You are on page 1of 10

Food Chemistry 141 (2013) 3827–3836

Contents lists available at SciVerse ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

Changes in nutritional and sensory properties of orange juice packed in


PET bottles: An experimental and modelling approach
Celine Bacigalupi a, Marie Hélène Lemaistre b, Naima Boutroy b, Christophe Bunel b, Stéphane Peyron a,
Valérie Guillard a, Pascale Chalier a,⇑
a
Joint Research Unit Agropolymers Engineering and Emerging Technologies-UMR 1208 Montpellier Supagro, INRA, UM2, CIRAD, CC 023, Pl. E. Bataillon, 34095 Montpellier
Cedex, France
b
Sidel, Octeville sur mer, BP 204, 76053 Le Havre Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: Sensitivity to oxidation of an orange juice was investigated through packaging in standard PET or active
Received 23 January 2013 PET with oxygen scavenger bottles. The evolution of dissolved oxygen was found to be similar in all bot-
Received in revised form 11 June 2013 tles, whereas ascorbic acid degradation was related to the oxygen transfer with higher losses in standard
Accepted 18 June 2013
PET (53%) against active PET (31%). Moreover, when juice was exposed to high intensity light, a fold faster
Available online 29 June 2013
degradation of ascorbic acid was observed compared to total darkness. Depending also on the light inten-
sity and regardless of the package permeability, changes in the aromatic profile of the juice were
Keywords:
observed due to the degradation of limonene and the formation of a-terpineol, an off-flavour. A mecha-
Packaging
PET
nistic model was developed to predict the shelf life of orange juice. This model, coupling O2 transfer and
Oxygen scavenger ascorbic acid oxidation reaction in the bottled juice, confirmed that oxygen permeation through packag-
Transfer ing material could not be neglected.
Ascorbic acid Ó 2013 Elsevier Ltd. All rights reserved.
Aroma compounds

1. Introduction gested to describe the ascorbic acid oxidation (Polydera, Stoforos, &
Taoukis, 2003; Singh, Heldman, & Kirk, 1976). Recently, a linear
The sensory and nutritional properties of orange juice and con- relationship was found between the first order ascorbic acid degra-
sequently its shelf life depend upon a large range of parameters re- dation constants and the initial headspace oxygen concentrations
lated to the initial product quality, the conditions of processing and (Van Bree et al., 2012). It is obvious after analysis of the above-ci-
the packaging properties (especially the ability of the packaging to ted works that no clear consensus actually exists on the type of
transmit O2 and light, both being involved in orange juice degrada- equation to use to represent and model the vitamin C oxidation
tion). Vitamin C is degraded by oxidative and non-oxidative path- in orange juice.
ways, which results in both nutritional and organoleptic losses Changes in colour and in aroma profile induce depreciation of
(Yuan & Chen, 1998; Zerdin, Rooney, & Vermue, 2003). The two sensory properties and the rejection of the orange juice by the con-
pathways contribute to the formation of reductones, which are in- sumer. The colour change is due to the browning of the juice in
volved in non-enzymatic browning reactions and lead to the for- relation to Maillard reactions (Kacem, Cornell, Marshall, Shireman,
mation of volatile compounds, such as furfural (Kennedy, Rivera, & Matthews, 1987; Solomon et al., 1995), and also to the isomeri-
Lloyd, Warner, & Jumel, 1992; Rouseff, Nagy, Naim, & Zahavi, sation and oxidation of carotenoids, which are influenced by heat,
1992), and brown pigments (Solomon, Svanberg, & Sahlström, light and oxygen (Baker & Gunter, 2004).
1995). Depending on the oxygen level, ascorbic acid degradation Although limonene is the aroma compound that displays the
would be a first-order reaction, with respect to ascorbic acid con- highest concentration in orange juice (50–130 ppm), the aroma
centration (high oxygen level) or to oxygen (low oxygen level) profile of orange juice is largely determined by compounds present
(Garcia-Torres, Ponagandla, Rouseff, Goodrich-Schneider, & at concentrations of less than 1 ppm such as terpenes or terpenols,
Reyes-De-Corcuera, 2009; Zerdin et al., 2003). However, at a lim- aldehydes or esters (Averbeck & Schieberle, 2009; Perez-Cacho &
ited level of dissolved oxygen, for instance, when different oxygen Rouseff, 2008). The aromatic profile of orange juice can be modified
barrier packagings were used, 2nd or zero-order kinetics were sug- by the formation of new compounds, through oxidation or acid-
catalysed reactions (Clark & Chamblee, 1992; Perez-Cacho &
Rouseff, 2008). For instance a-terpineol, which has a stale, musty
⇑ Corresponding author. Tel.: +33 467143891.
or pine-like off-flavour, is formed from the degradation of both
E-mail address: chalier@univ-montp2.fr (P. Chalier).

0308-8146/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.foodchem.2013.06.076
3828 C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836

limonene and linalool (Haleva-Toledo, Naim, Zehavi, & Roussef, 2. Material and method
1999).
Oxygen strongly affects all the aforementioned reactions (Kene- 2.1. Reagents
dy et al. (1992), Perez-Cacho and Rouseff (2008), Petersen, Tonder,
and Poll (1998), Polydera, Stoforos, and Taoukis (2005), Solomon Analytical metaphosphoric acid (33.5–36.5%), 2,6-dichlorophe-
et al. (1995)). Oxygen is present during bottling as dissolved O2 nolindophenol sodium salt hydrate (purity P 97%, based on dry
in the juice and gaseous O2 in the headspace but also diffuses from substance), sodium bicarbonate (purity P 99.7%), ferric chloride
the surrounding atmosphere through the packaging material into hexahydrate (98–102%) and ascorbic acid (purity P 99.0%) were
the juice with a rate depending on the permeability of the material supplied by Fluka.
and the difference in O2 partial pressure on both sides of the pack- Glacial acetic acid (99–100%), absolute ethanol (analytic grade),
aging. Adequate sizing of O2 permeability of the packaging mate- orthophenanthroline (99%) and ammonium acetate (purity P 98%)
rial can increase the shelf life of the juice by avoiding oxidative were obtained from Sigma–Aldrich, as well as aroma compounds
degradation. In this sense, high barrier materials are usually se- limonene, linalool, b-myrcene, a-pinene, furfural, limonene oxide,
lected, such as glass bottles, foil laminates in carton packs (e.g., carvone and 2-heptanol (purity > 99%).
Tetrapak) or flexible pouches and polyethylene terephthalate bot-
tles (PET) (Berlinet, Brat, Brillouet, & Ducret, 2006; Kennedy et al., 2.2. Packaging material
1992; Petersen et al., 1998; Polydera et al., 2003). PET is increas-
ingly used in beverage packaging, due to its excellent mechanical Three PET grades were used in this study: a standard monolayer
properties, UV resistance and good oxygen barrier properties, PET and two active monolayer PETs including an oxygen absorber,
which can be improved by combining different films (multilayer called PETOS1 and PETOS2 and provided by two different suppliers.
PET) or by adding oxygen scavengers to monolayer or multilayer Standard PET can be considered as a passive package compared to
PET (Berlinet, Brat, & Ducret, 2008; Ros-Chumilla, Belissario, Iguaz, the active packages containing oxygen absorbers. Preforms were
& Lopez, 2007; Zerdin et al., 2003). In these active packages, scav- kindly blow-moulded by Sidel (Le Havre, France) to obtain bottles
engers act by reducing the oxygen content dissolved in the juice with 25-cl capacity, weighing 14 g. The thickness of the bottles
and present in the headspace but also by limiting oxygen ingress (measured at body level) was 150 lm (±1.5). All bottles were
and increasing the shelf life. closed using polyethylene caps without any internal joint supplied
Besides oxygen, temperature and light strongly affect the reac- by Bericap (Longvic, France).
tion phenomena by activating diffusion and reaction rates. The ef-
fect of light exposure (fluorescence, UV) on orange juice quality has
2.3. Preparation of orange juice from concentrate and conditioning
not been much investigated and findings on vitamin C are often
contradictory (Conrad, Davidson, Mulholland, Britt, & Yada, 2005;
Concentrated orange juice (65° Brix) was supplied by LSDH (Lai-
Solomon et al., 1995). Its effect depends on the nature and the
terie de Saint Denis de l’Hôtel, France). Sidel prepared the juice and
intensity of light and on the UV barrier properties of packaging
filled the bottles using pilot machines. The concentrated juice was
but also on the nature of the juice. Conrad et al. (2005) showed that
diluted in a tank by adding water under agitation (600 rpm) to ob-
ascorbic acid degradation was dependent on light in apple juice
tain a ° Brix of 11.6. The juice was degassed at 65 °C under vacuum
but not in orange juice. Berlinet et al. (2008) did not observe any
(0.65 mbar) and then flash pasteurised for 20 s at 95 °C in the
effect of artificial light on the vitamin C content of a juice whatever
pasteurisation pilot (Sidel, Le Havre, France). The juice was cooled
the type of packaging material (standard PET or multilayer PET
to 20 °C and maintained in a sterile tank before filling at 5 °C and
with oxygen scavenger), probably because both materials display
under 130 mbar of pressure to avoid microbial contamination.
similar UV transmission properties.
Then, the orange juice was packed in the three different PET bottles
Numerous studies can be found dealing with the nutritional and
using a semi-automatic filler (Sidel, Le Havre, France). Bottles were
sensory quality of orange juice. However, as far as we know, no
previously sterilised by injection of peracetic solution (1800 ppm)
study has yet considered oxidation of vitamin C and changes in ar-
for 10 min, under a pressure of 2 bars at 54 °C and the caps were
oma profile with variation in dissolved oxygen and oxygen barrier
treated by ionisation. Before filling, the bottles were washed with
properties, and the effect of light exposure. Moreover, to rationally
sterilised water previously filtered through a 0.45-lm membrane.
design orange juice packaging, it is necessary to quantitatively
After filling, the headspace was renewed under N2 flux at 1 bar and
evaluate the influence of the oxygen content within the juice and
the bottles were manually closed with the caps. To verify the effi-
the impact of oxygen ingress in the bottle on the oxidation kinetics.
ciency of pasteurisation and filling, microbial analyses were carried
Upto now, dimensioning of O2 barrier properties of packaging
out by enumeration of total flora on PCA (mesophilic and anaero-
material is done empirically with a ‘‘pack-and-pray’’ approach,
bic) and evaluation of mould contamination on PDA.
which is expensive and time-consuming, due to the long shelf life
The samples packed in standard PET bottles were stored at
of pasteurised orange juice (>6 months). An alternative would be to
22 ± 0.5 °C in total darkness or under white light with a radian flux
propose a mathematical model coupling O2 transfer and oxidation
of 400 lux (40 W, 4000 K, 3200 lumens) or of 10,000 lux (150 W,
reactions in the bottle, permitting the prediction of the shelf life of
6000 K, 11,250 lumens) (S.B.F, France). The samples packed in
the juice on the basis of ascorbic acid losses and to identify the O2
PET with oxygen absorbers were stored at only 400 lux.
permeability that best suits the target product.
The objectives of this study were to evaluate the ascorbic acid
content and the change in aroma profile of an orange juice packed 2.4. Methods
in different lightweight PET-based bottles. We used a standard PET
vs. two PETs with oxygen scavengers during 6 months of storage at Analyses were carried out at different times: 0, 7, 15, 22, 30, 60,
three different light exposures. Changes in nutritional and sensory 90, 120, 150 and 180 days and performed in triplicate with 3 differ-
properties of the juice were related to O2 content in the bottle, O2 ent bottles under each condition or packaging.
barrier properties of the packaging and light exposure. These re-
sults were then used to tentatively validate a mathematical model 2.4.1. Oxygen transfer rate (OTR)
aiming at predicting evolution of orange juice quality during The oxygen transfer rate was measured on a MOCON OX-TRAN
storage. 2/20 instrument (MOCON, Minneapolis, MN) according to the
C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836 3829

ASTDM-D3985 standard method. It is a steady state, isostatic 2.4.4. Extraction of aroma compounds of orange juice
method using a coulometric sensor. The test specimen was held Fifty millilitres of orange juice (containing 200 ll of 2-heptanol
so that the two sides of a test chamber were separated: one side solution at 50 mg/100 ml of ethanol used as internal standard)
was exposed to nitrogen and the other to air (i.e. O2 at 21%). The were extracted with 2  50 ml of dichloromethane forduring 30
temperature was maintained at 23 °C and the relative humidity and 15 min under 500 rpm agitation at ambient temperature. After
at 50% whatever the test specimen (passive or active packaging). each homogenisation, the resulting emulsion was disrupted by
OTR was measured at to and after 1, 3 and 6 months of storage be- freezing for 2 h at 18 °C. The organic and aqueous phase were
cause the efficiency of active packaging can vary according to the separated by decantation and gathered before being dried on anhy-
capacity of the absorbers. Each test trial consisted of 5 replicates. drous sodium sulfate and concentrated under nitrogen flux to a fi-
nal volume of 1 ml.

2.4.2. Measurement of local O2 content in juice 2.4.5. Quantification of aroma compounds by GC-FID analysis
Local oxygen partial pressure was determined using a Fibox 3 GC-FID analysis was performed for the quantification of the ar-
fibre optic oxygen meter purchased from PreSens Precision Sensing oma compounds. A Varian 3800 GC equipped with an automatic
GmbH (Neuburg, Germany). sampler (Combipal; CTC, Zwingen, Switzerland), a DB-WaxÒ col-
The O2-sensitive optical sensors (spots of 4 mm diameter) se- umn (30 m  0.25 mm, film thickness of 0.25 lm; J & W Scientific,
lected for this study were the PSt3 sensors (PreSens Precision Sens- Folsom, CA) and a flame ionisation detector (FID; hydrogen,
ing GmbH) that can be used for a range of oxygen pressures 30 mL min1; nitrogen, 30 mL min1; air, 300 ml min1) was used.
ranging from 0% to 100% O2 (with a limit of detection at 0.03% Hydrogen was the carrier gas at a flow rate of 1.5 ml min1. The
and an accuracy of ±0.4% O2 at 20.9% O2 and ±0.05% O2 at 0.2% temperature was set at 250 °C for the injector and 300 °C for the
O2). A conventional two-point calibration in an oxygen-free envi- detector. Programmed thermal conditions were used: oven tem-
ronment (nitrogen or sodium sulfite), and saturated air environ- perature initially at 40 °C was maintained 3 min, then raised by
ment (i.e. 21% O2) was performed before use. A PSt3 sensor was 3 °C min1 to 250 °C, and was kept at 250 °C for 10 min. Selected
glued inside each bottle, using silicone glue, prior to closure. The aroma compounds were identified using known standards and
dot was located in the orange juice about 2 cm below the neck. the quantification was performed using the internal standard for
We measured dissolved O2 just at the interface between the bottle which the response coefficient of each compound was determined.
material and the juice. The interest of this local non-invasive mea-
surement was that it avoided modifying the diffusion rate of O2 2.4.6. Identification of aroma compounds by GC/MS analysis
within the juice by shaking the bottle to homogenise the O2 con- An Agilent 5973 series gas chromatograph equipped with an
tent before the measurement. The optical fibre which was con- autosampler (Combipal; CTC) and a DB-WaxÒ analytical fused-sil-
nected to the oxygen transmitter was placed on the outer surface ica column (30 m  0.25 mm  0.25 lm film thickness; J & W Sci-
of the bottle, right opposite the sensor spot and emitted an excita- entific) and coupled with a 5989A quadrupole mass spectrometer
tion light through the wall. Data acquisition was performed using in electron ionisation mode, was used for identification. The elec-
PST3 software (PreSens – Precision Sensing GmbH, Regensburg, tron impact (EI) energy was 70 eV and the quadrupole temperature
Germany). For each measurement, the temperature was fixed at was set at 250 °C. The injector was heated at 250 °C and used in
20 °C, and the oxygen measurements were compensated accord- split mode (20:1). The oven temperature was held at 40 °C for
ingly. This method proved to be very efficient for monitoring O2 in- 5 min, ramped to 250 °C at a rate of 4 °C/min and maintained at
gress in wine bottles during storage (Dieval, Vidal, & Olav, 2011). 250 °C for 10 min. The carrier gas was Helium 6.0 (Air Products,
Basingstoke, UK), with a flow-rate of 1.5 ml/min. The detection
was carried out in full scan mode covering a mass range (m/z) of
50–450 amu. The aroma compounds were identified from their
2.4.3. Ascorbic acid analysis
mass spectra using the Wiley 275.L library (Agilent Technologies).
Ascorbic acid contents of orange juice were measured by a
Attribution was performed on peaks showing a signal to noise S/
method previously developed by Pénicaud, Peyron, Bohuon, Gon-
N > 3 and identification of the compound was valid when the con-
tard, and Guillard (2010), and based on redox reaction between
fidence rating of mass spectral comparison was superior or equal
ascorbic acid and Fe (III), monitored by spectrophotometry at
to 95. This attribution was further confirmed using relative reten-
505 nm and using microplates. Vitamin C was extracted with aque-
tion indices and by injection of reference compounds. A standard
ous metaphosphoric acid 3% – acetic acid 8% solution, as recom-
solution of alkanes (C11 to C27) was analysed by GC/MS using the
mended in the official method (AOAC, 2007). In short, the
same GC conditions to calculate linear retention indices (ITs ).
extracted sample or fresh standard solutions (15 ll) were spread
out into a wells plate and reacted with 50 ll iron (III). The covered
plate was stored for 10 min at room temperature before the addi- 2.5. Statistical analyses
tion of orthophenanthroline (50 ll) then incubated for 15 min at
room temperature. Finally, ammonium acetate was added: 135 ll Statistical analyses were performed to examine the effect of
in the measurement wells, 185 ll in the blank wells. The plate packaging and storage conditions on vitamin C content and aroma
was incubated for 2 h at room temperature, and absorbences were compounds. Significant differences between samples were esti-
read at 505 nm using a microplate reader (Multiskan Spectrum, mated by analysis of variance (ANOVA). The statistical significance
Thermo Electron Corporation, Saint-Herblain, France). A standard level was set to 0.05. Statistical analyses were performed using R
curve could be obtained from the standard solutions values, and software (free access, http://www.r-project.org/).
from these data all well concentrations could be obtained. Each
analysis was performed in triplicate on three different bottles from 3. Results and discussion
the same storage time and condition. The limit of detection of this
method at 95% confidence level was found to be equal to 11 mg l1 The orange juice was packed in monolayer passive PET bottles
by Pénicaud, Guilbert, Peyron, Gontard, and Guillard (2010) and (standard PET) and in active PET containing oxygen scavengers
the limit of quantification at the same confidence level of (PETOS1 and PETOS2). All bottles were stored at 20 °C (±2 °C)
21 mg l1. and 400 (±5) lux to imitate the conditions of conventional shelf
3830 C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836

storage. To evaluate the effect of light, orange juice packed in stan- observed between the three PETs despite their varying barrier
dard PET bottles was stored at 20 °C in the dark and at 400 and properties. A concentration of oxygen around 0.37 ± 0.09 ppm
10,000 lux. was obtained for all the stored bottles at 400 lux at the end of
the storage, whatever the type of PET. The initial decrease of dis-
3.1. OTR evolution of PETs during storage solved oxygen would be due to its consumption by oxidative com-
ponents, such as vitamin C and aroma compounds, whereas the
The oxygen transfer rate (OTR) of each material (standard PET, further increase could be explained by a higher O2 ingress via the
PETOS1 and PETOS2) was measured just before filling and after dif- packaging rather than its consumption by reactions (Garcıa-Torres,
ferent storage times (0, 1, 3 and 6 months) at 20 °C and 400 lux, in Ponagandla, Rouseff, Goodrich-Schneider, & Reyes-De-Corcuera,
order to verify that the bottles maintained or not their oxygen bar- 2009). Variations in OTR values between passive (standard) and ac-
rier properties until the end of shelf life (6 months). As expected, tive (oxygen scavenger) PETs (i.e. 8–9 times lower for active pack-
both active PET (PETOS1 and PETOS2) showed very low initial OTR aging than for standard PET) did not affect the concentration of
(i.e. at time 0 before filling), since the value was below the detection local oxygen content in the orange juice. We noted that the posi-
limit of the method (0.0005 ± 0.0000), confirming the role of the tion of the dot in the bottle may vary from one bottle to another,
oxygen scavenger in both cases. In contrast an OTR equal to leading to difficulty in comparison between PETs. Because of this,
0.0291 ± 0.0005 cc bottle1 day1 was measured for standard PET we cannot make a conclusion about the average amount of O2 in
and remained unchanged during the 6 months of storage. For both the juice, which could differ between PET types
active PETs, as expected, OTR strongly increased, to 0.0041 ± In contrast, changes in light intensity during storage clearly af-
0.0010 cc bottle1 day1 for PETOS1 and 0.0035 ± 0.008 cc bot- fected the dissolved oxygen concentration, since in total darkness,
tle1 day1 for PETOS2 (up to 8 times) after 3 months of storage. the dissolved oxygen was stable throughout the storage but in-
Accordingly, active PETs lost a part of their initial barrier capacity creased at 400 lux and at 10,000 lux. At 6 months of storage, dis-
after 3 months of storage but remained more effective than standard solved oxygen was significantly different for the three conditions.
PET. It was obvious from these OTR data that oxidation phenomena This could be explained by a modification of OTR with light, due
would be more pronounced in standard PET than in active PET. to the photosensitivity of PET combined with the possible oxida-
tion of the cap composed of polyethylene.
3.2. Evolution of dissolved oxygen content in orange juice during
storage 3.3. Changes in concentration of ascorbic acid during storage

After filling, the initial concentration of dissolved oxygen in The concentration of ascorbic acid was monitored during the
juice was around 0.98 ± 0.17 ppm (or mg l1). It is important to storage of orange juice at 20 °C in different packages and light con-
note that for t > 0 the measurement of O2 content was representa- ditions (Fig. 2). The initial ascorbic acid content just after packag-
tive of the local concentration around the dot stuck at 2 cm below ing was found to be 521 ± 7 mg l1. Whatever the package and
the neck, i.e. very close to the juice/headspace interface, and on the the light conditions, a large decrease in ascorbic acid (25–34% of
internal face of the bottle. Whatever the package and light inten- initial content) was observed from the beginning of the storage
sity, the local oxygen content decreased dramatically during the to day 30 (Fig. 2). This trend coincided with the steep decrease of
first days of storage to reach 0.12 ppm (Fig. 1). When the bottles dissolved oxygen in the juice (Fig. 1) and can be attributed to the
were stored at 400 lux, the values of local dissolved oxygen were consumption of oxygen initially present in the bottle, i.e. dissolved
stable from the 15th day up to the 90th day and then began to in- in the juice and present in the headspace. After the first month, the
crease. After 6 months of storage, no significant differences were loss of ascorbic acid was more gradual as shown by other studies

1.20

1.00

0.80
O2 in ppm

0.60

0.40

0.20

0.00
0 30 60 90 120 150 180
time (day)

Fig. 1. Evolution of dissolved oxygen in orange juice packed in standard PET bottle and stored in total darkness (), at 400 lux ( ), at 10,000 lux (j) and at 20 °C or in PET
bottles with oxygen scavengers PETOS1 ( ) PETOS2 (D) and stored at 400 lux and at 20 °C.
C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836 3831

Fig. 2. Changes in ascorbic content of orange juice packed in standard PET bottles stored in darkness (), at 400 lux ( ), at10,000 lux (j) and at 20 °C or in PET bottles with
oxygen scavengers PETOS1 ( ) and PETOS2 (D) and stored at 400 lux and at 20 °C.

(Berlinet et al., 2006; Kennedy et al., 1992; Zerdin et al., 2003). Up ent increase in ascorbic acid degradation rate. Currently, according
to 90 days, ascorbic acid oxidation seemed independent of the oxy- to the scientific literature data, there is no evidence of an effect of a
gen barrier properties of the packages, whereas, after this duration, high level of light (>1000 lux) on the degradation of ascorbic acid.
significant high losses of ascorbic acid were found for juice packed The shelf life of orange juice is usually estimated from the vita-
in standard PET bottles compared to active PET bottles. Using stan- min C concentration at the end of storage, a value higher than
dard PET, the ascorbic acid degradation was worse and losses 200 mg l1 at the end of shelf life being recommended by AIJN
reached 53% of the initial quantity at 6 months of storage, whereas (2005). Our results showed that with both active PET in all condi-
losses were 41% and 31% at the end of storage for PETOS1 and PE- tions of light exposure and with standard PET (for light exposure
TOS2, respectively. At this time, the OTR of standard PET was 8-fold <400 lux), the vitamin C content was greater than 200 mg l1, i.e.
higher than those of both active PET bottles, leading to higher 314, 370 and 250 mg l1, respectively. In contrast, the juice stored
transfer of oxygen and probably greater losses of ascorbic acid at 10,000 lux in standard PET contained only 195 mg l1 of ascorbic
through oxidation. The OTR difference between the two PETOS acid after 6 months and consequently was no longer saleable.
did not significantly impact the ascorbic losses, except after
6 months of storage. At this time, the more efficient active PET (PE- 3.4. Evolution of aroma compounds during storage
TOS1) led to a reduction in losses of 25%.
The interest of O2 scavengers for preserving vitamin C from oxi- GC–MS analysis of the orange juice aroma compounds allowed
dation in orange juice had already been pointed out in the litera- us to identify 25 compounds including 12 terpenes and sesquiter-
ture. Indeed, Berlinet et al. (2006) found that the losses of penes, 6 terpenols, 1 aliphatic alcohol, 1 ester, 1 acid, 2 ketones and
ascorbic acid in an orange juice packed in a 0.33-l monolayer PET 2 aldehydes (Table 1). Quantification showed that limonene (citrus
bottle and a multilayer PET bottle containing an oxygen scavenger aroma) was the major compound (54.3 ± 1.4 mg l1) followed by b-
during 6 months of storage at 20 °C and under artificial light with myrcene, a-pinene, valencene, linalool, a-terpineol and terpinen-
an intensity of 750 lux reached 60% and 30%, respectively. Simi- 4-ol. The other compounds were present below 0.1 mg/l.
larly, Ros-Chumilla et al. (2007) reported higher losses of vitamin Limonene, linalool (floral) a-pinene and b-myrcene, identified as
C in orange juice packed at 25 °C in darkness with standard PET key odorants of orange juice (Averbeck & Schieberle, 2009), were
(1-l bottle) than with active PET. present in concentrations above their odour thresholds determined
Concerning the light conditions of storage, no significant differ- in water (34, 0.22, 5 and 14 lg l1, respectively). Their odour
ences were observed between 400 lux and darkness for orange activity values, OAV, calculated by dividing the concentration of
juice packed in standard PET, whereas a significantly higher degra- odorant by the odour threshold, reached 1597, 4545, 200 and 80,
dation of ascorbic acid was highlighted from 3 months of storage at respectively, confirming their odorant impact Ahmed, Dennison,
10,000 lux (Fig. 2). After 6 months of storage, 64% of ascorbic acid Dougherty & Shaw (1978).
was lost against 53% for both other conditions (400 lux and dark- Only certain compounds were quantified during storage: those
ness). This result was at variance with previous studies performed susceptible to degradation (limonene, linalool), those formed as a
using lower light intensities (1500–2000 lux), which emphasised result of degradation (limonene oxide, carvone, cis and trans-carve-
no light effect on ascorbic acid losses in PET bottles (Berlinet ol, a-terpineol, terpinen-4-ol and furfural), and those with an odor-
et al., 2008; Conrad et al., 2005) or other material (Solomon ant impact (a-pinene, b-myrcene). Whatever the packaging type,
et al., 1995). Two hypotheses may be expressed to explain higher the limonene level decrease was similar, with a faster degradation
losses of ascorbic acid obtained under the strong light conditions during the first month than the following 5 months (Fig. 3). No sig-
(10,000 lux) used in our study. First, it may be suggested that the nificant differences were observed between passive and active PET.
accelerated degradation of ascorbic acid is related to the strong in- Losses of limonene reached 50% of the initial quantity (i.e.
gress of oxygen observed for this storage condition (Fig. 1). Second, 28 mg l1) after 6 months of storage at 20 °C and 400 lux or in
a light-induced degradation of ascorbic acid may occur in addition darkness using the different PETs. Berlinet, Ducret, Brillouet, and
to the auto-oxidation phenomenon that would provoke an appar- Brat (2005) only reported 30% of losses, i.e. 40 mg l1 of losses,
3832 C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836

Table 1 sorial quality of juice. However, as for limonene, the degradation of


Volatile flavour components of orange juice just after filling (t = 0) these monoterpenes could not be related to the content and trans-
Component IR lita IR GC–MSb Concentration mg l1 fer of oxygen, since no significant differences were observed be-
a-Pinene 1038–1039 1029 1.06 ± 0.07 tween different packagings. It can be also suggested that the
Ethyl butanoate 1048–1040 1050 0.09 ± 0.03 effect of oxygen would be masked by acid-catalysed reaction (Per-
d-Carene 1148 1132 tr ez-Cacho & Rouseff, 2008). Accordingly, it was shown that oxygen
b-Myrcene 1148 1151 1.13 ± 0.14 removal did not improve the sensory properties of orange juice
Limonene 1192–1183 1185 54.26 ± 1.38
c-Terpinene 1224–1251 1228 0.07 ± 0.01
stored for 5 months (Trammell, Dalsis, & Malone, 1986). On the
p-Cymène 1261 1254 tr other hand, these compounds showed a great reactivity under
a-Terpinolene 1264–1284 1240 tr strong light, which could be associated with the formation of oxi-
Octanal 1267–1276 1268 tr dation products caused by allylic oxidation (Nguyen et al., 2009).
Hexanol (IS) 1351–1392 1361
The higher reactivity of b-myrcene relative to limonene or a-
Furfural 1432–1455 1450 tr
a-Copaene 1488 1464 0.10 ± 0.03 pinene could be associated with greater allylic hydrogen. It would
Linalool 1525–1537 1534 0.96 ± 0.06 have been of interest to study the effect of strong light in the pres-
Octanol 1537–1566 1542 0.08 ± 0.01 ence of oxygen absorbers.
b-Cubebene 1546 1560 0.10 ± 0.04 Linalool was degraded during storage and no difference was ob-
b-Caryophyllene 1562–1594 1565 0.15 ± 0.04
served between the different PETs. However, in contrast to terp-
1-terpine-4-ol 1591 1580 0.19 ± 0.02
b-Terpineol 1616 1605 tr enes, no significant effect of light was observed (Fig. 3). After
Sesquiterpene nd 1659 0.09 ± 0.01 6 months, losses reached 40% of the initial quantity of linalool. A
a-Terpineol 1688 1676 0.16 ± 0.04 more drastic decrease (49%) was observed by Berlinet et al.
Valencene 1751 1687 0.91 ± 0.13
(2005) after only 5 months using a multilayer PET with a 2-fold
Carvone 1720 1704 0.09 ± 0.01
Trans carveol 1801 1813 0.08 ± 0.02 higher initial concentration of linalool. In agreement with our re-
Cis carveol 1820 1801 tr sults, the final concentration of aroma compounds varied between
Hexanoic acid 1829 1829 0.10 ± 0.01 6% and 38% of the initial level after 12 weeks of storage for orange
Nootkatone 2250–2366 2473 tr juices obtained by high pressure or thermal treatment and condi-
tr trace (<0.05 mg/l), IS (internal standard). tioned in PET bottles and stored at 4 or 10 °C (Baxter, Easton, Sch-
a T
Is litterature : linear retention index were extracted from http://www.phero- neebeli, & Whitfield, 2005).
base.com/database/kovats or from Berlinet et al. (2005). a-Terpineol is a derivate of linalool and limonene. It was ini-
b T
Is calculated linear retention index from GC data.
tially present at a very low amount (0.16 ± 0.04 mg l1) but as ex-
pected, its concentration increased during storage for all the
different packages and storage conditions. After 6 months, an 8-
for orange juice packed in a monolayer PET and stored 5 months at fold increase was observed. In their study, Berlinet et al. (2005) ob-
20 °C in total darkness for a higher initial amount of limonene served only a 4-fold increase regardless of the packaging material
(130 mg/l). However, it was difficult to make a comparison with employed. It can be noted that the initial concentration was higher
the Berlinet et al. study, due to the different concentrations of lim- than in our study, probably due to its formation during processing.
onene, as well as there being no data regarding oxygen concentra- The detection threshold of this aroma compound imparting a stale,
tion. For orange juice stored at 10,000 lux, significant difference in musty or pine-like, odour is 2.5 ppm. After 6 months of storage, the
limonene content was observed from 2 months onwards. At the concentration largely remained below this value. Linalool is a more
end of storage, 65% of limonene was lost, i.e. more than 10 mg l1, reactive substrate than limonene for producing a-terpineol. How-
compared to the other conditions. ever, since there is more limonene than linalool in citrus juices,
It was observed that limonene could be rapidly degraded under a-terpineol appeared to be formed from both precursors (Haleva-
UV irradiation (48 h at 37 °C) resulting in the formation of a mix- Toledo et al., 1999). Regarding terpinen-4-ol, generated from a-
ture of products and a caraway-like and minty odour (Nguyen, pinene and limonene, no increase was observed during storage.
Campi, Jackson, & Patti, 2009), due to carvone enantiomers (Aver- Furfural, which can be formed from ascorbic acid decomposition,
beck & Schieberle, 2009). The carvone concentration was about was initially present (Table 1) but its quantification was possible
0.02 ± 0.01 mg l1 at 22 days, regardless of the storage conditions only after 7 days of storage. Its concentration slightly increased
and PET used, except at 10,000 lux (Fig. 3). In this case, the forma- up to 0.2 mg l1 whatever the packaging and the storage condi-
tion of carvone was 10 times higher (0.18 mg l1) at 22 days of tions. Its formation was not perturbed by the rate of ascorbic acid
storage and then decreased to reach 0.04 mg l1 at 3 months. We degradation depending on oxygen barrier properties of the packag-
can conclude that carvone formation is increased by intensive ing. However, it is an intermediate product, which can be also de-
light, which is in agreement with the high rate of carvone observed graded as quickly as it is formed.
when limonene was submitted to UV light (Nguyen et al., 2009). In short, for the 10 aroma compounds studied here, no signifi-
Among the other limonene degradation products, the presence cant differences were noticed in their degradation during storage
of limonene oxides and carveol was not evidenced during the en- of the juice in different barrier PET bottles. No impact of packaging
tire storage time. The limonene oxides and carvone were formed type on oxygen level was observed in agreement with the results of
in parallel but oxides have a relatively faster rate of formation than Berlinet et al. (2005), who highlighted that changes in aromatic
carvone (Haleva-Toledo et al., 1999). Moreover, limonene oxides profile of orange juice are independent of the oxygen permeability
can be converted into other products faster than they are formed. of the PET material. The main degradation pathways of aroma com-
Regarding the two other monoterpenes, a-pinene (pine-tree- pounds are probably due to acid catalysed reactions as well as to
like smell) and b-myrcene (geranium-like odour), a strong decrease aroma sorption in PET (Berlinet et al., 2005). However, a high level
in concentration was observed : the losses reached Fig. 3 100% and of light induced more drastic degradation and the formation of a
51%, respectively for all the juices stored in dark or at 400 lux. Un- higher amount of carvone, one of the oxidation products of limo-
der 10,000 lux, the degradation of the two compounds was acceler- nene. With regard to the losses of key aroma compounds, which
ated and 84% loss was noticed for b-myrcene after 6 months of can reach up to 40–100%, OAV were strongly decreased but they
storage. They are minor components but they contribute signifi- remained higher than 1, except for a-pinene (pine odour), which
cantly to the flavour and their losses can negatively affect the sen- was not detected after 6 months of storage. From this study, it is
C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836 3833

Fig. 3. Evolution of limonene, b-myrcene, a-pinene, linalool and a-terpineol in orange juice packed in standard PET bottles and stored in total darkness (), at 400 lux ( ), at
10,000 lux(j) and at 20 °C or in PET bottles with oxygen scavengers PETOS1 ( ),PETOS2 (D) stored at 400 lux and at 20 °C.

supposed that limited changes occur to the sensorial properties of tial order of reactions with respect to ascorbic acid and dissolved
the juice. Moreover the fast and continuous oxidation of ascorbic oxygen. In Eq. (1), the impact of both O2 and ascorbic acid content
acid led to a reduced content of oxygen and reactive oxygen spe- (affected by partial order of reaction) on the reaction rate are taken
cies and thus protected the aromatic profile against oxidation. into account, which is rare compared to all the previous studies car-
ried out on the oxidation kinetic of ascorbic acid. Indeed, ascorbic
3.5. Modelling the shelf life of bottled orange juice acid oxidation has been often described using first order kinetics,
with respect to AA concentration or zero order reaction (Conrad
The experimental results previously presented confirm that et al., 2005; Garcia-Torres et al., 2009; Polydera et al., 2003; Zedin
ascorbic acid losses can be rightly considered as a marker of the et al., 2003). By using Eq. (1), a direct coupling of a mathematical
sensitivity of a product to the level of oxygen and thus the evolu- model for O2 diffusion with AA oxidation equation could be made:
tion of quality of orange juice during storage. The possibility to pre- the concentrations of O2 at time t in the product being one of the
dict in advance ascorbic acid losses during storage of a packed outputs of the diffusion model (Fick’s model) and one of the inputs
orange juice would be very helpful to allow a correct design and required in Eq. (1).
sizing of the packaging material. In order to develop such a tool, In the present work, the model of Pénicaud et al. (2011) was up-
a mechanistic model previously developed by Pénicaud, Broyart, dated in order to take into account O2 ingress into the bottle
Peyron, Gontard, and Guillard (2011), coupling O2 transfer (Fick’s through the packaging material, which was represented as follows:
model) and ascorbic acid oxidation reaction, was upgraded and
dO2 P
adapted to the present study. These authors used Eq. (1) as the oxi- ¼ AðpO2 outside  S  O2d Þ ð2Þ
dation model. dt e

r AA ¼ k½AAa ½O2 b ð1Þ where P, e and A are, respectively, the permeability, thickness and
surface of the packaging material, pO2 outside is the external partial
where AA and O2 represent ascorbic acid and dissolved oxygen con- pressure of O2, S is the solubility of O2 in the juice and O2d is the
tent (mol l1) in food, k is the kinetics constant rate, a and b are par- concentration of dissolved oxygen at time t. Then at the interface
3834 C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836

Fig. 4. Description of the food/packaging system considered in the modelling approach.

Table 2
Input parameters required for predicting average concentration in ascorbic acid (AA) and dissolved O2 as a function of time during storage of orange juice at 20 °C and 400 lux.

Geometry of the system Diffusion model Reaction model


Bottle thickness 370  106a O2 solubility (mol l1 Pa1) 1.35  108b k (mol1.2 l1.2 s1) 3.0  102b
(m)
Bottle surface (m2) 26050  106 O2 diffusivity in juice 0.31  109b a (–) 1b
(m2 s1)
Headspace volume 17 AA diffusivity in juice 2.26  109b b (–) 1.2b
(mL) (m2 s1)b
Bottle width (mm) 160 O2 diffusivity in headspace 16  106c Initial O2 concentration in juice 0.98
(m2 s1) (mg l1) (±0.17)
Juice volume (mL) 250 Permeability 9.015  1018 for passive PET 1.308  1018 for Initial AA concentration in juice 521 (±7.1)
(mol m1 s1 Pa1) active PETOS1 (mg l1)
External O2 partial pressure 21,273 Initial O2 concentration in 21
(Pa) headspace (%)

Other parameters are from this study.


a
Supposed to be homogeneous. This average thickness taking into account the neck was calculated from the thickness profile of the bottle obtained from Sidel.
b
From Pénicaud et al. (2011).
c
From Sidel.

headspace/juice, O2 was assumed to be dissolved according to mental determination of the permeability of the cap being unfeasi-
Henry’s law (see Table 2 for the model parameters and Fig. 4 for ble, an identification of this parameter was attempted on the basis
the details of the system studied). To take into account more easily of O2 experimental results. With an average thickness value of
the 2-dimensional diffusion in the bottle (assumed to be a cylinder 1 mm, a permeability of 3  1017 mol m1 s1 Pa1 was found opti-
of 56 mm diameter and 160 mm height), the mathematical model mal to fit the O2 data (Fig. 6B – dotted line, RMSE lower than 10%).
was implemented in COMSOL3.5Ò (COMSOL Multiphysics) and then This value is in agreement with literature data about permeability
exported into MatlabÒ software (The Mathworks Inc., Natick, MA) of HPDE (2.49 and 4.98  1017 mol m1 s1 Pa1 (Monsivais-Bar-
to draw the figures. ron, Bonilla-Rios, Ramos de Valle, & Palacios, 2013). However, in that
In a first instance, the cap of the bottle was assumed to be case, the fitting of the ascorbic acid losses was less good; the RMSE
impermeable and only the OTR of the PET bottle was considered. increased from 34.2 to 44.6 mg l1. Discrepancies between experi-
As shown in Fig. 5A for standard PET bottles, the model used with mental and predicted ascorbic losses could be explained by the ex-
this assumption predicted quite well the experimental results of tremely simplified oxidation mechanism represented by Eq. (1).
ascorbic acid losses (RMSE1 of 34.2 mg l1, less than 10% of error Obviously impact of O2 on ascorbic acid losses was not completely
on the predicted ascorbic losses; Fig. 5A) whereas fitting of the local well represented by Eq. (1); i.e. less O2 was required in practise than
O2 content in the juice is obviously not as good (more than 30% of in theory to oxidise ascorbic acid in the standard PET bottle. The stoi-
error at 180 days; Fig. 5B). The high discrepancy between theoretical chiometry of the reaction used by Pénicaud et al. (2011) in their
and experimental results for O2 suggests an additional flux of O2 work carried out in aqueous solutions may not completely fulfil
would occur in the bottle that could come from the closure. Experi- the conditions encountered in orange juice. In addition, the oxidative
pathway may not be the unique and predominant mode of degrada-
tion for ascorbic acid in orange juice (i.e. non-oxidative pathway,
1
Root Mean Square Error between the experimental data and the predicted one. Maillard reactions, hydrolysis) and, in that case, it is understandable
C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836 3835

Position of the
O2sensor

500

Local oxygen content (mg/L)


0.8
400
Ascorbic acid (mg/L)

0.6
300

0.4
200

100 0.2

0 0
0 50 100 150 0 5 10 15
Time (s) 6
Time (d) x 10

A B

500

0.8
Local O2 content (mg/L)

400
Ascorbic acid (mg/L)

0.6
300

0.4
200

0.2
100

0 0
0 50 100 150 0 5 10 15
6
Time (d) Time (s) x 10
C D
Fig. 5. Experimental (symbols) and predicted (solid and dotted lines) average ascorbic acid concentration and local content of dissolved O2 in orange juice stored at 20 °C and
400 lux in standard PET (A and B) and in active PETOS1 (C and D) bottles of 250 ml. Solid lines were obtained assuming that the cap was impermeable and dotted lines for a
permeable cap with an adjusted oxygen permeability. Experimental measurement and prediction of O2 content was for a position of 2 cm below the neck and against the
interface of the bottle.

that Eq. (1) was not able to correctly represent degradation of ascor- of local concentration (close to the interface of juice/headspace),
bic acid. local O2 content was not so different between the two kinds of
In order to confirm the aforementioned hypothesis, the predic- PET in spite of differences in O2 average content and consequent
tions of the model were applied to a second set of data, those ob- ascorbic acid losses. Without mathematical modelling, this unex-
tained for the active PET (PETOS1). As shown in Fig. 5C and D, for pected behaviour of the O2 curve would not have been explained.
active PET, using the permeability of the cap previously identified, It is also important to note that uncertainty on the exact position
predictions of local oxygen content were quite good (RMSE of 12%) of the dot used to monitor local O2 made the analysis of O2 data dif-
whereas prediction of ascorbic acid losses was worse than for stan- ficult and revealed a limit of such non-invasive and non-destruc-
dard PET (RMSE of 60.0 mg l1). Indeed, experimental points varied tive methods.
on both sides of the theoretical curve: in practise, ascorbic acid
would degrade more rapidly than in theory during the first 15 days
of storage, then this degradation slowed down and became overes- 4. Conclusion
timated by the model. This confirms that the role of O2 in ascorbic
acid degradation in orange juice would be not as constant as pre- This study has confirmed that the ascorbic acid losses are the
dicted by Eq. (1) otherwise this experimental curve would have most relevant marker of orange juice ageing and quality evolution
been linear most of the time like the theoretical one. Further work during storage. O2 was obviously the limiting parameter in the
should be addressed in the future to clarify the modelling of ascor- reaction of ascorbic acid degradation and was consumed gradually
bic acid degradation in a real product when different mechanisms as it entered the package depending on its O2 permeability proper-
coexist. This mathematical model, even if imperfect in its predic- ties. Contrary to ascorbic acid, the aroma changes as a function of
tions, pointed out the role of the closure in the O2 ingress into time were not relevant markers of oxygen ingress and permeability
the packaging. Due to this phenomenon, and also the monitoring properties of the packaging. Indeed, the modification of the aroma
3836 C. Bacigalupi et al. / Food Chemistry 141 (2013) 3827–3836

profile seems mainly due to acid-catalysed reactions and to a lesser methods of removal. Comprehensive Reviews in Food Science and Food Safety, 8,
409–421.
extent to oxidation. After 6 months of storage, the losses of aroma
Haleva-Toledo, E., Naim, M., Zehavi, U., & Roussef, R. L. (1999). Formation of a-
compounds can be drastic since they reached 30–100% of the initial terpineol in citrus juices, model and buffer solutions. Food Chemistry and
level, depending on the light intensity and nature of the com- Toxicology, 64, 838–841.
pounds. Exposure to extreme light condition (=10,000 lux) was Kacem, B., Cornell, J. A., Marshall, M. R., Shireman, R. B., & Matthews, R. F. (1987).
Nonenzymatic browning in aseptically packaged orange drinks: Effect of
found to speed up ascorbic acid and aroma compound degradation, ascorbic acid, amino acids and oxygen. Journal of Food Science, 52, 1668–1672.
inducing the formation of oxidative derivatives. A first attempt of Kennedy, J. F., Rivera, Z. S., Lloyd, L. L., Warner, F. P., & Jumel, K. (1992). L-ascorbic
shelf life prediction was made using a mechanistic model coupling acid stability in aseptically processed orange juice in TetraBrik cartons and the
effect of oxygen. Food Chemistry, 45, 327–331.
O2 diffusion and the reaction of ascorbic acid degradation. Simula- Monsivais-Barron, A. J., Bonilla-Rios, J., Ramos de Valle, L. F., & Palacios, E. (2013).
tions confirmed that permeation through packaging material and Oxygen permeation properties of HDPE-layered silicate nanocomposites.
especially through the closure need to be considered and that Polymer Bulletin, 70, 939–951.
Nguyen, H., Campi, E. V., Jackson, R. W., & Patti, A. F. (2009). Effect of oxidative
mechanisms of ascorbic acid degradation must be better character- deterioration on flavour and aroma components of lemon oil. Food Chemistry,
ised, in order to propose more accurate predictions. At the mo- 112, 388–393.
ment, this model is able to predict the shelf life of orange juice Pénicaud, C., Broyart, B., Peyron, S., Gontard, N., & Guillard, V. (2011). Mechanistic
model to couple oxygen transfer with ascorbic acid oxidation kinetics in model
with less than 20% error and, used in a reverse manner, is able to solid food. Journal of Food Engineering, 104, 96–104.
identify the window of O2 permeability required for maintaining Pénicaud, C., Guilbert, S., Peyron, S., Gontard, N., & Guillard, V. (2010). Oxygen
a minimal level of ascorbic acid (chosen by the user). transfer in foods using oxygen luminescence sensors: Influence of oxygen
partial pressure and food nature and composition. Food Chemistry, 123,
1275–1281.
References Pénicaud, C., Peyron, S., Bohuon, P., Gontard, N., & Guillard, V. (2010). Ascorbic acid
in food: Development of a rapid analysis technique and application to
Ahmed, E. M., Dennison, R. A., Dougherty, R. H., & Shaw, P. E. (1978). Flavor and odor diffusivity determination. Food Research International, 43, 838–847.
thresholds in water of selected orange juice components. Journal of Agricultural Perez-Cacho, P. R., & Rouseff, R. L. (2008). Fresh squeezed orange juice odor: A
Food Chemistry, 26, 187–191. review. Critical Reviews in Food Science and Nutrition, 48, 681–695.
AIJN (2005). Code of practice, 6.1 reference guide orange, Absolute quality Petersen, X., Tonder, D., & Poll, L. (1998). Comparison of normal and accelerated
requirements. Association of Juices and nectars from fruits and vegetables of storage of commercial orange juice-changes in flavour and content of volatile
European union revised June 2003. compounds. Food Quality and Preference, 9, 43–51.
AOAC (2007). Vitamin C (ascorbic acid) in vitamin preparation and juices: 2,6 Polydera, A. C., Stoforos, N. G., & Taoukis, P. S. (2003). Comparative shelf life study
dichloroindophenol titrimetric method final reaction, Ch. 45.1.14, in Officials and vitamin C loss kinetics in pasteurised and high pressure processed
Methods of analysis. In W. Horwitz, & G. W. Latimer (Eds.), Association of Official reconstituted orange juice. Journal of Food Engineering, 60, 21–29.
analytical Chemist MD, US. Polydera, A. C., Stoforos, N. G., & Taoukis, P. S. (2005). Quality degradation kinetics of
Averbeck, M., & Schieberle, P. H. (2009). Characterisation of the key aroma pasteurised and high pressure processed fresh Navel orange juice: Nutritional
compounds in a freshly reconstituted orange juice from concentrate. European parameters and shelf life? Innovative Food Science and Emerging Technologies, 6,
Food Research Technology, 229, 611–622. 1–9.
Baker, R., & Gunter, C. (2004). The role of carotenoids in consumer choice and the Ros-Chumilla, M., Belissario, Y., Iguaz, A., & Lopez, A. (2007). Quality and shelf life of
likely benefits from their inclusion into products for human consumption. orange juice aseptically packaged in PET bottles. Journal of Food Engineering, 79,
Trends Food Science Technology, 15, 484–488. 234–242.
Baxter, I. A., Easton, K., Schneebeli, K., & Whitfield, F. B. (2005). High pressure Rouseff, R. L., Nagy, S., Naim, M., & Zahavi, U. (1992). Off-flavor development in
processing of Australian navel orange juices: Sensory analysis and volatile citrus juice products. In G. Charalambous (Ed.), Off-flavors in foods and beverages
flavor profiling. Innovative Food Science and Emerging Technologies, 6, 372–387. (pp. 211–217). Amsterdam, the Netherlands: Elsevier.
Berlinet, C., Brat, P., Brillouet, J.-M., & Ducret, V. (2006). Ascorbic acid, aroma Singh, R. P., Heldman, D. R., & Kirk, J. R. (1976). Kinetics of quality degradation:
compounds and browning of orange juice related to PET packaging materials Ascorbic acid oxidation in infant formula during storage. Journal of Food Science,
and pH. Journal of Science of Food and Agriculture, 86, 2206–2212. 41, 304–308.
Berlinet, C., Brat, P., & Ducret, V. (2008). Quality of orange juice in barrier packaging Solomon, O., Svanberg, U., & Sahlström, A. (1995). Effect of oxygen and fluorescent
material. Packaging Technology and Science, 21, 279–286. light on the quality of orange juice during storage at 8 °C. Food Chemistry, 53,
Berlinet, C., Ducret, V., Brillouet, J. M., & Brat, P. (2005). Evolution of aroma 363–368.
compounds from orange juice stored in polyetlylene terephthalate (PET). Food Trammell, D. J., Dalsis, D. E., & Malone, C. T. (1986). Effect of oxygen on taste,
Additives & Contaminants, 22, 185–195. ascorbic acid loss and browning for HTST-pasteurized, single-strength orange
Clark, B. C., Jr., & Chamblee, T. S. (1992). Acid-catalyzed reactions of citrus oils and juice. Journal of Food Science, 51, 1021–10233.
other terpene-containing flavors. In G. Charalambous (Ed.), Off-flavors in food Van Bree, I., Baetens, J. M., Samapundo, S., Devlieghere, F., Laleman, R.,
and beverages (pp. 229–284). Amsterdam: Elsevier. Vandekinderen, I., et al. (2012). Modelling the degradation kinetics of vitamin
Conrad, K. R., Davidson, V. J., Mulholland, D. L., Britt, I. J., & Yada, S. (2005). Influence C in fruit juice in relation to the initial headspace oxygen concentration. Food
of PET and PET/PEN blend packaging on ascorbic acid and color in juices Chemistry, 134, 207–214.
exposed to fluorescent and UV light. Journal of Food Science, 70, 19–25. Yuan, J. P., & Chen, F. (1998). Degradation of ascorbic acid in aqueous solution.
Dieval, J., Vidal, S., & Olav, A. (2011). Measurement of the oxygen transmission rate Journal of Agricultural and Food Chemistry, 82, 387–395.
of co-extruded wine bottle closures using a luminescence-based technique. Zerdin, K., Rooney, M. L., & Vermue, J. (2003). The vitamin C content of orange juice
Packaging Technology and Science, 24, 375–385. packed in an oxygen scavenger material. Food Chemistry, 82, 387–395.
Garcıa-Torres, R., Ponagandla, N. R., Rouseff, R. L., Goodrich-Schneider, R. M., &
Reyes-De-Corcuera, J. I. (2009). Effects of dissolved oxygen in fruit juices and

You might also like