Abstract 28

Introduction 29

Materials And Methods 30

Wine Preparation 30

Sensory Evaluation 32

Free-Choice Profiling 32
Statistical Analysis 34

Results And Discussion 35

Riesling Sensory Evaluation 35

Chardonnay Sensory Evaluation 55

Pinot Noir Sensory Evaluation 74

Conclusion 94

Literature Cited 94

CHAPTER 4 - CHEMICAL COMPOSITION OF RIESLING,


CHARDONNAY AND PINOT NOIR AND VOLATILE
COMPOSITION OF RIESLING AND CHARDONNAY
FERMENTED WITH DIFFERENT COMMERCIAL STRAINS
OF SACCHAROMYCES CEREVISIAE. 97

Abstract 98
TABLE OF CONTENTS (Continued)

Page

Introduction 99

Materials And Methods 101

Wine Preparation 101

Wine Analysis 103

Extraction Of Volatile Compounds 104

Gas Chromatography-Mass Spectrometry 106

Results And Discussion 107

Fermentation Rates 107

Composition Analysis 109

GC/MS Analysis Of Riesling And Chardonnay 115

Conclusion 138

Literature Cited 139

CHAPTER 5. SUMMARY 145

BIBLIOGRAPHY 146
LIST OF FIGURES

Figure Description Page

2.1 Sugar and amino acids metabolism during


grape fermentation by yeasts 6

3.1 Sample consensus plot for Free-Choice


Profiling of the aroma of Riesling (7 months)
fermented by different yeast strains 37

3.2 Sample consensus plot for Free-Choice


Profiling of the aroma of Riesling (20 months)
fermented by different yeast strains 41

3.3 Sample consensus plot for Free-Choice


Profiling of the flavor of Riesling (7 months)
fermented by different yeast strains 45

3.4 Sample consensus plot for Free-Choice


Profiling of the flavor of Riesling (20 months)
fermented by different yeast strains 48

3.5 Percentages of consensus and within (residual)


variation distributed over the Riesling
samples for the aroma at 7 months of age 51

3.6 Percentages of consensus and within (residual)


variation distributed over the Riesling
samples for the aroma at 20 months of age 52

3.7 Percentages of consensus and within (residual)


variation distributed over the Riesling
samples for the flavor at 7 months of age 53
LIST OF FIGURES (Continued)

Page

3.8 Percentages of consensus and within (residual)


variation distributed over the Riesling
samples for the flavor at 20 months of age 54

3.9 Sample consensus plot for Free-Choice


Profiling of the aroma of Chardonnay '
(7 months) fermented by different yeast
strains 57

3.10 Sample consensus plot for Free-Choice


Profiling of the aroma of Chardonnay
(20 months) fermented by different yeast
strains 60

3.11 Sample consensus plot for Free-Choice


Profiling of the flavor of Chardonnay
(7 months) fermented by different yeast
strains 65

3.12 Sample consensus plot for Free-Choice


Profiling of the flavor of Chardonnay
(20 months) fermented by different yeast
strains 67

3.13 Percentages of consensus and within (residual)


variation distributed over the Chardonnay
samples for the aroma at 7 months of age 70

3.14 Percentages of consensus and within (residual)


variation distributed over the Chardonnay
samples for the aroma at 20 months of age 71
LIST OF FIGURES (Continued)

Page

3.15 Percentages of consensus and within (residual)


variation distributed over the Chardonnay
samples for the flavor at 7 months of age 72

3.16 Percentages of consensus and within (residual)


variation distributed over the Chardonnay
samples for the flavor at 20 months of age 73

3.17 Sample consensus plot for Free-Choice


Profiling of the aroma of Pinot Noir
(7 months) fermented with different yeast
strains 76

3.18 Sample consensus plot for Free-Choice


Profiling of the aroma of Pinot Noir
(20 months) fermented with different yeast
strains 79

3.19 Sample consensus plot for Free-Choice


Profiling of the flavor of Pinot Noir
(7 months) fermented with different yeast
strains 84

3.20 Sample consensus plot for Free-Choice


Profiling of the flavor of Pinot Noir
(20 months) fermented with different yeast
strains 86

3.21 Percentages of consensus and within (residual)


variation distributed over the Pinot Noir
samples for the aroma at 7 months of age 90
LIST OF FIGURES (Continued)

Page
3.22 Percentages of consensus and within (residual)
variation distributed over the Pinot Noir
samples for the aroma at 20 months of age 91

3.23 Percentages of consensus and within (residual)


variation distributed over the Pinot Noir
samples for the flavor at 7 months of age 92

3.24 Percentages of consensus and within (residual)


variation distributed over the Pinot Noir
samples forthe flavor at 20 months of age 93

4.1 Fermentation rate of Riesling with different


yeast strains 108

4.2 Fermentation rate of Chardonnay with


different yeast strains 110

4.3 Fermentation rate of Pinot Noir with different


yeast strains 111

4.4 GC/MS Chromatogram of the extract of Riesling


fermented by Zymaflore VL-1 118

4.5 GC/MS chromatogram of the extract of Riesling


fermented by Epernay 2 CEG 119

4.6 GC/MS chromatogram of the extract of


Chardonnay fermented by EC-1118 Prise de
Mousse 120

4.7 GC/MS chromatogram of the extract of


Chardonnay fermented by D47 121
LIST OF TABLES

Table Description Page

3.1 Aroma attributes with positive or negative


loadings > 0.30 for Riesling at 7 months.by
each panelist 36

3.2 Aroma attributes with positive or negative


loadings > 0.30 for Riesling at 20 months by
each panelist 39

3.3 Summary of the main aroma descriptors of


Riesling fermented by different yeast strains
at 7 and 20 months of age 42

3.4 Flavor attributes with positive or negative


loadings > 0.30 for Riesling at 7 months by
each panelist 44

3.5 Flavor attributes with positive or negative


loadings > 0.30 for Riesling at 20 months by
each panelist 47

3.6 Summary of the main flavor descriptors of


Riesling fermented by different yeast strains
at 7 and 20 months of age 49

3.7 Aroma attributes with positive or negative


loadings > 0.30 for Chardonnay at 7 months
by each panelist 56

3.8 Aroma attributes with positive or negative


loadings > 0.30 for Chardonnay at 20 months
by each panelist 59
LIST OF TABLES (Continued)

Page
3.9 Summary of the main aroma descriptors of
Chardonnay fermented by different yeast
strains at 7 and 20 months of age 62

3.10 Flavor attributes with positive or negative


loadings > 0.30 for Chardonnay at 7 months
by each panelist 63

3.11 Flavor attributes with positive or negative


loadings > 0.30 for Chardonnay at 20 months
by each panelist 66

3.12 Summary of the main flavor descriptors of


Chardonnay fermented by different yeast
strains at 7 and 20 months of age 68

3.13 Aroma attributes with positive or negative


loadings > 0.30 for Pinot Noir at 7 months
by each panelist 75

3.14 Aroma attributes with positive or negative


loadings > 0.30 for Pinot Noir at 20 months
by each panelist 78

3.15 Summary of the main aroma descriptors of


Pinot Noir fermented by different yeast
strains at 7 and 20 months of age 81

3.16 Flavor attributes with positive or negative


loadings > 0.30 for Pinot Noir at 7 months
by each panelist 83
LIST OF TABLES (Continued)

Page

3.17 Flavor attributes with positive or negative


loadings > 0.30 for Pinot Noir at 20 months
by each panelist 85

3.18 Summary of the main flavor descriptors of


Pinot Noir fermented by different yeast
strains at 7 and 20 months of age 88

4.1 ANOVA of the means of the composition analysis


of Riesling fermented with different yeast strains 112

4.2 ANOVA of the means of the composition analysis


of Chardonnay fermented with different
yeast strains 114

4.3 ANOVA of the means of the composition analysis


of Pinot Noir fermented with different
yeast strains 116

4.4 Carboxylic acids identified by Kovats Indices


and GC-MS spectra and the range of
the relative concentration in Riesling
fermented with Zymaflore VL-1 and
Epernay 2 CEG 123

4.5 Carboxylic acids identified by Kovats Indices


and GC-MS spectra and the range of
the relative concentration in Chardonnay
fermented with EC-1118 Prise de
Mousse and D47 124
LIST OF TABLES (Continued)

Page

4.6 Alcohols identified by Kovats Indices


and GC-MS spectra and the range of
the relative concentration in Riesling
fermented with Zymaflore VL-1 and
Epernay2CEG 125

4.7 Alcohols identified by Kovats Indices


and GC-MS spectra and the range of
the relative concentration in Chardonnay
fermented with EC-1118 Prise de
Mousse and D47 127

4.8 Esters identified by Kovats Indices and GC-MS


spectra and the range of the
relative concentration in Riesling fermented
with Zymaflore VL-1 and Epernay 2 CEG 128

4.9 Esters identified by Kovats Indices and GC-MS


spectra and the range of the
relative concentration in Chardonnay
fermented with EC-1118 Prise de Mousse and
D47 130

4.10 Terpenes identified by Kovats Indices


and GC-MS spectra and the range of
the relative concentration in Riesling
fermented with Zymaflore VL-1 and
Epernay 2 CEG 131

4.11 Miscellaneous compounds identified by Kovats


Indices and GC-MS spectra and the range of
the relative concentration in Riesling
fermented with Zymaflore VL-1 and
Epernay 2 CEG 133
LIST OF TABLES (Continued)

Page
4.12 Miscellaneous compounds identified by Kovats
Indices and GC-MS spectra and the range
of the relative concentration in Chardonnay
fermented with EC-1118 Prise de
Mousse and D47 135

4.13 Tentatively identified and unidentified


compounds and the range of the relative
concentration in Riesling fermented with
Zymaflore VL-1 and Epernay 2 CEG 136

4.14 Unidentified compounds and the range of


the relative concentration in Chardonnay
fermented with EC-1118 Prise
de Mousse and D47 137
SENSORY AND CHEMICAL EVALUATION OF RIESLING,
CHARDONNAY AND PINOT NOIR FERMENTED BY DIFFERENT
STRAINS OF SACCHAROMYCES CEREVISIAE.

CHAPTER 1. INTRODUCTION

Winemaking often seems mysterious because of the tradition surrounding

the process. However, it is mainly a conversion of grape sugars into alcohol


and carbon dioxide upon the action of the yeast Saccharomyces cerevisiae.

Along with this process many volatile compounds are produced by yeasts

which give the wine its particular sensory attributes depending on precursors
found in each grape variety (Rapp and Mandery, 1986). The yeast strains are
selected on criteria such as fermentation rate and temperature, killer factor,

resistance to high alcohol and sulfites (Degre, 1992). Some attention is also
given to some strains that appear to influence the sensory development of the

wine. Controversies arose between researchers who believed that the yeast

strains would influence the aroma of the wine, and others who believed that
the sensory contribution was minimal, and disappeared with aging (Zoecklein

et al., 1990). The biochemical processes involved in the formation of volatile

compounds are very complex and several studies have demonstrated

different enzymatic activities in different yeast strains in the formation of esters,

alcohols, lactones, terpenes, sulfur-containing compounds etc. (Soles et al.,

1982; Stashenko et al., 1992; Etievant et al., 1983; Delfini, 1992; Lavigne et
al., 1992). Those compounds have been shown to be very important

contributors to the aroma of wine, and it is believed that more compounds are

yet to be found that are important.


The sensory evaluation of wines has been recognized as being very
important since it is used throughout the winemaking process, as well as the
final step before marketing. Sensory analysis of wines fermented by different

strains has not been extensively conducted. Work by Cavazza et al. (1993)
showed that yeast strains would be an important parameter since they

resulted in different aroma profiles.

Scientific sensory tastings are often laborious since they require training
and agreement of panelists on descriptors used before the actual tastings.

Since winemakers can be considered expert tasters for wine (Lawless, 1984),

and are important in the selection of yeast strains for fermentation, they are
ideal candidates for wine evaluation by Free-Choice Profiling to detect

differences or similarities produced by yeast strains. This method is different

from conventional descriptive analysis in that it requires little training and


individual winemakers are able to use descriptors with which they feel

comfortable, making the process more efficient. This technique has been
applied to coffee, chocolate, cheese, port and acids (Williams and Arnold,

1985; McEwan et al., 1989; Marshall and Kirby, 1988; Williams and Langron,

1984; Rubico, 1993), but never to wine. It is advantageous since it is a quick

and reliable means to describe wine with a panel of experts.

Many studies have reported several hundred volatile compounds in wines

(Schreier, 1979; Sefton et al., 1994; Montedoro and Bertuccioli, 1986). Some
studies were involved in determining different volatile compounds produced

by different yeast species and strains. Soles et al. (1982) studied different

ester production. Cabrera et al. (1988) studied the production of higher

alcohols and Delfini (1992) and Delcroix et al. (1994) focused their work on

some terpenes found in aromatic varieties such as Muscat and Riesling using
different commercial yeast strains. Many mechanisms were studied in yeasts
for the production of those volatile compounds. Enzymatic activities in
different yeast strains were often reported to vary form strain to strain. The
variation in enzymatic activity resulted in quantitative differences in volatile

compounds. The enzyme (E) substituted cinnamate decarboxylyase involved

in the production of vinylphenols varied from very high to almost nil in

commercial preparations of S.cerevisiae (Chatonnet et al., 1993). Esterase

activity (Suomalainen, 1981)) and alcohol dehydrogenase activity (Soles et


al., 1982) were also reported to vary between strains of S.cerevisiae, and
would have a direct effect in the production of esters and higher alcohols. The

quantitative differences could lead to different sensory perception in the

wines.
The research goals of this work were to study both the sensory profiles and

the chemical and volatile compositions of wine fermented with different yeast

strains. The work was divided into the following two main objectives. The first

objective was to determine the sensory profiles through use of Free-Choice

Profiling of three important commercial grape varieties in Oregon fermented

with several different yeast strains and to evaluate if the sensory profiles are

maintained with aging. The second objective was to investigate the important

chemical composition parameters of the three wine varieties including pH,

alcohol content, titratable acidity, volatile acidity, residual sugar, malate


content, as determinants of fermentation performance by different yeast

strains. In addition, two wines in the Riesling and Chardonnay trials were

selected for differences and similarities in their sensory profiles and their

volatile composition was studied by GC/MS.


CHAPTER 2. YEASTS AND WINES: AN OVERVIEW

Since 1906, commercial yeast preparations were recommended for


winemaking in California wineries (Bioletti, 1906). The benefits for using an

inoculum were to obtain a complete fermentation and to avoid unpleasant


flavors from "wild type" yeasts. Wine yeasts were recommended and not beer
or bread yeasts because they would impart a "beer" or "bread" flavor to the

wine (Bioletti, 1906).

Since then, the microflora of wine has been studied more extensively.

Several microorganisms are found in wines, including yeast, lactic acid

bacteria, acetic acid bacteria and molds (Fleet, 1990). Saccharomyces


cerevisiae is the principal yeast that conducts alcoholic fermentation. Other
yeasts such as Zygosaccharomyces, Kloeckera, Hanseniaspora and Candida

are also found (Fleet, 1992). With the availability of pure selected yeast

strains of S. cerevisiae in the form of active dry yeast, it has become easier to
assure a complete fermentation and prevent spoilage (Fleet and Heard,

1992). In winemaking, yeast strain selection is based on several parameters


such as volatile acidity and sulfide production, aromatic character,

completeness of fermentation (Fleet, 1990), tolerance to alcohol and

fermentation at low temperature (Degre, 1992).

Fermentation is a complex process where the activity of enzymes of the

yeast and the must components interact (Delfini, 1992). S.cerevisiae can use

the substrates found in grape juice. Under conditions of high concentration of


sugar and low pH of the must, yeasts follow a specific growth cycle and
metabolism. Fermentation activity is then slowly inhibited by the products of
sugar metabolism, primarily alcohol (Lafon-Lafourcades, 1986). Sugars, in
the form of glucose and fructose are the main growth and energy sources for
yeast. Nitrogen, in the form of ammonia and amino acids, phosphates,

minerals and vitamins are also present in the juice and are necessary for the
fermentation (Bisson, 1992). Once the wine is inoculated with a pure culture
of S. cerevisiae, sugar is transformed to ethanol and carbon dioxide via
glycolysis. Glucose and fructose are phosphorylated in the yeast cells, then

glucose-6-phosphate is transformed to ethanol via pyruvate by the enzyme

alcohol dehydrogenase (Bisson, 1992).

FLAVOR COMPOUNDS PRODUCED BY YEAST FERMENTATION

Hundreds of different compounds are responsible for the sensory

properties of wines. These compounds can originate from the grapes, from

processing, from fermentation and from aging (Rapp and Mandery, 1986).

The compounds coming from the grapes are mainly aldehydes, ketones,
some alcohols, sesquiterpenes alcohols, monoterpenes and their derivatives
(Rapp and Mandery, 1986).

During fermentation, most aroma/flavor compounds are produced (Rapp

and Mandery, 1986). The following diagram (Fig. 2.1) from Henschke and

Jiranek (1992) summarizes the route taken for the formation of most flavor

compounds. The grape constituents such as sugars and amino acids are
modified during fermentation (Fig. 2.1) to produce ethanol, fatty acids, esters,
higher alcohols, hydrogen sulfite and diacetyl.
sulfate
sugar sulfite

amino acids

keto acids

fatty acid CoA higher alcohols

1r *f

fatty acids esters

Fig.2.1 Sugar and amino acids metabolism during grape fermentation by


yeasts (Henschke and Jiranek, 1992).

Alcohols

Ethanol, methanol and glycerol are the most abundant alcohols found in

wines. Ethanol is important in the viscosity (body) of wine. It balances taste


sensation and acts as a fixer of odors (Rapp and Mandery, 1986). Other
alcohols are also present such as propan-1-ol (11-68 mg/L), 2-methyl propan-

1-ol (6-174 mg/L), pentan-1-ol (< 0.4 mg/L), 2-methylbutan-1-ol (19-96 mg/L),
3-methylbutan-1-ol (83-400 mg/L), hexan-1-ol (0.5-12 mg/L), octan-1-ol (0.2-
1.5 mg/L), 2-phenyl ethanol (25-105 mg/L), 2-(4-hydroxyphenol) ethanol (2-

4.8 mg/L), butane-2,3-diol (165-1250 mg/L), and 3-thiopropan-1-ol (0.05-2


mg/L) (Montedoro and Bertuccioli, 1986). At high concentration (>300 mg/L),
these alcohols bring a negative quality to the wines due to their fusel oil odor.
However, at lower concentration, they add a desirable complexity to the wine

(Montedoro and Bertuccioli, 1986). It was reported that different yeast strains

will influence the formation of fusel alcohols (Suomalainen and Lehtonen,

1979). The fusel alcohols are formed by yeast from exogenous amino acids
and from metabolism of sugars (Ayrapaa, 1971).

Aldehydes

Aldehydes are mostly detectable at the beginning of fermentation. They

are most probably reduced to alcohols and are also very important because of
their low sensory threshold (Suomalainen and Lehtonen, 1979). Aldehydes
can be produced from carbohydrate degradation (furfural, 5-

hydroxymethylfurfural), from lignin degradation (vanillin, cinnamaldehyde)

and can be formed during aging (Rapp and Mandery, 1986). Aldehydes not

reduced to alcohols during fermentation such as acetaldehyde (50-100 mg/L)

are also found in wines. Acetaldehyde is an intermediary product of yeast


metabolism from pyruvate, is responsible for the oxidized odor of table wine,
and is an indicator of the oxidation state of the wine (Schreir, 1979).

Acetaldehyde can also be formed by oxidation of ethanol during aging and

can also be produced by oxidative spoilage organisms. Acetaldehyde, along

with 2-keto acids are major by-products when the yeasts produce fusel

alcohols from amino acids and sugars. Once they are formed by the yeast
8

cells, they are transferred to wine (Nykanen, 1986). It was also found that total
aldehyde content seemed to depend on the yeast used. It would appear that
the amount of acetaldehyde produced is related to the activity of pyruvate

decarboxylase which may vary with yeast strain (Nykanen, 1986). Other

aldehydes such as hexanal, frans-hex^-enal, and c/s-hex-2-enal have a

typical "grassy" odor. Hexanals and hexenals are formed by enzymatic

cleavage of linoleic and linolenic acids respectively (Montedoro and


Bertuccioli, 1986). Those aldehydes of C-6 chains are formed by enzymes
during breaking of the grape cell (Rapp and Mandery, 1986).

Ketones

The ketones arising from the grapes are primarily 2-, and 3-alkanones.
Ketones resulting from the fermentation are found in small amounts. The
majority of ketones have little sensory influence except butanedione which is

found to be important in bacteria spoiled wine. Several acetals such as

formaldehyde, dimethylacetal and diethylacetal have been found in wines

although their contribution to wine flavor is unknown (Montedoro and

Bertuccioli, 1986). The sensory impact of ketones formed during fermentation,


such as acetoin, acetone and 2,3-pentadione seems very low. Diacetyl is one

of the important ketones and imparts a buttery odor to wine (Rapp and

Mandery, 1986).
Acids

Tartaric and malic acids are responsible for the acid taste in wine. Several

acids are also present but are too volatile to contribute to odor except for

acetic (vinegar), propionic (goaty) and butanoic (spoiled butter). Of the acids

produced during fermentation, the carboxylic acids are the most important.
These acids are often associated with unpleasant odors but usually have high

detection thresholds. Sugar acids such as gluconic, glucoronic and

galacturonic are also found in wines as well as formic, lactic, pyruvic, citric and

succinic acids (Montedoro and Bertuccioli, 1986).

Esters

Esters are found in grapes and are also a product of fermentation and

aging. The majority of fatty acid esters are formed during fermentation.

Different esters can contribute particular notes and influence the aroma.
Biosynthesis of esters is similar to the synthesis of fatty acids, a-keto acids

undergo oxidative decarboxylation to form acyl-CoA, and acetyl-CoA is


formed by the decarboxylation of pyruvic acid. The formation of fatty acid ethyl

esters occurs by the reaction of acyl CoA with ethanol. Several enzymes are
required for this reaction (Suomalainen and Lehtonen, 1979).

Ethyl hexanoate, octanoate, decanoate and dodecanoate (1-10 mg/L) are

all important in the flavor of wines. It was reported that aroma of wine
fermented at lower temperature is more fruity and soapy and that the amount
of fatty acid esters (Ce, Cs, C-io, C12) increased by decreasing the
10

fermentation temperature. Ethyl, isobutyl and isopentyl esters, as well as


monoacetates were a result of yeast fermentation as reported by Schreir
(1979). Nykanen and Nykanen (1977) reported that strains of S. cerevisiae

were found to produce more isopentyl acetate, ethyl hexanoate, ethyl

octanoate, ethyl decanoate and phenylethyl acetate than Saccharomyces


uvarum. Another experiment conducted by Soles et al. (1982) found

significantly different ester concentrations in white wines fermented with 14


different yeast strains. The esters identified were isoamyl acetate, ethyl

hexanoate, hexyl acetate, ethyl octanoate, 2-phenylethyl acetate and ethyl

decanoate. They also found that the S. cerevisiae strains had different
alcohol dehydrogenase activity and the formation of higher alcohols was
reported to be influenced by the yeast strains and higher alcohols and their

corresponding esters shared the same precursors. Herraiz and Ough (1993)

reported some differences in the production of ethyl esters of amino acids by

two different commercial yeast strains. It appeared that the ethyl esters came

mainly from amino acids synthesized by the yeasts. Other esters, such as
hydroxy and oxo acid esters have a slight influence on wine flavor. So far, up
to 170 various esters have been found in wines, among them also are ethyl

vanillate and ethyl cinnamate (Montedoro and Bertuccioli, 1986). Ethyl

cinnamate is not very abundant but exhibits a strong pleasant odor (Etievant

and Bayonove, 1983).

Terpenes

Monoterpenoids are important flavor compounds. They include free flavor

compounds, flavorless polyhydroxylated compounds (polyols) and glycosidic


11

derivatives (Williams et al., 1988). Monoterpenes, such as linalool, nerol,


geraniol, citronellol, a-terpineol and hotrienol are important in Muscat cultivars

and aroma related cultivars such as Riesling and are responsible for the

varietal character in those wines (Colagrande et al, 1993). Linalool is found in


highest quantity but the amount of terpenes varies indifferent grape varieties.
Linalool contributes a lavender-like flowery note to various Muscat varieties,

White Riesling and Gewurztraminer but is found in very small amounts in Pinot

Noir (Brander et al., 1980). Most of the terpenes are found in grapes but they

have also been reported to be produced by microorganisms (Montedoro and

Bertuccioli, 1986). Dugelay et al. (1992) reported that the terpene citronellol

(lime odor) was not formed in the absence of yeast. They noted that citronellol
was not formed in the presence of nerol and geraniol even in the presence of

glycosidic enzymes. However, they found that citronellol was produced in

Ugni Blanc wine inoculated with commercial yeasts and the concentration

was higher in must supplemented with geraniol. There was also an increase

in citronellol in must spiked with nerol. Yeast strain had an influence on the

amount of citronellol produced by fermenting the must with different


commercial S. cerevisiae .

Monoterpenes that are glycosidically-bound are odorless and flavorless,

and to release the terpene from the sugar moiety, glycoside enzymes are

necessary (Williams et al., 1988). Aryan et al. (1987) also showed that

glycosidases (found in grapes and in some yeast strains) could be used to


release glycosidically bound flavorants such as monoterpenes. The aroma
precusors are present, and are mainly glycosides of linalool, nerol and

geraniol. When hydrolysis of the glycoside moiety occurs, the aglycon is then

aroma active (Colagrande et al., 1994). Darriet et al. (1988) also found that
12

the B-glycosidases of yeasts would liberate the monoterpenes such as


geraniol and nerol and that the enzyme was glucose independent (the
enzyme in other fungi is inhibited by glucose). This last property is particularly

important in must fermentation considering the high amount of sugar present.

The work of Cabrera et al. (1988) showed differences in amount of terpenes


produced from the fermentation of two strains of S. cerevisiae . Of those two

strains, one had higher content of volatile acidity, lactic acid esters, terpenes
and isobutanol. Grossmann and Rapp (1988) reported that linalool
concentration increased after treatment with pure B-glycosidase. Nerol, oc-

terpineol and geraniol concentrations increased the most with treatment with a
glucanase containing 6-glycosidase activity after fermentation. The wines
made with this last treatment were ranked highest for quality and varietal

character. However, Delcroix et al. (1994) showed the release of the odorous
moiety is not only dependent on 6-glucosidase, but also on a-arabinosidase

and a-rhamnosidase activities, which were similar in three commercial yeast

strains that they studied. They also reported that the yeast B-glycosidase
stability was very low at wine pH since the maximum activity for this enzyme is
at pH 5.0. However differences in trans-iuran linalool oxide, nerol, geraniol,
citronellol and hotrienol were also reported. Citronellol was absent in the

juice but present in the wine since it is produced from nerol and geraniol by

yeast during fermentation.

Amines

There are two mechanisms for the formation of amines: one through

decarboxylation of amino acids and the other by transamination of an


13

aldehyde. Even though the importance on odor is unknown, there are several
volatile amines found in wines such as methylamine, dimethylamine and
ethylamine which appear to be a result of the fermentation by S. cerevisiae

(Montedoro and Bertuccioli, 1986).

Sulfur Compounds

There are several sulfur compounds produced during fermentation.

Usually they have an unpleasant odor, but they are found in small quantities,

below identification threshold. Some may contribute desirable odors to wines

(Montedoro and Bertuccioli, 1986). Their maximum concentration is found at

the end of the fermentation and are predominant in "reduced" wines (Lavigne

et al., 1992). The formation of sulfur compounds is closely linked to the


metabolism of yeast (Rapp and Mandery, 1986) and the production of

hydrogen sulfide is affected by the must composition as well as yeast strains.

Lavigne et al. (1992) studied the parameters influencing the formation of those

compounds during fermentation. One of the parameter studied was the yeast

strain. Six different commercial yeast strains of S.cerevisiae were found to

produce significant difference in methionol concentration, ranging from 250 to


430 |ig/L. They also concluded that the sulfur containing compounds

produced by the yeast were directly influenced by the composition of the must

as well as by fermentation conditions. Thomas et al. (1994) obtained similar


results when they showed that hydrogen sulfide production was related to

yeast strain-medium interaction in both a model medium and grape juice.

Hydrogen sulfide (rotten egg odor with a recognition threshold of lug/L), alkyl
sulfides (thioethers), thiolanes, esters of sulfur containing acids, thiazols,
14

mercaptal, acetamide and thiols are most commonly found. Sulfur containing
amino acids can be used by yeasts to produce dimethyl sulfide and 3-
methylthiopropan-1-ol (Montedoro and Bertuccioli, 1986; Rapp and Mandery,

1986).

Lactones

Lactones are formed by the dehydration of aliphatic hydroxy acids. Yeast

has been found to synthesize some lactones, and lactones such as 4-hydroxy-
3-methyloctanoic (y-lactone) is found in oak. The most common lactone found

in wine is y-butyrolactone (Montedoro and Bertuccioli, 1986). The y lactones

are important in the flavor of sherries where they occur in high concentration,
however, they are found in low amounts in wines (Rapp and Mandery, 1986).

Phenols

Phenols are extracted from skins, juice and pulp during processing by

yeast or bacteria metabolism and some phenols are produced by alcohol


extraction. All volatile phenolic compounds except acetovanillone have been

identified in wines but not in grapes. It was deduced that the volatile phenols

were either yeast metabolic products or cleavage products of phenols found in


grapes (Schreir, 1979). Suomalainen and Lehtonen (1979) reported that

yeasts play a role in the formation of phenolic compounds since they can

decarboxylate p-coumaric acid, ferulic acid and vanillin to respectively 4-ethyl


phenol, 4-ethyl guaiacol and 4-methyl guaiacol. Vinyl phenols are
15

guaiacol. Vinyl phenols are intermediates in this decarboxylation. They are


characterized by having a medicinal, phenolic, balsamic floral, spicy, vanilla
or clove-like odor. Chatonnet et al. (1993) reported the enzyme responsible
for the release of the vinylphenol was substituted (E) cinnamate

decarboxylyase and its activity would vary in different commercial yeast

strains. The enzyme possesses endocellular activity, is constitutive and only

expressed during alcohol fermentation. Grando et al.(1993) obtained similar


results with different yeast strains having different pof + (phenolic off odors

gene) activity. Williams et al. (1988) analyzed for volatile phenols following

glycosidase treatment of Muscat, White Riesling and Gewurztraminer and

found compounds such as vanillin, propiovanillone, methylvanillate,

zingerone, zingerol, homovanillyl alcohol, coniferyl alcohol and

dihydroconiferyl alcohol.

VOLATILE COMPOUNDS ANALYSIS

Wine Extraction

In order to recover the volatile compounds of wines, they must be first


extracted, then concentrated since some of the compounds are in minute

quantities. Several methods of recovery are available including distillation,

head space methods and extraction with Freon, an approved method in wine

flavor research (Tracey and Britz, 1989).

Freon 113 (1,1,2-trichloro-1,2,2-trifluoroethane) is a non polar solvent used


to extract the non-polar compounds in the wine. In wine research, it has been
16

successfully used to extract volatile compounds, and for identification (Akhtar


et al., 1985). The advantages of using Freon 113 is that it is inexpensive, easy
to handle, has low toxicity, is inert and won't extract water or ethanol (Tracey

and Britz, 1989; Miranda-Lopez, 1990). It also gives extracts of good sensory

properties (Blakesley and Loots, 1977). Other solvents such as

dichloromethane (Cobbs et al., 1978) and ethyl acetate (Acree, private


communication, 1993) can also be used.

Gas Chromatography

Gas Chromatographic (GC) analysis provides a means to separate,

quantify and identify the compounds present in wine extracts. The separation
is achieved by partitioning the components between the mobile phase (carrier
gas) and a stationary phase (Gordon, 1990). It has been the method of choice

for the analysis of the volatile composition of wines. Higher alcohols


(Holloway and Subden, 1991), alcohol acetates, ethyl esters of C4-C8 fatty

acids (Stashenko et al., 1992) and many volatile compounds are separated by
GC. Tentative identification can be done by retention index (Maga, 1990).
Retention index (Kovats index) is an index for a specific compound based on

its elution time. Under identical conditions, hydrocarbon standards and the

samples are injected (Maga, 1990) and the Kovats indices can be calculated

with the following formula:


17

Index = 100 logt'n - logt'x X100n

logt'n - logfn+l

Where
n = retention time of hydrocarbon before compound X

n+1 = retention time of hydrocarbon after compound X

x= retention time of compound X

(Kovatsetal., 1958)

When two compounds possess the same or close retention indices, spiking
the extract with the suspected standard can help to further identify the

compound. Using a different column can also help identify the compounds

since a column with a different stationary phase and a different polarity will

interact differently with the compounds. For each compound, you obtain two

retention indices and with reference tables it can be tentatively identified.

Gas Chromatography- Mass Spectrometry

Mass spectrometry (MS) is used for the structural elucidation of volatile

compounds (Jennings and Shibamoto, 1980). GC-MS has had extensive use

in wine composition analysis. It has been used to identify sulfur-containing

compounds (Rapp et al., 1985), and ethylphenols produced by spoilage


yeasts (Chatonnet et al., 1992). Stashenko et al. (1992) used GC-MS to study

the major compounds formed during yeast fermentation including volatile


compounds such as acetals, alcohols and ethyl esters of C4-C8 fatty acids.
18

Most often, GC-MS is used to profile specific wines. Nelson et al (1978)


identified 36 volatile compounds of Catawba wines with GC-MS after Freon

extraction. Quantitative estimation was also done against an internal


standard. Pinot Noir harvested at different maturities was also analyzed by
GC-MS, and characterized by aroma and flavor. McDaniel et al. (1989) and

Miranda-Lopez (1990) identified aroma active compounds in Pinot Noir by

GC-sniffing and then identified those compounds by GC-MS. Brander et al.

(1980) was able to identify 66 volatile compounds of Pinot Noir by GC-MS,

and the confirmation of the compounds was done by Kovats indices values for
each compound. With these two means of identification, it is possible to
confirm the identification of a compound. Avedovech et al. (1992), and later

Sefton et al. (1993) supplied important information on the composition of

Chardonnay. Sefton et al (1993) were able to identify 180 compounds found

in wines treated with various enzymes or acids to release glycosidically bound

volatile compounds. Avedovech et al. (1992) identified alcohols, esters,


aldehydes and ketones in Chardonnay before and after malo-lactic
fermentation.

SENSORY EVALUATION

In the sensory evaluation of wines, descriptive analysis is most often used.


This evaluation method requires training of panelists for several sessions with

specific standards associated with the product (Meilgard, 1991). Rosi and

Bertuccioli (1992) used descriptive analysis on experimental wines fermented

with different fatty acid mixtures, including acetate, ethyl esters, and medium

chain fatty acids. Those compounds are flavor active and usually described
19

by four aroma attributes (banana, artificial fruit, fusel alcohol, soapy). The
terms were chosen according to their importance in defining the sensory
attributes from a wine aroma standardized terminology (Noble et al., 1987).
An analysis of variance was used to confirm the influence of fatty acid mixtures
on the different pattern of flavor active compounds of wines.

Descriptive analysis can also be used to compare quality rating of wines.

Noble et al. (1984) used this method to describe Bordeaux wines and found

that differences could be detected across the wines only with green
bean/green olive aroma, and astringency and bitterness in flavor. They also

found no significant differences among the wines in quality rating by wine

experts.

McDaniel et al. (1989) used different techniques for the sensory evaluation

of Pinot Noir wines made from grapes harvested at different maturities. A

profile of the wines was achieved by descriptive analysis using a 15 point


intensity scale with trained panelists. The panelists were first trained with
aroma, flavor and intensity standards before evaluating the wines. A panel of

Oregon winemakers was also used to evaluate the wines in both intensity and

quality, using two different scales.

Free-Choice Profiling (FCP) is a new approach in sensory evaluation that

enables each individual panelist to use his or her own vocabulary to describe
a product (Williams and Langron, 1984). This method considerably reduces

the time needed to train a panel and obtain a complete agreement on the

descriptors to be used since panelist will often use them in a different way.

Other problems found during the training of a panel is the scoring variation of

the panelists; variation during sessions and different assessors may perceive
20

different stimuli in the same product (Arnold and Williams, 1986) and FCP will
help overcome these.

STATISTICAL ANALYSIS

Generalized Procrustes Analysis (GPA) is used for the data analysis after
Free-Choice Profiling. First, the centroids of each panelist's data are matched,
in order to eliminate the variation if different parts of the scales are used.

Afterwards, the differences in the scoring range used by the different panelists

are removed by the isotropic scale changes. And thirdly, close matches are

obtained for the configurations by rotation and reflection of the axes (Arnold

and Williams, 1986). A consensus configuration is obtained by matching as

closely as possible all the panelists' perceptual space.


Williams and Langron (1984) evaluated commercial ports with FCP. They

rated the ports in terms of appearance, flavor and aroma. They showed that

there was no need to develop a specific vocabulary since the analysis

revealed a fair degree of agreement in respect to the shape of the

multidimensional plots.

LITERATURE CITED

Acree, T. Private communication. 1993.

Akhtar, M., Subden, R.E., Cunningham, J.D., Fyfe, C, and A.G. Meiring.
Production of volatile yeast metabolites in fermenting grape musts. Amerine,
M.A., and C.S. Ough. Wine and Musts analysis. John Wiley & Sons. NY
(1980).
21

Arnold, G.M., and A.A. Williams. The use of Generalized Procrustes


techniques in sensory analysis. In: Statistical Procedures in Food Research.
Piggot, J.R. (Ed.), pp. 233-255. Elsevier Applied Science. London (1986).

Aryan, A.P., Wilson, B., Strauss, C.R., and P.J. Williams. The properties of
glycosidases of Vitis vinifera and a comparison of their B-glucosidase activity
with that of exogenous enzymes. An assessment o.f possible application in
enology. Am. J. Enol. Vitic. 38(3): 182-188 (1987).

Avedovech, R.M. Jr., McDaniel, M.R., Watson, B.T., and W.E. Sandine. An
evaluation of combinations of wine yeast and Leuconostoc oenos strains in
malo-lactic fermentation of Chardonnay wine. Am. J. Enol. Vitic. 43(3): 253-
260 (1992).

Ayrapaa, T. Biosynthetic formation of higher alcohols by yeast. Dependence


on the nitrogenous nutrient level of the medium. J. Inst. Brew. 77: 266-276
(1971).

Bioletti, FT. Pure yeast in wineries. University of California at Berkeley.


College of Agriculture. Circular No.23 (1906).

Bisson, L. Yeasts - Metabolism of sugars. In: Wine Microbiology and


Biotechnology. Fleet, G.H. (Ed), pp. 55-76. Hardwood Academic Publishers.
Switzerland (1992).

Blakesley, C.N., and J. Loots. A convenient method for multiple extraction of


volatile flavor components from food slurries and pulp using a two-chambered
glass bomb extractor and dichlorodifluoromethane (Freon 12) solvent. J.
Agric. Food Chem. 25(4): 961-963 (1977).

Brander, C.F., Kepner, R.E., and A.D. Webb. Identification of some volatile
compounds of wine of Vitis vinifera cultivar of Pinot Noir. Am.J. Enol. Vitic.
31(1): 69-75 (1980).

Cabrera, M.J., Moreno, J., Ortega, J.M., and M. Medina. Formation of ethanol,
higher alcohols, esters and terpenes by five yeast strains in musts from Pedro
Ximenez grapes in various degrees of ripeness. Am. J. Enol. Vitic. 39(4): 283-
287 (1988).
22

Chatonnet, P., Dubourdieu, D., Boidron, J.-N., and M. Pons. The origin of
ethylphenols in wines. J. Sci. Food Agric. 60: 165-178 (1992).

Chatonnet, P., Dubourdieu, D., Boidron, J.N., and V. Lavigne. Synthesis of


volatile phenols by Saccharomyces cerevisiae in wines. J. Sci. Food Agric.
62:191-202 (1993).

Cobbs, C.S., Bursey, M.M., and A.C. Rice. Major volatile components of
Aurore wine: a gas chromatographic - mass spectrometric analysis combining
ionization and electron impact ionization. J. Food Science. 43: 1822-1825
(1978).

Colagrande, O., Silva, A., M.D. Fumi. Recent applications of biotechnology in


wine production. Biotechnol. Prog. 10: 2-18.(1994).

Darriet, P., Boidron, J.-N., and D. Dubourdieu. Hydrolysis of the terpenic


heterosides of small Muscat grapes using the periplasmic enzymes of
Saccharomyces cerevisiae . Connaissance de la Vigne et du Vin. 22(3): 189-
192 (1988).

Degre, R. Selection and commercial cultivation of wine yeast and bacteria. In:
Wine Microbiology and Biotechnology. Fleet, G.H. (Ed), pp. 421-448.
Hardwood Academic Publishers. Switzerland (1992).

Delcroix, A., Gunata, Z., Sapis, J.-C, Salmon, J.-M., and C. Bayonove.
Glycosidase acitvities of three enological yeast strains during winemaking;
effect on the terpenol content of Muscat wine. Am. J. Enol. Vitic. 45(3):291-296
(1994).

Delfini, C. Innovative trends in oenology and in the selection of yeasts and


malolactic bacteria for wine industry. Riv. Vitic. Enol. 1:17-30 (1992).

Dugelay, I., Gunata, Z., Sapis, J.C., Baumes, R., and C. Bayonove. Etude de
I'origine du citronellol dans les vins. J. Int. Sci. Vigne Vin. 26(3): 177-184
(1992).

Etievant, P.X., and C.L. Bayonove. Aroma components of pomaces and wine
from a variety Muscat de Frontignac. J. Sci. Food Agric. 34: 393-403 (1983).
23

Fleet, G.H. Growth of yeasts during wine fermentation. J. Wine Res. 1(3): 211-
223 (1990).

Fleet, G.H. The microorganisms of wine making- isolation, enumeration and


identification. In Wine Microbiology and Biotechnology. Fleet, G.H. (Ed), pp.1-
26. Hardwood Academic Publishers. Switzerland. (1992).

Fleet, G.H., and G.M. Heard. Yeasts - Growth during fermentation. In: Wine
Microbiology and Biotechnology. Fleet, G.H. (Ed), pp. 27-54. Hardwood
Academic Publishers. Switzerland (1992).

Gordon, M.H. Principles of gas chromatography. In: Principles and


Applications of Gas Chromatography in Food Analysis, pp.11-58. Ellis
Horwood Ltd. (1990)

Grando, M.S., Versini, G., Nicolini, G., and F. Mattivi. Selective use of wine
yeast strains having different volatile phenols production. Vitis. 32:43-50
(1993).

Grossmann, M., and A. Rapp. Steigerung des sortentypischen Weinbuketts


nach Enzymbehandlung. Deutsche Lebensmittel-Rundschau. 84(2): 35-37
(1988).

Henschke, P.A., and V. Jiranek. Yeasts - Metabolism of nitrogen, compounds.


In: Wine Microbiology and Biotechnology. Fleet, G.H. (Ed), pp. 77-164.
Hardwood Academic Publishers. Switzerland. (1992).

Herraiz, T., and C.S. Ough. Formation of ethyl esters of amino acids by yeasts
during the alcoholic fermentation of grape juice. Am. J. Enol. Vitic. 44(1): 41-
48 (1993).

Holloway, P., and R.E. Subden. Volatile metabolites produced in a Riesling


must by wild yeast isolates. Can. Inst. Sci. Technol. J. 24(1/2): 57-59 (1991).

Jennings, W., and T. Shibamoto. Qualitative analysis of flavor and fragrance


volatile compounds by glass capillary gas chromatography. pp 22-24.
Academic Press Inc. NY (1980).
24

Kovats, E. von, Simon, W., and E. Heilbronner. Programm - Gesteverter Gas-


Chromatograph zur praparativen trennung von genishen organisher verbin
dungen. Helv. Chim. Acta. 32: 275- 278 (1958).

Lafon-Lafourcade, S. Applied microbiology. Experientia. 42: 904-914 (1986).

Lavigne, V., Boidron, J.N., and D. Dubourdieu. Formation des composes


soufres lourds au cours de la vinification des vins blancs sees. J. Int. Sci.
Vigne et du Vin. 26(2): 75-85 (1992).

Maga, J.A. Analysis of aroma volatile compounds. In: Principles and


Applications of Gas Chromatography in Food Analysis. Ellis Norwood Ltd.
(1990).

McDaniel, M.R., Miranda-Lopez, R., Watson, B.T., Michaels, N.J., and L.M.
Libbey. Pinot Noir aroma: A sensory/gas chromatography approach. In:
Flavors and Off-Flavors, Proceedings of the 6th International Flavor
Conference. Rethymnon Crete, Greece, 5-7 July, 1989. pp. 23-36 (1989).

Meilgard, M., Civille, G.V., and B. Carr. Sensory Evaluation Techniques. 2nd
Ed. CRC Press (1991).

Miranda-Lopez, R. A maturity trial study of Pinot Noir wines: Aroma profile by


sniffing gas chromatographic effluent. Thesis, Oregon State University (1990).

Montedoro, G., and M. Bertuccioli. The flavour of wines, vermouth and fortified
wines. In: Food Flavours. Part B. The flavour of beverages. Morton, I. and A.J.
MacLeod (Eds.) Elsevier Publishing Co., London (1986).

Nelson, R.R., Acree, T.E., and R.M. Butts. Isolation and identification of
volatiles from Catawba wines. J. Agric. Food Chem. 26(5): 1188-1190 (1978).

Noble, A.C., Arnold, R.A., Buechsenstein, J., Leach, E.J., Schmidt, J.O., and
P.M. Stern. Modification of a standardized system of wine aroma terminology.
Am. J. Enol. Vitic. 38(2): 143-146 (1987).
25

Noble, A.C., Williams, A.A., and S.P. Langron. Descriptive analysis and quality
ratings of 1976 wines from four Bordeaux communes. J. Sci. Food. Agric. 35:
88-98 (1984).

Nykanen, L. Formation and occurrence of flavor compounds in wine and


distilled alcoholic beverages. Am. J. Enol. Vitic. 37(1).: 84-96 (1986).

Nykanen, L, and I. Nykanen. Production of esters by different yeast strains in


sugar fermentations. J. Inst. Brew. 83: 30-31 (1977).

Rapp, A., Guntert, M., and J. Almy. Identification and significance of several
sulfur-containing compounds in wine. Am. J. Enol. Vitic. 36(3): 219-221
(1985).

Rapp, A., and H. Mandery. Wine aroma. Experientia. 42: 873-883 (1986).

Rosi, I., and M. Bertuccioli. Influences of lipid addition on fatty acid


composition of Saccharomyces cerevisiae and aroma characteristics of
experimental wines. J. Inst. Brew. 98: 305-314 (1992).

Schreir, P. Flavor composition of wines: a review. CRC Critical Reviews in


Food Science and Nutrition. 12: 59-111 (1979).

Sefton, M.A., Francis, I.L., and P.J. Williams. The volatile composition of
Chardonnay juices: A study by flavor precursor analysis. Am. J. Enol. Vitic.
44(4): 359-370 (1993).

Soles, R.M., Ough, C.S., and R.E. Kunkee. Ester concentration differences in
wine fermented by various species and strains of yeasts. Am. J. Enol. Vitic.
33(2): 94-98 (1982).

Stashenko, H., Macku, C, and T. Shibamato. Monitoring volatile chemicals


formed from must during yeast fermentation. J. Agric. Food Chem. 40: 2257-
2259 (1992).

Suomalainen, H., and M. Lehtonen. The production of aroma compounds by


yeast. J. Inst. Brew. 85: 149-156 (1979).
26

Thomas, C.S, Boulton, R.B., Silacci, M.W., and W.D. Gubler. The effect of
elemental sulfur, yeast strain, and fermentation medium on hydrogen sulfide
production during fermentation. Am. J. Enol. Vitic. 44(2): 211-216 (1993).

Tracey, R.P, and T.J. Britz. Freon 11 extraction of volatile metabolites formed
by certain lactic acid bacteria. App. Env. Microbiol. 55(6): 1617-1623 (1989).

Williams, A.A., and S.P. Langron. The use of Free-Choice Profiling for the
evaluation of commercial ports. J. Sci. Food Agric. 35: 558-568 (1984).

Williams, P.J., Strauss, C.R., and B. Wilson. Development in flavour research


on premium volatiles. In: Proceedings of the 2nd International Cool Climate
Viticulture and Oenology Symposium. Auckland, New Zealand (1988).
27

CHAPTER 3. SENSORY EVALUATION OF RIESLING,


CHARDONNAY AND PINOT NOIR FERMENTED WITH DIFFERENT
COMMERCIAL SACCHAROMYCES CEREVISIAE STRAINS BY
FREE-CHOICE PROFILING AT SEVEN
AND TWENTY MONTHS OF AGE

Ann Dumont, Mina R. McDaniel and Barney T. Watson

Department of Food Science and Technology, Wiegand Hall 100, Oregon


State University, Corvallis, OR 97331

Presented in Parts at the American Society of Enology and Viticulture,


Anaheim, CA, June 28-30, 1994
And Eastern American Society of Enology and Viticulture,
Cincinnati, OH, July 15-17 1994.
28

ABSTRACT

The 1992 Riesling, Chardonnay and Pinot Noir wines fermented with

different commercial strains of Saccharomyces cerevisiae were characterized

with the sensory technique of Free-Choice Profiling. The wines were tasted at

seven and twenty months of aging by a panel of expert winemakers to verify if


yeast strains had an influence on the sensory characteristics of the wines, and
if so, if the differences were maintained or evolved with time. The panelists'

scores were analyzed by Generalized Procrustes Analysis (GPA) to provide

information on the consensus profile of the wines in both aroma and flavor.

The results showed that in all three varieties, the panelists were able to

perceive differences, as well as similarities between the wines fermented with


the different yeast strains.

In the Riesling wines, GPA results indicated a separation of the wines

fermented with Simi White and CS-2 strains in both aroma at 7 and 20 months

and in flavor at 7 months. The grouping appeared different in flavor at 20

months. The other three strains, Zymaflore VL1, K1 and Epernay 2 CEG were

similar in principal axis 1, but were separated in principal axis 2 in aroma at 7


and 20 months, and in flavor at 7 months, but the differences were different in
flavor at 20 months. In the Chardonnay analysis, differences and similarities

were not maintained after the aging period in both aroma and flavor. Profiling

of the wines showed a less distinct separation compared to the Riesling

wines. The GPA results for the Pinot Noir at 7 months showed a significant

poor replication by the panelists making the profiling of the wines difficult. At
20 months of aging, panelists replicated well and differences and similarities

could be found in both aroma and flavor of most wines. This last finding
29

confirmed that the influence of the yeast strains was not lost during the period
of aging studied.

INTRODUCTION

In the sensory evaluation of wines, descriptive analyses have been the

method of choice (Noble, 1988). However, this method is time-consuming

since it involves the training of panelists. Free-Choice Profiling (FCP) can be

used in order to utilize a panel of expert winemakers as panelists to evaluate

the wine. There is no need for extensive training since the winemakers can
qualify as trained or expert since they have extensive experience with the
product (Lawless, 1984). This technique is relatively new and is used to

profile a product; it will consider that the panelist's perception will vary since

they are allowed to use their own words for description with minimal training

(Williams and Arnold, 1985). This method has been used previously to

describe whisky (Guy et al., 1989), commercial ports (Williams and Langron,

1984), lager beers (Gains and Thompson, 1990), chocolate (McEwan et al.,
1989) and acids (Rubico, 1993). The method was used to assess different

aspects of the products such as aroma and flavor, as well as appearance.

The panelists background ranged from experts, and trained panelists to

consumers, showing the adaptability of the technique. The results are

analyzed through Generalized Procrustes Analysis (GPA) in the following


way. The centroids of each panelists data are matched, therefore eliminating
variation if different parts of the scale are used. Then, the differences in the

scoring range used by the different panelists will be removed by the isotropic

scale changes. Lastly, close matches are obtained for the configurations by
30

rotation and reflection of the axes (Arnold and Williams, 1986). The result is a
consensus configuration that matched as close as possible all the panelists'

perceptual space.
Sensory evaluation is a crucial part of winemaking and should be a
convenient tool in research to optimize wine quality. Yeast selection for

fermentation has improved in the last twenty years and the role of yeast in the
conversion of sugar to alcohol and carbon dioxide is critical. Some studies

have reported that the yeast strain selected would not influence the sensory

development of the wine, and if it did, it would be lost with aging (Zoecklein et
al., 1990). Yeast strains are selected in different wine regions of the world and
their metabolic activity may differ in some cases. It could be assumed that the

production of volatile compounds during the fermentation may also be

different, and might result in wines with different sensory attributes.

In this study, Free-Choice Profiling is used to evaluate differences and

similarities in Oregon Riesling, Chardonnay and Pinot Noir fermented with

different commercial yeast strains. This technique was also used to evaluate if
the differences or similarities varied with aging of the wines.

MATERIALS AND METHODS

Wine Preparation

Three varieties of grapes were used for the research. Chardonnay and

Riesling were harvested from the Lewis Brown Farm, Corvallis, Oregon on

October 1, 1992. The grapes were crushed and destemmed, and pressed
31

using a Willmes bladder press. The must was pooled and 50 ppm SO2 was

added using 1% potassium metabisulfite solution. The must was settled


overnight at 40C, racked from the solids and separated into 10 lots (2 lots per
yeast) of 23.2 L each in glass fermentors, equipped with a fermentation lock.

The juice was inoculated with 20g/hL of active dry yeast.

The five yeasts used for Chardonnay were Saccharomyces cerevisiae

strains: Lalvin EC-1118 Prise de Mousse from Epernay, France; Lalvin QA23

from Portugal; Lalvin Bourgoblanc 3079 from Bureau International du Vin de


Bourgogne (BIVB), France; Lalvin M Montrachet, an isolate obtained from the
University of California, Davis; and Enoferm ICV D47 from C6tes-du-Rhone.

The five yeasts used for Riesling were Saccharomyces cerevisiae strains:
Enoferm Simi White, an isolate from the Pasteur Institute, France; Lalvin CS-2

from the Alsace region, France; Lalvin Zymaflore VL-1 from Bordeaux, France;

Lalvin K1 (V-1116) from Montpellier, France and Uvaferm Epernay 2 CEG


from the Rheingau region in Germany.
The white wines were fermented at 20oC. After fermentation was

completed, the wines were racked from the solids and 1% potassium
metabisulfite was added to obtain 20 ppm of free SO2. The wines were cold

stabilized for 1 month at 40C and fined with 0.36 g of bentonite per liter. After

settling, the wines were racked from the bentonite and 1% potassium
metabisulfite was added to obtain 20 ppm of free SO2. The wines were

bottled unfiltered in 750 mL bottles and stored at 40C.

The Pinot Noir grapes were harvested October 14, 1992 from Beaver Creek

Vineyard, Corvallis, Oregon. Thirty four kg/lot (2 lots per yeast) were crushed
and destemmed and 50 ppm of SO2 was added using 1% potassium

metabisulfite. The must was inoculated 2 hours after SO2 addition with 20
32

g/hL of the following six active dry yeast; Lalvin Wadenswil WSK-27 from
Switzerland; Enoferm Assmannshausen, from the Geisenheim Institute in
Germany; Lalvin 2056 from C6tes-du-Rhone, France and from Bourgogne,
France, the Lalvin Bourgorouge 212, Bourgorouge RA17 and Enoferm

Burgundy. The wines were fermented at 25-30oC on the skins for 14 days

prior to pressing. The wines were punched down twice daily. Pressing was

done with a Willmes bladder press, and the wines were settled, then racked
from the solids. The new wines were inoculated with 0.006 g per liter of
Leuconostoc oenos (Lalvin OSU strains). After completion of malo-lactic

fermentation, the wines were racked, 1% potassium metabisulfite solution was


added to obtain 20 ppm of free SO2 and the acidity was adjusted to pH 3.65

with the addition of tartaric acid. Before bottling the wines were cold

stabilized, then racked and 1% potassium metabisulfite was added to obtain


20 ppm of free SO2. At bottling, the wines were rough filtered with Seitz 200

filter pads (20 x 20 cm) first washed with 3% citric acid solution, then rinsed

with distilled water. Then 1% potassium metabisulfite was added to obtain 20


ppm of free SO2 before bottling in 750 mL wine bottles. The bottled wines

were stored at 40C.

Sensory Evaluation

Free-Choice Profiling

Sixteen wines were evaluated for aroma and flavor. The wines included:
Riesling (5), Chardonnay (5) and Pinot Noir (6) all fermented with different
33

yeast strains. The experiment was divided in two parts. The first tasting was
done on the young wines (7 months after fermentation), and the second
tasting on the older wines (20 months after fermentation).

A panel of 12 experienced commercial winemakers were selected. The

first wines tested were Riesling, followed by Chardonnay, then Pinot Noir.

Rest periods were allowed between each variety, as well as between


replicates. The panelists were first instructed on the principles of Free-Choice
Profiling (FCP). They were given a sample of each wine and were asked to

generate descriptors in aroma, then flavor, and to rate the intensity on a 15

points intensity scale. At the end of the training, the descriptors were

discussed and defined, and each panelist was instructed to use his/her own

descriptors and/or some terms described by other panelists during the


discussion following the training. The panelists were also instructed to keep

the same descriptors throughout the tastings of the wines of the same variety.

Two replications of each wine were evaluated. The samples were

presented randomly, according to MacFie et al. (1989) to balance the effect of

order of presentation. The wines were identified with 3-digit random numbers.

The samples were poured in a volume of 60 mL in wine tasting glasses, and


covered immediately with aluminum glass covers until tasting.

A second tasting was conducted after 20 months of aging of the wines. The

procedure was identical to the one described above. Of the original panelists,

9 out of the 12 were present. In the Chardonnay evaluation, the wines

fermented with the 3079 strain were omitted due to the spontaneous malo-

lactic fermentation that occurred.


34

Statistical Analysis

The data were analyzed using Procrustes Analysis with Procrustes-PC


Version 2.2 (Dijksterhuis and van Buuren, 1991) and by Minitab® Statistical

Software Release 8.2 (Ryan and al., 1985). The data of the 10 or 12 panelists

were arranged in a matrix of 10 rows (2 replications of 5 samples) by n

columns, where n equals the number of descriptors used by each panelist.

The generalized Procrustes analysis was done on the matrices. Differences


were determined through Analysis of Variance (ANOVA) and least significant

differences (LSD) were determined using the Fisher pairwise comparison on

the principal axis scores (transformed consensus mean of each sample). The

sample size was n = 5 for Riesling and Chardonnay, and n = 6 for Pinot Noir.

The principal axes (dimension) considered were the ones with eigenvalues >
1.00. The descriptors considered were the ones with loading values > 0.30.
Each descriptor was then listed per panelist for each principal axis used in the

analysis with the positive or negative loading values to evaluate any contrast

between the principal axes. Consensus plots were generated to locate the

position of the samples in a two-dimensional space where the percent of

variability in each dimension was indicated. The descriptors associated with


each principal axis were then applied to the consensus plot for evaluation of
their relationship with the samples.
35

RESULTS AND DISCUSSION

Riesling Sensory Evaluation

The first tasting was done on the wines 7 months after fermentation.

Throughout the tastings, the panelists used terms that were mostly in
accordance with wine aroma terminology (Noble, 1987). In the Riesling, the
attributes used for the profiling of the wines are shown in Table 3.1 for aroma

at 7 months. In this analysis, Principal axis 1 (PA1) and Principal axis 2 (PA2)

are described with attributes with both positive and negative loading values.

The negative or positive signs of the loading values do not infer a negative or

positive character in the wine, but rather an indication of the intensity or


position in that particular dimension. The attributes with loadings greater than
+/- 0.30 were reported. Analyses by Williams and Langron (1988) in the

evaluation of commercial ports and Williams and Arnold (1986) in the

evaluation of coffees, both by FCP, showed that each descriptor and its

loading value could show a contrast between the two principal axes. The

results of the Riesling aroma (Table 3.1) showed a contrast between yeasty,
reduced, musty, stinky, pine, herbal, spicy and few fruity descriptors in PA1
compared to more fruity, honey, honeysuckle/floral and very few musty,

yeasty, reduced descriptors in PA2.

The consensus plot in aroma (Fig. 3.1) of the 12 panelists showed that PA1

and 2 accounted for 56% of the variability, which is lower than results from

Williams and Langron (1984) where they obtained 96% with two dimensions
but similar to Dijksterhuis and Punter (1980), and McEwan et al. (1989). The
difference could be attributed to the different program they used for the
36

Table 3.1 Aroma attributes with positive or negative loadings > 0.30 for Riesling
at 7 months by each panelist.

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)

1 yeasty (+ 0.55), reduced (+0.52) herbal (-0.68), honey(+0.58)


honey (-0.46), peach (-0.33) reduced (+0.36)

2 garlic (+0.68), floral (-0.53) honey (-0.67), ripe pear (-0.47)


yeasty (+0.46) floral (+0.40)
3 grapefruit (+0.70), alcohol (-0.61), spicy (+0.48)
cardboard (-0.41),cheesy (+0.40) grapefruit (-0.37), honeysuckle
(-0.37)

4 pine (+0.59), mango (+0.56) pine (+0.49), lime (+0.47)


grapefruit (+0.40) mango (-0.43), apple (+0.41)

5 yeasty (-0.80) grapefruit (+0.74), skunky


pineapple (+0.31) (+0.39), peach (+0.37)

6 skunky (+0.64), floral (-0.58) floral (+0.69), pine (+0.51)


bubblegum (-0.36) grapefruit (-0.47)

7 burnt (+0.56), herbal (+0.49) floral (+0.62), citrus (-0.53)


spice (+0.40), musty (+0.35) orange (+0.34), musty (+0.30)

8 grassy (+0.72), citrus (-0.42) yeast (+0.66)


repressed (-0.40) melon (+0.61)

9 flint/steel (-0.51), stinky (+0.49) rhubarb (-0.53),apricot(+0.46)


cheesy (+0.39), apricot (-0.33) flint (-0.46), cheesy (+0.37)

10 dusty/musty (-0.70), herbal (+0.53) peas (+0.56), floral (+0.53)


floral (-0.34) dusty/musty (-0.46), fruity
(+0.37)

11 lemon (+0.58), burnt (+0.42) burnt (+0.60), pineapple


peach (-0.39), yeasty (+0.33) (+0.52), peach (+0.43),
apple (-0.32) apple (-0.30)

12 pine (+0.77), pineapple (-0.46) pineapple(+0.79),yeasty(+0.50)


yeasty (+0.43) floral/perfume (+0.32)
1.0
citrus 1a/1b Simi White
floral 2a/2b CS-2
tree fruit 3a/3b VL-1
yeasty 4a/4b K1
pine 5a/5b CEG
0.5
1 3b 1aB
CM
B
co
3a
2b
x Q 4b 2a
CO
D ■
0.0

IS
o "-' 1 4a
c 1b
13
a B
\ 5b 5a
-0.5

herbal
honey
alcohol
-1.0 T ' 1 ■ —i —r ■ r ■ 1
-1.0 yeasty/musty -0.5 0.0 0.5 yeasty 1.0
flint/steel Principal axis 1 skunky
floral (39%) herbal
fruity pine
tropical fruit

Fig. 3.1 Sample consensus plot for Free-Choice Profiling of the


aroma of Riesling (7 months) fermented by different yeast strains

CO
~vl
38

analysis (Rubico, 1993) and because some panelists would use highly
correlated descriptors, which is a common problem in FCP (Gains et al.,
1988).
The five wines showed good replication in aroma, demonstrated by the
closeness of each replicate, as well as no significantdifference was found by

ANOVA of the PA scores for the sample replicate at p < 0.05. The descriptors

at the bottom of each side of the axes helped differentiate the samples. They

are also listed in order of highest loading, and frequency of use by the

panelists. The Simi White and CS-2 wines appeared to be grouped, and
defined in terms of the positive spaces of PA1 and PA2. They were separated

from the other samples in terms of skunky, garlic, herbal, pine, tropical fruit

and citrus notes in PA1. In PA2, only CS-2 wines replicated well enough to be

associated with floral, citrus, tree fruit, yeasty and pine descriptors. Both

samples were opposed to the other three wines, VL-1, K1 and CEG. The VL-1

and CEG wines were both located in the same area of PA1, but in different
spaces for PA2. It could be interpreted that VL-1 and CEG wines were better

separated by fruity and floral characters than by musty, flint, citrus, floral and

fruity of PA1. The K1 wines were situated between VL-1 and CEG, and would

be described similarly as the two other wines. The ANOVA results of the PA

scores showed significant differences between treatments at the 0.05 level. A

Fisher pairwise comparison showed that differences were between CS-2, K1


and CEG, between VL1 and CEG, and between Simi White and CEG.

After 20 months of aging, the wines were tasted using the same method but
with 10 panelists. The attributes with loadings greater than +/- 0.30 are listed
in Tables 3.2 for aroma in the PA1 and PA2. In aroma, the attributes were

almost the same as for the seven months tasting, except for descriptors such
39

Table 3.2. Aroma attributes with positive or negative loadings > 0.30 for Riesling
at 20 months by each panelist

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)
1 reduced (+0.67), tropical fruit grass (+0.61), pine (-0.53)
(-0.47), grassy (+0.32) floral (-0.48)
2 earth (-0.79), honey (+0.43) linalool (+0.62), green tea
jasmine (+0.32) (-0.46), rose (+0.45),
honey (+0.36)
3 musty (+0.66) tropical fruit (-0.64),
sulfury (+0.61) peach/apricot, (-0.50), pine
(+0.42), musty (-0.37)

4 leesy (+0.76) citrus (+0.61), nutty (-0.49)


floral (-0.62) leesy (+0.45), floral (+0.43)
5 citrus/qrapefruit (+0.64), sulfide tropical fruit(+0.47),
(-0.51) honeysuckle (+0.32) peach (+0.44), herbal (-0.41),
musty (-0.33)

6 asparagus (+0.62), tropical fruit asparagus (+0.54), tropical fruit


(-0.55)lemon/lime (+0.39), paper (+0.54), green tea (-0.39),
(+0.30) melon , (+0.33)

7 earthy (-0.60), tropical fruit (+0.55) citrus (-0.71), orange peel


diesel (-0.48) (-0.62), tropical fruit (-0.32)

8 grass (+0.89) tropical fruit (+0.69), apricot


tropical fruit (-0.34) (-0.50), grass (+0.39)

9 musty (+0.56), yeasty (+0.56) citrus (-0.62), green tea (+0.47)


apple (-0.42) lime (+0.30)

10 yeasty (+0.53), canned fruit (-0.51) sharp (-0.60), musty (-0.42)


honey suckle (-0.43), talc (-0.36) peach/apricot (+0.42), tropical
fruit, (+0.36),
honeysuckle (+0.30)
40

as jasmine, asparagus, paper, canned fruit, diesel, green tea and linalool.
However, it might just reflect an evolution or change in the vocabulary of the
wine makers. The contrast between the two axes was harder to determine. It
would appear however that PA1 reflected more musty, yeasty, earthy

attributes, along with tropical fruit, citrus and floral notes to a lesser extent.

PA2 reflected more the tropical fruit, citrus, peach/apricot, floral and pine notes
in the wines.
In the consensus plot of aroma (Fig. 3.2), the same separation was

maintained after 20 months, and again the samples replicated well (p < 0.05)

as shown in the ANOVA results. Both axes explained 56% of the variability,

which is identical to the first tasting. Simi White and CS-2 wines were closely

associated and were principally defined by PA1, with more musty, asparagus,
citrus and jasmine and honey descriptors. Both wines did not appear to be

well defined in terms of PA2, where more fruity notes are found. The VL-1

wines were well defined in both axes. However, they were more defined in
terms of musty, floral, diesel, talc and tropical fruit in PA1. In PA2, citrus,

tropical fruit, peach/apricot, orange peel (more fruity) and some herbal/green

tea and pine descriptors were used to separate the wines. The K1 wines were
more closely associated with the CEG wines. Both wines are represented by
the same descriptors in PA1 and PA2, such as musty, floral, diesel, talc and

tropical fruit in PA1 and grassy, pine, floral, honey and tropical fruit in PA2.

Significant differences with ANOVA were found at the 0.05 level between the

samples. In the Fisher pairwise comparison, differences were found between

Simi White and VL-1, between K1 and CEG, between CS-2 and VL-1, and
between K1 and CEG.
10
citrus
la/lb Simi White
tropical fruit
peach/apricot 2a/2b CS-2
orange peel B 3a/3b VL-1
3a
herbal 4a/4b K1
pine 5a/5b CEG
0.5
musty
green tea
I Q3b

CM 1a
CO
0.0 a 2a lb
2b
4b
I 4a
B
"5
c
I 5b
-0.5 B
1 5a
grassy
floral
honey
pine -1.0 —T ■■-
citrus ! o earthy -0.5 o.o 0.5 reduced 1.0
tropical fruit floral musty
honeysuckle honeysuckle Principal axis 1 asparagus
tropical fruit (45%) jasmine
talc honey
apple grapefruit
diesel lemon/lime

Fig. 3.2 Sample consensus plot for Free-Choice Profiling of the


aroma of Riesling (20 months) fermented with different yeast strains
42

Table 3.3 summarizes the descriptors associated with each wine at both 7
and 20 months of age for aroma. All the young wines were described with

yeasty, skunky and musty terms which are often found in very young wines.

We can also see that two groups are formed.

Table 3.3. Summary of the main aroma descriptors of Riesling fermented by


different yeast strains at 7 and 20 months of age

Yeast strain Descriptors


7 months 20 months

Simi White yeasty, skunky, herbal reduced, musty, asparagus,


pine, tropical fruit jasmine, honey, citrus,
tropical fruit, fruity, orange
peel, herbal, pine, green tea

CS-2 yeasty, skunky, herbal reduced, musty, asparagus,


pine, tropical fruit, citrus, tropical fruit, fruity, orange
floral, tree fruit peel, herbal, pine, green tea

VL-1 yeasty, musty, flint/steel, earthy, floral, honeysuckle,


floral, fruity, citrus, tropical fruit, talc, apple,
tree fruit, pine diesel, citrus, fruity, orange
peel, herbal, pine, musty,
green tea

K1 yeasty, musty, flint/steel, earthy, floral, honeysuckle,


floral, fruity, citrus, tropical fruit, talc, apple,
tree fruit, pine diesel, grassy, honey, pine,
citrus

CEG yeasty, musty, flint/steel, earthy, floral, honeysuckle,


floral, herbal, honey, tropical fruit, talc, apple,
alcohol diesel, grassy, honey, pine,
citrus
43

The Simi White and CS-2 wines have comparable descriptors and are
different only with the descriptors of dimension 2. Simi White and CS-2 wines

are more predominant with herbal and pine descriptors. As those wines aged,

they were described identically in both PA2, shown by the same aroma
profiles and appeared to be more complex with a variety of descriptors, from
floral, honey and fruity descriptors but also herbal, pine descriptors. The VL-1,

K1 and CEG wines were described comparably in PA1, but are different in the
descriptors of PA2. The wines were more a contrast in terms of fruits, citrus

and floral terms compared to Simi White and CS-2. As the wines aged, they

still maintained the fruitiness, as well as floral (honeysuckle), and some grassy
and pine terms.
In the flavor analysis at 7 months of aging (Table 3.4), a contrast between

the two axes was observed; PA1 separated samples based on the flavor of the
fruit, as well as herbal/vegetal notes a few basic tastes (sweet, bitter, acid)

whereas PA2 separated samples on mouthfeel and the basic tastes, as well

as citrus characters and less fruit.


Flavor analysis of the young Riesling in the consensus plot (Fig.3.3)

showed that both axes explained 62% of the variability. Again, the samples

replicated well as shown by the ANOVA results (p < 0.05), in PA1, but not so
well in PA2, especially the K1 wines. Simi White and CS-2 were separated

from the other three wines. They were described by the more acid,

vegetal/herbal, bitter and astringent descriptors in PA1. They were harder to


define in PA2, where they tended to be positioned toward the center of the
dimension. Both samples were neutral regarding the intensity of the basic

tastes and mouthfeel characters compared to the other wines. Samples VL-1

were closely associated with the viscous, fruity and sweet descriptors of PA1
44

Table 3.4. Flavor attributes with positive or negative loadings > 0.30 for
Riesling at 7 months by each panelist.

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)
1 reduced (-0.52), apricot (+0.49) reduced (+0.67), herbal (-0.53)
honey (+0.47), flinty (-0.47) apricot (+0.34)

2 burnt rind (-0.60), honey/sweet (+0.46) soapy/leesy (+0.68),


floral (+0.45), green apple (+0.41) honey/sweet (+0.58)
apple (-0.34)

grapefruit (-0.53), heat/alcohol (-0.50) acid (+0.92)


leesy (-0.43), melon (+0.42)
sweet (+0.34)
4 mango (-0.71) honey (+0.59), lime (+0.49)
lime (-0.62) mushroom (-0.42), pine (+0.35)

5 acid/tart (-0.77), apple (-0.45) grapefruit (+0.97)


viscous (-0.43)

6 vegetal (-0.59), coarse (-0.47) tart (+0.52), sweet (+0.47)


sweet (+0.42), apricot (+0.35) vegetal (+0.42), coarse (-0.33)
tart (-0.33) musty (-0.32)

7 finish (-0.74) tart (+0.51), finish (-0.50)


tart (-0.47) citrus (+0.45), honey (+0.33)

8 vegetal/herbal (-0.76), bitter (-0.42) honey (+0.52), bitter (-0.50)


alcohol (+0.31) citrus (-0.47), orange (-0.34)

9 vegetal (-0.65), apple (+0.52) steely (-0.61), vegetal (-0.50)


bitter (-0.33) apricot (-0.34), peach pit
(+0.30)

10 citrus (-0.62), acid (-0.53) viscous (+0.75)


viscous (+0.44), astringent (-0.38) citrus (+0.63)

11 grapefruit (-0.47), pear (+0.43) sour (-0.68), coarse (-0.39)


peach (+0.40), astringent (-0.34) burnt (+0.38)

12 viscous (+0.87), yeasty (+0.34) toasty (+0.74), bitter (-0.50)


fruit/apple/pear (-0.33) fruit/apple/pear (+0.37)
1.0
sweet/honey 1a/1b Simi White
yeasty 2a/2b CS-2
tart 3a/3b VL-1
citrus 4a/4b K1
5a/5b CEG
0.5-

2bB
1
B
CM 1a 3b a
3a
B
4b
X
CO -—. D
0.0
D El
g.£!. lb
o 1 2a 5b
c
5a

-0.5--1 4a

bitter
herbal
musty -1.0
citrus -1.0
acid -0.5 0.0 0.5 viscous 1.0
coarse vegetal/herbal tree fruit
Principal axis 1
bitter sweet
citrus
(42%)

Fig. 3.3 Sample consensus plot for Free-Choice Profiling of the


flavor of Riesling (7 months) fermented by different yeast strains

en
46

and sweet, tart/citrus and yeasty descriptors of PA2. Similarly, CEG was
associated with the same descriptor in dimension 1 but to a lesser extent, and
associated more closely in PA2 to bitterness, herbal, musty and some citrus

and coarse notes. Once again, K1 could only be associated with the

descriptors of CEG and VL-1 in PA1, but it was ambiguous in terms of PA2

where it seemed to be less separated by those descriptors. The ANOVA

results between the PA scores of the samples showed significant differences


at the 0.05 level between the yeast strains. A Fisher pairwise comparison
showed difference between CS-2 and K1, and between VL1 and K1.

The attributes in flavor of the older wines with a loading > 0.30 are

represented in Table 3.5 and are similar to the ones used in the first tasting.

The contrast between the two axis is not as evident as in aroma. Both axis

seem to explain the variability between basic tastes, fruits and citrus.
However, PA1 would appear to explain differences in viscosity/body, whereas
PA2 would be more on the finish/harshness and less fruit.

The consensus plot for flavor of the older wines (Fig. 3.4) showed that 66%

of the variability was explained by both dimensions. Except for Simi White, all

the samples position in space and their interrelationships are different than

from the 7 month tasting. Among the samples, the flavor probably might have

evolved differently than the aroma or the winemakers used different terms to
describe the wines. Simi White was described in PA1 by acid, citrus and

bitterness, and in PA2 by finish, floral, honey, sweet and citrus/acid


descriptors. K1 wines appeared to have descriptors similar to Simi White in

PA1, but less intense, more neutral descriptors in terms of astringency, fruity

and finish in PA2. The CS-2 wines were more concentrated in the center of
both axis, making any association difficult. VL-1 and CEG both were
47

Table 3.5. Flavor attributes with positive or negative loadings > 0.30 for Riesling
at 20 months by each panelist

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)
1 sweet (+0.66) astringent (-0.92)
acidic (-0.53), viscous (+0.44) sweet (+0.35)

2 grapefruit (-0.57) green apple (+0.73)


peach (+0.46) honey (+0.54)
linalool (+0.43) pine (-0.33)
pine (-0.37), lime (-0.31)
3 body (+0.69) astringent (-0.54)
herbal (+0.47) acid (+0.47)
honey (+0.33) fruity (-0.44)
peach/apricot (+0.31) peach/apricot (-0.34)
body (+0.33)

4 leesy (-0.76) leesy (+0.58)


flowery (+0.64) flowery (+0.58), acid(+0.52)

5 peach (+0.68) finish (+0.78)


citrus (+0.54) citrus (+0.37)
sweet (+0.35), rich (+0.31)
6 body (+0.77), fruit (+0.45) finish (-0.79), harsh (+0.52)
7 acid (-0.68) viscous (+0.68)
viscous (+0.47) acid (+0.46)
lemon (+0.44), harsh (-0.36) harsh (-0.43), lemon (-0.37)

8 viscous (+0.62) apricot (-0.78)


apricot (+0.48) bitter (-0.45)
acid (-0.40), sweet (+0.37)
9 harsh (-0.65) body (+0.45)
tart (-0.50), sweet (+0.40) sweet (+0.44), tart (-0.43)
body (-0.35) citrus (+0.35)
peach/apricot (-0.34)

10 bitter (-0.49) acid (-0.73)


canned fruit (+0.43) canned fruit (-0.53)
yeasty (-0.32) viscous (+0.31)
10
finish layibSimi White
green apple 2a/2b CS-2
leesy • 3ay3b VL-1
floral 4a/4b K1
honey 5a/5b CEG
sweet 0.5
acid
citrus 0
la
CM
5bE
B
lb
«
5a a
S--. o.o
_ss 4a E, a Q
2> 0°
o.£J. •
0 3b 3a
o 4b
c
QL -0.5

astringent
finish H 2a
apricot
lemon
acid 1.0
fruity -1.0 leesy -0.5 0.0 0.5 1.0
grapefruit sweet
pine Principal axis 1 floral
acid
harsh (44%) peach
bitter apricot
pine viscous
lime fruity

Fig. 3.4 Sample consensus plot for Free-Choice Profiling of the


flavor of Riesling (20 months) fermented by different yeast strains

00
49

described similarly in PA1 with body/viscosity, sweet/honey, floral, herbal and


fruity descriptors. However, VL-1 had more of the astringent, fruity, citrus, acid

and finish descriptors, and CEG was higher in finish (mouthfeel after

swallowing), floral, honey, sweet and citrus and acid descriptors. Since both
wines seem to be located near the center of the PA2 space, it is harder to
make definite association with the descriptors of PA2.

Table 3.6 summarizes the descriptors associated with each wines at both 7

and 20 months of age for the flavor of Riesling. We can see that two groups

are formed in the young wines, as was found for the aroma.

Table 3.6. Summary of the main flavor descriptors of Riesling fermented by


different yeast strains at 7 and 20 months of age

Yeast strain Descriptors


7 months 20 months

Simi White acid, vegetal, herbal, leesy, grapefruit, acid, harsh,


bitter, citrus bitter, pine, lime, finish,
green apple, floral, honey,
sweet

CS-2 acid, vegetal, herbal, astringent, finish, apricot,


bitter, citrus lemon, acid, fruity, pine

VL-1 viscous, tree fruit, sweet, body, sweet, floral, fruity,


honey, yeasty, tart, citrus viscous, herbal, honey,
astringent, finish, lemon,
acid, pine

K1 viscous, tree fruit, sweet leesy, grapefruit, acid, harsh,


bitter, pine, lime, astringent,
finish, apricot, lemon, fruity

CEG viscous, tree fruit, sweet, body, sweet, floral, fruity,


bitter, herbal, musty, viscous, herbal, honey,
citrus, coarse finish, leesy, acid, citrus
50

The Simi White and CS-2 wines have comparable descriptors. The young
wines were more acid and herbal, and as they aged, the flavor of both wines
were described differently. Simi White was more citrus, acid, leesy, and

harsher, while CS-2 also had citrus and acid descriptors, but also more fruit,
and more astringency, although astringency and harshness might mean the

same thing. The young VL-1, K1 and CEG wines were generally described as

more viscous, fruity and sweet, a contrast with Simi White and CS-2. As the
wines aged, only VL-1 and CEG were described similarly, whereas K1 appear
to approach more the description of the Simi White wines. However, all five

wines remained different to some degree in both aroma and flavor at 7 and 20

months of age.
Fig. 3.5 through 3.8 illustrate the percentage consensus and percentage

within (residual) obtain from GPA. It can show how the panelists performed,
and how well they agreed in the tasting of the young and older wines. A high

percentage of the consensus (lower part) represents a high agreement within

the panelists. The lower the residual error, also means a higher agreement
within the panelists. In the aroma analysis of the young wines (Fig. 3.5), there

was high agreement between the panelists in all wines, as shown by the high

consensus, and low residual error. In the aroma analysis of the older wines
(Fig. 3.6), there was high agreement between the panelists in all wines, as
shown by the high consensus, and low residual error similar in all wines. In

the flavor analysis of the young wines (Fig. 3.7), the panelist had a high

agreement for the CS-2 and VL-1 wine, a medium agreement for K1 and CEG
wines and a lower agreement for the Simi White wines. In the flavor analysis

of the older wines (Fig. 3.8), the panelist also had a high agreement for all the
□ % within
% consensus

c
o

>
(0
o
o

CS-2 VL-1 K1 CEG


Riesling sample

Fig. 3.5 Percentages of consensus and within (residual) variation


distributed over the Riesling samples for the aroma at 7 months of age

en
c
o

CO
>
(O
o

m h P CS VL 1 K1 CEG
Riesling sample

Fig. 3.6 Percentages of consensus and within (residual) variation


distributed over the Riesling samples for the aroma at 20 months of age

l\5
c
o
.2
CO
>

o
O

CS-2 VL-1 K1
Riesling sample

Fig. 3.7 Percentages of consensus and within (residual) variation


distributed over the Riesling samples for the flavor at 7 months of age

en
CO
D % within
% consensus

simi Wh e CS 2 LI K 1 CEG
Riesling sample

Fig. 3.8 Percentages of consensus and within (residual) variation


distributed over the Riesling samples for the flavor at 20 months of age

en
4^
55

wines. The results show well that in general, aroma profiling was easier than
the flavor profiling by the winemakers.

The sensory profile of both young and older Rieslings shows that
differences are found in both and since yeast strains were the only variable in
the experiment, any differences can be attributed to them. We were able to

find patterns between yeast strains and sensory attributes. Generalized

Procrustes Analysis (GPA), as reported by Noble (1988), can be used as a

multivariate technique between sensory and instrumental analyses, in our

case, between yeast strains that perform differently and sensory analysis.
GPA can then provide information on the relationship between the samples

and the attributes (Rubico, 1993).

Chardonnay Sensory Evaluation

The attributes with loadings greater than +/- 0.30 are shown in Table 3.7 for
aroma in 7 month old Chardonnays. The contrast between the two axes was

not very well defined. However, axis 1 tended toward nutty, butter, reduced,

floral, estery/candied/aldehyde and tropical fruit descriptors, whereas PA2

tended more to separate on the basis of citrus, vanilla, yeasty/musty,

metallic/tin can and honey descriptors.


The consensus plot of Fig. 3.9 in aroma shows that both axes explained
55% of the variability. There was no significant difference between the

replicates as shown by ANOVA (p < 0.05) and by the position of the samples

in space. The EC-1118 and M wines tended to be located toward the center

of both axes, maybe indicating that the wines did not stand out with the

attributes used in both dimensions. However, EC-1118 tended toward the


56

Table 3.7. Aroma attributes with positive or negative loadings > 0.30 for
Chardonnay at 7 months by each panelist.

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)

1 reduced (+0.70), apple (-0.37) honey (-0.56), floral (-0.42)


herbal/sage (-0.32) apple (-0.39), pineapple (-0.32)
2 nutty (+0.75) vanilla (-0.55), metallic (+0.40)
apple (-0.42) yeast (-0.40), apple (+0.38)
stemmy/herbal (-0.35)

3 butter (+0.70), apple (+0.41) reduced (+0.83), yeasty (+0.52)


soapy (-0.41)

4 coconut (+0.62), wet straw (+0.43) honey (+0.55), apple(+0.54)


green apple (-0.37), rose (-0.37) coconut (+0.34), pear (+0.33)
5 chemical (+0.72), pear (-0.43) melon (-0.56), musty (-0.44)
coconut (-0.36) chemical (+0.34), vanilla
(+0.33)

tropical fruit (-0.54), butter (+0.49) peach (+0.57), yeasty (-0.57)


nutty (+0.42), reduced (+0.38) reduced (+0.37), tropical fruit
(-0.32)

7 candied (+0.54), aldehyde (+0.54) herbal (-0.63), lemon (+0.50)


lemon (-0.46) vanilla (-0.40), orange (-0.38)
8 sour (-0.94) repressed (-0.69), nutmeg
(+0.45), alcohol (+0.39)

9 leesy (-0.88) pear (-0.56), smokey (+0.53)


green (+0.36) butter (-0.43), grapefruit (-0.31)

10 smokey (+0.71), pineapple (+0.59) vegetal (+0.83)


floral (+0.31) vanilla (-0.47)

11 estery (-0.70), apple (+0.34) lemon (-0.60), apple (-0.57)


lemon (+0.31), floral (-0.30) tin can (+0.43)
12 fruity (-0.59), spicy (-0.56) spicy (+0.81)
yeasty (+0.48), toasty (+0.33) fruity (-0.51)
apple 1.0
honey 1a/1bEC-1118
metallic 2a/2b QA23
fruity 3a/3b 3079
coconut 4a/4b M
reduced 5a/5b D47
2a i
0.5 -
3b a
CM
(0
'£ 2b 1a
lb
0.0
o B4a
c 5b G
4b H 3a a

5a ■
-0.5
floral
apple
honey
tropical fruit
yeasty
vanilla -1.0 T I
fruity -1.0 apple -0.5 0.0 reduced x.o
0.5
citrus floral nutty
herbal Principal axis 1 butter
lemon (35%) lemon
tropical fruit tropical fruit
yeasty yeasty

Fig. 3.9 Sample consensus plot for Free-Choice Profiling of the aroma
of Chardonnay (7 months) fermented by different yeast strains

en
58

positive side of the axes, whereas the M wines were more towards the
negative side, and most likely less separated by those descriptors. The QA23
and D47 wines were both defined quite similarly in PA1 with apple, floral,

lemon, tropical fruit, herbal and yeasty descriptors. However, QA23 was more

separated in PA2 in terms of honey, metallic, coconut, fruity and reduced


terms. D47 wines on the other hand, were better separated in PA2 by honey,
vanilla, fruity and citrus descriptors. The 3079 wines were separated since

they underwent spontaneous malo-lactic fermentation. The wines had typical


aroma characters associated with whites that undergo ML fermentation such

as nutty and butter. The ANOVA analysis on the PA scores of the samples

was done without the 3079 wines, and it showed significant differences at the
0.05 level. Using the Fisher pairwise comparison, specific differences were
found between EC-1118, QA23 and D47, between QA23 and M, and between

M and D47 wines.

After 20 months of aging, the attributes with loadings > 0.30 (Table 3.8) are

shown for aroma. The aroma attributes showed a contrast between PA1 and

PA2, where PA1 separated the samples more in terms of musty/toasty/yeasty


and herbal/vegetal, and some in terms of citrus, tropical fruit and butter. PA2
separated the samples with more fruity terms such as apple/pear/peach, and

tropical fruit such as pineapple, and citrus fruit such as lime and grapefruit.
This axis also separated the wines in terms of herbal/vegetal, and yeasty.

The consensus plot in aroma (Fig. 3.10) showed that both axes explained

72% of the variability, which is an increase from the first tasting. The EC-1118,
QA23 and M samples replicated well but the ANOVA found significant
difference at the 0.05 level between the replicates, due to the poor replication

of the D47 wines. During the tasting, the panelists noted that one of the D47
59

Table 3.8. Aroma attributes with positive or negative loadings > 0.30 for
Chardonnay at 20 months by each panelist

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)

1 steely (+0.61) yeasty (+0.74)


menthol (-0.48) steely (-0.43)
vegetal (+0.39) menthol (-0.35)

2 musty (+0.78) peach (+0.84)


butter (-0.40), lime (-0.35) lime (-0.42)

3 musty (+0.68) vegetal (+0.58)


toasty (+0.44) apple (-0.57)
vegetal (+0.36), herbal (+0.35) citrus (-0.47)

4 musty (+0.82) herbal (+0.89)


tropical fruit (-0.48) tropical fruit (-0.33)

5 vanilla (+0.70) citrus (-0.67)


pear (+0.41) yeasty (+0.50)
citrus (-0.40) vanilla (-0.34)

6 corked (+0.90) pineapple (+0.69)


canned corn (-0.35) canned corn (-0.46)
citrus (+0.43)
steely (-0.30)

7 cork (+0.68) grapefruit (-0.53)


grapefruit (-0.40) cork (-0.53)
herbal (-0.33) vegetal (+0.52)
lemon (-0.32)

8 musty (+0.82), perfume (-0.42) canned corn (+0.88)

9 musty (+0.43) herbal (+0.68)


floral (-0.42) skunky (-0.33)
tropical fruit (-0.42) yeasty (-0.33)
skunky (+0.40) tropical fruit (-0.33)
yeasty (+0.40), herbal (+0.30)

10 cork (+0.72) pineapple (-0.83)


butter (-0.56) apple(0.35)
herbal 1a/1bEC-1118
canned corn
peach 2a/2b QA23
yeasty 4a/4bM
pineapple 5a/5b D47
citrus , 4a B

- Q4b
0.5
CM
B
• 5b
X
CO .-f 0.0
_ o^
Q.CM
tarn ^^
2b 2aB
u
c lb" 3
1a 5a B
-0.5

pineapple -1.0
citrus
apple
steely
cork -15 ■ 1 '
1
1
tropical fruit - 1.5 butter -1.0 -0.5 0.0 0.5 1.0 corked 1.5
tropical fruit musty
menthol Principal axis 1 vanilla
grapefruit steely
perfume/floral (48%) herbal
citrus vegetal

Fig. 3.10 Sample consensus plot for Free-Choice Profiling of the aroma of
Chardonnav (20 months) fermented by different veast strains

o
61

wines had a very corky aroma, not found in the second set. It was assumed
that one of the replicate of the D47 wines (D47a) underwent some changes in
the bottle, that did not occur in the other replicate. Because the replicate stood

out so much with the 'cork' aroma, which is a wine default, this descriptor had

a high loading and had to be introduced in the list of descriptors. It is shown


on the consensus plot that all the other samples were in the other perceptual
space (or close to) of dimension 1. The 3079 wines were removed from this

tasting due to the introduction of ML fermentation as another variable. The


EC-1118 and QA23 wines were closer than in the first tasting. They were both

described in PA1 with butter, tropical fruit, menthol, grapefruit, floral/perfume

and herbal, and in PA2 by pineapple, citrus, apple, steely and cork. This last
attribute would only be due to the presence of D47a, present on the negative

side of the PA1. The M wines seemed to have the same interrelationship as in

the 7 month tasting. Even though they did not replicate in the same quadrants

of PA1, they are nonetheless both located near the center of the PA1 space,

and might be more neutral in the descriptors of this dimension. In PA2, they

were separated from the other wines by herbal, canned corn, peach, yeasty,
pineapple and citrus descriptors. Analysis of variance however showed no

significant differences at the 0.05 level in aroma, and that is probably

explained by the poor replication of the D47 wine.


Table 3.9 summarizes the descriptors associated with each wine at both 7

and 20 months of age for the aroma of the Chardonnay. The young wines

were all described similarly with apple terms. QA23 and D47 were described
similarly with floral and herbal terms but differed in terms of the metallic,

reduced and coconut found in QA23. With aging, EC-1118 and QA23 were

described similarly, but differently from the M wines. The former wines had a
62

Table 3.9. Summary of the main aroma descriptors of Chardonnay fermented


by different yeast strains at 7 and 20 months of age

Yeast strain Descriptors


7 months 20 months

EC-1118 apple, honey, metallic, butter, tropical fruit,


fruity, coconut, reduced menthol, grapefruit,
perfume/floral, citrus,
herbal, pineapple, apple,
steely, cork

QA23 apple, floral, herbal, butter, tropical fruit,


lemon, tropical fruit, menthol, grapefruit,
yeasty, honey, metallic, perfume/floral, citrus,
fruity, coconut, reduced herbal, pineapple, apple,
steely, cork

M floral, apple, honey, herbal, canned corn, peach,


tropical fruit, yeasty, yeasty, pineapple, citrus
vanilla, fruity, citrus

D47 apple, floral, herbal, corked (0473), musty, vanilla,


lemon, tropical fruit, steely, herbal, vegetal, pear
honey, yeasty, vanilla,
fruity, citrus

more butter, fruity, citrus, menthol and floral predominant notes to describe

them, whereas M had more herbal, fruity, canned corn aroma.


In the flavor analysis of the young wine, Table 3.10 showed that a contrast

between PA1 that separated more on the basis of sweetness, yeasty/musty,


citrus, acid/tart, vanilla, fruity and some astringency whereas PA2 separated

also on the basis of citrus, acid/tart and fruity, but also in terms of bitterness

and body. The contrast did not seem to be well defined.


63

Table 3.10. Flavor attributes with positive or negative loadings > 0.30 for
Chardonnay at 7 months by each panelist.

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)

1 sweet (-0.74), honey (-0.41) apple (-0.48), flinty (-0.44)


yeasty (-0.31) yeasty (-0.44), green apple (-0.39)
2 green apple (+0.60), sweet (-0.50) grapefruit (+0.60), vanilla (-0.53)
soapy (+0.44), floral (-0.32) floral (+0.45)
vanilla (-0.30)

3 sweet (-0.73), acid (+0.43) heat (+0.62), leesy (+0.54)


heat (+0.38), astringent (+0.30) sweet (+0.44)

4 vanilla (-0.80), coconut (-0.39) smoke (-0.71), pear (+0.34)


green apple (+0.30) green apple (-0.30)
5 oxidized (-0.88) apple (-0.86)
musty (-0.33) musty (-0.48)

6 pear (-0.61), sweet (-0.50) coarse (+0.91)


grapefruit (+0.41), tart (+0.35) pear (+0.30)

7 citrus (+0.62), tart (+0.52) fruit (+0.81)


bitter (-0.53) tart (-0.56)

8 lime (+0.65), viscous (-0.45) citrus (-0.78)


orange peel (-0.42), nutmeg (+0.30) nutmeg (+0.40)

acidic (+0.58), oily/rich (-0.53) melon (+0.50), acidic (+0.45)


pineapple (-0.46) bitter (+0.37), sweet (-0.36)
pear (-0.35), oily/rich (+0.31)

10 sweet (-0.97) viscous (+0.83), pineapple


(+0.37), grapefruit (+0.31)

11 sour (+0.65), pineapple (+0.43) astringent (-0.61), sour (+0.46)


astringent (+0.40), yeast (+0.33) tin can (-0.34), lemon (-0.31)
sweet (-0.31)

12 sweet (-0.88) fruity (-0.65), lemon/grapefruit


body (+0.35) (-0.52), body (+0.42)
64

In the consensus plot for the flavor analysis of the young wines, (Fig. 3.11),
both dimensions explained 59% of the variability. The samples did not
replicate as well as the other tasting, shown by the position of the samples

spacially, as well as significant difference at the 0.05 level for the replication in

ANOVA of the PA scores of the samples. These results might be explained by

the young age of the wines where they are still harsh in texture, making it

more difficult for the panelists to taste the wines. The Chardonnay fermented
with EC-1118, QA23 and D47 replicated well in PA1, and all three samples

were associated with the acid/tart, citrus, pineapple, green apple and

astringent attributes. The M wines did not replicate well enough in any

dimension.

The flavor attributes of the older Chardonnay (Table 3.11) showed a

contrast between PA1 and PA2. The PA1 axis tended more toward sweetness
and acidity/sourness, as well as some yeasty/musty and citrus descriptors,
whereas PA2 describes the wines more in terms of fruit (pear, apple, melon),

tropical fruit such as pineapple, and citrus fruit such as lemon and grapefruit
and floral. Both axes described the wines with bitterness, body and astringent

descriptors as well.
The consensus plot (Fig. 3.12) of the flavor analysis showed both PA
accounting for 74% of the variability, and the samples replicated better than in

the first tasting shown by no significant difference in replication at the 0.05

level of the ANOVA, except again for the D47 wines. The EC-1118 and QA23
wines were again closely positioned. Both wines were described as citrus,

yeasty, vanilla, sweet and different in body in PA1, and astringent, straw, acid,

pine, bitter, citrus and yeasty in PA2. The M wines were well separated from
the other wines and were described in PA1 with musty, bitter, apple,
coarse 1.0
floral la/lb EC-1118
grapefruit 2a/2b QA23
sweet 3a/3b 3079
pear 4a/4b M
body 5a/5b D47
astringent 0 5
nutmeg 4a
CM 2a
3a
CO
'S _ 5a QQ
1a
_ S, 0.0
4b
'5 Ei3b
c „1b
B
2b
D

•0.5 5b
apple
yeasty
citrus
fruity
vanilla -1.0 ■ 1
smoky -1.0 honey -0.5 0.0 0.5 green apple 1.0
tart sweet acid
astringent floral Principal axis 1 tart
vanilla (40%) pineapple
pear citrus
viscous astringent
coconut nutmeg

Fig. 3.11 Sample consensus plot for Free-Choice Profiling of the flavor of
Chardonnay (7 months) fermented by different yeast strains

en
66

Table 3.11. Flavor attributes with positive or negative loadings > 0.30 for
Chardonnay at 20 months by each panelist

Panelist Attributes

Principal axis 1 Principal axis 2


(loading value) (loading value)

1 alcohol (+0.67) viscous (+0.74)


bitter (+0.60) acidic (-0.47)
sweet (-0.37) alcohol (-0.40)
2 musty (+0.78) honey (+0.77)
earthy (+0.61) pine (-0.48)
green apple (+0.37)

3 apple (+0.68) astringent (-0.76)


astringent (+0.56), body (+0.39) citrus (+0.48)
4 citrus (-0.86) sweet (+0.66)
earthy (+0.30) acid (-0.57)
citrus (-0.41)

5 bitter (+0.73) lime (+0.70)


acid (+0.62) sweet (+0.56)

6 vegetal (+0.74) fruity (+0.63)


sweet (-0.45) rich (+0.59)
acid (-0.32) bitter (+0.47)

7 bitter (+0.49) sour (+0.64)


tart (+0.49) bitter (-0.55)
body (-0.41) body (-0.32)
harsh (+0.37)

8 yeasty (-0.73) sweet (+0.63)


viscous (+0.50) acid (-0.49)
acid (+0.37) bitter (-0.39)
yeasty (-0.34)

9 musty (+0.76) apple (+0.75)


yeasty (+0.48) musty (-0.42)

10 bitter (+0.65) straw (-0.77)


vanilla (-0.54) acid (-0.41)
herbal (+0.35), sweet (-0.33) citrus (-0.32)
honey
apple 1a/1bEC-1118
viscous 2a/2b QA23
lime 4a/4bM
sweet 1.0-
a 5a/5b D47
fruity 4a
rich
sour
0.5- B 5b

<n 4b
(0 ^0 u
lb
2 co
5aH
o 0.5 - a la
c
2b

astringent 1.0 -
straw
acid
pine
bitter
citrus 1.5- 1 i —i 1 ■
yeasty -1.5 citrus -1.0 -0.5 0.0 0.5 1.0 musty 1.5
yeasty bitter
vanilla Principal axis 1 apple
sweet (43%) astringent
body acid
vegetal

Fig. 3.12 Sample consensus plot for Free-Choice Profiling of the flavor of
Chardonnay (20 months) fermented by different yeast strains
68

astringent, acid, vegetal and harsh terms. As for PA2, the descriptors used
were honey, apple, viscous, lime, sweet, fruity, rich and sour. No descriptors
could be associated with the D47 wines due to the poor replication.

Table 3.12 summarizes the descriptors associated with each wines at both

7 and 20 months of age for the flavor of the Chardonnay. In the young wines,
they were all described with the same terms. At 20 months of age, both EC-
1118 and QA23 were described similarly, as with the aroma results. The M

Table 3.12. Summary of the main flavor descriptors of Chardonnay fermented


by different yeast strains at 7 and 20 months of age

Yeast strain Descriptors


7 months 20 months

EC-1118 green apple, acid/tart, citrus, yeasty, vanilla,


pineapple, citrus, astringent, sweet, body, astringent,
nutmeg straw, acid, pine, bitter

QA23 green apple, acid/tart, citrus, yeasty, vanilla,


pineapple, citrus, astringent, sweet, body, astringent,
nutmeg straw, acid, pine, bitter

M green apple, acid/tart, musty, bitter, apple,


pineapple, citrus, astringent, astringent, acid, vegetal,
nutmeg harsh, honey, viscous,
lime, sweet, fruity, rich,
sour

D47 green apple, acid/tart, citrus, yeasty, vanilla


pineapple, citrus, astringent, sweet, body3
nutmeg

a Only for D47b, since D47b had a default in the wine and was an outlier
69

wine was described differently with more musty, bitter, apple, astringent and
acid terms, compared to citrus, yeasty, vanilla and sweet descriptors in the
other wines. The separation in the older Chardonnay wines, compared to the

Rieslings was not as dramatic. The Riesling, being a more aromatic variety

could have been more easily described than Chardonnay. Chardonnay could

be influenced more by processing variables such as aging on oak. In both

cases, the wines were easier to differentiate in the aroma, compared to the
flavor.

Fig. 3.13 through 3.16 illustrate the percentage consensus and percentage

within (residual) obtain from GPA in the evaluation of the younger and older

Chardonnay. In the aroma analysis of the young wines (Fig.3.13), there was

higher agreement among the panelists for the 3079, however this wine had

the ML fermentation, therefore was different due to an extra variable and


easier to differentiate. A high consensus, and low residual error were found in

the QA23, D47 and M wines, however, the EC-1118 had a low agreement

within the panelists, and it appeared that the panelists had difficulty
differentiating those wines, as shown by the position of the sample in Fig. 3.9

where they tend to be centered in the two-dimensional spaces. In the


evaluation of the aroma of older Chardonnay (Fig. 3.14), the samples were not

averaged together for the consensus and residual variations since some of
the values of the replicates were too different. Only the D47 wines did not

replicate well enough, and it is also seen in Fig. 3.10 where the wines are

separated in different ways. However, the others samples show a high

agreement between the panelists for their differentiation and description. In

the flavor analysis of the young wines (Fig. 3.15), the panelists had a high
agreement for all the wines, and the 3079 wines were easy to differentiate. In
□ % within
% consensus

c
o

CO
>
ra
o
o

EC-1118 QA23 3079 M D47


Chardonnay sample

Fig. 3.13 Percentages of consensus and within (residual) variation


distributed over the Chardonnay samples for the aroma at 7 months of age

o
30

□ % within
25- 0 % consensus

•= 20

CO
>
15-
CO
O

CO co CO m
^ eg CM
^ ^ < < a
T" ^ O a
o O
u tu
Chardonnay sample

Fig. 3.14 Percentages of consensus and within (residual) variation distributed


over the Chardonnay samples for the aroma at 20 months of age
EC-1118 QA23 3079 M D47
Chardonnay sample

Fig. 3.15 Percentages of consensus and within (residual) variation distributed


over the Chardonnay samples for the flavor at 7 months of age

-vl
30
D % within
% consensus
C 25
O

a 20

S 15
o
10-

CO 00 CO «
r- v- tM eg
r- T- < < Q
T- r-
o O
o o
LU LU Chardonnay sample

Fig. 3.16 Percentages of consensus and within (residual) variation distributed


over the Chardonnay samples for the flavor at 20 months of age

CO
74

the flavor analysis of the older wines (Fig. 3.16) the D47 replicates and the M
replicates were not similar. One of each of the M and D47 wines showed high

agreement within the panelists, except for one of the EC-1118 wines. Again,

the samples were not averaged together for the consensus and residual

variations since some of the values of the replicates were too different.

Pinot Noir Sensory Evaluation

Table 3.13 shows the aroma descriptors with loadings greater than +/- 0.30

used by each panelist for the young Pinot Noir wines. The aroma descriptors
of the new wines did not show a strong contrast between the two axes. Most

of the same descriptors were found to separate the wines in both dimensions,

such as cherry, pepper, musty/skunky/moldy/sweaty, smoky.

In the consensus analysis (Fig. 3.17), both axes explained 46% of the

variability. Dimensions 1 and 4 were used since they separated the wines the

best. The samples did not replicate well, as shown by their position in space,
as well as by significant differences between replication at the 0.05 level in the

Analysis of Variance of the PA scores of the sample. The BGY wines


replicated the best in both axes. The wine was described as cherry, herbal,
pepper, anise, mint, prune, sharp/VA (volatile acidity or acetic acid), alcohol

and musty. The RC212 wine replicated well in PA4, but remained in the

center of the dimension, giving it a more neutral character. The WSK wines
did not replicate well in both dimensions to be able to associate the samples

with any descriptors. The L2056, AMH and RA17 wines replicated well in
PA1. The AMH and L2056 wines tended toward berry, plum, musty, smoke

and spicy descriptors. The RA17 wines, even if they appeared to remain
75

Table 3.13. Aroma attributes with positive or negative loadings > 0.30 for Pinot
Noir at 7 months by each panelist.

Panelist Attributes
Principal axis 1 Principal axis 4
(loading value) (loading value)

1 moldy (+0.72), sharpA/A (+0.48) meaty (+0.51), berry (+0.42)


meaty (+0.33) cheesy (-0.39), sage/herbal (-0.39)
2 VA (+0.59), blackberry (-0.55) herbal (+0.62), VA (+0.55)
herbal (+0.39), prune (+0.33) prune (-0.34), pepper (-0.30)

3 black pepper (+0.51), heat (+0.50) rubber/reduced (+0.81), green


spicy (-0.41), berry (+0.39) olive, (-0.41), pencil lead (+0.31)
4 butter (+0.55), cherry (+0.49) tar (+0.62)
black pepper (+0.43), cooked meat fishy (+0.54)
(-0.39), smoke (-0.32)

5 resin (-0.90) chemical (+0.59), cherry (-0.52)


smoke (+0.36) smoke (-0.40), black cherry (-0.40)

6 earth (-0.67), musty (-0.51) anise (+0.54), pepper (+0.44)


anise (+0.35) musty (+0.39), skunky (+0.33)

7 cedar (-0.82), cherry (+0.39) mint (+0.71)


mint (+0.34) cherry (-0.68)

8 fish (-0.72) stemmy (+0.52), pepper (-0.48)


alcohol (+0.54) alcohol (+0.43), tobacco (+0.35)

9 sweaty (-0.46), tobacco (-0.45) pepper (+0.59), licorice (-0.45)


smoke (-0.36), musty (-0.35) leather (+0.41), sweaty (+0.34)
cherry (+0.33), pepper (+0.31) smoke (-0.31)

10 cherry (+0.67), metallic (-0.63) metallic (+0.63), cherry (+0.46)


berry (-0.35) herbal (+0.38), smoke (+0.33)

11 metallic (-0.67), floral (+0.51) cherry (+0.66), berry (+0.60)


plum (-0.37), berry (+0.33) alcohol (-0.31)

12 plum (-0.51), fruity (-0.49) reduced (+0.57), pungent


spice (-0.40), pungent (+0.37) (+0.48), berry (-0.43),
pepper (-0.43)
reduced -j Q
herbal la/lb WSK
mint 2a/2b AMH
metallic 3a/3b L2056
pepper 4a/4bRC212
cheesy 5a/5b RA17
musty 0.5 - 5a 6a/6b BGY
tar/tobacco
a lb

4a □
C? 0.0 4b
co <o

H
1a
2a Q
4 Q
o 6a
c &
3a a
-0.5--
prune
cherry
pepper

licorice
berry -1.0
-10 metallic -0.5 0.0 0.5 moldy i.o
berry Principal axis 1 black pepper
spicy alcohol/heat
smoke (30%) cherry
plum floral

Fig. 3.17 Sample consensus plot for Free-Choice Profiling of the aroma
of Pinot Noir (7 months) fermented by different yeast strains
77

toward the center of the dimension, would tend toward the more cherry,
pepper, herbal, anise and mint notes. No statistical differences were found by
ANOVA at the 0.05 level, which was not surprising considering the lack of

consistent replication.

After 20 months of aging, the panelists perceived 1he wines differently and

the assumption is that the wines have undergone some changes. Table 3.14

represents the significant attributes used by each panelist in FCP in aroma.


The aroma attributes between PA1 and PA2 showed contrast between berry
and more vegetal notes in PA1 and more fruit (especially cherry) and prune,

and herbal notes in PA2. Both axes separated the wines with pepper, spicy

and cola attributes.

In the consensus analysis (Fig. 3.18), for the aroma of the older wines, 55%

of the variability was explained by PA1 and PA2. The ANOVA of the PA
scores showed that the samples were better grouped between the replicates

(no significant difference at p < 0.05), and more separation was seen between

the samples in both dimensions (significant differences at p < 0.05). The

Fisher pairwise comparison shown differences more specifically between

RC212 and AMH, and between BGY and RA17 in PA1. In the PA2, the

differences were found between RC212 and AMH and L2056. The RA17
wines were described with berry, cherry, heat, vegetal, spice, fruity, pepper,
soy, green tea, prune/plum and herbal characters. Both AMH and RC212

wines were described by berry, cherry, heat (alcohol), vegetal, spice, fruity,
pepper, cheesy, tobacco, menthol, cola and woody characters. However,
since both wines are in different spaces in PA1, the characters of that

dimension were different for both wines. The L2056 and BGY wines were
both described with dish rag (musty), leather, herbal, cola, prune, musty,
78

Table 3.14. Aroma attributes with positive or negative loadings > 0.30 for Pi not
Noir at 20 months by each panelist

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)
1 berry (-0.82) tobacco (+0.53)
steely (+0.43) rhubarb (+0.52)
rhubarb (-0.33) vegetal (+0.40), berry (-0.38)
2 heat (-0.83) soy (-0.73)
tobacco (+0.34) aldehyde (-0.43),

3 vegetal (-0.65) herbal (-0.55)


cherry (+0.48), pungent (+0.40) floral (-0.54), plum (-0.44)

4 dish rag (+0.62) black pepper (+0.64)


leather (+0.46) vegetal (-0.59)
black pepper (-0.43) prune (-0.40)

5 vegetal (+0.46) pepper (+0.56)


herbal (+0.42) prune (-0.49)
prune (+0.40), cola (+0.36) cherry (+0.39), cola (+0.39)
chocolate (-0.33), cherry (-0.32)

6 cherry (-0.53) cherry cola (-0.54)


vegetal (+0.46) vegetal (+0.47)
cheese (+0.43) cheese (-0.42)
leather (+0.42) leather (-0.41)

7 cumin (+0.52) green tea (-0.72)


spice (-0.51), cork (+0.47) cumin (-0.51), spice (-0.32)

8 spice (-0.68) cheesy (+0.76)


musty (+0.54) berry (+0.47)
berry (-0.45) ethyl acetate (+0.39)

9 pepper (-0.52) menthol (+0.64)


berry (-0.49) green tea (+0.41)
herbal (-0.41), cherry (-0.32) woody (+0.33)

10 musty (+0.46) aldehyde (+0.55)


berry (-0.44) tea (-0.52), medicinal (+0.43)
cherry cola (-0.39) musty (-0.32)
aldehyde (+0.37), herbal (+0.34)
cheesy
black pepper10 1a/1b WSK
tobacco 2a/2b AMH
menthol 6a
3a/3b L2056
cherry 4a/4b RC212
cola
5a/5b RA17
berry 0.5 -
vegetal 6a/6b BGY
woodyN a 6b
.2 4b ,4a 2b,, 2a
« ^ o.o
_ ^s lb 3a 3b

c 5b a
a 1a
Q! -0.5 5a

soy
green tea
spice -1.0
prune/plum berry aisn rag
-1.0 -0.5 0.0 0.5 10
fruity cherry leather
herbal heat herbal
vegetal Principal axis 1 cola
spice (35%) prune
fruity musty

Fig. 3.18 Sample consensus plot for Free-Choice Profiling of the aroma
of Pinot Noir (20 months) fermented by different yeast strains

CD
80

cheesy, pepper, tobacco, menthol, cherry, berry, vegetal and woody. The
L2056 wines appeared to be more neutral in the descriptors of dimension 2.
The WSK replication was not close enough to assign the wines with specific
characters, however, due to the position of the replicates in PA1, dish rag,

leather, herbal, cola, prune and musty notes could be found in the wines. In

both young and older WSK wines, the WSKa wine was an outlier and it was

assumed that the profile of the wine should be more described in terms of the
WSKb. The WSKb wine was centered at the intersection of both axes in the
diagrams (Figs 3.17 and 3.18), and therefore difficult to differentiate from the

other wines.

When compared to the aroma results of the young wines, there seemed to

be an evolution in the wines that enabled the panelists to set them apart and

to replicate the samples as well. It would appear that Pinot Noir showed many
more difference when older than younger, which contradicts the belief that the
effect of yeast is less evident with aging ( Zoecklein et al., 1990). Table 3.15

represents the descriptors used in the time frame that we studied, and it shows

an evolution in the wine. The older wines were also easier to taste that the

young wines most likely because they were less 'rough' or 'harsh', and it was

shown by a better replication, as well as fewer "Not Determined" descriptors at


20 months. The two most important sensory dimensions (PA1 and 2) were
also better able to separate the older wines compared to the young wines

where PA1 and PA4 were needed to see a separation. At 7 months of age,

AMH and L2056 were described with the same terms, as were RC212 and

BGY except that BGY was further separated with berry, prune and licorice

descriptors. At 20 months of age, berry, cherry, heat, vegetal, spice, fruity and
pepper terms were used to describe AMH, L2056, RC212 and RA17, however,
81

Table 3.15. Summary of the main aroma descriptors of Pinot Noir fermented
by different yeast strains at 7 and 20 months of age

Yeast strain Descriptors


7 months 20 months

WSK NDa dish rag, leather, herbal,


cola, prune, musty, soy,
green tea, spice, fruity

AMH resin, metallic, berry, berry, cherry, heat,


spicy, smoke, plum, vegetal, spice, fruity,
tobacco pepper, cheesy,
tobacco, menthol, cola,
woody

L2056 resin, metallic, berry, berry, cherry, heat,


spicy, smoke, plum, vegetal, spice, fruity,
tobacco pepper,

RC212 volatile acidity, moldy, berry, cherry, heat,


black pepper, alcohol/heat, vegetal, spice, fruity,
cherry, floral, herbal, pepper
berry

RA17 ND berry, cherry, heat,


vegetal, spice, fruity,
pepper, soy, green tea,
prune, herbal

BGY volatile acidity, moldy, dish rag, leather, herbal,


black pepper, alcohol/heat, cola, prune, musty,
cherry, floral, herbal, black pepper, tobacco,
berry, prune, licorice menthol, cherry, cola,
berry, vegetal, woody
cheesy

a Not determined since both replicates stood in different location in both


dimensions

AMH and RA17 all had different profiles with some other descriptors that

separated them from the other wines. WSK and BGY also had similar terms
82

such as dish rag, leather, herbal, cola, etc. but were again separated further
by other descriptors. It also appeared that those two wines were less fruity
than the other four wines.

Table 3.16 shows the attributes in the flavor analysis of the new Pinot Noir

wines. Both axes appeared to explain the wines in terms of basic tastes such

as bitterness, acidity, sweetness and astringency. Both axes also separated

the wines with fruity, cherry, raspberry and jam attributes, as well as alcohol.

PA2 seemed to separate also in terms of the body of the wines. In general, no
strong contrast between the two axes was found. The consensus plot (Fig.

3.19) shows that 43% of the variability was explained by PA2, and PA3, again

chosen because they separated the wines better than the other dimensions.

The samples did not replicate well in both dimensions at the same time, but
good replications were seen in one or the other dimension. The WSK wines
replicated well in dimension 2, and tended towards astringent, bitter, acid and

berry attributes. The BGY wines also replicated well in dimension 2, but

tended towards more cherry, jammy, spice, alcohol and tobacco notes. The

AMH, L2056 and RA17 wines appeared to be centered on the PA3, making it

harder to associate any descriptors with the wines. The RC212 wines did not

replicate well enough to profile them. No statistical differences using ANOVA


were found at the 0.05 level, due again to poor replication.

The flavor attributes of the older wines are shown in Table 3.17. Both axes

defined the wines in terms of bitterness, cherry, vegetal, fruit and spice. A

contrast between viscosity in PA1 and astringent/harsh and earthy/musty in

PA2 was seen.


The consensus plot (Fig.3.20) of the older Pinot Noir flavor showed that
58% of the variability was explained by PA1 and PA2. Some wines replicated
83

Table 3.16. Flavor attributes with positive or negative loadings > 0.30 for Pinot
Noir at 7 months by each panelist.

Panelist Attributes
Principal axis 2 Principal axis 3
(loading value) (loading value)

1 moldy (+0.72) jam (+0.73), plum (-0.45)


cherry (+0.47) smoky (-0.34)
2 berry (-0.59), tanky (+0.59) bitter (+0.62), berry (+0.51)
pepper (-0.50) cherry (+0.36), tanky (+0.32)

3 bitter (+0.68), alcohol (+0.62) alcohol (-0.58), tar/tobacco


acid (+0.39) (+0.57), bitter (+0.53)
4 tart (-0.75), tobacco (-0.43) cherry (-0.76), butter (-0.45)
cherry (-0.34) tart (+0.31)

5 smoke (+0.63), alcohol (+0.50) smoke (+0.72), fruit (+0.42)


fruit (-0.43) tannin (-0.38), alcohol (-0.37)

6 earthy (+0.73), sweet (-0.40) sweet(-0.83)


raspberry (+0.38), VA (-0.38) raspberry (-0.48)

7 bitter (+0.89) spicy (-0.56), astringent (-0.45)


spicy (-0.30) bitter (-0.38), salty (-0.35)
8 thin (+0.75) spice (-0.65)
tobacco/tar (-0.57) tannin (-0.58)

9 alcohol (+0.73), metallic (-0.47) leather (-0.74), jam (-0.47)


jam (+0.36) bitter (+0.34)

10 lenght (+0.73), acid (+0.52) cherry (-0.67), astringent


astringent (-0.36) (+0.55), acid (+0.48)

11 musty (+0.66), bitter (+0.50) licorice (-0.69), yeasty (+0.38)


cherry (-0.34) cherry (+0.36)
dry (+0.33)

12 astringent (-0.55), body (-0.53) bitter (+0.63), fruity (-0.40)


cherry (-0.34), berry (-0.31) astringent (+0.38), berry (-0.31)
J3"1. 08
smoke la/lb WSK
bitter 2a/2b AMH
astringent} g. 1a B 3a/3b L2056
berry 4a/4bRC212
tar/tobacco 5a/5b RA17
cherry 0.4 6a/6b BGY
acid
n 0.2 6a a ,31.
1ba 2b
X 4bB
0.0
6b B 5b

5a 2a 3a1
c
Q.
-0.4
sweet
cherry 4a1
leather -0.6
licorice
spice
tannin -0.8 . . —i—
alcohol -0.8ta';t, -0.6 -0.4 -0.2 0.0 0.2 0.4 0€bltt ,5r, n^u
moldy/earthy
astringent astringenf Principal axis 2 smoke
raspberry {^f000 (23%) alcohol
cherry body
acid

Fig. 3.19 Sample consensus plot for Free-Choice Profiling of the flavoi
of Pinot Noir (7 months) fermented by different yeast strains

oo
4^
85

Table 3.17. Flavor attributes with positive or negative loadings > 0.30 for Pinot
Noir at 20 months by each panelist

Panelist Attributes
Principal axis 1 Principal axis 2
(loading value) (loading value)
1 astringent (-0.65) tanky (+0.75)
tanky (+0.53) astringent (+0.48)
bitter (-0.42) bitter (+0.33)
2 aldehyde (+0.85) bitter (+0.51)
bitter (-0.45) cherry (-0.48)
aldehyde (+0.45)
soy (+0.43), tea (+0.33)
3 cherry (-0.73) pepper (+0.54)
vegetal (-0.60) vegetal (-0.47)
astringent (+0.43)
smoky (+0.33), cherry (+0.31)
4 cherry (-0.95) cheesy (+0.78)
cheesy (+0.30) vegetal (+0.56)

5 bitter (+0.58) cola (-0.62)


viscous (-0.47) spice (-0.50)
cola (+0.44), cherry (-0.39) bitter (+0.46), cherry (-0.36)
6 bitter (+0.65) vegetal (+0.68)
fruity (-0.60), vegetal (+0.46) bitter (-0.60), rich (-0.37)
7 finish (-0.77) bitter (-0.89)
tart (+0.47), fruit (-0.34) fruit (-0.39)
8 berry (-0.68) cherry (-0.79)
earth (+0.50) body (+0.46)
spice (-0.35), cooked (+0.32) astringent (+0.33)
9 viscous (-0.53) yeasty (+0.67)
cherry (-0.52) earthy (+0.45)
fruity (-0.47), bitter (-035) musty (-0.39), harsh (+0.32)

10 bitter (-0.57) coffee (+0.51)


tart (+0.47) musty (-0.46)
coffee (+0.39) bitter (+0.39)
smoky (+0.33) spice (+0.37), tobacco (+0.35)
tanky i.o 1a/1b WSK
bitter 2a/2b AMH
yeasty/musty
astringent 3a/3b L2056
pepper 4a/4b RC212
vegetal 5a/5b RA17
0.5- D
B2b 6a/6b BGY
1a
■ 2a B
3b
CM
« B
3a
x
« ^^ 0.0
□ a lb
_ -^
S <» 4a Q
4b
o ^
c
-0.5 . ,6a Sb*
H
5b
cherry
fruity
cola
rich -1.0 ■ 1
spice -1.0 cherry -0.5 0.0 0.5 aldehyde 1
finish Principal axis 1 tart
viscous bitter
spicy (29%) smoky
astringent vegetal

Fig. 3.20 Sample consensus plot for Free-Choice Profiling of the flavor
of Pinot Noir (20 months) fermented by different yeast strains

CO
CD
87

well, whereas others did not, but still there was no significant differences
between the replicates at the 0.05 level. The BGY wines were well defined in
PA1 and PA2. Descriptors such as aldehyde (chemical), tart, bitter, smoky,

vegetal, cherry, fruity, cola, rich and spice were used to profile the BGY wines.
The RA17 wines appeared to have the same descriptors, however the

replication made it harder to determine the position of the wines. The L2056

wines were well replicated in both dimensions and were described with
aldehyde, tart, bitter, smoky, vegetal, tanky, yeasty/musty, astringent and
pepper. The RC212 wines would be described with cherry, finish, viscous,

spicy, astringent, fruit, bitter, cola and rich. Samples AMH could only be

described in PA2 with tanky, bitter, yeasty/musty, astringent, pepper and


vegetal. WSK wines could be described with the PA1 descriptors such as

aldehyde, tart, bitter, smoky and vegetal descriptors but did not replicate well.

Even without any statistical differences by the ANOVA (only barely since p =
0.052 for dim.2), it is evident that some wines are different since they were

described differently, and the poor replication of some wines might cause any

differences in the other wines to be hidden.

Table 3.18 represents the descriptors used in the flavor analysis of the

young and older Pinot Noir. At 7 months of age, only 3 wines were described,

and all three were described differently. At 20 months of age, L2056, RA17

and BGY were described similarly (RA17 and BGY identically). The AMH and

RC212 were described differently, where AMH was more bitter, yeasty and
astringent, whereas RC212 had a fruitier, more viscous finish, as well as some

spice, cola terms. Again, the wines show some differences depending on the

yeast strain that was used.


88

Table 3.18. Summary of the main flavor descriptors of Pinot Noir fermented by
different yeast strains at 7 and 20 months of age

Yeast strain Descriptors


7 months 20 months

WSK bitter, earthy, smoke, NDa


alcohol, body, acid, tanky,
cherry, jam, astringent,
berry, tar/tobacco

AMH ND tanky, bitter,


yeasty/musty,
astringent, pepper,
vegetal

L2056 ND aldehyde, tart, bitter,


smokey, vegetal, tanky,
yeasty/musty, astringent,
pepper

RC212 ND cherry, finish, viscous,


spicy, astringent, fruit,
bitter, cola, rich, spice

RA17 sweet, cherry, leather, aldehyde, tart, bitter,


licorice, spice, tannin, smokey, vegetal, cherry,
alcohol, astringent, fruity, cola, rich, spice
raspberry, plum, smokey

BGY tart, astringent, tobacco, aldehyde, tart, bitter,


berry, cherry smokey, vegetal, cherry,
fruity, cola, rich, spice

a Not determined since both replicates stood in different location in both


dimensions

Figs. 3.21 through 3.24 illustrate the percentage consensus and

percentage within (residual) obtained from GPA in the evaluation of Pinot


Noir. The samples were not averaged together for the consensus and
89

residual variations since some of the values of the replicates were too
different. In the aroma analysis of the young wines (Fig. 3.21), the panelists
did not agree on the differences of the wines as shown by the erratic total
variation, and some high residual, and very low consensus, for example in the

RA17 wines. In the aroma analysis of the older wines (Fig. 3.22), the
agreement between the panelists was higher than in the first tasting. Some of

the wines also replicated well for the agreement, such as L2056, RC212,

RA17. Still some wines such as WSK and BGY did not replicate well,

although it would appear that the consensus of the panelists was good. The

AMH wines had the lowest agreement, and a high residual associated with it.

The L2056 wines also had a high residual error associated with them.
However, both wines replicated well in the aroma space of PA1 and 2, but

were close to other wine samples, making them harder to differentiate. The

flavor results of the young wines (Fig. 3.23) indicate a higher agreement,
however, it is offset by the high residual, or greater error. One explanation for

the poor replication and agreement on the wines could by the winemaking

procedure. Since the grapes could not be pooled uniformly as juice like the
white varieties because of the fermentation on the skins of red grapes, the

grapes were distributed as uniformly as possible before crushing however, if

variation in the grape lots occurred, it would confuse the effect of the yeast
strains. In the flavor analysis of the older wines (Fig. 3.24), the total variation

was similar within the wines, except for the WSK replicates. The variation was

mostly situated between 5 and 10%, and the panelists appeared to have
agreed on the differentiation and profile of the older Pinot Noir wines

compared to the younger.


20

□ % within
% consensus

x X ID (0 CM CM 1^ r^- >- >■

CO CO s s U>
o
u>
o
T-
<N
T- ^ ^ C3 <■>
5 5 < CM CM O
CM
O
<
cc
<
cc
m m
-1 -1 cc CC
Pinot Noir sample

Fig. 3.21 Percentages of consensus and within (residual) variation distributed


over the Pinot Noir samples for the aroma at 7 months of age

CO
o
20
□ % within
M % consensus

15
c
o

(0
10
(0
o

<o
ir>
<£>
in
CM CM
T-
r^. r~ >- >
CO CO
T- r" r- (3 (3
5 5 o o CM
O
CM
o < < SD ffi
CM
_l
CM
_l cc oc
cc OC

Pinot Noir sample

Fig. 3.22 Percentages of consensus and within (residual) variation distributed


over the Pinot Noir samples for the aroma at 20 months of age

CD
□ % within
% consensus

c
o
.2
'^
>

r^ 1^ >- >•
CO S T- ^ C5 (3
o o CM CM < < ffl m
5 < CM CM O O CC cc
_l _l DC cc
Pinot Noir sample

Fig. 3.23 Percentages of consensus and within (residual) variation distributed


over the Pinot Noir samples for the flavor at 7 months of age

CD
c
o

CD
>
(0
o

X X to IO CM CM r- r^ >■ >-
CO CO £ 2 u>
o
in
o
y-
CM
T-
CM
^ am oCQ
5 5 < < CM o O
< <
_l
pj
_i cc CC
cc DC

Pinot Noir sample

Fig. 3.24 Percentages of consensus and within (residual) variation distributed


over the Pinot Noir samples for the flavor at 20 months of age

CD
CO
94

CONCLUSION

Free-Choice Profiling was a useful technique in determining differences in


Riesling, Chardonnay and Pinot Noir fermented by different S.cerevisiae

strains. It could be an advantageous tool to use in wine evaluation, especially

with expert tasters. This work showed that wines can be profiled, and that
differences or similarities can be found. The Riesling and Pinot Noir wines
showed that sensory differences occurred in both young and older wines,

making the yeast selection for fermentation an important parameter in

winemaking. In the Chardonnay wines, more similarities than differences

were seen among the yeast strains. Chardonnay is a less aromatic variety

and it could be that yeast strains would not be as important as other variables,
such as processing and aging on oak for example. Further studies with

multiple vintages and additional yeast strains would help answer the

questions.

LITERATURE CITED

Arnold, G.M., and A.A. Williams. The use of generalized Procrustes technique
in sensory analysis. In: Statistical Procedures in Food Research. Piggot, J.R.
(Ed.), pp 233-253. Elsevier Applied Science, London (1986).

Dijksterhuis, G.B., and S. van Buuren. Procrustes-PC v2.2 Manual. OP&P


Utrecht Holland (1991)

Gains, N., Krzanowski, W.J. and D.M.H. Thompson. A comparison of variable


reduction techniques in an attitudinal investigation of meat products. J.
Sensory Stud. 3: 37-48 (1988)
95

Gains, N., and D.M.H. Thompson. Sensory profiling of canned lager beers in
their own homes. Food Quality Preference. 2:39-47 (1990).

Guy, C, Piggott, J.R., and S. Marie. Consumer profiling of Scotch whisky.


Food Quality Preference. 2: 69-73 (1989).

Lawless, H.T. Flavor description of white wine by "expert" and nonexpert wine
consumers. J. Food Sci. 49: 120-123 (1984).

MacFie, H.J., Bratchell, N., Greenhoff, K, and L.V. Vallis. Designs to balance
the effect of order of presentation and first-order carry-over effects in Hall tests.
J. Sensory Studies. 4: 129-148 (1989).

McEwan, J.A., Colwell, J.S., and D.M. Thompson. The application of two free-
choice profiling methods to investigate the sensory characteristics of
chocolates. J. Sensory Stud. 3: 271-273. (1989).

Noble, A.C. Analysis of wine sensory properties. In: Wine Analysis. Linskens,
and Jackson (Eds.), pp 9-28. Springer-Verlag, New York (1988).

Noble, A.C, Arnold, R.A., Buechsenstein, J., Leach, E.J., Schmidt, J.O., and
P.M. Stern. 1987. Modification of a standardized system of wine aroma
terminology. Am. J. Enol. Vitic. 38(2): 143-146.

Rubico, S.M. Perceptual characteristics of selected acidulants by different


sensory and multivariate methods. Thesis, Oregon State University (1993).

Ryan, B.E., Joiner, B.L, and T.A. Ryan, Jr. Minitab Handbook. 2nd Ed. PWS-
Kent Publishing (1985).

Williams, A.A., and G.M. Arnold. A comparison of the aromas of six coffees
characterized by conventional profiling, free-choice profiling and similarly
scaling methods. J. Sci. Food Agric. 36: 204-214 (1985).

Williams, A.A., and S.P. Langron. The use of free-choice profiling for the
evaluation of commercial ports. J. Sci. Food Agric. 35: 558-569 (1984).
96

Zoecklein, B.W., Fugelsang, K.C., Gump, B.H., and F.S. Nury. Production wine
analysis, pp. 475. AVI Book. Van Norstrand Reinhold, New York (1990).
97

CHAPTER 4. CHEMICAL COMPOSITION OF RIESLING,


CHARDONNAY AND PINOT NOIR AND VOLATILE COMPOSITION
OF RIESLING AND CHARDONNAY FERMENTED WITH DIFFERENT
COMMERCIAL STRAINS OF SACCHAROMYCES CEREVISIAE

Ann Dumont1, Mina R. McDaniel1, Max Deinzer2 and Barney T. Watson1

1.Department of Food Science and Technology, Wiegand Hall 100


2. Department of Agricultural Chemistry,
Oregon State University, Corvallis, OR 97331

Presented in Parts at the American Society of Enology and Viticulture,


Anaheim, CA, June 28-30,1994
And Eastern American Society of Enology and Viticulture,
Cincinnati, OH, July 15-17 1994.
98

ABSTRACT

Compositional analysis were done on 1992 Riesling, Chardonnay and

Pinot Noir fermented with different commercial preparations of


Saccharomyces cerevisiae in order to assess if any differences were present.

The results showed differences in titratable acidity and residual sugar in all

three varieties. Differences were also seen in volatile acidity in both Riesling
and Chardonnay. The Pinot Noir wines were also analyzed for color

differences but no significant difference (p < 0.05) were found.

Volatile composition by GC/MS in two selected Riesling and Chardonnay


wines also revealed differences in composition. In the Riesling wines, 24

compounds out of 39 were positively identified in the wines fermented with

Lalvin Zymaflore VL1. In the wines fermented with Lalvin Epernay 2 CEG, 21

compounds out of 34 were positively identified. Similarities were seen

between the two wines in carboxylic acids. However, differences in the range

of relative concentration were seen in the alcohols and ester composition, as


well as some terpenes, lactones and sulfur compounds. The differences were

mainly quantitative (relative concentration), rather than qualitative. In the

Chardonnay wines, 21 compounds out of 32 were positively identified in the

Lalvin EC-1118 Prise de Mousse, whereas 16 compounds out of 23 were

positively identified in the Lalvin D47 wines. Differences in the range of

relative concentration were seen in the carboxylic acids, alcohols and esters.
Most often, they were found in higher relative concentration in the EC-1118

wines.
99

INTRODUCTION

The mixture of volatile and non-volatile compounds found in wines


contribute significantly to the aroma and flavor (Rapp and Mandery, 1986).

The chemical composition will influence how the wine is perceived in the

flavor-by-mouth, and to a lesser extent the aroma. Extraction and


identification of aroma compounds has been done in some studies on both

Riesling (Holloway and Subden, 1991) and Chardonnay (Sefton et al., 1994).

The Riesling is a more aromatic variety and contains more terpenes and their
precursors. Chardonnay, on the other hand will have high concentration of

compounds such as diacetyl and damacesnone (Sefton et al., 1993). Some

work has also been done to compare volatile production in wines fermented
by different Saccharomyces strains as well as other species of yeasts

(Holloway and Subden, 1991; Lavigne et al., 1992; Longo et al., 1992;

Delcroix et al., 1994). It was generally found that the wines would differ

quantitatively but not qualitatively. It was found that some strains produced

different amounts of acetaldehyde and acetic acid (Millan and Ortega, 1988;

Longo et al., 1992), esters, (Soles et al., 1982; Rosi et al., 1989), higher

alcohols (Ciani et al., 1991; Cabrera et al,. 1988), phenols (Chatonnet et al.,

1993; Grando et al., 1993) and sulfur-containing compounds (Lavigne et al.,

1992).
The differences in wines fermented with different strains of S.cerevisiae

were found to be associated with different enzymatic activities. Esterases

(Rosi et al., 1989; Suomalainen, 1981) for the hydrolysis and synthesis of
esters are found to vary with different commercial yeasts. This was of

considerable interest since esters contribute positively to the sensory


100

attributes of wines. The enzyme (E) cinnamate decarboxylyase varied in


experiments with commercial yeast strains (Chatonnet et al., 1993; Grando et
al., 1993). This enzyme will convert phenolic acids in the musts into

vinylphenols. Vinylphenols are described as medicinal, spicy odors which are

easily identifiable in wines, and not always desirable-depending on the variety

and the amount present.


The different metabolic activities of different strains of S.cerevisiae could

lead to different volatile composition. Little work has been done between the
relationship of the volatile content and the sensory properties. Miranda-Lopez

(1990) studied Pinot Noir volatile compounds compared to GC-sniffing

chromatograms. Akhtar et al. (1985) and Etievant et al. (1983) also studied
the correlation between composition and sensory properties. However, it is

considered a black box since it difficult to evaluate the contribution of each

compound, its threshold in wine, and the interactions with other volatile
compounds and non-volatile compounds.

The objective of this study was to determine the chemical composition of

Oregon varieties including Riesling, Chardonnay and Pinot Noir fermented

with different commercial yeast strains. Two selected wines of both Riesling

and Chardonnay fermented by different yeast strains were then analyzed for
volatile compounds which may affect sensory perception.
101

MATERIALS AND METHODS

Wine Preparation

Three varieties of grapes were used for the research. Chardonnay and

Riesling were harvested from the Lewis Brown Farm, Corvallis, Oregon on
October 1, 1992. The grapes were crushed and destemmed, and pressed
using a Willmes bladder press. The must was pooled and 50 ppm SO2 was

added using 1% potassium metabisulfite solution. The must was settled

overnight at 40C, racked from the solids and separated into 10 lots (2 lots per

yeast) of 23.2 L each in glass fermentors, equipped with a fermentation lock.

The chemical analysis for the juice for Chardonnay was 22.2 0Brix, pH 3.08
with titratable acidity of 7.05 g/L. The Riesling chemical analysis was 21.9
0
Brix, pH of 2.98 and a titratable acidity of 7.15 g/L. The juice was inoculated

with 20g/hL of active dry yeast.

The five yeasts used for Chardonnay were Saccharomyces cerevisiae


strains: Lalvin EC-1118 Prise de Mousse from Epernay, France; Lalvin QA23

from Portugal; Lalvin Bourgoblanc 3079 from Bureau International du Vin de


Bourgogne (BIVB), France; Lalvin M Montrachet, an isolate obtained from the

University of California, Davis; and Enoferm ICV D47 from C6tes-du-Rhone.

The five yeasts used for Riesling were Saccharomyces cerevisiae strains:
Enoferm Simi White, an isolate from the Pasteur Institute, France; Lalvin CS-2

from the Alsace region, France; Lalvin Zymaflore VL-1 from Bordeaux, France;

Lalvin K1 (V-1116) from Montpellier, France and Uvaferm Epernay 2 CEG


from the Rheingau region in Germany.
102

The white wines were fermented at 18-240C. After fermentation was


completed, the wines were racked from the solids and 1% potassium
metabisulfite was added to obtain 20 ppm of free SO2. The wines were cold

stabilized for 1 month at 40C and fined with 0.36 g of bentonite per liter. After

settling, the wines were racked from the bentonite and 1% potassium
metabisulfite was added to obtain 20 ppm of free SO2. The wines were

bottled unfiltered in 750 mL bottles and stored at 40C.

The Pinot Noir grapes were harvested October 14, 1992 from Beaver Creek

Vineyard, Corvallis, Oregon. Thirty four kg/lots (2 lots per yeast) were crushed
and destemmed and 50 ppm of SO2 was added using 1% potassium

metabisulfite. The chemical analysis of the must was 24.3 0Brix, pH 3.12 and
a titratable acidity of 6.71 g/L. The must was inoculated 2 hours after SO2

addition with 20 g/hL of the following six active dry yeast; Lalvin Wadenswil

WSK-27 from Switzerland; Enoferm Assmannshausen, from the Geisenheim

Institute in Germany; Lalvin 2056 from C6tes-du-Rhone, France and from

Bourgogne, France, the Lalvin Bourgorouge 212, Bourgorouge RA17 and

Enoferm Burgundy. The wines were fermented at 25-30oC on the skins for 14
days prior to pressing. The wines were punched down twice daily. Pressing

was done with a Willmes bladder press, and the wines were settled, then

racked from the solids. The new wines were inoculated with 0.006 g per liter

of Leuconostoc oenos (Lalvin OSU strains). After completion of malo-lactic

fermentation, the wines were racked and 1% potassium metabisulfite solution


was added to obtain 20 ppm of free SO2 and the acidity was adjusted to pH

3.65 with the addition of tartaric acid. Before bottling the wines were cold

stabilized, then racked and 1% potassium metabisulfite was added to obtain


20 ppm of free S02- At bottling, the wines were rough filtered with Seitz 200
103

filter pads (20 x 20 cm) first washed with 3% citric acid solution, then rinsed
with distilled water. Then 1% potassium metabisulfite was added to obtain 20
ppm of free SO2 before bottling in 750 mL wine bottles. The bottled wines

were stored at 40C.

Wine Analysis

The fermentation rate was monitored daily in the Pinot Noir by measuring
0
Brix (g sugar/100 g of water) using a hydrometer and every second day for

Riesling and Chardonnay. The results are reported as means of the Brix of
the replicate lots of each wine.

Alcohol was measured by ebulliometer (Dujardin-Salleron, Paris, France)

as described by Amerine and Ough (1980) by measuring the boiling point of

the wines (Zoecklein et al., 1990).

The Rebelein method was used (Zoecklein et al., 1990) to measure

reducing sugars as glucose and fructose. In the Rebelein procedure, excess


copper remaining after reaction with reducing sugars (glucose and fructose)

reacts to release iodine. The iodine is then titrated with sodium thiosulfate to a

white endpoint (Watson et al., 1980).

Volatile acidity (VA) is measured as acetic acid. It was done by steam

distillation of the volatile acid components followed by titration with 0.01667 N

NaOH using a 1% phenolphthalein indicator. The titer was then corrected for
error due to SO2 by titrating with 0.020 N iodine using a 1% starch indicator in

a solution previously rendered acidic with 1 + 3 H2SO4 (Watson et al., 1980).

Total phenols were measured in the Pinot Noir wines by the Folin-
Ciocalteu method described by Zoecklein et al. (1990). This method uses
104

gallic acid as the standard compound for reference, and therefore the results
are expressed as gallic acid equivalents (Zoecklein et al., 1990).
Anthocyanin content was obtained by measuring absorbance at 520 nm at
pH < 1 with a 1% extinction coefficient E1%1 cm = 380 nm (Singleton, 1980).

Titratable acidity was measured as tartaric acid as described by Amerine

and Ough (1980) and Watson et al. (1980). The pH was measured using a
glass electrode in the wine with a Corning pH Meter 125.

Malate was measured enzymatically in the form of L-malic acid. The assay

measures the formation of reduced NADH after the oxidation of L-malate by


the enzyme L-malate dehydrogenase (Zoecklein et al., 1990). The

absorbance values were measured using a UV Spectrophotometer.

The color intensity was measured as absorbance of the wine with no


dilution at 420 nm plus 520 nm at wine pH using a 1 mm path length cuvette.

Duplicate measurements of wine components were analyzed for

differences using an Analysis of Variance (ANOVA) on Minitab©. The means


of the duplicates were used for ANOVA, with n=5 or n=6 for

Riesling/Chardonnay and Pinot Noir respectively. If significant differences

were found (p < 0.05), means were analyzed by Least Significant Difference
using Fisher's pairwise comparison.

Extraction Of Volatile Compounds

The Riesling fermented with Zymaflore VL-1 and Epernay 2 CEG, and the

Chardonnay fermented with ICV D47 and EC-1118 Prise de Mousse were

chosen for extraction of volatile compounds. Duplicate lots of each wine were

pooled and bottled in 750 mL wine bottles. The wines were kept at 40C until
105

extraction. Two aliquots of 375 mL each were taken from each bottle and
extracted. The extraction procedure was done using Freon 113 (1,1,2-
trichloro-1,2,2-trifluoroethane), certified grade, obtained from Fisher Scientific,

Pittsburgh, PA. The wine and the Freon 113 were measured and mixed in

equal volumes of 375 mL and the head space was flushed with nitrogen gas
for 3 minutes. The mixture was stirred gently with a magnetic stirring bar of

approximately 7.5 cm long at 30 to 40 rpm for exactly 30 minutes each. The

two phases were transferred to a 1 L separatory funnel, and allowed to


separate for 30 minutes. The head space of the mixture was gently flushed

with nitrogen gas for 3 minutes. The Freon layer at the bottom of the

separatory funnel was dried using anhydrous sodium sulfate. The Freon layer
was removed by opening the bottom valve of the separatory funnel directly

onto a fluted filter paper (Whatman #4, 12.5 cm) containing 10 g of sodium

sulfate. The dried Freon was collected directly into a 500 mL Kuderna-Danish
apparatus fitted with a 5 mL collector vessel flushed with nitrogen. It was then

concentrated in the Kuderna-Danish apparatus at 570C to nearly 5 mL. The

three ball Snyder condenser was replaced by a smaller condenser, and the

mixture was concentrated to 1 mL. The extract was transferred quickly to a 2

mL vial with Teflon cap, and 0.5 mL of undecane and 5 mL of tetradecane

(both certified grade, Fisher Scientific, Pittsburgh, PA) were added as internal
standards. The mixture was flushed with nitrogen and kept at -20oC until

needed (Miranda-Lopez, 1990; Acree, 1993, personal communication). Two

injections per extraction were done on GC/MS.


106

Gas Chromatography-Mass Spectrometry

Mass spectroscopy was performed using a Finnigan model 4023

quadrupole mass spectrometer (MS) upgraded with a model 4500 source for

low resolution electron impact (El) mass spectral analysis and a Varian model

3400 gas chromatograph supplied by the Environmental Health Sciences


Center at Oregon State University. A Galaxy 2000 data system (LGG Co., San

Jose, CA) was used to control the mass spectrometer. The spectra from El
were obtained at an electron energy of 70 eV at sources temperatures of

140oC. The samples were injected in a 1 mL volume onto the column using a

splitless injection, at a temperature of 225 0C, and a transfer temperature of


230oC. The initial column temperature was 50oC, was held for 5 minutes, and

increased at a rate of 40C/minute to 200 0C and held for 30 minutes at this

temperature. An Alltech (Deerfield, II) capillary column 30 M in length, 0.25


mm id., bonded with 0.25 mm Carbowax was used and operated at injector

pressure of 9 psi. The carrier gas was purified oxygen-free helium. Two

injections per extraction of each wine were done (4 injections total per wine).
The retention time of a series of n-parrafins standards of C5-C30 (Alltech

Associates Inc., Deerfield, IL) was used to calculate the Kovats (retention)

indices (Kovats et al., 1958) for the volatile compounds. The relative
concentrations were measured by dividing the area of each chromatographic

peak by the area of internal standard peak, after normalizing for the amount of

internal standard added to the extract (Perez-Gonzalez et al., 1993; Akhtar et

al., 1985; Holloway and Subden, 1991). This method allowed for the relative
comparison of the volatile compounds present in the wines but did not allow

the calculation of the actual concentration by standard curves the pure


107

compounds. The results are shown as the range of relative concentration


found for each volatile compounds. No statistical tests were used to evaluate
differences, and the relative concentration was evaluated as being greater,
smaller or similar between the samples.

The identification was done by comparison of the experimental spectra with


the published spectra found in Jennings and Shibamoto (1980), Libbey
(1990) and the NIST spectra library. The comparison of the Kovats indices of

the compounds was done with those of the authentic compounds found in
Jennings and Shibamoto (1980) and Libbey (1990). The identification was

considered confirmatory if the Kovats indices and mass spectra of the wine

compound were similar (70-100% match) to the authentic compounds. The


compounds tentatively identified had a 50% match with the GC-MS spectra
and a match with the Kovats Indices.

RESULTS AND DISCUSSION

Fermentation Rates

In the Riesling (Fig. 4.1) there were differences in fermentation rates. The

fastest fermenting strains were CS-2 and K1, which fermented to -10Brix in 37
and 42 days, respectively. The other three strains fermented longer; Simi

White required 60 days, and VL-1 and CEG 65 days. CEG was reported

earlier by Avedovech et al. (1992) to take the longest time to complete


fermentation in an Oregon Chardonnay must.
108

30

■H— SIMI WHITE

■+— CS-2

20

><
DC
CD
10 -

-10' —r-
20 40 60 80

TIME (DAYS)

Fig. 4.1 Fermentation rate of Riesling


with different yeast strains
109

In the Chardonnay (Fig. 4.2) the EC-1118, QA23 and D47 strains took 47
days to complete fermentation. The M strain took 95 days, and the 3079
strain, 110 days. The 3079 strain stopped fermenting at 0.5oBrix.
The Pinot Noir wines were all fermented on the skin for the same period of
time, 12 days. They all completed fermentation within that time, although

some strains were faster to reach -10Brix (Fig. 4.3). The 2056 strain was the
fastest, followed by the 212, RA17, BGY (all three from the Bourgogne region

in France), followed by the AMH and WSK strains. The WSK strain was the

slowest and was left on the skins an extra day to reach -10Brix level.

Composition Analysis

In the Riesling wines (Table 4.1), no statistical differences ( p < 0.05)were

found in pH and % alcohol. The pH values were 2.88 to 2.90, which was

similar to the initial pH of the must of 2.98. The mean % alcohol of the wines
was 13.6-13.7 %. This is in accordance with some results from Cabrera et al.

(1988) where they found similar ethanol production in four yeast strains of

S.cerevisiae.
Statistical differences (p < 0.05) were found in titratable acidity (TA),

residual sugar (RS), volatile acidity (VA), and malate content. The mean TA

varied from 6.85 to 7.30 g/L. All the wines were statistically different except for
the Simi White and CEG wines with 7.27 and 7.30 g/L, respectively, which

were found to be similar using a Fisher pairwise comparison for the Least

Significant Difference (LSD). The wines were statistically different (p < 0.05)

in RS content varying from 2.50 to 7.80 g/L. Only Simi White and CS-2 were
not significantly different from each other by LSD with 3.00 and 2.50 g/L,
110

><

40 60 80 120
TIME (DAYS)

Fig. 4.2 Fermentation rate of Chardonnay


with different yeast strains
111

30

-Q— WSK

■+— AMH

20 "

RA17

10
CD

o-

-10 —f—

6 8 10 12 14
Time (days)

Fig. 4.3 Fermentation rate of Pinot Noir


with different yeast strains
Table 4.1 ANOVA of the means of the composition analysis of Riesling fermented with different
yeast strains

Yeast strain PH Alcohol TA1 RS2 VA3 Malate


% vol. g/L g/L g/L g/L

Simi White 2.88a 13.7a 7.30d 3.00a 0.33a 1.35a

CS-2 2.87a 13.6a 6.85a 2.50a 0.40^ 1.44b

VL-1 2.87a 13.6a 7.17C 7.80d 0.32a 1.70d

K1 2.88a 13.7a 7.00b 4.10b 0.45^ 1.66d

CEG 2.90a 13.6a 7.27d 5.80C 0.42b 1.58C

LSD5 NA* NA 0.10 0.10 0.005 0.07

The means with the same letters in the same column are not significantly different (p < 0.05) by ANOVA
Sample size = 5
1. Titratable acidity
2. Residual sugar
3. Volatile acidity (as acetic acid)
4. Not applicable
5. Least Significant Difference with Fisher pairwise comparison

rv>
113

respectively. Some statistical differences (p < 0.05) were found in VA. Simi

White and VL-1 were both similar with 0.32 and 0.33 g/L, but different from
CS-2, K1 and CEG with 0.40, 0.45 and 0.42 g/L, respectively. Cabrera et al.
(1988) also found that different yeast strains can produce different levels of

volatile acidity. The level of VA (as acetic acid) were all well below the legal
limit of VA in commercial wines (Watson et al., 1980). Statistical differences (p

< 0.05) were also seen in malate content. The malate varied from 1.35 to 1.70

g/L. VL-1 and K1 were not statistically different from one anther with 1.70 and

1.6 g/L, respectively, using LSD.


In the Chardonnay analysis (Table 4.2), statistical differences (p < 0.05)
were found in pH, TA, RS, VA and malate. The wines fermented with 3079

underwent spontaneous malo-lactic fermentation (MLF) shown by the higher

pH (3.25), lower TA (4.30 g/L) and much lower malate content (0.15 g/L) than

wines fermented with the other strains. MLF occurred most likely because the

3079 strain took the longest time to ferment and the wines were in contact with
yeast lees for the longest period of time, which can promote spontaneous
malo-lactic fermentation. The alcohol levels ranged from 13.8 to 14.2 % and

no differences were found among the yeast strains. Statistical differences (p <

0.05) were found among all the yeast strain in TA and RS. The TA varied from

4.30 (for 3079) to 7.25 g/L for the D47 wines. The RS varied from 1.60 for the

QA23 wines to 6.20 g/L for the M wines. The fermentation rates correlated
with the RS in that the strains that fermented the fastest (EC-1118, QA23, D47)
had the lowest residual sugar and the slowest (3079, M) had the highest RS.

The volatile acidity varied from 0.24 to 0.45 g/L and was statistically different

(p < 0.05) except between EC-118 and 3079, and between QA23 and M using

LSD. The D47 wines had the lowest VA with 0.24 g/L. The malate content
Table 4.2 ANOVA of the means of the composition analysis of Chardonnay fermented with different
yeast strains

Yeast strain PH Alcohol TAl RS2 VA3 Malate


% vol. g/L g/L g/L g/L

EC-1118 3.13b 14.2a 7.03d 2.60b 0.44c 3.1 yd

QA23 3.13b 14.2a 6.90C 1.60a 0.34b 2.98C

3079 3.25C 13.8a 4.30a 5.20d 0.45c 0.15a

M 3.08a 13.9a 6.45b 6.20e 0.34b 2.70b

D47 3.08^ 14.2a 7.25© 2.90C 0.24a 3.29d

LSD5 0.02 NA4 0.10 0.058 0.003 0.12 ,

The means with the same letters in the same column are not significantly different (p < 0.05) by ANOVA
Sample size = 5
1. Titratable acidity
2. Residual sugar
3. Volatile acidity (as acetic acid)
4. Not Applicable
5. Least Significant Difference with Fisher pairwise comparison
115

varied from 2.70 to 3.29 g/L and was statistically different (p < 0.05) with and
without the 3079 wines. The 3079 had the lowest, as expected after malo-
lactic fermentation, with 0.15 g/L. The EC-1118 and D47 were not different

from one another using LSD. Some wine yeasts are able to metabolize

malate during fermentation but generally in small amounts (3 to 45%

depending on the strain (Radler, 1993)).

Along with all the other composition analyses, the Pinot Noir wines (Table
4.3) were also analyzed for anthocyanins, total phenols and color intensity.

No statistical differences (p < 0.05) were seen in pH, alcohol or VA. The pH

values ranged from 3.82 to 3.91, the alcohol from 13.1 to 13.4 % and the VA

from 0.28 to 0.32 g/L. Statistical differences (p < 0.05) were seen in TA, but

not between 212 and WSK wines, between 2056 and RA17 wines, between
AMH and BGY wines, nor between WSK and RA17 wines using LSD. The
mean TA varied from 4.87 to 5.56 g/L. RS was also statistically different at p <

0.05 except between WSK, 2056 and 212 using LSD. The RS varied from

3.68 to 5.05 g/L. No statistical differences were also seen at p < 0.05 in

anthocyanin content, phenol content and color intensity It is not known

whether or not the yeast strain will vary in the ability to extract color or if once

extracted, color is less stable due to unknown interactions related to yeast


metabolism. For example, if present, B-glucosidase activity could degrade
pigments after extraction to less stable aglycon forms.

GC/MS Analysis Of Riesling And Chardonnay

Over 500 aroma substances have been identified in wines by GC-MS


(Sefton et al., 1994; Schreir, 1979; Nykanen and Suomalainen, 1983;
Table 4.3 ANOVA of the means of the composition analysis of Pinot Noir fermented with different
yeast strains

Yeast strain PH Alcohol TA1 RS2 VA3 Anthocyanin4 Phenol5 Color 6


% vol. g/L g/L g/L mg/L g/L Intensity

WSK 3.84a 13.1a 5.31 be 4.82C 0.30a 182a 2.10a 3.43a

AMH 3.87a 13.4a 4.87a 5.05d 0.30a 188a 2.24a 3.92a

L2056 3.85a 13.2a 5.56c 4.69c 0.29a 22ia 2.39a 5.0ia

212 3.86a 13.2a 5.25b 4.70C 0.32a 197a 2.36a 4.19a

RA17 3.82a 13.2a 5.45bc 3.68b 0.30a 164a 1.84a 3.23a

BGY 3.9ia 13.4a 4.87a 3.36a 0.28a 159a 1.95a 3.34a

LSD8 NA? NA 0.26 0.08 NA NA NA NA

The means with the same letter in the same column are not significantly different at the 0.05 level by ANOVA
Sample size = 6
1. Titratable acidity 5. Gallic acid equivalents
2. Residual sugar 6. Absorbance 420 + 520 nm, 1 mm path lenght, wine pH
3. Volatile acidity (as acetic acid) 7. Not Applicable
4. Absorbance at 520 nm, pH < 1, E10/o1 cm = 380 8. Least Significant Difference with Fisher pairwise
comparison

O)
117

Montedoro and Bertuccioli, 1986; Nykanen, 1986). Many authors have


expressed the importance in the selection of a proper yeast strain of

S.cerevisiae for the production of the volatile aroma compounds. Mateo et al.
(1992) determined potential differences in volatile production in order to
choose the best strain as an inoculum in an aromatic wine. Suomalainen and

Lehtonen (1979) stated that most of the aroma compounds were formed

during fermentation, and under the influence of yeasts. Muller et al. (1993)

recognized that the yeast strains were able to produce flavor constituents

during the fermentation, and also that the amount and the nature of the volatile
compounds depended not only on the fermentation temperature, the grape
used, the chemical and biochemical changes prior to fermentation, but also on

the type of yeast used.

Fig. 4.4 and 4.5 represent the GC/MS chromatograms of the Riesling

fermented with Zymaflore VL-1 and Epernay 2 CEG respectively. The

chromatograms are both similar and 39 compounds for VL-1 and 34

compounds for CEG were recorded by GC/MS. Kovats Indices coupled with
the mass spectra were used to identify the compounds.

In the VL-1 extracts, 24 compounds out of the 39 were positively identified


by matching Kovats Indices and mass spectra (70-100% match). Three
compounds were tentatively identified with the correct Kovats Indices and a

50-70% correct match in the mass spectra. Twelve compounds remained


unknown in the VL-1 samples. In the CEG extract, 21 compounds out of the
34 were positively identified, and three were tentatively identified and 10

remained unknown.

Fig. 4.6 and 4.7 represent the GC chromatograms of the extracted samples

of the EC-1118 and D47 Chardonnay wines. In the EC-1118 extract, 32


i 3-methyl-1-butanol 21 phenyl ethyl acetate
100 2 ethyl hexanoate 22 ethyl dodecanoate
3 hexyl acetate 24 hotrienol
25
std 4 n-pentanol 25 R-phenyl ethanol
29 5 . ethyl lactate 26 compound d
6 1-hexanol 27 compound e
7 acetic acid 28 3-phenyl propanol
75- 8 n-nonan-2-ol 29 octanoic acid
33 9 linalool 30 g-decalactone
10 2,2-butanediol 31 compound f
11 g-butyrolactone 32 compound g
12 ethyl decanoate 33 decanoic acid
23 34 compound h
13 compound a
50- 14 diethyl succinate 35 compound i
15 compound b 36 dodecanoic acid
16 nerol 37 methyl vanillate
14 17 a-tetpineol 38 compound j
18 compound c 39 compound k
19 methionol
25- 28 20 (E) linalool oxide pyranoid
13
21 35
15
30 31 34 38

J
' 10 24 I 26 27 39
37

T^K'U'ft
1000 2000 3000 4000
K-t,,,,!
5000 6000 7000 8000
V'II I'H ^i V^.U ■■*.,,
"5:00 v10: 00 '15:00 "20:00 "25:00 "30:00 "35:00 H0:00 "45:00 "50:00

Fig. 4.4 GC/MS chromatogram of the extract of Riesling fermented by


Zymaflore VL-1

00
1 3-methyl-1-butanol 26 g-decalactone
2 ethyl hexanoate 27 compound t
3 hexyl acetate 28 compound g
4 n-pentanol 29 decanoic acid
5 ethyl lactate 30 compound h
100 6 1-hexanol 31 compound i
«td
21 23
7 n-nonan-2-ol 32 dodecanoic acid
8 linalool 33 methyl vanillate
9 2,3-butanediol 34 compound k
10 dihyd rote rpinyI acetate
11 g-butyrolactone
I
112 ethyl decanoate
1 13 compound a
23 14 diethyl succinate
15 compound b
16 a-terpineol
17 compound c
50- 18 methionol
19 phenyl ethyl acetatei
20 hexanoic acid
20 21 B-phenyl ethanol
14 23 compound e
24 3-phenyl propanol
25H 25 octanoic acid
24
12 31

U~- ILuL 78 9 11 13] 16


iT^M
5000
19

iJ
6000
17000
I"!UJUu.
8000
80
82 S3 34

1000 2000 3000 4000


^5:00 '10:00 "15:00 "20:00 "25:00 "30:00 "35:00 '40:00 "45:00 "50:00

Fig. 4.5 GC/MS chromatogram of the extract of Riesling fermented by


Epernay 2 CEG

CD
1 3-methyl-1-butanol 23 3-phenyi-i-propanol
2 ethyl hexanoate 24 octanoic acid
3 n-pentanol 25 g-decalactone
4 ethyl lactate 26 compound i
5 1-hexanol 27 decanoicacid
100" 6 compound a 28 compound j
22 7 compound b 29 compound k
8 3-hexen-1-ol 30 compound I
9 acetic acid 31 comound m
10 compound c 32 dodecanoic acid
75-
24 11 compound d
12 compound e
13 2,3-butanediol
14 compound f
tatradcoarw
15 g-butyrolactone
16 ethyl decanoate
50- 17 diethyl succinate
18 nerol
21
27
19 methionol
20 phenyl ethyl acetate
21 hexanoic acid
22 13 phenyl ethanol
25-
17

20 23

li
IS,"
undtomm
JX
1000
3 *, 37 8
-f I
2000 3000
9 10 12 j 18 19
J,
4000
"-f
5000
■r*-'
6000
23

v
J
26

7000
2^ 31 32

:00 ''l 0:00 '15:00 "20:00 "25:00 "30:00 "35:00 40:00

Fig. 4.6 GC/MS chromatogram of the extract of Chardonnay fermented


by EC-1118 Prise de Mousse

o
1 3-methyl-1-butanol
2 ethyl hexanoate
3 n-pentanol
4 ethyl lactate
100 5 1-hexanol
13 6 compound c
7 g-butyrolactone
8 ethyl decanoate
9 diethyl succinate
16 10 methionol
75
1 11 phenyl ethyl acetate
12 hexanoic acid
I 13 6 phenyl ethanol
- 14 compound g
i 15 3-phenyl-1-propanol
16 octanoic acid
50^ 17 compound h
18 g-decalactone
19 compound i
20 decanoic acid
12 21 compound k
20
22 compound I
25
23 compound m
1 24 dodecanoic acid
22
15
11

0J v. a 1 ,i
1000 2000 3000
G ^j 10

4000
-A ^
5000
14
~M
6000
1718 19 21

7000
23
I
2*
' ' ' ■

"5:00 '10:00 '15:00 "20:00 "25:00 "30:00 "35: 00 HO: 00

Fig. 4.7 GC/MS chromatogram of the extract of Chardonnay fermented


by D47

K>
122

compounds were detected in addition to the two internal standards, undecane


and tetradecane (Fig.4.6). The D47 wine extract contained 23 volatile
compounds that were detected in addition to the internal standards (Fig.4.7).
The relatively low number of compounds detected can be attributed to the
extraction method. Freon (tnchlorofluoromethane) is often chosen because it

facilitates the recovery and concentration of volatile compounds in aqueous-


alcoholic solutions (Herraiz et al. 1991) and extracts more flavor compounds

than other solvents (Akhtar et al. 1985). However, extraction procedures are

often carried out for 14 to 24 hours and over 100 compounds can be detected
by GC/MS analysis (Park and Noble, 1993; Herraiz et al. 1991). Extraction
with Freon 113 stirred for 1 hour in Catawba wine resulted in the identification

of 36 compounds which represented 95.7 to 97.5 % of the total extractable


volatile compounds (Nelson et al. 1978) which is similar to our results in both

varieties using a shorter extraction period.

The relative concentrations of the detected volatile compounds was

measured, using the method of Tracey and Britz (1989) in order to make a
comparison between the yeast strains. In Riesling wines, volatile carboxylic

acid content (Table 4.4) was similar in relative concentration between VL-1

and CEG. Acetic acid was not detected in the CEG wines, which was
surprising since higher amount of VA was measured in this wine. One

explanation could be that VL-1 wines had a higher amount of acetic acid.

Hexanoic, octanoic, decanoic and dodecanoic acids were present in similar


relative concentrations. The sensory properties of most of these compounds

can be unpleasant at high level, however they have relatively high odor

thresholds and high amounts are necessary for sensory detection.


123

Table 4.4. Carboxylic acids identified by Kovats Indices and GC-MS spectra
and the range of the relative concentration in Riesling fermented with
Zymaflore VL-1 and Epernay 2 CEG.

Compound Kl VL1 CEG Sensory


(I20M) Relative concentration1 properties

Acetic acid 1415 3.5-4.4 NDa vinegarb

Hexanoic acid 1839 674-685 562-574 sickening, sweaty


rancid.sour, sharp
pungent, cheesyb

Octanoic acid 2057 1688-1932 850-2450 unpleasant, oily


fatty odorb

Decanoic acid 2323 753-977 983-994 fatty, unpleasant


rancid odor,
citrusb

Dodecanoic acid 2659 23.7-30.6 25.0-30.2 fattyb

1 Area of chromatographic peak/area of internal standards


a Not detected in the sample
b Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.

At sub-threshold level, they may add to the complexity and not necessarily
confer a negative aspect to the wine.

Five carboxylic acids (Table 4.5) were positively identified in the

Chardonnay wines. Acetic acid was detected in the EC-1118 wine but not in
the D47 wine. The acetic acid results correspond to the VA analysis since EC-

1118 also appeared to have more acetic acid compared to D47. The amount

of acetic acid has been previously reported to vary depending on the yeast
124

race (Millan and Ortega, 1988). Hexanoic acid was found in higher range of
relative concentration in the EC-1118 extract. Since this compound has an

unpleasant aroma, the EC-1118 wine might be affected by a big difference if


the concentrations were above threshold value. Decanoic acid also had a
higher range in EC-1118 and the same may apply since it also has an

unpleasant odor. Octanoic and dodecanoic acids were similar in both wines.

Table 4.5. Carboxylic acids identified by Kovats Indices and GC-MS spectra
and the range of the relative concentration in Chardonnay fermented with EC-
1118 Prise de Mousse and D47.

Compound Kl EC-1118 D47 Sensory


(I20M) Relative concentration1 properties

Acetic acid 1453 3.8-6.2 ND* vinegar^

Hexanoic acid 1791 1236-1334 420-428 sickening, sweaty


rancid,sour, sharp
pungent, cheesy,
fatty b

Octanoic acid 2007 2155-4399 2235-2319 unpleasant, oily


fatty odorb

Decanoic acid 2219 83-151 796-804 fatty, unpleasant


rancid odor,
citrusb

Dodecanoic acid 2416 5.2-11.9 6.2-8.7 fattyb

1 Area of chromatographic peak/area of internal standards


a Not detected in the sample
b Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
125

Alcohols (Table 4.6) were similar for both yeast strains in the Riesling
wines. 3-methyl-1-butanol is a isoamyl alcohol which is the main fusel alcohol
produced during fermentation that is 40-70% of total alcohol fraction

(Suomalainen and Lehtonen, 1979). It confers a fusel oil or whiskey odor at

high concentration. B-phenyl ethanol (an aromatic alcohol) is also found in

larger range of concentration in CEG wine compared to VL1 wine and its odor

Table 4.6. Alcohols identified by Kovats Indices and GC-MS spectra and the
range of the relative concentration in Riesling fermented with Zymaflore VL-1
and Epernay 2 CEG.

Compound Kl VL1 CEG Sensory


(I20M) Relative concentration1 properties

3-methyl-1-butanol 1182 3211-3589 2999-3521 fusel oil, whiskey*

n-pentanol 1263 7.5-10.9 11.4-16.4 strong, somewhat


sweetb

1-hexanol 1324 246-250 262-276 alcoholic.ethereal


medicinal3

B-phenyl ethanol 1906 2898-3062 3048-3330 floral, rose,


perfume^

3-phenyl propanol 2041 578-608 503-619 sweet, balsamic


floral, hyacinth^

1 Area of chromatographic peak/area of internal standards


a Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
b Miranda-Lopez, 1990.
126

B-phenyl ethanol (an aromatic alcohol) is also found in larger range of


concentration in CEG wine compared to VL1 wine and its odor is more floral
and desirable. Beyond a certain level, alcohols may impart a negative

character to the wine, such as fusel oil odor. (Rapp and Mandery, 1986).

Six alcohols were identified in the Chardonnay wines fermented with EC-
1118 and D47 (Table 4.7). The EC-1118 wines contained larger relative

concentrations of 3-methyl-butanol, 1-hexanol, 3-hexen-1-ol, 13-phenyl ethanol

and 3-phenyl propanol. The relative concentrations appear greater for the 3-

methyl-butanol, 1-hexanol and 6-phenyl ethanol found in the EC-1118 wines.

All of these alcohols may have an affect on the sensory character of the wines
at or above threshold level. All of these alcohols were also higher compared

to the Riesling wines except for 3-phenyl propanol. It could be that the

differences in production of some aromatic volatile compounds by yeasts may

vary with the grape variety due to different relative concentration of precursors.

EC-1118 was reported before by Holloway and Subden (1991) to produce

more higher alcohols than wild yeasts. Earlier work by Rankine (1967)

showed that the strain of yeast had a strong influence on the amount of the
various higher alcohols formed during lab and pilot scale fermentations. Later

work by Giudici et al. (1989) found, in a study of a 100 strains of S.cerevisiae,

that the production of higher alcohols was strain characteristic. Higher

alcohols were shown to be formed, in complex amino acid mixtures partly or

entirely from amino acids through Ehrlich's mechanism. Ayrapaa (1968)

found different patterns that indicated that the utilization of the exogenous
amino acid as a source of nitrogen, and also the extent of regulation of the

biosynthetic pathway would vary with species as well as with strains.


127

Table 4.7. Alcohols identified by Kovats Indices and GC-MS spectra and the
range of the relative concentration in Chardonnay fermented with EC-1118
Prise de Mousse and D47.

Compound Kl EC-1118 D47 Sensory


(I20M) Relative concentration1 properties

3-methyl-butanol 1164 5766-8970 4083-4195 fusel oil, whiskeyb

n-pentanol 1259 56-59 52-55 strong, somewhat


sweet0

1-hexanol 1283 336-538 NDa alcoholic.ethereal


medicinal'3

3-hexen-1-ol 1309 6.5-10.6 ND Unavailable

B-phenyl ethanol 1861 6037-9093 5005-5655 floral, rose,


perfume0

3-phenyl propanol 1992 197-341 175-217 sweet, balsamic,


floral, hyacinth0

1 Area of chromatographic peak/area of internal standards


a Not detected in the sample
b Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
c Miranda-Lopez, 1990.

Stashenko et al. (1992) reported that esters are some of the major

compounds formed during fermentation. Their formation by yeasts were often


associated with quality of white wines (Rosi et al. 1989). Table 4.8 represents

seven esters found in the Riesling wines. The volatile esters are responsible

for the fruity and floral characters in wines (Soles et al. 1982). VL-1 showed
higher relative concentrations of ethyl hexanoate, phenyl ethyl acetate and
128

Table 4.8. Esters identified by Kovats Indices and GC-MS spectra and the
range of the relative concentration in Riesling fermented with Zymaflore VL-1
and Epernay 2 CEG.

Compound Kl VL1 CEG Sensory


(I20M) Relative concentration1 properties

Ethyl hexanoate 1199 252-256 181-189 powerful, fruity


wine-like, apple
banana, brandyb

Hexyl acetate 1232 4.6-9.3 9.6-12.8 apple, cherry,


pear, floral0

Ethyl lactate 1318 88-110 114-120 fruity, buttery


butterscotch13

Ethyl decanoate 1598 162-170 157-181 brandy, oily, fruit


grapeb

Diethyl succinate 1649 301-354 407-445 faint, pleasant0

Phenyl ethyl 1808 130-134 50-57 sweet, honey,


acetate fruity, rose0

Ethyl dodecanoate 1847 2.1-6.2 NDa fruity, floral0

1 Area of chromatographic peak/area of internal standards


a Not detected in the sample
b Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
c Miranda-Lopez, 1990.

ethyl dodecanoate. CEG showed higher relative concentration of hexyl


acetate and diethyl succinate. It appeared that ethyl lactate and ethyl
decanoate were similar in both wines. All the descriptors for the esters are
129

fruity, floral, sweet, honey and butterscotch related. Since most esters are
found above threshold level, they are most likely important contributors to the
aroma of the wines (Herraiz et al., 1991).

Five esters were identified (Table 4.9) in the Chardonnay wines. Ethyl
lactate and ethyl decanoate levels appear to be greater in the EC-1118 and

D47 wines. Ethyl lactate is produced from the esterification of lactic acid

(Simpson and Miller, 1984) and has fruity, buttery or butterscotch characters.
It was found to be present at almost twice the concentration in EC-1118
compared to D47. Ethyl decanoate, described as having a brandy, oily, fruit or

grape odor, was found at a little less than twice the concentration in EC-1118
compared to D47. Ethyl hexanoate was also higher in EC-1118 than D47.
Diethyl succinate and phenyl ethyl acetate were similar in both wines. Soles

et al. (1982) showed that fourteen wine yeast in white wines gave significantly

different ester concentrations. They measured each ester quantitatively


(isoamyl acetate, hexyl acetate, ethyl hexanoate, ethyl octanoate, 2-phenyl

ethyl acetate and ethyl decanoate) and found differences (p < 0.001) due to
yeast species and strain. These investigations also found different banding

patterns using gel electrophoresis for the enzyme alcohol dehydrogenase in

different species and strains, and also found that higher alcohols and their
corresponding esters shared many of the same precursors. The different

banding patterns could indicate differences in the enzyme which might be

related to the activity and specificity. It could influence the formation of higher
alcohols, and therefore could influence the formation of the corresponding

esters. Rosi et al. (1989) explained the production of different concentrations

of esters by showing that yeasts exhibit a wide range of esterase activity, and
that the extent of the hydrolysis was strictly strain dependent. Delfini (1992)
130

Table 4.9. Esters identified by Kovats Indices and GC-MS spectra and the
range of the relative concentration in Chardonnay fermented with EC-1118
Prise de Mousse and D47.

Compound Kl EC-1118 D47 Sensory


(I20M) Relative concentration1 properties

Ethyl hexanoate 1182 326-484 255-283 powerful, fruity


wine-like, apple
banana, brandy3

Ethyl lactate 1278 212-282 127-141 fruity, buttery


butterscotch3

Ethyl decanoate 1571 251-265 162-176 brandy, oily, fruit


grape3

Diethyl succinate 1605 312-546 421-447 faint, pleasantb

Phenyl ethyl acetate 1749 485-360 196-240 sweet, honey,


fruity, roseb

1 Area of chromatographic peak/area of internal standards


a Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
b Miranda-Lopez, 1990.

showed that those esterases were bound to the cellular walls of yeast, and

were differently distributed from strain to strain.


The terpenes linalool and cc-terpineol were similar in both Riesling wines

(Table 4.10). Linalool has a low sensory threshold compared to nerol and ct-

terpineol (Park and Noble, 1993), and is therefore an important sensory


component. Nerol was not detected nor was (E) linalool oxide pyranoid in the

CEG wine. The latter has a very high sensory threshold, and is therefore less
131

Table 4.10. Terpenes identified by Kovats Indices and GC-MS spectra and
the range of the relative concentration in Riesling fermented with Zymaflore
VL-1 and Epernay 2 CEG

Compound Kl VL1 CEG Sensory


{\20M) Relative concentration1 properties

Linalool 1518 20.6-21.7 18.2-25.8 refreshing, light


floral, citrus,
lemon, orange
fragrant, sweet*3

a-terpineol 1662 12.5-14.8 11.9-16.6 fragrant, floral


lilacb

nerol 1671 6.14-6.69 NDa orange, peelc

(E) linalool oxide 1717 3.05-3.29 NDa Unknown


pyranoid

1 Area of chromatographic peak/area of internal standards


a Not detected in two samples
b Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
c Miranda-Lopez, 1990.

likely to contribute to the aroma of the wine to the same extent as the other

three terpenes (Park and Noble, 1993). Geraniol was absent most likely
because the juice did not ferment on the skins. This terpene has been found

to arise mainly from wines fermented on skins (Herraiz et al., 1991). Wilson et

al. (1986) also showed that geraniol, as well as nerol were associated mostly

with the skins, whereas free linalool is uniformly distributed in the skins and
132

the juice. Citronellol was also not detected in the wine. This terpene is
reported to be transformed from geraniol and nerol by yeasts (Dugelay et al.,
1992), and since geraniol was absent, and nerol was present in a low relative

concentration, citronellol could not be produced or was in very low


concentration. The terpenes are very fragrant, flowery or citrus like and are

desirable in Riesling (Herraiz et al., 1991). No terpenes were detected in the


Chardonnay because there are very few in the must compared to the Riesling

(Simpson and Miller, 1984).

The terpenes have important aroma precursors in the glycosides of

terpenes. Those terpenes are released in two stages, and the released
compounds are odorous. In the first stage, depending on the sugar moiety,

arabinofuranosidase, rhamnopyranosidase or apiofuranosidase present in


the grape juice or molds will cleave the intersugar linkage and the

corresponding monoterpenyl-6-D-glucosides will be released. Then the 6-

glucosidase will liberate the monoterpene from the monoterpenyl-G-D-

glucoside (Colagrande et al., 1994). Saccharomyces cerevisiae possesses 6-


glucosidases however, they lack the enzymes of the first stage whose

hydrolytic activity on the bound terpene molecule must precede the second

stage to release the odorous compound (Delfini, 1992). Delcroix et al. (1994)
showed that Muscat fermented with three commercial yeast strains had similar

levels of terpenols They found that the level of activity of the enzymes (3-
glucosidase, a-arabinosidase and a-rhamnosidase were similar in the yeast

strains. They also found that the activity of the yeast's B-glucosidase is
maximum at pH 5.0, which is higher than wine pH, resulting in a decrease of

activity as the pH is reduced. However, they found differences in trans-furan


linalool oxide, nerol, geraniol, citronellol and hotrienol. They also reported
133

that citronellol was absent in the juice and present in the wine, resulting from
geraniol and nerol during yeast fermentation, as reported previously by
Dugelay et al. (1992).

Some miscellaneous compounds (Table 4.11) were also positively


identified in the Riesling wines. Methionol (or 3-methylthio-1-propanol) is an
important compound since it has a low threshold and will result in a cabbage,
stinky odor in the wine. It is also one of the most abundant sulfur containing

compound found in wines (Herraiz et al., 1991). It appeared like VL-1 had a

Table 4.11. Miscellaneous compounds identified by Kovats Indices and GC-


MS spectra and the range of the relative concentration in Riesling fermented
with Zymaflore VL-1 and Epernay 2 CEG.

Compound Kl VL1 CEG Sensory


(I20M) Relative concentration1 properties

methionol 1692 26-30 40-56 vegetative, stinky


cabbage^

Y-butyrolactone 1595 198-217 135-143 faint, sweet,


caramel^

y-decalactone 2143 121-141 159-179 fruity, peach^

methyl vanillate 2803 38.2-42.6 37.8-45.3 vanilla, herbal,


spicy, caramel^

1 Area of chromatographic peak/area of internal standards


a Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
b Miranda-Lopez, 1990.
134

slightly lower relative concentration of methionol compared to CEG.


Methionol was also reported to be lower in wines fermented with VL-1
compared to CEG in a work by Lavigne et al. (1992). The lactones, y-

butyrolactone and y-decalactone were slightly different in both wines. VL-1


appeared to have more y-butyrolactone, with its sweet, caramel odor, and
CEG appeared to have more y-decalactone which has a more fruity odor.

Methyl vanillate was similar in both wines.

Five miscellaneous compounds were identified in the Chardonnay (Table

4.12). The sulfur containing compound 2,3-butanediol and the terpene nerol

were found the EC-1118 but not D47 extract. Methionol was found in slightly
higher relative concentration in EC-1118. Two lactones, y-butyrolactone and
y-decalactone were also higher in EC-1118, however, y-decalactone

appeared to be similar in both wines, y-lactones are formed by the reduction

of ketoacids and further cyclization during the anaerobic phase of the yeast

fermentation (Etievant and Bayonove, 1983).

No volatile phenolic compounds were found in either variety. More

phenols appear in wines fermented on the skins (Herraiz et al., 1991). Even if
phenols were not detected, it is important to mention that Chatonnet et al.

(1993) showed that different commercial strains of S.cerevisiae had different

intensity of activity of the substituted cinnamate decarboxylyase. This enzyme

is responsible for the decarboxylation of phenolic acids into vinylphenols.

Vinylphenols have phenolic and medicinal aromas that can be desirable or


undesirable depending on the concentration. The VL-1 strain was reported
(also by Grando et al., 1993) to have very low activity compared to the UCD

522 Montrachet (identical to the M strain in the Chardonnay). This would be

more important in varieties fermented on the skins, such as red wines.


135

Table 4.12. Miscellaneous compounds identified by Kovats Indices and GC-


MS spectra and the range of the relative concentration in Chardonnay
fermented with EC-1118 Prise de Mousse and DAT.

Compound Kl EC-1118 D47 Sensory


(I20M) Relative concentration1 properties

2,3-butanediol 1516 6.3-9.1 NDa Unavailable

y-butyrolactone 1550 99-129 81-89 faint, sweet,


caramel0

nerol 1613 5.0-13.3 ND orange, peel0

methionol 1646 24.1-25.7 15.9-18.0 vegetative, stinky


cabbage0

y-decalactone 2107 23.3-24.6 21.9-22.0 fruity, peachb

1 Area of chromatographic peak/area of internal standards


a Not detected in two samples
b Sensory properties of pure standards from Aldrich Flavors and Fragrances.
Aldrich Chemical Company Inc. Milwaukee, Wl.
c Miranda-Lopez, 1990.

In the Riesling wines, four compounds were tentatively identified (Table


4.13). N-nonan-2-ol, and dihydroterpinyl acetate were not detected in the VL-

1 wine, and the terpene hotrienol was not detected in the CEG wine. Similar

relative concentrations of 2,3-butanediol were seen. Compounds a, b, d and k


were all found in slightly higher relative concentrations in the VL-1 wine

(compound k was not detected in the CEG wine). Compounds c, e, g and h

were similar in both wines. The CEG wine appeared to have more in relative
concentration of compounds f, i and j.
136

Table 4.13. Tentatively identified and unidentified compounds and the range
of the relative concentration in Riesling fermented with Zymaflore VL-1 and
Epernay 2 CEG.

Compound Kl VL1 CEG


(I20M) Relative concentration1

n-nonan-2-ol 1515 Not Detected 16.6-20.1


2,3-butanediol 1549 15.4-17.3 17.4-20.7
dihydroterpinyl 1573 Not Detected 9.2-10.2
acetate
hotrienol 1865 5.0-9.0 Not Detected

Compound a 1605 10.2-11.2 1.8-3.3

Compound b 1643 55.3-64.1 34.3-50.5


Compound c 1680 4.6-7.3 6.7-7.3

Compound d 1965 7.5-11.5 3.9-6.5

Compound e 1978 5.5-10.8 2.8-8.7


Compound f 2266 16.3-17.5 21.8-28.5

Compound g 2310 35.9-44.1 37.3-43.7

Compound h 2459 32.1-40.1 29.6-39.6


Compound i 2553 122.1-143.9 190.6-225.7

Compound j 2837 4.1-6.5 Not Detected


Compound k 2853 29.6-32.2 38.6-42.5
1 Area of chromatographic peak/area of internal standards

Thirteen unidentified compounds were also detected in Chardonnay (Table

4.14). Compounds i, and m were similar in both wines. Compounds a, b, d, e,


137

f, and j were not detected in the D47 extract, however, except for compounds e
and j, the other compounds were in very low relative concentrations.

Table 4.14. Unidentified compounds and the range of the relative


concentration in Chardonnay fermented with EC-1118 Prise de Mousse and
D47.

Compound Kl EC-1118 D47


(I20M) Relative concentration1

Compound a 1297 2.2-5.4 Not detected

Compound b 1302 1.1-1.5 Not detected

Compound c 1482 36-62 13.6-13.7

Compound d 1489 3.0-7.1 Not detected

Compound e 1511 10.3-18.3 Not detected

Compound f 1536 2.2-3.6 Not detected

Compound g 1901 Not detected 9.5-11.3

Compound h 2049 Not detected 5.5-9.9

Compound i 2160 14.9-20.2 18.6-18.9

Compound j 2242 25.2-27.8 Not detected

Compound k 2268 31-355 36-49

Compound I 2313 19-535 279-297

Compound m 2319 8-42 24-30

1 Area of chromatographic peak/area of internal standards


138

Compounds g and h were not detected in the EC-1118 but were present in
the D47 extract. Compounds c and k were found to have higher relative

concentration in EC-1118 and were detected in the D47 extract.

CONCLUSION

Yeast strains appeared to influence the formation of some volatile

compounds, which correlated with the work done by other researchers.

Whether or not the differences in the compounds will lead to significant

aroma/flavor perception in extracts was not established, and could only be

done so by GC-sniffing, as shown by Miranda-Lopez (1990) in her work with


the aroma of Pinot Noir and by Acree (1984). GC-sniffing can identify
potentially important aroma active compounds. In order to know the sensory

impact of a volatile compound, it would then need to be added to wines in

varying concentrations to determine aroma thresholds and generate


descriptors. A comparison of the concentrations found in the wines with the

sensory threshold of each volatile compounds could also indicate its relative
importance in the wine. The complex mixture of volatile compounds and non-
volatile compounds and their interactions could also influence the impact on

the sensory attributes of the wine. Further research needs to focus on the

relationship between the volatile composition of a wine and its sensory profile

and the role of the yeast strain with respect to specific aroma compounds.
139

LITERATURE CITED

Acree, T.E., Barnard, J., and D.G. Cunningham. A Procedure for the sensory
analysis of gas chromatography effluents. Food Chem. 14: 273-286 (1984).

Akhtar, M., Subden, R.E., Cunningham, J.D., Fyfe, C. and A. Mejering.


Production of volatile yeast metabolites in fermenting grape musts. Can. Inst.
Food Sci. Technol. J. 18(4):280-283 (1985).

Amerine, M.A., and C.S. Ough. Wine and Musts analysis. John Wiley & Sons.
NY (1980).

Avedovech, R.M.Jr., McDaniel, M.R., Watson, B.T., and W.E. Sandine. An


evaluation of combination of wine yeast and Leuconostoc oenos strains in
malo-lactic fermentation of Chardonnay wine. Am. J. Enol. Vitic. 43(3) :253-
260 (1992).

Ayrapaa, T. Formation of higher alcohols by various yeasts. J. Inst. Brew. 74:


169-178 (1968).

Cabrera, M.J., Moreno, J., Ortega, J.M., and M. Medina. Formation of ethanol,
higher alcohols, esters and terpenes by five yeast strains in musts from Pedro
Ximenez grapes in various degrees of ripeness. Am. J. Enol. Vitic. 39(4):283-
287 (1988).

Chatonnet, P., Dubourdieu, D., Boidron, J.N., and V. Lavigne. Synthesis of


volatile phenols by Saccharomyces cerevisiae in wines. J. Sci. Food Agric.
62:191-202 (1993).

Ciani, M., Palpacelli, V., and G. Rosini. Oenological aptitude of some


Saccharomyces cerevisiae cultures: fermentative and metabolic characters
verified during grape juice fermentations. Ital. J. Food Sci. 4:281-290 (1991).

Colagrande, O., Silva, A. and M.D. Fumi. Recent applications of biotechnology


in wine production. Biotechnol. Prog. 10(1):2-18 (1994).
140

Delcroix, A., GLinata, Z., Sapis, J.-C, Salmon, J.-M., and C. Bayonove.
Glycosidase activities of three enological yeast strains during winemaking;
effect on the terpenol content of Muscat wine. Am. J. Enol. Vitic. 45(3):291-296
(1994).

Delfini, C. Innovative trends in oenology and in the selection of yeasts and


malolactic bacteria for wine industry. Riv. Vitic. Enol. 1:17-30 (1992).

Dugelay, I., Gunata, Z., Sapis, J.C., Baumes, R. and C. Bayonove. Etude de
I'origine du citronellol dans les vins. J. Int. Sci. Vigne Vin. 26(3): 177-184
(1992).

Etievant, P.X., and C.L. Bayonove. Aroma components of pomaces and wine
from the variety Muscat de Frontignan. J. Sci. Food Agric. 34: 393-403 (1983).

Etievant, P.X., Issanchou, S.N., and C.L. Bayonove. The flavor of Muscat wine:
the sensory contribution of some volatile compounds. J. Sci. Food Agric.
34:497-504 (1984).

Giudici, P., Romano, R. and C. Zambonelli. A biometric study of higher alcohol


production in Saccharomyces cerevisiae Can. J. Microbiol. 36:61-64 (1989).

Grando, M.S., Versini, G., Nicolini, G., and F. Mattivi. Selective use of wine
yeast strains having different volatile phenols production. Vitis. 32:43-50
(1993).

Herraiz, T., Reglero, G., Martin-Alvarez, P.J., Herraiz, M., and M.D. Cabezudo.
Identification of aroma components of Spanish Verdejo wine. 55:103-116
(1991).

Holloway, P. and R.E. Subden. Volatile metabolites produced in a Riesling


must by wild yeast isolates. Can. Inst. Food Sci. Technol. J. 24(1/2):57-59
(1991).

Jennings, W., and T. Shimamoto. Qualitative analysis of flavor and fragrance


volatile compounds by glass capillary gas chromatography. Academic Press
Inc., New York. (1980).
141

Kovats, E. von, Simon, W., and E. Heilbronner. Programm - Gesteverter Gas-


Chromatograph zur praparativen trennung von genishen organisher verbin
dungen. Helv. Chim. Acta. 32: 275- 278 (1958).

Lavigne, V., Boidron, J.N., and D. Dubourdieu. Formation des composes


soufres lourds au cours de la vinification des vins blancs sees. J. Int. Sci.
Vigne Vin. 26(2):75-85 (1992).

Libbey, L. A Paradox database for GC-MS data on components of essential


oils and other volatile compounds. J. Ess. Oil Res. 3: 193-194 (1990).

Longo, E., Velazquez, J.B., Sieiro, C, Cansado, J., Calo, P., and T.G. Villa.
Production of higher alcohols, ethyl acetate, acetaldehyde and other
compounds by 14 Saccharomyces cerevisiae wine strains isolated from the
same region (Salnes, N.W. Spain). World J. Microbiol. Biotechnol. 8: 539-541
(1992).

Mateo, J.J., Jimenez, M., Huerta, T., and A. Pastor. Comparison of volatile
compounds produced by four Saccharomyces cerevisiae strains isolated form
Monastrell musts. Am. J. Enol. Vitic. 43(2):206-209 (1992).

Millan, C, and J.M. Ortega. Production of ethanol, acetaldehyde, and acetic


acid in wine by various yeast races: role of alcohol and aldehyde
dehydrogenase. Am. J. Enol. Vitic. 39(2):107-112 (1988).

Miranda-Lopez, R. A maturity trial study of Pinot Noir wines: Aroma profile by


sniffing gas chromatographic effluent. Thesis, Oregon State University (1990).

Montedoro, G., and M. Bertuccioli. The flavour of wines, vermouth and fortified
wines. In Food Flavours. Part B. The flavour of beverages. Morton, I. and A.J.
MacLeod (Eds.) Elsevier Publishing Co. (1986).

Muller, C.J., Wahlstrom, V.L., and K.C. Fugelsang. Capture and use of volatile
flavor constituents emitted during wine fermentation. In: Beer and Wine
Production. Analysis, Characterization, and Technological Advances. ACS
Symposium Series, pp. 219-232. B.H. Gump & D.J. Pruett Eds. (1993).
142

Nelson, R.R., Acree, T.E. and R.M. Butts. Isolation and identification of volatile
compounds from Catawba wine. J. Agric. Food Chem. 26(5):1188-1190
(1978).

Nykanen, L. Formation and occurrence of flavor compounds in wine and


distilled alcoholic beverages. Am.J. Enol.Vitic. 37(1): 84-96 (1986).

Nykanen, L, and H. Suomalainen. Aroma of beer, wine and distilled alcoholic


beverages. D. Reidel Publishing Company, Dordrecht, Holland (1983).

Park, S.K. and A.C. Noble. Monoterpenes and monoterpene glycosides in


wine aromas. Beer and Wine Production. Analysis, Characterization, and
Technological Advances. ACS Symposium Series. B,H. Gump & D.J. Pruett
Eds. (1993).

Perez-Gonzalez, J.A., Gonzalez, R., Querol, A., Sendra, J., and D. Ramon.
Construction of a recombinant wine yeast strain expressing I3-(1,4)-
endoglucanase and its use in microvinification processes. App. Env.
Microbiol. 59(9): 2801-2806 (1993).

Radler, F. Yeasts-Metabolism of organic acids. In Wine Microbiology and


Biotechnology. Fleet, G.H. (Ed), pp.165-182. Hardwood Academic Publishers.
Switzerland. (1992).

Rankine, B.C. Formation of higher alcohols by wine yeasts, and relationship to


taste thresholds. J. Sci. Food Agric. 18: 583-589 (1967).

Rapp, A., and H. Mandery. Wine aroma. Experientia. 42: 873-883 (1986).

Rosi, I., Contini, M., and M. Bertuccioli. Relationship between enzymatic


activities of wine yeasts and aroma compound formation. In: Flavors and Off-
Flavors. Proceedings of the 6th International Flavor conference, pp 103-120.
Elsevier Science Publishers, Amsterdam (1989).

Schreir, P. Flavor composition of wines: a review. CRC Critical Reviews in


Food Science and Nutrition. 12: 59-111 (1979).
143

Sefton, M.A., Francis, I.L., and P.J. Williams. The volatile composition of
Chardonnay juices: a study by flavor precursors analysis. Am. J. Enol. Vitic.
44(4):359-370 (1993).

Simpson, R.F. and G.C. Miller. Aroma composition of Chardonnay wine. Vitis.
23: 143-158 (1984).

Singleton, V. Grapes and wine phenolics; background and prospects. Proc.


Grape and Wine Centennial Symposium. A.D.Webb Ed. pp. 215-227.
University of California Press, Berkeley (1980).

Soles, R.M, Ough, C.S., and R.E. Kunkee. Ester concentration differences in
wine fermented by various species and strains of yeasts. Am. J. Enol. Vitic.
33(2):94-98 (1982).

Stashenko, H., Macku, C, and T. Shibamoto. Monitoring volatile chemicals


formed from must during yeast fermentation. J. Agric. Food Chem. 40:2257-
2259 (1992).

Suomalainen, H. Yeast esterases and aroma esters in alcoholic beverages. J.


Inst. Brew. 87:296-300 (1981).

Suomalainen, H., and M. Lehtonen. The production of aroma compounds by


yeast. J. Inst. Brew. 85:149-156 (1979).

Tracey, R.P., and T.J. Britz. Freon 11 extraction of volatile metabolites formed
by certain lactic acid bacteria. App. Env. Microbiol. 55(6):1617-1623 (1989).

Watson, B.T. Jr., Dodd, E.E., and D.A. Heatherbell. Basic Chemical Analysis of
Musts and Wines. Table Wine Workshop. Department of Food Science &
Technology, Oregon State University, OR. May 16-17, (1980).

Wilson, B., Strauss, C.R., and P.J. Williams. The distribution of free and
glycosidically-bound monoterpenes among skin, juice and pulp fractions of
some white grape varieties. Am. J. Enol. Vitic. 37(2): 107-111 (1986).
144

Zoecklein, B.W., Fugelsang, K.C., Gump, B.H., and F.S. Nury. Production wine
analysis. 475 pp. AVI Book. Van Norstrand Reinhold, New York (1990).
145

Chapter 5. Summary

From this study, it was shown that yeast strain selection for winemaking is

an important step to obtain wine with desirable sensory characters. The wine

yeast will influence the sensory development of the wine, and if selected

properly, will enhance particular characters in different varieties. Free-Choice

Profiling was also shown to be a suitable method for the evaluation of wines,
and could be use in other work for a faster and accurate sensory technique.

The differences in sensory perception of wines fermented with different strains

of Saccharomyces cerevisiae could eventually be explained by different


chemical and volatile composition of wines. Studies have to continue to

determine the contribution esters, alcohols, terpenes and numerous other

compounds produced by yeasts, and show how they are related to sensory
differences perceive in wines.
146

Bibliography

Acree, T. Private communication. 1993.

Acree, T.E., Barnard, J., and D.G. Cunningham. A Procedure for the sensory
analysis of gas chromatography effluents. Food Chem. 14: 273-286 (1984).

Akhtar, M., Subden, R.E., Cunningham, J.D., Fyfe, C, and A.G. Meiring.
Production of volatile yeast metabolites in fermenting grape musts. Amerine,
M.A., and C.S. Ough. Wine and Musts analysis. John Wiley & Sons. NY
(1980).

Arnold, G.M., and A.A. Williams. The use of Generalized Procrustes


techniques in sensory analysis. In: Statistical Procedures in Food Research.
Piggot, J.R. (Ed.), pp. 233-255. Elsevier Applied Science. London (1986).

Aryan, A.P., Wilson, B., Strauss, C.R., and P.J. Williams. The properties of
glycosidases of Vitis vinifera and a comparison of their B-glucosidase activity
with that of exogenous enzymes. An assessment of possible application in
enology. Am. J. Enol. Vitic. 38(3): 182-188 (1987).

Avedovech, R.M. Jr., McDaniel, M.R., Watson, B.T., and W.E. Sandine. An
evaluation of combinations of wine yeast and Leuconostoc oenos strains in
malo-lactic fermentation of Chardonnay wine. Am. J. Enol. Vitic. 43(3): 253-
260 (1992).

Ayrapaa, T. Formation of higher alcohols by various yeasts. J. Inst. Brew. 74:


169-178 (1968).

Ayrapaa, T. Biosynthetic formation of higher alcohols by yeast. Dependence


on the nitrogenous nutrient level of the medium. J. Inst. Brew. 77: 266-276
(1971).

Bioletti, F.T. Pure yeast in wineries. University of California at Berkeley.


College of Agriculture. Circular No.23 (1906).
147

Bisson, L. Yeasts - Metabolism of sugars. In: Wine Microbiology and


Biotechnology. Fleet, G.H. (Ed), pp. 55-76. Hardwood Academic Publishers.
Switzerland (1992).

Blakesley, C.N., and J. Loots. A convenient method for multiple extraction of


volatile flavor components from food slurries and pulp using a two-chambered
glass bomb extractor and dichlorodifluoromethane (Freon 12) solvent. J.
Agric. Food Chem. 25(4): 961-963 (1977).

Brander, C.F., Kepner, R.E., and A.D. Webb. Identification of some volatile
compounds of wine of Vitis vinifera cultivar of Pinot Noir. Am.J. Enol. Vitic.
31(1): 69-75 (1980).

Cabrera, M.J., Moreno, J., Ortega, J.M., and M. Medina. Formation of ethanol,
higher alcohols, esters and terpenes by five yeast strains in musts from Pedro
Ximenez grapes in various degrees of ripeness. Am. J. Enol. Vitic. 39(4): 283-
287 (1988).

Cavazza, A., lacono, F., Stefanini, M., Nicolini, G. and F. Romano. The
environmental adaptability of clones: Influence of the yeast strains and must
clarifying in the modification of wine quality. Vitic. Enol. Sci. 48: 203-207
(1993).

Chatonnet, P., Dubourdieu, D., Boidron, J.-N., and M. Pons. The origin of
ethylphenols in wines. J. Sci. Food Agric. 60: 165-178 (1992).

Chatonnet, P., Dubourdieu, D., Boidron, J.N., and V. Lavigne. Synthesis of


volatile phenols by Saccharomyces cerevisiae in wines. J. Sci. Food Agric.
62:191-202 (1993).

Ciani, M., Palpacelli, V., and G. Rosini. Oenological aptitude of some


Saccharomyces cerevisiae cultures: fermentative and metabolic characters
verified during grape juice fermentations. Ital. J. Food Sci. 4:281-290 (1991).

Cobbs, C.S., Bursey, M.M., and A.C. Rice. Major volatile components of
Aurore wine: a gas chromatographic - mass spectrometric analysis combining
ionization and electron impact ionization. J. Food Science. 43: 1822-1825
(1978).
148

Colagrande, O., Silva, A., M.D. Fumi. Recent applications of biotechnology in


wine production. Biotechnol. Prog. 10: 2-18.(1994).

Darriet, P., Boidron, J.-N., and D. Dubourdieu. Hydrolysis of the terpenic


heterosides of small Muscat grapes using the periplasmic enzymes of
Saccharomyces cerevisiae . Connaissance de la Vigne et du Vin. 22(3): 189-
192 (1988).

Degre, R. Selection and commercial cultivation of wine yeast and bacteria. In:
Wine Microbiology and Biotechnology. Fleet, G.H. (Ed), pp. 421-448.
Hardwood Academic Publishers. Switzerland (1992).

Delcroix, A., Gunata, Z., Sapis, J.-C, Salmon, J.-M., and C. Bayonove.
Glycosidase acitvities of three enological yeast strains during winemaking;
effect on the terpenol content of Muscat wine. Am. J. Enol. Vitic. 45(3):291-296
(1994).

Delfini, C. Innovative trends in oenology and in the selection of yeasts and


malolactic bacteria for wine industry. Riv. Vitic. Enol. 1: 17-30 (1992).

Dijksterhuis, G.B., and S. van Buuren. Procrustes-PC v2.2 Manuel. OP&P


Utrecht Holland (1991).

Dugelay, I., GLinata, Z., Sapis, J.C., Baumes, R., and C. Bayonove. Etude de
I'origine du citronellol dans les vins. J. Int. Sci. Vigne Vin. 26(3): 177-184
(1992).

Etievant, P.X., and C.L. Bayonove. Aroma components of pomaces and wine
from a variety Muscat de Frontignac. J. Sci. Food Agric. 34: 393-403 (1983).

Etievant, P.X., Issanchou, S.N., and C.L. Bayonove. The flavor of Muscat wine:
the sensory contribution of some volatile compounds. J. Sci. Food Agric.
34:497-504 (1984).

Fleet, G.H. Growth of yeasts during wine fermentation. J. Wine Res. 1(3): 211-
223 (1990).
149

Fleet, G.H. The microorganisms of wine making- isolation, enumeration and


identification. In Wine Microbiology and Biotechnology. Fleet, G.H. (Ed), pp.1-
26. Hardwood Academic Publishers. Switzerland. (1992).

Fleet, G.H., and G.M. Heard. Yeasts - Growth during fermentation. In: Wine
Microbiology and Biotechnology. Fleet, G.H. (Ed), pp. 27-54. Hardwood
Academic Publishers. Switzerland (1992).

Gains, N., Krzanowski, W.J. and D.M.H. Thompson. A comparison of variable


reduction techniques in an attitudinal investigation of meat products. J.
Sensory Stud. 3: 37-48 (1988).

Gains, N., and D.M.H. Thompson. Sensory profiling of canned lager beers in
their own homes. Food Quality Preference. 2: 39-47 (1990).

Giudici, P., Romano, R. and C. Zanbonelli. A biometric study of higher alcohol


production in Saccharomyces cerevisiae Can. J. Microbiol. 36:61-64 (1989).

Gordon, M.H. Principles of gas chromatography. In: Principles and


Applications of Gas Chromatography in Food Analysis, pp.11-58. Ellis
Horwood Ltd. (1990)

Grando, M.S., Versini, G., Nicolini, G., and F. Mattivi. Selective use of wine
yeast strains having different volatile phenols production. Vitis. 32:43-50
(1993).

Grossmann, M., and A. Rapp. Steigerung des sortentypischen Weinbuketts


nach Enzymbehandlung. Deutsche Lebensmittel-Rundschau. 84(2): 35-37
(1988).

Guy, C, Piggott, J.R., and S. Marie. Consumer profiling of Scotch whisky.


Food Quality Preference. 2: 69-73 (1989).

Henschke, P.A., and V. Jiranek. Yeasts - Metabolism of nitrogen, compounds.


In: Wine Microbiology and Biotechnology. Fleet, G.H. (Ed), pp. 77-164.
Hardwood Academic Publishers. Switzerland. (1992).
150

Herraiz, T., and C.S. Ough. Formation of ethyl esters of amino acids by yeasts
during the alcoholic fermentation of grape juice. Am. J. Enol. Vitic. 44(1): 41-
48 (1993).

Herraiz, T., Reglero, G., Martin-Alvarez, P.J., Herraiz, M., and M.D. Cabezudo.
Identification of aroma components of Spanish Verdejo wine. 55:103-116
(1991).

Holloway, P., and R.E. Subden. Volatile metabolites produced in a Riesling


must by wild yeast isolates. Can. Inst. Sci. Technol. J. 24(1/2): 57-59 (1991).

Javelot, C, Girard, P., Colonna-Ceccaldi, B., and B. Vladescu. Introduction of


terpene-producing ability in a wine strain of Saccharomyces cerevisiae . J.
Biotechnol. 21: 239-251 (1991).

Jennings, W., and T. Shibamoto. Qualitative analysis of flavor and fragrance


volatile compounds by glass capillary gas chromatography. pp 22-24.
Academic Press Inc. NY (1980).

Kovats, E. von, Simon, W., and E. Heilbronner. Programm - Gesteverter Gas-


Chromatograph zur praparativen trennung von genishen organisher verbin
dungen. Helv. Chim. Acta. 32: 275- 278 (1958).

Lafon-Lafourcade, S. Applied microbiology. Experientia. 42: 904-914 (1986).

Lavigne, V., Boidron, J.N., and D. Dubourdieu. Formation des composes


soufres lourds au cours de la vinification des vins blancs sees. J. Int. Sci.
Vigne et du Vin. 26(2): 75-85 (1992).

Lawless, H.T. Flavor description of white wine by "expert" and nonexpert wine
consumers. J. Food Sci. 49: 120-123 (1984).

Libbey, L. A Paradox database for GC-MS data on components of essential


oils and other volatile compounds. J. Ess. Oil Res. 3: 193-194 (1990).
151

Longo, E., Velazquez, J.B., Sieiro, C, Cansado, J., Calo, P., and T.G. Villa.
Production of higher alcohols, ethyl acetate, acetaldehyde and other
compounds by 14 Saccharomyces cerevisiae wine strains isolated from the
same region (Salnes, N.W. Spain). World J. Microbiol. Biotechnol. 8: 539-541
(1992).

MacFie, H.J, Bratchell, N., Greenhoff, K, and L.V. Vallis. Designs to balance
the effect of order of presentation and first-order carry-over effects in Hall tests.
J. Sensory Studies. 4: 129-148 (1989).

Maga, J.A. Analysis of aroma volatile compounds. In: Principles and


Applications of Gas Chromatography in Food Analysis. Ellis Norwood Ltd.
(1990).

Marshall, R.J., and S.P.J. Kirby. Sensory measurement of food texture by free-
choice profiling. J. Sensory Stud. 3: 63-69 (1988).

Mateo, J.J., Jimenez, M., Huerta, T., and A. Pastor. Comparison of volatile
compounds produced by four Saccharomyces cerevisiae strains isolated form
Monastrell musts. Am. J. Enol. Vitic. 43(2):206-209 (1992).

McDaniel, M.R., Miranda-Lopez, R., Watson, B.T., Michaels, N.J., and L.M.
Libbey. Pinot Noir aroma: A sensory/gas chromatography approach. In:
Flavors and Off-Flavors, Proceedings of the 6th International Flavor
Conference. Rethymnon Crete, Greece, 5-7 July, 1989. pp. 23-36 (1989).

McEwan, J.A., Colwell, J.S., and D.M. Thompson. The application of two free-
choice profiling methods to investigate the sensory characteristics of
chocolates. J. Sensory Stud. 3: 271 (1989).

Meilgard, M., Civille, G.V., and B. Carr. Sensory Evaluation Techniques. 2nd
Ed. CRC Press (1991).

Millan, C, and J.M. Ortega. Production of ethanol, acetaldehyde, and acetic


acid in wine by various yeast races: role of alcohol and aldehyde
dehydrogenase. Am. J. Enol. Vitic. 39(2):107-112 (1988).
152

Miranda-Lopez, R. A maturity trial study of Pinot Noir wines: Aroma profile by


sniffing gas chromatographic effluent. Thesis, Oregon State University (1990).

Montedoro, G., and M. Bertuccioli. The flavour of wines, vermouth and fortified
wines. In: Food Flavours. Part B. The flavour of beverages. Morton, I. and A.J.
MacLeod (Eds.) Elsevier Publishing Co., London (1986).

Muller, C.J., Wahlstrom, V.L., and K.C. Fugelsang. Capture and use of volatile
flavor constituents emitted during wine fermentation. In: Beer and Wine
Production. Analysis, Characterization, and Technological Advances. ACS
Symposium Series, pp. 219-232. B,H. Gump & D.J. Pruett Eds. (1993).

Nelson, R.R., Acree, T.E., and R.M. Butts. Isolation and identification of
volatiles from Catawba wines. J. Agric. Food Chem. 26(5): 1188-1190 (1978).

Noble, A.C. Analysis of wine sensory properties. In: Wine Analysis. Linskens,
and Jackson (Eds.), pp 9-28. Springer-Verlag, New York (1988).

Noble, A.C, Arnold, R.A., Buechsenstein, J., Leach, E.J., Schmidt, J.O., and
P.M. Stern. Modification of a standardized system of wine aroma terminology.
Am. J. Enol. Vitic. 38(2): 143-146 (1987).

Noble, A.C, Williams, A.A., and S.P. Langron. Descriptive analysis and quality
ratings of 1976 wines from four Bordeaux communes. J. Sci. Food. Agric. 35:
88-98 (1984).

Nykanen, L. Formation and occurrence of flavor compounds in wine and


distilled alcoholic beverages. Am. J. Enol. Vitic. 37(1).: 84-96 (1986).

Nykanen, L., and I. Nykanen. Production of esters by different yeast strains in


sugar fermentations. J. Inst. Brew. 83: 30-31 (1977).

Nykanen, L., and H. Suomalainen. Aroma of beer, wine and distilled alcoholic
beverages. D. Reidel Publishing Company, Dordrecht, Holland (1983).
153

Park, S.K. and A.C. Noble. Monoterpenes and monoterpene glycosides in


wine aromas. Beer and Wine Production. Analysis, Characterization, and
Technological Advances. ACS Symposium Series. B,H. Gump & D.J. Pruett
Eds. (1993).

Perez-Gonzalez, J.A., Gonzalez, R., Querol, A., Sendra, J., and D. Ramon.
Construction of a recombinant wine yeast strain expressing 6-(1,4)-
endoglucanase and its use in microvinification processes. App. Env.
Microbiol. 59(9): 2801-2806 (1993).

Radler, F. Yeasts-Metabolism of organic acids. In Wine Microbiology and


Biotechnology. Fleet, G.H. (Ed), pp.165-182. Hardwood Academic Publishers.
Switzerland. (1992).

Rankine, B.C. Formation of higher alcohols by wine yeasts, and relationship to


taste thresholds. J. Sci. Food Agric. 18: 583-589 (1967).

Rapp, A., Guntert, M., and J. Almy. Identification and significance of several
sulfur-containing compounds in wine. Am. J. Enol. Vitic. 36(3): 219-221
(1985).

Rapp, A., and H. Mandery. Wine aroma. Experientia. 42: 873-883 (1986).

Rosi, I., and M. Bertuccioli. Influences of lipid addition on fatty acid


composition of Saccharomyces cerevisiae and aroma characteristics of
experimental wines. J. Inst. Brew. 98: 305-314 (1992).

Rosi, I., Contini, M., and M. Bertuccioli. Relationship between enzymatic


activities of wine yeasts and aroma compound formation. In: Flavors and Off-
Flavors. Proceedings of the 6th International Flavor conference, pp 103-120.
Elsevier Science Publishers, Amsterdam (1989).

Rubico, S.M. Perceptual characteristics of selected acidulants by different


sensory and multivariate methods. Thesis, Oregon State University (1993).

Ryan, B.E., Joiner, B.L, and T.A. Ryan, Jr. Minitab Handbook. 2nd Ed. PWS-
Kent Publishing (1985).
154

Schreir, P. Flavor composition of wines: a review. CRC Critical Reviews in


Food Science and Nutrition. 12: 59-111 (1979).

Schreir, P. Wine aroma composition: Identification of additional volatiles


constituents of red wine. J. Agric. Food Chem. 28(5): 926-928 (1990).

Sefton, M.A., Francis, I.L., and P.J. Williams. The volatile composition of
Chardonnay juices: A study by flavor precursor analysis. Am. J. Enol. Vitic.
44(4): 359-370 (1993).

Simpson, R.F. and G.C. Miller. Aroma composition of Chardonnay wine. Vitis.
23: 143-158 (1984).

Singleton, V. Grapes and wine phenolics; background and prospects. Proc.


Grape and Wine Centennial Symposium. A.D.Webb Ed. pp. 215-227.
University of California Press, Berkeley (1980).

Soles, R.M., Ough, C.S., and R.E. Kunkee. Ester concentration differences in
wine fermented by various species and strains of yeasts. Am. J. Enol. Vitic.
33(2): 94-98 (1982).

Stashenko, H., Macku, C, and T. Shibamato. Monitoring volatile chemicals


formed from must during yeast fermentation. J. Agric. Food Chem. 40: 2257-
2259 (1992).

Suomalainen, H. Yeast esterases and aroma esters in alcoholic beverages. J.


Inst. Brew. 87:296-300 (1981).

Suomalainen, H., and M. Lehtonen. The production of aroma compounds by


yeast. J. Inst. Brew. 85: 149-156 (1979).

Thomas, C.S, Boulton, R.B., Silacci, M.W., and W.D. Gubler. The effect of
elemental sulfur, yeast strain, and fermentation medium on hydrogen sulfide
production during fermentation. Am. J. Enol. Vitic. 44(2): 211-216 (1993).

Tracey, R.P, and T.J. Britz. Freon 11 extraction of volatile metabolites formed
by certain lactic acid bacteria. App. Env. Microbiol. 55(6): 1617-1623 (1989).
155

Watson, B.T. Jr., Dodd, E.E., and D.A. Heatherbell. Basic Chemical Analysis of
Musts and Wines. Table Wine Workshop. Department of Food Science &
Technology, Oregon State University, OR. May 16-17, (1980).

Williams, A.A., and G.M. Arnold. A comparison of the aromas of six coffees
characterized by conventional profiling, free-choice profiling and similarly
scaling methods. J. Sci. Food Agric. 36: 204-214 (1985).

Williams, A.A., and S.P. Langron. The use of Free-Choice Profiling for the
evaluation of commercial ports. J. Sci. Food Agric. 35: 558-568 (1984).

Williams, P.J., Strauss, C.R., and B. Wilson. Development in flavour research


on premium volatiles. In: Proceedings of the 2nd International Cool Climate
Viticulture and Oenology Symposium. Auckland, New Zealand (1988).

Wilson, B., Strauss, C.R., and P.J. Williams. The distribution of free and
glycosidically-bound monoterpenes among skin, juice and pulp fractions of
some white grape varieties. Am. J. Enol. Vitic. 37(2): 107-111 (1986).

Zoecklein, B.W., Fugelsang, K.C., Gump, B.H., and F.S. Nury. Production wine
analysis, pp. 475. AVI Book. Van Norstrand Reinhold, New York (1990).

You might also like