Uploaded by Jiaping Lv

赵某明 搅打稀奶油FC

Food Chemistry 347 (2021) 128997
Contents lists available at ScienceDirect
Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem
Whipping properties and stability of whipping cream: The impact of fatty
acid composition and crystallization properties
Pingli Liu a, Lihua Huang a, Tongxun Liu a, Yongjian Cai a, Di Zeng a, Feibai Zhou a,
Mouming Zhao a, b, Xinlun Deng c, Qiangzhong Zhao a, b, *
a
b
c
School of Food Science and Engineering, South China University of Technology, Guangzhou 510640, People’s Republic of China
Research Institute for Food Nutrition and Human Health, Guangzhou 510640, People’s Republic of China
Guangdong Wenbang Biotechnology Co., Ltd., Guangzhou 511458, People’s Republic of China
A R T I C L E I N F O
A B S T R A C T
Keywords:
Crystallization properties
Fat crystal morphology
Partial coalescence
Whipping properties
Whipping cream
In this study, five fats (hydrogenated palm kernel oil, HPKO-A and HPKO-B; refined vegetable oils, RVO-A and
RVO-B; transesterification oil, TO) were used to prepare whipping creams. HPKO-A and RVO-A which rich in
lauric and myristic acids facilitated the formation of small crystals and dense crystal network, while higher
stearic acid content of HPKO-B formed large spherical crystals. The richness in palmitic acid (RVO-B and TO) and
oleic acid (TO) led to the formation of weak crystal network. Higher partial coalescence was correlated to higher
collision frequency of fat globules and crystal connection, therefore, the overruns, firmness and stability of
creams prepared by HPKO-A and RVO-A were higher than those of HPKO-B and RVO-B. The least stability of
cream prepared by TO was related to the weak crystal networks. In summary, higher lauric and myristic acids
content resulted in dense crystal networks, promoting partial coalescence and improving the cream quality.
1. Introduction
Whipping cream, a typical kind of foamy food, is comprised of a 3dimensional structure throughout the aqueous phase that could sup­
port the air trapped through whipping possess (Warren & Hartel, 2018;
Hotrum, Stuart, van Vliet, Avino & van Aken, 2005). The popular
application of whipping cream is the decoration of cakes and desserts,
which necessitates good whipping properties (Munk & Andersen, 2015).
For these products, partial coalescence fulfills an essential role in the
whipping process (Petrut, Danthine & Blecker, 2016). Under the shear
force, the adjacent fat globules approach each other, and as long as the
fat crystal of one fat globule pierces the interfacial layer between the fat
globules, a crystal link between fat globules is established (Fredrick
et al., 2013). A network of partially coalesced fat globules foamed by fat
crystals that protrude the fat globule membranes ensures an efficient
coverage of the air, resulting in desirable overrun and good stability of
the whipping cream (Wang et al., 2019).
Fat is probably the ingredient that greatly influences the properties
of the oil–water emulsion or whipping cream (Anihouvi, Danthine,
Karamoko & Blecke, 2012). Whipping cream formulated from vegetable
fat is growing in popularity for its lower cost and higher stability
compared to whipping cream prepared with anhydrous milk fat (Munk
& Andersen, 2015). However, the selection of vegetable fat for whipping
cream becomes limited due to the need for partial coalescence. Partial
coalescence takes place only when, within oil-in-water emulsions which
contain a certain percentage of partially crystalline fat, two or more fat
globules collide and connect crystal links. Next, this fat crystal links will
be wetted and reinforced by the remaining liquid fat, forming large
clumps of irregular shapes at final (Phan, Moens, Le, Van der Meeren &
Dewettinck, 2014; Fredrick, Pieter, Walstra & Dewettinck, 2010).
Therefore, the oil-in-water emulsions cannot transform into foam
structures through whipping if the fat is completely crystallized or noncrystallized. The rate and extent of partial coalescence would be varying
due to the ratio of solid versus liquid of fat (Zhao et al., 2013, Sung &
Goff, 2010).
Researches on partial coalescence and properties of oil-in-water
emulsion or whipping cream mainly focused on the effect of interfa­
cial components and especially on the comparison between crystallizing
and non-crystallizing surfactants (Cheng, Dudu, Li & Yan, 2020, Jiang,
Jing, Xiong & Liu, 2019, Fredrick et al., 2013). As for fat crystallization
behaviors, many previous studies focused on the influence of SFC of fat
on partial coalescence of whipping cream. However, the optimal SFC for
* Corresponding author at: School of Food Science and Engineering, South China University of Technology, Guangzhou 510640, People’s Republic of China.
E-mail address: qzzhao@scut.edu.cn (Q. Zhao).
https://doi.org/10.1016/j.foodchem.2020.128997
Received 8 September 2020; Received in revised form 26 November 2020; Accepted 30 December 2020
Available online 7 January 2021
0308-8146/© 2021 Elsevier Ltd. All rights reserved.
P. Liu et al.
Food Chemistry 347 (2021) 128997
partial coalescence from previous studies were diverse due to their
various formulation or processing conditions (Liu et al., 2019; Hana­
zawa, Sakurai, Matsumiya, Mutoh & Matsumura, 2018; Fuller, Con­
sidine, Golding, Matia-merino & MacGibbon, 2015; Mendez-Velasco &
Goff, 2012). This implied that SFC was not the only reason for the
different behavior of fat towards partial coalescence. Fat composition,
crystal melting and recrystallization during thermal cycles and crystal
morphology might also have important effects on partial coalescence
and determine the properties of the final products.
To date, few works have been carried out on fat composition, crys­
tallization properties and its mechanisms involved in partial coalescence
of whipping cream formulated with different fats. Munk & Andersen
(2015) observed the behavior of oil-in-water emulsion containing
different fats under stable shearing. They found that the elastic modulus
and rate of partial coalescence of the destabilized emulsion were both
correlated with the ratio of palmitic acid over oleic acid. Hubbes, Braun
and Foerst (2020) investigated how fat crystallization, the solid fat
content via solids volume fraction (Φ) and the final properties of nougat
cremes were interrelated. They concluded that a tripalmitin (PPP)-rich
fraction of palm fat can improve elastic properties of the underlying fat
crystal network in nougat cremes. Thus, proper structural properties of
the final nougat cremes could be possibly acquired. Moens, Tavernier
and Dewettinck (2018) studied the crystallization characteristics of
different fats at 15 ◦ C, 20 ◦ C, and 25 ◦ C as well as their effects on partial
coalescence of recombined cream. They demonstrated that fat compo­
sition could lead to different microstructure of the fat crystal network
that appeared to influence partial coalescence. From a theoretical point
of view, fat crystallization properties are closely related to property of
final products. However, it is still unclear how fatty acids composition in
combination with crystallization properties and the final properties of
whipping cream are correlated.
This study examined the crystallization properties of fats with
different SFC profiles, and then explored the mechanism of fatty acids
composition and crystallization behavior on whipping cream quality.
This current study will provide more understanding on the selection of
raw materials for whipping cream with desirable properties to meet
consumer demand.
60 ◦ C and held for 2 min, then increased from 60 ◦ C to 160 ◦ C at 20 ◦ C/
min, the temperature was then increased from 160 ◦ C to 240 ◦ C at 5 ◦ C/
min and held for 7 min. Helium flow rate and nitrogen flow rate were
2.25 mL/min and 1.5 mL/min, respectively.
2.2.2. Differential scanning calorimetry (DSC)
The melting and crystallization behaviors of the five bulk fats were
studied by differential scanning calorimetry (DSC) (TA Instruments,
New Castle, Delaware, USA). 4–6 mg of fat samples were sealed in an
aluminium pan and an empty pan was used as reference. The fat samples
were heated to 60 ◦ C (20 ◦ C/min) for 10 min to ensure complete melting,
then cooled by 5 ◦ C/min to − 20 ◦ C, held for 10 min, finally, heated by
5 ◦ C/min to 60 ◦ C. Nitrogen was used to purify the system.
2.2.3. Solid fat content (SFC)
A NM-2 nuclear magnetic resonance (NMR) instrument (New Mai
Electronic Technology Co., Ltd. Shanghai, China) was used to measure
SFC of the five bulk fats. 3–5 mg of fat samples were poured into the
NMR tubes and were tempered in water bath with 60 ◦ C for 30 min to
erase crystal memory. The SFC of fat was determined at temperatures
ranging from 0 to 40 ◦ C (at 5 ◦ C intervals) by equilibrating the NMR
tubes at these temperatures for 30 min before measurement (Long et al.,
2015).
2.2.4. X-ray diffraction (XRD)
XRD diffractions of bulk fats were implemented on an X-ray
diffractometer (Xpert Powder, PANAalytical B. V., Netherland) equip­
ped with CuKα ray (λ = 1.5406 Å) at generating voltage of 40 kV and an
incident current of 40 mA. Samples were scanned from 10 to 35 deg at
2◦ /min for wide-angle XRD (WAXD) analysis. The XRD patterns were
further analyzed using MDI Jade 6.5 and were presented as a function of
d (Å), with d = λ/2sinθ (2θ is the scattering angle and λ is the wave­
length of the incident beam).
2.2.5. Polarized light microscopy (PLM)
To visualize fat crystals of bulk fats, each of the fat sample was
heated to 60 ◦ C to complete melt and erase crystal memory. One droplet
of liquid fat was placed on a preheated glass slide and covered with a
coverslip. Next, the microscopic slide was stored at 4 ◦ C for 12 h before
being visualized. Finally, samples were imaged with a polarized light
microscope (ULH100HG, Olympus Coro Tokyo, Japan) equipped with a
digital camera attached (Olympus Coro Tokyo, Japan).
2. Materials and methods
2.1. Materials
Two types of fat, hydrogenated palm kernel oil (HPKO-A and HPKOB) were purchased from COFCO Donghai grain and oil industry Co., Ltd.
(Jiangshu, China) and Dongguan Jiaji grain and Oil Co., Ltd. (Dongguan,
China), respectively. The refined vegetable oils (RVO-A and RVO-B)
were supported by Ahuskels Oil Co., Ltd. (Shanghai, China). The
transesterification oil (TO) was purchased from Dongguan Jiaji grain
and Oil Co., Ltd. (Dongguan, China). Sodium caseinate (90% protein)
was purchased from New Zealand Milk Products (NZMP, Fonterra Cooperative Group Limited, New Zealand). Oil Red O was purchased
from Shanghai Macklin Biochemical Ltd. (China).
All the chemicals or solvents for Gas Chromatograph-Mass Spec­
trometer (GC–MS) analysis were GC–MS grade and the other reagents
were of analytical grade.
2.3. Preparation of whipping cream
Five types of fat, namely HPKO-A, HPKO-B, RVO-A, RVO-B and TO,
were used to prepare the whipping cream, respectively. Fat (16%, w/w)
was melted at 60 ◦ C, then 0.5% (w/w) sodium caseinate and 1.5% (w/w)
stabilizers were added to obtain fat phase. The mixtures of glucose (18%,
w/w), starch syrup (10%, w/w), and sucrose (2.5%, w/w) were dis­
solved in the water under heating conditions to obtain the aqueous
phase. All percentage presented above was based on the total mass of the
fat phase and the aqueous phase. Then mixed the two phases at 60 ◦ C for
30 min under continuous stirring. The mixtures were homogenized
twice by a high-pressure homogenizer (APV-1000 High Pressure Ho­
mogenizer, Denmark) at a pressure of 50 MPa and then were frozen at
− 18 ◦ C for 24 h to obtain frozen emulsion.
Each of the frozen emulsion was thawed to 0–4 ◦ C and was whipped
at a speed of 160 rpm in an adjustable speed mixer (UK KENWOOD) at
room temperature. The whipped creams that included the five types of
fat were named Cream-HA, Cream-HB, Cream-RA, Cream-RB and
Cream-TO, respectively.
2.2. Characteristic of bulk fat
2.2.1. Fatty acid (FA) composition
FA compositions of bulk fat was determined by gas chromatographymass spectrometer (GC–MS) equipped with an Agilent 19091N-133: 2
capillary column (30 m × 250 μm × 0.25 μm, USA) and gas chroma­
tography triple quadrupole tandem mass spectrophotometry (GCTQ/
MS, Agilent, Santa Clara, CA). The fatty acid methyl esters were pre­
pared by the method of Wang et al. (2010) with modification. The
temperature programs were set as follows: the column was heated to
2
P. Liu et al.
Food Chemistry 347 (2021) 128997
2.4. Measurement of whipping cream
were placed on wooden board at room temperature for two hours. The
degree of roughening of foam structure inside the cream was observed
and photographed to record.
2.4.1. Partial coalescence of fat
The partial coalescence of fat in whipping cream was determined
according to Petrut et al. (2016). Oil Red O solution (0.001 wt% in oil)
was prepared using corn oil with gentle stirring for at least 12 h to allow
complete dispersion. Mixtures consisting of 20 g whipped cream and 10
g Oil Red O solution were centrifuged at 10,000g and 30 ◦ C for 30 min
by a temperature-controlled centrifuge (Model GL-21 M, Xiangyi In­
strument Co. Ltd., Changsha, China). The partial coalesced fat in
whipping cream would dissolve in the dye solution while the emulsified
fat would not. Then used corn oil as the blank and measured the
absorbance of dye solution at 520 nm by a UV–visible spectrophotom­
eter (Perkin-Elmer, Lambda 3, Norwalk, CT). The change in absorbance
of the dye solution indicated the mass of partial coalesced fat, which was
calculated as Eq. (1).
M1 × (α − 1)
φ=
M2 × Φ
2.5. Statistical analysis
All tests were performed at least in triplicate and reported as mean ±
standard deviation. Statistical analysis was performed using version
SPSS 22.0. One-way ANOVA was applied to identify significant differ­
ences between samples and the effects of treatments (p < 0.05).
3. Results and discussion
3.1. Characterization of bulk fat
3.1.1. Fatty acid (FA) composition
The FA composition of each fat sample is presented in Table 1, all fats
studied were composed of various fatty acids. The dominant saturated
fatty acids (SFAs) present in HPKO-A, HPKO-B and RVO-A were lauric
acid (C12:0, 54.42 ± 0.30%, 51.09 ± 0.06% and 47.15 ± 0.11%,
respectively). Moreover, HPKO-A and RVO-A were rich in myristic acid
(C14:0, 24.42 ± 0.30% and 21.30 ± 0.21%, respectively), while HPKO-B
contained abundant stearic acid (C18:0, 17.26 ± 0.18%). RVO-B and TO
were dominated by palmitic acid (C16:0, 40.40 ± 0.28% and 32.30 ±
0.21%, respectively), and Fat E was rich in oleic acid (C18:1, 17.54 ±
0.32%). The variation of FAC would influence the fat crystallization
properties, which will be demonstrated in the following study by the
crystallization and melting profiles, SFC, polymorphic crystalline phase
and morphology of fat crystal.
(1)
where M1 (g) is the weight of added Oil Red O solution, M2 (g) is the
weight of whipping cream, α (dimensionless) is the ratio of absorbance
of Oil Red O solution before and after centrifugation, and Φ (w/w %) is
the mass fraction of partial coalesced fat in whipped cream.
2.4.2. Particle size distribution of fat
Particle size distributions of fat globules in the whipping cream were
determined by light scattering using a Mastersizer 2000 (Malvern In­
struments, Malvern, UK) followed the method of Cai et al. (2021) with
modification. Each of the whipped cream was collected immediately
after reaching the optimum whipping time, then added it drop by drop
to the laser diffraction instrument. Distilled water was used as the
dispersant phase, which diluted the sample at approximately 1:1000 in
the sample chamber. Measurements were performed at ambient tem­
peratures and repeated in triplicate.
3.1.2. Fat crystallization and melting characteristics
The crystallization and melting temperatures of fat determined by
DSC curves are noteworthy for the preparation of whipping cream, as
the formation of fat crystal plays a key role in the formation and stability
of whipping cream (Jiang et al., 2019). The onset temperature (To),
offset temperature (Tf), and temperature range (the difference between
To and Tf) of crystallization and melting are summarized (Table 2).
Moreover, the crystallization and melting profiles of different fats are
presented in supplementary information (Fig. S1, Supplementary
information).
During crystallization process, both HPKO-A and RVO-A had a lower
To (19.85 ± 0.60 ◦ C and 17.25 ± 0.55 ◦ C, respectively). The major
crystallization peak was relatively steeper, indicating that HPKO-A and
2.4.3. Whipping properties
2.4.3.1. The optimal whipping time (top). The frozen emulsions were
whipped after being thawed until a defined observable end point. The
optimum whipping time (top) was defined as the time at which a cream
that was obtained broke away from the wires and the bowl. Whipping
experiments were repeated minimally three times for each whipping
cream, for each whipping experiment the overrun and firmness were
determined.
Table 1
Fatty acid (FA) composition of different fats.
2.4.3.2. Overrun. Overrun is a common measure for the amount of air
introduced in whipping cream and was measured according to the
method of Zhao et al. (2013). The computational formula was described
as the Eq. (2).
M1 − M2
Overrun (%) =
× 100
M2
(2)
Fatty
acid
HPKO-A
(%)
RVO-A (%)
HPKO-B
(%)
RVO-B (%)
TO (%)
C8:0
3.21 ±
0.15b
3.80 ±
0.14b
51.09 ±
0.06b
15.30 ±
0.21c
8.91 ±
0.06d
17.26 ±
0.18a
N.D.
C18:2
N.D.
1.82 ±
0.13c
3.20 ±
0.14c
47.15 ±
0.11c
21.30 ±
0.21b
17.24 ±
0.17c
1.79 ±
0.15d
6.19 ±
0.13c
0.98 ±
0.01a
1.11 ±
0.08d
1.76 ±
0.17d
32.28 ±
0.20d
12.10 ±
0.07d
40.40 ±
0.28a
2.82 ±
0.13c
7.78 ±
0.16b
1.18 ±
0.13a
9.70 ±
0.21a
10.72 ±
0.20a
29.74 ±
0.18e
N.D.
C18:1
1.57 ±
0.30c
2.95 ±
0.04c
54.42 ±
0.30a
24.42 ±
0.30a
8.86 ±
0.10d
7.46 ±
0.33b
N.D.
C10:0
C12:0
where M1 (g) and M2 (g) represent the mass of the cream before whip­
ping and after whipping, respectively.
C14:0
C16:0
2.4.3.3. Firmness. Deformation measurements of whipping creams
were conducted with TA-XT Texture Analyzer (Stable Micro Systems,
UK). Measurements in compression mode and A/BE probe (35 mm
diameter) were selected. Puncture tests were performed at a rate of 1
mm⋅s− 1 over a distance of 25 mm in the sample. The trigger value for the
start of the measurement was set to 0.20 N. The force (N) required to
reach this depth was defined as the firmness of the whipped cream.
C18:0
N.D.
32.30 ±
0.21b
N.D.
17.54 ±
0.32a
N.D.
HPKO-A and HPKO-B represent two types of hydrogenated palm kernel oil; RVOA and RVO-B represent two types of refined vegetable oil; TO represents the
transesterification oil. Values followed by different letters in the same line
represent significant differences at p < 0.05, N.D. means not detected.
2.4.4. Stability
Freshly prepared whipped creams were piled into a hill shape and
3
P. Liu et al.
Food Chemistry 347 (2021) 128997
interpret the melting profile due to polymorphism of liquid fat and solid
fat while the crystallization profile was only affected by the FA
composition of fat.
Table 2
The onset, offset temperatures and temperature ranges derived from the DSC
curves of different fats.
Curve
Sample
To (◦ C)
Tf (◦ C)
Temperature
range
Crystallization
HPKOA
RVO-A
HPKOB
RVO-B
19.85 ± 0.60d
7.05 ± 0.12b
12.80 ± 0.49c
17.25 ± 0.55e
28.65 ± 0.06b
4.05 ± 0.51c
12.05 ±
0.25a
12.65 ±
0.73a
2.15 ± 0.11d
36.95 ±
0.28d
37.42 ±
0.57d
47.62 ±
0.56a
44.32 ±
0.12c
45.22 ±
0.22b
13.20 ± 1.05c
16.60 ± 0.28b
Melting
25.55 ± 0.25c
TO
HPKOA
RVO-B
43.35 ± 0.61a
24.62 ± 0.23b
HPKOB
RVO-B
22.22 ± 0.20c
TO
− 14.57 ±
0.12e
20.62 ± 0.61d
25.92 ± 0.12a
3.1.3. Solid fat content (SFC)
SFC, the solid fat index at a certain temperature in a fat or fat blend,
determines the characteristics and application range of fat (Basso et al.,
2010). As shown in Fig. 1, SFC decreased with the increase of temper­
ature from 0 ◦ C to 40 ◦ C for each of the fat. As temperature increased
from 0 ◦ C to 20 ◦ C, the SFC of HPKO-A, HPKO-B, RVO-A and RVO-B did
not show an obvious decrease but a small drop. Among them, the SFC of
HPKO-A was the highest (93.33% at 20 ◦ C), followed by HPKO-B
(85.30% at 20 ◦ C), RVO-A (81.23% at 20 ◦ C) and RVO-B (81.14% at
20 ◦ C). The lowest SFC was observed for TO, which was only 28.09% at
20 ◦ C. Moreover, the SFC profiles at 0–20 ◦ C of RVO-A and RVO-B were
similar. These results could be explained by the higher proportion of
SFAs in HPKO-A, HPKO-B, RVO-A and RVO-B mentioned above. How­
ever, TO contained less SFAs and higher amount of lower melting
fractions such as capric, caprylic and oleic acids. As a result, TO had the
lowest SFC in low temperatures and, also had the least content of fat
crystals.
Between 20 ◦ C and 25 ◦ C, the SFC of HPKO-A showed a small
decrease (3.73%) indicating that HPKO-A had lower decreasing rate of
SFC at 20–25 ◦ C. Then was the RVO-B (7.73%), TO (11.39%), HPKO-B
(12.73%) and RVO-B (13.16%). As temperature increased from 25 ◦ C
to 35 ◦ C, the higher reduction in SFC was observed for HPKO-A
(84.99%), RVO-A (61.65%) and HPKO-B (53.08%). RVO-B and TO
had lower decreasing rate of SFC at 25–35 ◦ C, the SFC of RVO-B
decreased from 75.14% at 25 ◦ C to 38.73% at 35 ◦ C while that of TO
reduced by only 15.23%. This could be related to the fact that HPKO-A,
HPKO-B and RVO-A were dominated by lauric acid, while RVO-B and
TO were rich in palmitic acid, the later one had a higher melting point.
Large part of low melting point components will melt quicker in higher
temperature region (Hubbes et al., 2019; Viriato, Queirós, Neves,
Ribeiro & Gigante, 2019).
12.90 ± 0.49c
41.20 ± 0.11a
12.33 ± 0.45e
16.80 ± 0.20d
25.40 ± 1.21b
18.40 ± 0.29c
59.75 ± 0.26a
HPKO-A and HPKO-B represent two types of hydrogenated palm kernel oil; RVOA and RVO-B represent two types of refined vegetable oil; TO represents the
transesterification oil. Abbreviations are: To, onset temperature of crystalliza­
tion and melt, respectively, Tf, offset temperature of crystallization and melt,
respectively. Values followed by different letters in the same column represent
significant differences at p < 0.05.
RVO-A had a faster nucleation period and a shorter time to achieve
equilibrium crystals, which was in accordance with Liu et al. (2019).
This result could be related to higher content of lauric and myristic acid
of HPKO-A and RVO-A, whereby a lower temperature was needed to
achieve the required driving crystallization force and induce more
nucleation sites (Danthine, 2012, Moens, 2018). The To of HPKO-B and
RVO-B appeared to be shifted to higher temperatures (28.65 ± 0.06 ◦ C
and 25.55 ± 0.25 ◦ C, respectively) and both exhibited two separate
exothermic peaks. This could be related to HPKO-B was rich in stearic
and RVO-B was dominated by palmitic acid, those high melting fractions
would firstly crystallize from the melt (Hubbes, Danzl & Foerst, 2019).
Finally, TO exhibited two overlapping exothermic peaks with low in­
tensity, indicating a low compactness of crystal structure. The crystal­
lization peak at lower temperature was related with the low melting
fractions such as capric, caprylic and oleic acids. The onset crystalliza­
tion peak had the highest To of crystallization (43.35 ± 0.61 ◦ C) due to
the highest concentration of palmitic acid of TO. These results indicated
that longer carbon chain saturated fatty acids led to a higher crystalli­
zation temperature, which in turn slowed down the crystallization rate.
This was in line with the research on the isothermal crystallization of
palm kernel stearin (Liu et al., 2019).
During melting process, the To of HPKO-A and RVO-A was 24.62 ±
0.23 ◦ C and 20.62 ± 0.61 ◦ C, respectively, and both exhibited only one
endotherm peak with a narrow melting range within 20–35 ◦ C. The Tf of
HPKO-B and RVO-B shifted to higher temperature and their temperature
ranges were higher than HPKO-A and RVO-B. For HPKO-B, the melting
curve showed a distinct separation in two peaks (27.92 ◦ C and 34.72 ◦ C,
respectively) and a small fusion peak at higher temperature (44.22 ◦ C).
For RVO-B, a major endotherm peak (40.12 ◦ C) with a small shoulder
peak at lower temperature (20.22 ◦ C) was observed. Moreover, the
lowest melting point (-14.57 ± 0.12 ◦ C) with a wide melting tempera­
ture range was found for TO. The differences present in the melting
behavior during heating could be related with the dissimilarities in
crystallization behavior during cooling. The mixed low and middlemelting crystals melted firstly upon heating, and the higher melting
crystals melted at higher temperature after recrystallized into a more
stable form (Bootello et al., 2013). However, as indicated by Bercikova,
Simkova, Huadecova, Kyselka, Filip, and Hradkova (2020), it is hard to
3.1.4. Polymorphic crystalline phase
The polymorphic crystalline phases are characterized by the shortspacings of the crystal lattices (typically between 3 Å and 6 Å) as
observed in XRD patterns and the results are presented in Fig. 2I. HPKOA and RVO-A showed WAXD intensities at 3.80, 4.05, 4.20 and 4.35 Å
while the WAXD intensities of HPKO-B and RVO-B located at 3.80 and
4.20 Å, respectively. The high intensity at 3.80 Å along with the low
intensity at 4.35 Å indicated β’1 crystalline phase whereas a β’2 crys­
talline phase was represented by the peaks with intensities at 4.20 and
3.80 Å (Truong, Morgan, Bansal, Palmer & Bhandari, 2015). Therefore,
Fig. 1. The solid fat content (SFC) versus temperature for different fats. (HPKO:
hydrogenated palm kernel oil, RVO: refined vegetable oils, TO: trans­
esterification oil.)
4
P. Liu et al.
Food Chemistry 347 (2021) 128997
Fig. 2. XRD spectra for short spacings and PLM images of different fats taken after an isothermal period of 24 h at 4 ◦ C, respectively. (Values on the top of curves in
λ
, 2θ is the scattering angle and λ is the wavelength of the incident beam, the scale bar in Fig. 2II is 50 μm; HPKO: hydrogenated palm
Fig. 2I represent d, d = 2sinθ
kernel oil, RVO: refined vegetable oils, TO: transesterification oil.)
HPKO-A and RVO-A contained β’1 and β’2 crystalline phases, while
HPKO-A and RVO-A formed β’2 crystalline phase. A distinct separation
between the β’1- and β’2-peaks in HPKO-A and RVO-A indicated purer
crystals and better organized crystal structure and the mixed β’-crystals
can be more stable than β-crystals (Moens, 2018). This could also be
proved by a distinct steep crystallization peak of HPKO-A and RVO-A
(Fig. S1). The short spacing patterns of TO showed WAXD peaks at
3.80, 4.20 and 4.60 Å, characteristics of the β’2 and β crystalline phase.
This could be owing to the fact that TO was rich in oleic acid which tends
to form crystals in the β-form (Bugeat et al., 2011).
distributed, leading to incomplete crystal networks. These results were
mainly due to HPKO-B and RVO-B were rich in stearic and palmitic acid
respectively and both had separated crystallization regions. The long
chain molecules firstly formed high melting crystals and then cocrystallized with low melting crystals upon cooling, thereby forming
mixed and large crystals. However, HPKO-B had a higher lauric acid
content than RVO-B, so the fat crystals of HPKO-B were more connected
than that of RVO-B. Similarly, the lamellar and sparsely-distributed
crystals were also visualized in TO, where the crystals were smaller
and harder to discernible as compared to that of RVO-B. This phenom­
enon could be resulted from the lower SFC and the overlapping crys­
tallization peaks of TO, leading to fewer, sparsely-distributed crystal
clusters . The crystal clusters would form a loose and weak fat crystal
network. These results in morphology conformed to the polymorphic
crystalline phase obtained in the XRD analysis.
3.1.5. Morphology of the fat crystal
The crystalline morphologies of different fats are shown in Fig. 2II.
HPKO-A contained smaller crystals and the hollower spaces in the fat
crystal network were occupied by small needle-like crystals, forming a
densely-packed microstructure. Needle-like crystals were also observed
in RVO-A, but they were larger in size than that of HPKO-A. This result
could be related to the higher concentration of lauric and myristic acid
and a sharp crystallization region in lower temperatures of HPKO-A and
RVO-A. This accelerated the period of nucleation, inducing more small
crystals. Large spherical shape crystals were more dominant in HPKO-B
and RVO-B appeared large lamellar crystal clusters which were sparsely
3.2. Characteristics of whipping cream
3.2.1. Partial coalescence of whipping cream
Fig. 3I indicates the extent of partial coalescence of whipping cream
prepared with different fats. As can be seen, the extent of partial coa­
lescence of Cream-HA was the greatest (44.02 ± 0.64%) at the optimum
5
P. Liu et al.
Food Chemistry 347 (2021) 128997
Fig. 3. Partial coalescence and particle size distribution of fat of whipping cream with different fats. (I ~ II represent partial coalescence and particle size distribution
of fat of whipping cream, respectively. Different superscript letters on the top of columns in I represent significant differences at p < 0.05.)
whipping time, followed by Cream-RA (40.25 ± 1.95%), Cream-HB
(35.08 ± 0.33%), Cream-RB (29.44 ± 0.94%) and Cream-TO (15.12
± 0.79%).
The particle size distributions of fat globules in whipping cream are
showed in Fig. 3II. The charts of particle distribution of all whipping
cream simples were bimodal. The first particle size distribution parts of
all samples were between 0.02 and 1 µm. The second particle size dis­
tribution parts of all simples were situated at larger diameters, which
demonstrated fat aggregates formed by partial coalescence during
whipping (Petrut et al., 2016). The dynamic light scattering results
recorded the average particle size of whipping cream as follows: CreamHA (29.698 ± 1.745 μm) > Cream-RA (26.997 ± 0.716 μm) > Cream-HB
(18.683 ± 0.543 μm) > Cream-RB (15.716 ± 0.905 μm) > Cream-TO
(8.475 ± 0.513 μm), which was consistent with the results in Fig. 3I.
The higher extent of partial coalescence for Cream-HA and CreamRA could be attributed to the following two reasons. On the one hand,
HPKO-A and RVO-A had higher SFC at 0–20 ◦ C (Fig. 1), providing
adequate fat crystals (Fredrick et al., 2010; Boode, Walstra & GrootMostert, 1993). The number of protruding crystals was raised when
the content of fat crystal increased, which could enhance the suscepti­
bility to partial coalescence of fat globules. On the other hand, HPKO-A
and RVO-A formed a dense fat crystal network (Fig. 2II). Then as
whipping proceeds, the dense fat crystal network could prevent the
protruding crystals being pushed back into the globules when fat glob­
ules collide and the completely coalescence of fat globules would be
limited. As a consequence, the extent of partial coalescence could be
enhanced greatly.
On the contrary, although the SFC of HPKO-B and RVO-B maintained
at a higher value at 0–20 ◦ C, Cream-HB and Cream-RB exhibited lower
extent of partial coalescence than Cream-HA and Cream-RA. This could
be explained by the fact that HPKO-B and RVO-B appeared larger fat
crystal clusters compared with HPKO-A and RVO-A, leading to a weaker
6
P. Liu et al.
Food Chemistry 347 (2021) 128997
fat crystal network. Though the larger crystals had longer protrusion
distance, the weaker fat crystal network could not provide appropriate
mechanical strength to hinder the complete merging of fat globules,
which would restrict partial coalescence. Furthermore, the extent of
partial coalescence of Cream-HB was higher than that of Cream-RB due
to the fact that HPKO-B appeared more intensive crystal networks than
RVO-B. Finally, Cream-TO showed the lowest extent of partial coales­
cence. This could be explained by the fact that TO had neither adequate
fat crystals due to its low SFC at 0–20 ◦ C nor formation of a dense fat
crystal network due to its slow crystallization rates.
Table 3
Whipping properties of whipping cream prepared by different fats.
Sample
Cream-HA
Cream-RA
Cream-HB
Cream-RB
Cream-TO
top (s)
208.0
230.0
305.0
338.0
305.0
Overrun (%)
d
± 9.9
± 5.8c
± 13.7ab
± 12.0a
± 2.9b
328.9 ±
295.8 ±
307.5 ±
285.3 ±
284.2 ±
a
6.2
5.3b
6.4a
3.0b
2.6b
Firmness (N)
1.96
1.78
1.68
1.20
0.53
± 0.03a
± 0.03d
± 0.01b
± 0.01c
± 0.01e
Cream-HA, Cream-HB, Cream-RA, Cream-RB and Cream-TO represent whipping
cream prepared with HPKO-A, HPKO-B, RVO-A, RVO-B and TO, respectively.
The top represents the optimal whipping time. Values followed by different let­
ters in the same column represent significant differences at p < 0.05.
3.2.2. Whipping properties of whipping cream
The comparisons of whipping properties between each of the whip­
ping cream are listed in Table 3. Cream-HA had the highest overrun
(328.9 ± 6.2%) in a shortest top (208.0 ± 9.9 s), followed by Cream-RA
(295.8 ± 5.3% in 230.0 ± 5.8 s). However, Cream-RB and Cream-TO
both obtained lower overrun (285.3 ± 3.0% and 284.2 ± 2.6%,
respectively) in a longer top (338.0 ± 12.0 s and 305.0 ± 2.9 s, respec­
tively). A shorter top demonstrates an increased rate of partial coales­
cence at the air interface in cream (Hotrum et al. 2005). Those results
indicated that partial coalescence at the air interface of Cream-HA and
Cream-RA was greater than that of Cream-RB and Cream-TO. The
overrun seemed to be lower when a longer top was need to obtain an
acceptable whipping cream.
During whipping, the small needle-like crystals are then no longer
kept in the crystal network, instead, they moved to the interface and
pierce through the membrane as temperature increase (Fredrick et al.,
2010). Adequate fat crystals combined with denser fat crystal networks
could increase collision frequency and facilitate faster creation of effi­
cient persistent crystal connection when crystals penetrate interfacial
membrane between the fat globule. Besides, mixed β’ crystalline phase
has fine crystal grain and can wrap more air than β crystalline phase
during whipping (Tzompa-Sosa, Ramel, Van Valenberg & Van Aken,
2016). Fat globules would spread surrounding air bubbles and create a
stronger fat crystal network by higher extent of partial coalescence to
stabilize air bubbles. Consequently, the whipping process had been
promoted and the overrun had been increased. Conversely, though
larger crystals could penetrate farther and create larger holes in the
interfacial membrane, fat globules would completely merge without
denser fat crystal networks to counteract. This could largely reduce the
collision frequency of fat globules and form incomplete fat crystal
network, allowing less air inclusion. Moreover, larger holes in the
interfacial membrane could result in more liquid fat being extruded
under a mechanical force to spread on the air bubble surface. Much
liquid fat that spread on the air bubble surface was detrimental for foam
stability, resulting in a lower overrun. Similar result was reported by
Moens (2018) for recombined dairy cream.
sufficient strength to whipping cream, leading to higher firmness. The
stronger network of partially coalesced fat globules was efficient for
immobilization of air bubbles. Moreover, the differences of firmness and
stability could also be explained by the melting behavior of crystals,
which mainly manifested at the SFC versus temperature and the To and
Tf in melting profiles. HPKO-A had a lower decreasing rate in SFC at
20–25 ◦ C than HPKO-B and RVO-A, which could promote efficient foam
stabilization of whipping cream at room temperature. Therefore, the
foam structure of Cream-HA was smoother and denser than that of
Cream-HB and Cream-RA. The lowest firmness and storage stability of
Cream-RB could be related to the lower melting point of β’2-crystals
compared to β’1-crystals. For Cream-TO, the crystal network would be
merge upon storage temperature due to the lower To and wider tem­
perature range of melt, leading to lower firmness and stability.
4. Conclusions
This study demonstrated the crystallization properties of fat with
different fatty acid composition and clarified the effect of crystallization
behavior on the quality of whipping cream. The chemical composition
was a dominant factor in determining the crystallization properties of
fat, which contributed greatly to partial coalescence and whipping
properties of whipping cream. HPKO-A and RVO-A dominated by lauric
acid and myristic acid formed a sharp crystallization region and dense
fat crystal networks. HPKO-B was rich in stearic acid formed lager
spherical crystals. However, RVO-B and TO dominated by palmitic acid
formed looser fat crystal networks, especially in the presence of oleic
acids which accounted for the lower solid fat of TO. As whipping pro­
ceeds, adequate fat crystals combined with dense fat crystal networks
could increase collision frequency of fat globules and facilitate faster
creation of efficient persistent crystal connection. This promoted partial
coalescence at the air interface and formation of a stronger network of
partially coalesced fat globules. Therefore, whipping cream based on
HPKO-A and RVO-A obtained shorter whipping time, higher overrun
and firmness. On the contrary, though larger crystals could penetrate
farther, fat globules would completely merge without the protection of
denser fat crystal networks. This could greatly reduce the collision fre­
quency of fat globules and form uncomplete networks of partially coa­
lesced fat globules. Thus, the whipping cream based on RVO-B and TO
had lower overrun in longer whipping time as well as lower firmness.
This study indicated the formulation of the fat could be a powerful
tool to control and improve quality of foamy food and provide a theo­
retical basis for the preparation and selection of special fats for whipping
cream.
3.2.3. Firmness and stability
As shown in Table 3, firmness showed significant differences (p <
0.05) for each of the whipping creams. The firmness of all creams
decreased in the following order: Crean-HA (1.96 ± 0.03 N) > Cream-RA
(1.78 ± 0.03 N) > Cream-HB (1.68 ± 0.01 N) > Cream-RB (1.20 ± 0.01
N) > Cream-TO (0.53 ± 0.01 N), which was in line with the results in the
extent of partial coalescence. In addition, the shape retention and foam
structure of whipping cream after two hours were photographed
(Fig. S2, Supplementary information). It was observed that Cream-HA,
Cream-HB and Cream-RA kept firm shape and showed no expulsion of
aqueous phase for over two hours, which indicated that Cream-HA,
Cream-HB and Cream-RA had higher storage stability. However,
Cream-RB and Cream-TO showed lower storage stability as evidenced
not only by the occurrence of softening of foam structure but also by the
separation of fat phase from aqueous phase.
Firmness is related with the extent of partial coalescence of whipping
cream (Fredrick et al., 2013). Higher extent of partial coalescence could
form a stronger network of partially coalesced fat globules to impose
Declaration of Competing Interest
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
7
P. Liu et al.
Food Chemistry 347 (2021) 128997
Acknowledgements
Hotrum, N. E., Stuart, M. A. C., Vliet, T.van., Avino, S. F., & van Aken, G. A. (2005).
Elucidating the relationship between the spreading coefficient, surface-mediated
partial coalescence and the whipping time of artificial cream. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 260(1-3), 71–78.
Hubbes, S.-S., Braun, A., & Foerst, P. (2020). Crystallization kinetics and mechanical
properties of nougat creme model fats. Food Biophysics, 15(1), 1–15.
Hubbes, S.-S., Danzl, W., & Foerst, P. (2019). Crystallization kinetics of palm oil of
different geographic origins and blends thereof by the application of the Avrami
model. LWT - Food Science and Technology, 93, 189–196.
Jiang, J., Jing, W., Xiong, Y. L., & Liu, Y. (2019). Interfacial competitive adsorption of
different amphipathicity emulsifiers and milk protein affect fat crystallization,
physical properties, and morphology of frozen aerated emulsion. Food Hydrocolloids,
87, 670–678.
Liu, N., Cui, J., Li, G., Li, D., Chang, D., Li, C., & Chen, X. (2019). The application of high
purity diacylglycerol oil in whipped cream: Effect on the emulsion properties and
whipping characteristics. CyTA-Journal of Food, 17(1), 60–68.
Long, Z., Zhao, M., Liu, N., Liu, D., Sun-Waterhouse, D., & Zhao, Q. (2015).
Physicochemical properties of peanut oil-based diacylglycerol and their derived oilin-water emulsions stabilized by sodium caseinate. Food Chemistry, 184, 105–113.
Méndez-Velasco, C., & Goff, H. D. (2012). Fat structure in ice cream: A study on the types
of fat interactions. Food Hydrocolloids, 29(1), 152–159.
Moens, K. (2018). Fat crystal networks in relation to partial coalescence (PhD thesis).
Belgium: Ghent University.
Moens, K., Tavernier, I., & Dewettinck, K. (2018). Crystallization behavior of emulsified
fats influences shear-induced partial coalescence. Food Research International, 113,
362–370.
Munk, M. B., & Andersen, M. L. (2015). Partial coalescence in emulsions: The impact of
solid fat content and fatty acid composition. European Journal of Lipid Science and
Technology, 117(10), 1627–1635.
Petrut, R. F., Danthine, S., & Blecker, C. (2016). Assessment of partial coalescence in
whippable oil-in-water food emulsions. Advances in Colloid and Interface Science, 229,
25–33.
Phan, T. T. Q., Moens, K., Le, T. T., Van der Meeren, P., & Dewettinck, K. (2014).
Potential of milk fat globule membrane enriched materials to improve the whipping
properties of recombined cream. International Dairy Journal, 39(1), 16–23.
Sung, K. K., & Goff, H. D. (2010). Effect of solid fat content on structure in ice creams
containing palm kernel oil and high-oleic sunflower oil. Journal of Food Science, 75
(3), 274–279.
Truong, T., Morgan, G. P., Bansal, N., Palmer, M., & Bhandari, B. (2015). Crystal
structures and morphologies of fractionated milk fat in nanoemulsions. Food
Chemistry, 171, 157–167.
Tzompa-Sosa, D. A., Ramel, P. R., Van Valenberg, H. J. F., & Van Aken, G. A. (2016).
Formation of beta polymorphs in milk fats with large differences in triacylglycerol
profiles. Journal of Agricultural and Food Chemistry, 64(20), 4152–4157.
Viriato, R. Lázaro. S., Queirós, M.de. S., Neves, M. I. L., Ribeiro, A. P. B., &
Gigante, M. Lúcia. (2019). Improvement in the functionality of spreads based on
milk fat by the addition of low melting triacylglycerols. Food Research International,
120, 432–440.
Wang, Y., Yuan, D., Li, Y., Li, M., Wang, Y., Li, Y., & Zhang, L. (2019). Thermodynamic
and whipping properties of milk fat in whipped cream: A study based on DSC and
TD-NMR. International Dairy Journal, 97, 149–157.
Wang, Y., Zhao, M., Song, K., Wang, L., Tang, S., & Riley, W. W. (2010). Partial
hydrolysis of soybean oil by phospholipase A (1) (Lecitase Ultra). Food Chemistry,
121(4), 1066–1072.
Warren, M. M., & Hartel, R. W. (2018). Effects of emulsifier, overrun and dasher speed on
ice cream microstructure and melting properties. Journal of Food Science, 83(3),
639–647.
Zhao, Q., Kuang, W., Long, Z., Fang, M., Liu, D., Yang, B., & Zhao, M. (2013). Effect of
sorbitan monostearate on the physical characteristics and whipping properties of
whipped cream. Food Chemistry, 141(3), 1834–1840.
The authors are grateful for the financial support from the Thir­
teenth-five National Key R&D Program of China (2016YFD0400803),
the Guangdong Key R&D Program (2020B020226010), the National
Natural Science Foundation of China (No. 31571883, 31701539), and
the 111 Project (B17018).
Appendix A. Supplementary data
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.foodchem.2020.128997.
References
Anihouvi, P. P., Danthine, S., Karamoko, G., & Blecker, C. (2012). Vegetal creams: An
alternative to dairy creams. A review. Biotechnologie Agronomie Societe Et
Environnement, 16(3), 344–359.
Basso, R. C., Ribeiro, A. P. B., Masuchi, M. H., Gioielli, L. A., Gonçalves, L. A. G.,
Santos, A. O. D., … Grimaldi, R. (2010). Tripalmitin and monoacylglycerols as
modifiers in the crystallisation of palm oil. Food Chemistry, 122(4), 1185–1192.
Bercikova, M., Simkova, A., Hudecová, K., Kyselka, J., Filip, V., & Hrádková, I. (2020).
Physical properties of structured fats and fat blends based on the long chain fatty
acids. Food Biophysics, 15(1), 143–152.
Boode, K., Walstra, P., & de Groot-Mostert, A. E. A. (1993). Partial coalescence in oil-inwater emulsions 2. Influence of the properties of the fat. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 81, 139–151.
Bootello, M. A., Hartel, R. W., Levin, M., Martínez-Blanes, J. M., Real, C., Garces, R., …
Salas, J. J. (2013). Studies of isothermal crystallization kinetics of sunflower hard
stearin-based confectionery fats. Food Chemistry, 139(1–4), 184–195.
Bugeat, S., Briard-Bion, V., Perez, J., Pradel, P., Martin, B., Lesieur, S., … Lopez, C.
(2011). Enrichment in unsaturated fatty acids and emulsion droplet size affect the
crystallization behaviour of milk triacylglycerols upon storage at 4 degrees C. Food
Research International, 44(5), 1314–1330.
Cai, Y. J., Huang, L. H., Chen, B. F., Zhao, X. J., Zhao, M. M., Zhao, Q. Z., … Paul. (2021).
Effect of alkaline pH on the physicochemical properties of insolublesoybean fiber
(ISF), formation and stability of ISF-emulsions. Food Hydrocolloids, 111, Article
106188.
Cheng, J., Dudu, O. E., Li, X., & Yan, T. (2020). Effect of emulsifier-fat interactions and
interfacial competitive adsorption of emulsifiers with proteins on fat crystallization
and stability of whipped-frozen emulsions. Food Hydrocolloids, 101, 105491. https://
doi.org/10.1016/j.foodhyd.2019.105491.
Danthine, S. (2012). Physicochemical and structural properties of compound dairy fat
blends. Food Research International, 48(1), 187–195.
Fredrick, E., Heyman, B., Moens, K., Fischer, S., Verwijlen, T., Moldenaers, P., …
Dewettinck, K. (2013). Monoacylglycerols in dairy recombined cream: II. The effect
on partial coalescence and whipping properties. Food Research International Journal,
51(2), 936–945.
Fredrick, E., Walstra, P., & Dewettinck, K. (2010). Factors governing partial coalescence
in oil-in-water emulsions. Advances in Colloid and Interface Science, 153(1-2), 30–42.
Fuller, G. T., Considine, T., Golding, M., Matia-Merino, L., & MacGibbon, A. (2015).
Aggregation behavior of partially crystalline oil-in-water emulsions: Part II-Effect of
solid fat content and interfacial film composition on quiescent and shear stability.
Food Hydrocolloids, 51, 23–32.
Hanazawa, T., Sakurai, Y., Matsumiya, K., Mutoh, T.-aki., & Matsumura, Y. (2018).
Effects of solid fat content in fat particles on their adsorption at the air-water
interface. Food Hydrocolloids, 83, 317–325.
8