You are on page 1of 9

2258

International Journal of Food Science and Technology 2013, 48, 22582266

Original article
Mango wine aroma enhancement by pulp contact and
b-glucosidase
Xiao Li,1 Sien Long Lim,1 Bin Yu,2 Philip Curran2 & Shao-Quan Liu1,3*
1 Food Science and Technology Programme, Department of Chemistry, National University of Singapore, 3 Science Drive 3, Singapore City
117543, Singapore
2 Firmenich Asia Pte Ltd, Tuas 638377, Singapore
3 National University of Singapore (Suzhou) Research Institute, No. 377 Linquan Street, Suzhou Industrial Park, Suzhou, Jiangsu 215123,
China
(Received 6 March 2013; Accepted in revised form 11 May 2013)

Summary

This study determined the inuence of pulp maceration and b-glucosidase on mango wine physico-chemical
properties and volatile proles. The distinction in pH, sugars and organic acids among dierent treatments
was not statistically signicant. All wine samples reached around 8% (v/v) ethanol from about 16% (w/v)
of sugars. The wine with pulp contact contained about ten times higher a-terpinolene and up to three times
higher acetate esters than the wine without pulp contact, but mitigated the production of medium-chain
fatty acids and relative ethyl esters by up to six times. b-glucosidase enhanced terpenols by up to ten times
and acetate esters by up to three times. Furthermore, enzyme treatment mitigated, by up to ve times, the
formation of medium-chain fatty acids and ethyl esters to moderate levels. Sensory evaluation showed pulp
contact, and b-glucosidase not only improved the intensity and complexity of wine aroma but balanced
odour attributes.

Keywords

Flavour, juice, mango wine, pulp, volatiles, b-glucosidase.

Introduction

Many reports revealed that, besides free aromatic compounds, there are nonodorous and nonvolatile glucosides that represent an important source of fragrant
compounds in fruit juices and wines (Shoseyov et al.,
1988; Gueguen et al., 1996). The aglycone moieties of
glycosides include monoterpene hydrocarbons, terpenols, C13-norisoprenoids, benzene derivatives and
aliphatic alcohols, most of which possess pleasant oral and fruity aromas with low perception thresholds
(Palomo et al., 2005). Sarry & G
unata (2004) summarised that enzymatic and/or acid hydrolysis of glycosidically bound volatiles could allow the liberation of free
volatiles (aglycones) in dierent fruit juices or wines
such as grape, apple and passion fruits.
Acid hydrolysis of glycosides occurs very slowly and
may induce terpenol rearrangements (G
unata et al.,
1988). Enzymatic hydrolysis can be much more eective and specic in liberating aromatic compounds
from glycosides and minimise terpenol rearrangements
(G
unata et al., 1988). However, glycosidases from
fruits or Saccharomyces cerevisiae are not suciently
*Correspondent: Fax: +65 6775 7895; e-mail: chmLsq@nus.edu.sg

active because these enzymes show poor stability under


wine conditions (Palomo et al., 2005). In contrast,
glycosidases from fungi such as Aspergillus niger have
shown very good stability and activity in wine (Sarry
& G
unata, 2004).
The glycosidically bound volatiles such as terpenols
are frequently found in mango cultivars (Adedeji
et al., 1992; Drider et al., 1994; Sakho et al., 1997;
Lalel et al., 2003). These studies revealed an abundance of glycosidically bound volatiles, which included
monoterpene hydrocarbons, terpenols, alcohols, aldehydes, acids, esters, C13-norisoprenoids and benzene
derivatives in mangoes. Furthermore, Lalel et al.
(2003) demonstrated the role of b-glucosidase and
hemicellulase in the release of glycosidically bound
volatiles during ripening process in both pulp and skin
of the Kensington Pride mango. They discovered that
most glycosidically bound aroma compounds increased
in the pulp as maturity progressed. This suggests a
possible method for enhancement of mango wine
aroma by fermenting both juice and pulp with the use
of glycosidases. However, there is a lack of systematic
investigations on the application of exogenous glycosidases in mango wine. Therefore, the aim of this
study was to determine potent aroma compounds in

doi:10.1111/ijfs.12212
2013 The Authors. International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

Mango wine with and without pulp and b-glucosidase X. Li et al.

mango wine on the basis of odour activity values


(OAV) and the eect of application of exogenous
b-glucosidase on wine aroma composition. For further
enrichment of free aroma compounds and bound
aroma precursors (glycosides), mango pulp was not
discarded but included throughout the fermentation to
mimic the maceration step in grape wine fermentation.
The outcome of this study would aid further research
on pulp contact and ascertain the eectiveness of
b-glucosidase application in mango wine avour modulation. The ultimate aim was to enable mango wine
avour intensication, diversication and/or dierentiation so as to achieve unique mango wine style and
distinction.
Materials and methods

Enzymes

The commercial enzyme Novarom Blanc (Novozymes,


Bagsvaerd, Denmark) used in this study was chosen out
of three other enzyme preparations: Lallzyme Beta and
Lallzyme Cuvee Blanc (Lallemand SA, Blagnac,
France) and Macer8 W (Biocatalysts, Parc Nantgarw,
United Kingdom). This selection was based on an
enzyme activity test where Novarom Blanc (a commercial enzyme prepared from Aspergillus niger) exhibited a
consistently higher b-glucosidase activity under pH and
temperature ranges of pH 27 and 2070 C, respectively (data not shown). The Novarom Blanc enzyme
was mainly composed of maltodextrin-encapsulated
b-glucosidase. The dosage of the enzyme used in this
experiment was 0.4 g L 1 to accelerate the enzymatic
process (vs. the recommended dosage of 0.1 g L 1 for
about 23 weeks).
Reagents and chemicals

Food-grade DL-malic acid (>97%) was purchased


from Suntop Ltd. (Singapore). Potassium metabisulte
(>97%) was bought from Goodlife Homebrew centre
(Norfolk, England). Standards of glucose, fructose,
sucrose, tartaric, citric, malic, succinic, lactic and pyruvic acids (>99%) were purchased from Sigma-Aldrich
(St. Louis, MO, USA); glycerol (>99%) was bought
from Merck (Darmstadt, Germany). Firmenich Asia
(Singapore) provided the following standards: hexanoic, octanoic, decanoic and dodecanoic acids; ethanol; isobutyl alcohol; active amyl alcohol; isoamyl
alcohol; (Z)-3-hexenol; 2-phenylethyl alcohol; linalool;
a-terpineol; b-citronellol; nerol; geraniol; ethyl acetate;
isobutyl acetate isoamyl acetate; (Z)-3-hexenyl acetate;
citronellyl acetate; 2-phenylethyl acetate; ethyl hexanoate; ethyl octanoate; ethyl decanoate; ethyl dodecanoate; and a-terpinolene (>98%). Ethanol (>99%) was
bought from Fisher Chemical, Singapore.

Preparation of mango juice

Chok Anan mangoes from Malaysia were purchased


from a local market in Singapore and kept at 25 C
until fully ripened before use. The ripened mangoes
were juiced and divided into two lots. The rst lot was
centrifuged at 41 415 g and 4 C for 10 min to remove
the pulp (Beckman Coulter Allegra 64R Centrifuge,
Brea, CA, USA). The second lot was not centrifuged.
The sugar concentration of centrifuged and uncentrifuged juice was 158 and 165 g L 1, respectively.
Preculture media prepared from the mango juice were
sterilised at 100 C for 3 min and cooled to room temperature. The media were inoculated with 10% (v/v) of
S. cerevisiae MERIT.ferm (Chr.-Han., Horsholm, Denmark, maintained in nutrient broth) and incubated for
48 h at 25 C until yeasts grew to at least 107 cfu mL 1.
The mango juice used for fermentation was adjusted to
pH = 3.5 with food-grade 50% (w/v) DL-malic acid
and was treated with 100 ppm (w/v) potassium metabisulphite and then heated at 60 C for 15 min and
cooled to room temperature.
Fermentation and enzyme treatment

Triplicate mango juice fermentations were carried out


in 500-mL sterile Erlenmeyer conical asks (plugged
with cotton wool and then wrapped with aluminium
foil), and each ask contained 450 mL of centrifuged
mango juice or uncentrifuged pulpy mango juice. All
samples were inoculated with 1% (v/v) preculture of
S. cerevisiae MERIT.ferm, and the fermentation was
conducted at 20 C statically for 10 days. The samples
were divided into two 225-mL portions at the end of
fermentation. One portion was subjected to enzyme
treatment (enzyme dosage of 0.4 g L 1) and incubated
at 20 C for another 4 days. The other portion was
kept as control and incubated at 20 C. All enzyme
treatments and controls were carried out in triplicate.
Sampling was done at the end of enzyme treatment.
Analysis of sugars and organic acids

Cell-free samples were obtained by centrifugation and


ltration through a 0.2-lm regenerated cellulose lter
membrane (Sartorius Stedim Biotech, Goettingen, Germany). Sugars were measured by Shimadzu ultra-fast
liquid chromatography (UFLC, Shimadzu Asia Pacic, Singapore) according to the method of Li et al.
(2012) using Agilent Zorbax carbohydrate column
(150 9 4.6 mm, 5 lm, Santa Clara, CA, USA) connected to an ELSD-LT detector. External standards of
glucose, fructose, sucrose and glycerol were used for
identication and quantication. Organic acids were
also determined by Shimadzu UFLC according to
Li et al. (2012) using a Supelcogel C-610H column

2013 The Authors


International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

International Journal of Food Science and Technology 2013

2259

2260

Mango wine with and without pulp and b-glucosidase X. Li et al.

(7.8 9 300 mm, 9 lm, Bellefonte, PA, USA) connected to a photodiode array detector. Identication
and quantication were done by comparing the retention time and peak area with those obtained from
external standards of tartaric, citric, malic, succinic,
lactic and pyruvic acids at 210 nm. All samples were
analysed in duplicate.
Analysis of volatiles

Volatiles were analysed using a headspace (HS) solidphase microextraction (SPME) method coupled with
gas chromatography (GC)mass spectrometer (MS)
and ame ionisation detector (FID). The system was
Agilent (Palo Alto, CA, USA) 6890 N network gas
chromatography system equipped with a DB-FFAP
capillary column (60 m 9 0.25 mm 9 0.25 lm, Agilent, Woodbridge, USA), 5975 inert mass selective
detector (MSD) and FID. Adsorption of volatiles was
carried out with an 85 lm carboxenpolydimethylsiloxane bre (Supelco, Bellefonte, PA, USA). Details of
analysis were described in Li et al. (2012). The volatiles
were identied using Chemstation integrated with Wiley mass spectral library. The linear retention indices
(LRI) of the compounds were used to express their
retention times on the gas chromatographic column relative to a homologous series of n-alkanes. For semiquantication of the volatiles, FID peak area was used
to represent the amount of each volatile. A total of
twenty-six major volatiles were quantied by comparison with available external standards. The method was
based on Chen et al. (2006). All standards were dissolved in 10% v/v diluted mango juice with water,
except for ethanol that was dissolved in 100% v/v
mango juice. The same HS-SPME-GC-MS/FID condition was used for quantitative analysis, and good linearity was obtained for all standard curves (R2 > 0.98).
All samples were analysed in duplicate. Thereafter,
odour activity values (OAVs) of quantied volatiles
were calculated according to their known threshold levels in the literature (Guth, 1997; Lambrechts & Pretorius, 2000; Yamamoto et al., 2004; Zemni et al., 2007;
Bartowsky & Pretorius, 2008).
Sensory analysis

Four wine samples were assessed by a panel of seven


trained avourists from Firmenich Asia Pte. Ltd., Singapore. The four wine samples were assessed in the
following order: (i) non-enzyme-treated wine from centrifuged juice; (ii) non-enzyme-treated wine from noncentrifuged pulpy juice; (iii) enzyme-treated wine from
centrifuged juice; and (iv) enzyme-treated wine from
noncentrifuged pulpy juice. A set of descriptive terms
for nine attributes were rated on a ve-point scale for
the intensity perceived, where zero indicated that the

International Journal of Food Science and Technology 2013

descriptor was not perceived and ve indicated a very


high intensity.
Statistical analysis

The statistical dierences of the eect of enzyme treatment on the volatiles of mango wine fermented with
and without pulp were evaluated using analysis of
variance (ANOVA). All tests of signicance were conducted at a probability level of P < 0.05. Means and
standard deviations were obtained from triplicate fermentation samples. The volatile and aroma proles for
enzyme-treated mango wines and control were further
analysed using principal component analysis (PCA) to
characterise the multidimensional data.
Results and discussion

Physico-chemical properties of mango wine with


maceration and enzyme treatment

The pH increased slightly in macerated (with pulp


contact) wines and viable cell count reached about
7 9 107 8 9 107 cfu mL 1 (Table 1) for all four
wines. The dierence of total soluble solids was not
signicant in all four treatments (Table 1), but the
Brix value was slightly higher for the macerated wine
compared with the nonmacerated for both enzyme and
nonenzyme treatments, which is logically due to a
higher content of pectins and other soluble solids
(Rouse et al., 1974). In addition, there was no signicant dierence for the organic acids between the control and enzyme-treated wine (Table 1), and lactic acid
was not detected in all samples. The residual sucrose
and fructose content was 0.080.19 g L 1 higher in the
enzyme-treated wines compared with the controls. The
higher sugar content could be due to b-glucosidase
and pectinase activity from the enzyme treatment,
resulting in the hydrolysis and release of glycosidically
bound saccharides into the solution (Joshi et al., 2011;
Pei et al., 2012). The higher sugar content could also
lead to signicantly higher glycerol level in the
enzyme-treated wine, which was shown in Table 1,
and this is probably because high osmolar concentrations of solutes (e.g. sugar) would enhance glycerol
synthesis to counteract cell dehydration (Lopes et al.,
2000). Glycerol does not contribute to the aroma of
the wine due to its nonvolatile nature but does contribute to sweetness, smoothness and viscosity of wine
(Eustace & Thornton, 1987). The typical concentration
of glycerol in wine is from 1 to 15 g L 1 (Scanes et al.,
1998). The threshold taste level of glycerol is observed
at 5.2 g L 1 in wine (Noble & Bursick, 1984). Glycerol
in macerated mango wine with enzyme addition was
remarkably improved, and its concentration was
higher than the reported threshold level (Table 1). In

2013 The Authors


International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

Mango wine with and without pulp and b-glucosidase X. Li et al.

Table 1 Physico-chemical properties, alcohol, organic acid and sugar concentrations of nonmacerated and macerated mango wine, with and
without enzyme treatment
Control
Fermentation medium
Physico-chemical properties
Ph
Total soluble solids (Brix)
Cell count (106 cfu mL 1)
Alcohol
Ethanol (% v/v)
Glycerol (g L 1)
Organic acids (g L 1)
Citric acid
Tartaric acid
Malic acid
Pyruvic acid
Succinic acid
Reducing sugars (g L 1)
Fructose
Glucose
Sucrose

Enzyme

Nonmacerated

Macerated

Nonmacerated

Macerated

3.57  0.00ab
5.58  0.14ab
89.75  28.64a

3.70  0.06c
5.98  0.13a
73.88  24.57a

3.50  0.01a
5.09  0.16b
79.75  28.64a

3.64  0.01bc
5.87  0.13a
85.13  6.19a

7.87  0.83a
4.62  0.11a

8.08  0.41a
5.33  0.30ab

8.49  0.31a
5.94  0.21b

8.14  0.47a
8.46  0.52c

2.21
0.28
3.81
0.10
1.21







0.30a
0.07a
0.41a
0.01a
0.07a

N. D.
0.36  0.00a
0.39  0.01a

1.72
0.25
3.13
0.05
1.33







0.02b
0.01a
0.45b
0.00b
0.08a

1.95
0.26
3.62
0.08
1.11







0.04ab
0.01a
0.35ab
0.00c
0.05a

0.17  0.00a
0.45  0.02a
0.47  0.02ab

N. D.
0.48  0.01a
0.39  0.05ab

1.93
0.27
3.27
0.03
1.19







0.03ab
0.01a
0.01ab
0.00d
0.04a

0.19  0.00b
0.49  0.01a
0.47  0.03b

ANOVA (n = 6) at 95% confidence level with same letter indicating no significant difference.
N.D.: not detected,  standard deviation.

abcd

addition, the nal ethanol reached 7.878.49% (v/v)


for four wines and was slightly higher in the enzymetreated wines (Table 1); however, this dierence was
not signicant. The slight dierence could be due to
more sugars being released from glycoside hydrolysis
in enzyme-treated wine, or its higher fermentation
activity suggested by higher viability of cells (Table 1).
Volatiles of mango wine with pulp contact and enzyme
treatment
Terpenols and monoterpene hydrocarbons

Few terpenols in the control wine were at levels above


their odour thresholds and thus scarcely contribute to
the overall aroma (Table 2). Enzyme treatment signicantly enhanced the concentrations of free terpenols
(Table 2), but pulp contact only facilitated enrichment
of b-citronellol and nerol but not all terpenols. This
result was slightly dierent from the report of Cabaroglu et al. (2003). This was probably due to some extent
of interconversion of terpenols catalysed by Saccharomyces cerevisiae as summarised by King & Dickinson
(2000). The concentration of geraniol increased signicantly in both nonmacerated and macerated wine with
enzyme addition, achieving an OAV of 14 and 8,
respectively. In addition, enzyme treatment increased
b-citronellol to 0.1 and 0.23 mg L 1 in the nonmacerated and macerated wines, respectively, which were at
the level of or higher than its odour threshold
(Table 2). On the other hand, linalool was found to

have an OAV of 12 and 7.3 in both nonmacerated and


macerated wines, respectively, with enzyme treatment.
The concentrations of a-terpineol and nerol were also
enhanced by enzyme treatment but were still below
their threshold levels. The result was consistent with
the results obtained in fruit juices (Gueguen et al.,
1996) and grape wine (Rogerson et al., 1999; Cabaroglu et al., 2003). Terpenols such as geraniol, linalool
and citronellol have been found to be important in the
contribution of oral and citrus notes to wine especially those from aromatic grape varieties such as
Muscat and Riesling (Marais, 1984; Swiegers et al.,
2005).
Monoterpene hydrocarbons (C10H16, most importantly a-terpinolene in Chok Anan mango), the key
volatile constituents of the fresh mango aroma, are largely formed in the mango during ripening. These
compounds signicantly contributed to typical fresh
mango aroma. Unfortunately, monoterpene hydrocarbons were signicantly lost throughout the fermentation (Li et al., 2011, 2012) possibly due to their
volatility. In this experiment, pulp contact strongly
retained the level of a-terpinolene (Table 2), but
surprisingly, its concentration decreased with enzyme
treatment. Although other important monoterpene
hydrocarbons such as d-3-carene, a-phellandrene,
a-terpinene, limonene and sabinene were not quantied
due to a lack of standard compounds, they also showed
highest FID peak areas in the macerated control sample relative to the enzyme-treated sample. The decrease

2013 The Authors


International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

International Journal of Food Science and Technology 2013

2261

Acetic acid
Hexanoic acid
Octanoic acid
Decanoic acid
Dodecanoic acid
Isobutyl alcohol
Active amyl alcohol
Isoamyl alcohol
(Z)-3-Hexenol
2-Phenylethyl alcohol
Linalool
a-Terpineol
b-Citronellol
Nerol
Geraniol
Ethyl acetate
Isobutyl acetate
Isoamyl acetate
(Z)-3-Hexenyl acetate
Citronellyl acetate
2-Phenylethyl acetate
Ethyl hexanoate
Ethyl octanoate
Ethyl decanoate
Ethyl dodecanoate
a-Terpinolene

Acid

International Journal of Food Science and Technology 2013


1408
1780
1990
2188
2393
981
1173
1173
1350
1944
1489
1713
1770
1833
1834
858
979
1061
1206
1595
1773
1158
1374
1575
1785

LRI
0.31
2.70
0.71
0.09
0.07
0.98
0.15
5.57
1.50
5.38
0.00
0.06
0.10
0.00
0.00
0.02
0.04
23
6.25
0.00
9.40
24
1695
56
5.11

61.1 
0.81 
6.21 
1.32 
0.74 
39.3 
9.91 
167 
0.60 
53.8 
N. D.
0.02 
0.01 
N. D.
N. D.
0.13 
0.07 
0.07 
0.05 
N. D
2.35 
0.34 
3.39 
11.2 
6.13 
0.036 
0.51a
0.04a
0.54a
1.7a
0.25a
0.000a

0.02a
0.01a
0.02a
0.00a

0.00a
0.00a

6.5a
0.09a
0.53a
0.35a
0.11a
7.2a
1.82a
17.1a
0.07a
8.0a

OAV*

Nonmacerated
135 
0.39 
1.81 
0.24 
0.39 
67.1 
18.5 
182 
1.05 
76.2 
0.02 
0.06 
0.05 
N. D.
N. D.
0.17 
0.10 
0.13 
0.03 
N. D.
1.85 
0.18 
1.10 
2.3 
1.25 
0.488 
0.27a
0.03b
0.02b
0.5b
0.20b
0.012b

0.02b
0.02ab
0.02b
0.00b

12.4b
0.09b
0.21b
0.02b
0.07b
4.5b
2.31b
6.3a
0.08b
8.6b
0.00a
0.01b
0.00b

Macerated
0.68
1.30
0.21
0.02
0.04
1.68
0.28
6.07
2.63
7.62
1.33
0.18
0.50
0.00
0.00
0.02
0.06
43
3.75
0.00
7.40
13
550
12
1.04

OAV
75.2
0.71
4.40
0.89
0.15
45.5
15.5
192
0.96
58.5
0.18
0.12
0.10
0.03
0.42
0.37
0.10
0.17
0.06
0.005
2.13
0.62
1.75
1.9
0.63
0.008



























8.5a
0.04a
0.23c
0.00c
0.04c
6.3a
1.75b
22.7a
0.09b
7.3a
0.04b
0.03c
0.02c
0.01a
0.04a
0.03a
0.01ab
0.03bc
0.01a
0.001a
0.47a
0.04c
0.04c
0.4b
0.08c
0.001c

Nonmacerated

Enzyme (mg L 1)

N.D., not detected; LRI, Linear retention index.


abcd
ANOVA (n = 6) at 95% confidence level with same letter indicating no significant difference.
*Odour activity values (OAV) were calculated by dividing concentration by the odour threshold value of the compound.

The odour threshold obtained from Guth (1997).

The odour threshold obtained from Bartowsky & Pretorius (2008).

The odour threshold obtained from Lambrechts & Pretorius (2000).

The odour threshold obtained from Zemni et al. (2007).


**The odour threshold obtained from Yamamoto et al. (2004).

The odour threshold obtained from Pino & Queris (2011).

Monoterpene hydrocarbon

Ethyl ester

Acetate ester

Terpenol

Alcohol

Compound

Group

Control (mg L 1)

0.38
2.37
0.50
0.06
0.02
1.14
0.24
6.40
2.40
5.85
12.0
0.36
1.00
0.08
14.0
0.05
0.06
57
7.50
0.02
8.52
44
875
9.5
0.53

OAV
244
0.30
1.46
0.12
0.09
71.2
22.1
204
0.96
83.8
0.11
0.10
0.23
0.07
0.24
0.41
0.14
0.21
0.04
0.017
1.95
0.26
0.49
0.73
0.33
0.030



























9.1c
0.06b
0.13b
0.03d
0.02c
7.1b
5.28b
16.8a
0.07b
8.4b
0.03b
0.01c
0.03d
0.01b
0.09b
0.03b
0.03b
0.05c
0.01ab
0.003b
0.21a
0.01ab
0.01d
0.02c
0.06d
0.001a

Macerated

Table 2 Major volatiles (mg L 1) and their odour activity values (OAVs) for nonmacerated and macerated wine, with and without enzyme treatment

1.22
1.00
0.17
0.01
0.01
1.78
0.34
6.80
2.40
8.38
7.33
0.30
2.30
0.18
8.00
0.05
0.09
70
5.00
0.07
7.80
19
245
3.7
0.28

OAV

200
0.3
8.8
15
10
40
65
30
0.4
10
0.015
0.33
0.1
0.4
0.03
7.5
1.6
0.003
0.008
0.25**
0.25
0.014
0.002
0.2
1.2
Not applicable

Odour
threshold (mg L 1)

2262
Mango wine with and without pulp and b-glucosidase X. Li et al.

International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

2013 The Authors

Mango wine with and without pulp and b-glucosidase X. Li et al.

in monoterpene hydrocarbons as a result of enzyme


treatment was possibly due to monoterpene hydrocarbons being bound by the enzyme protein via hydrophobic interactions or being trapped by maltodextrin (the
carrier substance for b-glucosidase in the Novarom
Blanc enzyme preparation). The nonpolar monoterpene
hydrocarbons became less volatile when interacting
with protein and/or maltodextrin (Misharina, 2010).
Alcohols

(Z)-3-Hexenol was considered as an important aglycone in fresh mango juice, and it could be eciently
released by either exogenous glycosidase or glycosidase
from fruits (Sakho et al., 1997; Lalel et al., 2003).
(Z)-3-Hexenol has also been reported to be present
above their odour threshold values in some mango
cultivars, contributing to the oralcitrus odour and
green grassy odour (Pino & Mesa, 2006). In this study,
(Z)-3-Hexenol, which was already present above its
odour threshold in both wines, was signicantly
increased in the nonmacerated wine after the enzyme
treatment. This was consistent with the result in other
reports (Sakho et al., 1997; Lalel et al., 2003). However, there was no statistically signicant dierence
between the concentrations of other alcohols (except
active amyl alcohol) after enzyme treatment, but the
macerated wine was higher in fusel alcohols, which
could be related to the higher availability of relative
amino acids in pulpy juice.
Esters and fatty acids

Esters are the principal odourants in mango wine due


to their high OAVs (Table 2). The esters are mainly of
two types: ethyl esters of medium-chain fatty acids
and acetate esters of higher alcohols, and both types
are formed enzymatically or chemically during fermentation, but chemical formation is very slow and can be
ignored (Lambrechts & Pretorius, 2000). Major acetate
esters such as isoamyl acetate and 2-phenylethyl acetate give banana-like and rose-like aromas, respectively. Major ethyl esters such as ethyl hexanoate and
ethyl octanoate engender an apple-like character
(Rojas et al., 2003). However, an overproduction of
acetate esters and ethyl esters may lead to nail polisher
and soapy o-avours (Lambrechts & Pretorius,
2000).
Compared with non-Saccharomyces yeasts such as
Hanseniaspora and Pichia (Rojas et al., 2001), the production of acetate esters is generally moderate in
Saccharomyces yeasts. However, Saccharomyces yeasts
are able to vigorously produce ethyl esters (Li et al.,
2011), which is also shown in Table 2. According to
Saerens et al. (2008), production of ethyl esters is positively correlated with the production of respective fatty
acids, and this is consistent with our results. Enzyme
addition (except for ethyl hexanoate) or pulp contact

seemed to induce decreased production of fatty acids,


which correspondingly led to more moderate production of ethyl esters (Table 2). Nonetheless, the detailed
mechanism involved remains to be elucidated.
Minor volatile compounds

Some quantitatively minor volatile compounds were


identied, such as (E)-b-damascenone, p-cymen-8-ol,
nerol oxide, geranyl ether, and 1,8-menthadien-4-ol.
They could not be quantied because the reference
compounds are hardly commercially available, but the
FID peak area of these compounds showed that they
were signicantly enhanced by b-glucosidase treatment
[FID chromatogram of (E)-b-damascenone was shown
in Fig. 1]. More importantly, some of these volatiles
are potent odour-active compounds such as
(E)-b-damascenone, which has a owery, quince-like
odour, and its odour threshold is 2 ng L 1 in water
and 50 ng L 1 in 10% alcoholic solution (Cabaroglu
et al., 2003).
Sensory characteristics of mango wine

The sensory proles of the mango wines were represented in a spider web plot as shown in Fig. 2. From
the analysis of variance (ANOVA), no signicant dierences were found for all sensory attributes except for
yeasty attribute. This was likely due to high standard
deviations, which were probably caused by variations
of panellists sensitivity to dierent aroma attributes,
small panellist size of seven and a small ve-point
scale.
The control samples for nonmacerated wine were
signicantly more yeasty than the rest. This was probably related to its higher concentration of ethyl esters
and fatty acids. Yeasty note has been used to characterise the complexity of sparkling wine (Torrens
et al., 2010), but this aroma attribute has been determined to be unfavourable in other alcoholic drinks
such as Tannat grape wines (Varela & G
ambaro,
2006), and it may aect the fruitiness of wine. Also,
the sensory result concomitantly showed that pulp
contact and enzyme treatment mitigated the protruding yeasty note in control sample. In addition, the
terpenic or woody aroma here referred to the characteristic aromas associated with fresh mango. The macerated wines were perceived as more terpenic
compared with the nonmacerated wines, where the
macerated wine without enzyme treatment achieved
the highest score, and this was consistent with its highest concentrations of a-terpinolene (Table 2). Macerated wines were also considered as winey and this
correlated with their higher concentration of higher
alcohols. The enzyme-treated wines had higher scores
for rosy (oral) and fruity attributes compared with
those without enzyme treatment (Fig. 2). The rosy and

2013 The Authors


International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

International Journal of Food Science and Technology 2013

2263

Mango wine with and without pulp and b-glucosidase X. Li et al.

Abundance

215 000
210 000
205 000
200 000
195 000
190 000
185 000
180 000
175 000
170 000
165 000
160 000
155 000
150 000
145 000
140 000
135 000
130 000

Signal: Juice A Enz.D\FID1A.ch


Signal: Juice A Ctrl.D\FID1A.ch
Signal: Pulp A Ctrl.D\FID1A.ch
Signal: Pulp A Enz.D\FID1A.ch

(E)--damascenone

Macerated with Enzyme


Non-macerated with Enzyme
Macerated control
Non-macerated control

Time--> 31.20

31.21

31.22

31.23

31.24

31.25

31.26

31.27

Figure 1 Flame ionisation detector (FID) chromatogram of (E)-b-damascenone in nonmacerated control, macerated wine control, nonmacerated with enzyme and macerated wine with enzyme.
0.4

Green
5
Acidic

21

0.2

JEnz

2
4 3 25

2
Yeasty

15
11

20

0.3

Fruity

Waxy

22
0

18

JCtrl
10

23
26

0.1

Sweet

Terpenic/Woody

PEnz

1
5

16
7

0.2

PCtrl
27

0.3

Winey

12
17
14 9
19
2413

0.1
PC 2 (25%)

2264

Rosy/Floral

Figure 2 Aroma prole of mango wines: nonmacerated control


(), macerated wine control (), nonmacerated with enzyme (),
and macerated wine with enzyme (9).

fruity notes were probably ascribed to the higher level


of short-chain acetate esters such as isobutyl acetate
and isoamyl acetate, as well as terpenols such as linalool, geraniol and b-citronellol (Table 2).
Principal component analysis of volatiles in mango wine

Principal component analysis (PCA) was applied to


discriminate the volatile proles of the mango wines
(Fig. 3). The rst principal component (PC1)
accounted for 62.7% of the total variance in the data
set, while the second principal component (PC2)

International Journal of Food Science and Technology 2013

0.4
0.4

0.3

0.2

0.1

0
0.1
PC 1 (64%)

0.2

0.3

0.4

Figure 3 Biplot of principal component analysis of mango wines:


nonmacerated control (JCtrl, ), macerated wine control (PCtrl, ),
enzyme-treated nonmacerated wine (JEnz, ) and enzyme-treated
macerated wine (PEnz, 9): (1) acetic acid; (2) hexanoic acid; (3)
octanoic acid; (4) decanoic acid; (5) dodecanoic acid; (6) ethanol; (7)
isobutyl alcohol; (8) active amyl alcohol; (9) isoamyl alcohol; (10)
(Z)-3-hexenol; (11) linalool; (12) a-terpineol; (13) b-citronellol; (14)
nerol; (15) geraniol; (16) 2-phenylethyl alcohol; (17) ethyl acetate;
(18) isobutyl acetate; (19) isoamyl acetate; (20) (Z)-3-hexenyl acetate;
(21) ethyl hexanoate; (22) ethyl octanoate; (23) ethyl decanoate; (24)
citronellyl acetate; (25) 2-phenylethyl acetate; (26) ethyl dodecanoate;
(27) a-terpinolene.

accounted for 23.2% of the total variance. The PCA


biplot separated the twenty-six dierent volatile compounds and the four dierent samples.

2013 The Authors


International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

Mango wine with and without pulp and b-glucosidase X. Li et al.

The rst principal component (PC1) separated the


nonmacerated control wine from other samples based
on the higher concentrations of medium-chain fatty
acids (octanoic, decanoic and dodecanoic acids), medium-chain ethyl esters (ethyl esters of octanoic, decanoate and dodecanoate) 2-phenylethyl acetate. PC1 also
separated the macerated wine with enzyme addition
from other samples due to higher concentrations of
acetic acid, most of higher alcohols, short-chain acetate esters such as isoamyl acetate as well as some
terpenols and their acetate esters such as b-citronellol
and citronellyl acetate. Furthermore, PC2 separated
the nonmacerated wine with enzyme addition and
macerated wine without enzyme addition based on
higher concentrations of ethanol and isoamyl alcohol
in the former and higher a-terpinolene in the latter.
The compounds that correlated with the nonmacerated wine were those imparting waxy, soapy, fatty, oral, green and fruity notes. The macerated wines were
associated with compounds with winey, fruity, terpenic, citrus and oral odours, but also with the acidic
odour. Moreover, the enzyme addition generally
further intensied the aroma dierence between nonmacerated and macerated wines.
Conclusion

This work assessed the use of exogenous b-glucosidase


and pulp contact in mango wine for enhancing avour.
Pulp contact could enhance monoterpene hydrocarbons, higher alcohols and acetate esters. The addition
of b-glucosidase could accelerate the release of odouractive volatiles such as terpenols, acetate esters, benzene derivatives and C13-norisoprenoids. Furthermore,
both b-glucosidase addition and pulp contact could
mitigate production of excess fatty acids and their
ethyl esters (yeasty note in sensory test). The application of b-glucosidase with mango pulp contact was
eective in intensication, diversication and also
balancing of mango wine aroma prole.
Acknowledgment

This work was supported, in part, by an ARF grant


from Ministry of Education of Singapore (WBS No.
R-143-000-507-112).
References
Adedeji, J., Hartman, T.G., Lech, J. & Ho, C.T. (1992). Characterization of glycosidically bound aroma compounds in the African
mango (Mangifera indica L.). Journal of Agricultural and Food
Chemistry, 40, 659661.
Bartowsky, E.J. & Pretorius, I.S. (2008). Microbial formation and
modication of avor and o-avor compounds in wine. In:
Biology of Microorganisms on Grapes, in Must and in Wine (Edited

by H. Konig, G. Uden & J. Frohlich). Pp. 215217. Berlin, Germany: Springer-Verlag.


Cabaroglu, T., Selli, S., Canbas, A., Lepoutre, J.-P. & G
unata, Z.
(2003). Wine avor and enhancement through the use of exogenous
fungal glycosidases. Enzyme and Microbial Technology, 33,
581587.
Chen, Y., Begnaud, F., Chaintreau, A. & Pawliszyn, J. (2006).
Quantication of perfume compounds in shampoo using solidphase microextraction. Flavour and Fragrance Journal, 21, 822832.
Drider, D., Janbon, G., Chemardin, P., Arnaud, A. & Galzy, P.
(1994). Enzymatic hydrolysis of monoterpene glycosides of passion
fruit and mango with a b-glucosidase from yeast. Bioresource Technology, 49, 243246.
Eustace, R. & Thornton, R.J. (1987). Selective hybridization of wine
yeasts for higher yields of glycerol. Canadian Journal of Microbiology, 33, 112117.
Gueguen, Y., Chemardin, P., Janbon, G., Arnaud, A. & Galzy, P.
(1996). A very ecient b-glucosidase catalyst for the hydrolysis of
avor precursors of wines and fruit Juices. Journal of Agricultural
and Food Chemistry, 44, 23362340.
G
unata, Y.Z., Bitteur, S.M., Brillouet, J.M., Bayonove, C.L. & Cordonnier, R.E. (1988). Sequential enzymic hydrolysis of potentially
aromatic glycosides from grape. Carbohydrate Research, 184,
139149.
Guth, H. (1997). Quantitation and sensory studies of character
impact odorants of dierent white wine varieties. Journal of Agricultural and Food Chemistry, 45, 30273032.
Joshi, V.K., Parmar, M. & Rana, N. (2011). Purication and characterization of pectinase produced from apple pomace and evaluation
of its eciency in fruit juice extraction and clarication. Indian
Journal of Natural Products and Resources, 2, 189197.
King, A. & Dickinson, J.R. (2000). Biotransformation of monoterpene alcohols by Saccharomyces cerevisiae, Torulaspora delbrueckii
and Kluyveromyces lactis. Yeast, 16, 499506.
Lalel, H.D.J., Singh, Z. & Tan, S.C. (2003). Glycosidically-bound
aroma volatile compounds in the skin and pulp of Kensington
Pride mango fruit at dierent stages of maturity. Postharvest Biology and Technology, 29, 205218.
Lambrechts, M.G. & Pretorius, I.S. (2000). Yeast and its importance
to wine aroma a review. South African Journal of Enology and
Viticulture, 21, 97129.
Li, X., Yu, B., Curran, P. & Liu, S.-Q. (2011). Chemical and volatile
composition of mango wines fermented with dierent Saccharomyces cerevisiae strains. South African Journal of Enology and Viticulture, 32, 117128.
Li, X., Chan, L.J., Yu, B., Curran, P. & Liu, S.-Q. (2012). Fermentation of three varieties of mango juices with a mixture of Saccharomyces cerevisiae and Williopsis saturnus var. mrakii. International
Journal of Food Microbiology, 158, 2835.
Lopes, M.B., Rehman, A., Gockowiak, H., Heinrich, A., Langridge,
P. & Henschke, P.A. (2000). Fermentation properties of a wine
yeast over-expressing the Saccharomyces cerevisiae glycerol 3-phosphate dehydrogenase gene. Australian Journal of Grape and Wine
Research, 6, 208215.
Marais, J. (1984). Terpenes in the aroma of grapes and wines: a
review. South African Journal of Enology and Viticulture, 4, 4956.
Misharina, T. (2010). Headspace analysis of aroma compounds using
porus adsorbents. Chemistry and Chemical Technology, 5, 347354.
Noble, A.C. & Bursick, G.F. (1984). The contribution of glycerol to
perceived viscosity and sweetness in white wines. American Journal
of Enology and Viticulture, 35, 110112.
 & Pereznas, M.A.
Palomo, E.S., Hidalgo, M.C.D.-M., Gonzalez-Vi~
Coello, M.S. (2005). Aroma enhancement in wines from dierent
grape varieties using exogenous glycosidases. Food Chemistry, 92,
627635.
Pei, J., Pang, Q., Zhao, L., Fan, S. & Shi, H. (2012). Thermoanaerobacterium thermosaccharolyticum b-glucosidase: a glucose-tolerant

2013 The Authors


International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

International Journal of Food Science and Technology 2013

2265

2266

Mango wine with and without pulp and b-glucosidase X. Li et al.

enzyme with high specic activity for cellobiose. Biotechnology for


Biofuels, 5, 3140.
Pino, J.A. & Mesa, J. (2006). Contribution of volatile compounds to
mango (Mangifera indica L.) aroma. Flavour and Fragrance Journal, 21, 207213.
Pino, J.A. & Queris, O. (2011). Analysis of volatile compounds of
mango wine. Food Chemistry, 125, 11411146.
Rogerson, F.S.S., Grande, H.J. & Silva, M.C.M. (1999). Free and
enzyme enhanced monoterpenol content of Portuguese red wines
from the Douro. Ci^
encia e Tecnologia de Alimentos, 2, 169173.
Rojas, V., Gil, J.V., Pi~
naga, F. & Manzanares, P. (2001). Studies on
acetate ester production by non-Saccharomyces wine yeasts. International Journal of Food Microbiology, 70, 283289.
Rojas, V., Gil, J.V., Pi~
naga, F. & Manzanares, P. (2003). Acetate
ester formation in wine by mixed cultures in laboratory fermentations. International Journal of Food Microbiology, 86, 181188.
Rouse, A.H., Albrigo, L.G., Huggart, R.L. & Moore, E.L. (1974).
Viscometric measurement and pectic content of frozen concentrated orange juices for citrus futures. Proceedings of the Florida
State Horticultural Society, 87, 293296.
Saerens, S.M.G., Delvaux, F., Verstrepen, K.J., Van Dijck, P.,
Thevelein, J.M. & Delvaux, F.R. (2008). Parameters aecting ethyl
ester production by Saccharomyces cerevisiae during fermentation.
Applied and Environmental Microbiology, 74, 454461.
Sakho, M., Chassagne, D. & Crouzet, J. (1997). African mango
glycosidically bound volatile compounds. Journal of Agricultural
and Food Chemistry, 45, 883888.

International Journal of Food Science and Technology 2013

Sarry, J.-E. & G


unata, Z. (2004). Plant and microbial glycoside
hydrolase: volatile release from glycosidic aroma precursors. Food
Chemistry, 87, 502521.
Scanes, K.T., Hohmann, S. & Priori, B.A. (1998). Glycerol production
by the yeast Saccharomyces cerevisiae and its relevance to wine: a
review. South African Journal of Enology and Viticulture, 19, 1724.
Shoseyov, O., Bravdo, B.-A., Ikan, R. & Chet, I. (1988). Endo-bglucosidase from Aspergillus niger grown on a monoterpene glycoside-containing medium. Phytochemistry, 27, 19731976.
Swiegers, J.H., Bartowsky, E.J., Henschke, P.A. & Pretorius, I.S.
(2005). Yeast and bacterial modulation of wine aroma and avour.
Australian Journal of Grape and Wine Research, 11, 139173.
Torrens, J., Rlu-Aumatell, M., Vichi, S., L
opez-Tamames, E. &
Buxaderas, S. (2010). Assessment of volatile and sensory proles
between base and sparkling wines. Journal of Agricultural and Food
Chemistry, 58, 24552461.
Varela, P. & Gambaro, A. (2006). Sensory descriptive analysis of
Uruguayan Tannat wine: correlation to quality assessment. Journal
of Sensory Studies, 21, 203217.
Yamamoto, T., Shimada, A., Ohmoto, T., Matsuda, H., Ogura, M.
& Kanisawa, T. (2004). Olfactory study on optically active citronellyl derivatives. Flavour and Fragrance Journal, 19, 121133.
Zemni, H., Souid, I., Fathalli, N. et al. (2007). Aromatic composition
of two Muscat grape cultivars cultivated in two dierent regions of
Tunisia. International Journal of Fruit Science, 7, 97112.

2013 The Authors


International Journal of Food Science and Technology 2013 Institute of Food Science and Technology

You might also like