Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Available online at www.sciencedirect.

com

Journal of Food Engineering 88 (2008) 315–322


www.elsevier.com/locate/jfoodeng

Thermophysical properties of processed meat and poultry products


Michèle Marcotte a,*, Ali R. Taherian a, Yousef Karimi a,b
a
Food Research and Development Centre, Agriculture and Agri-Food Canada, 3600 Casavant Blvd West. St. Hyacinthe, Quebec, Canada J2S 8E3
b
Department of Food Science and Agricultural Chemistry, McGill University, Macdonald Campus, 21,111 Lakeshore Road,
Ste Anne de Bellevue, Quebec, Canada H9X 3V9

Received 6 September 2007; received in revised form 7 February 2008; accepted 13 February 2008
Available online 10 March 2008

Abstract

Thermophysical properties of various meat and poultry emulsions were evaluated at four temperatures (20, 40, 60 and 80 °C).
Thermal conductivities (0.26–0.48 W m1 K1) increased linearly with temperature between 20 and 60 °C. Between 60 and 80 °C, it
remained constant for most products except bologna. Curves for thermal conductivity as a function of temperature could be roughly
grouped into two different categories: products containing meat particles and those containing meat emulsions. The application of var-
ious models was investigated for thermal conductivity prediction. It was found that a three phase structural based Kirscher model had
the potential for predicting thermal conductivities with acceptable accuracy. Densities decreased slightly as a function of temperature
from 20 to 40 °C. A transition phase was observed from 40 to 60 °C, which was followed by a decrease from 60 to 80 °C. There was
a decrease of about 50 kg m3 between the density of a raw product at room temperature (at maximum 1070 kg m3) and the product
heated to 80 °C (at minimum 970 kg m3), due to the gelation or setting of the structure. After a transition period from 10 to 30 °C,
the heat capacity increased linearly from 30 to 80 °C, and ranged from 2850 to 3380 J kg1 °C1, respectively. Densities and heat capac-
ities were strongly influenced by the carbohydrate content (i.e. as the carbohydrate content increased the density decreased). The salt
content adversely affected thermal conductivity and thermal diffusivity values. However, these parameters increased with moisture
content.
Ó 2008 Published by Elsevier Ltd.

Keywords: Meat and poultry emulsions; Meat and poultry products; Thermal conductivity; Thermophysical properties; Modeling; Correlation matrix

1. Introduction (Unklesbay et al., 1999). As an example, the temperature


at the core of a typical sausage must be above a certain
Thermophysical properties (specific heat, thermal con- level (72 °C) by the end of heating and below certain
ductivity, thermal diffusivity and density) of food are temperature (15 °C) at the end of cooling in order to
important parameters in describing various thermal achieve microbiological stability of the product (Akterian,
processes, optimizing the design and the operation of heat- 1997). Therefore, there is a need to evaluate the heating
ing, cooking, freezing and cooling systems (Karunakar characteristics of the products, known as thermophysical
et al., 1998). Thermal properties are also essential for the properties.
modeling and evaluation of food processing operations There are many methods available to measure thermo-
involving heat transfer, especially when energy costs, food physical properties such as guarded hot plate method,
quality and safety are the main considerations. These prop- differential scanning calorimeter (DSC) attachment
erties are especially important to ensure food safety method, capped column test device, line heat source
probe method, temperature history and transient hot
strip (THS) method (Baik et al., 2001). While the first
*
Corresponding author. Tel.: +1 450 768 3295; fax: +1 450 773 2888. three mentioned methods are applied at steady state,
E-mail address: marcottem@agr.gc.ca (M. Marcotte). the last two methods are applied for transient conditions.

0260-8774/$ - see front matter Ó 2008 Published by Elsevier Ltd.


doi:10.1016/j.jfoodeng.2008.02.016
316 M. Marcotte et al. / Journal of Food Engineering 88 (2008) 315–322

Nomenclature

Cp specific heat at constant pressure (kJ m3 SD Standard error


o 1
C ) T temperature
E predicted value
f a distribution factor Greek symbols
I current (mA) a thermal diffusivity (m2 s1)
k thermal conductivity (W m1 K1) U volume fraction
m slope of the linear part of the temperature vs. ln q density (kg m3)
(time) graph
N number of samples Subscripts
O experimental value a air
Q heat flux (W m1) i component i
Rth probe resistance (m1) s solid
r2 coefficient of determination w water

Thermal conductivity is highly temperature dependent ture, even though they often contain empirical parame-
especially over a temperature range where a phase change ters. The selection of the appropriate model for a given
occurs. According to Karunakar et al. (1998), at a low food product is not an easy task. As reported by Carson
temperature range (0–40 °C), the thermal conductivity et al. (2006), there is a huge disagreement between ther-
does not exhibit very significant difference between vari- mal conductivity values predicted from various models.
ous temperatures. At high temperatures (>50 °C), it It can be concluded that, to-date, no single model or pre-
increases gradually as the temperature increases (Pan diction procedure has been found with a universal appli-
and Singh, 2001). Both thermal conductivity and heat cability. Therefore, it must be evaluated on a case-by case
capacity are known to increase with moisture content basis.
increase (Shmalko et al., 1996). Water content will affect In recent decades, there have been many published
the heat capacity more than other components, the lower research work on experimental values of thermophysical
heat capacity values generally occur with the lower mois- properties of foods and mathematical models to represent
ture content values (Unklesbay et al., 1999). Thermal these data (Polley et al., 1980; Sanz et al., 1987; Lind,
conductivity and density of foods vary with temperature 1991; Karunakar et al., 1998; Baik et al., 2001; Gonzo,
during thermal processing due to the changes in texture 2002; Wang et al., 2006; Carson, 2006). However, there is
and/or composition (Karunakar et al., 1998). A decrease very little work done on food emulsions. The overall goal
in density values will become important due to its effect of this study was to measure thermophysical properties of
on other thermal properties (Mohsenin, 1980). Most various commercial meat and poultry emulsions as a func-
changes in meat products occur during heating, shrink- tion of temperature. A more specific objective was also to
age, tissue hardening, moisture loss, fat loss and discolor- investigate the ability of mathematical models to predict
ation, and are caused by the changes in muscle protein the thermal conductivity of these emulsions.
denaturation (Pan and Singh, 2001). All these changes
in the meat will affect the thermophysical properties. 2. Materials and methods
Mathematical models can be used to study and under-
stand the relationship between thermophysical properties 2.1. Products
of complex food systems and temperature. For the pre-
diction of the thermal conductivities, a large number of Five types of meat products were used: fine emulsion of
models have been suggested in the literature (Gonzo, bologna and wieners, coarse emulsion of pepperoni, turkey
2002; Wang et al., 2006; Carson, 2006), used for compos- emulsion and flaky ham. Turkey emulsions and the flaky
ite or heterogeneous materials based on composition. ham contained muscle parts. Raw products were collected
Most of the proposed models are highly material based from a typical industrial plant and measurements were
and contain material-specific parameters. Although, some made the day after. The products were kept at 4 °C cold
models are considered to have a more general applicabil- room until they were analyzed. All experiments were per-
ity, their parameters are still empirically determined. formed three times and with three different batches of
Generic models have been proposed by some scientists products. Thermophysical properties of meat and poultry
(Gonzo, 2002; Wang et al., 2006; Carson, 2006), where emulsions were gathered at different temperatures from
a set of equations, usually based on a conceptual ‘parent’ raw product to cooked product temperature. The composi-
model, are derived. These models are developed so that tion of meat and poultry emulsions studied is listed in
they can account for variations in composition and struc- Table 1.
M. Marcotte et al. / Journal of Food Engineering 88 (2008) 315–322 317

Table 1 Having the slope, it was possible to calculate the thermal


Composition of various meat and poultry emulsions (%) conductivity using the following equations:
Food Moisture Fat Salt Ash Protein Carbohydrate
emulsion Q ¼ Rth I 2 ð1Þ
Bologna 61.59 20.31 2.39 4.22 11.49 2.39 Q
k¼ ð2Þ
Pepperoni 57.26 21.72 2.53 3.48 12.32 5.22 4pm
Wieners 60.51 20.9 2.43 4.19 12.4 2.00
Turkey 74.88 1.67 1.54 6.69 15.46 1.30 where Q is the heat flux (W m1), Rth the probe resistance
Ham 72.70 6.35 2.78 7.08 11.62 2.25 (X m1), k the thermal conductivity (W m1 K1), I the
current (mA) and m the slope of the linear part of the tem-
perature vs. ln (time) graph.

2.2. Thermal conductivity measurements 2.3. Density measurement

Thermal conductivity measurements of various meat Densities were determined from the mass of the samples
samples were performed using the probe method based inserted in the copper cylinder and the volume of the
on the line-heat source approach developed by Sweat cylinders, which were pre-determined for six cylinders
(1974). In the probe method, there was a heater wire insu- and the average value was considered. The mass of the
lated over its length and a thermocouple at the center of samples was measured at the end of the treatment to verify
this length. The probe was 38 mm long with an outside if there was any weight change during the treatment. The
diameter of 0.66 mm. It consisted of a constantan heater density was calculated as the ratio of the final mass of
wire and a chromel–constantan thermocouple (type E) the sample to its volume.
(Sweat, 1995). The probe was connected to a power supply
(Hewlett-Packard, 6236B) and to a multimeter (TES 2610 2.4. Heat capacity
multimeter) in order to read the current more precisely;
the multimeter was set to read in mA DC. The thermocou- A modulated differential scanning calorimeter (MDSC
ple wires were connected to a data acquisition system (Data 2910, TA Instruments Inc., New Castle, DE) was used with
Shuttle by Strawberry Tree) connected to a computer (Baik a nitrogen cooling system. MDSC has the ability for applica-
et al., 1999). The software ‘‘Workbench for Windows ver- tion of two independent heating rates, isothermal and mod-
sion 3” was used to convert the analog signal of the ther- ulated, which is an advantage over conventional DSC that
mocouple into a digital signal, to set the acquisition rate carries only isothermal. This allowed determination of the
at 1 reading every 2 s. The probe had a theoretical internal heat capacity at various temperatures. At the start, the cell
resistance of 226.67 X m1. It was calibrated with glycerol. constant of the instrument was evaluated by running an
Values obtained were within 10% of the literature value of experiment with Sapphire (Al2O3). The ratio of the experi-
0.284 W m1 K1 at 20 °C. mental heat capacity to the theoretical heat capacity of Sap-
Constant temperature baths set in increments of 20 °C phire gave a cell constant of 1.935. A minimum mass of 200 g
were used for the experiments: water baths at 20 °C, product was then homogenized using a ‘‘Polytron” (Poly-
40 °C and 60 °C, and an 80 °C oil bath to prevent evapora- tron, PT 10–35 by Kinematica) in order to get a representa-
tion. A copper cylinder (12.7 cm height and 2.54 cm inside tive sample, especially for coarse emulsion, turkey and ham,
diameter with a maximum wall thickness of 0.159 cm) with and to fill up the experimental container which was small in
a high thermal conductivity was used. Emulsions samples capacity. A sample of 10–13 mg was placed in the aluminium
were inserted using a syringe, whereas the turkey and the pan which was then hermetically sealed using an encapsulat-
flaky ham were introduced by hand. A rubber cover was ing press from TA Instruments Inc. An empty pan that had
placed at both ends of the cylinder to insulate them and been previously weight-matched with the sample pan was
maintain heat flow only from the side of the cylinder. An also sealed for reference. The method consisted of equilibrat-
infinite cylinder was assumed for thermal conductivity cal- ing the sample at 5 °C, starting the modulation for a 60-sec-
culations. Three cylinders were placed in each of the four ond period with an amplitude of ±1 °C, keeping it
temperature controlled water bath. When the core temper- isothermal for 5 min, and heating the sample at a rate of
ature of the samples reached equilibrium with the water 2.0 °C/min from 5 °C to 90 °C. Helium was used as a purging
bath, the top rubber cover was replaced by a thinner one gas at a flow of 25 ml min1. The instrument automatically
and the probe was inserted in a small hole made at the cen- gave heat capacity curves from 5 to 90 °C.
ter. The probe was placed at the core of the sample. The Furthermore, the heat capacities (Cp) of the emulsions
data acquisition was switched on for 8 s to record the initial were related to temperature (T) using following linear
temperature. Then the power was turned on to pass a cur- regression equation:
rent of 200 mA for 2 min of data acquisition before being
C p ¼ a þ bT ð3Þ
stopped. The thermal conductivity was calculated by plot-
ting the temperature versus the natural logarithm of the where a is the intercept and b the slope of the regression
time and taking the slope of the linear part of this graph. line.
318 M. Marcotte et al. / Journal of Food Engineering 88 (2008) 315–322

2.5. Thermal diffusivity calculation Table 3


Thermal conductivity of food components as a function of temperature
(40 °C < T < (150 °C)
The thermal diffusivity values were calculated from the
experimental values of density, thermal conductivity and Component Thermal conductivity (W m1 K1)
specific heat using the following equation: Protein 1.7880  101 + 1.1958  103 T  2.7178  106 T 2

Fat 1.8071  101  2.7604  104 T  1.7749  107 T 2

k Carbohydrate 2.0141  101 + 1.3874  103 T  4.3312  106 T


a¼ ð4Þ
qC p ash
Ash 3.2962  101 + 1.4011  103 T  2.9069  106 T 2
where a is the thermal diffusivity (m2 s1), k the thermal Water 5.7109  101 + 1.7625  103 T – 6.7036  106 T 2
conductivity (W m1 K1), q is density (kg m3), Cp is spe- Air 2.3820  102 + 6.75  105 T
cific heat at constant pressure (kJ m3 °C1). Adapted from Ramaswawy and Marcotte (2006).

2.6. Thermal conductivity modeling


Carson et al. (2006). The component thermal conductivity
A generic three phase model, known as Kirscher’s model was determined using the functions presented in Table 3.
(Maroulis et al., 2002; Wang et al., 2006), was selected and Since the Kirscher’s weighing factor was not available for
used in this study. Kirscher’s model is one of the most meat and poultry emulsions, it was adjusted to get the best
widely used, and it is basically a combination of the predic- model performance i.e. by matching predicted and experi-
tion of the Series and Parallel models (Carson et al., 2006) mental values.
containing a weighting parameter (f) (also named distribu- In order to examine the performance of the applied
tion factor) which has to be adjusted for different food sys- model, the correlation between thermal conductivity values
tems. The model was formulated as follows: predicted by the model and those calculated from the
1 experimental data were evaluated using the coefficient of
ke ¼ ð5Þ determination (r2), standard errors and standard percent-
ð1f þ kfse Þ
kp age errors. The standard error was obtained by the follow-
k p ¼ /s k s þ /w k w þ /a k a ð6Þ ing equation:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P ffi
1 ðEi  Oi Þ
k se ¼ /s
ð7Þ SD ¼ ð8Þ
ks
þ /kww þ /kaa N
where kw, ks, ka are the thermal conductivities of water, sol- where Ei is the estimated value of thermal conductivity; Oi
ids and air, respectively, f is a distribution factor and Uw, is its observed (experimental) value; and N is the number of
Us, Ua are the volume fractions of water, solids and air, samples. The relative standard error is defined as the ratio
respectively. Thermal conductivities of pure food compo- of standard error to the average experimental value of ther-
nents are available in the literature as a function of temper- mal conductivity and expressed as a percentage.
atures. The volume fractions (Uw, Us, Ua) are dependent on
the material state. 3. Results and discussion
Since the compositions of meat and poultry emulsions
were determined and presented on a weight basis, the vol- 3.1. Thermal conductivity
ume fraction of each component was evaluated using its
density. The densities of various components, as a function As shown in Fig. 1, the thermal conductivity values
of temperature, are shown in Table 2 (Ramaswamy and increased with temperature for all products. For a fine
Marcotte, 2006). The thermal conductivity of the solid bologna emulsion, the thermal conductivity increased from
phase was calculated by combining the thermal conductiv- 0.304 ± 0.028 W m1 K1 at 22 °C to 0.459 ± 0.100
ity of each individual component (i.e., protein, carbohy- W m1 K1 at 80 °C; and for a fine wieners emulsion, the
drate, ash, and fat) using the procedure suggested by values varied from 0.281 ± 0.022 to 0.415 ± 0.066 W
m1 K1. For a coarse pepperoni emulsion, the thermal
conductivity varied between 0.272 ± 0.019 to 0.402 ±
Table 2 0.044 W m1 K1 from 22 to 79 °C. In the case of products
Density of food components as a function of temperature
(40 °C < T < (150 °C) containing muscle parts, it was found that the thermal con-
ductivity of turkey product varied between 0.332 ± 0.040
Component Density (kg m3)
and 0.482 ± 0.086 W m1 K1; and that for the ham pro-
Protein 1329.9–0.5184 T duct varied from 0.339 ± 0.037 to 0.437 ± 0.058
Fat 925.59–0.41757 T
Carbohydrate 1599.1–0.31046 T
W m1 K1. Measurements were performed using samples
Ash 2423.8–0.28063 T from three different production batches. Average values are
Water 997.18 + 3.1439  103 T – 3.7574  103 T 2
reported since there was a good agreement between the
Air 1.2847–3.2358  103 T results of these batches. Two sets of measurements were
Adapted from Ramaswamy and Marcotte (2006). performed on each batch of products to establish repeat-
M. Marcotte et al. / Journal of Food Engineering 88 (2008) 315–322 319

0.50

0.45

Thermal Conductivity (W m -1 K-1)


0.40

0.35

Bologna
0.30 Ham
Pepperoni
Turkey
Wieners
0.25

0.20
20 30 40 50 60 70 80 90
Temperature (ºC)

Fig. 1. Thermal conductivity of meat and poultry emulsions.

ability. Differences between values were found to be of the a small increase in the volume of the sample during the
same order of magnitude as the measurements undertaken heating stage of the emulsions. At the start of heating,
on two individual batches indicating that our methodology a slight increase in density, associated with all the tested
was highly repeatable. The observed differences between emulsions, was observed which may be due to the fat
batches were probably due to variation in the composition content changing phase from solid to liquid. Neverthe-
of each batch, which may occur during production of each less, the densities decreased slightly with temperature in
batch at industrial level. general (Fig. 2). For each product, the density of the final
product heated to 80 °C was lower by approximately
3.2. Density 50 kg m3 when compared with the initial raw product
kept at room temperature. The density of the raw pep-
The density changes observed during cooking were peroni and ham products were 1031 and 1037 kg m3
very small. There was no significant mass loss for these and for the cooked products were 969 and 1014 kg m3,
products. Gelation or setting of the solid structure caused respectively.

1080

1060

1040
Density (kg m-3)

1020

1000 Bologna
Pepperoni
Wieners
Turkey
980
Ham

960
20 30 40 50 60 70 80 90
Temperature (ºC)

Fig. 2. Density of meat and poultry emulsions.


320 M. Marcotte et al. / Journal of Food Engineering 88 (2008) 315–322

3500

3400

3300

Heat capacity (J kg-1 k-1)


3200

3100

3000

2900
Bologna
Pepperoni
2800 Wieners
TURKEY
Ham
2700
0 10 20 30 40 50 60 70 80 90
Temperature (°C)

Fig. 3. Heat capacity of meat and poultry emulsions.

3.3. Heat capacity Table 4


Correlation of heat capacity with temperature for various emulsions
It is well known that the heat capacity varies with tem- Emulsion Heat capacity (J kg1 K1) r2
perature, which was also confirmed by our results shown in Bologna 2836.8 + 4.306 T 0.996
Fig. 3. The heat capacity values of meat emulsions at con- Pepperoni 2719.1 + 3.760 T 0.978
stant pressure increased linearly between 35 °C and 82 °C Wieners 2962.4 + 4.282 T 0.990
Turkey 3003.3 + 4.567 T 0.998
with a gradient of 4.34 J kg1 K1. Amongst the emulsions Ham 2940.3 + 4.234 T 0.996
studied, the lowest heat capacity was found for the Pepper-
oni emulsion: 2812 J kg1 K1 at 10 °C and 3027 J kg1
K1 at 82 °C. This emulsion showed a peak of 3160 regression parameters along with the coefficient of determi-
J kg1 K1 at 26 °C. For the bologna emulsion, the heat nation (r2) values. High values of r2 indicate that heat
capacity was found to be around 2933 J kg1 K1 at capacity and temperature are noticeably related.
10 °C and 3191 J kg1 K1 at 82 °C. The heat capacity
curve for this emulsion also indicated a maximum value 3.4. Thermal diffusivity
of 3055 J kg1 K1 at 20 °C. The minimum was 2933
J kg1 K1 at 10 °C. The heat capacity curve for wieners Thermal diffusivity is another property that changes
had a similar shape compared to bologna, but the values with temperature (Fig. 4), since it is related to all three ther-
were approximately 150 J kg1 K1 higher over the curve mophysical properties described above: conductivity, heat
length. For the turkey product, the curve was almost linear capacity and density. The thermal diffusivity curves were
over the entire range of temperatures studied, i.e. 10 to very similar to thermal conductivity curves because other
82 °C with a slope of 3.9 J kg1 K1 and an ordinate of properties are not as sensitive to temperature. The thermal
3050 J kg1 K1. The ham product had a maximum heat diffusivity values of the emulsions varied from 8.30 
capacity of 3287 J kg1 K1 at 82 °C and a minimum of 108 m2 s1 (for pepperoni and wieners) at room tempera-
3031 J kg1 K1 at 10 °C, this curve also included a peak ture to 1.36  107 m2 s1 (for pepperoni) at 80 °C.
at 20 °C (3224 J kg1 K1). The average standard deviation
for all curves was 115.5 J kg1 K1. It has been reported 3.5. Correlation between composition and thermophysical
that the composition of product (i.e. fat, protein, water properties
and salt contents) will influence the values of these proper-
ties (Zhang et al., 2007). The peak values observed varied Values of thermophysical properties were measured for
directly with fat content, suggesting that samples contain- a variety of meat and poultry emulsions. Significant differ-
ing higher fat content absorb more thermal energy for ences were found that may be attributed to the meat and
changing the fat from solid to liquid state. poultry compositions. Average values of thermophysical
In order to be able to predict heat capacity of the emul- properties were therefore, correlated to the chemical com-
sions, the values of heat capacity were related to the tem- position. Table 5 shows the correlation matrix. A strong
perature in the range 34–82 °C. Table 4 shows the proportional relationship would be indicated by a value
M. Marcotte et al. / Journal of Food Engineering 88 (2008) 315–322 321

1.600E-07

1.500E-07

1.400E-07

Thermal diffusivity (m2 s-1)


1.300E-07

1.200E-07

1.100E-07
Bologna
Pepperoni
1.000E-07 Wieners
Turkey
Ham
9.000E-08

8.000E-08
20 30 40 50 60 70 80 90
Temperature (ºC)

Fig. 4. Thermal diffusivity of meat and poultry emulsions.

Table 5
Correlation matrix for thermophysical properties vs. emulsion composition
k q Cp a Moisture (%) Fat (%) Salt (%) Protein (%) Carbohydrate (%)
k 1.000
q 0.383 1.000
Cp 0.744 0.734 1.000
a 0.959 0.155 0.525 1.000
Moisture (%) 0.960 0.530 0.743 0.899 1.000
Fat (%) 0.978 0.387 0.668 0.957 0.986 1.000
Salt (%) 0.652 0.099 0.488 0.638 0.450 0.503 1.000
Protein (%) 0.776 0.055 0.491 0.799 0.569 0.654 0.927 1.000
Carbohydrate (%) 0.655 0.828 0.972 0.419 0.678 0.580 0.471 0.393 1.000

close to 1 or 1; and a value close to 0 indicates weak ally air) on a volume basis. These products are therefore,
correlation. three phase systems containing air (a), water (w), and solids
From a compositional point of view, the proportion of (s). To predict an effective thermal conductivity of the meat
moisture was inversely correlated to the amount of fat. or poultry emulsions, the applicability of various two
The proportion of salt was also inversely related to the per- phase-based models such as: Series model, Parallel model,
centage of protein in these products. The density and the Kopelman model, Maxwell model, Levy’s model, EMT
heat capacity were strongly influenced by carbohydrate model (Carson et al., 2005) was investigated. None of these
content (i.e. as the carbohydrate content increased, the models was found to be adequate (results not shown).
density decreased). As the moisture content increased, the The thermal conductivities of the food products were
thermal conductivity and diffusivity increased. As the fat calculated using Kirscher’s model (Eqs. (5)–(7)). The distri-
content increased, the value of k and a decreased. Contrary bution factor f was adjusted for each product so that pre-
to the expectations, the proportion of salt did not have any dicted and experimental values matched. The f values
significant effect on the density, heat capacity, thermal con- vary from 0.35 (for pepperoni) to 0.45 (for Bologna and
ductivity and diffusivity. This was probably due to the lim- turkey). These values are much lower than the value of
ited amount of salt (approximately 2%) added to the 0.743 reported for dried apple by Maroulis et al. (2002)
emulsions. The proportion of proteins did not significantly where the samples had very low moisture content. However
influence these properties as correlation coefficients were they are comparable with the values reported by Hamdami
smaller than 0.9. et al. (2003). The value of the distribution factor depends
on the moisture content and decrease with the increase in
3.6. Thermal conductivity modeling moisture content as has been reported by Hamdami et al.
(2003). A comparison of predicted and experimental ther-
Meat and poultry emulsions can be considered to be mal conductivities demonstrated a good agreement. Statis-
porous media containing a significant amount of gas (usu- tical parameters of the comparison are summarized in
322 M. Marcotte et al. / Journal of Food Engineering 88 (2008) 315–322

Table 6 Baik, O.D., Sablani, S.S., Marcotte, M., Castaigne, F., 1999. Modeling the
Comparison of statistical parameters for predicting thermal conductivities thermal properties of a cup cake during baking. Journal of Food
using Kirscher’s model along with the weighting factor (f) Science 64, 295–299.
Food Weighting Coefficient of Standard Relative standard Baik, O.D., Marcotte, M., Sablani, S.S., Castaigne, F., 2001. Thermal and
physical properties of bakery products – a review. Critical Review in
emulsion parameter determination error error (%)
Food Science CRC 41 (5), 321–352.
(f) (r2)
Carson, J.K., 2006. Review of effective thermal conductivity models for
Bologna 0.45 0.916 0.0234 4.89 foods. International Journal of Refrigeration 29, 958–967.
Pepperoni 0.35 0.984 0.0047 1.39 Carson, J.K., Lovatt, S.J., Tanner, D.J., Cleland, A.C., 2005. Thermal
Wieners 0.40 0.986 0.0043 1.24 conductivity bounds for isotropic, porous materials. International
Turkey 0.45 0.993 0.0033 2.10 Journal of Heat and Mass Transfer 48, 2150–2158.
Ham 0.40 0.906 0.0187 3.43 Carson, J.K., Lovatt, S.J., Tanner, D.J., Cleland, A.C., 2006. Predicting
the effective thermal conductivity of unfrozen, porous foods. Journal
of Food Engineering 75, 297–307.
Table 6. Low values of the standard error along with the Gonzo, E.E., 2002. Estimating correlations for the effective thermal
high r2 values for all products indicated that the model per- conductivity of granular materials. Chemical Engineering Journal 90,
299–302.
formed well, where the highest and lowest values of r2 were Hamdami, N., Monteau, J.Y., Le Bail, A., 2003. Effective thermal
0.993 and 0.906 for turkey and ham, respectively. The conductivity of a high porosity model food at above and sub-
relative standard error was less than 5% for all products. freezing temperatures. International Journal of Refrigeration 26,
809–816.
4. Conclusions Karunakar, B., Mishra, S.K., Bandyopadhyay, S., 1998. Specific heat and
thermal conductivity of shrimp meat. Journal of Food Engineering 37,
345–351.
Thermophysical properties are important because of Lind, I., 1991. The measurement and prediction of thermal properties of
their influence on the thermal exchanges between smoke- food during freezing and thawing – a review with particular reference
houses or cookers and meat and poultry emulsions, since to meat and dough. Journal of Food Engineering 13, 285–319.
the long cooking–cooling process relies heavily on conduc- Maroulis, Z.B., Krokida, M.K., Rahman, M.S., 2002. A structural
generic model to predict the effective thermal conductivity of fruits
tion heating. It is also important to identify the movement and vegetables during drying. Journal of Food Engineering 52 (1),
of heat through the product since proper cooking–cooling 47–52.
cycles are generally established using core temperature Mohsenin, N.N., 1980. Thermal Properties of Foods and Agricultural
measurement that must reach a pre-determined legal Materials. Gordon and Breach Science Publishers, New York.
requirement at the end of the process. Results of this paper Pan, Z., Singh, R.P., 2001. Physical and thermal properties of ground beef
during cooking. Food Science and Technology 34, 437–444.
have shown that there are significant differences between Polley, S.L., Synder, 0.P., Kotnour, P., 1980. A compilation of thermal
thermophysical properties of various emulsions. The corre- properties of foods. Food Technology 34 (1l), 76–94.
lation matrix revealed that these differences may be attrib- Ramaswamy, H., Marcotte, M., 2006. Food processing. In: Principles and
uted to the composition of the emulsions. To predict the Applications. Taylor & Francis Publishers, Boca Raton, Florida (pp.
thermal conductivity of the product different models were 62–66).
Sanz, P.D., Alonso, M.D., Mascheroni, R.H., 1987. Thermophysical
applied and it was found that a generic three phase struc- properties of meat products: general bibliography and experimental
ture based model is best suited to predict thermal conduc- values. Transactions of ASAE 30 (l), 283–289.
tivities. Therefore, it is possible to formulate meat and Shmalko, M.E., Morawicki, R.O., Ramallo, L.A., 1996. Simultaneous
poultry emulsions with optimal thermophysical properties determination of specific heat and thermal conductivity using the
in order to maximize efficiency of cooking-cooling cycles. finite-difference method. Journal of Food Engineering 31, 531–
540.
Moreover, these properties can also be used as an input Sweat, V.E., 1974. Experimental values of thermal conductivity of selected
for modeling of heat transfer during the cooking–cooling fruits and vegetables. Journal of Food Science 39, 1080–1083.
process. Sweat, V.E., 1995. Thermal properties of foods. In: Rao, M.A., Rizvi,
S.S.H. (Eds.), Engineering Properties of Foods. Marcel Dekker, NY–
Basel, pp. 99–138.
Acknowledgments Unklesbay, N., Unklesbay, K., Clarke, A.D., 1999. Thermal properties of
restructured beef snack sticks throughout smokehouse processing.
The authors would like to thank the Canadian Meat Food Science and Technology 32, 527–534.
Council for their support. Wang, J.F., Carson, J.K., North, M.F., Cleland, D.J., 2006. A new
approach to the modeling of the effective thermal conductivity of
heterogeneous materials. International Journal of Heat and Mass
References Transfer 49 (17–18), 3075–3083.
Zhang, L., Lyng, J.G., Brunton, N.P., 2007. The effect of fat, water and
Akterian, S.G., 1997. Control strategy using functions of sensitivity for salt on the thermal and dielectric properties of meat batter and its
thermal processing of sausages. Journal of Food Engineering 31, 449– temperature following microwave or radio frequency heating. Journal
455. of Food Engineering 80 (1), 142–151.

You might also like