THERMAL PROPERTIES - DCX
THERMAL PROPERTIES - DCX
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Module 2. Lesson 4
THERMAL PROPERTIES
General Objectives
Specific Objectives
Contents
Title Page
I. Introduction 3
II. Thermal Properties of Food 4
Thermal Properties of Food Constituents 5
Thermal Property Models 5
III. Factors Affecting Thermophysical Properties of Food 6
Water Content 6
Initial Freezing Point 6
Ice Fraction 10
Density 12
IV. Thermal Properties Affecting Thermal Effects 12
Heat Capacity and Specific Heat 15
Specific Heat of Foods 15
Specific Heat of Unfrozen Food 17
Specific Heat of Frozen Food 17
Enthalpy and Constant Pressure Specific Heat 19
Enthalpy of Unfrozen Foods 19
Enthalpy of Frozen Foods 20
Enthalpy and Latent Heat 22
V. Measurement of Specific Heat 23
VI. Fourier’s Law of Heat Conduction 24
VII Thermal Conductivity 25
1|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
2|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Lesson 4
THERMAL PROPERTIES
I. Introduction
Almost every process in the food processing operations involved thermal effects
such as blanching, cooking, pasteurization, and sterilization which are temperature-
dependent biochemical or chemical changes. Thermal properties of particular
importance affecting thermal effects are specific heat, thermal conductivity, thermal
diffusivity, latent heat of phase transition and emissivity (Table 1)
When heat is added to a material (heating), the temperature of that material will
increase so long as it is not undergoing a change in phase. The extent of temperature
rise is governed by the heat capacity of the material. When heat is removed from a
material (cooling), and transferred to a surrounding heat exchange medium at a lower
temperature, the temperature of the material will decrease. (Figura and Tiexiera, 2007)
To lengthen the shelf life of foods, most of the food processing operations involved
heating foods to temperatures capable of inactivating microbial and enzymatic activity.
These heat treatments are based on controlled heat transfer that depends upon thermal
properties of the food materials. In order to increase the internal temperature of a food
product, heat must first be transferred to the outer surface of the food, and then
transmitted through the food material in order to reach the center of the food product.
This is an example of heat transfer.
Understanding thermal properties of materials are very important. An energy
balance for heating or cooling unit operations cannot be made and the temperature
profile within the material cannot be determined without having knowledge of the
thermal properties of the material, (Kutz, 2019).
In food processing, thermal process operations are very important for food
safety. The safety and quality of foods critically depend on correct temperature
regimes; for example, in canning, the classical problem is finding the optimum heating
regime that inactivates any microorganism while still preserving nutritional quality
(avoiding over-processing and destruction of vitamins). In water-containing foods, heat
transfer is often accompanied by a significant water transfer. Thus, the quality and
safety of foods critically depend on the entire temperature history and the state and
distribution of water in the food. Other physical properties and variables such as
pressure, low, electric fields, and water activity also greatly influence processing of
foods. (Rao et. al, 2014)
3|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The information on the thermal behaviors of foods are also needed in food product
characterizations.
Data in thermal properties of foods are required in engineering and process designs
(Table 2). Specifically, knowledge of thermal properties is needed in the design
of food storage and refrigeration equipment, estimation of process times for
refrigerating, freezing, heating, or drying of foods as well as in cooling load
calculations.
Thermal properties of foods and beverages must be known to perform the various
heat transfer calculations involved in designing storage and refrigeration
equipment and estimating process times for refrigerating, freezing, heating, or
drying of foods and beverages. Because the thermal properties of foods and beverages
strongly depend on chemical composition and temperature, and because many types of
food are available, it is nearly impossible to experimentally determine and tabulate the
thermal properties of foods and beverages for all possible conditions and compositions.
However, composition data for foods and beverages are readily available from sources such
as Holland et al. (1991) and USDA (1975). These data consist of the mass fractions of the
major components found in foods. Thermal properties of foods can be predicted by using
these composition data in conjunction with temperature-dependent mathematical models
of thermal properties of the individual food constituents.
4|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Thermophysical properties often required for heat transfer calculations include density,
specific heat, enthalpy, thermal conductivity, and thermal diffusivity. In addition, if the food
is a living organism, such as a fresh fruit or vegetable, it generates heat through respiration
and loses moisture through transpiration. Both of these processes should be included in
heat transfer calculations. Tables of measured thermophysical property data for various
foods and beverages are also provided in this module from ASHRAE 2016.
Constituents commonly found in foods include water, protein, fat, carbohydrate, fiber,
and ash. Choi and Okos (1986) developed mathematical models for predicting the thermal
properties of these components as functions of temperature in the range of –40 to 150°C
(Table 1); they also developed models for predicting the thermal properties of water and
ice (Table 2). Table 3 lists the composition of various foods, including the mass percentage
of moisture, protein, fat, carbohydrate, fiber, and ash (USDA 1996).
Table 2 Thermal Property Models for Water and Ice (−40 t 150°C)
Thermal Property Thermal Property Model
Thermal conductivity, W/(m·K) kw = 5.7109 × 10–1 + 1.7625 × 10–3t – 6.7036 × 10–6t 2
Water Thermal diffusivity, m2/s = 1.3168 × 10–7 + 6.2477 × 10–10t – 2.4022 × 10–12t 2
Density, kg/m3 rw = 9.9718 × 102 + 3.1439 × 10–3t – 3.7574 × 10–3t 2
5|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The initial freezing point of a food is somewhat lower than the freezing point of pure
water because of dissolved substances in the moisture in the food. At the initial freezing
point, some of the water in the food crystallizes, and the remaining solution becomes more
concentrated. Thus, the freezing point of the unfrozen portion of the food is further
reduced. The temperature continues to decrease as separation of ice crystals increases
the concentration of solutes in solution and depresses the freezing point further. Thus, the
ice and water fractions in the frozen food depend on temperature. Because the
thermophysical properties of ice and water are quite different, thermophysical properties
of frozen foods vary dramatically with temperature. In addition, the thermophysical
properties of the food above and below the freezing point are drastically different.
Water Content
Because water is the predominant constituent in most foods, water content significantly
influences the thermophysical properties of foods. Average values of moisture content
(percent by mass) are given in Table 3. For fruits and vegetables, water content varies
with the cultivar as well as with the stage of development or maturity when harvested,
growing conditions, and amount of moisture lost after harvest. In general, values given in
Table 3 apply to mature products shortly after harvest. For fresh meat, the water content
values in Table 3 are at the time of slaughter or after the usual aging period. For cured or
processed products, the water content depends on the particular process or product.
6|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
they freeze, a food’s initial freezing point must be known to model its thermophysical
properties accurately. Experimentally determined values of the initial freezing point of foods
and beverages are given in Table 3.
Table 3 Unfrozen Composition Data, Initial Freezing Point, and Specific Heats of Foods*
7|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
8|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
9|Page
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Ice Fraction
( ) (1)
=
10 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
where:
xs = mass fraction of solids in food
Ms = relative molecular mass of soluble solids, kg/kmol
R = universal gas constant = 8.314 kJ/(kg mol·K)
To = freezing point of water = 273.2 K
Lo = latent heat of fusion of water at 273.2 K = 333.6 kJ/kg
tf = initial freezing point of food, °C
t = food temperature, °C
The relative molecular mass of the soluble solids in the food may be estimated as follows:
where xwo is the mass fraction of water in the unfrozen food and xb is the mass fraction of
bound water in the food (Schwartzberg 1976). Bound water is the portion of water in a
food that is bound to solids in the food, and thus is unavailable for freezing.
where xp is the mass fraction of protein in the food. Substituting Equation (2) into Equation
(1) yields a simple way to predict the ice fraction (Miles 1974):
Because Equation (4) underestimates the ice fraction at temperatures near the initial
freezing point and overestimates the ice fraction at lower temperatures, Tchigeov (1979)
proposed an empirical relationship to estimate the mass fraction of ice:
Fikiin (1996) notes that Equation (5) applies to a wide variety of foods and provides
satisfactory accuracy.
Example 1. A 150 kg beef carcass is to be frozen to –20°C. What are the masses of the
frozen and unfrozen water at –20°C?
Solution:
From Table 3, the mass fraction of water in the beef carcass is 0.58 and the initial
freezing point for the beef carcass is –1.7°C. Using Equation (5), the mass fraction of ice
is
11 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Density
Modeling the density of foods and beverages requires knowledge of the food porosity,
as well as the mass fraction and density of the food components. The density r of foods
and beverages can be calculated accordingly:
where ε is the porosity, xi is the mass fraction of the food constituents, and ρi is the
density of the food constituents. The porosity ε is required to model the density of
granular foods stored in bulk, such as grains and rice. For other foods, the porosity is
zero.
Different substances respond to heat in different ways. If a metal chair sits in the bright
sun on a hot day, it may become quite hot to the touch. An equal mass of water in the
same sun will not become nearly as hot. We would say that water has a high heat
capacity (the amount of heat required to raise the temperature of an object by 1oC. Water
is very resistant to changes in temperature, while metals in general are not. The specific
heat of a substance is the amount of energy required to raise the temperature of 1 gram
of the substance by 1oC.
=
where is the energy required to produce a temperature change. Ordinarily, heat
capacity is specified per mole of material (e.g, J/mol-K, or cal/mol-K).
When heat capacity is defined only in this way, it will also depend upon the mass of
the material sample, and serves as a property only of the specific sample size measured.
For this reason, we normally measure and report the heat capacity on the basis of a
12 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
common unit of mass. When we do this, we call it the specific heat capacity. Sometimes
this property is called specific heat of the material but this should be avoided because
dQ/dm = q is specific heat.
In order to help better understand heat capacity, let us assume we wish to determine
how much heat is needed to raise the temperature of one liter of water at 21oC up to 23oC
(by 2 K). When we do this, we measure the heat required to be 8.36 kJ.
When heat is added to a system like this (liter of water), the water molecules experience
an increase in their kinetic energy. They move in both rotational and translational motion
at faster rates. If we insert our finger (or a thermometer) into this liter of water, we can
sense this increased thermal energy level by a warming sensation on our finger, and a rise
in the temperature scale on the thermometer. Therefore, we use temperature as a measure
of increased thermal energy. In this case, the temperature increase was ΔT = 2K.
= =
= =
= =
13 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Specific heat capacity at constant pressure, Cp, and its temperature integral, enthalpy,
H, are the fundamental properties for heat balance calculations. Water has a much higher
specific heat and thermal conductivity than the other major food constituents. Therefore,
water greatly influences the thermal properties of foods, albeit not as dramatically in
unfrozen food compared with frozen foods, where the water–ice phase change dominates
Calorimetry and more conveniently, differential scanning calorimetry (DSC) are the usual
techniques for measuring the specific heat capacity and phase transitions. Table 7.3 shows
representative data for various foods. (Rao et al, 2014)
The specific heat capacity of the various components of the food can be calculated from
empirical equations
cp = a + bxw
where xw is the mass fraction of water in the food, and a and b are empirical constants
specific to the food. Riedel reined this by introducing temperature de dependence and
introducing a nonlinear term improving predictions at low water contents:
.
= + 4.19 ((α + 0.001 t) (1 – ) – βexp (–43 )
Riedel determined the constants α and β for eight different foods. However, if mean
values of α = 0.37 and β = 0.9 are taken, the maximum deviation of Cp from the individual
14 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
eight equations does not exceed 10%, and food engineers may find this approximation
useful for some calculations.
As the specific heat capacity is an additive property, one can use the summation
formula to predict the specific heat capacity of a food with known composition:
where Cp is the specific heat capacity of the mixture, the values of xi are the mass fractions
of the various constituents, and the values of Cpi are the specific heat capacities of the
constituents. (Rao et. al. 2014)
Often in thermodynamics it is necessary to distinguish between internal energy terms
in which displacement work is present or not present. When there is no displacement work,
the volume of the system remains constant (V = const., dQ = dU), and the subscript V is
used with the heat capacity term. When displacement is present, then the pressure of the
system remains constant (p = const., dQ = dH), and the subscript p is used with the heat
capacity term (see Table 2).
Key Points
Heat capacity is the ratio of the amount of heat energy transferred to an object to
the resulting increase in its temperature.
Molar heat capacity is a measure of the amount of heat necessary to raise the
temperature of one mole of a pure substance by one degree K.
Specific heat capacity is a measure of the amount of heat necessary to raise the
temperature of one gram of a pure substance by one degree K.
Further Readings & video:
http://www.differencebetween.net/science/difference-between-specific-heat-and-heat-
capacity/
https://youtu.be/TqJFIBODrjM
https://youtu.be/yhNHJ7WdT8A
Specific heat is a measure of the energy required to change the temperature of a food
by one degree. Therefore, the specific heat of foods or beverages can be used to
calculate the heat load imposed on the refrigeration equipment by the cooling
or freezing of foods and beverages. In unfrozen foods, specific heat becomes slightly
lower as the temperature rises from 0°C to 20°C. For frozen foods, there is a large decrease
in specific heat as the temperature decreases. Table 3 lists experimentally determined
values of the specific heats for various foods above and below freezing.
Specific heat is defined as the amount of heat required to increase the temperature
of a unit mass of a material by a unit or one degree. Accordingly, its unit is J/kg-K in the
SI system or Btu/lbm- F in English system. The specific heat depends on the nature of the
process of heat addition in terms of either a constant pressure process (Cp) or a
constant volume process (Cv). However, because specific heats of solids and liquids do
not depend on pressure much, except extremely high pressures, and because pressure
changes in heat transfer problems of food materials are usually negligible, the specific heat
15 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
at constant pressure is considered (Mohsenin, 1980). Like the thermal conductivity, the
specific heats of foodstuffs are greatly dependent on their composition. Knowing
the specific heat of each component of a mixture is usually sufficient to predict the
specific heat of the matrix (Sweat, 1995). For instance, the specific heat of high-
moisture foods is largely dominated by water content. Specific heat data for different food
materials below and above freezing points were reported by Rahman (1995).
The temperature dependence of specific heat of major food components has also been
studied. The specific heat of pure water, carbohydrate (CHO), protein, fat, ash, and ice
at different temperatures can be expressed empirically in J/kgoC according to Choi and
Okos (1986) as indicated below:
It is generally true that experimentally determined specific heat is higher than the
predicted value. The difference might be due to the presence of bound water, variation of
specific heat of the component phases with the source, and interaction of the component
phases (Rahman, 1995).
Where X represents the mass fraction of each of the component groups (Rahman, 1995).
For mixtures that approximate solutions of sugar in water (e.g. fruit juices), Eq. (1.3)
becomes:
16 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Another frequently used model assigns to the total dry matter of the mixture a single
relative specific value of 0.837. The resulting approximate empirical expressions for
temperatures above and below freezing are given in Eq. (1.5):
The specific heat of a food, at temperatures above its initial freezing point, can be
obtained from the mass average of the specific heats of the food components. Thus, the
specific heat of an unfrozen food cu may be determined as follows:
cu = ∑ci xi (7)
Where ci is the specific heat of the individual food components and xi is the mass fraction
of the food components.
A simpler model for the specific heat of an unfrozen food is presented by Chen (1985).
If detailed composition data are not available, the following expression for specific heat of
an unfrozen food can be used:
Where cu is the specific heat of the unfrozen food in kJ/(kg·K) and xs is the mass fraction
of the solids in the food.
Below the food’s freezing point, the sensible heat from temperature change and the
latent heat from the fusion of water must be considered. Because latent heat is not released
at a constant temperature, but rather over a range of temperatures, an apparent specific
heat must be used to account for both the sensible and latent heat effects. A common
method to predict the apparent specific heat of foods is:
17 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The specific heat of the food above the freezing point may be estimated with Equation
(7) or (8).
Where
cf = specific heat of fully frozen food (typically at –40°C)
to = freezing point of water = 0°C
tf = initial freezing point of food, °C
t = food temperature, °C
Lo = latent heat of fusion of water = 333.6 kJ/kg
Experimentally determined values of the specific heat of fully frozen foods are given in
Table 3.
A slightly simpler apparent specific heat model, which is similar in form to that of
Schwartzberg (1976), was developed by Chen (1985). Chen’s model is an expansion of
Siebel’s equation (Siebel 1892) for specific heat and has the following form:
where
ca = apparent specific heat, kJ/(kg·K)
xs = mass fraction of solids
R = universal gas constant
To = freezing point of water = 273.2 K
Ms = relative molecular mass of soluble solids in food
t = food temperature, °C
If the relative molecular mass of the soluble solids is unknown, Equation (2) may be
used to estimate the molecular mass. Substituting Equation (2) into Equation (11) yields
Example 2. One hundred fifty kilograms of lamb meat is to be cooled from 10°C to 0°C.
Using the specific heat, determine the amount of heat that must be removed from the
lamb.
Solution:
From Table 3, the composition of lamb is given as follows:
xwo = 0.7342 xf = 0.0525
xp = 0.2029 xa = 0.0106
cw = 4.1762 – 9.0864 × 10–5(5) + 5.4731 × 10–6(5)2
= 4.1759 kJ/(kg·K)
cp = 2.0082 + 1.2089 × 10–3(5) – 1.3129 × 10–6(5)2
= 2.0142 kJ/(kg·K)
18 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The change in a food’s enthalpy can be used to estimate the energy that must be added
or removed to effect a temperature change. Above the freezing point, enthalpy consists of
sensible energy; below the freezing point, enthalpy consists of both sensible and latent
energy. Enthalpy may be obtained from the definition of constant-pressure specific heat:
where Hi is the enthalpy of the individual food components and xi is the mass fraction of
the food components.
In Chen’s (1985) method, the enthalpy of an unfrozen food may be obtained by
integrating Equation (8):
where
H = enthalpy of food, kJ/kg
Hf = enthalpy of food at initial freezing temperature, kJ/kg
t = temperature of food, °C
tf = initial freezing temperature of food, °C
xs = mass fraction of food solids
19 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The enthalpy at initial freezing point Hf may be estimated by evaluating either Equation
(17) or (18) at the initial freezing temperature of the food, as discussed in the following
section.
Teacher’s Note:
For additional exercises and information. Please refer to this book:
Physical Properties of Food by Sahin and Sumnu pp. 107-151)
For foods below the initial freezing point, mathematical expressions for enthalpy may
be obtained by integrating the apparent specific heat models. Integration of Equation (9)
between a reference temperature Tr and food temperature T leads to the following
expression for the enthalpy of a food (Schwartzberg 1976):
Where
H = enthalpy of food
R = universal gas constant
To = freezing point of water = 273.2 K
Substituting Equation (2) for the relative molecular mass of the soluble solids Ms
simplifies Chen’s method as follows:
where
H = enthalpy of food, kJ/kg
20 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Fruit/Vegetable Group:
Juice Group:
In addition, the enthalpy of the food at its initial freezing point is required in Equation
(19). Chang and Tao (1981) suggest the following correlation for determining the food’s
enthalpy at its initial freezing point Hf :
Table 4 presents experimentally determined values for the enthalpy of some frozen
foods at a reference temperature of –40°C as well as the percentage of unfrozen water in
these foods.
21 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The enthalpy of the beef carcass at –20°C is given by Equation (18) for frozen foods:
The enthalpy of the beef carcass at the initial freezing point is determine by
evaluating Equation (18) at the initial freezing point:
( . . )( . )( . )
= Hf [−1.7 − (−40] 1.55 + (1.26)(0.41790) − ( )( . )
= 243.14kJ/kg
The enthalpy of the beef carcass at 10°C is given by Equation (15) for unfrozen
foods:
H10 = 243.14 + [10 – (–1.7)] × [4.19 – (2.30)(0.4179) – (0.628)(0.4179)3 – ]
= 280.38 kJ/kg
Thus, the amount of heat removed during the freezing process is
Enthalpy and latent heat are the two other thermal properties of great importance in
characterizing food products and processing systems. Enthalpy is the heat content in a sys-
tem per unit mass and hence the unit is J/kg in the SI system. It is a thermodynamic
property that depends only on the state of the system and it is expressed in terms of
internal energy, pressure, and volume as:
where H is the total heat content (enthalpy), U is internal energy, and PV is pressure and
volume.
Latent heat, on the other hand, is the heat released or absorbed by a chemical
substance or a thermodynamic system during a change of state that occurs without a
change in temperature. Latent heat is a hidden energy as its impact on the temperature of
the system is not observed. The effect of latent heat on the system is noted only as a
phase change such as the melting of ice or the boiling of water and their reverse processes.
22 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
23 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Detailed Descriptions
1. Method of Mixture
2. Guarded Plate
3. Comparison Calorimeter,
4. Adiabatic Agricultural Calorimeter
5. Differential Scanning Calorimeter (DSC),
6. Calculated Specific Heat
(3.1)
Temperature T0. At time t = 0, one side of the wall is suddenly brought to a slightly
higher temperature T1 and maintained at that temperature. Heat is conducted through the
wall as a result of the temperature difference, and as time proceeds, the temperature
profile in the wall changes. Finally, linear steady-state temperature distribution is achieved
as shown in Fig. 3.1.
The driving force for the heat transfer to occur is the temperature difference:
While the rate of heat conduction through the wall is proportional to the heat transfer
area (A), the thickness of the wall (X) provides resistance to heat transfer. In addition, the
ability of the wall material to conduct heat should be considered. Each material has a
different ability to conduct heat.
24 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The responses of steel and wood to heating are not the same when they are exposed
to the same amount of heat. This material property is named thermal conductivity (k).
Considering all these parameters, the resistance to heat transfer can be written as:
(3.3)
When the steady-state condition has been reached, the rate of heat flow (Q) through
the wall can be written by substituting Eqs. (3.2) and (3.3) into Eq. (3.1):
(3.4)
(3.5)
where Qx is the rate of heat flow in the x-direction. Heat is conducted in the direction
of decreasing temperature and the temperature gradient becomes negative when
temperature decreases with increasing x. Therefore, a negative sign is added to Eq.
(3.5).
= +
where is total thermal conductivity; is free electron; and is lattice wave component.
In pure metals, heat conduction is based mainly on the flow of free electrons and the
effect of lattice vibrations is negligible. In alloys and nonmetallic solids, which have few
free electrons, heat conduction between molecules is due to lattice vibrations. Therefore
metals have higher thermal conductivities than alloys and also than nonmetallic solids.
The regularity of the lattice arrangement has an important effect on the lattice
component of thermal conductivity. For example, diamond has very high thermal
conductivity because of its well-ordered structure. On the other hand, in porous solids such
as foods and other agricultural materials, thermal conductivity depends mostly on
composition but also on many factors that affect the heat flow paths through the material.
Some of the limiting factors include void fraction, shape, size, and arrangement of void
spaces; the fluid contained in the pores; and homogeneity (Sweat, 1995).
25 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Thermal conductivity in foods having fibrous structures such as meat cannot be the
same in different directions (anisotropy) because heat flow paths through the material
change with respect to direction. Thermal conductivity is known to increase with
moisture content. Thermal conductivities of food materials vary between that of water
(kwater = 0.614 W/moC at 27oC) and that of air (kair=0.026 W/moC at 27oC), which in many
food matrices are the most and the least conductive components, respectively. Thermal
conductivity of ice is nearly four times greater than that of water (kice= 2.24 W/moC at 0oC).
This partly accounts for the difference in freezing and thawing rates of food materials.
Thermal conductivity of food matrices can be predicted using mathematical models that
take the chemical compositions into account. Predictive models have been used to estimate
the effective thermal conductivity of foods. It is important to include the effect of air in the
porous foods and ice in the case of frozen ones. Temperature dependence of thermal
conductivities of major food components has been studied. Thermal conductivities of pure
water, carbohydrate (CHO), protein, fat, ash, and ice at different temperatures can be
empirically expressed according to Choi and Okos (1986) as summarized below:
where thermal conductivities (k) are in W/moC; temperature (T) is in oC and varies between
0o C and 90o C in these equations.
The table below shows representative value of thermal conductivity. Source: Rao et al,
2014)
26 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Thermal conductivity relates the conduction heat transfer rate to the temperature
gradient. A food’s thermal conductivity depends on factors such as composition, structure,
and temperature. Early work in the modeling of thermal conductivity of foods and
beverages includes Eucken’s adaption of Maxwell’s equation (Eucken 1940).
This model is based on the thermal conductivity of dilute dispersions of small spheres
in a continuous phase:
where
k = conductivity of mixture
kc = conductivity of continuous phase
kd = conductivity of dispersed phase
a = 3kc /(2kc + kd)
b = Vd /(Vc + Vd)
Vd = volume of dispersed phase
Vc = volume of continuous phase
In an effort to account for the different structural features of foods, Kopelman (1966)
developed thermal conductivity models for homogeneous and fibrous foods.
Differences in thermal conductivity parallel and perpendicular to the food fibers are
accounted for in Kopelman’s fibrous food thermal conductivity models. For an isotropic,
two-component system composed of continuous and discontinuous phases, in which
thermal conductivity is independent of direction of heat flow, Kopelman (1966)
developed the following expression for thermal conductivity k:
where kc is the thermal conductivity of the continuous phase and L3 is the volume fraction
of the discontinuous phase. In Equation (27), thermal conductivity of the continuous phase
is assumed to be much larger than that of the discontinuous phase. However, if the
opposite if true, the following expression is used to calculate the thermal conductivity of
the isotropic mixture:
where M = L2(1 – kd /kc) and kd is the thermal conductivity of the discontinuous phase.
For an anisotropic, two-component system in which thermal conductivity depends on the
direction of heat flow, such as in fibrous food materials, Kopelman (1966) developed two
expressions for thermal conductivity. For heat flow parallel to food fibers, thermal
conductivity k= is
27 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
where N2 is the volume fraction of the discontinuous phase. If the heat flow is perpendicular
to the food fibers, then thermal conductivity is:
where P = N(1 – kd /kc). Levy (1981) introduced a modified version of the Maxwell- Eucken
equation. Levy’s expression for the thermal conductivity of a two-component system is as
follows:
where Λ is the thermal conductivity ratio (Λ = k1/k2), and k1 and k2 are the thermal
conductivities of components 1 and 2, respectively. The parameter F1 introduced by Levy
is given as follows:
where:
28 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Numerous researchers have proposed using parallel and perpendicular (or series)
thermal conductivity models based on analogies with electrical resistance. The parallel
model is the sum of the thermal conductivities of the food constituent multiplied by their
volume fractions:
where is the volume fraction of constituent i. The volume fraction of constituent i can
be found from the following equation:
The perpendicular model is the reciprocal of the sum of the volume fractions divided
by their thermal conductivities:
These two models have been found to predict the upper and lower bounds of the
thermal conductivity of most foods. Tables 5 and 6 list the thermal conductivities for many
foods (Qashou et al. 1972). Data in these tables have been averaged, interpolated,
extrapolated, selected, or rounded off from the original research data. Tables 5 and 6 also
include ASHRAE research data on foods of low and intermediate moisture content (Sweat
1985).
Example 4. Determine the thermal conductivity and density of lean pork shoulder meat
at -40°C. Use both the parallel and perpendicular thermal conductivity models.
Solution:
From Table 3, the composition of lean pork shoulder meat is:
xwo = 0.7263 xf = 0.0714
xp = 0.1955 xa = 0.0102
In addition, the initial freezing point of lean pork shoulder meat is –2.2°C. Because the
pork’s temperature is below the initial freezing point, the fraction of ice in the pork must
be determined. Using Equation (4), the ice fraction becomes
=( − ) 1− = −.4 1−
.
= [0.7263 − (0.40)(0.1955] 1 −
= 0.6125
The mass fraction of unfrozen water is then,
=( − ) = 0.7263 – 0.6125 = 0.1138
29 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Using the equations in Tables 1 and 2, the density and thermal conductivity of the food
constituents are calculated at the given temperature –40°C:
r = 9.9718 x 10 + 3.1439 x 10 (-40)-3.7574 x 10 (-40)2
= 991.04 kg/m3
r = 9.1689 x 102- 1.3071 x 10-1(-40)
= 922.12 kg/m3
r = 1.3299 x 103- 5.1840 x 10-1(-40)
= 1350.6 kg/m3
r = 9.2559 x 102 – 4.1757 x 10-1x (-40)
= 942.29 kg/m3
r = 2.4238 x 103 – 2.8063 x 10-1(-40)
= 2435.0 kg/m3
= 5.7109 × 10–1 + 1.7625 × 10–3(–40) – 6.7036 × 10–6(–40)2
= 0.4899 W/(m ·K)
= 2.2196 – 6.2489 × 10–3(–40) + 1.0154 × 10–4(–40)2
= 2.632 W/(m ·K)
= 1.7881 × 10–1 + 1.1958 × 10–3(–40) – 2.7178 × 10–6(–40)2
= 0.1266 W/(m ·K)
= 1.8071 × 10–1 – 2.7604 × 10–3(–40) – 1.7749 × 10–7(–40)2
= 0.2908 W/(m ·K)
= 3.2962 × 10–1 + 1.4011 × 10–3(–40) – 2.9069 × 10–6(–40)2
= 0.2689 W/(m ·K)
Using Equation (6), the density of lean pork shoulder meat at –40°C can be determined:
Using Equation (36), the volume fractions of the constituents can be found
30 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Using the parallel model, Equation (35), the thermal conductivity becomes
Using the perpendicular model, Equation (37), the thermal conductivity becomes
Example 5. Determine the thermal conductivity and density of lean pork shoulder meat
at a temperature of –40°C. Use the isotropic model developed by Kopelman (1966).
Solution:
From Table 3, the composition of lean pork shoulder meat is
xwo = 0.7263 xf = 0.0714
xp = 0.1955 xa = 0.0102
In addition, the initial freezing point of lean pork shoulder is –2.2°C. Because the pork’s
temperature is below the initial freezing point, the fraction of ice within the pork must be
determined. From Example 4, the ice fraction was found to be
xice = 0.6125
The mass fraction of unfrozen water is then
xw = xwo – xice = 0.7263 – 0.6125 = 0.1138
31 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
32 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
33 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
34 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
35 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Using the equations in Tables 1 and 2, the density and thermal conductivity of the food
constituents are c1alculated at the given temperature, –40°C (refer to Example 4):
ρw = 991.04 kg/m3 kw = 0.4899 W/(m·K)
ρice= 922.12 kg/m3 kice = 2.632 W/(m·K)
ρp = 1350.6 kg/m3 kp = 0.1266 W/(m·K)
ρf = 942.29 kg/m3 kf = 0.2908 W/(m·K)
ρa = 2435.0 kg/m3 ka = 0.2689 W/(m·K)
Now, determine the thermal conductivity of the ice/water mixture. This requires the
volume fractions of the ice and water:
36 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Note that the volume fractions calculated for the two-component ice/water mixture are
different from those calculated in Example 4 for lean pork shoulder meat. Because the ice
has the largest volume fraction in the two-component ice/water mixture, consider the ice
to be the “continuous” phase. Then, L from Equation (27) becomes
Because kice > kw and the ice is the continuous phase, the thermal conductivity of the
ice/water mixture is calculated using Equation (27)
Next, find the thermal conductivity of the ice/water/protein mixture. This requires the
volume fractions of the ice/water and the protein:
Note that these volume fractions are calculated based on a two component system
composed of ice/water as one constituent and protein as the other. Because protein has
the smaller volume fraction, consider it to be the discontinuous phase.
37 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Next, find the thermal conductivity of the ice/water/protein/fat mixture. This requires
the volume fractions of the ice/water/protein and the fat:
Finally, the thermal conductivity of the lean pork shoulder meat can be found. This
requires the volume fractions of the ice/water/protein/fat and the ash:
Thus, the thermal conductivity of the lean pork shoulder meat becomes
38 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
39 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
measurements can be made in situ. This makes the method suitable for use in industry,
especially if the measurement is automated so that non-specialists can use it. For small
temperature rises of the heated probe, convection currents do not arise in fluid samples.
This makes the method suitable for liquid foods.
Detailed Description
1. Steady State Methods
a. Description of steady state method
b. Steady State Methods
Longitudinal Heat Flow Method (guarded hot plate method)
Radial Heat Flow Methods (Concentric Cylinder Method; Concentric Cylinder
Comparative Method; Sphere with Central Heating Source
Heat of Vaporization Method
Heat Flux Method
Differential Scanning Calorimeter (DSC
2. Transient State Methods
a. Description of transient state method
b. transient state method
thermal conductivity probe method,
transient hot wire method,
modified Fitch method,
point heat source method, and
Comparative method.
40 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
a material accepts or gives away heat. The main idea behind thermal diffusivity is the rate
at which heat diffuses throughout a material.
=
r
Where k is the thermal conductivity, r is the density, and c is the specific heat capacity at
constant pressure, r is referred to as the volumetric heat capacity.
ủ= u
Where u is a measure of some property, ủ is its derivative with respect to time, and is
its Laplace operator (the divergence of the gradient).
In the case of heat transfer through a homogeneous (uniform) body, u could represent
temperature and α would be the same as above.
=
One benefit of this equation is that can often be written independently of any coordinate
system. In this form it is clear to see that thermal diffusivity is a scaling factor, meaning it
directly controls the speed at which temperature changes.
https://thermtest.com/thermal-diffusivity-overview
For transient heat transfer, thermal diffusivity α, which appears in the Fourier equation:
41 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
This is the same as the thermal diffusivity for solids, where is thermal diffusivity
2
(m /s) in SI derived units, k is thermal conductivity [W/(m·K)] , ρ is density [kg/m³], and
c is specific heat [J/(kg·K)] .
Thermal diffusivity of selected foodstuff is shown in the next table.
42 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Equation (3.121) can be used to calculate thermal diffusivity by trial and error. To
ensure convergence, the first 40 terms of equation have to be evaluated if the set of
thermocouples were placed in such location that 0.16 < β < 3.1 over the time interval used
(Nix et al., 1967). This method is suitable for biological materials since the test duration is
short (about 300 s) and the temperature change imposed on the sample is small. The exact
distance r of the two thermocouples is critical. Therefore, a calibration of the system may
be required.
43 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Dickerson Method
Dickerson (1965) described an apparatus to measure thermal conductivity. The
cylindrical container of radius R with high thermal conductivity is filled with sample and
placed in a constant-temperature agitated water bath. Both ends of the container are
insulated with rubber corks so that only a radial temperature gradient exists. The
temperatures at the surface and center of the cylinder were monitored with thermocouples.
The time–temperature data are recorded until constant rate of temperature rise is obtained
for both the inner and outer thermocouples. Then, the Fourier equation is given as:
Many industries rely on thermal diffusivity to determine the most suitable materials to
optimize efficient heat flow. Insulation is an example of a material that requires a low
thermal diffusivity so that a minimal amount of heat is passing through it at any one time.
A heat sink is an appliance that is designed to carry the heat out and away from another
piece of equipment. A heat sink is required to have a very high thermal diffusivity that
enables the quick transport of the heat. If slow heat transfer were to occur, the area
accepting the heat would heat up and not permit as much heat flow per unit time. Heat
sinks are used in almost every piece of electrical equipment. An increase in temperature in
certain components can lead to an increased electrical resistance and unexpected behavior.
44 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
45 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
46 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
All living foods respire. During respiration, sugar and oxygen combine to form CO2,
H2O, and heat as follows:
In most stored plant products, little cell development takes place, and the greater part
of respiration energy is released as heat, which must be taken into account when cooling
and storing these living commodities (Becker et al. 1996). The rate at which this chemical
reaction takes place varies with the type and temperature of the commodity.
Becker et al. (1996) developed correlations that relate a commodity’s rate of carbon
dioxide production to its temperature. The carbon dioxide production rate can then be
related to the commodity’s heat generation rate from respiration. The resulting correlation
gives the commodity’s respiratory heat generation rate W in W/kg as a function of
temperature t in °C: (41)
.
= + 32
The respiration coefficients f and g for various commodities are given in Table 8.
Fruits, vegetables, flowers, bulbs, florists’ greens, and nursery stock are storage
commodities with significant heats of respiration.
Dry plant products, such as seeds and nuts, have very low respiration rates. Young,
actively growing tissues, such as asparagus, broccoli, and spinach, have high rates of
respiration, as do immature seeds such as green peas and sweet corn. Fast-developing
fruits, such as strawberries, raspberries, and blackberries, have much higher respiration
rates than do fruits that are slow to develop, such as apples, grapes, and citrus fruits.
In general, most vegetables, other than root crops, have a high initial respiration rate
for the first one or two days after harvest. Within a few days, the respiration rate quickly
lowers to the equilibrium rate (Ryall and Lipton 1972).
47 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Fruits that do not ripen during storage, such as citrus fruits and grapes, have fairly
constant rates of respiration. Those that ripen in storage, such as apples, peaches, and
avocados, increase in respiration rate. At low storage temperatures, around 0°C, the rate
of respiration rarely increases because no ripening takes place. However, if fruits are stored
at higher temperatures (10°C to 15°C), the respiration rate increases because of ripening
and then decreases. Soft fruits, such as blueberries, figs, and strawberries, decrease in
respiration with time at 0°C. If they become infected with decay organisms, however,
respiration increases.
Table 9 lists the heats of respiration as a function of temperature for a variety of
commodities, and Table 10 shows the change in respiration rate with time. Most
commodities in Table 9 have a low and a high value for heat of respiration at each
temperature. When no range is given, the value is an average for the specified temperature
and may be an average of the respiration rates for many days.
When using Table 9, select the lower value for estimating the heat of respiration at
equilibrium storage, and use the higher value for calculating the heat load for the first day
or two after harvest, including precooling and short-distance transport. In storage of fruits
between 0°C and 5°C, the increase in respiration rate caused by ripening is slight. However,
for fruits such as mangoes, avocados, or bananas, significant ripening occurs at
temperatures above 10°C and the higher rates listed in Table 9 should be used. Vegetables
such as onions, garlic, and cabbage can increase heat production after a long storage
period.
The most abundant constituent in fresh fruits and vegetables is water, which exists as
a continuous liquid phase in the fruit or vegetable. Some of this water is lost through
transpiration, which the transport of moisture through the skin, evaporation, and
convective mass transport of the moisture to the surroundings
(Becker et al. 1996b).
The rate of transpiration in fresh fruits and vegetables affects product quality. Moisture
transpires continuously from commodities during handling and storage. Some moisture loss
is inevitable and can be tolerated. However, under many conditions, enough moisture may
be lost to cause shriveling. The resulting loss in mass not only affects appearance, texture,
and flavor of the commodity, but also reduces the salable mass (Becker et al. 1996a).
Many factors affect the rate of transpiration from fresh fruits and vegetables. Moisture
loss is driven by a difference in water vapor pressure between the product surface and the
environment. Becker and Fricke (1996a) state that the product surface may be assumed
to be saturated, and thus the water vapor pressure at the commodity surface is equal to
the water vapor saturation pressure evaluated at the product’s surface temperature.
However, they also report that dissolved substances in the moisture of the commodity tend
to lower the vapor pressure at the evaporating surface slightly.
48 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
Evaporation at the product surface is an endothermic process that cools the surface,
thus lowering the vapor pressure at the surface and reducing transpiration. Respiration
within the fruit or vegetable, on the other hand, tends to increase the product’s
temperature, thus raising the vapor pressure at the surface and increasing transpiration.
Furthermore, the respiration rate is itself a function of the commodity’s temperature
(Gaffney et al. 1985). In addition, factors such as surface structure, skin permeability, and
airflow also effect the transpiration rate (Sastry et al. 1978).
Becker et al. (1996c) performed a numerical, parametric study to investigate the
influence of bulk mass, airflow rate, skin mass transfer coefficient, and relative humidity
on the cooling time and moisture loss of a bulk load of apples. They found that relative
humidity and skin mass transfer coefficient had little effect on cooling time, whereas bulk
mass and airflow rate were of primary importance.
Moisture loss varied appreciably with relative humidity, airflow rate, and skin mass
transfer coefficient; bulk mass had little effect. Increased airflow resulted in a decrease in
moisture loss; increased airflow reduces cooling time, which quickly reduces the vapor
pressure deficit, thus lowering the transpiration rate.
The driving force for transpiration is a difference in water vapor pressure between the
surface of a commodity and the surrounding air. Thus, the basic form of the transpiration
model is as follows:
̇ = ( − ) (42)
where ̇ is the transpiration rate expressed as the mass of moisture transpired per unit
area of commodity surface per unit time. This rate may also be expressed per unit mass of
commodity rather than per unit area of commodity surface. The transpiration coefficient kt
is the mass of moisture transpired per unit area of commodity, per unit water vapor
pressure deficit, per unit time. It may also be expressed per unit mass of commodity rather
than per unit area of commodity surface. The quantity ( − )is the water vapor pressure
deficit. The water vapor pressure at the commodity surface is the water vapor saturation
pressure evaluated at the commodity surface temperature; the water vapor pressure in the
surrounding air pa is a function of the relative humidity of the air.
49 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
where is the air film mass transfer coefficient and is the skin mass transfer coefficient.
The variable describes the convective mass transfer that occurs at the surface of the
commodity and is a function of airflow rate. The variable ks describes the skin’s diffusional
resistance to moisture migration.
The air film mass transfer coefficient ka can be estimated by using the Sherwood-Reynolds-
Schmidt correlations (Becker et al. 1996b). The Sherwood number is defined as follows:
where is the air film mass transfer coefficient, d is the commodity’s diameter, and δ is
the commodity coefficient of diffusion of water vapor in air. For convective mass transfer
from a spherical fruit or vegetable, Becker and Fricke (1996b) recommend using the
following Sherwood-Reynolds-Schmidt correlation, which was taken from Geankoplis
(1978):
Re is the Reynolds number ( = ) and Sc is the Schmidt number (Sc = ν/δ), where u is
the free stream air velocity and is the kinematic viscosity of air. The driving force for is
concentration. However, the driving force in the transpiration model is vapor pressure.
Thus, the following conversion from concentration to vapor pressure is required:
where is the gas constant for water vapor and T is the absolute mean temperature of
the boundary layer. The skin mass transfer coefficient ks, which describes the resistance to
moisture migration through the skin of a commodity, is based on the fraction of the product
surface covered by pores. Although it is difficult to theoretically determine the skin mass
transfer coefficient, experimental determination has been performed by Chau et al. (1987)
and Gan and Woods (1989).
(Heat of respiration for fresh fruits and vegetables at various temperatures are readily
available in ASHRAE Handbook any edition, refer to Thermal properties as well as change
of respiration rates with time including transpiration coefficients)
Although the surface heat transfer coefficient is not a thermal property of a food or
beverage, it is needed to design heat transfer equipment for processing foods and
beverages where convection is involved. Newton’s law of cooling defines the surface heat
transfer coefficient h as follows:
where is the heat transfer rate, is the surface temperature of the food, is the
surrounding fluid temperature, and is the surface area of the food through which the
heat transfer occurs.
50 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
The surface heat transfer coefficient h depends on the velocity of the surrounding fluid,
product geometry, orientation, surface roughness, and packaging, as well as other factors.
Therefore, for most applications ℎ must be determined experimentally. Researchers have
generally reported their findings as correlations, which give the Nusselt number as a
function of the Reynolds number and the Prandtl number. (Recall previous lesson in
Heat Transfer)
Problems: (Source: Physical Properties of Food by Sahin and Sumnu pp151 – 152.
1. The thermal conductivity of a Royal Gala apple is measured at 25◦C by guarded hot
plate method. The apple samples are cut into chips with area of 305 mm × 305 mm
and thickness of 15 mm. The temperature difference between the hot and cold surfaces
is kept at 2◦C and the measured rate of heat input is 6.1 W. Calculate the thermal
conductivity of the apple.
2. Determine the thermal conductivity of hazelnut having a porosity of 0.45 at 24.8◦C
using the isotropic model ofKopelman. Determine the thermal conductivity using the
order: water (phase 1), carbohydrate (2), protein (3), fat (4), ash (5), and air (6). The
composition data and density (ρ) and thermal conductivity (k) of each component are
given in Table P.3.2.1.
3. The line heat source probe method was used to determine the thermal conductivity of
a food sample. The sample container is filled with the sample with the probe inserted
at the center and placed in a constant temperature bath at 25◦C for equilibration. After
the equilibrium is reached, the probe heater was activated. The electrical resistance of
heated source per unit length is 223.1 Ω/m and the electrical current was measured as
0.14 A. Calculate the thermal conductivity from the data given in Table P.3.3.1.
4. A modified Fitch device was used for measuring the thermal conductivity of a strawberry
sample. The sample was shaped to obtain a cylinder with 6.35 mm diameter and 2.25
mm height, which is suitable for the device. The initial temperature of both the sample
and the copper plug was 25◦C. After equilibrium was reached, the copper rod, which
had a constant temperature of 35◦C, was lowered, making good contact with the
sample surface, and the temperature variation of copper plug was recorded and given
in Table P.3.4.1. Calculate the thermal conductivity of the sample if the mass and
specific heat of the copper plug are 0.010762 kg and 384.9 J/kg K, respectively.
51 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
5. To determine the specific heat of the apple, 100 g of apple was heated from 20◦C to
30◦C for 10 min via the guarded plate method. The voltage and the current supplied to
the heater were 0.5 V and 12 A, respectively. Calculate the specific heat of the apple.
6. The comparison calorimeter is used to determine the specific heat of milk. The first cup
is filled with 100 g of distilled water having a specific heat of 4.18 kJ/kg K. The other
cup is filled with 105 g of milk. Both cups are heated to the same temperature and then
placed in the calorimeter to cool. The rates of cooling are 10◦C/s and 10.3◦C/s for water
and milk, respectively. Both cups have the same mass (50 g) and specific heat of 0.95
kJ/kg K. The calorimeter is well insulated and the heat loss to the surroundings is
negligible. Calculate the specific heat of milk.
7. The specific heat of cucumber was determined using DSC. An 18-mg sample was used.
The deflection from the baseline was 87.78 for the sample and 19 for the sapphire at
32◦C in the thermogram. The specific heat of the sapphire is 0.191 kJ/kg K. If the mass
of reference material is 82 mg, calculate the specific heat of cucumber at 32◦C.
52 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
8. The Dickerson method was used to measure thermal diffusivity of a sample. The
variation of surface (Ts) and center (Tc) temperatures of a sample during measurement
is shown in Table P.3.8.1. A cylindrical container in which the sample was placed has a
diameter of 3.6 cm. The center temperature (Tc) and surface temperature (Ts ) were
recorded at 2-min intervals until the surface temperature of the sample reached to
60◦C. Calculate the thermal diffusivity of the sample.
Books
1. ASHRAE. 2014 ASHRAE Handbook. Refrigeration SI Edition. 2014. ASHRAE. GA USA
2. Figura, L.O. and Teixeira, A.A. Food Physics Measurement Physical Properties-
Measurement and Applications. 2007. Springer-Verlag Berlin Heidelberg NY USA.
3. Kutz, M. Handbook of Farm, Dairy and Food Machinery Engineering 3rd Edition.
2019. Academic Press. CA USA.
4. Sahin, S. & Sumnu, S.G. Physical Properties of Foods. 2006. Springer Science and
Business Media, LLC. USA
5. Rao, M.A. et. al. Engineering Properties of Foods 4th Edition. 2014. CRC Press Taylor
& Francis Group. FL USA
Suggested Readings:
1. Arana, I. Properties of Foods Novel Measurement Techniques and Applications.
2012. CRC Press Taylor & Francis Group. NY USA.
2. Berk, Z. Food Process Engineering and Technology. 2009. Academic Press. NY
USA.
3. Singh, R.P. & Heldman, D.R. Introduction to Food Engineering, 5th Edition, 2014.
Academic Press. CA USA
Online Sources:
https://www.cae.tntech.edu/~jbiernacki/CHE%204410%202016/Thermal%20Properties
%20of%20Foods
https://thermtest.com/thermal-diffusivity-overview
https://www.quora.com/What-is-relationship-between-specific-heat-capacity-and-heat-
capacity
https://youtu.be/TqJFIBODrjM
53 | P a g e
Properties of ABE Material
Module 2 Lesson 4
Thermal Properties
Engr. S. Balsote, 09/2020
http://ecoursesonline.iasri.res.in/mod/page/view.php?id=1012
54 | P a g e