Вы находитесь на странице: 1из 10

J Am Oil Chem Soc

DOI 10.1007/s11746-017-2973-3

ORIGINAL PAPER

Crystallization Behavior and Kinetics of Chocolate‑Lauric Fat


Blends and Model Systems
Margaret Geary1 · Richard Hartel1 

Received: 8 August 2016 / Revised: 26 February 2017 / Accepted: 11 March 2017


© AOCS 2017

Abstract  Binary mixtures of cocoa butter and lauric fats kernel oil, to produce a wide array of confectionery prod-
have widespread use in chocolates and confections, yet ucts, including chocolate truffle centers, chocolate melta-
incompatibilities between these fats can present formula- ways, and fillings for sandwich-type biscuits. Such products
tion and processing constraints. This study examined the derive their silky smooth textures and melt-in-your-mouth
phase behavior and crystallization kinetics of cocoa butter- sensations from these fat blends. However, cocoa butter
lauric fat model systems and chocolate-lauric fat blends. and lauric fat exhibit different melting and crystallization
Solid fat content (SFC) profiles and isosolid diagrams behaviors, which pose limitations when developing confec-
confirmed eutectic and diluent interactions, indicating a tions made from these mixtures. Principally, the presence
softening of cocoa butter by lauric fat addition. Crystalliza- of lauric fat in a cocoa butter system can result in bloom
tion kinetics of model systems adhered to an exponential or excessive softening (eutectic) in the finished product [1].
growth model. High lauric fat levels delayed crystal growth The phase behavior of blends of cocoa butter and lau-
and reduced equilibrium SFC of cocoa butter. Coconut ric fat can be studied to understand the compatibility and
and palm kernel oils altered the solidification mechanisms equilibrium conditions of a binary system. Eutectic effects
of cocoa butter to a greater extent than fractionated palm can be visualized through solid fat content (SFC) profiles
kernel oil. Chocolate systems displayed multi-step crys- or isosolid diagrams, which depict the relationship between
tal growth that contrasted with the exponential growth fat composition and temperature with lines of constant
observed in the model systems. At high lauric fat levels SFC and illustrate fat compatibility. Several authors have
(30%), crystallization onset was significantly lengthened. investigated the phase behavior of cocoa butter and lauric
Blends with high lauric fat contents showed low G′max and fat using isosolid diagrams [2–4]. In these studies, eutectic
did not achieve final equilibrium after 60 min of cooling, formations were observed at intermediate levels of lauric
indicating incomplete crystallization. fat. Eutectic interactions are said to be due to differences
in thermal properties, triacylglycerol profiles, and polymor-
Keywords  Chocolate · Lauric fat · Crystallization phic forms of fats [5]. However, these studies were specific
to cocoa butter-palm kernel oil blends and did not com-
pare the phase behaviors of different types of lauric fat in
Introduction cocoa butter systems, specifically coconut oil or a lauric fat
fraction.
In the confectionery industry, cocoa butter can be blended The crystallization kinetics of cocoa butter has been
with lauric vegetable fats, such as coconut oil and palm a subject of intense focus [6, 7], yet there is a paucity of
research on the effect of lauric fat on cocoa butter crystal-
lization. Ribeiro et al. [8] studied the kinetics of isother-
* Richard Hartel mal crystallization of cocoa butter-palm kernel oil blends
rwhartel@wisc.edu
by analyzing the increase in SFC over cooling time. The
1
Department of Food Science, University of Wisconsin- addition of low levels of lauric fat effectively delayed the
Madison, 1605 Linden Dr, Madison, WI 53706, USA onset of cocoa butter crystallization and decreased the

13
J Am Oil Chem Soc

final equilibrium SFC. Due to dissimilarities in the tria- dry-conched chocolate mass and CB. Fats were melted at
cylglycerol chain lengths and polymorphism, co-crystalli- 65 °C for 1 h and mixed thoroughly. To prepare model sys-
zation of lauric fat and cocoa butter is markedly weak [9]. tems, lauric fats were blended with CB at 10% composi-
Long-chain triacylglycerols (16–18 carbons) comprise the tion intervals from 0 to 100% wt/wt for SFC profiles and
majority of cocoa butter whereas short- and medium-chain isosolid diagrams, and from 0 to 30% wt/wt levels for crys-
triacylglycerols (12–14 carbons) dominate lauric fat. Fur- tallization curves. Chocolate systems were made at 0–30%
thermore, the formation of a β polymorph in cocoa butter wt/wt levels to the same final fat contents of 40.8% to
and β′ polymorph in lauric fat has been noted in X-ray dif- account for variation in solids. Lauric fat, chocolate mass
fraction patterns by Quast et al. [4]. While these differences (32.5% fat content) and CB comprised 10 and 20% blends,
have recognized inhibitory effects on the crystallization and 30% blends contained lauric fat and chocolate mass.
of cocoa butter, no studies have sought to quantify these
effects by the application of empirical growth models. SFC Profiles and Isosolid Diagrams of Model Systems
The rheological properties of chocolate can be deter-
mined using small amplitude oscillatory rheology, which A Bruker nuclear magnetic resonance (NMR) Minispec Mq
does not disrupt the formation of a fat network during crys- instrument (Bruker, The Woodlands, TX, USA) equipped
tallization and measures the storage modulus (G′) and loss with Minispec software was used to evaluate equilibrium
modulus (G″) to indicate the relationship between solid- SFC. Calibration was performed using oil with 0% sol-
like and liquid-like properties at a given time [10]. Specifi- ids and paraffin-based standards of 30 and 70% solids.
cally, G′ values can be used to compare the effects of cool- Melted fats (2 mL) were filled and capped into NMR tubes
ing conditions on the solidification kinetics of fat systems. (10-mm diameter) and tempered according to the IUPAC
These rheological cooling curves depict an initial lag time, method 2.150 (IUPAC 1987). Fats were exposed to 80 °C
an increase in G′, and a final plateau region. When melted for 20 min, 0 °C for 90 min, 26 °C for 40 h, and 0 °C for
cocoa butter is cooled to 15 °C, G′ increases over time as 90 min and stabilized for 60 min at each test temperature
growing crystals interact and form a matrix, which occurs (10, 20, 25, 30, 35, and 40 °C) prior to NMR measurement.
rapidly initially but slows toward a state of thermodynamic All samples were prepared in triplicate and average SFC
equilibrium. A high equilibrium G′ corresponds to a final was determined from the three measurements. SFC profiles
crystal network with high hardness. Small amplitude oscil- were constructed for decrease in SFC over temperature.
latory rheology has been used to evaluate the effect of an Isosolid diagrams were prepared using Matlab R2015A
additive on the solidification kinetics of chocolate [11, 12]. software and a customized quadratic interpolation code
However, these studies documented the effects of milk after a program written by Timms [13]. Isosolid lines were
fat or lecithin on crystallization, and virtually no research drawn at constant SFC values of 5, 10, 20, 30, 40, 60, and
exists for chocolate-lauric fat systems. 80%.
In order to optimize the use of lauric fat in chocolate
products in industrial settings, it is important to understand Crystallization Curves of Model Systems
their effects on the crystallization behavior of cocoa but-
ter. A knowledge of crystallization properties is useful to Fats were melted and stirred at 45 °C. A cyclothermic tem-
maintain efficiency in chocolate manufacture and evade pering method described by Kleinert [14] was applied to
conditions that might cause inconsistencies or defects in a samples with exact times and temperatures adjusted by trial
product. The objective of this work was to study the crys- and error. After melting at 50 °C, samples were cooled to
tallization behavior and kinetics of chocolate-lauric fat 24 °C (0 and 10% lauric fat) or 23 °C (20 and 30% lau-
blends and cocoa butter-lauric fat model systems. ric fat) and maintained until viscosity reached a maximum.
The temperature was raised to 27 °C until texture appeared
uniformly smooth, lowered again to 26 °C until viscos-
Experimental Procedures ity reached a maximum, and reheated to 31 °C. To test for
proper tempering, samples were spooned into a temperme-
Materials ter cup set in an ice bath, and a temperature–time cooling
curve was measured at 10-s intervals. Well-tempered sam-
Lauric fats coconut oil (CO; Columbus Vegetable Oils, Des ples displayed cooling curves with a sharp temperature
Plaines, IL, USA), palm kernel oil (PKO; AarhusKarlshamn decrease, a plateau region, and a final temperature drop.
USA, Port Newark, NJ, USA), and fractionated palm kernel Once a well-tempered profile was obtained in triplicate,
oil (FPKO; AarhusKarlshamn USA) were used in model the tempering regimen was adopted as the pre-treatment
systems with cocoa butter (CB; Archer Daniels Midland for SFC determination. As a second test, a thin layer of
Cocoa, Milwaukee, WI, USA) and chocolate systems with tempered sample was spread onto a strip of wax paper and

13
J Am Oil Chem Soc

maintained at room temperature. A well-tempered sample using Matlab R2015A software and slopes were calculated
hardened after 1–2 min, easily peeled away from the wax using an Excel spreadsheet. Statistical analyses were based
paper, and had a glossy surface. Tempered fats were filled on analysis of variance (ANOVA) and Tukey’s test to deter-
into NMR tubes, equilibrated to 31 °C, then quickly cooled mine significant differences of the means (p < 0.05).
to the crystallization temperature. The increase in SFC
over crystallization time at 10, 15, and 20 °C was moni-
tored at 2-min intervals for 60 min or until SFC reached a Results and Discussion
maximum. Average cooling rates were recorded and ranged
from 0.05 to 0.1 °C/min. All samples were prepared in trip- SFC Profiles and Isosolid Diagrams of Model Systems
licate and the average SFC was determined from the three
NMR measurements. Crystallization curves were plotted SFC profiles were used to compare the melting properties
and fit to an empirical model using Origin 9.1 software. of the fat blends. Figure 1 shows the SFC profile for CB
The model equation was selected based on the exponential and blends of CB with (a) 10%, (b) 20%, and (c) 30% lau-
growth shape of data and quality of fit (Eq. 1): ric fat. Blends with 10% lauric fat (Fig. 1a) had lower SFC
than CB at a given temperature, and the three blends exhib-
SFC = A1 − A2 e−kt , (1) ited similar SFC profiles across the melting range. Below
where SFC = solid fat content (%), A1 = equilibrium solid 25 °C, FPKO had a slightly higher SFC than PKO and CO,
fat content (%), A2 = pre-exponential factor (%), t = time which were not different at any temperature. However, the
(min), and k = rate parameter (min−1). The fit of the model SFC of the three blends were nearly identical above 25 °C.
was analyzed using linear regression to compare the SFC The SFC profiles of blends with 20% lauric fat (Fig. 1b)
predicted by the model to the SFC determined experimen- showed greater deviation from the SFC profile of CB than
tally. Standard error was used to evaluate the fit of each 10% blends and displayed wider melting curves. At 20 °C,
constant within individual experiments. Constants were distinctions were observable between types of fat, with
compared across different experiments using the standard FPKO having the highest SFC followed by PKO and sub-
error and evaluating the z-score for multiple comparisons sequently CO. Compared to CB, which had a slow decrease
(α < 0.05). followed by a rapid decline in SFC at 25 °C, blends of 30%
lauric fat (Fig. 1c) exhibited significant decreases in SFC
Rheology of Chocolate Systems at 10 °C with more gradual melting at higher temperatures.
Whereas the SFC profile of CB displayed a more convex,
A Discovery Hybrid Rheometer 2 (TA Instruments, Wood outwardly inflected shape due to sharp melting properties,
Dale, IL, USA) equipped with parallel-plate geometry the SFC profiles of the 30% blends appeared concave or
and TRIOS software was used to evaluate the rheological inward-facing. The differences in solid fat profiles can be
properties of chocolate systems. The distance between the explained by the triacylglycerol compositions of each fat.
upper and lower plates was calibrated prior to each sample Lauric fats are rich in short- and medium-chain triacylglyc-
loading. Each sample was subjected to cyclothermic tem- erols, which have lower melting points than the long-chain
pering according to Kleinert’s method and proper temper- triacylglycerols of CB, resulting in faster melting behavior
ing was confirmed using the two tests described previously. [15]. The exception to the trend of lower SFC with high
Approximately 0.8 g of sample were loaded and slightly lauric fat addition was found in FPKO at 10 °C. While
compressed between parallel plates set to a 1000-µm gap going from 0 to 10% reduced SFC at first, increasing FPKO
and temperature of 31 °C. Small amplitude oscillatory rhe- beyond 10% increased the SFC of the blend. It is likely that
ology was analyzed using a temperature sweep program the fractionation process yielded a palm kernel stearin with
from 31 to 12 °C at a cooling rate of 0.3 °C/min. Dynamic a melting point virtually the same as that of CB. When CB
oscillatory test parameters storage modulus (G′) and loss was blended with another fat of the same melting point, a
modulus (G″) were calculated using TRIOS software. depression in SFC corresponding to a lower melting point
Experiments were performed on three different samples, of the mixture was observed when going from 0 to 10% but
and average moduli were determined from the three meas- not from 20 to 30%.
urements. Crystallization curves of G′ and G″ over time Figure 2 shows the isosolid diagram for the addition of
were plotted. To compare data, G′ curves were sectioned (a) CO, (b) PKO, and (c) FPKO to CB. In both CO and
and assigned parameters: time of initial crystallization PKO systems (Fig. 2a, b), eutectic and dilution interactions
onset (t1), slope of initial crystallization phase (Δ1), time were evident across the composition range with the most
of secondary crystallization onset (t2), slope of secondary pronounced eutectics occurring at the 60% lauric fat addi-
crystallization phase (Δ2), and maximum G′ (G′max). Time tion levels. At high PKO levels (Fig. 2b), the isosolid lines
values were determined from inflection points on the curves were farther separated from each other, indicating broad

13
J Am Oil Chem Soc

10% PKO 20% PKO 30% PKO


80 10% CO 80 80 30% CO
20% CO
10% FPKO 30% FPKO
20% FPKO
CB CB
CB

Solid Fat Content (%)


Solid Fat Content (%)
60

Solid Fat Content (%)


60 60

40 40 40

20 20 20

0 0 0
10 30 10 30 10 30
Temperature (oC) Temperature (o C) Temperature (oC)
(a) (b) (c)

Fig. 1  Solid fat profiles for blends of 10% (a), 20% (b), and 30% (c) coconut oil (CO), palm kernel oil (PKO), and fractionated palm kernel oil
(FPKO) with cocoa butter (CB). Data represent means of measurements of three different mixtures with error bars indicating standard deviation

Fig. 2  Isosolid diagrams for the addition of (a) coconut oil (CO), (b) palm kernel oil (PKO), and (c) fractionated palm kernel oil (FPKO) to
cocoa butter (CB). Each line represents a constant solid fat content (SFC) given by the number above the line

melting ranges for these fat blends. It can be reasoned that forms of the two fats. The isosolid diagram for the blends
CO and PKO are both incompatible with CB, and the great- of CB with FPKO (Fig. 2c) showed a clear eutectic at the
est extent of incompatibility occurs at the 60% addition 70% addition level, which was a slightly higher value than
level. The eutectic interaction and incompatibility of CB that observed in CO and PKO blends. Unlike CO and PKO,
and lauric fat has been extensively reported in the literature FPKO and CB had about the same melting points across
[1–3] and can be attributed to the weak co-crystallization the range of temperatures. As explained previously, these
of shorter-chain PKO triacylglycerols and longer-chain CB findings may be due to the fractionation process yielding
triacylglycerols in addition to differences in polymorphic a palm kernel stearin with high melting properties similar

13
J Am Oil Chem Soc

to CB. According to Calliauw et al. [16], unhydrogenated


palm kernel stearin has exceptionally sharp melting and
crystallization behavior that make it ideal for use as a CB
substitute. However, it still shows eutectic phase behavior
when mixed with CB due to differences in triacylglycerol
profile and polymorphism.

Crystallization Curves of Model Systems

An empirical model was applied to the crystallization


curves in order to describe the isothermal crystallization
kinetics of fat mixtures. By fitting data to a model, the crys-
tallization properties of the blends could be compared. The
model equation (Eq. 1) was chosen based on the exponen-
tial growth shape of data:

SFC = A1 − A2 e−kt , (1)


Fig. 4  Representative solid fat content (SFC) cooling curves for the
where SFC = solid fat content (%), A1 = equilibrium solid crystallization at 15 °C of cocoa butter (CB) and blends with 20%
fat content (%), A2 = pre-exponential factor (%), t = time coconut oil (CO) palm kernel oil (PKO), and fractionated palm kernel
(min), and k  = rate parameter (min−1). Once the model oil (FPKO). The continuous lines represent the fit of the experimental
data using the model equation (Eq. 1). Data points represent means
equation was verified, its overall quality of fit was assessed
of measurements of three different mixtures with error bars indicat-
by linear regression analysis. Figure 3 depicts the compari- ing standard deviation
son between the SFC predicted by the model to the SFC
measured experimentally. The relationship between pre-
dicted versus measured SFC was linear with an R2 value after 60 min. Although the Avrami model has been widely
of 0.99657, indicating that the model equation provided used to quantify the solidification kinetics of confectionery
a good fit and could accurately predict SFC for the data. fats [17, 18], this theory is based on isothermal crystalli-
The fit of the exponential model to the experimental SFC zation without any pre-crystallization (tempering or seed-
data can be seen in Fig. 4 using the crystallization of CB ing). As the fats used in this study were tempered prior to
and blends at 15 °C for illustration. All crystallization SFC measurement, a lag or nucleation phase was absent,
curves displayed an initial exponential increase in SFC and crystallization was underway before the hold tempera-
followed by an asymptote at equilibrium SFC. For all fat ture was reached. Average cooling rates ranged from 0.05
blends, crystallization slowed down after approximately to 0.1 °C/min, which varied based on the crystallization
30 min, with most blends having reached equilibrium SFC temperature, with crystallization starting immediately upon
cooling. Thus, the crystallization curves in this study did
100 not conform to the Avrami model. Further, a simpler model
with a single A value also did not fit the data very well.
Measured Solid Fat Content (%)

Ultimately, the model in Eq. (1) was selected as the best for


80 y = 0.9976x + 0.0843
data analysis.
R² = 0.9966
Kinetic parameters of crystallization are compared in
60 Table 1. At all three cooling temperatures, pure CB reached
the highest equilibrium SFC (A1) compared to blends with
40 lauric fats. For blends containing CO or PKO, increasing the
concentration of lauric fat (10–30%) decreased equilibrium
20 SFC, which is in agreement with previous research [4, 8].
Triacylglycerols that contain long-chain saturated fatty acids
tend to crystallize more quickly and have higher melting
0
0 20 40 60 80 100 points than those with shorter chains. Thus, fats with higher
Solid Fat Content Predicted by Model (%) proportions of long-chain triacylglycerols (CB and FPKO)
have higher SFC and crystallize earlier during cooling. Simi-
Fig. 3  Relationship between the measured solid fat content (SFC) for
lar studies on fat blend crystallization showed that the most
all experiments in triplicate and the SFC predicted by the exponential pronounced effects on kinetics were observed when the chain
model (Eq. 1) length of the additive was markedly dissimilar from that of

13
J Am Oil Chem Soc

Table 1  Equilibrium SFC (A1), pre-exponential factor (A2), and rate parameter (k) for crystallization of cocoa butter (CB) and blends with coco-
nut oil (CO), palm kernel oil (PKO), or fractionated palm kernel oil (FPKO)
10 °C % in blend Fat A1 (%) A2 (%) k (min−1)

100 CB 80.4 (±0.1)A 56.5 (±0.5)E 0.28 (±0.01)A


10 CO 71.0 (±0.2)F 53.7 (±0.6)F 0.19 (±0.00)B
PKO 71.8 (±0.2)E 55.7 (±0.7)E 0.18 (±0.00)B
FPKO 76.6 (±0.2)D 60.4 (±0.6)C 0.16 (±0.00)C
20 CO 69.8 (±0.2)G 58.8 (±0.5)D 0.12 (±0.00)E
PKO 70.7 (±0.2)F 59.8 (±0.5)CD 0.13 (±0.00)D
FPKO 78.5 (±0.4)C 67.7 (±0.4)B 0.12 (±0.00)E
30 CO 68.6 (±0.2)H 61.8 (±0.4)C 0.08 (±0.00)H
PKO 68.8 (±0.2)H 61.1 (±0.5)C 0.09 (±0.00)G
FPKO 79.1 (±0.1)B 72.6 (±0.4)A 0.11 (±0.00)F
15 °C % in blend Fat A1 A2 k
A BC
100 CB 73.4 (±0.2) 48.0 (±0.7) 0.17 (±0.00)B
10 CO 61.5 (±0.1)E 46.2 (±0.5)D 0.17 (±0.00)B
PKO 61.9 (±0.2)D 45.6 (±0.6)D 0.19 (±0.01)A
FPKO 66.0 (±0.2)C 49.3 (±0.4)B 0.11 (±0.00)C
20 CO 55.1 (±0.3)H 43.3 (±0.6)E 0.08 (±0.00)D
PKO 58.7 (±0.1)F 48.0 (±0.2)C 0.08 (±0.00)E
FPKO 69.4 (±0.4)B 58.4 (±0.6)A 0.07 (±0.00)F
30 CO 53.0 (±0.7)I 45.3 (±0.7)D 0.04 (±0.00)H
PKO 56.9 (±0.2)G 49.3 (±0.3)B 0.05 (±0.00)G
FPKO 67.8 (±1.0)BC 57.5 (±1.2)A 0.04 (±0.00)H
20 °C % in blend Fat A1 A2 k

100 CB 64.4 (±0.2)A 42.2 (±0.4)A 0.12 (±0.00)A


10 CO 51.3 (±0.1)B 35.1 (±0.2)C 0.09 (±0.00)C
PKO 51.3 (±0.1)B 35.5 (±0.3)C 0.10 (±0.00)B
FPKO 51.7 (±0.2)B 35.1 (±0.4)C 0.09 (±0.00)C
20 CO 41.2 (±0.4)D 29.8 (±0.5)D 0.06 (±0.00)D
PKO 41.8 (±0.1)D 31.5 (±0.2)D 0.06 (±0.00)D
FPKO 50.9 (±0.7)B 40.6 (±0.6)B 0.04 (±0.00)E
30 CO 34.2 (±0.3)E 26.9 (±0.3)E 0.04 (±0.00)F
PKO 36.2 (±1.8)E 28.7 (±1.6)DE 0.03 (±0.01)FG
FPKO 47.9 (±1.0)C 40.2 (±0.9)B 0.03 (±0.00)G

Kinetic parameters were obtained by fitting the model equation (Eq. 1) to data. Different letters denote significant differences (p < 0.05) within
columns. Standard error is located in parenthesis next to each mean value

the continuous lipid phase. Dissimilarities in chain length or initial region of the SFC cooling curves. Blends with
introduced some degree of disorder during crystal network higher concentrations of lauric fats had lower rate parameter
formation [19]. Conversely, for FPKO, increasing concentra- values, indicating delayed crystal growth of the blends rela-
tion increased equilibrium SFC at 10 °C. At all temperatures, tive to CB. In confectionery applications, crystallization delay
the order of equilibrium SFC value was as follows: CB (great- is highly undesirable, as slower crystallization can advance
est), FPKO, CO or PKO (least). Pre-exponential factor (A2), fat bloom formation [9]. Kinetic parameters of crystallization
the parameter that describes the inflection of the exponential were also compared across the three cooling temperatures.
model, increased for each blend with increasing addition level As temperature was increased from 10 to 20 °C, significant
of lauric fat at 10 °C, indicating less-defined curvature of reductions in both A1 and k values were observed, ratifying
crystallization curves with increasing lauric fat concentration. slower growth rates and diminished equilibrium SFC of the
Rate parameter (k) correlates to the period of crystal growth fats. Also, at high temperatures, intact PKO and CO displayed

13
J Am Oil Chem Soc

identical crystallization profiles and the same values of A1 crystallization of CB releases latent heat to balance cool-
and k. Lower temperatures shifted the crystallization behavior ing and maintain constant temperature briefly (the plateau
of the lauric fat blends more closely to that of pure CB and in the tempermeter curve), followed by another decrease in
accelerated crystal growth. In general, the addition of lauric temperature as solidification continues [20]. Furthermore,
fat had an inhibitory effect on crystallization, but the extent of the rheometer controlled temperature using a cooling ramp
this effect depended on the temperature of cooling. at a set rate of 0.3 °C/min, so all chocolate systems were
cooled at a constant rate. This contrasts with the crystal-
Rheology of Chocolate Systems lization of the model systems that was performed under
temperature-hold conditions that resulted in variable cool-
Chocolate blends required proper tempering prior to rhe- ing profiles. These discrepancies resulted in differently
ological analysis. Each blend was subjected to an indi- shaped G′ curves, and possibly influenced crystallization
vidualized tempering profile with two temperature cycles curves as well, of the chocolate and model systems. A sec-
to ensure proper pre-crystallization (Table 2). The exact ond explanation for the unusual curves might be based on
temperatures of tempering steps differed slightly among the mechanism of seed crystal growth during cooling of
blends. In general, blends with higher lauric fat contents tempered CB. Spherulites grow uninhibitedly at first, but
required lower temperatures in order to achieve good increasingly begin to overlap with each other. This overlap
temper, confirming the results from the model systems in consequently slows further growth as the liquid phase is
which lauric fats inhibited CB crystallization. Also, in the squeezed out between spherulites, then crystallization lev-
SFC cooling curves, equilibrium SFC values were lower at els off at equilibrium [21]. The implications of this mecha-
a given cooling temperature for blends with greater lauric nism may be reflected in the G′ curves. After a lag period, a
fat levels, as these blends required lower temperatures than critical time was reached (t1) at which G′ rapidly increased.
CB to achieve stable crystallization during cooling. Repre- This initial rapid burst (Δ1) would correspond to an ini-
sentative tempermeter curves indicated that chocolate and tial, unimpeded 3D growth of spherulites. As growth pro-
the lauric fat blends were properly tempered. ceeds, a crystallization rate change occurs at a later point
Figure  5a–d shows examples of G′ curves obtained for (t2), signifying a region of impeded 3D growth (Δ2) where
(a) chocolate and blends with (b) 10%, (c) 20%, and (d) spherulites butt into each other. Once the liquid phase has
30% PKO added. G′ increased relative to G″ over time for been depleted between spherulites, G′ slows to an equilib-
all systems, reflecting an increase in solid character due rium (G′max). Variations arose between chocolate blends in
to the formation of a crystal network, but multiple slopes the kinetics of these curve stages. In the final portion of the
were observed over the crystal growth period. Choco- curve, chocolate showed a stable G′ that was clearly greater
late (Fig. 5a) exhibited a multi-step increase in G′: a lag than G″, which decreased slightly as the liquid phase was
phase, initial rapid growth (Δ1), secondary gradual growth incorporated into the crystal matrix. Both 10 and 20% lau-
(Δ1), and equilibrium. The unusually shaped curve was ric fat (Fig. 5b, c) depicted multi-step crystallization simi-
due to the temper step prior to rheological testing. Cool- lar to chocolate. As lauric fat level was increased in a blend
ing causes the temperature of chocolate to decrease, but the (10–30%), the final portion of a curve was less defined,
decrease in G″ was less evident, and stable G′ was reached
Table 2  Temperatures for tempering profiles of blends of chocolate later during cooling or not at all. Blends with 30% lauric
(CHOC) with 10, 20, and 30% addition levels of coconut oil (CO), fat did not reach stable G′ or clear equilibria after 60 min
palm kernel oil (PKO), and fractionated palm kernel oil (FPKO)
had elapsed (Fig. 5d), indicating incomplete crystallization
% in blend Fat T1 T2 T3 T4 within the 60-min cooling study. Furthermore, these results
confirm the changes to cooling curve shape and behavior
100 CHOC 25.5 29.5 28.0 31.0
that were observed in model systems, confirming that high
10 CO 25.0 29.0 27.5 31.0
concentrations of lauric fat can dramatically alter the over-
PKO 25.0 29.0 27.5 31.0
all crystallization of both model and chocolate systems.
FPKO 25.0 29.0 27.5 31.0
G′ curves were evaluated according to regions of ini-
20 CO 24.5 28.5 27.0 31.0
tial onset time (t1), initial growth slope (Δ1), secondary
PKO 24.5 28.5 27.0 31.0
onset time (t2), slope of secondary growth phase (Δ2), and
FPKO 24.5 28.5 27.0 31.0
maximum G′ (G′max). Kinetic parameters for crystalliza-
30 CO 24.0 28.0 26.5 31.0
tion of chocolate and blends are tabulated in Table 3. Initial
PKO 24.0 28.0 26.5 31.0
onset time (t1) was determined from the point of inflection
FPKO 24.0 28.0 26.5 31.0 where the lag period transitioned into a sharp increase in
T1 is initial cooling temperature, T2 is warming temperature, T3 is G′. The highest t1 values were observed in 30% CO and
recooling temperature, and T4 is final holding temperature PKO, suggesting that the two blends required the longest

13
J Am Oil Chem Soc

Fig. 5  The storage modulus (G′) and loss modulus (G″) tial onset time (t1), initial growth slope (Δ1), secondary onset time
for crystallization of a chocolate with b 10%, c 20%, and d 30% (t2), secondary growth slope (Δ2), and maximum G′ (G′max). Data
palm kernel oil (PKO). Each G′ curve was divided into phases and represent means of triplicate measurements with error bars indicating
ascribed parameters related to different stages of crystallization: ini- standard deviation

cooling time to initiate rapid crystal growth and validat- triacylglycerol chain lengths relative to those of the lauric
ing what was seen in the crystallization of CB/lauric fat fat blends. In the model systems, pure CB consistently dis-
model systems. Despite parallels between data sets, some played stronger crystallization characteristics than any mix-
contradictions were evident. Chocolate had the highest t1 ture. Chocolate did show the highest value for secondary
and t2 values behind only 30% CO and PKO. In princi- growth slope (Δ2), but its maximum G′ (G′max) was not the
ple, pure chocolate would be expected to have the short- absolute highest as would be expected. The superior crys-
est t1 and t2, or fastest times, given the homogeneity of its tallization properties that differentiated pure CB from its

13
J Am Oil Chem Soc

Table 3  Initial onset time % Fat t1 (min) Δ1 [log(Pa)/min] t2 (min) Δ2 [log(Pa)/min] G′max [log(Pa)]
(t1), initial growth slope (Δ1),
secondary onset time (t2), 100 CB 19.1 (±0.5)B 0.47 (±0.01)D 24.6 (±0.4)B 0.14 (±0.02)B 7.40 (±0.01)A
secondary growth slope (Δ2),
10 CO 14.4 (±0.4)DE 0.58 (±0.02)AB 19.6 (±0.6)D 0.09 (±0.07)C 7.40 (±0.05)A
and maximum G′ (G′max) for
crystallization of chocolate PKO 14.9 (±0.2)CDE 0.62 (±0.03)A 19.4 (±0.3)D 0.07 (±0.03)CD 7.39 (±0.02)A
with cocoa butter (CB) added FPKO 14.0 (±0.8)E 0.56 (±0.04)ABC 19.5 (±0.9)D 0.07 (±0.02)CD 7.40 (±0.05)A
and blends of 10, 20, and 30% 20 CO 17.1 (±1.4)BC 0.49 (±0.02)CD 21.3 (±1.3)C 0.07 (±0.02)CD 7.35 (±0.02)A
coconut oil (CO), palm kernel
oil (PKO), and fractionated PKO 16.8 (±0.8)BCD 0.52 (±0.04)BCD 20.5 (±0.7)C 0.07 (±0.01)CD 7.35 (±0.01)A
palm kernel oil (FPKO) FPKO 18.4 (±0.6)B 0.45 (±0.01)D 22.8 (±0.5)C 0.06 (±0.04)D 7.28 (±0.07)AB
30 CO 24.1 (±1.1)A 0.44 (±0.02)D 30.2 (±1.6)A 0.06 (±0.09)E 7.10 (±0.02)C*
PKO 23.3 (±0.3)A 0.45 (±0.04)D 28.3 (±0.8)B 0.16 (±0.01)A 7.17 (±0.08)BC*
FPKO 19.0 (±1.4)B 0.51 (±0.03)BCD 24.4 (±2.0)B 0.06 (±0.01)E 7.34 (±0.03)A*

Point values (t1, t2) were determined from inflection points on the cooling curves and used to calculate
slopes (Δ1, Δ2). Different letters denote significant differences (p < 0.05) within columns. The standard
deviation is located in parentheses next to each mean value
*
  G′max values that do not reflect equilibrium within 60 min of cooling

blends in model systems did not always translate to choco- indicating that the addition of lauric fat appeared to accel-
late in the rheology data. This may be due to differences in erate initial crystallization rate. This result was unexpected
cooling rate and control of cooling between rheology and as it goes against the inhibition theory. For CO and PKO,
NMR studies. The rheometer used in the study of chocolate as the amount of lauric fat was increased from 10 to 30%,
cooled all blends at a set ramp of 0.3 °C/min, while CB/ slope values decreased, verifying the slower crystallization
lauric fat model systems were cooled at slower rates overall rates with greater addition levels of lauric fat observed in
that ranged from 0.06 to 0.1 °C/min depending on crystalli- the model systems.
zation temperature. Moreover, the model systems were not The slope of the secondary growth phase (Δ2) was cal-
cooled at a ramp; the temperature-hold conditions under culated for the region between t2 and the next inflection
which crystallization was performed led to variable cooling point in a curve. These results were more complimentary
profiles. overall to the results of the model systems and the expected
The time of secondary onset was computed from the effects of lauric fat on crystallization behavior. Aside from
point of curve inflection succeeding the linear Δ1 region. 30% CO, chocolate had the highest Δ2 of 0.14 log(Pa)/min,
Like t1, the highest values of t2 were for the systems made which was significantly higher than any of the blends with
with 30% CO and 30% PKO, and t2 increased with increas- 0.06–0.09 log(Pa)/min. This ratifies previous findings that
ing amount of lauric fat. For both t1 and t2, no relationship the addition of lauric fat resulted in slower rates of crys-
could be identified between type of fat and onset of crys- tal growth and the consequent interference on the solidi-
tallization stages, which challenged the general order of fication process of chocolate. Thus, the inhibition theory
FPKO-PKO-CO crystallization that was manifested in CB/ applied during the secondary growth phase, with higher
lauric fat model systems. Therefore, in chocolate systems, lauric fat levels corresponding to slower rates of secondary
the addition level of lauric fat is a more influential factor on crystallization and pure chocolate showing the overall fast-
crystallization kinetics and behavior than specific fat type est secondary growth rate. However, the exception was the
added. A similar conclusion was reached by Baldino et al. unusually high Δ2 value for 30% PKO. As discussed, the
[11], who investigated the rheological properties of choco- cooling curve for 30% PKO did not reach a final equilib-
late in blends with anhydrous milkfats of different melting rium and appeared to be still in the process of crystallizing
points and found that addition level had a more signifi- after 60 min had elapsed, which could impact its Δ2 value.
cant impact on chocolate crystallization than the variety of This might be because 30% PKO was still undergoing rapid
milkfat used. While milkfat is not a lauric fat, similarities crystallization during its secondary growth phase, while
exist between the two fats in their properties in chocolate the secondary growth rates of the other blends had slowed
systems, as milkfat also forms eutectics with CB and can down because their crystallization processes were nearer to
impede its crystallization [17]. completion.
The slope of the initial growth phase (Δ1) was defined as G′max described the maximum G′ reached by each blend
the slope of the linear region of the curve between the t1 and after crystallization had elapsed for 60 min. In fat systems,
t2 inflection points. The three 10% blends along with 20% G′max is often correlated directly to equilibrium SFC [22].
CO and PKO showed higher slopes than pure chocolate, G′ is listed in units of log(Pa) in Table 3 for clarity and ease

13
J Am Oil Chem Soc

of comparison. G′max was a stable value for blends with 2. Ali ARM, Dimick PS (1994) Melting and solidification char-
lower lauric fat contents. However, the G′max of the 30% acteristics of confectionery fats: anhydrous milk fat, cocoa
butter and palm kernel stearin blends. J Am Oil Chem Soc
blends do not reflect final equilibrium values and cannot be 71(8):803–806
compared as indicated in Table 3. Chocolate and all blends 3. Williams SD, Ransom-Painter KL, Hartel RW (1997) Mixtures
reached relatively high G′max of 7.10–7.40 log(Pa), indi- of palm kernel oil with cocoa butter and milk fat in compound
cating more solid-like than liquid-like character with high coatings. J Am Oil Chem Soc 74:357–366
4. Quast LB, Luccas V, Ribeiro APB, Cardoso LP, Kieckbusch
hardness. The high G′max value of chocolate was expected TG (2013) Physical properties of tempered mixtures of cocoa
given the equilibrium SFC of pure CB in model systems. butter, CBR and CBS fats. J Food Sci Technol 48:1579–1588
These results also can be related to the isosolid diagrams, 5. Paulicka FR (1973) Phase behavior of cocoa butter extenders.
which were used to analyze the equilibrium conditions in Chem Ind 17:835–839
6. Hachiya I, Koyano T, Sato K (1989) Seeding effects on crystal-
blends of lauric fat with CB. The isosolid diagrams showed lization behavior of cocoa butter. Agric Bio Chem 53:327–332
that pure CB had the highest equilibrium SFC at a given 7. Himawan C, Starov VM, Stapley AGF (2006) Thermodynamic
temperature than any of the lauric fat blends. In rheology and kinetics aspects of fat crystallization. Adv Colloid Inter-
data, the outliers with significantly lower G′max were 30% face 122:3–33
8. Ribeiro APB, Basso RC, dos Santos AO (2013) Hardfats as
CO and PKO. This authenticates the data from the model crystallization modifiers of cocoa butter. Eur J Lipid Sci Tech-
systems, in which 30% CO and PKO reached the low- nol 115:1462–1473
est equilibrium SFC at 10 °C. Additionally, in the isosolid 9. Lonchampt P, Hartel RW (2004) Fat bloom in chocolate and
diagrams, 30% blends had lower SFC across the range of compound coatings. Eur J Lipid Sci Technol 106:241–274
10. Wright AJ, Scanlon MG, Hartel RW, Marangoni AG (2001)
measurement temperatures than 10 or 20%. Therefore, par- Rheological properties of milk fat and butter. J Food Sci
allels can be drawn between phase behavior of model sys- 66:1056–1071
tems, crystallization kinetics of model systems, and crystal- 11. Baldino N, Gabriele D, Migliori M (2010) The influence of
lization kinetics of chocolate systems studied by rheology, formulation and cooling rate on the rheological properties of
chocolate. Eur Food Res Technol 231:821–828
particularly in terms of equilibrium SFC/G′max and the influ- 12. Rigolle A, Gheysen L, Depypere F, Landuyt A, Van Den

ence of lauric fat on final equilibrium. Abeele K, Foubert I (2015) Lecithin influences cocoa butter
crystallization depending on concentration and matrix. Eur J
Lipid Sci Technol 117:1722–1732
13. Timms RE (1979) Computer program to construct isosolid dia-
grams for fat blends. Chem Ind 7(4):257–258
Conclusions 14. Kleinert J (1970) Cocoa butter and chocolate: the cor-

relation between tempering and structure. Rev Int Choc
Delayed crystal growth and reduction in equilibrium SFC 25(11):386–399
15. Karabulut I, Turan S, Ergin G (2004) Effects of chemical inter-
by higher lauric fat levels were manifested in both model esterification on solid fat content and slip melting point of fat/oil
and chocolate systems, effectively bolstering the claim that blends. Eur Food Res Tech 218:224–229
the use of lauric fat in cocoa butter systems is commercially 16. Calliauw G, Foubert I, De Greyt W, Dijckmans P, Kellens M,
limited. Solidification kinetics of the cocoa butter model Dewettinck K (2005) Production of cocoa butter substitutes via
two-stage static fractionation of palm kernel oil. J Am Oil Chem
systems adhered to an exponential model. Thus, although Soc 82(11):783–789
a sigmoidal model may be suitable for untempered fats, an 17. Metin S (1997) Crystallization behavior of blends of cocoa but-
exponential model is most fitting for tempered fats. By con- ter and milk fat or milk fat fractions. PhD Thesis, University of
trast, chocolate systems displayed multi-step crystalliza- Wisconsin-Madison
18. Marangoni AG (1998) On the use and misuse of the Avrami
tion behavior that precluded the fitting of a singular math- equation in the characterization of the kinetics of fat crystalliza-
ematical model. When applied to chocolate crystallization tion. J Am Oil Chem Soc 75(10):1465–1467
kinetics, oscillatory rheology can shed light on fat matrix 19. Smith KW, Bhaggan K, Talbot G, Van Malssen KF (2011) Crys-
formation under different conditions and model real-world tallization of fats: influence of minor components and additives.
J Am Oil Chem Soc 88:1085–1101
processing in a controlled cooling environment. Therefore, 20. Hartel RW (1998) Phase transitions in chocolate and coatings.
the results for chocolate crystallization measured by oscil- In: Rao MA, Hartel RW (eds) Phase/state transitions in foods:
latory rheology are more translatable to industrial processes chemical, structural, and rheological changes. Marcel Dekker
and useful in product development applications. Inc, New York, pp 227–240
21. Kinta Y, Hartel RW (2010) Bloom formation on poorly tem-
pered chocolate and effects of seed addition. J Am Oil Chem Soc
87(1):19–27
References 22. Marangoni AG, Narine SS, Acevedo NC, Tang D (2012) Nano-
structure and microstructure of fats. In: Marangoni AG, Wesdorp
1. Stewart IM, Timms RE (2002) Fats for chocolate and sugar LH (eds) Structure and properties of fat crystal networks. CRC
confectionery. In: Rajah KK (ed) Fats & Food Technology. Press, Boca Raton, pp 173–227
Sheffield Academic Press, Sheffield, pp 159–191

13

Вам также может понравиться