You are on page 1of 23

5.

0 STEADY STATE HEAT TRANSFER

5.1 Fourier equation for steady state one dimensional conduction

5.1.1 Conduction through a uniform homogeneous slab

Steady state conditions imply that time has no influence on the temperature distribution within an

object, and temperature may be different at different locations within the object.

The heat flux (heat flow density or heat flow rate intensity) due to conduction in the x-direction

through a uniform homogeneous slab of material (Fig. 5.1) is given by Fourier’s first law of

conduction

T1 T2

dT/dX q

x x

Fig. 5.1 Steady-state conduction through a uniform homogeneous slab

Where T1 and T2 are the surface temperatures of the respective faces of the slab and X is its

thickness in the x-direction. The rate of heat transfer is given by

Q = KA (T1 −T2) / X . . . . . . . . . . . . . . (5.1)

Eqn. 5.1 is also known as Fourier’s law, which can be rearranged in the form

Q = A▲T
X/ K . . . . . . . . . . . . . . . . . . (5.2)

1
Equation (5.2) shows that for uni-dimensional, steady-state conduction, the rate of heat transfer

Q, is proportional to the temperature difference across the slab (ΔT) and the area through which

heat is transferred, and is inversely proportional to the quantity X/K or thermal resistance.

Thermal resistance (X/K) is a measure of the difficulty with which heat is conducted through a

solid. It depends upon the thermal conductivity of the material, and the dimension in the

direction of heat flow, in this case the slab thickness. Note that thermal resistance has units of m 2

KW−1. Table 5.1 lists the thermal conductivities of a range of food and non-food materials.

Example 5.1

In an experiment to measure the thermal conductivity of meat, beef was formed into a square

section block 5cm × 5cm and 1 cm thick. The edges of the block were insulated and heat was

supplied continuously to one face of the block at a rate of 0.80W. The temperatures of each face

were measured with thermocouples and found to be 28.5 and 23.3◦C, respectively. What is the

thermal conductivity of beef?

The integrated form of Fourier’s law, Eq. (5.1), can be rearranged explicitly in terms of thermal

conductivity:

K = Q X / A (T1 −T2) . . . . . . . . . . . . . . (5.2)


(Take care to convert both the area and the slab thickness into SI units)

K = 0.80×0.01 / (0.05)2(28.5−23.3) Wm−1 K−1

K=

5.2 Series and Parallel arrangement of Conductors

2
5.2.1 Conduction in a Composite Slab (Total Conductance)

If the slab in 5.1.1 is a composite of materials of both different thickness and different thermal

conductivity, as in Fig. 5.2, Eq. (5.1) cannot be used for the whole slab; because there is no

single value of either X or K, and thermal conductivities cannot be added together. However,

applying Eq. (5.2) to each layer in turn gives the respective temperature differences as

▲T1 = Q1 / A * X1 / K1

▲T2 = Q2 / A *X2 / K2

▲T3 = Q3 /A * X3 / K3 . .. . . . . . . … . . . . . .. . . . . . . (5.3)

where Q1, Q2 and Q3 are the rates of heat transfer in each layer, respectively.

Now

▲T1 =T1 −T2

▲T2 =T2 −T3

▲T3 =T3 −T4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5.4)

The overall temperature difference across the composite slab is

▲T = T1 −T4 . . . . . . . . . . . . . . . . . . . . . . (5.6)

And

▲ ▲T = ▲T1 +▲T2 +▲T3 . . . . . . . . . . . . . . . . . . (5.7)

3
T1 T2 T3 T4

X1 X2 X3

K1 K2 K3

Fig. 5.2 Steady-state conduction through a composite slab

Also, there is no generation or accumulation of heat at the interfaces between each layer,

therefore,

Q1 = Q2 = Q3 = Q . . . . . . . . . . . . . . . . . . . . . . . . . . . (5.8)

Thus, substituting from Eqs. (5.3) and (5.8) into Eq. (5.7), the overall temperature difference

becomes

▲T = Q / A ( X1/ K1 + X2 / K2 + X3 / K3) (5.9)

Therefore,

Q = A ▲T
( X1 / K1 + X2 / K2 + X3 / K3) (5.10)

The rate of heat transfer through a composite slab is now proportional to area and to the overall

temperature difference across the whole slab and is inversely proportional to the sum of the ther-

mal resistances of each layer. This is an extremely important point: thermal resistance is an

additive property, whereas thermal conductivity is not. A general version of Eq. (5.10) is now

4
Q = area ×∑ temperature difference
∑ thermal resistance (5.11)

This relationship can be applied to any steady-state problem involving either conduction or

convection or combinations of the two.

Example 5.2

It is proposed to build a cold store having an outer wall of concrete (100 mm thick) and an inner

wall of wood (10 mm thick), with the space in between (100 mm) filled with polyurethane foam.

If the inner wall temperature is 5◦C and the outer wall is maintained at the ambient air

temperature of 20◦C, calculate the rate of heat penetration. The wall area is not specified and

therefore only the rate of heat penetration per unit area can be determined.

Hence, from Eq. (5.10),

Q = ▲T
( X1 / K1 + X2 / K2 + X3 / K3) 5.12

The overall temperature difference across the composite wall is the difference between 20 and

5◦C. Let the subscripts 1, 2 and 3 represent concrete, insulation and wood, respectively. Thus,

taking thermal conductivity data from Table 5.1,

Q /A = (20−5)
0.10 / 0.80 + 0.10 / 0.025 + 0.01 / 0.17 Wm−2

Q/A= (20−5)
(0.125 + 4.0 + 0.0588) Wm−2

Q / A = 3.59Wm−2

5
3 Note the magnitudes of the thermal resistances. That of the insulation is significantly greater

than those of the structural materials.

5.2.2 Radial Conduction

Fig. 5.3 Steady-state conduction through a thick-walled cylinder

Many problems in food engineering involve the conduction of heat in a radial direction through

the walls of pipes and tubes (Fig. 5.3). If a hot liquid is conveyed along a pipe, heat will be lost

from the liquid to the surroundings, assuming the surroundings are at a lower temperature, even

if the pipe is insulated. E.g, in some heat exchangers, in which heat is deliberately transferred

from one fluid to another, the fluids are carried in tubular conduits or pipes.

Here, the temperature varies with radius r of the pipe, but is constant along the length of the pipe

L. The rate of heat transfer is then given by

Q = 2πr2L (T1 −T2)


r2/K ln (r2/r1) (5.13)

Where,

r2/K ln (r2/ r1) = the thermal resistance of the cylinder wall

2πr2L is the (external) surface area through which heat is transferred.

It is possible to base the rate of heat transfer on the internal surface area, but it is conventional to

use the external area. The thermal resistance is inversely proportional to thermal conductivity

and is a function of the inner and outer radii of the cylinder which define the distance in the r-

6
direction through which heat is transferred. As before this has units of m 2KW−1. It is essential to

realize that the temperatures in Eq. (5.13) are surface temperatures and not the temperatures of

the fluids (the hot food fluid and the air surrounding the pipe).

Example 5.3

Water flows through an uninsulated pipe which has a diameter of 0.05 m, and a wall thickness of

0.01 m. The thermal conductivity of the pipe wall is 50Wm −1 K−1 and the inside and outside

surface temperatures of the pipe are 70 and 69.5◦C, respectively. Calculate the radial heat loss

per metre length. Adapting Eq. (5.13), the rate of heat transfer in a radial direction per unit length

of pipe is

Q / L = 2πr2 (T1 −T2)


r2 / K ln (r2/ r1)

For convenience, the external radius may be cancelled from this expression and therefore

Q / L = 2π (70−69.5)
1 /50 ln (0.035/ 0.025) Wm−1

Q / L = 466.8Wm−1

5.2.3 Conduction in a Composite Cylinder

Fig. 5.4 Steady-state conduction through a composite cylinder

7
In order to reduce heat loss from a pipe conveying a hot food fluid a layer of insulation material

of low thermal conductivity may be placed around the pipe (Fig. 5.4). If the insulation has a

thermal conductivity kB and an outer radius r3 then the rate of heat transfer through it is

Q = 2πr2L▲T
r2 / kA ln r2 / r1+ r2/ kB ln r3 / r2 (5.14)

Note that Eq. (5.14) is based on the external surface area of the pipe 2πr2L.

5.3 Heat Transfer by Convection

Introduction

Every moving fluid carries an associated energy that causes heat to be transferred from one point

to another; these points are at different temperatures due to such movement. This type of heat

transmission is called convection. If a fluid is in contact with a solid that is at a greater

temperature, the fluid receives heat that is transferred within it by movement of the particles of

the fluid. This movement causes heat transport by convection to happen, and it can occur in

natural or forced form. The first case occurs when there is no mechanical agitation and is

attributed to density differences at different points of the fluid caused by the effect of

temperature. On the other hand, forced convection occurs when the movement of the fluid is

produced mechanically using devices such as agitators and pumps, among others. Heat transfer

by convection is very important when studying the heat exchange between two fluids separated

by a wall in such a way that one of them gives up heat to the other, so that the first fluid cools

while the second heats up. The devices in which this heat transmission is performed are called

heat exchangers.

8
5.3.1 Film Heat Transfer Coefficient

Imagine a container of liquid to which heat is supplied as indicated in Fig.5.5. Further, Imagine a

small element or packet of liquid, perhaps only a few molecules thick, in contact with the heated

surface of the container. Heat is transferred by conduction through the container wall and

initially by conduction to the liquid packet. The temperature of the liquid increases and

consequently it experiences thermal expansion and a decrease in density. Thus the packet of

liquid rises (because of buoyancy forces) and as it moves away from the hot container wall it is

replaced by cooler liquid which will in turn be heated and will rise through the body of liquid.

The circulation pattern which is established in this way is known as a convection current and is

responsible for the transfer of heat throughout the mass of a fluid (either liquid or gas). The rate

at which fluid is circulated, and thus the rate at which heat is transferred, can be increased by

deliberately agitating the fluid, for example by the addition of an impeller. A current which is

established solely because of thermal expansion is called free or natural convection, whilst

deliberately enhanced circulation is called forced convection.

In pipe flow, forced convection can be introduced by ensuring that the flow is turbulent.

However, it is accepted that, there is a region of more slowly moving and often laminar fluid

adjacent to the wall of the pipe which may be termed the boundary layer. Beyond this layer the

fluid has a much greater velocity and may well be turbulent. We may assume that heat is

transferred by conduction through this layer and it follows that the layer contains almost all of

the resistance to heat transfer. This is because most fluids (Non Newtonian) have relatively low

thermal conductivities, and also because of the rapid heat transfer from the edge of the boundary

layer into the bulk of the fluid. Thus it is valid to write for this layer

9
Q = A▲T
x/k (5.15)

But, because x (the boundary layer thickness) can neither be predicted nor measured easily, the

thermal resistance cannot be determined and x / k is replaced with the term 1 / h, where h is a

film heat transfer coefficient. Eq. (5.15) now becomes

Q = hA▲T (5.16)

Where

h=k/x

Heat transfer coefficients are thus empirical measures of the ability to transfer heat in particular

circumstances and are used to describe heat transfer within the entire fluid both across any

boundary layer and across turbulent or higher velocity regions. Values of the heat transfer

coefficient must be determined in order to solve any realistic industrial heat transfer problem and

much effort in process engineering research has been, and is, directed to this end. Increasing the

velocity of the fluid has the effect of promoting turbulence (Reynolds’ experiment) and thus of

reducing the thickness of the boundary layer. Therefore, as x / k can be approximated to 1 / h, the

heat transfer coefficient increases with increasing fluid velocity. Values of h depend also upon

the physical properties of the fluid and upon geometry. For example, there is a difference

between flow in a circular cross-section pipe and flow in an annulus with consequent differences

in the rate at which heat is transferred even though materials and temperatures may be the same

in each case. By inspecting Eq. (5.16) it is clear that the units of a film heat transfer coefficient

must be W m−2 K−1.

10
5.3.2 Radial Convection

The analysis of convection can now be extended to radial geometries. Consider again the

problem of conveying a hot food fluid in a circular cross-section pipe. The rate of heat transfer

from the hot fluid (temperature Ti) inside the pipe to the inner pipe surface (temperature T 1) is

Q= 2πr2L ▲T
r2 r1hi + r2 ln r2 / r1 + 1/ ho
k (5.17)

Where

2πr1L = the inner surface area

ho = the film heat transfer coefficient

hi = the relevant film heat transfer coefficient

2πr2L = the external pipe surface area

K = thermal conductivity

▲T = change in Temperature

r1 = the internal tube radius

r2 = the external tube radius

Example 5.4

Steam at 100◦C condenses on the outside of an alloy tube of thermal conductivity 180 W m−1 K−1

through which water flows at a velocity such that the tube-side film heat transfer coefficient is

4000 W m−2 K−1. The film heat transfer coefficient for condensing steam may be assumed to be

10,000 W m−2 K−1. The tube is 5 m long, has an external diameter of 25 mm and a wall thickness

of 1 mm. If the mean temperature of the water is 15◦C, calculate the rate of heat transfer to the

11
water. It is instructive to consider the magnitudes of the three thermal resistances involved. The

external tube radius r2 is 0.0125 m and the internal radius r1 is 0.0115 m.

This problem may be solved simply by substituting the relevant quantities into Eq. (5.17).

Therefore,

The conductive resistance is

r2/ k ln r2 /r1 = 0.0125 /180 ln 0.0125 / 0.0115 m2 KW−1

= 5.79×10−6 m2 KW−1

The tube-side thermal resistance is,

r2 / r1hi = 0.0125 / 0.0115×4000 m2 KW−1

= 2.72×10−4 m2 KW−1

The film resistance on the outside of the tube is = the reciprocal of the film coefficient i.e 1/h o

= 10−4 m2 KW−1.

The total thermal resistance is therefore

5.79×10−6 + 2.72×10−4 + 10−4

= 3.78×10−4 m2 KW−1

Substituting for the temperature driving force and the external tube surface area (2πr 2L) gives the

rate of transfer of heat to the water as

Q = 2π0.0125×5(100−15) W

3.78×10−4

Q = 88.42kW

In a condenser, as in any heat exchanger, the conductive resistance should be minimised and

this is achieved by using thin-walled tubes of high thermal conductivity.

12
5.4 Overall Heat Transfer Coefficient

Thermal resistance is a measure of the difficulty with which heat is transferred and is inversely

proportional to the film heat transfer coefficient. It is possible to rewrite Eq. (5.11), replacing the

sum of thermal resistance term with the reciprocal of an overall heat transfer coefficient, U. Thus

Q = UA▲T (5.18)

The overall heat transfer coefficient includes both conductive and convective resistances and for

the case of heat transfer to or from a fluid in a circular cross-section pipe is given by

1 / U = resistance of fluid i + resistance of pipe wall + resistance of fluid o (5.19)

Which becomes

1 / U = 1/ hi + r1 ln r2 / r1 + r1 / r2 ho (5.20)
k

Example 5.5

Calculate the overall heat transfer coefficient for the condenser in Example 5.4.

The overall heat transfer coefficient is the reciprocal of the sum of thermal resistance.

U = 1 / (2.72×10−4) + (5.79×10−6) + (10−4) Wm−2 K−2

U = 2647Wm−2 K−1

5.5 Heat Transfer by Radiation

5.5.1 Principles of Radiation

Radiation is the term given to the transfer of energy by electromagnetic waves; specifically, ther

mal radiation is the transfer of energy by that part of the electromagnetic spectrum with

wavelengths between 10−7 and 10−4 m. All wave forms are propagated at what is known as the

13
velocity of light c, which has a value of approximately 3×108 ms−1. The relationship between the

velocity of propagation, wavelength λ and the frequency υ is then

c = λυ (5.21)

Where frequency is measured in Hz. Note that wavelengths are often quoted in nanometers (nm)

or in Angstroms (Å) where 1Å = 10−10 m.

All bodies at a temperature above absolute zero emit thermal radiation but the quantities

involved only become significant at temperatures well above ambient. Unlike conduction and

convection, radiation can be propagated through a vacuum and the most important example for

human life is that of solar radiation; that is, thermal radiation travelling through space (as an

electromagnetic wave) from a body with a surface temperature of about 6000 K. On reaching the

earth this energy is then dissipated by conduction and convection. In many industrial examples

of heat transfer conduction and convection occur simultaneously with radiation, despite the very

dissimilar mechanisms involved. However, thermal radiation is the dominant heat transfer

mechanism in a number of food processing applications such as drying, especially where the

product is in a granular form with a relatively low initial moisture content. Some examples are

tunnel or conveyor driers, vacuum band or shelf driers used for the manufacture of ‘puffed’ or

expanded cereal products and freeze driers. Radiation is also the principal mechanism in solar

drying which is used extensively in the developing world for the drying of food in the open air.

In addition, mention may be made of baking, roasting and heat shrink packaging.

5.5.2 Absorption, Reflection and Transmission

All radiation falling upon a surface is either absorbed, reflected or transmitted. The fraction of

incident radiation which is absorbed is called the absorptivity of that surface and is denoted by α.

14
Similarly, the reflectivity ρ and transmissivity τ of the surface are the fractions of incident

radiation which are reflected from the surface or transmitted through the body respectively. It

therefore follows that

α + ρ +τ =1 (5.22)

Most solids, except those that are transparent to visible light (e.g. quartz, glass and some plastics)

do not transmit radiation. Hence

α + ρ =1 (5.23)

For gases, reflectivities are zero and therefore

α + τ =1 (5.24)

At this point it is essential to understand that the emission, absorption and reflection of radiation

are surface phenomena. The rate at which thermal radiation is emitted is a function of the surface

temperature and surface characteristics such as colour. Equally, the colour of a surface

determines its absorptivity. The absorption of radiation takes place only at the surface and

thereafter heat transfer through the body is by conduction. The nature of a surface, especially its

roughness, has a considerable effect upon reflection. For a specular reflector the angle of

incidence equals the angle of reflection. This will be the case where the roughness dimension is

very much smaller than the wavelength (Fig. 5.5a). However, if the roughness dimension is

considerably greater than the wavelength of the incident radiation, the reflection will be diffuse,

that is it is reflected in all directions and with varying intensities (Fig. 5.5b). For a specular

Fig. 5.6 Surface characteristics: (a) Specular surface, (b) Diffuse surface

15
surface, for example polished metals, the reflectivity approaches unity and the absorptivity will

be close to zero whereas for a diffuse surface the absorptivity approaches unity.

5.5.3 Black Body Radiation

Emissive power E is defined as the total emitted radiant thermal energy leaving a surface per unit

time and per unit area, over all wavelengths and in all directions. This should not be confused

with radiosity which is the total radiant thermal energy leaving a surface, per unit time and per

unit area, and which includes both emitted and reflected radiation. A black body is defined as one

which absorbs all radiation falling upon it (regardless of directional or spectral differences), that

is,

α B = 1.

The emissive power of a black body depends upon the fourth power of its absolute temperature,

thus

EB = σT4 (5.25)

Where σ is the Stefan−Boltzmann constant and has a value of 5.67×10−8 Wm−2 K−4.

Example 5.5

Determine the total emissive power of a black body at 1027◦C. Emissive power depends upon

the absolute temperature.

Therefore, absolute temperature = 1027 + 273

= 1300

Thus, from Eq. (5.25),

EB =5.67×10−8 (1300)4 kW m−2

16
EB =161.9kW m−2

5.5.4 Emmisivity

Values of emissivity must be known in order to calculate rates of radiative heat transfer between

surfaces. Kirchoff’s law states ‘the emissivity of any surface equals its absorbtivity when it is in

thermal equilibrium with its surroundings’ This can be shown to be the case by considering two

bodies in an adiabatic enclosure. Let I be the incident radiation rate, A1 and A2 the respective

areas of the two bodies and α1 and α2 the respective absorbtivities. Now, at equilibrium, the

radiation falling on body 1 will equal the radiation emitted by body 1, and therefore, for body 1

IA1α1 =E1A1 (5.26)

and for body 2

IA2α1 = E2A2 (5.27)

where E1 and E2 are the respective emissive powers. Thus, from Eq.s. (5.26) and (5.27),

E1 / α1 = E2 / α2 = E / α (5.28)

In other words for all bodies the ratio of emissive power to absorptivity is the same and depends

only on temperature. It must also be the case that

E / E B = α / αB (5.29)

Now by definition α is equal to unity and therefore

e=α (for all bodies) (5.30)

17
Table 5.5 Approximate values of emissivity for various surfaces

Dull black surface 1.0

Non metallic/non polished 0.90

Oxidised metals 0.60–0.85

Highly polished metals 0.03–0.06

Most foods 0.80–0.90

Dough 0.85

Lean beef 0.74

Beef fat 0.78

Water 0.95

Ice 0.97

5.5.5 Radiative Heat Transfer

Imagine two infinite parallel plates 1 and 2. The rate at which radiant heat is emitted by surface 1

and absorbed by surface 2 is E 1A 1α 2 whilst heat is emitted by surface 2 and absorbed by surface

1 at a rate E 2A 2α 1. The net rate of heat transfer between the two surfaces is then

Q1→2 =E 1A 1α 2 −E 2A 2α 1 (5.31).

18
The emissive powers E 1 and E 2 are fractions of the emissive powers of a black body radiator and

therefore

E 1 =e 1EB1

E2 =e2EB2

(5.32)

Thus, on substitution, Eq. (7.113) becomes

Q1→2 =e1EB1A1α2 −e2EB2A2α1 (5.33)

However, from the Stefan–Boltzmann law

EB1 = σT14

EB2 = σT24

(5.34)

Consequently,

Q1→2 = σ(e1A1α2T14 −e2A2α1T24) (5.35)

Now the areas of surfaces 1 and 2 are equal (they are both infinite plates of area A) and the

absorbtivities and emissivities are the same (e and α, respectively) for similar surfaces, hence

Q1→2 = σeAα(T14 −T24) (5.36)

Therefore the heat flux is given by

q1→2 = σeα(T14 −T24) (5.37)

And for the case of black surfaces where α = e = 1 the above reduces to

q1→2 = σ(T14 −T24) (5.38)

19
Example 5.6

Two infinite parallel plates are held at 850 and 700◦C, respectively. If each plate approximates to

a black body, calculate the rate of heat exchange between them. For a black body e B = αB =1 and

therefore the net heat flux is given by.

q = 5.67×10−8 [(1123)4 − (923)4] Wm−2

q = 49.03kW m−2

6.0 UNSTEADY STATE HEAT TRANSFER

In all heat transfer operations, it is assumed that the temperature of any point of the solid remains

constant with time. However, there are cases in which temperature inside the solid, besides

changing with position, also changes with time. Example is the case of freezing and thawing, in

which it is desirable to know the time needed to reach a certain temperature at a determined point

of the solid, or to know the temperature of such a point after some time. This process is

developed under unsteady state regime.

Unsteady state (or transient) heat transfer is that phase of the heating and cooling process when

the temperature changes as a function of both location and time. By contrast, in steady state heat

transfer, temperature varies only with location. In the initial unsteady state period, many

important reactions in the food may take place, e.g in pasteurization and food sterilization

20
processes, the unsteady state period is an important component of the process. Analysis of

temperature variations with time during the unsteady state period is essential in designing such a

process.

Since temperature is a function of two independent variables, time and location, then for a one

dimensional heat transfer, the following equation applies for conduction;

▲T / ▲t = k / ▲cprn * 1/ ▲r * (rn ▲T /▲r)

Where;

T = Temperature (oC)

t = Time (s)

r = distance from center location

k / ρcp = is defined as the thermal diffusivity

For convection, the equation is given as;

K * ▲T / ▲r = h (Ta – Ts)

Where

h = convective heat transfer coefficient

Ta = Temperature of heating or cooling medium far away from the surface

Ts = Temperature at the surface

7.0 THERMAL CONDUCTIVITY

Thermal conductivity is an indication of how good a material is in conducting heat. In

quantitative terms, this property gives the amount of heat that will be conducted per unit time

through a unit thickness of material if a unit temperature difference exists across the thickness.

Hence, materials with high thermal conductivities are best for transferring heat, while those with
21
low thermal conductivities are better suited as insulators. Example, metals are highly conducting

in nature, therefore have high thermal conductivities; while materials like asbestos, wood, and

cotton are poor conductors, and therefore have low thermal conductivities. Most high moisture

foods have thermal conductivity values closer to that of water, while the thermal conductivities

of dried, porous foods is influenced by the presence of air with its low thermal conductivity

value. The thermal conductivity of ice is almost four times that of water, and that is why

naturally frozen foods have higher thermal conductivities than normal foods, which makes them

lose heat very quickly when exposed to the outside environment.

The thermal conductivity values of certain materials are tabulated or predetermined. However,

values can also be determined using empirical predictive equations for the purpose of process

calculations where temperature may be changing. Some of these equations are given below:

1. Fruits and Vegetables:

For fruits and vegetables with a water content greater than 60%, the following equation had

been proposed by Sweat;

k = 0.148 + 0.493Xw

Where

k = thermal conductivity

Xw = Water content expressed as a fraction

2. Meats and Fish:

For meats and fish at a temperature range between 0 to 60oC, water content 60-80% wet

basis, Sweat proposed the following equation

k = 0.08 + 0.52Xw

3. Solid and liquid foods

22
Another empirical equation developed by Sweat for solid and liquid foods is as follows:

K = 0.25Xc +0.155Xp + 0.16Xf +0.135Xa + 0.58Xw

Where,

X = Mass fraction

Subscripts c,p,f,a, and w are carbohydrate, protein, fat, ash, and water respectively.

The coefficients are thermal conductivity values of the pure component. It should be

noted however that the thermal conductivity of water in a food is different from that of

pure water.

4. Choi and Okos equation:

Choi and Okos gave the following expression that includes the influence of product

composition and temperature;

n
k = ∑ kiYi
i=1

Where a food material has n components, ki is the thermal conductivity of the ith

component, Yi is the volume fraction of the ith component, obtained as follows;

Yi = Xi / ρi
n
∑ (Xi / ρi)
i=1

Where,

Xi = Weight fraction

ρ = density of the ith component

23

You might also like