Sei sulla pagina 1di 10

Effects of maltodextrin and pulp on the water sorption, glass transition, and caking

properties of freeze-dried mango powder


Suwalee Fongina,b, Alex Eduardo Alvino Granadosb, Nathdanai Harnkarnsujaritc, Yoshio
Hagurab, Kiyoshi Kawaib,∗
a
Division of Food Science and Technology, School of Agriculture and Natural Resource, University of Phayao, 19 Phaholyothin Rd, Muang Phayao, Phayao, 56000,
Thailand b Graduate School of Biosphere Science, Hiroshima University, 1-4-4 Kagamiyama, Higashi-Hiroshima, Hiroshima, 739-8528, Japan
c
Department of Packaging and Materials Technology, Faculty of Agro-Industry, Kasetsart University, 50 Ngam Wong Wan Rd, Lat Yao, Chatuchak, Bangkok, 10900,
Thailand

ARTICLEINFO ABSTRACT
Keywords: Effect of maltodextrin on the water sorption behavior and glass transition temperature (Tg) of freeze-dried mango solute (water-
Dry mango soluble mango material) was investigated. In addition, the caking property of freeze-dried mango puree and mango solute with
Maltodextrin and without maltodextrin was investigated. The Tg of mango solute (32 °C) increased discontinuously with maltodextrin weight
Water sorption fraction with a step-point between 0.6 and 0.7. The pulp of mango decreased the hygroscopicity and increased the T g of mango
Glass transition powder to 49 °C. The degree of caking was reduced by the additions of maltodextrin and pulp because of the increased Tg. The
Caking powder samples except for maltodextrin showed a small degree of caking (below 13%) even below their Tg. An empirical model
to predict the effect of water content on Tg was applied to the mango solute-maltodextrin mixtures. The predicted Tg values were
in good agreement with the measured values (coefficient of determination = 0.919).

1. Introduction content. When the Tg becomes lower than room temperature as a result of water
sorption, the glass turns into the rubbery state. The water content and the a w for
Mango (Mangifera indica L.) is a seasonal fruit grown in tropical and which the amorphous food materials exhibit glass transition at 25 °C have been
subtropical regions, particularly Asia (Berardini et al., 2005). Mango possesses commonly described as critical water content (Wc) and critical aw (awc) (Fabra
an outstanding sweet and succulent taste along with an exotic flavor. Mango is et al., 2011; Roos, 1995b). Water content or water activity has a big influence
also considered to be a functional food because of its nutritional value (high on the Tg of dried food materials, thus, gives important criteria for their
dietary fiber, vitamins, minerals, and phenolic compounds) (Kim et al., 2009; processing (Goula and Adamopoulos, 2008a; Omar and Roos, 2007), storage
Pott et al., 2003; Rocha Ribeiro et al., 2007; Valente et al., 2011). Because of stability (Telis and Martínez-Navarrete, 2009; Wang et al., 2008), and texture
its high moisture content, mango is classified as a highly perishable fruit. control (Djendoubi Mrad et al., 2012; Kurozawa et al., 2012).
Therefore, preservation methods such as drying are employed to create less It is generally difficult to obtain physically stable powder from dry fruits
perishable forms of mango. because they are prone to caking during processing and storage (Aguilera et al.,
Dried fruits are commonly in an amorphous state, and amorphous solids 1995). Caking is a deteriorative phenomenon that involves powder
suddenly undergo evident rheological changes at the glass transition agglomeration, consolidation, and adhesion (Palzer and Sommer, 2011).

∗ Corresponding author. Postal address: Department of Biofunctional Science and Technology, Graduate School of Biosphere Science, Hiroshima University, 1-4-4 Kagamiyama,
Higashi-Hiroshima, Hiroshima, 739-8528, Japan. E-mail address: kawai@hiroshima-u.ac.jp (K. Kawai).

https://doi.org/10.1016/j.jfoodeng.2018.11.027
Received 26 June 2018; Received in revised form 26 November 2018; Accepted 30 November 2018 Available online 03
December 2018
0260-8774/ © 2018 Elsevier Ltd. All rights reserved.
temperature (Tg), which is the temperature at which glass (rigid) transitions to Caking is a serious problem in the powder food industry as it negatively affects
the rubber (soft) state (Roos, 1995b; Roos and Silalai, 2011; Telis and food processing and handling (Brennan et al., 1971; Paterson and Bröckel,
Martínez-Navarrete, 2012). Glass transition also occurs as a result of an 2015) along with the shelf life (Peleg and Mannheim, 1969). Poorly rehydrated
increase in water content or water activity (aw), even at a constant temperature, powder, reduced sensory properties, and shortened shelf life are negative
because the Tg of amorphous solids decreases with an increase in the water qualities that have been attributed to caking in food powders (Aguilera et al.,
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

1995). Since caking of amorphous powders occurs in the rubbery state, the 2.2. Preparations of freeze-dried mango puree (MP) and mango solute
caking property of dry fruit powders is expected to be improved by the mixing (MS)
of high-Tg amorphous materials.
The Tg of mango powder can be increased by mixing with maltodextrin To prepare freeze-dried MP powder, MP was used without further
(MD) (Fongin et al., 2017). MD is commonly used for this purpose because of treatment. For freeze-dried MS, the pulp of the MP was removed as follows.
its high Tg, bland flavor, and low price. The addition of MD can reduce caking MP (25 g) and distilled water (50 g) were mixed until homogeneity was
in food powders, thereby improving the flowability of the powder (Palzer and achieved. The suspension of MP was centrifuged at approximately 1000×g for
Sommer, 2011). For example, oven-dried grugru palm (Acrocomia aculeata) 30 min (Model He103N series, Kokusan Inc., Tokyo, Japan), and the
powder containing 8% MD exhibited significantly less caking and greater supernatant was collected. The supernatant was mixed with distilled water (50
stability compared to the powder without MD (Oliveira et al., 2014). Bhusari g), and the solution was centrifuged again. The obtained supernatant was
et al. (2014) evaluated the effects of different concentrations of MD (40%, 50%, treated as “mango solution”.
and 60%) on the powder properties of spray-dried tamarind pulp and found that The MP and mango solution were distributed in an aluminum chamber and
the higher MD contents resulted in a more free-flowing powder. Working with radially frozen in a freezer at −20 °C. The frozen samples were then transferred
spray-dried date powders with MD contents of 30%, 40%, 50%, and 60% (dry to a pre-cooled chamber and freeze-dried at a pressure lower than 70 Pa as the
basis of date syrup), Farahnaky et al. (2016) showed that increasing the MD temperature was increased from −35 °C to 5 °C in a stepwise fashion over a 48
content increased Tg, reduced caking, and enhanced stability during processing, h period. The freeze-dried solids were powdered using a mixer.
handling, and storage. Jaya and Das (2004) studied the properties of vacuum-
dried mango powders with various MD contents (0.25, 0.35, 0.45, 0.55, and 2.3. Preparations of freeze-dried mango puree-maltodextrin (MP-MD) and
0.65 kg/kg dry mango solid). They found that the most desirable powder mango solute-maltodextrin (MS-MD)
properties were obtained when MD content was in the range of 0.43–0.55 kg/kg
The freeze-dried MP powder and MS powder were mixed with MD at
dry mango solid. The addition of MD has also been shown to prevent caking in
various dry-weight fractions of MD. Samples are identified by the MD weight
commercial onion (Peleg and Mannheim, 1969), spray-dried tomato (Goula and
fraction; for example, the mixture with MS: MD = 0.8: 0.2 is referred to as “0.2
Adamopoulos, 2008b), oven-dried bael (Sagar and Kumar, 2014), spray-dried
soursop (Costa et al., 2014), vacuum-dried honey (Sahu, 2008), freeze-dried MD”. The powders were diluted into 20% (w/w) aqueous systems by the
dragon fruit pulp (Molina et al., 2014), spray-dried barberry fruit (Akhavan addition of distilled water and then freeze-dried according to the above-
Mahdavi et al., 2016), spray-dried orange juice (Goula and Adamopoulos, mentioned procedures. The freeze-dried solids were powdered using a mixer.
2010; Islam et al., 2017), and spray-dried soy sauce (Wang and Zhou, 2012).
2.4. Water sorption isotherm
Despite the large number of works highlighted above, no systematic study
has focused on the effects of MD on dried fruit powders. Systematic
The freeze-dried MS-MD samples (0.5 g) were placed in an aluminum dish
experimental data allow the prediction of Tg and control over the physical
(diameter =40 mm) and vacuum-dried at 80 °C for 6 h. The fully dried samples
stability of dried fruit powders. In our previous study, it was reported that the
were exposed to various aw conditions at 25 °C for at least 7 d. The value of a w
Tg of freeze-dried mango puree (MP) increased with increasing MD content and
was controlled using the following saturated salt solutions: LiCl (a w = 0.113),
decreasing water content (Fongin et al., 2017). The results for mango puree
CH3COOK (aw =0.225),
were compared with those for other dried fruits, and an empirical model to
MgCl2 (aw = 0.328), K2CO3 (aw =0.432), Mg(NO3)2 (aw =0.529), NaBr (aw
predict the plasticizing effect of water on the Tg of dried fruits was proposed.
=0.576), and NaCl (aw = 0.753). The equilibrium water content (We) of each
This model will be useful for the prediction of physical quality change induced
sample was evaluated gravimetrically by drying at 105 °C for 16 h. The
by water sorption. Although the model was suggested to be affected by the pulp measurements were performed in triplicate, and the results were averaged.
content of mango puree, there was no quantitative evidence. In addition, The water sorption isotherm was analyzed by the Guggenheim, Anderson,
relatively little was known about the caking property of complex food system, and de Boer model (Eq. (1)),
though it is known that caking of powder can be characterized by glass
transition. For example, Jaya and Das (2004) investigated effect of MD and
We W CK am⋅ ⋅ ⋅ w
other additives on the caking property of vacuum-dried mango, but there was
no consideration of glass transition. On the other hand, Farahnaky et al. (2016) (1)
investigated effect of MD content on the Tg and caking of date syrup powder, where We (g-water/g-solid) is the equilibrium water content, Wm (gwater/g-
but the caking was investigated at a constant aw-condition. solid) is the amount of water adsorbed strongly to specific sites on the surface
For a better understating of the physical properties of freeze-dried mango of the material (i.e., monolayer water), and C and K are correction factors for
powder, mango solute (MS) of which pulp was removed from MP was monolayer and multilayer sorption properties, respectively (Cano-Higuita et al.,
prepared, and the water sorption behavior and Tg of freezedried MS-MD 2015).
mixtures were investigated systematically in this study. In addition, the caking
property of freeze-dried MP-MD mixtures and MSMD mixtures was 2.5. Glass transition temperature (Tg)
investigated. Furthermore, the applicability of an empirical model to predict the
effect of water content on Tg was confirmed. The effect of water content on the Tg of MS-MD samples was investigated
using differential scanning calorimetry (DSC; DSC 120, Seiko Instruments Inc.,
2. Materials and methods Tokyo, Japan) with alumina powder as a reference material. The DSC
temperature and heat flux were calibrated using distilled water and indium. The
2.1. Materials
samples (5–10 mg) were placed into aluminum DSC pans and hermetically
MS (Mangifera indica L. var. Alphonso cultivated in India, 19.2 °Bx) was enclosed. To evaluate the Tg of the anhydrous sample (Tg(as)), the vacuum-dried
provided by Taiyo Kagaku Co., Ltd., Mie, Japan. MD with a dextrose sample was placed into a DSC pan and then further dried at 105 °C for 12 h
equivalent of 17–21 was provided by San-ei Sucrochemical, Co., Ltd., Aichi, before enclosing the pan. DSC measurements were carried out at a heating rate
of 3 °C/min. To reset of the thermal history of glassy samples, DSC scan was
Japan.
repeated two times. In the first scan, DSC measurement was stopped at a
slightly higher temperature than an endothermic shift expected as glass
96
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

transition, and then cooled down. In the second scan, DSC measurement was
stopped at a much higher temperature than the Tg suggested by first scan. The
Tg was determined from the onset point of the endothermic shift observed in the
second scanning. The measurements were conducted in duplicate, and the
results were averaged.
Effect of water content on the Tg was analyzed by the
Gordon–Taylor (GT) model (Eq. (2)),

(1 − X Tw) ⋅ g as( ) + k X T⋅ w⋅ g w( )

Tg =
(1 − Xw) + kX⋅ w (2)

where Tg(as), Tg(w), Xw, and k are the Tg of the anhydrous sample (K), the Tg of
water (K), the weight fraction of water (dimensionless), and a constant
(dimensionless), respectively. Tg(as) was determined experimentally, and Tg(w)
was set to 136 K based on the literature value (Johari et al., 1987; Sastry, 1999).

2.6. Degree of caking

The degree of caking was evaluated in both MP-MD and MS-MD samples
using a previously reported method (Farahnaky et al., 2016) with minor Fig. 1. Water sorption isotherms of MS-MD samples at 25 °C. The solid lines were obtained
modifications. The freeze-dried powder was vacuum-dried at 60 °C for 6 h, and by fitting the GAB equation to the experimental data. The values are expressed as mean ±
SD (n = 3).
the fully dried powder was then manually sieved (aperture size = 0.5 mm). Each
sample (approximately 0.5 g) was equilibrated under various a w conditions at
25 °C according to the above-mentioned procedures. The equilibrated samples Table 1
were vacuumdried at 25 °C for 6 h. The dried samples were then weighed Wi GAB parameters for MS-MD and MP-MD samples.
and sieved (aperture size = 1.4 mm) at the vibration amplitude of 4.5 mm for 5
min using a sieve machine (model MVS-1, AS ONE Corp., Osaka, Japan). The
sample retained on the sieve was weighed (Wr), and the degree of caking was
calculated as.

Degree of caking (%) = 100 × Wr/Wi (3)

The measurements were performed in triplicate, and the results were


averaged. 0.2 MD 0.130 2.303 0.922 0.999 0.0026

0.4 MD 0.098 2.268 0.958 0.999 0.0024


3. Results and discussion
0.6 MD 0.081 2.640 0.950 0.998 0.0030
3.1. Effects of MD and pulp on the water sorption isotherm of freeze-dried
mango powder 0.8 MD 0.064 3.414 0.915 0.997 0.0033
MP-MD
0.0 MD (MP) 0.087 8.841 1.009 0.999 0.0051 Fongin et al.
The water sorption isotherms of the freeze-dried MS-MD samples at 0.2 MD 0.077 8.052 1.007 0.999 0.0033 (2017)
25 °C are shown in Fig. 1. The water content increased with increasing aw, and 0.4 MD 0.079 2.995 0.98 0.999 0.0029
the water sorption isotherm shifted downwards with increased MD content.
When the crystallization of solutes occurs during water sorption, water content 0.6 MD 0.072 2.123 0.979 0.999 0.0026

of the sample drastically decreases (Fukami et al., 2016). Thus, it is thought 0.8 MD 0.055 8.844 0.956 0.997 0.0041
that the MS-MD samples maintained mostly amorphous during water sorption
as discussed in the latter section. The solid lines in Fig. 1 were obtained by 1.0 MD 0.043 9.192 0.995 0.999 0.0028
fitting the GAB model (Eq. (1)) to the data. The obtained GAB parameters (Wm,
C, and K) are summarized in Table 1.
The Wm values of the MS-MD samples varied from 0.043 to 0.161 gwater/g-
R2: the coefficient of determination; RMSE: the root means square error.
solid and decreased with increasing MD content. This relationship between MD
content and Wm is similar to those reported for other dried fruits such as freeze-
2005). Similar results have been reported for various dried fruits (Fabra et al.,
dried borojó (Mosquera et al., 2010), spray-dried noni (Fabra et al., 2011),
2011; Farahnaky et al., 2016; Mosquera et al., 2010).
spray-dried mango (Cano-Higuita et al., 2015) and drum-dried date syrup
For comparison, the previously reported GAB parameters of the MPMD
(Farahnaky et al., 2016). The value of C increased with increasing MD content.
samples are also listed in Table 1. The values of Wm and C were higher for MS-
Higher values of C indicate stronger bonds between water molecules and the
MD than for MP-MD. Since MP contains a large amount of water-insoluble
binding sites (Quirijns et al., 2005). Thus, the effects of MD on Wm and C can
components (e.g., insoluble pectin, hemicellulose, and cellulose), the
be attributed to the fact that MD is much less hygroscopic than mango solute.
equilibrium water content of MP at a given aw is lower than that of MS.
In contrast, K ranged from 0.915 to 0.995 with only small variations for
different MD contents. The K values close to 1 indicate that the multilayer water
tends to behave as liquid water (Quirijns et al., 3.2. Effects of MD and pulp on the anhydrous Tg of freeze-dried mango
powder

97
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

Typical DSC thermograms for the anhydrous MS-MD samples are shown 2008a), spray-dried bayberry (Fang and Bhandari, 2012), and vacuum spray-
in Fig. 2-a (first scan) and -b (second scan). As mentioned above, DSC scan dried orange juice (Islam et al., 2017). In addition, the Tg(as) curve of the MS-
was repeated two times. In the first scan, glassy samples may show an MD system showed discontinuous behavior with a step-point between MD
endothermic peak (recovery of relaxed enthalpy) at their Tg depending on the weight fractions of 0.6 and 0.7. Similar characteristics were observed for the
thermal history (preparation temperature and time); the endothermic peak MP-MD system along with
affects their Tg value. However there was no or little effect of thermal history

Fig. 2. Typical DSC thermograms of the first scan (a) and second scan (b) for anhydrous MS-MD samples.
on the glass transition behavior of the MSMD samples (Fig. 2-a), because the
samples were held at 105 °C for the fully dehydration; the thermal history of
the glassy samples was reset at the dehydration stage. In the second scan, glass
transition of the glassy samples cooled at 3 °C/min is detected. As the result,
clear endothermic shift was detected, and anhydrous Tg (Tg(as)) was evaluated
from the onset point.
Since MS has a large amount of low-molecular-weight carbohydrates
(sucrose, glucose, and fructose), there was a possibility of their crystallization
during the dehydration at 105 °C. This possibility could be excluded primarily
from the clear glass transition behavior in the DSC thermograms (Fig. 2-b).
Other water soluble materials (e.g., water soluble pectin and oligosaccharide)
would have prevented crystallization of the low-molecular-weight
carbohydrates.
From the second scan, it was found that the MS (0.0 MD) and 0.2 MD
samples showed a slightly larger magnitude of glass transition in a lower
temperature range than the others. The temperature range of glass transition for
the 0.4 MD and 0.6 MD samples was extended. The 0.8 MD sample was lower
magnitude of glass transition than the others. These observations can be
explained by the change of amorphous structure in the MS-MD samples as
discussed in the latter part of this section.
Effect of MD content on the Tg(as) of freeze-dried MS and MP (Fongin et al.,
2017) is shown in Fig. 3. The Tg(as) for MS without MD was 37.2 °C, which falls Fig. 3. Effect of MD on the Tg(as) of MS-MD and MP-MD samples. The solid lines are
guides for the eyes.
within the range of Tg(as) for low-molecular-weight carbohydrates (e.g., 62 °C
for sucrose, 31 °C for glucose, and 5 °C for fructose) (Roos, 1995b), which are
major water-soluble carbohydrates in mango (Jaya and Das, 2008). glucose-MD and maltose-MD (Kawai and Hagura, 2012) and was discussed in
The Tg(as) values of the MS-MD samples were lower than those of the MP- more detail in our previous study (Fongin et al., 2017). Based on the discussion,
MD samples (Fig. 3). This result is discussed with model drawings shown in the model drawings of anhydrous MS-MD system are shown in Fig. 5. At MD
Fig. 4. Since pulp materials (e.g., insoluble pectin, hemicellulose, and cellulose) weight fractions exceeding 0.7, the amorphous mango materials are entrapped
are intrinsically crystalline, low-molecularweight carbohydrates will not by the large amount of MD and so MD was plasticized by the amorphous
interact with pulps; Tg(as) should be independent from pulp content. Although mango materials. Thus, the Tg(as) of the system decreased with increasing MS
there is a similar report on effect of pulp on the Tg of blueberry powder (Tao et content. When the MD weight fraction was less than 0.6, a part of the
al., 2018), the molecular mechanism is unclear. As one of possibilities, pulps amorphous mango materials is not entrapped by MD; the amorphous mixture
have some amorphous parts such as the tips of pulp segments. Consequently, contained both amorphous mango materials-MD domain and amorphous
pulp materials are thought to increase slightly the Tg of amorphous mango part mango materialsrich domain. The lower Tg(as) of the amorphous mango
(mainly low-molecular-weight carbohydrates). materials-rich domain compared to the amorphous mango materials-MD
As expected, Tg(as) increased with increasing MD content (Fig. 3). Similar domain will have caused the abrupt change in Tg(as) observed between MD
results have been reported for spray-dried tomato (Goula and Adamopoulos, weight fractions of 0.6 and 0.7. The change of amorphous structure can be

98
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

corresponded to the extended glass transition behavior observed in DSC curves is also supported by other researchers (Hughes et al., 2016, 2018; Tedeschi et
for 0.4 MD and 0.6 MD; this will be a result of continuous glass transitions. al., 2016). For example, Hughes et al. (2018) investigated the Tg of
The heterogeneous glass transition observed in sugarpolysaccharide mixtures hydrophobically modified starch-sucrose

Fig. 4. Model drawings for anhydrous MP and MS.

Fig. 5. Model drawings for anhydrous MS-MD.


Fig. 6. Effect of water content on the Tg of MS-MD samples. The solid lines were obtained
by fitting the GT equation to the experimental data.
mixtures by DSC, and reported a contentious glass transition behavior. They not show discontinuous drop (Fig. 1). As discussed above, other water soluble
divided the broad glass transition into two glass transitions by Gaussian curves. materials will have prevented their crystallization.
In addition, two different glass transitions were found at low sucrose contents Effect of water content on the Tg of the MS-MD samples is shown in Fig. 6.
and/or high aw conditions. The Tg decreased with increasing water content because of the plasticizing
effect of water. In addition, Tg increased with increasing MD content. The
3.3. Effects of MD and pulp on the Tg curve of freeze-dried mango powder solid lines in Fig. 3 were obtained by fitting the GT model (Eq. (2)) to the
data. The Tg(as) and k values of the MS-MD and MP-MD samples are listed in
Table 2.
The effect of water content on the Tg of MS-MD samples was investigated
For both MS and MP, k increased with increasing MD content. As k reflects
in the aw range between 0.00 and 0.432 at intervals of 0.2 MD weight fraction.
the sensitivity to the plasticizing effect of water, a higher k value corresponds
DSC thermograms showed a clear endothermic shift due to glass transition
to more Tg-depression resulting from water sorption (Ahmed and Rahman,
(data not shown), and Tg was evaluated from the onset point as similar to Fig.
2014). The k values of MS and MP exhibited only small differences attributed
2. From the glass transition behavior, there was no clear evidence for
to the difference in Tg(as) between MS and MP; it is empirically known that k
crystallization of the low-molecular-weight carbohydrates during water
sorption. This is also supported by the fact that water sorption isotherms did and Tg(as) are linearly related (Roos, 1993a).

99
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

As mentioned above, the water content and aw at which Tg = 25 °C are Fig. 7-a. For MP samples with MD weight fractions of 0 and 0.2, the degree of
commonly denoted as Wc and awc respectively (Fabra et al., 2011; Roos, 1993b, caking increased with increasing aw at higher aw than 0.11. When the MD weight
1995b). Higher values of Wc and awc indicate greater resistance to physical fraction was 0.4–1.0, the degree of caking increased drastically above a certain
deterioration (e.g., the caking of amorphous powder and recrystallization of aw condition. The aw value at which the degree of caking begins to increase
sugar) induced by water uptake (Palzer, drastically corresponds to awc (Table 2). This indicates that the degree of caking
Table 2 of an amorphous powder depends strongly on Tg (Aguilera et al., 1995; Peleg,
Tg(as), k, Wc, and awc, for MS-MD and MP-MD samples. 1993; Roos and Karel, 1991).
The effect of aw on the degree of caking of MS-MD is shown in Fig. 7-b.
The MS caked completely, even at the lowest aw. Although the addition of MD
reduced the degree of caking, the MS-MD samples were more sensitive to
(°C) (g- caking than MP-MD, likely because of their different
water/gsolid)
Tg values.
To confirm the relationship between the degree of caking and Tg, the degree
MS-MD of caking is plotted against 25−Tg (°C) in Fig. 8-a and -b. The 25 °C was
employed as a typical room temperature. When 25−Tg is negative, the powder
0.0 MD (MS) 37.2 3.86 0.997 2.103 0.019 0.074 This study
is in a glassy state, and a free-flowing powder can be expected (Aguilera et al.,
0.2 MD 47.1 4.07 0.982 3.913 0.033 0.120
1995; Foster et al., 2006; Haider et al., 2014). Conversely, when 25−Tg is
0.4 MD 52.3 4.95 0.996 1.935 0.034 0.160 positive, the sample is in rubbery state, and some degree of caking will occur
depending on the value of 25−Tg. As expected, the degree of caking of freeze-
0.6 MD 62.1 5.57 0.979 4.848 0.041 0.215
dried mango powder changed drastically at 25−Tg =0. Although it is expected
0.8 MD 92.9 7.18 0.979 4.979 0.058 0.345 that amorphous powder is physically stable below Tg, some MP-MD and
MP-MD MSMD samples except for MD showed a small degree of caking even below
0.0 MD (MP) 48.9 3.13 0.985 3.449 0.047 0.095 Fongin Tg (Fig. 8-c and -d).
0.2 MD 58.1 4.55 0.992 3.605 0.045 0.115 et al.
69.3 5.56 0.993 3.377 0.049 0.235 (2017)
Two possible reasons for the small degree of caking below Tg are suggested.
0.4 MD
0.6 MD 76.1 6.08 0.993 3.282 0.052 0.310 First reason is resulted from the slow dynamics below Tg. Tg
is a temperature of which viscosity becomes 1012 Pa s (Angell et al., 1994). This
0.8 MD 108.5 6.44 0.998 2.274 0.080 0.430 is too high viscosity, but not freezing. For example, slow molecular dynamics
1.0 MD 162.6 9.00 0.995 4.330 0.094 0.575
of amorphous materials below Tg can be observed as volume relaxation
(densification) and enthalpy relaxation (Lourdin et al., 2002; Noel et al., 2005;
Kawai et al., 2005). As a result of the progress of relaxation, physical strength
of the amorphous materials is enhanced (Lourdin et al., 2002; Noel et al., 2005).
R2: coefficient of determination; RMSE: the root means square error. Recently, Schulnies and Kleinschmidt (2018) investigated effect of storage
time on the consolidation of skim milk near the Tg. From the results, it was
2005; Roos, 1995a). The Wc and awc values of MS-MD samples varied from found that physical strength of the consolidated skim milk increased with an
0.019 to 0.094 g-water/g-solid and 0.074–0.575, respectively. As observed for increase in storage time even below the Tg. They suggested that relaxation
MP, Wc and awc increased with increasing MD content. Similar results have process contributed to the enhancement of the consolidated skim milk below
been reported for strawberries (Mosquera et al., 2012) and grapefruit (Telis and Tg due to small changes in the contact area between particles. It should be noted
Martínez-Navarrete, 2009). Compared to the MP-MD samples, Wc and awc were that the degree of caking is affected by the physical strength of cake (Fitzpatrick
lower for the MS-MD samples. These results are attributed to the increase in Tg et al., 2007; Schulnies and Kleinschmidt, 2018). If the powder forms a fragile
resulting from the addition of pulp. cake, the cake will be destructed by the sieving treatment. In this case, there is
no contribution to the degree of caking even though caking occurs. However,
the small degree of caking may be observed when physical strength of the cake
3.4. Effects of MD and pulp on the degree of caking of freeze-dried mango
is enhanced by the progress of relaxation. The reason why the MD sample
powder
showed negligible degree of caking below the Tg will be that the MD sample
Although the degree of caking will depend on the sieving conditions (e.g., was more fragile solid than MP-MD and MP-MS samples. For example, it is
sieving speed and amplitude) to some extent, comparing the degree of caking known that physical structure of amorphous polymer can be more packed by
at a constant sieving condition among different samples can provide valuable the addition of low-molecular-weight materials (Figueroa et al., 2016).
information. The effect of aw on the degree of caking of MP-MD is shown in

100
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

Fig. 7. Effect of aw on the degree of caking of MP-MD samples (a) and MS-MD samples (b). The values are expressed as mean ± SD (n = 3).

Fig. 8. The relationship between 25-Tg and the degree of caking for MP-MD samples (a) and MS-MD samples (b). The degree of caking lower than 20% is emphasized for MP-MD samples
(c) and MS-MD samples (d). The values are expressed as mean ± SD (n = 3).
The other reason is originated from the heterogeneous molecular mobility the Tg depending on the storage conditions. The small degree of caking below
of amorphous mixtures. The Tg for mixture systems does not always critical Tg was negligible at approximately 20 °C lower temperature than Tg (Fig. 8-c
parameter for the molecular mobility. For example, it is suggested that low- and -d). When the mango powder samples are stored at 25 °C, it is desirable
molecular-weight materials in polymer have relatively high molecular mobility that the Tg is more than 45 °C.
even at a temperature below the Tg (Kawai et al., 2004, 2006). Since caking of
amorphous powder is time dependent phenomenon (Bhandari and Howes, 3.5. Prediction of the Tg curve of freeze-dried mango powder
1999; Palzer, 2005), low-molecular-weight materials having relatively high
molecular mobility may have contributed to the small degree of caking below
101
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

As mentioned above, k and Tg(as) are linearly related (Roos, 1993a). We Acknowledgements
obtained previously the following empirical linear relationship between k
(dimensionless) and Tg(as) (K) for many dry fruits including MP-MD samples This study was financially supported by JSPS KAKENHI (Grant-inAid for
(Fongin et al., 2017). Scientific Research C: 15K07453). We wish to acknowledge Taiyo Kagaku
Co., Ltd. and San-ei Sucrochemical, Co., Ltd. for supporting alphonso mango
k = 0.0825Tg(as) − 22.518 (4)
puree and maltodextrin, respectively.
When Eq. (4) is combined with Eq. (2), the effect of water content on Tg
References
can be calculated using only Tg(as). It is suggested that Eq. (4) can be applied to
dry fruits including pulp, because completely water soluble materials (e.g., low- Aguilera, J., del Valle, J., Karel, M., 1995. Caking phenomena in amorphous food powders.
molecular-weight carbohydrate and MD having high dextrose equivalent) Trends Food Sci. Technol. 6 (5), 149–155. https://doi.org/10.1016/S09242244(00)89023-8.
deviated from the behavior. Ahmed, J., Rahman, M.S., 2014. Glass transition in foods. In: Rao, M.A., Rizvi, S.S.H., Datta,
A.K., Ahmed, J. (Eds.), Engineering Properties of Foods, fourth ed. CRC Press, pp. 93–120.
In order to obtain new insights into the subject, the Tg values of MSMD Akhavan Mahdavi, S., Jafari, S.M., Assadpoor, E., Dehnad, D., 2016. Microencapsulation
samples were calculated at each water content from experimental Tg(as) values optimization of natural anthocyanins with maltodextrin, gum Arabic and gelatin. Int. J. Biol.
(Table 2), and they were plotted against their measured Tg values (Fig. 9-a). It Macromol. 85, 379–385. https://doi.org/10.1016/j.ijbiomac.2016.01.011.
Angell, C.A., Bressel, R.D., Green, J.L., Kanno, H., Oguni, M., Sare, E.J., 1994. Liquid fragility
was confirmed that the predicted Tg values were in good agreement with the
and the glass transition in water and aqueous solutions. J. Food Eng. 22 (1), 115–142.
measured values (coefficient of determination, R2 = 0.919). Furthermore, the https://doi.org/10.1016/0260-8774(94)90028-0.
Wc values of MS-MD samples were calculated, and they were converted to awc Berardini, N., Fezer, R., Conrad, J., Beifuss, U., Carle, R., Schieber, A., 2005. Screening of mango
values using Eq. (1) with GAB parameters (Table 1). From Fig. 9-b, it was (Mangifera indica L.) cultivars for their contents of flavonol o- and xanthone cglycosides,
anthocyanins, and pectin. J. Agric. Food Chem. 53 (5), 1563–1570.
confirmed that the predicted awc values were in good agreement with the https://doi.org/10.1021/jf0484069.
experimentally determined awc values (R2 =0.960). From the results, it was Bhandari, B.R., Howes, T., 1999. Implication of glass transition for the drying and stability of
noted that Eq. (4) can be applied to dry fruit powders with independent of pulp. dried foods. J. Food Eng. 40 (1–2), 71–79. https://doi.org/10.1016/S02608774(99)00039-4.
This model will be useful for understanding the physical stability (e.g., degree Bhusari, S.N., Muzaffar, K., Kumar, P., 2014. Effect of Carrier agents on physical and
microstructural properties of spray dried tamarind pulp powder. Powder Technol.
of caking) of dried fruit powders based on a minimum experiment (i.e., 266, 354–364. https://doi.org/10.1016/j.powtec.2014.06.038.
determination of Tg(as)). Since amorphous materials in dry fruit powders are Brennan, J.G., Herrera, J., Jowitt, R., 1971. A study of some of the factors affecting the spray
mainly sucrose, glucose, and fructose, it is practically effective to understand drying of concentrated orange juice, on a laboratory scale. Int. J. Food Sci.
Technol. 6 (3), 295–307. https://doi.org/10.1111/j.1365-2621.1971.tb01618.x.
the Tg(as) of sucrose-glucose-fructose ternary system.
Cano-Higuita, D.M., Villa-Vélez, H.A., Telis-Romero, J., Váquiro, H.A., Telis, V.R.N.,
2015. Influence of alternative drying aids on water sorption of spray dried mango mix
powders: a thermodynamic approach. Food Bioprod. Process. 93, 19–28. https://doi.
4. Conclusion org/10.1016/j.fbp.2013.10.005.
Costa, J.d.P.d., Rocha, É.M.d.F.F., Costa, J.M.C.d., 2014. Study of the physicochemical
The water sorption, glass transition, and caking characteristics of freeze- characteristics of soursop powder obtained by spray-drying. Food Sci. Technol. 34 (4),
dried MP-MD and MS-MD powders were investigated. The results 663–666. https://doi.org/10.1590/1678-457x.6380.
Djendoubi Mrad, N., Bonazzi, C., Boudhrioua, N., Kechaou, N., Courtois, F., 2012. Influence of
demonstrate that both MD and pulp increase the Tg and improve the caking
sugar composition on water sorption isotherms and on glass transition in apricots. J. Food
property of freeze-dried mango powder. Although the caking of freeze-dried Eng. 111 (2), 403–411. https://doi.org/10.1016/j.jfoodeng.2012.
mango powders could be characterized by their Tg, minor 02.001.

Fig. 9. The relationship between predicted Tg and measured Tg for MS-MD samples (a) and the relationship between predicted awc and experimentally determined
awc for MS-MD samples (b). Fabra, M.J., Márquez, E., Castro, D., Chiralt, A., 2011. Effect of maltodextrins in the water-
content–water activity–glass transition relationships of noni (Morinda citrifolia L.) pulp
powder. J. Food Eng. 103 (1), 47–51. https://doi.org/10.1016/j.jfoodeng.
caking was observed even below their Tg. An empirical predictive approach for 2010.09.018.
determining the Tg curve of dry fruits was shown to be applicable to MS-MD Fang, Z., Bhandari, B., 2012. Comparing the efficiency of protein and maltodextrin on spray
samples. This approach provides a simple way to evaluate the degree of caking drying of bayberry juice. Food Res. Int. 48 (2), 478–483. https://doi.org/10.
1016/j.foodres.2012.05.025.
of dried fruit powders. Farahnaky, A., Mansoori, N., Majzoobi, M., Badii, F., 2016. Physicochemical and sorption
isotherm properties of date syrup powder: antiplasticizing effect of maltodextrin. Food
Bioprod. Process. 98, 133–141. https://doi.org/10.1016/j.fbp.2016.01.003.

102
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

Figueroa, Y., Guevara, M., Pérez, A., Cova, A., Sandoval, A.J., Müller, A.J., 2016. Effect of sugar Mosquera, L.H., Moraga, G., Martínez-Navarrete, N., 2012. Critical water activity and critical
addition on glass transition temperatures of cassava starch with low to intermediate moisture water content of freeze-dried strawberry powder as affected by maltodextrin and Arabic
contents. Carbohydr. Polym 146, 231–237. https://doi.org/10. gum. Food Res. Int. 47 (2), 201–206. https://doi.org/10.1016/j.foodres. 2011.05.019.
1016/j.carbpol.2016.03.054. Noel, T.R., Parker, R., Brownsey, G.J., Farhat, I.A., MacNaughtan, W., Ring, S.G., 2005.
Fitzpatrick, J.J., Hodnett, M., Twomey, M., Cerqueira, P.S.M., O'Flynn, J., Roos, Y.H., 2007. Physical aging of starch, maltodextrin, and maltose. J. Agric. Food Chem. 53 (22), 8580–
Glass transition and the flowability and caking of powders containing amorphous lactose. 8585. https://doi.org/10.1021/jf0580770.
Powder Technol 178 (2), 119–128. https://doi.org/10.1016/j.powtec. 2007.04.017. Oliveira, D.M., Clemente, E., da Costa, J.M., 2014. Hygroscopic behavior and degree of caking
Fongin, S., Kawai, K., Harnkarnsujarit, N., Hagura, Y., 2017. Effects of water and maltodextrin of grugru palm (Acrocomia aculeata) powder. J. Food Sci. Technol. 51 (10), 2783–2789.
on the glass transition temperature of freeze-dried mango pulp and an empirical model to https://doi.org/10.1007/s13197-012-0814-9.
predict plasticizing effect of water on dried fruits. J. Food Eng. Omar, E.A.M., Roos, Y.H., 2007. Glass transition and crystallization behaviour of freezedried
210, 91–97. https://doi.org/10.1016/j.jfoodeng.2017.04.025. lactose-salt mixtures. LWT - Food Sci. Technol. (Lebensmittel-Wissenschaft -Technol.) 40
Foster, K.D., Bronlund, J.E., Paterson, A.H.J., 2006. Glass transition related cohesion of (3), 536–543. https://doi.org/10.1016/j.lwt.2005.12.007.
amorphous sugar powders. J. Food Eng. 77 (4), 997–1006. https://doi.org/10.1016/ Palzer, S., 2005. The effect of glass transition on the desired and undesired agglomeration of
j.jfoodeng.2005.08.028. amorphous food powders. Chem. Eng. Sci. 60 (14), 3959–3968. https://doi.org/
Fukami, K., Kawai, K., Takeuchi, S., Harada, Y., Hagura, Y., 2016. Effect of water content on 10.1016/j.ces.2005.02.015.
the glass transition temperature of calcium maltobionate and its application to the Palzer, S., Sommer, K., 2011. Caking of water-soluble amorphous and crystalline food powders.
characterization of non-arrhenius viscosity behavior. Food Biophys. 11 (4), 410–416. In: Aguilera, J.M., Simpson, R., Welti-Chanes, J., Bermudez-Aguirre, D., Barbosa-
https://doi.org/10.1007/s11483-016-9455-2. Canovas, G. (Eds.), Food Engineering Interfaces. Springer New York, New York, NY, pp.
Goula, A.M., Adamopoulos, K.G., 2008a. Effect of maltodextrin addition during spray drying of 491–514.
tomato pulp in dehumidified air: I. Drying kinetics and product recovery. Paterson, A.H.J., Bröckel, U., 2015. Caking development in lemon juice powder. Procedia Eng
102, 142–149. https://doi.org/10.1016/j.proeng.2015.01.117.
Dry. Technol. 26 (6), 714–725. https://doi.org/10.1080/07373930802046369.
Peleg, M., 1993. Glass transition of the physical stability of food powders. In: Blanshard, J.M.V.,
Goula, A.M., Adamopoulos, K.G., 2008b. Effect of maltodextrin addition during spray drying of
Lillford, P.J. (Eds.), The Glassy State in Foods. Nottingham University Press, Leicestershire,
tomato pulp in dehumidified air: II. Powder properties. Dry. Technol. 26 (6), 726–737.
pp. 435–451.
https://doi.org/10.1080/07373930802046377.
Peleg, Y., Mannheim, C.H., 1969. Caking of onion powder. Int. J. Food Sci. Technol. 4 (2),
Goula, A.M., Adamopoulos, K.G., 2010. A new technique for spray drying orange juice
157–160. https://doi.org/10.1111/j.1365-2621.1969.tb01509.x.
concentrate. Innovat. Food Sci. Emerg. Technol. 11 (2), 342–351. https://doi.org/10.
Pott, I., Marx, M., Neidhart, S., Mühlbauer, W., Carle, R., 2003. Quantitative determination of
1016/j.ifset.2009.12.001. β-carotene stereoisomers in fresh, dried, and solar-dried mangoes (Mangifera indica L.). J.
Haider, C.I., Hounslow, M.J., Salman, A.D., Althaus, T.O., Niederreiter, G., Palzer, S., 2014.
Agric. Food Chem. 51 (16), 4527–4531. https://doi.org/10. 1021/jf034084h.
Influence of environmental conditions on caking mechanisms in individual amorphous food
Quirijns, E.J., van Boxtel, A.J.B., van Loon, W.K.P., van Straten, G., 2005. Sorption isotherms,
particle contacts. AIChE J. 60 (8), 2774–2787. https://doi.org/10. 1002/aic.14490.
GAB parameters and isosteric heat of sorption. J. Sci. Food Agric. 85 (11), 1805–1814.
Hughes, D., Tedeschi, C., Leuenberger, B., Roussenova, M., Coveney, A., Richardson, R., https://doi.org10.1002/jsfa.2140.
Ubbink, J., 2016. Amorphous-amorphous phase separation in hydrophobicallymodified Rocha Ribeiro, S.M., Queiroz, J.H., Lopes Ribeiro de Queiroz, M.E., Campos, F.M.,
starch–sucrose blends II. Crystallinity and local free volume investigation using wide-angle Pinheiro Sant'ana, H.M., 2007. Antioxidant in mango (Mangifera indica L.) pulp. Plant
X-ray scattering and positron annihilation lifetime spectroscopy. Food Hydrocolloids 58, Foods Hum. Nutr. 62 (1), 13–17. https://doi.org/10.1007/s11130-006-0035-3.
316–323. https://doi.org/10.1016/j.foodhyd.2016.02.024. Roos, Y.H., 1993a. Melting and glass transitions of low molecular weight carbohydrates.
Hughes, D.J., Bönisch, G.B., Zwick, T., Schäfer, C., Tedeschi, C., Leuenberger, B., Ubbink, J., Carbohydr. Res. 238, 39–48. https://doi.org/10.1016/0008-6215(93)87004-C.
2018. Phase separation in amorphous hydrophobically modified starch–sucrose blends: glass Roos, Y.H., 1995a. Characterization of food polymers using state diagrams. J. Food Eng.
transition, matrix dynamics and phase behavior. Carbohydr. Polym. 199, 1–10. 24 (3), 339–360. https://doi.org/10.1016/0260-8774(95)90050-L.
https://doi.org/10.1016/j.carbpol.2018.06.056. Roos, Y.H., Karel, M., 1991. Water and molecular weight effects on glass transitions in
Islam, M.Z., Kitamura, Y., Kokawa, M., Monalisa, K., Tsai, F.-H., Miyamura, S., 2017. Effects amorphous carbohydrates and carbohydrate solutions. J. Food Sci. 56 (6), 1676–1681.
of micro wet milling and vacuum spray drying on the physicochemical and antioxidant https://doi.org/10.1111/j.1365-2621.1991.tb08669.x.
properties of orange (Citrus unshiu) juice with pulp powder. Food Bioprod. Process. 101, Roos, Y.H., 1993b. Water activity and physical state effects on amorphous food stability. J. Food
132–144. https://doi.org/10.1016/j.fbp.2016.11.002. Process. Preserv. 16 (6), 433–447. https://doi.org/10.1111/j.1745-4549. 1993.tb00221.x.
Jaya, S., Das, H., 2004. Effect of maltodextrin, glycerol monostearate and tricalcium phosphate
Roos, Y.H., 1995b. Phase Transition in Foods. Academic Press, San Diego.
on vacuum dried mango powder properties. J. Food Eng. 63 (2), 125–134. Roos, Y.H., Silalai, N., 2011. Glass transitions: opportunities and challenges. In: Aguilera, J.M.,
https://doi.org/10.1016/s0260-8774(03)00135-3. Simpson, R., Welti-Chanes, J., Bermudez-Aguirre, D., Barbosa-Canovas, G.
Jaya, S., Das, H., 2008. Glass transition and sticky point temperatures and stability/ mobility (Eds.), Food Engineering Interfaces. Springer New York, New York, NY, pp. 473–490.
diagram of fruit powders. Food Bioprocess Technol. 2 (1), 89–95. https:// Sagar, V.R., Kumar, R., 2014. Effect of drying treatments and storage stability on quality
doi.org/10.1007/s11947-007-0047-5. characteristics of bael powder. J. Food Sci. Technol. 51 (9), 2162–2168. https://doi.
Johari, G.P., Hallbrucker, A., Mayer, E., 1987. The glass–iquid transition of hyperquenched org/10.1007/s13197-012-0727-7.
water. Nature 330 (6148), 552–553. https://doi.org/10.1038/ 330552a0. Sahu, J.K., 2008. The effect of additives on vacuum dried honey powder properties. Int. J.
Kawai, K., Hagiwara, T., Takai, R., Suzuki, T., 2004. Maillard reaction rate in various glassy Food Eng. 4 (8), 1–26. https://doi.org/10.2202/1556-3758.1356.
matrices. Biosci. Biotechnol. Biochem. 68 (11), 2285–2288. https://doi.org/ Sastry, S., 1999. Supercooled water: going strong or falling apart? Nature 398 (6727), 467–469.
10.1271/bbb.68.2285. https://doi.org/10.1038/18982.
Kawai, K., Hagiwara, T., Takai, R., Suzuki, T., 2005. Comparative investigation by two analytical Schulnies, F., Kleinschmidt, T., 2018. Time consolidation of skim milk powder near the glass
approaches of enthalpy relaxation for glassy glucose, sucrose, maltose, and trehalose. transition temperature. Int. Dairy J 85, 105–111. https://doi.org/10.1016/j.
Pharmaceut. Res. 22 (3), 490–495. https://doi.org/10.1007/s11095-0041887-6. idairyj.2018.05.005.
Kawai, K., Hagura, Y., 2012. Discontinuous and heterogeneous glass transition behavior of Tao, Y., Wu, Y., Yang, J., Jiang, N., Wang, Q., Chu, D.-T., Zhou, J., 2018. Thermodynamic
carbohydrate polymer–plasticizer systems. Carbohydr. Polym. 89 (3), 836–841. sorption properties, water plasticizing effect and particle characteristics of blueberry powders
https://doi.org/10.1016/j.carbpol.2012.04.018. produced from juices, fruits and pomaces. Powder Technol. 323, 208–218.
Kawai, K., Suzuki, T., Oguni, M., 2006. Low-temperature glass transitions of quenched and https://doi.org/10.1016/j.powtec.2017.09.033.
annealed bovine serum albumin aqueous solutions. Biophys. J. 90 (10), 3732–3738. Tedeschi, C., Leuenberger, B., Ubbink, J., 2016. Amorphous–amorphous phase separation in
https://doi.org/10.1529/biophysj.105.075986.
hydrophobically-modified starch–sucrose blends I. Phase behavior and thermodynamic
Kim, Y., Lounds-Singleton, A.J., Talcott, S.T., 2009. Antioxidant phytochemical and quality
characterization. Food Hydrocolloids 58, 75–88. https://doi.org/10.1016/j.
changes associated with hot water immersion treatment of mangoes (Mangifera indica L.).
foodhyd.2016.02.021.
Food Chem. 115 (3), 989–993. https://doi.org/10.1016/j. foodchem.2009.01.019.
Telis, V., Martínez-Navarrete, N., 2012. Biopolymers Used as Drying Aids in Spray-Drying and
Kurozawa, L.E., Hubinger, M.D., Park, K.J., 2012. Glass transition phenomenon on shrinkage of
Freeze-drying of Fruit Juices and Pulps Biopolymer Engineering in Food Processing. CRC
papaya during convective drying. J. Food Eng. 108 (1), 43–50. https://
Press, pp. 279–326.
doi.org/10.1016/j.jfoodeng.2011.07.033. Telis, V.R.N., Martínez-Navarrete, N., 2009. Collapse and color changes in grapefruit juice
Lourdin, D., Colonna, P., Brownsey, G.J., Noel, T.R., Ring, S.G., 2002. Structural relaxation and powder as affected by water activity, glass transition, and addition of carbohydrate polymers.
physical ageing of starchy materials. Carbohydr. Res. 337 (9), 827–833.
Food Biophys. 4 (2), 83–93. https://doi.org/10.1007/s11483-009-9104-0.
https://doi.org/10.1016/S0008-6215(02)00064-2.
Valente, A., Albuquerque, T.G., Sanches-Silva, A., Costa, H.S., 2011. Ascorbic acid content in
Molina, R., Clemente, E., Scapim, M.R.d.S., Vagula, J.M., 2014. Physical evaluation and
exotic fruits: a contribution to produce quality data for food composition databases. Food
hygroscopic behavior of dragon fruit (Hylocereus undatus) lyophilized pulp powder.
Res. Int. 44 (7), 2237–2242. https://doi.org/10.1016/j.foodres.2011.02. 012.
Dry. Technol. 32 (16), 2005–2011. https://doi.org/10.1080/07373937.2014. 929587.
Wang, H., Zhang, S., Chen, G., 2008. Glass transition and state diagram for fresh and freeze-
Mosquera, L.H., Moraga, G., Martínez-Navarrete, N., 2010. Effect of maltodextrin on the stability
dried Chinese gooseberry. J. Food Eng. 84 (2), 307–312. https://doi.org/10.
of freeze-dried borojó (Borojoa patinoi Cuatrec.) powder. J. Food Eng. 97 (1), 72–78.
1016/j.jfoodeng.2007.05.024.
https://doi.org/10.1016/j.jfoodeng.2009.09.017.

103
S. Fongin et al. Journal of Food Engineering 247 (2019) 95–103

Wang, W., Zhou, W., 2012. Characterization of spray-dried soy sauce powders using
maltodextrins as Carrier. J. Food Eng. 109 (3), 399–405. https://doi.org/10.1016/j.
jfoodeng.2011.11.012.

104

Potrebbero piacerti anche