0% found this document useful (0 votes)
57 views19 pages

Thermal Properties of Foods Guide

Propiedades termicas

Uploaded by

carlos
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
57 views19 pages

Thermal Properties of Foods Guide

Propiedades termicas

Uploaded by

carlos
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 19

Thermal Property of Food

Related terms:

Physical Property of Food, Infrared Heating, Frozen Food, High Fat Food, Retort
Pouches, Tomato Paste, Raw Food, Pasteurization, Canned Food

View all Topics

FREEZING | Principles
B.R. Becker, B.A. Fricke, in Encyclopedia of Food Sciences and Nutrition (Second
Edition), 2003

Thermal Properties of Foods


The thermal properties of foods are important in the design of food storage and
refrigeration equipment as well as in the estimation of process times for refrig-
erating, freezing, heating, or drying of foods. Because the thermal properties of
foods are strongly dependent upon chemical composition and temperature, the
most viable option is to predict these thermal properties using mathematical models
that account for the effects of chemical composition and temperature.

Composition data for foods are readily available in the literature. These data consist
of the mass fractions of the major food components: water, protein, fat, carbohy-
drate, fiber, and ash. Food thermal properties can be predicted by using these com-
position data in conjunction with temperature-dependent mathematical models of
the thermal properties of the individual food components.

Equations for predicting the thermal properties of these food components have
been developed as functions of temperature in the range of − 40 to 150 °C. These
equations are presented in Table 1. Because water is the predominant constituent
in most food items, the water content of food items significantly influences the
thermophysical properties of foods. Therefore, equations for predicting the thermal
properties of water and ice have also been developed. These equations are presented
in Table 2.

Table 1. Thermal property equations for food components (−40 °C ≤ t ≤ 150 °C)
Thermal property Food component Thermal property model
Thermal conductivity (W m−1 K- Protein k = 1.7881 × 10−1 + 1.1958 × 10-
−1)
−3t − 2.7178 × 10−6t2

Fat k = 1.8071 × 10−1 − 2.7604 × 10−3t − 1.7749 × 10-


−7t2

Carbohydrate k = 2.0141 × 10−1 + 1.3874 × 10−3t − 4.3312 × 10-


−6t2

Fiber k = 1.8331 × 10−1 + 1.2497 × 10−3t − 3.1683 × 10-


−6t2

Ash k = 3.2962 × 10−1 + 1.4011 × 10−3t − 2.9069 × 10-


−6t2

Density (kg m−3) Protein = 1.3299 × 103 − 5.1840 × 10-


−1t

Fat = 9.2559 × 102 − 4.1757 × 10−1t

Carbohydrate = 1.5991 × 103 − 3.1046 × 10−1t

Fiber = 1.3115 × 103 − 3.6589 × 10−1t

Ash = 2.4238 × 103 − 2.8063 × 10−1t

Specific heat (J kg−1 K−1) Protein cp = 2.0082 × 103 + 1.2089t −-


1.3129 × 10−3t2

Fat cp = 1.9842 × 103 + 1.4733t − 4.8008 × 10−3t2


Carbohydrate cp = 1.5488 × 103 + 1.9625t − 5.9399 × 10−3t2
Fiber cp = 1.8459 × 103 + 1.8306t − 4.6509 × 10−3t2
Ash cp = 1.0926 × 103 + 1.8896t − 3.6817 × 10−3t2

Table 2. Thermal property equations for water and ice (− 40 °C ≤ t ≤ 150 °C)

Thermal property Thermal property model


Water Thermal conductivity (W m−1- kw = 5.7109 × 10−1 + 1.7625 ×-
K−1)
10−3t − 6.7036 × 10−6t2

Density (kg m−3) w = 9.9718 × 102 + 3.1439 × 10−3t − 3.7574 × 10−3

t2

Specific heat (J kg−1 K−1)a cw = 4.0817 × 103 − 5.3062t + 9.9516 × 10−1t2

Specific heat (J kg−1 K−1)b cw = 4.1762 × 103 − 9.0864 × 10−2t + 5.4731 × 10-
−3t2
Ice Thermal conductivity (W m−1- kice = 2.2196 − 6.2489 × 10−3t +-
K−1)
1.0154 × 10−4t2

Density (kg m−3) ice = 9.1689 × 102 − 1.3071 × 10−1t

Specific heat (J kg−1 K−1) cice = 2.0623 × 103 + 6.0769t

a For the temperature range of −40 to 0 °C.

b For the temperature range of 0 to 150 °C.

In general, the thermophysical properties of a food item are well behaved when the
temperature of the food item is above its initial freezing point. However, below the
initial freezing point, the thermophysical properties of a food item vary dramatically
with temperature.

> Read full chapter

Food Freezing Technology


Kasiviswanathan Muthukumarappan, ... V. Sunkesula, in Handbook of Farm, Dairy
and Food Machinery Engineering (Third Edition), 2019

15.8.3 Effect of Freezing on Food Thermal Properties


Knowledge of thermal properties of food products is needed in design of cooling,
freezing processes, and equipment, as well as cooling load calculations. Data on
thermal properties of some foods are given in Table 15.3. Thermal conductivity of
ice (k=2.24 W/m K) is around four times that of water (k=0.56 W/m K). Consequently,
thermal conductivity of frozen foods will be three to four times higher than that of
unfrozen foods. During the initial stages of freezing, increase in thermal conductivity
is rapid. For high fat foods, the variation in thermal conductivity with temperature is
negligible. For meats, orientation of fibers greatly influences thermal conductivity.
Thermal conductivity measured along the fibers is 15%–30% higher than that mea-
sured across the fibers in meats (Dickerson, 1968). Thermal conductivities of several
food products at different temperatures are available in the literature (Woodams and
Nowrey, 1968; Lentz, 1961; Smith et al., 1952).

Table 15.3. Thermal Properties of Frozen Foods (Earle, 1983; Rahman and Velez-Ruiz,
2004)

Food Water Content (%) Specific Heat (kJ/kg K) Latent Heat (kJ/kg)
Apple 84 1.88 280
Banana 75 1.76 255
Beef 75 1.67 255
Bread 32–37 1.42 108.7–221.2
Cabbage 92 1.96 305.1
Carrot 88 1.88 292.6
Chicken – 1.77 247
Egg – 1.67 275.9
Fish 70 1.67 275.9
Green beans 89 1.96 296.8
Ice cream – 1.63 210
Milk 87.5 2.05 288.4
Oranges 85.9 1.94 291
Peaches 87 1.92 288.4
Pork 60 1.59 196.5
Shrimp 75.3 1.89 277
Strawberries – 1.97 301
Tomato 92.3 2.02 314
Turkey 92.8 1.65 214
Water – 2.01 334
Watermelon 92 2.0 305.1

Specific heat of ice (2.1 kJ/kg K) is only half of the specific heat of water (4.218 kJ/kg K).
On freezing, specific heat of foods decreases. Measurement of specific heat is
complicated because there is continuous phase change from water to ice. Latent heat
of fusion for any food product can be estimated from the water fraction of the food
(Fennema et al., 1973). Solute concentration in foods is so small that latent heat of
freezing of solutes is generally ignored while estimating the cooling loads. Thermal
diffusivity of frozen foods can be calculated from density, specific heat, and thermal
conductivity data. The thermal conductivity of ice is around four times higher than
that of water and its specific heat is half that of water. This leads to an increase of
around nine to ten times in thermal diffusivity values of frozen foods when compared
with unfrozen ones (Desrosier and Desrosier, 1982).

> Read full chapter

Heat processing
P.J. Fellows, in Food Processing Technology (Third Edition), 2009

10.1.1 Thermal properties of foods


Three important thermal properties of foods are specific heat, thermal conductivity
and thermal diffusivity. Specific heat is the amount of heat needed to raise the
temperature of 1 kg of a material by 1 °C. It is found using Equation 10.1 and specific
heat values for selected foods and other materials are given in Table 10.1.

Table 10.1. Specific heat of selected foods and other materials

Material Specific heat (kJ kg − 1 °C − 1) Temperature (°C)


Foods – solid
Apples 3.59 Ambient
Apples 1.88 Frozen
Bacon 2.85 Ambient
Beef 3.44 Ambient
Bread 2.72 –
Butter 2.04 Ambient
Carrots 3.86 Ambient
Cod 3.76 Ambient
Cod 2.05 Frozen
Cottage cheese 3.21 Ambient
Cucumber 4.06 Ambient
Flour 1.80 –
Lamb 2.80 Ambient
Lamb 1.25 Frozen
Mango 3.77 Ambient
Milk – dry 1.52 Ambient
Milk – skim 3.93 Ambient
Potatoes 3.48 Ambient
Potatoes 1.80 Frozen
Sardines 3.00 Ambient
Shrimps 3.40 Ambient
Foods – liquid
Acetic acid 2.20 20
Ethanol 2.30 20
Milk – whole 3.83 Ambient
Oil – maize 1.73 20
Oil – sunflower 1.93 20
Orange juice 3.89 Ambient
Water
Water 4.18 15
Water vapour 2.09 100
Ice 2.04 0
Non-foods – solid
Aluminium 0.89 20
Brick 0.84 20
Copper 0.38 20
Glass 0.84 20
Glass wool 0.7 20
Iron 0.45 20
Stainless steel 0.46 20
Stone 0.71-0.90 20
Tin 0.23 20
Wood 2.4–2.8 20
Non-foods – gases
Air 1.005 Ambient
Carbon dioxide 0.80 0
Oxygen 0.92 20
Nitrogen 1.05 0

Adapted from Anon (2005c, 2007a), Singh and Heldman (2001a) and Polley et al.
(1980)

10.1

where cp (J kg− 1 °C− 1) = specific heat of food at constant pressure, Q (J) = heat gained
or lost, m (kg) = mass and 1 – 2 (°C) = temperature difference.

The specific heat of compressible gases is usually quoted at constant pressure, but
in some applications where the pressure changes (e.g. vacuum evaporation (Chapter
14, section 14.1) or high-pressure processing (Chapter 8) it is quoted at constant
volume (Cv). The specific heat of foods depends on their composition, especially the
moisture content (Equation 10.2). Equation 10.3 is used to estimate specific heat
and takes account of the mass fraction of the solids contained in the food:

10.2

where M = moisture content (wet-weight basis, expressed as a fraction not percent-


age),

10.3

where X = mass fraction and subscripts w = water, p = protein, f = fat, c = carbohydrate


and a = ash.

Thermal conductivity is a measure of how well a material conducts heat. It is


the amount of heat that is conducted through unit thickness of a material per
second at a constant temperature difference across the material and is found using
Equation 10.4.

10.4

where k (J s− 1 m− 1 °C− 1 or W m− 1 °C− 1) = thermal conductivity and t (s) = time.

Thermal conductivity is influenced by a number of factors concerned with the


nature of the food (e.g. cell structure, the amount of air trapped between cells,
moisture content), and the temperature and pressure of the surroundings. A formula
to predict thermal conductivity based on the composition of foods is shown in
Equation 10.5:

10.5

where kW (W m− 1 °C− 1) = thermal conductivity of water, Xw = mass fraction of water,


ks = (W m− 1 °C− 1) = thermal conductivity of solids (assumed to be 0.259 W m− 1 °C− 1).

A reduction in moisture content causes a substantial reduction in thermal conduc-


tivity. This has important implications in unit operations which involve conduction of
heat through food to remove water (e.g. drying (Chapter 16), frying (Chapter 19) and
freeze drying (Chapter 23)). In freeze drying the reduction in atmospheric pressure
also influences the thermal conductivity of the food.

Ice has a higher thermal conductivity than water and this is important in determining
the rate of freezing and thawing (Chapter 22). The importance of thermal conduc-
tivity is shown in sample problem 10.1 and sample problem 11.1 (Chapter 11). The
thermal conductivities of some materials found in food processing are shown in
Table 10.2.

Sample problem 10.1


Part 1: In a bakery oven, combustion gases heat one side of a 2.5 cm steel plate at
300 °C and the temperature in the oven is 285 °C. Assuming steady state conditions,
and a thermal conductivity for steel of 17 W m− 2 °C− 1, calculate the rate of heat
transfer per m2 through the plate.

Part 2: The internal surface of the oven is 285 °C and air enters the oven at 18 °C.
Calculate the surface heat transfer coefficient per m2, assuming the rate of heat
transfer is 10.2 kW.

Solution to sample problem 10.1

Part 1:

From Equation 10.11,

Part 2:

From Equation 10.13,

This value indicates that natural convection is taking place in the oven.

Table 10.2. Thermal conductivity of selected foods and other materials

Material Thermal conductivity (Wm− 1 Temperature (°C)


oC− 1)

Food
Acetic acid 0.17 20
Apple juice 0.56 20
Avocado 0.43 28
Beef, frozen 1.30 –10
Bread 0.16 25
Carrot 0.56 40
Cauliflower, frozen 0.80 –8
Cod, frozen 1.66 –10
Egg, frozen liquid 0.96 –8
Ethanol 0.18 20
Freeze dried foods 0.01–0.04 0
Green beans, frozen 0.80 –12
Ice 2.25 0
Milk, whole 0.56 20
Oil, olive 0.17 20
Orange 0.41 15
Parsnip 0.39 40
Peach 0.58 28
Pear 0.59 28
Pork 0.48 3.8
Potato 0.55 40
Strawberry 0.46 28
Turnip 0.48 40
Water 0.57 20
Gases
Air 0.024 0
Air 0.031 100
Carbon dioxide 0.015 0
Nitrogen 0.024 0
Packaging materials
Cardboard 0.07 20
Glass 0.52 20
Polyethylene 0.55 20
Poly(vinylchloride) 0.29 20
Metals
Aluminium 220 0
Copper 388 0
Stainless steel 17–21 20
Other materials
Brick 0.69 20
Concrete 0.87 20
Insulation 0.026–0.052 30
Polystyrene foam 0.036 0
Polyurethane foam 0.026 0

Adapted from Anon (2007a,b), Choi and Okos (2003), Singh and Heldman (2001a)
and Lewis (1990)
Although, for example, stainless steel conducts heat ten times less well than alumini-
um (Table 10.2), the difference is small compared with the low thermal conductivity
of foods (20 to 30 times lower than steel) and does not limit the rate of heat transfer.
Stainless steel is much less reactive than other metals, and is therefore used in most
food processing equipment that comes into contact with foods.

Thermal diffusivity is a measure of a material’s ability to conduct heat relative to its


ability to store heat. It is a ratio involving thermal conductivity, density and specific
heat, and is found using Equation 10.6:

10.6

where (m2s− 1) = thermal diffusivity and (kg m− 3) = density. Thermal diffusivity is


used to calculate time-temperature distribution in materials undergoing heating or
cooling and selected examples are given in Table 10.3.

Table 10.3. Thermal diffusivity of selected foods

Food Thermal diffusivity (× l0− 7 m2- Temperature (°C)


s− 1)
Apples 1.37 0–30
Avocado 1.24 41
Banana 1.18 5
Beef 1.33 40
Cod 1.22 5
Ham, smoked 1.18 5
Lemon 1.07 0
Peach 1.39 4
Potato 1.70 25
Strawberry 1.27 5
Sweet potato 1.06 35
Tomato 1.48 4
Water 1.48 30
Water 1.60 65
Ice 11.82 0

Adapted from Singh and Heldman (2001a) and Murakami (2003)

The thermal diffusivity of foods is influenced by their composition, especially their


moisture content, and it can be estimated using Equation 10.7:

10.7

where X = mass fraction and subscripts w = water, f = fat, p = protein and c = carbo-
hydrate. For example, every 1% increase in the moisture content of vegetables cor-
responds to a 1–3% increase in their thermal diffusivity (Murakami 2003). Changes
in the volume fraction of air can also significantly alter the thermal diffusivity of
foods. During heating, the temperature does not have a substantial effect on thermal
diffusivity, but in freezing the temperature is important because of the different
thermal diffusivities of ice and water.

‘Sensible’ heat is the heat needed to raise the temperature of a food and is found
using Equation 10.4, rearranged from Equation 10.1:

10.8

where Q (J) = sensible heat, m (kg) = mass, cp (J kg− 1 °C− 1 or K− 1) = specific heat of
food at constant pressure and (°C) = temperature with subscripts 1 and 2 being
initial and final values.

Phase changes in water are important in many types of food processing including
steam generation for process heating (section 10.2), evaporation by boiling (Chapter
14, section 14.1), loss of water during dehydration, baking and frying (Chapters 16,
18, 19) and in freezing (Chapter 22). ‘Latent’ heat is the heat used to change phase
(e.g. latent heat of fusion to form ice, or latent heat of vaporisation to change water to
vapour) where the temperature remains constant while the phase change takes place.
A phase diagram (Fig. 23.2 in Chapter 23) shows how temperature and pressure
control the state of water (solid, liquid or vapour).

Vapour pressure is a measure of the rate at which water molecules escape as a gas
from the liquid. Boiling occurs when the vapour pressure of the water is equal to
the external pressure on the water surface (boiling point = 100 °C at atmospheric
pressure at sea level). At reduced pressures below atmospheric, water boils at lower
temperatures as shown in Chapter 14 (Fig. 14.1).

The changes in phase can be represented on a pressure–enthalpy diagram (Fig. 10.1)


where the bell-shaped curve shows the pressure, temperature and enthalpy rela-
tionships of water in its different states. Left of the curve is liquid water, becoming
subcooled the further to the left, and right of the curve is vapour, becoming super-
heated the further to the right. Inside the curve is a mixture of liquid and vapour. At
atmospheric pressure, the addition of sensible heat to liquid water increases its heat
content (enthalpy) until it reaches the saturated liquid curve (A–B in Fig 10.1). The
water at A is at 80 °C and has an enthalpy of 335 kJ kg− 1 and when heated to 100 °C
the enthalpy increases to 418 kJkg− 1. Further addition of heat as latent heat causes a
phase change. Moving further across the line (B–C) indicates more water changing
to vapour, until at point C all the water is in vapour form. This is then saturated steam
that has an enthalpy of 2675 kJ kg− 1 (i.e. the latent heat of vaporisation of water is
2257 (2675 – 418) kJ kg− 1 at atmospheric pressure while the temperature remains
constant at 100 °C). Within the curve along B–C, the changing proportions of water
and vapour are described by the ‘steam quality’. For example at point E, the steam
quality is 0.9, meaning that 90% is vapour and 10% is water. The specific volume
of steam with a quality < 100% can be found using Equation 10.9. Further heating
(C–D) produces superheated steam. At point D it is at 250 °C and has an enthalpy of
2800 kJ kg− 1.

Fig. 10.1. Pressure–enthalpy diagram for water: Hc = enthalpy of condensate; Hv =-


enthalpy of saturated vapour; Hs = enthalpy of superheated steam (from Straub and
Scheibner 1984, with kind permission of Springer Science and Business Media).

10.9

where Vs (m3 kg− 1) = specific volume of steam, xs (%) = steam quality, V1 (m3 kg− 1-
) = specific volume of liquid and Vv (m3 kg− 1) = specific volume of vapour. The data
summarised in Fig. 10.1 is also available as steam tables (Keenan et al. 1969), and
selected values are shown in Table 10.4 (‘steam’ is another term for hot water vapour).

Table 10.4. Properties of saturated steam

Temperature (°C) Vapour pressure Latent heat (kJ kg- Enthalpy (kJ kg− 1) Specific volume
(kPa) − 1) (m3 kg− 1)
Liquid Saturated vapour Liquid Saturated vapour
30 4.246 2431 125.79 2556.3 0.001 004 32.89
40 7.384 2407 167.57 2574.3 0.001 008 19.52
50 12.349 2383 209.33 2592.1 0.001 012 12.03
60 19.940 2359 251.13 2609.6 0.001 017 7.67
70 31.19 2334 292.98 2626.8 0.001 023 5.04
80 47.39 2309 334.91 2643.7 0.001 029 3.41
90 70.14 2283 376.92 2660.1 0.001 036 2.36
100 101.35 2257 419.04 2676.1 0.001 043 1.67
110 143.27 2230 461.30 2691.5 0.001 052 1.21
120 198.53 2203 503.71 2706.3 0.001 060 0.89
130 270.1 2174 546.31 2720.5 0.001 070 0.67
140 316.3 2145 589.13 2733.9 0.001 080 0.51
150 475.8 2114 632.20 2746.5 0.001 091 0.39
160 617.8 2083 675.55 2758.1 0.001 102 0.31
170 791.7 2046 719.21 2768.7 0.001 114 0.24
180 1002.1 2015 763.22 2778.2 0.001 127 0.19
190 1254.4 1972 807.62 2786.4 0.001 141 0.15
200 1553.8 1941 852.45 2793.2 0.001 156 0.13
250 3973.0 1716 1085.36 2801.5 0.001 251 0.05
300 8581.0 1405 1344.0 2749.0 0.001 044 0.02

Adapted from Singh and Heldman (2001b) original data from Keenan, J.H., Keyes,
F.G., Hill, P.G. and Moore, J.G., (1969), Steam tables metric units, Wiley, New York,
copyright John Wiley &amp; Sons

When a phase change from water to vapour occurs, there is a substantial increase
in the volume of vapour. In some unit operations, such as dehydration, this is not
important, but in freeze drying (Chapter 23, section 23.1) and evaporation (Chapter
14, section 14.1) the removal of large volumes of vapour requires special equipment
designs.

In steam production using boilers, the vapour produced by the phase change is con-
tained within the fixed volume of the boiler vessel and there is therefore an increase
in vapour (or steam) pressure. Higher pressures result in higher-temperature steam
(moving further right of the curve in the superheated vapour section of Fig. 10.1).
The required pressure and temperature of process steam are controlled by the rate
of heating in the boiler (see also section 10.2).

> Read full chapter

Physical Properties of Food Materials


Zeki Berk, in Food Process Engineering and Technology, 2009

1.3 Thermal Properties


Almost every process in the food industry involves thermal effects such as heating,
cooling or phase transition. The thermal properties of foods are therefore of consid-
erable relevance in food process engineering. The following properties are of partic-
ular importance: thermal conductivity, thermal diffusivity, specific heat, latent heat
of phase transition and emissivity. A steadily increasing volume of information on
experimental values of these properties is available in various texts (e.g. Mohsenin,
1980; Choi and Okos, 1986; Rahman, 1995) and electronic databases. In addition,
theoretical or empirical methods have been developed for the prediction of these
properties in the light of the chemical composition and physical structure of food
materials.

Specific heat cp (kJ.kg−1.K−1) is among the most fundamentals of thermal properties. It


is defined as the quantity of heat (kJ) needed to increase the temperature of one unit
mass (kg) of the material by one degree (°K) at constant pressure. The specification
of ‘at constant pressure’ is relevant to gases where the heat input needed to cause
a given increase in temperature depends on the process. It is practically irrelevant
in the case of liquids and solids. A short survey of the methods for the prediction
of specific heat is included below. Most of the other thermal properties of foods are
discussed in detail in Chapter 3, dealing with transport phenomena.

The definition of specific heat can be formulated as follows:

(1.2)

The specific heat of a material can be determined experimentally by static (adiabatic)


calorimetry or differential scanning calorimetry or calculated from measurements
involving other thermal properties. It can be also predicted quite accurately with the
help of a number of empirical equations.

The simplest model for solutions and liquid mixtures assumes that the specific heat
of the mixture is equal to the sum of the pondered contribution of each compo-
nent. The components are grouped in classes: water, salts, carbohydrates, proteins,
lipids. The specific heat, relative to water, is taken as: salts=0.2; carbohydrate=0.34;
proteins=0.37; lipids=0.4; water=1. The specific heat of water is 4.18 kJ.kg−1.K−1. The
specific heat of a solution or liquid mixture is therefore:

(1.3)

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:

(1.4)

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):
(1.5)

> Read full chapter

Heat Transfer in Food Processing


R. Paul Singh, Dennis R. Heldman, in Introduction to Food Engineering (Fifth
Edition), 2014

Heat transfer is ubiquitous in food processing. Heat exchangers are commonly used
for the purposes of heating and cooling foods. In designing heating or cooling
systems, thermal properties of foods and food contact materials are required. The
key thermal properties include specific heat, thermal conductivity, and thermal
diffusivity. Heat exchange between a heating or cooling medium and food occurs
by conduction, convection and/or radiation. Mathematical expressions are useful in
determining the rate of heat transfer and for designing process equipment. Heat
transfer calculations are conducted for steady state and unsteady state conditions.
Most common shapes of heat exchangers used in food processing are either tubular
or plate. Fouling is a common occurrence in thermal operations and its impact
on heat transfer can be determined with appropriate expressions. Dimensionless
relationships involving Biot number, Prandtl number, and Fourier number are used
in determining unsteady state heat transfer. Understanding mechanisms that are
involved in heating foods in a microwave field is necessary to develop novel foods
for microwave applications.

> Read full chapter

MEAT | Preservation
D.A. Ledward, in Encyclopedia of Food Sciences and Nutrition (Second Edition),
2003

Product Cooling
Refrigerated meats are generally classified as chilled or frozen. However, there
may be several steps involved in reducing a hot product to chill or frozen storage
temperature, with several additional links before final consumption. The design and
operation of equipment to perform these functions require an understanding of
the thermal properties of foods and an appreciation of their complexity and that of
prevailing legislation.
Immediately after heat treatment or cooking, it will be necessary to commence
cooling under controlled conditions. Filtered ambient air or mains water are suitable
media for cooling the product temperature to about 35°C. Naturally, the latter would
only be used where there is a hermetically sealed skin packaging. These represent
cheaper energy sources than mechanical refrigeration or expendable refrigerants,
which are required for achieving the statutory temperatures, which lie below 10°C.

> Read full chapter

Neural network method of modeling


heat penetration during retorting
C. Chen, in In-Pack Processed Foods, 2008

Publisher Summary
Conventional thermal processing can be divided into two types: retort processing
and aseptic processing. The retort processing method is one of the most mature
processing technologies. In retort thermal processing, the heat is transferred by
conduction and/or convection from the heating medium to the food, depending
on the type of foods being processed. The temperature inside the food during
heating will be determined by a variety of processing conditions, including the
type of heating medium and its temperature, initial product temperature, thermal
properties of food being heated, and rheological properties for liquid foods. The-
oretically, it is possible to apply a mathematical modeling method combined with
modern computation techniques for the simulation of thermal processing of solid
or particulate liquid foods, provided all the processing conditions can be discovered
and all the thermo-physical properties of the food obtained by independent experi-
ments. However, the biggest challenge that food modeling researchers are facing
is that, unlike other engineering materials, food materials have variable thermal
and/or physical properties, most of which are temperature and processing time
dependent. This means that it is very difficult to discover the properties and their
changes with processing temperature and time under conditions simulating the real
processes. In recent years, Artificial Neural Networks (ANNs) have opened alternative
pathways for modeling of complex and nonlinear processes. The advantages of
ANNs over conventional mathematical methods in modeling performance have
been recognized and confirmed by many research reports. This chapter focuses on
an introduction to the basic principles of neural networks, the development of neural
network models, and their application advances in food thermal processing areas.
> Read full chapter

Electrical properties
M.J. Lewis, in Physical Properties of Foods and Food Processing Systems, 1996

12.4 ELECTRICAL ENERGY


When a current flows through a potential difference, a quantity of energy is dis-
sipated in the form of heat. This is used to define the unit of potential difference
as follows. One volt (1 V) is the difference of electric potential between two points
of a conducting wire carrying a constant current of 1 A, when the power dissipated
between these points is equal to 1 W (Js− 1). Instruments used for measuring potential
difference are known as voltmeters.

The total amount of electrical energy E evolved when a current I flows over a potential
difference V for a time t is given by

The electrical power rating P is given by VI (W). In this way, electrical energy is
converted to thermal energy. Electrical methods are used for estimating thermal
properties of foods (see sections 8.5 and 9.7).

Thus an electric bulb rated at 60 W running off a power source of 240 V would carry
a current of 60/240 = 0.25 A. Energy would be dissipated at the rate of 60 J s− 1.

The amount of electrical energy used by the consumer is measured in terms of


kilowatt hours or units. The number of units used is equal to the product of the
power rating in kilowatts and the time in hours:

Thus a lamp rated at 60 W alight for 20 h would use (60/1000) × 20 = 1.2 units.
Currently the cost of electricity to the UK domestic consumer is 5.2p per unit.

The pricing system for industrial users is not so straightforward, being based on
the maximum demand as well as the total number of units. Electricity costs can be
reduced by trying to spread the load throughout the day, thereby ensuring that there
are no excessive peaks in electrical demand.

It is perhaps interesting to note that one unit of electricity is equivalent to l000


Js− 1 × 3600 s, i.e. 3.6 × 106 J (3.6 MJ) of energy. The latent heat of vaporization of water
at atmospheric pressure is 2.257 MJ kg− 1; so the evaporation of 1 kg of water vapour
by direct electrical methods would require 0.63 units of electricity and would cost
3.28p (assuming no heat losses and ignoring sensible heat changes). In contrast,
heating 1 kg of milk from 5 °C to 72 °C would require 1 × 67 × 4000 = 0.268 MJ and
would cost 0.387p.
Other expressions for electrical power can be derived from Ohm’s law:

Electrical power is measured using a wattmeter, the most common types being those
to measure AC power consumption for household supply.

> Read full chapter

Methodology
Yrjö H. Roos, Stephan Drusch, in Phase Transitions in Foods (Second Edition), 2016

3.4.1.2 DSC and DTA


DTA and DSC are closely related methods, which are probably the most common
techniques in the determination of phase transitions in inorganic, organic, polymer-
ic, and also food materials. DTA measures temperature of a sample and a reference
as a function of temperature. A phase transition causes a temperature difference
between the sample and the reference, which is recorded. DSC may use the same
principle, but the temperature difference between the sample and the reference is
used to derive the difference in the energy supplied. DSC may also measure the
amount of energy supplied to the sample and the reference. The use of DTA and
DSC in food applications was reviewed by Biliaderis (1983) and Lund (1983). DSC
and DTA in various modifications, such as modulated DSC and hyperheating DSC,
have been used to determine melting properties of sugars, lipids, and various other
thermal properties of food components and foods (Roos et al., 2013).

DTA and DSC are used to detect endothermal and exothermal changes that occur
during a dynamic measurement as a function of temperature or isothermally as
a function of time. The thermograms obtained show the heat flow to the sample
and DSC data can be used to calculate enthalpy changes and heat capacities.
First-order phase transitions produce peaks and a step change in heat flow occurs at
second-order transitions. As shown in Figure 3.10 thermograms showing first-order
transitions can be analyzed to obtain transition temperatures. The latent heat of
the transition is obtained by peak integration. Thermograms showing second-order
transitions can be used to derive transition temperatures and changes in heat
capacity as shown for glass transition in Figure 3.11. The transition occurs over
a temperature range of 10–30°C. Both the onset and midpoint temperatures of
the glass transition temperature range are commonly referred to as Tg. Enthalpy
recovery of an amorphous material after annealing around Tg may also be used to
indirectly estimate molecular mobility (Baird and Taylor, 2012).
Figure 3.10. A schematic DSC thermogram showing an endothermal, first-order
phase transition, for example, melting. The onset of the transition occurs at To, which
is the transition temperature. In broad melting transitions the peak temperature of
the endotherm, Tp, and the endset temperature, Te, may also be determined. To and
Te are obtained from the intercept of tangents drawn at the point at which deviation
from the baseline occurs. Peak integration is used to obtain the latent heat of the
transition, ΔHl.

Figure 3.11. Determination of second-order and glass transition temperatures, Tg,


and change in heat capacity, ΔCp, that occurs over the glass transition temperature
range from DSC thermograms. The endothermal step change in heat flow during
heating of glassy materials occurs due to ΔCp at the second-order transition tem-
perature.

Applications of DSC in the determination of phase transitions in foods include such


changes as crystallization and melting of water, lipids, and other food components;
protein denaturation and gelatinization; and retrogradation of starch. The samples
are usually placed in pans that can be hermetically sealed. Therefore, the method
can be used to observe phase transitions and to determine transition temperatures
without changes in water content. A constant water content is extremely important
in the determination of phase transitions of food materials. Water has an enor-
mous effect on transition temperatures, and its impact on food behavior cannot be
overemphasized.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like