food process engineering

79
Food Process Engineering An introduction by Alan Friis & Jørgen Risum BioCentrum-DTU Septemner 2002 First edition (0.5 version)

Upload: dmora1965183

Post on 26-Oct-2014

184 views

Category:

Documents


6 download

TRANSCRIPT

Page 1: Food Process Engineering

Food Process Engineering

An introduction

by

Alan Friis & Jørgen Risum

BioCentrum-DTU Septemner 2002 First edition (0.5 version)

Page 2: Food Process Engineering

Table of contents

Introduction ........................................................................................................ 1 Food Process Design and Quality................................................................... 1 A strategy for Process Design of a Food Production Line.............................. 1 About the teaching material........................................................................... 2

Flow sheets ........................................................................................................ 4 Process description......................................................................................... 4 Flow sheets..................................................................................................... 4

Mass and energy balances................................................................................. 6 Mass Balances ................................................................................................ 6

Material balance with accumulation ........................................................ 10 Energy balances ........................................................................................... 11

Continues systems.................................................................................... 11 Composition of food products .......................................................................... 16 Physical properties of food products................................................................ 18

Density .......................................................................................................... 18 Specific heat capacity................................................................................... 19 Heat conductivity.......................................................................................... 20 Thermal diffusivity ........................................................................................ 22 Viscosity (Newtonian liquids) ...................................................................... 22

The influence of temperature on viscosity................................................ 23 Other characterizing parameters ................................................................. 23 Rheology of liquid foods ............................................................................... 24

Non-Newtonian fluids ............................................................................... 25 Time dependency...................................................................................... 25 Time independent liquids ......................................................................... 26 Shear-thinning fluids................................................................................. 28 Shear-thickening fluids ............................................................................. 31 Fluids exhibiting an apparent yield stress ............................................... 31 The role of elasticity and time effects ....................................................... 32 Temperature dependency......................................................................... 32

Rheology solid and semi-solid food products............................................... 32 Fluid mechanics ............................................................................................... 33

Newtonian fluids .......................................................................................... 33 Laminar flow ............................................................................................. 35 Turbulent flow........................................................................................... 37

Non-Newtonian fluids................................................................................... 38 Laminar flow ............................................................................................. 40 Turbulent flow........................................................................................... 43

Flow through pipe fittings............................................................................ 43 Heat transfer..................................................................................................... 45

Steady state conduction heat transfer ......................................................... 46 Steady state convection heat transfer.......................................................... 49

Natural convection .................................................................................... 51 Forced convection ..................................................................................... 52

Unsteady state heat transfer ........................................................................ 56

Page 3: Food Process Engineering

- ii -

Cooking and condensation ........................................................................... 56 Kondenserende vanddamp ....................................................................... 56 Kogning ..................................................................................................... 56

Heat transfer by radiation ............................................................................ 57 Black body radiation ................................................................................. 57

Heating and cooling ......................................................................................... 58 Effects on processing ....................................................................................... 59 References ........................................................................................................ 60 List of symbols ................................................................................................. 61 Appendix 1. Usefull calculations and conversions .......................................... 63 Appendix 2. Units conversions ........................................................................ 64 Appendix 3. Density equations........................................................................ 65 Appendix 4. Heat capacity equations.............................................................. 66 Appendix 5. Heat conductivity equations ....................................................... 67

Beregninger af ikke homogene systemer ..................................................... 67 Appendix 6. Measuring viscosity and flow curves .......................................... 69 Appendix 7. Moody diagram............................................................................ 70 Appendix 8. Data for water, air + foods .......................................................... 71 Appendix 9. Equations for natural convection ................................................ 72 Appendix 10. Equations for forced convection ................................................ 73

Free falling droplets .................................................................................. 73 Drying........................................................................................................ 73 Agitated tanks........................................................................................... 73

Page 4: Food Process Engineering
Page 5: Food Process Engineering

- 1 -

Introduction Producing foods of high quality is an obvious objective for food engineers. Hence a methodology for objective determination of the relationships between the characteristics of food products and the involved manufacturing processes (including storage and transportation) is needed. The term ‘characteristic’ pertaining to food products is much more operational than the term ‘quality’ when dealing with process design and validation. Characteristics can be determined objectively (quantitatively or qualitatively), however, definition of quality is subjective. Two persons who like a specific product might like it for different reasons. Hence each individual define his or her own preferred characteristic to determine the quality of the product.

Food Process Design and Quality Some definitions relating to quality and characteristics of foods is given below:

§ Food companies producing high quality foods will first assure safety (microbiological, chemical and physical) of their products and next focus on uniform properties and characteristics over time. This, however, does not define the quality level. The quality of a product must comply with generally understood perceptions amongst consumers in a specific area. Hence we find that quality conveniently can be defined as what the consumers wants or what the consumers are made to believe they want and are willing to pay for.

§ In the consumer opinion quality is linked to aspects like nutritional content, physical appearance, consistency, structure, taste, safety issues, convenience, ecology, minimal, careful production, etc. For food producers the perceived quality must be translated into measurable characteristics to be able to design products, processes and control systems in a proper manner.

§ In the field of food process engineering we focus on characteristics of food products. We appreciate the total quality of a food product pertains to a specific situation and as such is seen in a subjective light.

A strategy for Process Design of a Food Production Line The toolbox needed when designing food process lines and determining sensible operating conditions origins from the general field of process engineering. The necessary calculations tools are presented in this book in a way, which pertains specifically to food production.

It can be beneficiary to begin whit a procedure of designing processes in order relate to the application of the theoretical subjects. Such a strategy is presented below:

1. Deliverables. Develop a product specification data sheet and assign preliminary limits to measurable characteristics

Page 6: Food Process Engineering

- 2 -

2. Raw Materials. Identifying raw material (quality parameter) requirements. Identify the key transformations and processes operations needed to produce the product from these materials. Some interactions between the product and the process operations may be identified in this step as well

3. Flow Sheets. Draw a flowsheet to illustrate the flow of different material streams through the process. Heating and cooling media streams should be included on the flowsheet. Remember that different flow sheets are produced to serve different purposes. While we rarely put all information in one sheet we need one concerning flow of material through the process and one concerning temperature and energy as a minimum

4. Mass Balances. Choose a basis and calculate preliminary mass balance estimates for all streams. These calculations can be refined later as the design process proceeds

5. Energy Balances. Energy balances determine the requirements for heating, cooling and mechanical power. Initial energy balance calculations are performed operation by operation. However, reuse or regeneration of energy must be considered in later to design an economically feasible process

6. Design Refinements. Refine the flow sheet if needed. The level of detail should at this point show each unit operation in separate steps. Refine mass and energy balances if needed

7. Process Selection based on Product-Process Interactions. Identify key interactions between product and process. Evaluate potential processes to produce the desired product. More processes may be suitable in theory; this stage ends with the choice of a suitable process made on a qualified basis

8. Equipment Sizing and Selection. For each part of the process, estimate the size of equipment you require. Calculate pumping, heating and cooling requirements

9. Capital Costs. Estimate the purchase costs (first costs) of your process line

10. Operating Costs. Estimate how much it costs each process to run (second costs)

11. Solution Optimization. Repeat steps 6 to 10 until a desired outcome is achieved

12. Communicate the result. Present the result and confirm the next steps.

The procedure listed describes an intuitive way to attack the design of a production line and will be the basis we often refer to throughout the food process engineering curriculum.

About the teaching material The text is divided in chapters pertaining to Flow Sheets (of processes), Mass and Energy Balances, Composition of Food Products, Physical Properties of Food (including Rheology), Fluid Mechanics, Heat Transfer, Heating and Cooling and Effects of Processing.

Page 7: Food Process Engineering

- 3 -

The text aims at explaining the theoretical information needed to perform process engineering calculations in a practical applicable way. The objective is to enable the student to be able to validate and design food process operations. The examples given in the text is designed to support this objective.

When doing calculus it is sensible to bring all quantities to SI units before computation is carried out (do always check that the dimensions of equations are correct). In this text SI units are stated for all of the applied physical parameters.

Page 8: Food Process Engineering

- 4 -

Flow sheets

Process description

Describing a food production process and the plant in which it takes place is a somewhat complicated task. Several questions are relevant to ask before starting the task.

o What is the purpose of the description?

o Who is going to use it?

o In how much detail do you need to describe it?

The purpose of the description may be to:

o inform a government office about your production in relation to an application for increasing or altering the production.

o inform the board of directors about a possible enlargement of the facilities.

o give an entrepreneur information in relation to a change in a production line.

o perform a critical analysis of the production facility with respect to wastewater and energy consumption.

The four examples require different kinds of information and the level of detail will be different. Regardless of purpose the first thing to do is to present the overall plan of production. This is best done with a flow sheet or block diagram. However, this may be very simple at first. Basically all we need is a recipe for the product as is used in normal households and a production procedure, but with the exception that the apparatus and amount of materials are much different.

Flow sheets A flow sheet is but a graphical representation of a process. The purpose is to clarify different streams to and from the processes in the production plant. The simplest flow sheet is perhaps: Raw materials Product However, it is not very informative. It indicates the need for raw materials to make the product. Further specifications will enlighten us somewhat more: Culture Raw milk skim milk Cream

Processing

Separation Fermentation Packing of ’buttermilk’

Page 9: Food Process Engineering

- 5 -

It should be recognised that the more detailed the flow sheet becomes and the more information we put in we risk loosing the comprehension of the process. Therefore a good principle is KISS:

Keep It Simple Stupid The information needed must be presented, but be critical, only include what is needed. We will give you an example of a superfluous description.

Example 1. A first flow sheet

The task is to produce a (continental) breakfast at 7 o’clock in the morning

Menu: Orange juice

Coffee Buttered toast and marmelade

Yoghurt Breakfast cereals and milk.

What do we need to know?

o How to produce orange juice from oranges? No!

o How to brew coffee? Yes! Because it this case we have a process. Water 8°C, 800 ml

40 g coarse milled roasted coffee beans coffee-grounds (waste)

650 ml 80°C coffee

Similarly we need to define how the bread is toasted, but not how to produce the bread (if we are buying it). Again we do not need to know how to milk a cow.

Boil water 100°C

Extraction, 5 min

Page 10: Food Process Engineering

- 6 -

Mass and energy balances

Mass Balances

The mass balance helps us to make bookkeeping concerning streams passing through a production line much easier.

Introduction example

The company Panada Aps are producing breaded products like fish sticks. You are going to introduce a new production of Turkey Cordon Bleu.

You are planning a series of analysis (1000) for the content of glucose. You are going to use the method applying Glucoseoxidase.

You are employed on a poultry processing plant and are given the task of informing the environmental authorities of the environmental impact caused by the production.

What is the common denominator of these three tasks?

Comments: They are complex and involve numerous different sub-processes. The solution requires knowledge of the amount of different materials being used or produced and the processes involved. There may even be a necessary sequence of processes in time.

How would you attack the problems?

Draw a flow sheet to illustrate the overall process. Include the needed information.

The first example requires at least the recipe for the product.

The second example requires knowledge about amounts of chemicals, temperatu-res and times needed and environmental conditions for disposal of waste.

The third example implies very complicated overviews as it concerns streams of different components (live chicken, water, feathers, blood, offal, waste, waste water and finished product) as well as energy consumption for scalding, chilling, freezing etc (electric power, gas, oil etc.).

When we are making an accounting for such tasks we use the laws of conservation of mass and constant energy.

For a process in steady-state we have:

IN = OUT

What comes in must come out.

If the process is unsteady (or transient) we may observe an accumulation of some materials:

IN = OUT + ACCUMULATED

or when a component is produced or consumed:

IN + PRODUCED = OUT

Page 11: Food Process Engineering

- 7 -

In generic terms we get:

IN + PRODUCED = OUT + ACCUMULATED

Correspondingly energy cannot just disappear. It may be converted to other kinds of energy (mechanical, thermal etc.) as we ignore the more exotic such as nuclear and quantum physical phenomena.

These observations are valid for absolute measures (mass) as well as for streams (mass per unit time etc.)

The generic procedure for establishing a mass balance goes:

1. Choose a basis (eg. 100 kg of raw material or product) 2. Draw the flow sheet with all in-and out-going streams. 3. Identify the streams of known size. 4. Mark the streams of unknown amounts 5. Identify the relationships between different streams 6. Construct the equations describing the relationships 7. If different units are used, transform to one and only one unit. 8. Identify that the established relations are independent. 9. Determine the number of variables (nv) 10. Determine the number of independent equations (ne) 11. Determine the degrees of freedom (df = nv-ne) in the system 12. If df = 0 solve the system of equations 13. Calculate the numeric value of necessary variables 14. Validate the solution. It must to be reasonable.

Ad. 1 The chosen basis may be a stream (kg per time)

Ad. 2 It is very important to define the boundaries of the system clearly.

Ad. 5 & 6 State equations and inequalities relating the variables to one another. In the ideal case it should give us a system of linear equations.

Ad. 12 If df = 0 there is only one solution to the set of equations. If df < 0 the system is over-determined and we have to evaluate the relations established. Some of the information may be redundant or more equations are linear combinations of each other. If df > 0 we may find a solution by assigning definite values to some of the variables, i.e. we are seeking a feasible combination of outcomes. This situation is sometimes preferred to the situation with only one solution. However, we might want to fix some variables.

Ad. 14 The validation is necessary as negative masses are not allowed, but may occur in the solution. In such a situation the system of equations and constraints has to be revised.

Energy balances are established in exactly the same way, including heat capacities etc. and a lot of problems require a simultaneous solution of both.

Page 12: Food Process Engineering

- 8 -

Please note that the set of equations and constraints will often be so complex that a manual solution becomes difficult. In spreadsheet programs like Excel and QuatroPro the function Solver may be used to solve such problems.

Example 2. Mass balance – salted butter

Butter produced continuously is unsalted. To produce normal salted Danish butter salt is added as brine. Determine the amount of brine that has to be added and the allowable water content of the unsalted butter.

Data: Brine 60 g salt per 100 g solution, salted butter: 15 g water per 100 g butter, 1,4 g salt per 100 g butter.

Solution: The basis is selected as 100 kg salted butter.

Indices: b: brine, f: fat, s: salt, sb: salted butter, ub: unsalted butter, w: water

Assumptions: Fat is only part of the unsalted butter and salt is only part of brine.

Constrains: All mass fractions must be > 0.

The sum of mass ratios must be unity (here for salted butter): (a) xsb,w + xsb,s + xsb,f = 1

and for brine : (a’) xb,w+xb,s = 1 and for unsalted butter : (a’’) xub,w + xub,f = 1 Total mass balance: (b) Mb+ Mub = Msb Balances for components: Salt: (c) Mb*xb,s = Msb* xsb,s

Fat: (d) Mub,f *xub,f = Msb * xsb,f Water: (e) Mb*xb,w + Mub*xub,w = Msb*xsb,w

Above the known quantities are marked with bold face.

(a) => xsb,f = 1 – 0.155 – 0.014 = 0.831 (a’) => xb,w = 1 – 0.60 = 0.40 (c) => Mb= (100*0.014)/0.60 = 2.333 kg brine (c) in (b) => Mub = 100 – 2,3333 = 97.667 kg unsalted butter in (d) => xub,f = (100*0.831)/97.667 = 0.851 g fat per g unsalted butter and (a’’) => xub,w = 1 – 0.851 = 0.149 g water per g unsalted butter

Per 100 kg salted butter we mix 97.67kg unsalted butter with 2.33kg 60% brine. The composition of unsalted butter (per 100g) must be 0.851g fat and 0.149g water.

Comment: This example shows the general principle. It may be solved directly by identifying the simple information and combining. The problem is a little peculiar as it puts strong limitations on the composition of unsalted butter. The natural choice would be to adjust the brine to suit the composition of the unsalted butter. Actually it is the same equations, but other unknowns. Note that equation (e) was never used. It is a linear combination of (c) and (d) combined with the (a)’s.

Page 13: Food Process Engineering

- 9 -

Example 3. Mass balance – raw milk separation

Data: Raw milk is to be separated in skimmed milk and cream in a centrifuge. The problem is to determine how much cream and skimmed milk and of what composition is produced under given demands on the composition of the cream.

Solution: The basis is chosen to be 100kg raw milk.

Indices: c: cream, f: fat, l: lactose (carbohydrates), p: protein, nf: non fat dry matter, rm: raw milk, s: salts, sm: skimmed milk, w: water.

Assumptions: All components are water soluble except the fat, ie. that the cream may be considered as a system composed of skimmed milk and fat. The composition of the aqueous phase is constant.

Constraints: All sums of mass fractions must equal unity and all mass fractions must be zero or positive.

Equations:

Definition of non fat dry matter: (a) xnf = xp + xl + xs

Sum of mass fractions: (b) xrm,nf + xrm, f + xrm,w = 1

(c) xc,nf + xc,f + xc,w = 1

(d) xsm,nf + xsm,f + xsm,w = 1

Total mass balance: (e) Mrm = Msm + Mc

Balances for components:

Fat: (f) Mrm * xrm,f = Msm * xsm,f + Mc * xc,f

Non fat dry matter: (g) Mrm * xrm,nf = Msm * xsm,nf + Mc * xc,nf

Water : (h) Mrm * xrm,w = Msm * xsm,w + Mc * xc,w

Constant concentration in aqueous phase :

(i) , , ,

, , ,

rm nf sm nf c nf

rmw smw c w

x x xx x x

= =

A common assumption in many textbooks is to consider the skimmed milk to be totally fat free. This gives us a convenient system of linear equations, but as we do know that skimmed milk is not void of fat it is a gross error (and the system is not generic any more). Introducing the assumption of constant composition of the aqueous phase complicates the system. It is no longer a set of linear equations. Solutions must be found by iterative methods. Fortunately this is a simple task in spreadsheet programs like Excel and Quatro Pro using the Solver option. But care must taken in validating the results. Below is shown an example of a solution.

Page 14: Food Process Engineering

- 10 -

Cream separation Composition Raw milk Skim milk Cream Lactose Protein Salts Non fat dm 0.077 0.080 0.052Fat 0.045 0.003 0.350Water 0.878 0.916 0.598Mass 100 87.98 12.02NF const 0.0877 0.0877 0.0877 Target 100 Constrains Mass fractions = 1, SM 1.000 = 1Mass fractions = 1, C 1.000 = 1Mass fractions >= 0, SM, nf 0.080 >= 0Mass fractions >= 0, SM, fat 0.003 >= 0Mass fractions >= 0, SM, water 0.916 >= 0Mass fractions >= 0, C, nf 0.052 >= 0Mass fractions >= 0, C, fat 0.350 >= 0Mass fractions >= 0, C, water 0.598 >= 0Mass fractions <= 1, SM, nf 0.080 <= 1Mass fractions <= 1, SM, fat 0.003 <= 1Mass fractions <= 1, SM, water 0.916 <= 1Mass fractions <= 1, C, nf 0.052 <= 1Mass fractions <= 1, C, fat 0.350 <= 1Mass fractions <= 1, C, water 0.598 <= 1Constant conc. In aq. SM 0.0877 = 0.0877Constant conc. In aq. C 0.0877 = 0.0877Demand on a component eg fat in C 0.350 = 0.350 Mass < total SM 88.0<= 100Mass < total C 12.0<= 100 Mass balance fat 0.00000 = 0Mass balance nf 0.00000 = 0Mass balance water 0.00000 = 0

Material balance with accumulation

Drying of foodstuffs in a cabinet dryer will often start with a period with constant drying rate (Rc kg water/(m2 s)). Describing this is simple :

IN = OUT

Ma* Xa,w,i = Ma * Xa,w,o

= Ma * Xa,w,i + Rc * A

Page 15: Food Process Engineering

- 11 -

The condition being stable until the drying rate changes. Accumulation of water in the system is negative as water is disappearing.

Energy balances Setting up an energy balance requires knowledge of the process and the possible losses. If we assume no loss to the surroundings (adiabatic conditions) the heat balance of a heat exchanger can be expressed as :

IN = OUT

MI * HI,in + MII * HII,in = MI * HI,out + MII * HII,out

MI* cp,I* TI,in + MII* cp,II* TII,in = MI* cp,I* TI,out + MII* cp,I* TII,out

This holds for a closed system, but for systems with changes of phase it becomes a little bit more complicated as in drying. Considering the cabinet dryer in steady state we have an energy consumption for the evaporation of water (∆Hevap) and we have to make use of the fact that enthalpy is at state function defining our basis as liquid water at 0°C. The enthalpy balance becomes (Y is kg water per kg dry air):

IN = OUT

Ma * Ha,in = Ma * Ha,out

( )( )

, , , , ,

, , , , ,

a p a a in p v a in a in evap

a p a aout p v aout aout evap

M c X c T X H

M c X c T X H

⋅ + ⋅ ⋅ + ⋅∆= ⋅ + ⋅ ⋅ + ⋅∆

This is only useful if we know Xa,out, because we then may calculate Ta,out. Fortunately we have an expression from the materials balance:

Ma* Xa,in = Ma * Xa,in + Rc * A

Combining the two gives us the solution:

Xa,out = Xa,in + Rc * A/Ma

Continues systems

An example of a continuous process is spray drying. Liquid milk is introduced to the dryer together with hot air. Leaving the dryer is moist air and a milk powder with some water. The balance for both mass and heat is given below based on dry matter streams (i.e. dry matter and dry air).

Notation :

Ms: solids flow (kg dry matter per s) Y: moisture in solid (kg water per kg dry matter)

Ma: air flow (kg dry air per s) X: moisture in air (kg water per kg dry air)

Indices: p: product, w: water, s: solids, a: air, v: water vapour

IN = OUT

Page 16: Food Process Engineering

- 12 -

Mass balance for water:

Ms * Yin + Ma * Xin = Ms* Yout + Ma * Xout

Energy balance:

Ms * Hp,in + Ma * Ha,in = Ms* Hp,out + Ma * Ha,out + losses

Hp,in = (cp,s + Yin * cp,w)*∆Tp,in

Hp,out = (cp,s + Yout * cp,w)*∆Tp,out

Ha,in = (cp,a + Xin * cp,v ) * ∆Ta,in + Xin * ∆Hevap

Ha,out = (cp,a + Xout * cp,v ) * ∆Ta,out + Xout * ∆Hevap

Example 4. Energy balance for a spray drying process

In a plant for producing spray dried milk. The capacity of the plant is 5,000 kg skimmed milk per hour. The plant engineer is worried about his energy consumption (measured to ca. 625 kW) in the spray-drying operation and would like to validate his measurements by theoretically estimating the necessary mass flow rate of air and energy supplied to the dryer.

Data: The composition of skimmed milk is (g per 100 g) : protein 3.5, lactose 6, salts 0.5, rest is water. The skimmed milk is concentrated by evaporation to 45 g dry matter per 100 g concentrate and are then dried to 4 g dry matter per 100 g product in the spray-dryer. Density of concentrated skim milk is 1040 kg/m3. Air : Cp,a = 1,00 kJ/(kg K), water : Cp,w = 4,18 kJ/(kg K), water vapour : Cp,vap = 1,88 kJ/(kg K),

Measurements: The concentrated milk has a temperature of 60°C when entering the spray-dryer and the dried milk leaves at 70°C and are cooled to 30°C in a special packer/dryer unit. The air is 20°C, with a humidity of 0.010 g water per g dry air (approx. 70% RH). The air is heated to 200°C before entering the spray dryer. Moist air leaves the dryer at 80°C.

Assumptions: The first assumption is that the dryer is operating adiabatically, i.e. no loss to the environment.

Solution: Calculate the thermophysical data for the dry matter assuming no temperature dependence:

1, but 0,035+0,06+0.005 = 0.10 so we correct it ix =∑

0.035/0.1 0.35

0.060/0.1 0.600.005/0.1 0.05

protein

lactose

salt

x

xx

= == =

= =

, 2.0 0.35 1.1 0.60 1.1 0.05 1.72 /( )p solidC kJ kg K= ⋅ + ⋅ + ⋅ = ⋅

Page 17: Food Process Engineering

- 13 -

The moisture content expressed on dry basis (indices s for solids) :

1 1.00 0.451.222

0.45s

ins

xY kg water per kg dry solid

x− −= = =

0.040.042

1 1.00 0.04w

outw

xY kg water per kg dry solid

x= = =− −

The mass flow of skimmed milk expressed as dry solids flow :

5000

0.45 0.139 sec3600sM kg solids per ond= ⋅ =&

The materials balance for water :

IN = OUT

s in a in s out a outM Y M X M Y M X⋅ + ⋅ = ⋅ + ⋅& & & &

0.139 1.222 0.010 0.139 0.042a a outM M X⋅ + ⋅ = ⋅ + ⋅ ⇒& &

0.1640.010out

a

XM

= +&

The enthalpy balance :

IN = OUT

, , , ,s s in a a in s s out a a outM H M H M H M H⋅ + ⋅ = ⋅ + ⋅& & & &

, , , ,( ) (1.72 1.222 4.18) (60 0)

409.7 / s in p s in pw s inH C Y C T

kJ kg dry solids

= + ⋅ ⋅∆ = + ⋅ ⋅ −=

, , , ,( ) (1,72 0.042 4.18) (70 0)

123.2 / s out p s out pw s outH C Y C T

kJ kg dry solids

= + ⋅ ⋅∆ = + ⋅ ⋅ −=

, , , ,( )

(1.00 0.010 1.88) (200 0) 0.010 2500 228.8 /

a in p a in pvap a in in evapH C X C T X H

kJ kg dry air

= + ⋅ ⋅∆ + ⋅∆= + ⋅ ⋅ − + ⋅=

, , , ,( )

(1.00 1.88) (80 0) 250080 2650.4 /

aout p a out pvap a out out evap

out out

out

H C X C T X H

X XX kJ kg dry air

= + ⋅ ⋅∆ + ⋅∆= + ⋅ ⋅ − + ⋅= + ⋅

0.139 409.7 228.8 0.139 123.2 (80 2650.4)a a outM M X⋅ + ⋅ = ⋅ + ⋅ + ⋅& &

0.16456.95 228.8 18.45 80 0.010 2650.4a a

aM M

M

+ ⋅ = + ⋅ + + ⋅ & & &

Page 18: Food Process Engineering

- 14 -

56.95 228.8 18.45 80 434,67 26.5a a aM M M+ ⋅ = + ⋅ + + ⋅& & &

434.67 18.45 56.95 396.173.24 sec

228.8 80 26.5 122.3aM kg dry air per ond+ −= = =− −

&

0.1640.010 0.061

3.24outX kg water per kg dry air= + =

Comments: The plant engineer happily confronts the plant manager with the results : mass of dry air needed is 3.24 kg per second and the exhaust air is humid with 0.061 kg water per kg dry air. The plant manager answers : ‘and so what. Explain to me is the plant running all right or …’

The analysis has just provided us with two figures but we have to elaborate a little bit more.

The needed energy input can be calculated to:

, ,( ) ( )

3.24 (1.00 0.010 1.88) (200 20) 594

a p a in pvap in envQ M C X C T T

kW

= ⋅ + ⋅ ⋅ −= ⋅ + ⋅ ⋅ −=

& &

The measured value was 625 kW so the loss would be approx. (625-594) = 31 kW or 5% of the input energy. Which is not bad.

The use of this energy should be as efficient as possible. Expressed as how close to the theoretical heat of evaporation are we in specific energy consumption. The theoretical ∆Hevap,65°C = 2346 kJ/kg. The mass of water evaporated is calculated as:

( ) 0.139 (1.222 0.042)0.164 sec

s in outW M Y Ykg water evaporated per ond

= ⋅ − = ⋅ −=& &

The theoretical specific energy consumption:

594

0.164 3,622 /

QSpecific energy consumption

WkJ kg water evaporated

⋅= ==

&

This is +55% of the theoretical. An enormous loss, but perhaps accountable for :

The enthalpy of the air leaving the system is:

, (1.00 0.061 1.88) 80 0.061 2500 89.17 152.5 241.7 /aoutH kJ kg= + ⋅ ⋅ + ⋅ = + =

∆Ha = 241.7-228.8 = 12.9 kJ/kg => 41.8 kJ/s

∆Hs = 123.2-409.7 = -286.5 kJ/kg => -39.8 kJ/s

The energy balance is not perfect. We need to account for 2 kJ/s or 5%. The data is assumed to be perfect, but are they are not and our assumptions are not perfect. The temperatures are measured, but how accurately. If we refined all data taking into account the temperature dependence we might get a better estimate. BUT

Page 19: Food Process Engineering

- 15 -

with this rough calculation a discrepancy of less than 5% is not at all bad, taken into consideration how little effort it takes.

The + 50% energy consumption compared to the theoretical is caused by having an exhaust temperature of 80°C. The heat in this stream is lost to the surroundings. Lowering the exhaust temperature could give us some serious problems with mist because of super saturation of the exhaust air.

All in all his plant manager should be happy as they have very low losses.

Page 20: Food Process Engineering

- 16 -

Composition of food products The composition of food products must be quantified before physical properties of food products can be determined. Hence we need to identify ways of determining the basic constituents of foods.

The calculation of declarations of nutrition is often based on knowledge concerning the composition of the ingredients. Usually this is not solved by designing experiments but by looking up known values in tables like:

o USDA (US department of Agriculture) Handbook no. 8 (USDA), which is one of the largest www.nal.usda.gov/fnic/cgi-bin/nut_search.pl. This collection contains approx. 6000 entries.

o Souci.Fachmann.Kraut, Food Composition and Nutrition Tables 1986/87 (SFK), 3.ed, Wissenschaftliche Verlagsgesellschaft mbH, Stuttgart, 1986, ISBN 3-8047-0833-1.

o The danish Levnedsmiddelstyrelsens Levnedsmiddeltabeller, 1989 (LST). These can also be found in the DanKost programme (Slank, Sport), where composition of meal can be calculated.

The composition of food products is the basis of estimating important parameters for engineering calculations: density, specific heat capacity and heat conductivity.

Example 5. Calculation of composition of a poultry pâté

Data: The recipe for production of a poultry pâté is as follows per 100kg: *fat from hen or pork 22.2 kg *Chicken liver 29.6 kg *Machine deboned chicken meat 14.8 kg *Boiling water 22.2 kg Caseinat (HV) 2.03 kg *Wheat flour 2.03 kg *Skim milk powder 2.54 kg *Raw onion 2.23 kg *Cocking salt 0.93 kg *Sugar 0.15 kg Spices 0.27 kg White pepper 0.17 kg *Anchovies 0.85 kg The ingredients marked * can be found directly in LST (levnedsmiddeltabellerne). The caseinat is declared from the producer: 5% water, 85% protein, and 10% salt.

Solution: The spices are assumed not to have any significant influence on the result. The calculations are conducted based on the content of water, protein, fat, carbohydrate and ash, per 100 kg product.

Page 21: Food Process Engineering

- 17 -

The result is presented below, where the x is the mass fraction of each basic component:

Database xw xprotein xfedt xkulhydrat xaske

SFK 0.600 0.136 0.207 0.028 0.021

LST 0.601 0.128 0.183 0.033 0.020

USDA 0.622 0.114 0.207 0.041 0.020

Comments: There some minor deviation concerning the water content between the two European databases and the American. Further differences are most certainly due to the fact that the product is slightly different (especially the anchovies). The influence of the differences on calculation of the physical properties will be shown.

Example 6. Calculation of energy content

Data: To give an example of practical use of the data presented in Example 5 we calculate the nutritional energy of the poultry pâté. This is based on the following energy content in the basic constituents:

Protein 17 kJ per gram Fat 38 kJ per gram Carbohydrate 17 kJ per gram

Solution: The energy content (values per 100g): The calculation based on data from SFK is given as example:

Total energy=17*0.136+38*0.207+17*0.028 = 10.65kJ (1065kJ per 100g)

SFK LST USDA

Total energy kJ per 100g 1065 970 1050

Energy percent:

Protein 22 21 18

Fat 74 72 75

Carbohydrate 4 6 7

Comment: Differences are observed between the databases. For most practical applications these differences will not influence the final result much, however, we suggest that data is taken from the database most suitable for the specific situation. Such judgement will often be based on geographical origin.

Page 22: Food Process Engineering

- 18 -

Physical properties of food products Food process engineering calculations depends on the availability of reliable information on the physical properties of foods. Important properties are density, specific heat capacity, thermal conductivity, thermal diffusivity, viscosity and rheological behaviour. It should be realised that we determine the physical properties on a need basis and that a complete characterisation rarely is needed for the process engineer, however such characterisation might be needed for other purposes (e.g. product development).

Density For the purpose of engineering calculations we need a reasonable estimate of the density ρ [kg/m3] of food products, however, we do not need very high accuracy.

This means that we should only invest little time in this task and quickly moves on to more relevant issues of process calculations.

In cases where the food product in question has a high water content the density of water being 1000kg/m3 can be applied. If one wishes a more accurate value equation (1) below can be applied using the values of the density of the basic constituents at 20°C listed in Table 1.

ρ =ρ∑

1i

i

x (1)

In equation (1) the subscript i denotes the different constituents and x is the mass fraction of each constituent. Remember that the sum of ix must always equal 1.

Table 1. Density of basic food constituents at 20°C

Component Density [kg/m3]

Protein 1320 Fat 920

Carbohydrate 1600 Fibre 1300

Ash 2420 Water 1000 Ice 910

The values stated above are rounded of to the nearest multiply of 10 for the ease of calculations. For reference to the relationships between the density of the basic food constituents and the temperature please refer to appendix 3.

Page 23: Food Process Engineering

- 19 -

Example 7. Determining density

Data: We utilise the data earlier presented for the poultry pâté.

Solution: Mass fractions from Example 5 and densities from Table 1 are inserted in equation (1) using data from SFK in the calculation example below:

ρ = = ≈ + + + + ρ∑31 1

1048 /0.600 0.136 0.207 0.028 0.0211000 1320 920 1600 2420

i

i

kg mx

The result from LST is ρ = 1080kg/m3 and USDA is ρ = 1034kg/m3.

Comment: For process engineering calculations these differences are insignificant.

Specific heat capacity

Determination of the specific heat capacity pC [kJ/(kg⋅K)] can conveniently be

based on a set of empirical expressions each valid for specific situations. Of cause the heat capacity can be determined experimentally but the approach chosen here is that it is better to calculate a value that you know is a bit inaccurate than measure one which might be much more off.

Below in Table 2 some general mathematical equations for pC are given based on

the water content (1 )w dx x= − , where dx is the mass fraction of dry matter.

Table 2. Equations for the specific heat capacity

Product Equation

Product above the freezing point 1.8 2.4p wC x= + ⋅ (2)

Frozen product (all water is considered frozen, i.e. being ice)

1.8 0.3p wC x= + ⋅ (3)

Products with a high water content 0.837 3.349p wC x= + ⋅ (4)

Meat and fish (> 25g water pr. 100g product)

1.67 2.5p wC x= + ⋅ (5)

Fruit and vegetables (more than 50g water pr. 100g product)

1.67 2.5p wC x= + ⋅ (6)

A more detailed approach is based on the total composition:

unfrozen product: 1.6 2.0 2.0 1.1 4.2p c p f a wC x x x x x= + + + + (7)

frozen product: 1.6 2.0 2.0 1.1 2.1p c p f a wC x x x x x= + + + + (8)

Page 24: Food Process Engineering

- 20 -

The mass fractions of constituents (x) used in equation (7) and (8) are: cx for carbohydrates, px for proteins, fx for fat and ax for ash.

From the equations listed above it can be seen that the specific heat capacity increases with the content of water, where =, 4.182p wC kJ/(kg⋅K) at 20°C.

The formula stated here will be sufficient in most engineering calculations. Should more elaborate expressions be needed please refer to appendix 4.

Example 8. Determining specific heat capacity

Data: Again we refer to Example 5 for product data.

Solution: Our product the poultry pâté has a water content of about 60%. Hence two equations are applicable to this case namely equation (5), which is specially derived for meat products and the general relation for unfrozen food products given by equation (7). Equation (5) yields a Cp = 3.170 kJ/(kg⋅K) using data from SFK.

However using equation (7) we get from SFK:

1.6 0.028 2.0 0.136 2.0 0.207 1.1 0.021 4.2 0.600 3.27 kJ/(kg K)pC = ⋅ + ⋅ + ⋅ + ⋅ + ⋅ ≈ ⋅

The results using both equations and data from all three databases are presented in the table below for Cp [kJ/(kg⋅K)].

SFK LST USDA

Equation (5) 3.17 3.17 3.23

Equation (7) 3.27 3.22 3.34

Comments: Again we find only small differences between the equations and databases. Hence the conclusion again is that these differences have no influence on engineering calculations.

Heat conductivity

The heat conductivity k [W/(m⋅K)] cannot be calculated as we have just seen for

the specific heat capacity. Thus we need to rely on experimental determinations based on measurements we either perform ourselves or we utilise other people’s measurements. We often turn to the last option.

The heat conductivity of a material depends on the composition and the structure. For food products we often find an orientation inside the material (e.g. fibre orientation in vegetables and meat), as well as food products may consist of layers of different composition. On top of this some foods contains air or are inhomogeneous in some other way. All of these are conditions we must take into account when determining the heat conductivity of foodstuffs.

Some relatively simple relations are determined as can be seen in Table 3 below.

Page 25: Food Process Engineering

- 21 -

Table 3. Equations for heat conductivity

Product Equation

Fruit and vegetables containing more than 60 % of water

0.148 0.493 wk x= + ⋅ (9)

Meat containing 60-80% water (0 – 60°C) 0.08 0.52 wk x= + ⋅ (10)

A more general equation based on the heat capacity is given by equation (11) where wk is the conductivity of water and ,p wC the specific heat capacity of water and ,p foodC the heat capacity of the food product in question:

,

,

p foodfood w

p w

Ck k

C= (11)

More relations for the heat capacity can be found in appendix 5.

Example 9. Determining heat conductivity

Data: Again we refer to Example 5 for product data. Kw,20°C=0,59W/(m⋅K)

Solution: The poultry pâté has water content of about 60%. Hence two equations are applicable to this case equation (10), which is special for meat products and the general relation given by equation (11). We use data from SFK for demonstration purposes and present all data in a table below, listing k in [W/(m⋅K)].

Equation (10): 0.08 0.52 0.600 0.39 /( )k W m K= + ⋅ ≈ ⋅

Equation (11): 3.27

0.59 0.46 /( )4.182

k W m K= ≈ ⋅

SFK LST USDA

Equation (10) 0.39 0.39 0.40

Equation (11) 0.46 0.45 0.47

Comments: The difference between the databases are negligible, however, here we find large discrepancies between the two equations. The conclusion is that the most reliable relations are the specific. The general equation will yield a result suitable for process engineering calculations, however, it should be used with care.

Page 26: Food Process Engineering

- 22 -

Thermal diffusivity

The thermal diffusivity α [m2/s] is a ratio that involves the thermal conductivity,

the density and the specific heat capacity an is given by:

p

kC

α =ρ ⋅

(12)

Thermal diffusivity may be calculated using the expressions for thermal conductivity, density and specific heat capacity derived earlier.

Viscosity (Newtonian liquids) In relation to fluid flow (i.e. in processing equipment) the most important fluid property is the internal resistance to flow when a shear force (driving force) is applied. This resistance is called the shear viscosity or often simply the viscosity and sometimes even the thickness. The higher the resistance to flow is the higher the viscosity of the fluid.

For Newtonian fluids the viscosity is invariable to the force applied. The best-known Newtonian fluid is water but also skimmed milk, edible oils, sugar syrups and some honey products belong to this group of fluids. Despite the similarities according to the flow behaviour there can be a difference of several decades in the magnitude of the viscosity of these fluids.

The flow behaviour of Newtonian fluids is simple to describe, as they follow Newton's law of viscosity given by equation (13).

τ = −µ ⋅ γ& (13)

Here τ [Pa] is the shear stress, µ [Pa ⋅ s] the Newtonian viscosity and γ& [s-1] the shear rate. The shear rate is defined as the rate of deformation γ = γ& /d dx .

The stringent definition of fluids exhibiting Newtonian behaviour is that:

• The shear viscosity is independent of the magnitude of the shear rate (i.e. the viscosity is constant for all situations).

• The viscosity must not show any time dependency when the material is sheared and the viscosity falls immediately to zero when the shearing stops.

• The viscosity is the same in repeated measurements.

It is important to acknowledge that the definition given above requires both temperature and pressure to be kept constant.

Occasionally the viscosity is given in the unit of Poise [P], however the SI unit is [Pa⋅s] or [N⋅s/m2]. The relation to Poise is 1 mPa s 1 cP⋅ = (equal to the viscosity of water at 20 °C).

Page 27: Food Process Engineering

- 23 -

Table 4. Viscosity of selected fluids

Fluid Temperature (°C) Viscosity (Pa⋅s)

Water 20 1.0 ⋅ 10-3

Raw milk 20 1.99 ⋅ 10-3

Corn oil 25 5.65 ⋅ 10-2

Grape juice 60° Brix 27 0.11

Honey, buckwheat 18.6% (total solids) 24.8 3.86

The magnitude of the viscosity may vary greatly from fluid to fluid. The viscosity for water designates the lower limit practically found for fluids. As seen in Table 4 the viscosity can be several orders of magnitude higher than that of water.

Examples of practical determination of viscosity are given in appendix 5.

The influence of temperature on viscosity

The viscosity of fluids, Newtonian as well as non-Newtonian (explained later), depends on the temperature. Most often a decrease of the viscosity is found with increasing temperature. An acceptable correlation applicable in most cases for the temperature dependence is an Arrhenius relationship as displayed in equation (14), which relates the viscosity to the inverse of the absolute temperature T by fitting A and B.

−µ = ⋅ /B TA e (14)

In most cases the temperature dependency of the flow curves must be incorporated in the shear viscosity model by applying an additional fit of the model parameters to their variation with temperature.

Other characterizing parameters Later on when dealing with heat transfer the Prandlt number proves itself as a convenient dimensionless number. The expression involves the heat capacity pC ,

the viscosity µ and the thermal conductivity.

⋅µ= pCPr

k (15)

The Prandlt number describes the ability of a fluid to take part in the heat exchange process. Thus the Prandlt number is a fluid property and not related to the process. It is so that the higher the Prandlt number the better the heat transfer will be a fluid and a solid surface. We will later on show that the viscosity µ can be

substituted by other expressions for the viscosity than the Newtonian.

Page 28: Food Process Engineering

- 24 -

Rheology of liquid foods The Newtonian fluids are only a small subdivision of the whole spectre of liquid foods. They are by tradition given much attention in relation to fluid mechanics and calculation of processing performance since they by nature are relatively simple to do calculus for. However, in order to determine the relevant processing parameters and process performance measures for the wide range of liquid foods not obeying Newton’s Law of viscosity we must turn to the field of rheology.

Rheology is the science of deformation and flow of matter. Rheology applies to all materials from gasses to solids. The subject might be considered rather philosophical when it is claimed that considering the right time scale (thousands of years) even the mountain flows. In reality, however, it is of utmost important to take into account the relation between applied forces, the time scale observed and the deformation applied.

When classifying materials two ideal types of material behaviour are observed. These are the true Hookean elastic and the purely viscous fluid. In a strict sense very few materials are ideal, however, in praxis groups of nearly ideal materials are found. Fluids such as water or air are very simple as they possess only viscous properties and billiard balls and steel springs are examples of near-ideal elastic bodies. Food materials, on the other hand, are more complex and the internal structure depends to a great deal on the applied processing steps and as such the rheological behaviour of a final food product is a function of the total processing history. The complex structure causes most foods to exhibit a combination of viscous and elastic behaviour. This so-called viscoelastic behaviour is most pronounced for food products where a gel structure is formed. For practical reasons a distinction has to be made between structure rheology (viscoelasticity) where internal molecular bonds are of the greatest importance and shear flow rheology where the mechanical properties of the product as a bulk governs the process. The latter is the main scope in this chapter, where as the structural phenomena is dealt with in details in the chapter on Food Structure.

The time scale over which a deformation process is studied is very important in rheology since a specific material might exhibit significantly different rheological behaviour depending on the time a given deformation is applied. A silicone material nicknamed “Bouncing Putty” has a very high viscosity and will given the appropriate time level out in a container. However, a ball of this material will bounce when dropped on the floor. It can therefore be concluded that this material behaves like a fluid for rather long experiment times and as an elastic-solid for very short times.

In order to quantify the relation between the time scale of the material and the experiment the Deborah number (De) is defined:

= τ ΓDe / (16)

τ is a characteristic time for the material in question and Γ is a characteristic time for the applied deformation process. For a Hookean elastic the time τ is infinite and for the Newtonian viscous fluid τ is practically zero. Therefore high

Page 29: Food Process Engineering

- 25 -

Deborah numbers correspond to dominantly solid behaviour, where as, low Deborah numbers implies a fluid-like behaviour. The earlier mentioned viscoelastic materials lie between the two ideal states and will therefore often depend greatly on the time scale of the applied deformation process.

Even though elastic behaviour plays a very important role in structuring of foods and how we perceive them, the elastic components are to a large extend disregarded in the following. The complexity of the physics needed to describe such phenomena in a way suitable for inclusion in equations determining the nature of fluid flow cannot be justified when determining the performance of the majority of steady state continuous processing of foods.

Non-Newtonian fluids

Non-Newtonian fluids are simply defined as any fluid, which does not obey Newton's Law of viscosity, meaning that the viscosity depends on the processing conditions. However, important the Non-Newtonian fluids are it must be realised that in reality Newtonian behaviour can be assumed for a group of foods products, which exhibits near Newtonian characteristics.

For Non-Newtonian fluids the viscosity no longer holds a constant value but depends on the magnitude of the applied shear rate. Hence, a viscosity value must in each case be related to a specific shear rate in order to make any sense. The correct term to use in relation to viscosity values for non-Newtonian fluids is shear dependent viscosity or apparent viscosity. In order to give a more complete picture of the flow behaviour of non-Newtonian fluids a so-called flow curve showing the relationship between shear rate and the apparent viscosity must be recorded experimentally.

The general expression describing the shear and time dependent viscosity of non-Newtonian fluids is given by equation (17):

η = γ&( , )f t (17)

where η [Pa ⋅ s] is the shear dependent viscosity, &γ [s-1] is the shear rate and t the

time. This rather general equation can for some types of non-Newtonian fluid behaviour be expressed as a mathematical relationship between the shear rate, the shear stress and the shear dependent viscosity.

Time dependency

A convenient subdivision of the non-Newtonian fluids pertains to whether they show significant time dependency or not.

The time-dependent fluids cover a wide area of rheological behaviour where fluids possess both elastic and viscous abilities and where both reversible and irreversible effects can be observed. Hence making the discovery of the relevant rheological behaviour a complex task. Some very ideal relationships between viscosity and shearing for time dependent fluids reversible as well as irreversible are shown in Figure 1.

Page 30: Food Process Engineering

- 26 -

Thixotropic fluids show a viscosity decreases with the duration of shearing. When the shearing is ended the initial higher viscosity is regained. Should the breakdown be irreversible to some extend the viscosity is only regained partly. Examples of such fluids are applesauce and most creams.

Rheopectic fluids exhibit the opposite behaviour, which means that the viscosity increases with the time of shearing. An example is whipping cream where the structure build up is mainly irreversible.

Figure 1. Classification of time dependent non-Newtonian fluids with respect to shear time and time of rest (Hallström et al., 1998).

Most liquid food products or food constituents will at some point exhibit time dependency to a greater or lesser extend. However, time effects are often observed in the initial phase of shearing and can be neglected in most practical applications related to fluid flow. Hence time effects may play a role in start-up of a process and must as such be recognised when determining the characteristics of the food product. Please refer to appendix 6 on Measuring viscosity and flow curves for examples of practically observed time effects.

Time independent liquids

The apparent viscosity of time independent fluids held at constant temperature and pressure changes only with the magnitude of the shear rate.

In general the time independent non-Newtonian fluids show either shear-thinning (pseudoplastic) or shear-thickening (dilatant) behaviour. Figure 2 compares idealised curves for the relation between shear stress and shear rate (upper graph) and between apparent viscosity and shear rate (lower graph) for selected types of fluid behaviour.

The Bingham (or Bingham plastic) type of fluid included in Figure 2 is and example of a yield stress fluid. This effect is often referred to as the tomato ketchup-effect since this is the best-known fluid, which possesses such flow behaviour. This type of fluid has an apparent initial resistance to flow due to a solid-like structure and in reality flow will only be measurable when the stress applied exceeds the apparent yield stress value. This effect will be covered in more detail later on.

Page 31: Food Process Engineering

- 27 -

Figure 2. Comparison of typical Newtonian, shear-thinning and shear-thickening power-law fluids and a Bingham plastic fluid. (Fryer et al., 1997)

In Figure 2 the n denoted along side the flow curves origins from the power-law model. This model is simplest and most frequently used model to describe shear-thinning or shear-thickening behaviour. The power-law model for the relation between shear rate and shear viscosity is the given by:

−η = γ& ( 1)( ) nm (18)

where η [Pa ⋅ s] is the shear dependent viscosity, &γ [s-1] is the shear rate, the

power-law exponent n is a dimensionless number and the consistency index m has the unit of [Pa ⋅ sn]. Note that multiplication with &γ on each side in equation (18) yields the power law relation between shear stress and shear rate (this equation is given in appendix 1).

Depending on the value of the power-law exponent n the power-law model describes three different types of fluid behaviour. Shear-thinning for n < 1, shear-thickening when n > 1 and lastly the case where n = 1 and = µm here the power-law model is identical to Newton's law of viscosity.

The power-law model and other models presented later here are merely convenient engineering representations of certain rheological behaviour of liquids over a specific range of shear rates. These models are therefore not suitable for extrapolation outside the range investigated experimentally.

Page 32: Food Process Engineering

- 28 -

Shear-thinning fluids

The shear-thinning flow behaviour is by far the most important as most fluid foods exhibit such rheological characteristic, where as, only very few are shear-thickening. Examples of shear thinning food products are banana purée, apple purée, mayonnaise, orange juice concentrates and starch suspensions.

The general shape of a flow curve for a shear-thinning fluid is shown in Figure 3. The curve can be separated in three parts with the shear-tinning region in the middle and two Newtonian plateaus - one in each end.

Figure 3. Typical shear-thinning behaviour (Fryer et al., 1997).

Even though the shear-thinning regime dominates the flow curve for these fluids the regions with constant viscosity at low respectively high shear rates must be recognised. They are called the lower and the upper Newtonian region respectively. The levels refer to the magnitude of the shear rate and not the viscosity. These regions can be more or less pronounced and might some times lie outside the range of practically measurable shear rates.

Far to the left (low shear rates) on the curve in Figure 3 the internal molecular forces dominate the structure and thereby the flow properties of the liquid; i.e. we have a random ordering and therefore a constant viscosity. As we move from left to right hydrodynamic forces from the process enforces increased structuring and alignment of molecules and this causes the viscosity to drop. In the end when the fully aligned structure is reached the viscosity again becomes constant. Beyond the upper Newtonian region increased shear rates can cause the structure to breakdown. This is for example intended in mixing processes.

The power-law model can with its two parameters (n and m) only fit the shear-thinning part of the flow curve shown in figure 3. The model is thus a relatively simplistic way to describe shear-thinning behaviour. Additionally a graphical interpretation of n can be found when plotting the flow curve in a double logarithmic plot, n is namely the slope of the shear-thinning region. This leads to the appreciation that n denotes the degree of shear-thinning behaviour and in praxis values of n between 0.2 and 1 is found for liquid foods. The consistency index m is a measure of the viscosity level. Examples of shear-thinning behaviour are given in Table 5.

Page 33: Food Process Engineering

- 29 -

Table 5. Power-law parameters for selected liquid foods (Fryer et al, 1997)

Food T (°C)

n (-)

m (Pa sn)

Range of (s-1)

Banana purée 22 0.28 107.3 28 – 200

Apple sauce 26 0.45 7.32 0.78 – 1260

Mayonaise 25 0.60 4.2 40 – 1100

The values of n and m shown above are determined by fitting the power law model to experimental data sets as described in the example below.

Example 10. Determining the parameters of the power law model

Data: A traditional fat rich mayonnaise has been subjected to repeated measurements of the flow curve. In the figure below data from three repeats is plotted along with lines representing a power law fit for each data set.

Shear dependent viscosity of mayonaise

Shear rate [s-1]

10 100 1000

Vis

cosi

ty [P

a s]

0,1

1

10

Data set 1Model 1Data set 2Data set 3Model 2Model 3

Solution: The fitting parameters for the power law models are given below:

Data set 1 Data set 2 Data set 3

n 0.42 0.43 0.46

m 13,1 15.2 10.1

Page 34: Food Process Engineering

- 30 -

Comments: We see that the three repeated measurements coincide pretty well and only small deviations are found in the model fit. In the case presented we could easily have pooled the data and made one model.

We have seen the inability of the power-law model to describe the full flow curve shown in Figure 3. We therefore turn to models with more fitting parameters in realisation of that four fitting parameters are needed to describe the plateaus at each end of the curve.

In Figure 4, the basic shapes of the shear thinning flow curve is expressed by the power-law, Sisko and Cross models respectively. We see that a three-parameter model like the Sisko model (not stated here) can fit the upper Newtonian plateau. The last model presented in the figure is the often applied Cross model which has the four parameters needed.

Figure 4. Typical viscosity curves from rheological models (Barnes et al., 1989).

The Cross model which is given by equation (19):

( )( )∞∞

η−η =η −η + γ& 10 1

1

1 nK , (19)

here 0η and ∞η is the Newtonian viscosity at low and high plateau respectively. K1 is a constant with the dimension [s] and n1 is a dimensionless constant equivalent to the power-law exponent. It is seen that the Cross model holds fitting capabilities similar to the power-law model.

A popular variant to the Cross model is the Carreau model, which has a slightly different implementation of the fitting parameters.

( )( )∞∞

η−η =η −η + γ& 2 / 2202

1

1n

K (20)

The constants K2 and n2 are similar to K1 and n1 in the Cross model.

Page 35: Food Process Engineering

- 31 -

When selecting a proper model to describe non-Newtonian flow behaviour it is important to identify the specific application needs. The power-law model can be perfectly appropriate when comparing different fluids, studying a process, which operate at shear rates in the shear-thinning region or for use in simplified fluid mechanical calculations. Where as the Cross and Carreau models gives much more detailed information and can be required especially in order to describe the viscosity levels at low or high shear rates.

A drawback in using the more detailed models is that the increased number of fitting parameters (from 2 to 4) requires more experimental data points in order to fit a reliable model. It is difficult to give guideline regarding how many data points one needs to fit a given model, however, a good rule of thumb is that the experimental data points must describe the shape of the part of flow curve one wishes to model very well. If lacking data points an alternative approach can be to fix some of the fitting parameters before the data fit is performed. The viscosities for the plateau’s 0η and/or ∞η can either be measured or assigned assumed values in order to reduce the number of fitting parameters in the Cross and Carreau models. An other approach often taken is to assign n1 in the Cross model the value of 0.66. This might seem as a major restraint on the fit, however, this is found yield good fits for many liquid foods.

Choosing the right model depends on the application. Always apply the model, which describes the part of the flow curve in interest. Even though we have realised that the power law model does not describe the full flow curve this model is sufficient to compare the flow characteristics of non-Newtonian fluids.

Shear-thickening fluids

Shear-thinning behaviour is rarely found for liquid foods; however, one example is highly concentrated starch suspensions.

Fluids exhibiting an apparent yield stress

In appreciation of the fact that some fluid foods at rest exhibit an extremely high resistance to deformation and the fact that the shear stress in such cases must exceed a certain value, the apparent yield stress yτ (Pa), before notable flow

occurs this phenomenon needs to be addressed. It is an apparent yield stress we observe when we try to get tomato ketchup to flow out of a bottle. Everybody has experienced that a high force or even violence needs to be applied in order to get the flow started but once it suddenly comes we often get more than we intended. This is an excellent example of a process where we need to apply a force above the apparent yield stress before the fluid flows, however, we easily observe that a much lower shear stress is needed to spread the ketchup on a plate. This is why an apparent yield stress only applies to a specific process and why reported yields stresses are not universal figures.

Some scientists argue whether the phenomena yield stress really exists or if the flow in the initial phase is just too small to be measured with the rheological instruments available today and yet some claims that yield stress is a proven fact. No matter who is right it is of practical relevance to operate with the term

Page 36: Food Process Engineering

- 32 -

apparent yield stress for some materials (like the ketchup and tooth paste) in certain processes.

Two common models, which are useful when describing fluids exhibiting an apparent yield stress, are presented here.

The Bingham plastic fluid has an apparent yield stress yτ followed by a

Newtonian flow region. The Bingham model is given by equation (21):

τ = τ +µ⋅γ&y (21)

Examples of Bingham fluids are: toothpaste and tomato products.

For fluids where an apparent yield stress is combined with shear-thinning behaviour (i.e. yoghurt) the Hershel-Bulkley model given by equation (22) can be applied:

τ = τ + ⋅ γ&( )ny m (22)

Examples of food products obeying the Hershel-Bulkley model are: minced fish paste, raisin paste and some yoghurt types.

The role of elasticity and time effects

As stated earlier many liquid foods possesses both viscous and elastic behavior as they to different degrees are what we call viscoelastic. This might be a bit difficult to comprehend, however, if we consider a fruit gel (like jam or marmalade) we can by compressing it slightly with a finger observe an elastic effect. The same product can be brought to flow in a tube given that an appropriate shear stress is applied. In most steady state processing applications we choose to neglect the elastic component since the elastic effects die out after a start-up phase in the processing. However, there are processes like extrusion of liquorices where contraction and pressure build-up and other force induced forming processes where elastic components play an important role when selecting the proper process parameters. Such processes are, however, not included here.

Temperature dependency

In most cases the temperature dependency of the flow curves must be incorporated in the shear viscosity model by applying an additional fit of the model parameters to their variation with temperature. See under Newtonian fluids for reference to a model for temperature dependency (page 23).

Rheology solid and semi-solid food products This part is unfortunately still under construction

Page 37: Food Process Engineering

- 33 -

Fluid mechanics Transport of matter is an essential operation in manufacturing plants. When it comes to fluids they are pumped, flow in pipes, are mixed and filled into containers etc. These are all operations, which are based on the flow behaviour of the fluids. The actual flow conditions found in processing equipment are determined where fluid mechanics is combined with the product rheology. Fluid mechanics of Newtonian and Non-Newtonian liquids is therefore a basic necessity to understand in order to work qualified with operation and design of processing equipment.

The basics of fluid mechanics are introduced for Newtonian fluids and extended to cover the most relevant Non-Newtonian fluids as well.

Newtonian fluids The nature of the flow pattern depends greatly on the rate of the flow. At low rates fluids flow in an orderly manner, which is highly reproducible. The flow is said to be laminar and ideally no exchange of fluid takes place in the height of the tube. This type of flow can be pictured as parallel layers, which flows smoothly side-by-side along the length of the tube and for flow in a tube with a circular cross-section the velocity profile is parabolic which can be seen to the left in Figure 5. In laminar flow the viscous forces dominates this means that the fluid viscosity is of higher importance than the hydrodynamics of the flow it self.

At higher flow rates the flow becomes turbulent. The flow pattern is now formed randomly, there is an increased exchange of fluid in the height of the tube and the flow is no longer reproducible but fluctuates continuously. The turbulent velocity profile is shown to the right in Figure 5. The hydrodynamic forces are now dominant and hence the fluid properties less important for the nature of the flow. Between the true laminar and fully developed turbulent flow regimes a transition takes place, in this zone it is difficult to determine the nature of the flow.

Figure 5. Laminar and turbulent velocity profiles for pipe flow. Here vav. is the average velocity (Leniger & Beverloo, 1975).

Page 38: Food Process Engineering

- 34 -

The parabolic velocity profile for the laminar can be derived from exact mathematics. However, there is no such exact mathematical equation for the turbulent velocity profile since random movements dominate the flow. By examining the two profiles we can easily see that the turbulent velocity profile is more flat nosed than the laminar and that the difference between the maximum and the average velocity is much less for turbulent flow compared to laminar.

When observing the velocity profiles we see that both touch the wall of the tube. This means that in this position where the lowest velocity is observed the fluid velocity matches that of the tube wall. Thus the velocity is zero here. This is the generally accepted boundary condition is that there is no slip between fluid and any hard surface (e.g. a tube wall). In reality this is of course an assumption but it yields realistic results in must situations where the slip is negligible. However, there are situations where lubrication of the surface takes place. In such cases the flow characteristics must be determined by experiments.

The Reynolds number (Re) is a dimensionless number, which yields the ratio of inertia to viscous forces and thus determine the nature of the flow. For low values of the Reynolds number the flow is laminar and for higher values first transition takes place and in the end fully developed turbulent flow develops. The limits for the Reynolds number determining the flow type depends on the cross section profile of the geometry studied. In a cylindrical pipe laminar flow is found for values of Re below 2100 (different sources reports the limit to be anywhere in the interval of 2000 to 2300). This value is called the critical Reynolds number (Rec) and is highly geometry depended. The flow is considered fully turbulent when the Reynolds number exceeds 4000. Hence transition flow is found between these two values.

The Reynolds number for Newtonian fluids is defined by:

⋅ ⋅ ρ= µRev d

(23)

where v (m/s) is the average velocity of the flow, d (m) is a characteristic distance (for a pipe the diameter), ρ is the density of the fluid and µ the Newtonian

viscosity. Alternatively the expression for the Reynolds can be rewritten to depend on the volume flow (V& ) instead of the average velocity.

⋅ρ ρ= =µ π ⋅µπ⋅& &2

4 4V d VRe

dd (24)

The expression for the volume flow is found in appendix 1.

Example 11 – Calculation of Reynolds number

Data: We calculate the Reynolds number for water flowing in a 2” (0.0508m) pipe at an average velocity of 2m/s at 25 °C. Physical properties for water at 25°C: ρ = 1000kg/m3 and µ = 9⋅10-4 N⋅s/m3. Hence we get: V& ≈ 0.00405m3/h

Page 39: Food Process Engineering

- 35 -

Solution: −⋅ ⋅= ≈ ⋅⋅ ⋅ ⋅

54

4 0.00405 10001.13 10

3.14 0.0508 9.0 10Re

It is seen that the flow is very much in the turbulent flow regime.

Comments: Had the fluid been a grape juice the viscosity would be 0.11 Pa⋅s (taken from Table 4), hence yielding a Reynolds number of 923, in which case the flow is laminar.

We have assumed that the density does not change, which can seem a bit careless and inaccurate at first. However, applying the right value will only change the Reynolds number slightly and does by no means change the result that the flow is laminar. By carrying out few calculations we can conclude that the density only plays a minor role in determining the flow characteristics.

Looking at the expression for the Reynolds number given in equation (24) the significant parameters influencing the magnitude are identified as the flow rate, the pipe diameter and the viscosity. In creasing the flow rate leads to an increase in the Reynolds number, where as an increase in either the diameter of the pipe or the viscosity would decrease the magnitude.

The nature of the flow has great implications for the process performance such as heat transfer, residence time and mixing. Turbulent flow will for example yield better heat transfer performance, however, the pressure needed to drive the liquid trough the process plant will increase at the same time and might limit the practical possibility to go to turbulent flow if for example the viscosity is very high. This is indeed the case for many liquid food products as they have a relatively high viscosity and thus yielding moderate Reynolds numbers. This is in contrast to classical chemical processes where high Reynolds numbers are found for flows involving gasses and low viscosity fluids.

Laminar flow

The parabolic velocity profile of laminar flow can described by equation (25):

∆ = − ⋅ ⋅µ 2 2( )

4P

v r R rL

(25)

where v(r) [m/s] is the velocity in position r [m] in the tube, ∆P [Pa] is the pressure gradient over the tube, L [m] the length of the tube, R [m] the radius of the tube and r [m] the actual radial position in the tube.

In engineering calculations the exact shape of the velocity profiles is often not as important as the relation between key characteristics of the flow and selected characteristic parameters.

The magnitude of matter flowing is specified by: a volumetric flow rate V& [m3/s], a mass flow rate &m [kg/s] or an average velocity v [m/s]. The rate of the flow is of interest as it is related to the pressure gradient, which can be looked upon as the pump energy requirement needed to keep the fluid flowing.

Page 40: Food Process Engineering

- 36 -

Equation (26) yields the relation between the average velocity and the pressure gradient:

∆ ∆ µ ⋅= ⇔ =µ2

232

32d P P v

vL L d

(26)

The relationship between the average velocity and the maximum velocity given by equation (27) is relevant when evaluating the effect and choosing process parameters of continuous processes (please refer to appendix 1 for more relevant relationships):

= ⋅max 2v v (27)

Example 12. Calculating pressure drop in laminar flow

Data: If we look back at the previous example with the grape juice where the flow were found to be laminar it is now possible to calculate the pressure needed to drive the product through a pipe.

Solution: As we do not know the length of interest we choose to calculate the pressure drop for pr. meter pipe.

⋅ ⋅∆ = ≈ ≈232 0.11 2

/ 2728 / 0,027 /0.0508

P L Pa m bar m

Comments: Which is a reasonable magnitude for a pressure drop. Had the process on the other hand taken place in a 1” pipe the pressure drop pr. meter pipe rises to 0,11bar (Re=1847), which can give cause to considerations as the pressure needs to be build up by a pump. A practical limit in the pressure is considered in the range of 4 to 7bar (or in some cases up to 10bar) depending on the product.

Assuming the pressure limit is 7bar yields a maximum length between two pumps of 64m for the 1” pipe, where as it is 257m for the 2” pipe (4bar limit yields a maximum length of 147m). In this case the smaller pipe would most certainly not be applied.

The shear stress relation is given by equation (13) and is proportional to the velocity gradient. For the parabolic velocity profile we know that the first derivative is a straight line and furthermore that in the top point this value is zero. We can now intuitively deduct that the shear stress has its maximum at the wall of the tube and that it from this point decreases linearly to zero in the centre of the tube. We will later appreciate the importance of this when dealing with non-Newtonian fluid mechanics. The wall shear stress wτ can be expressed by:

⋅µ⋅τ = 16w

vR

(28)

Page 41: Food Process Engineering

- 37 -

Turbulent flow

No exact mathematical analysis exists for turbulent flow, though some semi-theoretical expressions have been suggested. A suitable empirical formulation has shown that the velocity at any position in the cross-section of a tube will be proportional to the 1/7 power of the distance from the wall. The velocity profile can now be expressed using the notation applied in equation (25) for the laminar flow (Coulson & Richardson, 1996).

= ⋅ − 1/7( )

1.22 1v r r

v R (29)

The relation between the maximum and the average velocity is found in the centre of the pipe where r equals 0:

= ⋅max 1.22v v (30)

The relation between the average velocity and the pressure gradient can be expressed by the Fanning friction equation, which can be derived from equation (26) by introducing the expression for the Reynolds number (Fryer et al., 1997 & Toledo, 1991):

∆ ⋅ ⋅ ⋅ ρ=22P f v

L d (31)

In equation (31) f is the friction factor, which for laminar flow is 16/Re and for turbulent flow in a smooth pipe f can be computed from equations (32) and (33):

−= < <0.20 4 60.048Re 10 10f Re (32)

−= ⋅ < <0.35 3 40.193Re 3 10 10f Re (33)

Equation (33) can even though the lower limit in the Reynolds number is 3000 it reasonable to assume that this equation is valid down to a Reynolds number of 2100.

We see from equations (32) and (33) that the influence of an increase in the Reynolds number on the pressure loss becomes less important the higher the magnitude of the Reynolds number. The value for the friction factor can also be determined from the Moody diagram (see appendix 7), which also holds information for rough pipes.

Example 13. Calculating pressure drop in turbulent flow

Data: In

Example 11 we determined a flow of water to turbulent. Pipe diameter 2” (0.0508m), average velocity: 2m/s, temperature 25 °C, ρ = 1000kg/m3 and µ = 9⋅10-4 N⋅s/m3, V& ≈ 0.00405m3/h and Re = 1.13*105

Page 42: Food Process Engineering

- 38 -

Solution: Given the magnitude of the Reynolds number the friction factor is

computed from equation (32), hence −= ⋅ =4 0.200.048(1.13 10 ) 0,00468f .

The pressure drop pr. meter pipe is then determined by insertion in equation (31):

∆ ⋅ ⋅ ⋅= ≈ ≈=22 0,00468 2 1000

738 / 0,0074 /0,0508

PPa m bar m

L

Comments: This pressure drop is according to Example 12 very much reasonable.

Had the average flow rate instead been 3m/s we would get: V& ≈ 0.0061m3/h, Re = 1.7*105, f =0,04317

∆ ⋅ ⋅ ⋅= ≈ ≈=22 0,0043 3 1000

1530 / 0,015 /0,0508

PPa m bar m

L

This new value is twice as high as the previous but still of an applicable magnitude.

From the example we find that the magnitude of the friction factor is not much affected by the change in the Reynolds number caused rise in the average velocity as would be the case for any change in the Reynolds number (e.g. as the result of altering the viscosity). The main factor affecting the pressure drop is in this case the average velocity it self. A similar effect could have been seen if the pipe diameter was reduced by a factor of two.

The wall shear stress still assumes it maximum value at the wall of the tube and the symmetry still causes the value to be zero in the centre of the tube even though the decay is no longer linear. The wall shear stress can be calculated from the general equation given below.

∆τ =4wd P

L (34)

Non-Newtonian fluids

When applying a model of non-Newtonian behaviour in fluid mechanics a relation between the fluid behaviour and the constitutive equation of the flow need to be determined. Since the velocity profile depends on the viscosity it is of utmost importance to describe the viscosity change with the shear rate, which in some cases be as much as 2 to 4 decades. Such a change cannot be ignored when calculating velocity profiles and pressure drops in process equipment.

One of the earliest and most often implemented models is that of the generalised Newtonian fluid. The generalised Newtonian fluid is a minor modification of the Newtonian fluid where the viscosity is allowed to be a function of the shear rate. The practical application of non-Newtonian models in fluid mechanics is handled by implementing the power-law model from equation (18) into the governing equations for fluid flow (Bird et al., 1987). The mathematics, the theoretical consideration and assumptions of applying the generalised Newtonian fluid are not covered here. The reader must refer to Bird et al., 1987 for this information.

Page 43: Food Process Engineering

- 39 -

Analogous to Newtonian fluid mechanics we can compute the Reynolds number. A rewritten equation for the Reynolds number incorporating the exponent n and consistency index m from the power-law model is given by equation (35).

−− ⋅ ⋅ρ= + (2 )

3Re 23 1

n nn n v d

n m (35)

where v [m/s] is the average velocity of the flow, d [m] a characteristic length of the geometry being considered, ρ [kg/m3] the density.

Example 14. Calculating Reynolds numbers for non-Newtonian fluids

Data: Utilising the product data from Table 4 on page 29 we can calculate the Reynolds number. Banana pure (n=0.28 and m=107.3), pipe diameter 1”, average velocity: 1,1m/s, ρ = 1000kg/m3 and V& ≈ 2.0m3/h.

Solution: According to equation (35) the Reynolds number yields:

−− ⋅ ⋅= ≈ ⋅ + (2 0.28) 0.28

3 0.28 0.28 1.1 0.0254 1000Re 2 3.9

3 0.28 1 107.3

Comments: Such low Reynolds numbers are no uncommon when dealing with non-Newtonian fluids exhibiting a high viscosity level and as is seen from the table in Example 16 all cases considered proves well within the laminar flow regime.

We can also deduct an equation for determination of the critical Reynolds number where shear-thinning behaviour is considered.

( )( )

( )+ +=

+ 24 2 5 3

Re 21003 3 1

cn n

n (36)

Following equation (36) it can be observed that Rec is higher than 2100 for n < 1 which denotes shear thinning behaviour. We find that the limit between laminar and turbulent flow shifts to higher Reynolds numbers as n decreases, where as it shifts to lower values for n>1. The relationship between Rec and n is plotted in Figure 6.

Example 15. Critical values of the Reynolds number for power law fluids

It is very important not to confuse this number with the Reynolds number.

Data: As stated above the critical Reynolds number designates the limit between the flow being laminar or turbulent.

Solution: For a fluid with an n of 0.5 a Rec of 2463 is found which means that the flow stay laminar for higher flow rates compared to the Newtonian case, where the critical values is 2100. For a dilatant fluid where n is equal to 2 the critical limit drops to 1857.

Page 44: Food Process Engineering

- 40 -

Comments: We see that laminar flow can be found for higher values of the Reynolds number for shear-thinning fluids where as the transition to turbulent flow occurs earlier for the shear-thickening fluids.

Power law index (n)

0 1 2 3 4 5

Re c

1500

2000

2500

3000

3500

n<1

n>1

Figure 6. Critical Reynolds numbers Rec as a function of the power law index n.

Laminar flow

For laminar flow of a Newtonian fluid in a tube we found that the velocity profile takes the form of a parabola. That result was deducted from the constitutive equations for the flow. By incorporating the power-law model for the viscosity in the same set of equations we can derive an analytical mathematical formulation for the velocity profile for a power-law fluid in a pipe. The equation describing the velocity profile is given by equation (37).

+ + ∆ = − + ⋅ ⋅1/

1 1/ 1 1/( )1 2

nn nn P

v r R rn L m

(37)

where v(r) [m/s] is the velocity in position r, ∆P [Pa] is the pressure gradient over the tube, L [m] the length, R [m] the radius of the tube and r [m] the actual radial position in the tube. We see that for the case of a Newtonian fluid where n = 1 and = µm equation (37) reduces to the parabolic form shown in equation (25).

The expression for the relationship between the average velocity and the pressure drop is given by:

+ ∆ = +1/

( 1)/14 6 2

nn nP n

v dL m n

(38)

Page 45: Food Process Engineering

- 41 -

Example 16. Calculating pressure drop for non-Newtonian fluids

Data: Banana pure (n=0.28 and m=107.3), pipe diameter 1”, average velocity: 1.1m/s, ρ = 1000kg/m3 (as in the example Example 14).

Solution: Equation (38) needs to be rearranged before being able to calculate the pressure drop.

+

−+

∆ = ⋅ ⋅ +

∆ ⋅ ⋅= ≈ ⋅ = ⋅ +

c

0.280.28 (0.28 1)

0.285

0.28(0.28 1)

1 0.281.1 0.0254

4 107.3 6 0.28 2

1.1 4 107.31.0 10 / 1.0 /

0.280.0254

6 0.28 2

PL

PPa m bar m

L

This is a high pressure drop pr. meter as we can only pump the product up to 7 meter, which will be considered to short in most processing applications.

Utilising the data for all three products presented in Table 5 on page 29 a number of examples are given to the relation between the flow rate, the Reynolds number and the pressure drop.

Process characteristics for selected power law fluids

Product n

m

[Pa⋅sn]

v

[m/s]

d

[m]

V&

[m3/h]

Re /P L∆

[bar/m]

Banana pure 0.28 107.3 1.1 0.0254 2.0 3.9 1.00

0.28 107.3 0.125 0.0762 2.1 0.1 0.13

Apple sauce 0.45 7.32 1.4 0.0508 10.2 68 0.07

0.45 7.32 2.8 0.0508 20.4 198 0.10

Mayonaise 0.6 4.2 1.0 0.0254 1.8 30 0.23

0.6 4.2 0.445 0.0381 1.8 12 0.07

Comments: All examples apply realistic volume flow for industrial applications; hence suitable values of the average velocity and pipe diameter are selected in order to achieve pressure drops with in a realistic range.

By normalising equation (37) using the average velocity v given in equation (38) we can eliminate all dependencies but n and the radial position in the tube. This allows us to compare the general shape of velocity profiles for different power-law fluids. The normalised velocity profiles are calculated from equation (39):

+ + + = − +1 1/ 1 1/( ) 3 1

1n nv r n

R rv n

(39)

Page 46: Food Process Engineering

- 42 -

In Figure 7 velocity profiles for shear-thinning as well as shear-thickening fluids are compared to the parabolic profile of a Newtonian fluid.

We see that the velocity profiles for the shear-thinning fluids have a more flat nosed profile similar to that seen for turbulent flow. For the shear-thickening fluids, however, the velocity profile assumes a more triangular shape. The most obvious result is that the relation between the average and the maximum velocity increases for shear-thickening fluids where it decreases for shear-thinning fluid.

Figure 7. Velocity for power-law fluids for different values of the power-law exponent n (Holdsworth, 1992).

The velocity profile for n = 0 is identical to that of plug flow. Plug flow is an ideal case, which never is practically obtainable in the whole cross-section of the tube.

For the power-law fluid the relation between the maximum and the average velocity is by the following equation:

⋅max3 +1

=+1n

v vn

(40)

If we consider flow of a Bingham fluid described by equation (21) we can show that at a radius less that ry the fluid stress does not exceed the apparent yield stress value and material in the centre of the tube moves as a plug as can be seen in Figure 8.

The value of ry can be determined from equation (41).

τ= ∆4

/y

yr P L (41)

Page 47: Food Process Engineering

- 43 -

Figure 8. Velocity profile in a tube for a Bingham plastic fluid (Fryer et al., 1997).

Turbulent flow

When it comes to turbulent flow there are strong indications that non-Newtonian fluids show flow behaviour similar to that of Newtonian fluids. We have earlier shown that the fluid properties are less important to the nature of the flow in the turbulent region. On top of this it has been reported that if we assume a non-Newtonian fluid to be Newtonian in turbulent flow we slightly over-predict the pressure drop. It can therefore be concluded that there is a safety factor build in when treating turbulent flow of non-Newtonian fluids as if it was a Newtonian fluid. Hence we choose to calculate the Reynolds number using equation (35) and after worth determine the friction factor using either equation (32) or equation (33). The pressure drop is now determined by insertion in equation (31).

Flow through pipe fittings Often we need to calculate the pressure drop for other equipment parts than straight pipes. This is performed in a way that the equivalent length of straight pipe is computed using the dimensionless ratio of L’/D. Below L’/D is listed for some often used fittings.

Table 6. Equivalent L’/D ratios for fittings. (Adapted from Toledo, 1991)

Fitting type L’/D

90° elbow 35

45° elbow 15

Tee (used as a coupling), flow straight through 20

Tee (used as an elbow) 60

Gate valve, fully open (butterfly ventil) 10

Globe valve, fully open (kugle ventil) 290

Diaphram valve, fully open (membrane ventil) 105

Couplings and unions neglible

Page 48: Food Process Engineering

- 44 -

The length L’ is calculated based on a known diameter and subsequently added to the total length of straight pipe in the system.

Example 17. Determining the pressure drop in fittings 1

Data: We have a 90° elbow of a dimension of 1” we get a value of L’/D of 35.

Solution: L’ yields:

= ⋅ = ⋅ =' 35 0.0254 35 0.9L D m m m

This length is to be added to the total length of straight pipes and other fittings and subsequently the pressure prop must be computed using the relevant equation.

Comments: We se that an elbow is almost equivalent to a whole metre of tube

Example 18. Determining the pressure drop in fittings 2

Data: An other example is the Globe valve, where L’/D is 290

Solution: In this case L’ is as much as:

= ⋅ = ⋅ =' 290 0.0254 35 7.4L D m m m

Comments: We expect that there is a significant pressure drop over a valve and here the maximum values is given.

Page 49: Food Process Engineering

- 45 -

Heat transfer Processes involving heat transfer are highly important to food processing as processes involving heating and cooling is the most common in food processing plants. All processed foods shall be subjected to pasteurisation in order to prevent the presence of pathogenic microorganisms and to reduce the risk of microbial or enzymatic degradation of foods during transport and storage. We also apply heat or cold to introduce changes to the products in order to achieve desired characteristics of a food product e.g. flavour, texture, structure, colour etc.

Examples of processes were heating and cooling are involved are: pasteurisation, sterilisation, cooking, dehydration, drying, freezing and storage. A few examples of the physics involved in such processes are given as an introduction.

Pasteurisation and sterilisation processes apply fast heating and cooling, which implies that very high rates of transport of heat energy in and out of a product are of interest. In the case of a cool storage room or tempered storage tank insulation is applied to minimise the energy loss to the surroundings. Hence the interest is how low energy transfer rates can be achieved.

Transferring heat from one media to another is called heat transfer and it is comprehensive to separate steady state and unsteady state heat transfer. In steady state heat transfer the temperature at a specific location in the equipment is constant over time, hence the process itself is considered to be in steady state. Unsteady state heat transfer involves changing temperatures over time in the equipment and the product, hence yielding an unsteady process.

Example 19. Steady state heat transfer 1

For a freeze storage room the temperature inside the room and in the environs can be considered constant. In order to maintain constant temperature inside the freezer we must remove energy constantly. Hence we have a steady state heat transfer process with a transport of energy even though the temperatures are constant.

Example 20. Steady state heat transfer 2

In a continuous heating operation product flow through a heat exchanger at a constant rate and both inlet and outlet temperatures are constant. Thus we need to transfer energy to the process at a constant rate. The process involves a continuous flow of product but the temperature in a specific location is constant. Hence we have a steady state heat transfer process.

Example 21. Unsteady state heat transfer

A piece of meat is put in cooking brine. The temperature of the meat is constant throughout the piece before cooking. When the cooking starts the temperature changes gradually over time. We can observe this by cutting the meat and looking

Page 50: Food Process Engineering

- 46 -

at the colour when the cooking is stopped. If we wait long enough the temperature will be the same all through the piece but at shorter times we might find raw meat near the centre while the outer layers are cooked. This process involves a constant cooking temperature and changing temperature in the meat over time. Hence the heat transfer is unsteady.

Transmission at the surface in the freeze storage, at the surfaces in the heat exchanger and at the surface of the meat is controlled by convection whilst the transmission of energy through walls and inside the meat is due to conduction. These to cases of heat transfer will be dealt with separately.

Steady state conduction heat transfer Transport of thermal energy in solid matter can be expressed by Fouriers equation:

= = − ⋅& [J/s]dQ dT

Q k Adt dx

(42)

Where &Q [J/s] or [W] is the rate of energy transferred, k [W/(m⋅K)] is the thermal

conductivity, A [m2] is the area over which the energy is transferred, dT [K] is the driving force and /dT dx [K/m] is the temperature gradient.

Example 22. Steady state conduction heat transfer

Data: An illustration is the case where a stick connects two reservoirs of different temperature T1 and T2 ,where T1 > T2 . The cross section area of the stick is A [m2] and the length is L [m]. The two reservoirs hold a constant temperature (or has a sufficiently high heat capacity that the temperature does not change during the process).

Solution: In this case the driving force is (T1 - T2) [K] and the temperature gradient (T1 - T2)/L [K/m]. This makes it possible to determine the temperature at any position in the stick since we know the temperature at both ends and the distance to the point in question.

Comments: However simple this example demonstrates the driving temperature gradient in pratice.

In the steady state situation equation (42) can be written as:

− ∆= − ⋅ ⋅ − ⋅ ⋅& 1 2 = [J/s]T T T

Q k A k AL L

(43)

The energy flux [W/m2] is often an interesting characteristic:

= − ⋅ −& 2

1 2( ) [W/m ]Q k

T TA L

(44)

Page 51: Food Process Engineering

- 47 -

We see that the energy flux is proportional to the temperature difference and to (k/L). Hence this term yields valuable information regarding the characteristics of such transport processes.

We make an analogy to Ohms law (from electricity) and get that the energy flux is equivalent to the current (I) and that the temperature difference the voltage (U). Ohms law is written as: = ⋅U R I or = /I U R . Hence (k/L) must be equivalent to the resistance R, which can be described as the resistance towards transport of energy in solid material. We therefore call (k/L) the thermal resistance.

Equation (43) can now be written as:

⋅∆=& [J/s]A T

QR

(45)

If we have a number of layers in close contact with each other we can describe the total resistance as the accumulated sum of the individual resistances. This is again analogous to Ohms law where summations of resistors in series also are applied.

Assuming that the cross section area is constant for the layers we get:

= = + + + = + + + =∑ ∑1 21 2

1 2... ... ji

j ii j

xx x xR R R R R

k k k k (46)

The assumption is of cause that the layers are in immediate contact and that the resistance between the layers therefore can be neglected. Furthermore the temperature of the surfaces is maintained constant (in reality the outer surfaces). If there is a resistance in the transition between layers (and not infinitely good heat conductivity) this has to be expressed by an additional resistance. This is often the case where heat is transferred between fluids or gasses in motion and solid material. Due to friction a layer of the fluid near the solid material will not be in motion or move very slowly. Thus creating an isolating film towards the solid material. This is in reality convection heat transfer, which will be covered in detail later. At this point we merely extend to resistance analogy to cover transfer processes in general.

A wall where the surfaces are in contact with a warm and a cold fluid respectively (like in a heat exchanger) the total resistance is expressed by an expression similar to that in equation (46).

= + +, ,total film cold wall film warmR R R R (47)

The resistance in the thin film layers can be described by the following equation:

= filmfilm

film

xR

k (48)

Page 52: Food Process Engineering

- 48 -

The heat conductivity filmk of the fluids can be determined, however, the film thickness filmx is not easily estimated. Hence the term film heat transfer

coefficient h is introduced as an expression of this resistance:

= ⋅ = =2 1 [W/(m K)] and fluid film

filmfilm fluid

k xh R

x h k (49)

The result from equation (49) is inserted in equation (47), which yields an expression for the total resistance in the transfer of heat between two fluids through a solid material.

= + + = + +∑1 1 1 1jtotal wall

cold warm cold j warm

xR R

h h h k h (50)

It is convenient to introduce the overall heat transfer coefficient U, which is the inverse of the total resistance ( = 1/ totalU R ). Hence equation (45) can be

rewritten to the more appropriate form:

= ⋅ ⋅∆& [J/s]Q U A T (51)

Hence the calculation of the resistances can be written in a general way as:

= + +∑1 2

1 1 1j

j

x

U h k h (52)

The reason the reciprocal values are applied is psychological since the larger the magnitude of U the more energy can be transferred. The same is of cause the case concerning the film heat transfer coefficients.

The remainder of this chapter concentrates to a great extend on determination of film heat transfer coefficients and over all heat transfer coefficients, since these numbers determine the efficiency of most of the processes where &Q are of interest to us.

Example 23. Validation of energy transfer in a freezer

Determine the maximum amount of energy (heat) needed to be removed from a domestic freezer on a summer day.

Data: temperature of the environs Tenv=25°C, temperature in the freezer Tfreeze=-18°C, the thickness of the insulation layer (polyurethane foam) is 5cm and kfoam=0,033W/(m⋅K). The outer dimensions of the freezer are: H=80cm, D=60cm and

L=120cm.

Solution: Since we are interested in the maximum effect we only need to know the resistance towards heat transfer in the insulation layer. We set the resistance to zero on both inner and outer surface. Hence the film heat transfer coefficients are infinite.

Page 53: Food Process Engineering

- 49 -

The resistance towards heat transfer is calculated from the reduced form of equation (50):

= = = ⋅⋅0.05

1.515( )/0.033 /( )

foamfoam

foam

x mR m K W

k W m K

Hence the maximal energy flux yields by insertion in equation (45):

∆ − −= = =& 225 ( 18)

28.4 /1.515

Q TW m

A R

The outer area of the freezer: = ⋅ + ⋅ + ⋅ = 22((0.8 0.6) (0.8 1.2) (0.6 1.2)) 4.32A m

The total amount of energy then yields:

= ⋅ = ⋅ =&& 28.4 4.32 122.7

QQ A W W

A

Comments: At first it might not be all that obvious that we can neglect the resistance towards heat transfer from the wall to either side. We consider what would change if we include these resistance contributions. R would increase and

/Q A& would decrease. Hence Q& would decrease as well not yielding the maximum energy requirement anymore. It is true though that in most practical situations the amount of energy we need to remove from the freezer is lower that the maximum value.

Steady state convection heat transfer Several cases of convection are of interest to us in food process engineering calculations.

Natural convection occurs when there are differences in density in the bulk of a fluid. This can be exemplified by a hot fluid being in contact with a cold wall can cause such differences. The fluid near the wall will be cooled and thus the density increases. This will cause the cold fluid to move downwards near the wall and hence a natural convection flow occurs.

Forced convection is found where fluid or air is transported past a surface (i.e. by a pump or a ventilator). This type of convection is dominant in flow in closed processing equipment such as heat exchangers or in ventilation systems.

The third case is convection due to cooking or condensation.

The magnitude of energy transported from a solid surface to a fluid (or vice versa) depends on the temperature difference and the transport velocity (the degree of convection). The heat transfer conditions in convection cannot be calculated directly, however many experimental studies have produced numerous suitable empirical relations. Having many empirical expressions it is especially important to know the background and relevant assumptions for these before using them.

Page 54: Food Process Engineering

- 50 -

The task is to determine the film heat transfer coefficient h for this purpose a dimensionless relation between heat transfer due to convection and conduction has been introduced. This number is the Nusselt number (Nu), which relates to h through:

= ⋅Nufluid

dh

k (53)

In the following a relevant selection of the empirical expressions for the Nusselt number is presented. The equations determining the Nusselt number are based on other dimensionless numbers. For obvious reasons the Reynolds number (introduced on page 34) is important for forced convection and the Prandlt number (introduced on page 23) yields information on the fluids applied. The additional dimensionless numbers applied in the Nusselt number expressions are given below.

Grashofs number Gr describes the relation between the buoyancy forces and viscous forces. In other words this is the relation between gravity forces and density differences due to temperature differences. Hence the Grashof number is only of significance for natural convection.

⋅ ⋅β ⋅ρ ∆=µ

3 2

2Grd g T

(54)

where d is a charteristic dimension [m], g is the acceleration of gravity [m/s2], β is

the volumetric expansion coefficient, ∆T is the temperature difference between wall and the bulk fluid and µ is the viscosity. The physical parameters are evaluated at the film temperature = − )( / 2f wall bulkT T T .

Graetz number Gz is a convenient short form of the Reynolds number, the Prandlt number and the diameter to length ratio:

= ⋅ ⋅ Gz Re Pr 4

dl

p (55)

Also the Rayleigh number Ra is combining two other dimensionless numbers for the purpose of conveniences:

⋅ ⋅ ⋅ ∆⋅ = ⋅3 2

Ra=Gr Prd g T

kb r

m (56)

The equations for the Nusselt number take the general form of:

= Re,Pr, , , b

w

dNu f Gr

Lmm (57)

where d /L is the diameter to length ratio which appears in laminar flow, µb the bulk viscosity and µw the viscosity at the wall. The physical properties of the fluid

Page 55: Food Process Engineering

- 51 -

are determined using the average temperature of the fluid unless otherwise indicated (e.g. for µw which is the evaluated at the wall temperature).

Natural convection

For natural convection to air, water or other fluids many relations for the Nusselt number exists covering both highly specialised and more general cases. Most equations are on the form, where the parameters a and b can be found in Table 7:

= ⋅Nu Raba (58)

however, some equations are more elaborate.

Table 7. Relations for Nu valid in natural convection Geometry Characteristic

dimension d in Gr Limits for Ra=Gr⋅Pr

a b Type of flow

Vertical plate Height > 1m < 104 1.36 0.2 Laminar

104<Ra<109 0.55 0.25 Laminar >109 0.13 0.333 Turbulent Vertical cylinder

As vertical plate if: D>(35*height)/Gr0,25

d<0,2 m 103<Ra<109 0.53 0.25 Laminar Sphere &

horizontal cylinder >109 0.13 0.333 Turbulent

Vertical cylinder D 10-5<Ra<1012

⋅ = + +

2

1 6

8/279 16

0.3870.60.5591

Pr

RaNu (59)

Sphere ½πd Ra < 1011 and

Pr > 0,7

⋅ = + +

1/4

4 / 99/16

0.58920.4691

Pr

RaNu (60)

Horizontal plates

d=A/perimeter or edge length ⇐ Applies to the rows below

105<Ra<2⋅107 0,54 0,25 Laminar Cold ceiling or hot floor 2⋅107<Ra<3⋅1010 0,14 0,333 Turbulent Hot ceiling or cold floor

3⋅105<Ra<3⋅1010 0,27 0,25 Laminar

Page 56: Food Process Engineering

- 52 -

The parameters a and b can be found in Table 7. A selection of the most relevant relations is presented along with the required limitations and assumptions in Table 7 and some less often used relations are available in appendix 9. The collection of suitable equations is very large. Therefore it is important to study the intended uses of these carefully before choosing one.

Example 24. Natural convection

The example follows in a separate collection of examples

Forced convection

The diameter referred to in the Nusselt number equations is the inner diameter of the pipe. The relations can be applied for noncircular ducts when substituting the pipe diameter d with the hydraulic diameter Dc:

⋅= 4 (free flow area)wetted perimetercD (61)

Example 25. Calculation of the hydraulic diameter Dc

For a rectangle with sides of the length 0.2m and 0.3m we get:

⋅ ⋅ ⋅ ⋅= = = ⋅ =⋅ ⋅4 ( ) 4 (0.2 0.3 ) 0,06

2 0.242 (W+H) 2 (0.2m+0.3m) 0.05c

W H m mD m m

Laminar flow Newtonian fluids in pipes

Fully developed flow with constant surface temperature:

=Nu 3.66 (62)

Fully develop flow with constant surface heat flux:

=Nu 4.36 (63)

For both the pipe entrance region and fully developed flow conditions:

= ⋅ ⋅ ⋅

0.140.33Nu 1.86 Re Pr b

w

dL

mm (64)

The only physical property not evaluated at the average fluid temperature is µw which is the evaluated at the wall temperature. The constant here 1.86

depends on source reporting the equation and other equally valid constants or modifications to equation (64) can be found in the literature like the one presented in equation (65), where the constant in effect is 1.62 when substituting the expression for Gz (equation (55)) into equation (65):

Page 57: Food Process Engineering

- 53 -

= ⋅

0.141/3Nu 1.75 Gz b

w

mm (65)

Example 26 Laminar flow of Newtonian fluids

Data: The fluid considered is a grape juice with the physical properties. Density 1000 kg/m3, Specific heat capacity: 3.7 kJ/(kg·K) and thermal conductivity 0.5 W/(m2·K) and viscosity 0.113 Pa·s. The flow velocity is 2 m/s and the process takes place in a tube with a diameter of 0.05 m the length is 2 m.

Solution:

⋅= =3700 0.11Pr 814

0.5

⋅ ⋅= ≈0.05 2 1000Re 909

0.11

The flow is laminar. Hence we turn to equation (64) or (65) to determine the Nusselt number. For the ease of calculations we neglect the viscosity correction term.

= ⋅ ⋅ = 0.330.05

Nu 1.86 909 814 47.62

= ⋅ ⋅ ⋅ ⋅ 1/30.05Nu 1.75 ( 909 814 ) = 41.4

4 2p

Comments: We a small difference between the two equations but in practise this will not influence the design of a process plant much.

Transition flow of Newtonian fluids in pipes

For Reynolds numbers between 2100 and 10.000 the following equation is valid:

( ) ( )( ) ( )⋅ − ⋅

=+ ⋅ ⋅ −0.5 2/3

Re 1000 Pr81 12.7 8 Pr 1

fNu

f (66)

where the friction factor f is obtained for smooth pipes using equation (67):

( )=

⋅ − 21

0.790 ln(Re) 1.64f (67)

Turbulent flow of Newtonian fluids in pipes

For Reynolds numbers greater than 10.000 the equation below can be applied:

= ⋅ ⋅ ⋅

0.140.8 0.33Nu 0.023 Re Pr b

w

mm (68)

Page 58: Food Process Engineering

- 54 -

Example 27. Turbulent flow of Newtonian fluids

Data: The fluid considered is milk with the physical properties. Density 1030 kg/m3, Specific heat capacity: 3.9 kJ/(kg·K) and thermal conductivity 0.56 W/(m2·K) and viscosity 1.99·10-3 Pa·s. The flow velocity is 2 m/s and the process takes place in a tube with a diameter of 0.05 m .

Solution:

−⋅ ⋅= =33900 1.99 10

Pr 13.860.56

−⋅ ⋅= ≈⋅ 3

0.05 2 1030Re 51800

1.99 10

The flow is turbulent. Hence we turn to equation (68) to determine the Nusselt number. Again we neglect the viscosity correction.

= ⋅ ⋅ =0.8 0.330.023 51800 13.86 323.6Nu

Adapting calculations of heat transfer coefficients to power law fluids.

When doing calculations for power law fluids we conveniently compute a term, which can substitute the viscosity term (equation (69)) in the relations for Newtonian fluids.

− − ⋅ ⋅ + = = ⋅ (1 ) ( 1)3 1

Re 4

nn nd v m n

R vn

rh (69)

Hence the Prandlt number can be determined by inserting the viscosity in equation (15).

Furthermore we need to be able to evaluate the viscosity in the bulk and at the wall. The apparent viscosity at the wall must follow the theoretical derivations for power law fluids. Hence we substitute the expression for the shear rate at the wall in to the power law equation for the apparent viscosity (equation (18)) an get:

− − + = ( 1) ( 1)2 3 1n n

wv n

md n

h (70)

Laminar flow of power law fluids in pipes

Nusselt number for non-Newtonian fluids is similar to the Newtonian fluid relation in equation (65):

+ =

0.141/31/33 1

Nu 1.75 Gz 4

b

w

nn

hh (71)

Page 59: Food Process Engineering

- 55 -

By comparing equations (65) and (71) it can be seen that considering shear-thinning behaviour is highly relevant as the Nusselt number increases for decreasing n and thereby yields better heat transfer.

Example 28. Comparing Newtonian and non-Newtonian Nusselt numbers

= ⇒ = ⋅0.33 1.15 Newtonskn Nu Nu

Transition flow and turbulent flow of power law fluids in pipes

The procedure is to calculate Reynolds number, Prandlt number and the apparent viscosity and insert in Newtonian equations to obtain a working solution.

Flow over spheres

( ) = + ⋅ + ⋅ ⋅ ⋅

0.140.5 0.667 0.42 0.4 Re 0.060 Re Prd d d

sNu

mm (72)

⋅d s0.71 < Pr < 380; 3.5 < Re < 7.6 104; 1 < ( / ) < 3.2m m

Flow over cylindrical pipes

For flow over cylindrical pipes Churchill and Bernstein proposes the following equation for the Nusselt number:

( )

⋅ ⋅ = + ⋅ + +

4/55/80.5 1/3

1/42/3

0.62 Re Pr Re0.3 1

2820001 0.4 / Pr

dNu (73)

where Re⋅Pr > 0.2

Nusselt number for laminar flow over plates

= ⋅ ⋅0.5 1/30.664 Re PrL LNu , Pr > 0.6 (74)

Nusselt number for turbulent flow over plates

= ⋅ ⋅4/5 1/30.0296 Re Prx xNu , 0.6 < Pr < 60 (75)

Nusselt number for infinite plates

= ⋅ ⋅4/5 1/30.037 Re PrL LNu (76)

Agitated jacketed tanks

For agitated tanks several special case correlations is deducted, however, we just bring a single very general equation here:

Page 60: Food Process Engineering

- 56 -

=

⋅ ⋅ = ⋅ ⋅ ⋅ ∑2/3 0.142

1/3 2

10.36 Pr

nb

D iw i

L NNu Xr mm m (77)

D is the tank diameter (used in Nu), L is the impeller diameter, N is the revolutions of the impeller (per second) and the index w denotes the surfaces of the heating jacket. No specification of the impeller type is specified for the above equation.

Heating and cooling of agitated vessels

An example of a relation valid for both heating and cooling of tanks with specified impeller type is given below (more relations are found in appendix 10):

Propeller impeller (45° blade pitch)

= ⋅ ⋅

0.140.67 0.330.54 Re Pr b

wNu

mm (78)

Unsteady state heat transfer This subject is covered in a separate note – Ikke-stationær varmeledning.

Cooking and condensation

Kondenserende vanddamp

Opvarmning med kondenserende vanddamp er en meget almindelig metode i lev-nedsmiddelindustrien. Dette skyldes dels, at vanddamp indeholder meget energi (ca. 2250 kJ/kg ved kondensation (fortætning)) og altså er et meget energiintensitivt varmeoverføringsmdie, og dels at man kan opnå meget store varmeovergangstal ved kondensation af vanddamp. I lukkede systemer uden luft (ikke kondenserbare gasser) haves ved filmkondensation en h = 12 kW/(m2 K). Hvis der optræder luft falder h kraftigt: ved 3% luft i vanddampen er h ca 3,5 kW/(m2 K) og med 6% ca 1,2 kW/(m2 K).

Sker kondensationen ved at der dannes små dråber, der dannes altså ikke en isole-rende film på overfladen, kan man opnå varmeovergangstal på 25-100 kW/(m2 K).

Kogning

Kogning som følge af varmetilførsel fra en overflade til en væske er en så kompliceret proces, at man ikke kan regne sig frem til brugbare varmeovergangstal. Typen af kogning afhænger af temperaturdifferensen mellem hedefladen og væsken. Er ∆T fra 1-6°C vil der kun dannes få dampbobler og man får et varmeovergangstal på 0,5-2 kW/(m2 K). Ved stigende temperaturdifferens vil der dannes flere bobler der stiger op gennem væsken og forårsager en kraftig omrøring. H ligger i området 2 – 50 kW/(m2 K). Maksimum nås ved en ∆T på 25-75°C. Den eksakte kritiske temperatur afhænger af systemet, specielt overfladens beskaffenhed. Ved større temperaturforskelle vil der begynde at dannes en isolerende dampfilm som medfører at h falder til 1,2 – 1,5 kW/(m2 K). Der nås et minimum ved Leidenfrost punktet, hvorefter stråling begynder at gøre sig gældende og h stiger igen.

Page 61: Food Process Engineering

- 57 -

Heat transfer by radiation

Black body radiation

Alle legemer med en temperatur over det absolute nul-punkt TA = 0 K udsender elektromagnetisk stråling fra overfladen. Ligeledes er disse overflader i stand til at absorbere elektromagnetisk stråling. Den udsendte strålings bølgelængde afhænger af temperaturen, fx får man fra overflader der er varmere end 500°C udsendt synligt lys (glødende legemer).

Stefan – Boltzmanns lov bestemmer den udsendte energimængde ved temperaturen TA K:

−= ⋅ ⋅ ⋅ ⋅& 8 45.72 10 AQ A Te (79)

hvor ε er emisiviteten, der der er 1,000 for et absolut sort legeme og mindre for alle

andre legemer, dog altid positiv. Emisiviteten af et legeme afhænger af bølgelængden af den udsendte stråling.

Ligesom der udsendes stråling absorberes der stråling afhængig af overfladens beskaffenhed. Der udveksles altså hele tiden stråling mellem alle legemer.

Ved beregning af strålingstabet fra udstyr der udstråler til omgivelserne, regnes disse som et absolut sort rum :

−= ⋅ ⋅ ⋅ ⋅ −& 8 4 4, ,5.72 10 ( )A body A environmentQ A T Te (80)

For to planparallelle flader I og II med emisiviteterne εI og εII, sker udvekslingen :

( )−→ ⋅ = ⋅ ⋅ ⋅ − − − ⋅ −

& 8 4 4, ,5.72 10

1 (1 ) (1 )I II

I II A I A III II

Q T Te ee e (81)

For to flader der omslutter hinanden (eksempelvis kugle- eller cylinderflader fås tilsvarende :

( )( )−→

⋅ = ⋅ ⋅ ⋅ ⋅ − + ⋅ ⋅ −

& 8 4 4, ,5.72 10

1

I III II I A I A II

III I II

II

Q A T TAA

e ee e e

(82)

Page 62: Food Process Engineering

- 58 -

Heating and cooling

This part is covered in a separate note

Page 63: Food Process Engineering

- 59 -

Effects on processing The residence time of fluid passing through a continuous process is not uniform for a whole bulk. This is obvious when considering that there is a velocity distribution across the cross-section of a tube. We have explained found that the velocity has its maximum value in the centre of the tube and that it is very close to zero at the wall. This means that we get what is called a residence time distribution. Even though this subject requires much attention when designing processes for production of quality foods it is not covered in to great detail here. Even though no proof is presented here it is not to difficult to accept that the residence time distributions is more narrow for the shear-thinning fluid flow and turbulent flow cases than it for Newtonian fluids in laminar flow and that the situation is the opposite for the flow of shear-thickening fluids. Since we aim to give the most homogeneous processing possible we can conclude that turbulent flow or laminar flow of shear-thinning fluids proves advantageous.

When calculating the operation conditions of a continuous heating process we want to assure that the fastest moving part of the fluid receives sufficient processing as this is the part of the fluid having the shortest residence time.

Page 64: Food Process Engineering

- 60 -

References • Arpaci, V. & Larsen, P.S., 1984. Convection Heat Transfer, Prentice-Hall, Inc.,

Englewood Cliffs, USA.

• Barnes, H.A., Hutton J.F. & Walters, K., 1989. An Introduction to Rheology, Elsevier, Amsterdam, Holland.

• Bird, R.B., Armstrong, R.C., & Hassager, O., 1987. Dynamics of Polymeric Liquids, Volume 1: Fluid Mechanics, John Wiley & Sons, Inc., New York, USA.

• Bolmstedt, U. Web-site (http://chemeng1.kat.lht.se/staff/ulf_b/ulf_b.htm#mainprof)

• Coulson, J.M. & Richardson, J.F., 1996. Chemical Engineering, Volume 1, Fifth edition: Fluid Flow, Heat Transfer and Mass Transfer, Butterworth-Heinemann Ltd, Oxford, UK.

• Fryer P.J., Pyle, D.L. & Rielly, C.D., 1997. Chemical Engineering for the Food Industry, Blackie Academic & Professional, London, UK.

• Hallström, B., Skjöldebrand, C., & Trägårdh, C., 1988. Heat Transfer & Food Products, Elsevier Applied Science, London, UK.

• Holdsworth, S.D., 1992. Aseptic Processing and packaging of Food products, Elsevier Applied Science, London.

• Leniger, H.A. & Beverloo, W.A., 1975. Food Process Engineering, D. Reidel Publishing Company, Dordrecht, Holland.

• Singh, R.P & Heldman, D.R., 2001. Introduction to Food Engineering, Academic Press, London, UK.

• Toledo, R.T., 1991. Fundamentals of food process engineering (2. ed.), Chapman & Hall, New York, USA.

Page 65: Food Process Engineering

- 61 -

List of symbols

A Area [m2]

A, B Fitting parameters [-]

a, b Parameters in natural convection Nusselt numbers

Cp Specific heat capacity [J/(kg K)]

Dc Hydraulic diameter [m]

d Diameter [m]

f friction factor [-]

Hi Enthalpy [J/kg]

K1, K2 Fitting parameters [-]

k Thermal conductivity [W/(m K)]

L Length of tube [m]

n Power-law index [-]

n1, n2 Fitting parameters [-]

Mi Mass [kg]

m Consistency index in power-law model [Pa sn]

&m Mass flow rate [kg/s]

∆P Pressure gradient [Pa] &Q Energy rate [J/s]

Q Energy [J]

R Resistance towards heat transfer [K m2/W]

Rc Drying rate [kg/(m2s]

R, r radius [m]

ry radius of plug flow kernel for Bingham plastic fluids [m]

xi Mass fraction

t Characteristic time [s]

T Temperature [°C or K]

U Overall heat transfer coefficient [W/(m2 K)] &V Volume flow [m3/s]

vmax Maximum velocity in tube [m/s]

v Average velocity in tube [m/s]

a Thermal difussivity [m2/s]

γ& Shear rate [s-1]

η Apparent viscosity [Pa s]

Page 66: Food Process Engineering

- 62 -

0, ∞η η Plateau viscosities (shear-thinning fluids) [Pa s]

bη Bulk viscosity [Pa s]

wη Wall viscosity [Pa s]

µ Newtonian viscosity [Pa s]

ρ Density [kg/m3]

τ Shear stress [Pa]

yτ Apparent yield stress [Pa]

wτ Wall shear stress [Pa]

Γ Characteristic time [s]

De Deborah number

Gr Grashofs number

Gz Graetz number

Nu Nusselt number

Pr Prandtl number

Ra Rayleigh number

Re Reynolds number

Page 67: Food Process Engineering

- 63 -

Appendix 1. Usefull calculations and conversions Relation between volume flow and average velocity:

=& 2

4d

V vp

Relation between mass flow rate and volume flow:

= ⋅& &V m r

The power law model:

Apparent viscosity: −= & ( 1)( ) nmh g

Shear stress: = &( )nmt g

Relation between distance, speed and time:

⋅ =v t L

⋅ = ⋅ ⇒ =1 21 1 2 2

2 1

v tv t v t

v t

Page 68: Food Process Engineering

- 64 -

Appendix 2. Units conversions

Viscosity Pa s = N s/ m2 More to come

Page 69: Food Process Engineering

- 65 -

Appendix 3. Density equations The specific desnsity is found from (T is in °C)

= ⋅ − ⋅ ⋅= ⋅ − ⋅ ⋅

= ⋅ − ⋅ ⋅= ⋅ − ⋅ ⋅

= ⋅ − ⋅ ⋅= ⋅ + ⋅ ⋅ − ⋅

3 1

2 1

3 1

3 1

3 1

2 3

1.3299 10 5.1840 10

9.2559 10 4.1757 10

1.599 10 3.1046 10

1.3115 10 3.6589 10

2,4238 10 2,8063 10

9.9718 10 3.1439 10 3.7574 1

proteine

fat

carbohydrate

fibre

ash

water

T

T

T

T

T

T

rrrrrr −

−⋅

= ⋅ − ⋅ ⋅

3 2

2 1

0

9.1689 10 1.3071 10ice

T

Tr

Page 70: Food Process Engineering

- 66 -

Appendix 4. Heat capacity equations Baseret på sammensætning opdelt i fedt og fedtfrit tørstof (ffts) (Toledo)

Over frysepunktet

= ⋅ + ⋅ + ⋅, 1674.72 837.36 4186.8p snit fat ffts wc x x x J/(kg K)

fuldstændig frosset :

= ⋅ + ⋅ + ⋅, 1674.72 837.36 2093.4p frozen fat ffts wc x x x J/(kg K)

Temperaturafhængig beregning (for flydende produkter) (T is in °C)

− −

− −

− −

= + ⋅ ⋅ − ⋅ ⋅= + ⋅ ⋅ − ⋅ ⋅

= + ⋅ − ⋅= + ⋅ − ⋅ ⋅

= +

3 6 2,

3 6 2,

3 6 2,

6 2,

,

2.0082 1.2089 10 1.3129 10

1.5488 1.9625 10 5.9399 10

1.8459 1.8306 10 * 4.6509 10 *

1.9842 1.4733 4.8008 10

1.0926 1.8896

p protein

p carbohydrate

p fibre

p lipid

p ash

c T T

c T T

c T T

c T T

c −

− −− °− −

− − °−

⋅ − ⋅ ⋅= − ⋅ ⋅ + ⋅ ⋅= − ⋅ ⋅ + ⋅ ⋅

= + ⋅ ⋅= ⋅∑

6 2

5 6 2, ,(0 150 )

3 4 2, ,( 40 0 )

3,

, ,

3.6817 10

4.1762 9.0864 10 5.4731 10

4.0817 5.3062 10 9.9516 10

2.0623 6.0769 10

p w C

p w C

p ice

p product p i i

T T

c T T

c T T

c T

c c x

Leddet for fibre tages kun med hvis de ikke er inkluderet i mængden af kulhydrat.

For frosne produkter skal der tages hensyn til hvor meget af vandet der er udfrosset. Antager vi at alt vand er udfrosset kan man erstatte cp for vand i formlerne med cp for is i den del af vandet der ikke er frosset.

Beregning af dette se senere (vanddel i frosset produkt)

Page 71: Food Process Engineering

- 67 -

Appendix 5. Heat conductivity equations For flydende produkter har Choi og Okos (1986) foreslået følgende måde til bestemmelse af varmeledningsevnen :

= ⋅

=

Vi i

i

iV ii

i

i

k k X

x

Xx

r

r

De specifikke størrelser af varmeledningsevner findes af (T is in °C):

− − −

− − −

− − −

− − −

= ⋅ + ⋅ ⋅ − ⋅ ⋅= ⋅ − ⋅ ⋅ − ⋅ ⋅

= ⋅ + ⋅ ⋅ − ⋅ ⋅= ⋅ + ⋅ ⋅ − ⋅ ⋅

=

1 3 6 2

1 3 7 2

1 3 6 2

1 3 6 2

1.7881 10 1.1958 10 2.7178 10

1.8071 10 2.7604 10 1.7749 10

2.0141 10 1.3874 10 4.3312 10

1.8331 10 1.2497 10 3.1683 10

protein

fat

carbohydrate

fiber

ash

k T T

k T T

k T T

k T T

k − − −

− − −

− −

⋅ + ⋅ ⋅ − ⋅ ⋅= ⋅ + ⋅ ⋅ − ⋅ ⋅= − ⋅ ⋅ + ⋅ ⋅

1 3 6 2

1 3 6 2

3 4 2

3.2962 10 1.4011 10 2.9069 10

5.7109 10 1.7625 10 6.7036 10

2.2196 6.2489 10 1.0154 10

w

ice

T T

k T T

k T T

Beregninger af ikke homogene systemer En del produkter kan betragtes som bestående af lag af ellers homogene komponenter. Beregningerne af disses varmeledningsevner sker ved :

Lag i parallel (jf. figur) :

= ⋅= ⋅ + ⋅

+ =

∑1 1 2 2

1 2 1

i i

parallel

k k

k k k

ee e

e e

epsilon er porositet eller volumenfraktion.

Lag i serie (jf. figur) :

=

= +

+ =

∑1 2

1 2

1 2

1

1

1

i

i

series

k k

k k k

e

e e

e e

En fase dispergeret i en anden kontinuerlig fase :

Page 72: Food Process Engineering

- 68 -

⋅ + − ⋅ − = ⋅ ⋅ + + ⋅ −

1 2 2 1 21

1 2 2 1 2

2 2 ( )2 ( )disp

k k k kk k

k k k ke

e

For produkter der indeholder luft (er svulmet eller opskummet) kan man bestemme porositeten (ε) ud fra massefylderne :

= −

1 producta

product without air

re r

hvor ρprodukt bestemmes ud fra aktuelt volumen og vægt efter svulmning (det kan

være en designparameter) og varmeledningsevnen ellers bestemmes ud fra de foregående ligninger.

Page 73: Food Process Engineering

- 69 -

Appendix 6. Measuring viscosity and flow curves Under construction

Page 74: Food Process Engineering

- 70 -

Appendix 7. Moody diagram Under construction

Page 75: Food Process Engineering

- 71 -

Appendix 8. Data for water, air + foods Please refer to Fellows

Page 76: Food Process Engineering

- 72 -

Appendix 9. Equations for natural convection More equations for h.

Table XX Simplified equations determining h in natural convection for air or water

Geometry Fluid Equation

Air ∆ = ⋅

0.251.3196

Th

d

Horizontal cylinder heated or cooled

Water ∆ = ⋅

0.25291.1

Th

d

Fluid above heated horizontal plate Air ( )= ⋅ ∆ 0.252.4493h T

Fluid below heated horizontal plate Air ( )= ⋅ ∆ 0.251.3154h T

Fluid above cooled horizontal plate Air ( )= ⋅ ∆ 0.251.3154h T

Fluid below cooled horizontal plate Air ( )= ⋅ ∆ 0.252.4493h T

Air ∆ = ⋅

0.251.3683

Th

d

Vertical cylinder heated or cooled

Water ∆ = ⋅

0.25127.1

Th

d

Air ∆ = ⋅

0.251.3683

Th

d

Vertical plate heated or cooled

Water ∆ = ⋅

0.25127.1

Th

d

Page 77: Food Process Engineering

- 73 -

Appendix 10. Equations for forced convection Free falling droplets

For en frit faldende dråbe I luft har Ranz og Marshall benyttet :

= + ⋅ ⋅0.5 1/32 0.6 Re Prd dNu

Drying

Ved tørring (dvs. ofte med meget fugtig luft) er det foreslået for strømning omkring legemer at følgende gælder :

= ⋅ 0.50.75 Rex xNu

Hvor x findes af :

Genstandens form Karakteristisk dimension

X

Flad plade Længde L L Cylinder på tværs Diameter d d*(π/2) Kugle Diameter d d*(π/2) Cirkulær skive Diameter d d*(π/4) Trekantede prismer Kantlængden b b*(3/2) Åbne kantede legemer Tværsnit b B

H korrigeres for fugtigheden med faktoren : (Pr/0,7)0,33

Agitated tanks

For en beholder med indvendige varmespiraler (rør):

⋅ ⋅ = ⋅ ⋅ ⋅

2/3 0.1421/30.90 Pr b

Dw

L NNu r mm m

D is the tank diameter (used in Nu), L is the impeller diameter, N is the revolutions (per s) and the index s denotes the surfaces of the heating spiral.

Der er ikke oplyst noget om hvilken type omrører der er anvendt.

For opvarmning og køling i beholdere med mere specificerede omrørere gælder :

Turbineomrører med flade blade:

= ⋅ ⋅ ⋅

0.140.67 0.330.54 Re Pr b

wNu

mm , Re < 400, beregnet på hvilken

hastighed ?

= ⋅ ⋅ ⋅

0.140.67 0.330.74 Re Pr b

wNu

mm , Re > 400, beregnet på

Turbiner med bagudrettede blade (retreating blade)

Page 78: Food Process Engineering

- 74 -

µ = ⋅ ⋅ ⋅ µ

0.140.67 0.330.68 Re Pr b

wNu

Spiralblandere (helical ribbon)

− − = ⋅ ⋅ ⋅ ⋅ ⋅

0.14 0.22 0.280.5 0.330.248 Re Pr b

w

e iNu

D Dmm , Re < 130

e er afstanden (lysning) mellem beholdervæg og omrører, i er stigningen i spiralen

− = ⋅ ⋅ ⋅ ⋅

0.14 0.250.67 0.330.238 Re Pr b

w

iNu

Dmm , Re > 130

Paddel (paddle) :

= ⋅ ⋅

0.240.67 0.330.415 Re Pr ´b

wNu

mm , 20 < Re < 4000

= ⋅ ⋅

0.240.67 0.330.36 Re Pr b

wNu

mm , Re > 4000

Ankeromrører (anchor) :

= ⋅ ⋅ ⋅

0.180.67 0.331.0 Re Pr b

wNu

mm , 30 < Re < 300, afstand til væg < 2,54

cm

= ⋅ ⋅ ⋅

0.180.67 0.330.38 Re Pr b

wNu

mm , 300 < Re < 4000, afstand til væg <

2,54 cm

= ⋅ ⋅ ⋅

0.140.67 0.330.55 Re Pr b

wNu

mm , 4000 < Re < 37.000, afstand til væg

mellem 2,54 og 14 cm.

For beholdere med interne varmerør (spiraler) :

Tilbagebøjede blade :

= ⋅ ⋅

0.140.62 0.331.4 Re Pr b

wNu

mm

Propeller :

Page 79: Food Process Engineering

- 75 -

= ⋅ ⋅

0.140.62 0.330.078 Re Pr b

wNu

mm

Padler :

= ⋅ ⋅

0.140.62 0.330.87 Re Pr b

wNu

mm

Eksempler