modeling notes complete

144
INTRODUCTION TO MATHEMATICAL MODELING OF CHEMICAL PROCESSES Emad Ali, AbdelHamid Ajbar and Khalid Alhumaizi King Saud University College of Engineering Chemical Engineering department June 2006

Upload: harsha

Post on 04-Mar-2015

3.816 views

Category:

Documents


6 download

TRANSCRIPT

Page 1: Modeling Notes Complete

INTRODUCTION TO MATHEMATICAL

MODELING OF CHEMICAL PROCESSES

Emad Ali, AbdelHamid Ajbar and Khalid Alhumaizi

King Saud University College of Engineering

Chemical Engineering department

June 2006

Page 2: Modeling Notes Complete

i

Table of Contents 1. Chapter 1: Fundamentals 1

1.1. Introduction 1 1.2. Incentives for Process Modeling 1.3. Systems 3

1.3.1. Classification based on Thermodynamics 3 1.3.2. Classification based on number of phases 4

1.4. Classification of Models 4 1.5. State Variables and State Equations 5 1.6. Classification of theoretical models 6

1.6.1. Steady State vs Unsteady State 6 1.6.2. Lumped vs Distributed Parameters 6 1.6.3. Linear vs Non-Linear 7 1.6.4. Continuous vs Discrete 7 1.6.5. Deterministic vs Probabilistic 8

1.7. Building Steps for a Mathematical Model 1.8. Conservation Laws 9

1.8.1. Total Mass Balance 10 1.8.2. Component Balance 11 1.8.3. Momentum Balance 11 1.8.4. Energy Balance 12

1.9. Microscopic Balance 12 1.10. Macroscopic Balance 14 1.11. Transport rates 14

1.11.1. Mas Transport 15 1.11.2. Momentum Transport 16 1.11.3. Energy Transport 18

1.12. Thermodynamic relations 19 1.13. Phase Equilibrium 21 1.14. Chemical Kinetics 24 1.15. Control Laws 25 1.16. Degrees of Freedom 26 1.17. Model Solution 27 1.18. Model Validation 27

2. Chapter 2: Examples of Mathematical Models for Chemical processes 29 2.1. Examples of Lumped parameter Systems 29

2.1.1. Liquid Storage tank 29 2.1.2. Stirred Tank Heater 33 2.1.3. Isothermal CSTR 39 2.1.4. Gas-Phase Pressurized CSTR 43 2.1.5. Non-Isothermal CSTR 45 2.1.6. Mixing Processes 48 2.1.7. Heat Exchanger 53 2.1.8. Heat Exchanger with Steam 56

Page 3: Modeling Notes Complete

ii

2.1.9. Single Stage Heterogeneous Systems 58 2.1.10. Two-Phase Reactor 60 2.1.11. Reaction with Mass Transfer 64 2.1.12. Multistage Heterogeneous Systems 67 2.1.13. Binary Absorption Column 71 2.1.14. Multi-Component Distillation Column 73

2.2. Examples of Distributed Parameter Systems 82 2.2.1. Liquid Flow in Pipe 82 2.2.2. Velocity profile in a Pipe 84 2.2.3. Diffusion with Chemical Reaction in a Slab catalyst 86 2.2.4. Temperature Profile in a heated Cylindrical Rod 89 2.2.5. Isothermal Plug Flow Reactor 91 2.2.6. Non-Isothermal Plug-Flow Reactor 94 2.2.7. Heat Exchanger: Distributed Parameter Model 97 2.2.8. Mass Exchange in a Packed Column 99

3. Chapter 3: Equations of Change 103 3.1. Total Mass Balance 103 3.2. Component Balance Equation 106 3.3. Momentum balance Equation 109 3.4. Energy balance 113 3.5. Conversion between the Coordinates 116

3.5.1. Balance Equations in Cartesian Coordinates 117 3.5.2. Balance equation in Cylindrical Coordinates 118 3.5.3. Balance Equation in Spherical Coordinates 119

3.6. Examples of Application of equations of Change 121 3.6.1. Liquid Flow in a Pipe 121 3.6.2. Diffusion with Chemical Reaction in a Slab Catalyst 121 3.6.3. Plug Flow Reactor 122 3.6.4. Energy Transport with Heat Generation 123 3.6.5. Energy Transport in a Circular Tube 125 3.6.6. Unsteady State Heat Generation 127 3.6.7. Laminar Flow Heat Transfer with Constant Wall Temperature 127 3.6.8. Laminar Flow and Mass Transfer 129

Problems (Lumped Parameter Systems) 131 Problems (Distributed Parameter Systems) 139

Page 4: Modeling Notes Complete

1

Chapter 1 : Fundamentals

1.1 Introduction

The objective of mathematical modeling is the development of sets of

quantitative (mathematical) expressions that capture the essentials aspects of an existing

system. A mathematical model can assist in understanding the complex physical

interactions in the system and the causes and effects between the system variables

.Mathematical models are valuable tools since they are abstract equations that can be

solved and analyzed using computer calculations. It is therefore safer and cheaper to

perform tests on the model using computer simulations rather than to carry out

repetitive experimentations and observations on the real system. This becomes vital if

the real system is new, hazardous, or expensive to operate . Modeling, thus prevails the

field of science, engineering and business. It is used to assist in the design of equipment,

to predict behavior, to interpret data, to optimize resources and to communicate

information.

1.2 Incentives for process modeling

In the chemical engineering field, models can be useful in all the phases, from

research and development to plant operation. Models and their simulation are tools

utilized by the chemical engineer to help him analyze the process in the following ways:

• Better understanding of the process

Models can be used to study and investigate the effects of various process

parameters and operating conditions on the process behavior. It can also be used

to evaluate the interactions of different parts of the process. This analysis can be

carried out easily on a computer simulation without interrupting the actual

process, thus avoiding any delay or upsets for the process.

• Process synthesis and design

Page 5: Modeling Notes Complete

2

Model simulation can be utilized in the evaluation of equipment's size and

arrangements and in the study of alternative process flow-sheeting and strategies.

Furthermore, to verify the reliability and safety of the process design tests can be

carried out even prior to plant commissioning.

• Plant operators training

Models can be used to train plant personnel to simulate startup and shutdown

procedures, to operate complex processes and to handle emergency situations and

procedures .

• Controller design and tuning

Models help in developing and evaluating better controller structure and

configuration. Dynamic simulation of models is usually employed for testing and

assessing the effectiveness of various controller algorithms. It is worthwhile to

mention that models play a vital part in designing advanced model-based control

algorithms such as model predictive and internal model controllers .Moreover it

is a common practice of many control engineers to determine the optimum values

of the controller settings through dynamic simulation .

• Process optimization

It is desirable from economic standpoint to conduct process optimization before

plant operation to determine the optimum values of the process key parameters

or/and operating conditions that maximizes profit and reduces cost. Process

optimization is also performed during process operation to account for variations

in the feed-stock and utilities market and for changing environmental regulations.

It is worth mentioning that despite all their usefulness, models at their best are no more

than approximation of the real process since they do not necessarily incorporate all the

features of the real system . Therefore modeling can not eliminate completely the need

for some plant tests, especially to validate developed models or when some poorly

known parameters in the process need to be experimentally evaluated.

Models can be classified in a number of ways. But since mathematical models

are developed from applying the fundamental physical and chemical laws on a specific

system,

Page 6: Modeling Notes Complete

3

1

we review first the classification of systems since their nature affect the modeling

approach and the resulting model.

1.3 Systems

A system is a whole consisting of elements or subsystems. The system has

boundaries that distinguish it from the surrounding environment (external world) as

shown in Figure 1.1.The system may exchange matter and/or energy with the

surrounding through its boundary. Consequently, the state of a system can be defined or

understood via the interactions of its elements with the external world. A system may

be classified in different ways, some of which are as follows:

1.3.1 Classification based on thermodynamic principles

• Isolated system

This type of system does not exchange matter nor energy with the surrounding.

Adiabatic batch reactor is an example of such systems.

• Closed system

This type of system does not exchange matter with the surrounding but it does

exchange energy. Non-adiabatic batch reactor is an example of such systems.

• Open system

This system exchanges both matter and energy with the external environment.

An example of this system is the continuous stirred tank reactor (CSTR).

Page 7: Modeling Notes Complete

4

System

Boundary

Figure 1.1: system and its boundary

Suroundings

1.3.2 Classification based on number of phases

• Homogeneous system

This is a system that involves only one phase such as gas-phase or liquid-phase

chemical reaction processes.

• Heterogeneous system

This is a system that involves more than one phase. This kind of systems exists in

multi-phase reaction processes and in phase-based separation processes.

1.4 Classification of Models

Models can be classified according to how they are derived:

• Theoretical models.

These are models that are obtained from fundamental principles, such as the laws

of conservation of mass, energy, and momentum along with other chemical

principles such as chemical reaction kinetics and thermodynamic equilibrium,

etc. This first-principle model is capable of explaining the underlying physics of

the process and is often called `phenomenological model'. For this reason it is

particularly suitable for process design and optimization . Theoretical models are,

however, generally difficult to obtain and sometimes hard to solve.

Page 8: Modeling Notes Complete

5

• Empirical models

These models are based on experimental plant data. These models are developed

using data fitting techniques such as linear and non-linear regression. Models

obtained exclusively form experimental plant data are also known as black-box

models . Such models do not provide detailed description of the underlying

physics of the process. However, they do provide a description of the dynamic

relationship between inputs and outputs. Thus they are sometimes more adequate

for control design and implementation.

• Semi-empirical models

These models are somehow between the two previous models where uncertain or

poorly known process parameters are determined from plant data.

This book focuses only on developing theoretical models. The interested reader in

empirical modeling is referred to books listed in the references.

1.5 State variables and state equations

Once the system has been classified, developing a theoretical model for it

amounts to characterizing its behavior at any time and at any spatial position. For most

processing systems a number fundamental quantities are used to describe the natural

state of the system. These quantities are the mass, energy and momentum. Most often

these fundamental quantities can not be measured directly thus they are usually

represented by other variables that can be measured directly and conveniently. The most

common variables are density, concentration, temperature pressure and flow rate. There

are conveniently called 'state variables' since they characterize the state of the

processing system.

In order to describe the behavior of the system with time and position, the state

(dependent) variables should be linked to the independent variables (time, spatial

position) through sets of equations that are derived from writing mass, energy and

momentum balances. The set of equations describing these variables are called 'state

equations '.

Page 9: Modeling Notes Complete

6

1.6 Classification of theoretical models

With this in mind, theoretical models may be further classified in more practical

ways as discussed in the following.

1.6.1 Steady state Vs. unsteady state

When the physical state of the processing system remains constant with the time,

the system is said to be at steady state. Models that describe steady state situations are

also called static, time-invariant or stationary models. Basically, almost all chemical

process unit designs are carried out on static models. On the other hand unsteady state

processes represent the situation when the process state (dependent variables) changes

with time. Models that describe unsteady-state situations are also called dynamic and

transient models .Such models are useful for process control design and development.

Process dynamics are encountered in practice during startup, shutdown, and upsets

(disturbances )

1.6.2 Lumped Vs. distributed parameters

Lumped parameters models are those in which the state variables and other

parameters have/or assumed to have no spatial dependence, i.e. they are considered to

be uniform over the entire system. In this case the time (for unsteady state models) is the

only independent variable. The chemical engineering examples for this case include the

perfectly mixed CSTR, distillation columns. etc. Conceptually, these models are

obtained through carrying out a macroscopic balance for the process as it will be

discussed in chapter 2 . On the other hand, distributed parameters models are those in

which states and other variables are function of both time and spatial position. In this

case, modeling takes into account the variation of these variables with time and from

point to point throughout the entire system. Some examples of such systems include

plug flow reactor, heat exchangers, and packed columns. These models are essentially

obtained through writing microscopic balances equations for the process as it will be

discussed later in chapter 2.

Page 10: Modeling Notes Complete

7

1.6.3 Linear Vs non-linear

Linear models have the important property of superposition whereas nonlinear

models do not. Superposition means that the response of the system to a sum of inputs is

the same as the sum of responses to the individual inputs .In linear models all the

dependant variables or their derivatives appear in the model equations only to the first

power. These properties do not hold for nonlinear models . In this respect, it is

important to recognize the fact that most physical and chemical systems are nonlinear.

Linear models are commonly obtained through linearization of the nonlinear model

around a certain steady state. Linear models obtained this way are valid approximation

of the original non linear model only in the neighborhood of the selected state .

1.6.4 Continuous Vs discrete

When the dependant variables can assume any values within an interval the

model is called continuous .When on the other hand one or several variables are

assumed to take only discrete values, the model is called discrete. In chemical

engineering discrete models arise for example when some variables are required to take

only integer values, for example the number of stages in a distillation column, the

number of heat exchanger in a plant, etc.

1.6.5 Deterministic Vs probabilistic

Deterministic models are those in which each variable can be assigned a definite

fixed number, for a given set of conditions. On the other hand in probabilistic or

stochastic models some or all the variables used to describe the system are not precisely

known. They are considered as random variables. Probabilistic models are often

encountered when modeling systems that are subject to noise.

Page 11: Modeling Notes Complete

8

1.7 Building steps for a mathematical model

Building a theoretical mathematical model of a processing system requires the

knowledge of the physical and chemical interactions taking place within the boundaries

of the system. With a given degree of fundamental knowledge of the system at a certain

stage one can build different models with different degree of complexity depending on

the purpose of the model building and the level of rigor and accuracy required .The

choice of the level of rigor and degree of sophistication is in itself an art that requires

much experience. Theoretical modeling of chemical processes may encounter

difficulties that can be classified as follows:

• Problems arising from processes that exhibit complicated physical and chemical

phenomena. Such problems appear for instance in multi-component interactive

processes .

• Problems arising from imprecisely known process parameters. This situation can

be handled by periodically estimating the unknown parameters from plant data.

• Problems arising from the size and complexity of the ensuing model. These can

be overcomed by proper use of simplifying assumptions.

Consequently a modeler should practice careful utilization of the simplifying

assumptions based on engineering sense and experience. Failure to do so, the modeler

may fall into one of the two extremes, i.e. creating rigorous but over complicated model

or creating oversimplified model that does not capture all the critical features of the true

process.

The general procedure for building up a mathematical model includes the following

steps:

Page 12: Modeling Notes Complete

9

• Identification of the system configuration, its surrounding environment and the

modes of interaction between them, the identification of the relevant state

variables that describe the system and identification of the process taking place

within the boundaries of the system

.

• Introduction of the necessary simplifying assumptions.

• Formulation of the model equations based on principles of mass, energy and

momentum balances appropriate to the type of the system. This also requires the

determination of the fundamental quantitative laws (chemical kinetics,

thermodynamic relations…) that govern the rates of the process in terms of the

state variables .

• Determination of the solvability of the model using degree of freedom analysis.

• Development of the necessary numerical algorithms for the solution of the model

equations.

• Validation of the model against experimental results to ensure its reliability and

to re-evaluate the simplifying assumptions which may result in imposing new

simplifying assumptions or relaxing others.

1.8 Conservation Laws

As mentioned in previous section, the state equation forming the mathematical

model defines a relationship between the state variables (dependent variables) and the

independent variables, i.e., time and spatial variables of the system. These equations are

derived, from applying the conservation law for a specific fundamental quantity say S,

on a specific system with defined boundaries (see Figure 1.2) as follows:

Page 13: Modeling Notes Complete

10

System

Figure 1.2

Sin

S

Sout

Total flow

rate of (S)

into the

system

+

Generation

rate of (S)

within system

=

Total flow

rate of (S) out

of the system

+

Accumulation

rate of (S)

within system

±

amount of (S)

exchanged with

the surrounding

(1.1)

The quantity S can be any one of the following quantities:

• Total Mass

• Component Mass (Mole)

• Total Energy

• Momentum

1.8.1 Total mass balance

Since mass is always conserved, the balance equation for the total mass (m) of a given

system is:

Rate of mass in = rate of mass out + rate of mass accumulation (1.2)

Page 14: Modeling Notes Complete

11

The mass balance equation has the SI unit of kg/s.

1.8.2 Component balance

The mass balance for a component A is generally written in terms of number of moles

of A. Thus the component balance is

flow of moles

(A) in

+

Rate of Generation

of moles of (A)

=

Flow of moles

of (A) out

+

Rate of Accumulation

of moles of (A) (1.3)

The component balance has the unit of moles A/s. It should be noted that unlike the total

mass, the number of moles of species A is not conserved. The species A can be

generated or consumed by chemical reaction.

1.8.3 Momentum balance

The linear momentum (π) of a mass (m) moving with velocity (v) is defined as:

π = mv (1.4)

Since the velocity v is a vector, the momentum, unlike the mass is also a vector. The

momentum balance equation using (Eq 1.1) is:

Rate of

momentum in

+

Rate of Generation

of momentum

=

Rate of

momentum out

+

Rate of Accumulation

of momentum

(1.5)

The momentum balance has the unit of kg.m/s2. The momentum balance equation is

usually written using the Newton's second law. The law states that the time rate of

change of momentum of a system is equal to the sum of all forces F acting on the

system,

Page 15: Modeling Notes Complete

12

∑= Fdtmvd )(

(1.6)

1.8.4 Energy balance

The energy balance for a given system is:

Rate of

energy

in

+

Rate of

Generation of

energy

=

Rate of

energy

out

+

Rate of

accumulation

of energy

± amount of

energy

exchanged

with the

surrounding

(1.7)

The energy generated within a system includes the rate of heat and the rate of work. The

rate of heat includes the heat of reaction (if a reaction occurs in the system), and the heat

exchanged with the surroundings. For the rate of work we will distinguish between the

work done against pressure forces (flow work) and the other work such as the work

done against the gravity force, against viscous forces and shaft work. The reason for this

distinction will appear clearly in the next chapter.

For a given system the general conservation law (Eq. 1.1) can be carried out either on

microscopic scale or macroscopic scale.

1.9 Microscopic balance

In the microscopic case, the balance equation is written over a differential

element within the system to account for the variation of the state variables from point

to point in the system, besides its variation with time. If we choose for example

cartesian coordinates, the differential element is a cube as shown in Figure 1.3. Each

state variable V of the system is assumed to depend on the three coordinates x,y and z

plus the time. i.e. V = V(x,y,z,t). The microscopic balance can be also written in

cylindrical coordinates (Figure 1.4) and in spherical coordinates (Figure 1.5). The

selection of the appropriate coordinates depends on the geometry of the system under

Page 16: Modeling Notes Complete

13

study . It is possible to convert from one coordinate system to an other as it will be

discussed in chapter 3.

∆x

∆z∆y

x

z

y

Figure 1.3: Cartesian coordinates

ry

z

x

θ

z

Figure 1.4: Cylindrical coordinates

Page 17: Modeling Notes Complete

14

r zθ

φ

y

x

z

Figure 1.5: Spherical coordinates

1.10 Macroscopic balance

In some cases the process state variables are uniform over the entire system, that

is each state variable does not depend on the spatial variables, i.e. x,y and z in cartesian

coordinates but only on time t. In this case the balance equation is written over the

whole system using macroscopic modeling . When modeling the process on

microscopic scale the resulting models consists usually of partial differential equation

(PDE) where time and one or more spatial position are the independent variables. At

steady state, the PDE becomes independent of t and the spatial positions are the only

independent variables .When, on the other hand, the modeling is based on macroscopic

scale the resulting model consists of sets of ordinary differential equations (ODE) . In

the next chapter we present examples of applying macroscopic and microscopic

balances to model various chemical processes.

The fundamental balance equations of mass, momentum and energy already

discussed are usually supplemented with a number of equations associated with

transport rates and thermodynamic relationships. In the following we present an

overview of some of these relations.

1.11 Transport rates

Page 18: Modeling Notes Complete

15

Transport of the fundamental quantities, mass, energy and momentum occur by two

mechanisms

• Transport due to convection or bulk flow

• Transport due to molecular diffusion or potential difference .

In many cases the two transport mechanism occur together. Therefore, the flux due to

the transport of any fundamental quantity is the sum of a flux due to convection and a

flux due to diffusion.

1.11.1 Mass Transport

The total flux nAu (kg/m2s) of species A of density ρA (kg/m3) flowing with velocity vu

(m/s) in the u-direction is the sum of the two terms:

nAu = jAu + ρAvu (1.8)

total flux = diffusive flux + bulk flux (1.9)

The diffusive flux jAu (kg/m2s) for a binary mixture A-B is given by Fick's law:

dudwDj A

ABAu ρ−= (1.10)

where wA = ρA/ρ is the mass fraction of species Α, ρ (kg/m3) the density of the mixture

and DAB (m2/s) is the diffusivity coefficient of A in the mixture. In molar unit the flux

JAu (mol A/m2s) is given by:

dudxCDJ A

ABAu −= (1.11)

Page 19: Modeling Notes Complete

16

where C is the total concentration of A and B (Kg(A+B)/m3) and xA = CA/C is the mole

fraction of A in the mixture. For constant density the flux Eqs.( 1.10 and Eq. 1.11)

become:

dudDj A

ABAuρ

−= (1.12)

dudC

DJ AABAu −=

(1.13)

1.11.2 Momentum transport

Momentum is also transported by convection and diffusion. But unlike the mass,

the linear momentum π = mv is a vector. We have to consider then its transport in all

directions (x,y,z) of a given system. Let consider for instant the transport of the x-

component of the momentum. Similar analysis can be carried out for the transport of the

y and z component.

The flux due to convection of the x-component of the momentum in the y-direction, for

instant, is

(ρvx)vy (kg.m/s2) (1.14)

To determine the diffusion flux denoted by τyx of the x-component of the momentum in

the y-direction, consider a fluid flowing between two infinite parallel plates as shown in

Figure 1.6. At a certain time the lower plate is moved by applying a constant force Fx

while the upper plate is maintained constant. The force Fx is called a shear force since it

is tangential to the area Ay on which it is applied (Fig. 1.6).

Page 20: Modeling Notes Complete

17

∆ Vx

∆ y F, force

x

y

Figure 1.6: Momentum transfer between two parallel plates

The force per unit area

)/.( 2smkgAF

y

x (1.15)

is called a stress and denoted τyx (kg.m/s). It is also a shear stress since it is tangential.

The force Fx imparts a constant velocity V = vx (y = 0 ) to the layer adjacent to the plate.

Because of molecular transport, the layer above it has a slightly slower velocity vx(y)

and so on as shown in Figure 1.6. Therefore, there is a transport by diffusion of the x-

component of the momentum in the y-direction. The flux of this diffusive transport is in

fact the shear stress τyx.

Therefore, the total flux πyx of the x-component in the y-direction is the sum of the

convection term (Eq. 1.14) and diffusive term τyx

πyx = τyx + (ρvx)vy (1.16)

momentum flux = diffusive flux + bulk flux (1.17)

For a Newtonian fluid the shear stress τyx is proportional to the velocity gradient:

yvx

yx ∂∂

−= µτ (1.18)

Page 21: Modeling Notes Complete

18

where µ (kg/m.s) is the viscosity of the fluid.

1.11.3 Energy transport

The total energy flux eu (J/s.m2) of a fluid at constant pressure flowing with a velocity vu

in the u-direction can be expressed as:

eu = qu + (ρCpT)vu (1.19)

energy flux = diffusive flux + bulk flux (1.20)

The heat flux by molecular diffusion, i.e. conduction in the u-direction is given by

Fourier's law:

uTkqu ∂

∂−=

(1.21)

where k (J/s.m.K) is the thermal conductivity.

The three relations (Eq. 1.10, 1.18, 1.21) show the analogy that exists between mass,

momentum and energy transport. The diffusive flux in each case is given by the

following form:

Flux = − transport property × potential difference

(gradient)

(1.22)

The flux represents the rate of transfer per area, the potential difference indicates the

driving force and the transport property is the proportionality constant . Table 1.1

summarizes the transport laws for molecular diffusion.

Table 1.1: One-dimensional Transport laws for molecular diffusion

Transport

Type

Law Flux Transport

property

gradient

Page 22: Modeling Notes Complete

19

Mass Fick's JAu D du

dC A

Heat Fourrier qu k dudT

Momentum Newton τux µ dudvx

When modeling a process on macroscopic level we can also express the flux by a

relation equivalent to (Eq. 1.22). In this case the gradient is the difference between the

bulk properties, i.e. concentration or temperature in two medium in contact, while the

transport property represents an overall transfer coefficient. For example for mass

transfer problems, the molar flux can be expressed as follows:

AA CKJ ∆×= (1.23)

where K an overall mass transfer coefficient.

The heat flux on the other hand is expressed as

TUq ∆×= (1.24)

where U is an overall heat transfer coefficient.

As for the momentum balance, the macroscopic description generally uses the pressure

drop as the gradient while the friction coefficient is used instead of the flux. The

following relation is for instance commonly used to describe the momentum laminar

transport in a pipe

PLv

Df ∆=ρ22

(1.25)

f is the Fanning friction factor, ∆P is the pressure drop due to friction, D is the diameter

of the pipe, L the length and v is the velocity of the fluid.

1.12 Thermodynamic relations

Page 23: Modeling Notes Complete

20

An equation that relates the volume (V) of a fluid to its temperature (T) and pressure

(P) is called an equation of state. Such equations are used to determine fluid densities

and enthalpies .

• Densities :

The simplest equation of state is the ideal gas law:

PV = nRT (1.26)

which can be used to determine the vapor (gas) density

ρ = MwP/RT (1.27)

where Mw is the molecular weight . As for liquids, tabulated values of density

can be used and can be considered invariant unless large changes in composition

or temperature occur.

• Enthalpies :

Liquid and vapor enthalpies for pure component can be computed from simple

formulas, based on neglecting the pressure effect as follows:

)(~refp TTCh −= (1.28)

λ+−= )(~refp TTCH (1.29)

Where h~ is the liquid specific enthalpy, H~ is its vapor specific enthalpy, pC

the average liquid heat capacity and λ is the latent heat of vaporization

Note that the reference condition is taken to be liquid at temperature Tref. If heat

of mixing is negligible, the enthalpy of a mixture can be taken as the sum of the

specific enthalpies of the pure components multiplied by their corresponding

Page 24: Modeling Notes Complete

21

mole fractions Note that the above enthalpy functions are valid for small

temperature variation and/or when the heat capacities of fluids are weak function

of temperature. In general cases the heat capacity Cp can be taken as function of

temperature such as:

Cp = a + bT + cT2 + dT3 (1.30)

The specific enthalpies for vapors and liquids are computed as integrals as

follows:

∫=T

Tp

ref

dTCh~ (1.31)

λ+= ∫T

Tp

ref

dTCH~ (1.32)

• Internal energy :

The internal energy U is a fundamental quantity that appears in energy balance

equation. For liquids and solids the internal energy can be approximated by

enthalpy .This can be also a good approximation for gases if the pressure change

is small.

1.13 Phase Equilibrium

A large number of chemical processes involve more than one phase. In many

cases the phases are brought in direct contact with each other such as in packed or tray

towers. When the transfer of either mass or energy occurs from a fluid phase to another

phase the interface between fluid phases is usually at equilibrium. In the case of heat

transfer between phase I and Phase II, the equilibrium dictates that the temperature is

the same at the interface. This is not the case for mass transfer. Figure 1.7 shows for

instant the mass transfer of species A from a liquid to a gas phase. The concentration of

Page 25: Modeling Notes Complete

22

the bulk gas phase yAG decreases at the interface. The liquid concentration increases, on

the other hand, from xAL to xAi. At the interface, an equilibrium exists and yAi and xAi are

related by a relation of the form of

yAi = F(xAi) (1.33)

The equilibrium relations are in general nonlinear. However, quite satisfactory results

can be obtained through the use of simpler relations that are derived form assumptions

of ideal behavior of the two phases:

Gas-phase mixtureof A in gas G

Liquid -phase solutionof A in liquid L

xAL

xAi

yAiyAG

NA

Interface

Figure 1.7: Equlibrium at the interface

• For vapor phases at low concentration, Henry's law provides the following

simple equilibrium relation:

pA = HxA (1.34)

Where pA (atm) is the partial pressure of species A in the vapor, xA is the mole

fraction of A in the liquid and H is the Henry's constant (atm/mol fraction)

Dividing by the total Pressure (P) we get another form of Henry's law:

yA = H~ xA (1.35)

Page 26: Modeling Notes Complete

23

The constant ( H~ ) depends on temperature and pressure.

• Raoult's law provides a suitable equilibrium law for ideal vapor-liquid mixtures

S

AAA PxPy = (1.36)

Where Ax is the liquid-phase mole fraction, Ay is the vapor-phase mole

fraction, PAs is the vapor pressure of pure A at the temperature of the system,

and P is the total pressure on gas-phase side.

The dependence of vapor pressure PAs on temperature can be approximated by

the Antoine equation

TCBAP S

A +−=)ln(

(1.37)

where A, B and C are characteristic parameters of the fluid.

• Raoult’s law can be modified to account for non-ideal liquid and vapor behavior

using the activity coefficients γi and φi of component (i) for liquid and vapor

phase respectively. The following equilibrium relation, also known as the

Gamma/Phi formulation, can be used:

S

iiiii PxPy γφ =

(1.38)

Page 27: Modeling Notes Complete

24

The activity coefficients can be determined using correlations found in standard

thermodynamics text books. Equation (1.38) is reduced to Raoult law for ideal

mixture i.e. ( iφ = iγ =1).

• In some cases the transfer occurs in liquid-liquid phases (such as liquid

extraction) or liquid-solid (such as ion exchange). In such cases an equilibrium

relation similar to Henry's law can be defined:

yA = KxA (1.39)

where K is the equilibrium distribution coefficient that depends on pressure,

temperature and concentration.

Phase equilibrium relations are commonly used in the following calculations:

• Bubble point calculations :

For a given molar liquid compositions (xi) and either T or P, the bubble point

calculations consist in finding the molar vapor composition (yi) and either P or

T.

• Dew point calculations :

For a given molar vapor compositions (yi) and either T or P, the dew point

calculations consist in finding the molar liquid compositions (xi) and either P or

T.

• Flash calculation :

For known mixture compositions (zi) and known (T) and (P), the flash

calculations consist in finding the liquid and vapor compositions via

simultaneous solution of component balance and energy balance equations.

1.14 Chemical kinetics

The overall rate R in moles/m3s of a chemical reaction is defined by:

Page 28: Modeling Notes Complete

25

dtdn

VvR i

i

1=

(1.40)

where

ni the number of moles

νi the stochiometric coefficient of component i

V is the volume due to the chemical reaction.

The rate expression R is generally a complex relation of the concentrations (or partial

pressures) of the reactants and products in addition to pressure and temperature.

For a general irreversible reaction

νr1R1 + νr2R2 + …+ νrrRrr νp1P1 + νp2P2 + …+ νppRpp (1.41)

The law of mass action stipulates that the reaction rate is a power law function of

temperature and concentration of reactants, i.e.

rrRRR CCkCR ζζζ= L2

211

(1.42)

The powers ξi are determined experimentally and their values are not necessarily

integers. The temperature dependence comes from the reaction rate constant k given by

Arrhenius law

RTE

oekk−

= (1.43)

where ko is the pre-exponential factor, E the activation energy, T the absolute

temperature and R is the ideal gas constant

1.15 Control Laws

Although the discussion of feedback control systems is beyond the scope of this

book, the existence of control loops in any processing system provides extra relations

Page 29: Modeling Notes Complete

26

for the process. Specifically in many cases some key process variables are required to

be controlled, i.e. maintained within desired range. This control objective can be

achieved by closing the control loop, i.e., relating the controlled variables to inputs

(manipulated variables). This introduces additional independent equations.

1.16 Degrees of Freedom

A key step in the model development and solution is checking its consistency or

solvability, i.e. the existence of exact solution. This is done by checking the degrees of

freedom of the model after the equations have written and before attempting to solve

them .

For a processing system described by a set of Ne independent equations and Nv

variables, the degree of freedom f is

F = Nv − Ne (1.44)

Depending on the value of f three cases can be distinguished:

• f = 0. The system is exactly determined (specified) system .Thus, the set of

balance equation has a finite number of solutions (one solution for linear

systems)

• f < 0. The system is over-determined (over-specified) by f equations. f

equations have to be removed for the system to have a solution.

• f > 0. The system is under-determined (under-specified) by f equations. The set

of equation, hence, has infinite number of solution .

To avoid the situations of over-specified or under-specified systems it is advised to

follow the following steps while checking the consistency of the model.

1. Determine known quantities of the model that can be fixed such as equipment

dimensions, constant physical properties, etc.

Page 30: Modeling Notes Complete

27

2. Determine other variables that can specified by the external world, for example,

variables that are the outcome of an upstream processing units, and/or variables

that can be used as forcing function or manipulated variables.

1.17 Model solution

After the solvability of the model has been checked, the next step is to solve the

model. The purpose of the solution of the model is be able to obtain the variations of

the state variables with the model independent variables (time, spatial positions..). The

solution of the model permits also a parametric investigation of the model, that is a

study of the effects of the changing the value of some parameters. It would be ideal to

be able to solve the model analytically, that is to get closed forms of the state variables

in term of the independent variables. Unfortunately this seldom occurs for chemical

processes . The reason is that the vast majority of chemical processes are nonlinear.

They may be a sets of nonlinear partial differential equations (PDE) as it is the case for

distributed parameter models, or sets of nonlinear ordinary differential equations

(ODE) or nonlinear algebraic equations as it is the case for lumped parameter models.

However most non linear problems can not be solved analytically. In fact the only class

of differential equations for which there is a well-developed framework are linear ODE .

However linearizing the original nonlinear model and solving it is not always

recommended (expect for control purposes) since the behavior of the linearized model

matches the original nonlinear model only around the state chosen for linearization .For

these reasons the solution of process models is usually carried out numerically, most

often through a computer programs.

1.18 Model validation

Model verification (validation) is the last and the most important step of model

building. Reliability of the obtained model depends heavily on faithfully passing this

test. Implementation of the model without validation may lead to erroneous and

Page 31: Modeling Notes Complete

28

misleading results. So, it is essential, as it is saves a lot of effort, time and frustration, to

verify the model against plant operating data, experimental data, or at least published

correlations. If the model failed in the test, then it might be necessary to adjust some of

the model parameters, which is believed to be poorly known, in order to minimize the

mismatch between the model and the true plant. In worst cases, a modeler may need to

reconsider some of the simplifying assumptions used or the neglected modeling parts.

However, it should be kept in mind that the model is no more than approximation of

the real world, thus some degree of mismatch will remain and could be overlooked.

Page 32: Modeling Notes Complete

29

Chapter 2: Examples of Mathematical Models for Chemical Processes

In this chapter we develop mathematical models for a number of elementary

chemical processes that are commonly encountered in practice. We will apply the

methodology discussed in the previous chapter to guide the reader through various

examples. The goal is to give the reader a methodology to tackle more complicated

processes that are not covered in this chapter and that can be found in books listed in

the reference. The organization of this chapter includes examples of systems that can be

described by ordinary differential equations (ODE), i.e. lumped parameter systems

followed by examples of distributed parameters systems, i..e those described by partial

differential equations (PDE). The examples cover both homogeneous and heterogeneous

systems. Ordinary differential equations (ODE) are easier to solve and are reduced to

simple algebraic equations at steady state. The solution of partial differential equations

(PDE) on the other hand is a more difficult task. But we will be interested in the cases

were PDE's are reduced to ODE's. This is naturally the case where under appropriate

assumptions, the PDE's is a one-dimensional equation at steady state conditions. It is

worth to recall, as noted in the previous chapters, that the distinction between lumped

and distributed parameter models depends sometimes on the assumptions put forward

by the modeler. Systems that are normally distributed parameter can be modeled under

appropriate assumptions as lumped parameter systems. This chapter includes some

examples of this situation.

Page 33: Modeling Notes Complete

30

2.1 Examples of Lumped Parameter Systems

2.1.1 Liquid Storage Tank

Consider the perfectly mixed storage tank shown in figure 2.1. Liquid stream

with volumetric rate Ff (m3/s) and density ρf flow into the tank. The outlet stream has

volumetric rate Fo and density ρο. Our objective is to develop a model for the variations

of the tank holdup, i.e. volume of the tank. The system is therefore the liquid in the

tank. We will assume that it is perfectly mixed and that the density of the effluent is the

same as that of tank content. We will also assume that the tank is isothermal, i.e. no

variations in the temperature. To model the tank we need only to write a mass balance

equation.

Figure 2-1 Liquid Storage Tank

Since the system is perfectly mixed, the system properties do not vary with position

inside the tank. The only variations are with time. The mass balance equation can be

written then on the whole system and not only on a differential element of it. This leads

to therefore to a macroscopic model.

We apply the general balance equation (Eq. 1.2), to the total mass m = ρV. This yields:

Mass flow in:

ρf Ff (2.1)

Page 34: Modeling Notes Complete

31

Mass flow out:

ρo Fo (2.2)

Accumulation:

dtVd

dtdm )(ρ

=

(2.3)

The generation term is zero since the mass is conserved. The balance equation yields:

dtVdFF ooff

)(ρ+ρ=ρ

(2.4)

For consistency we can check that all the terms in the equation have the SI unit of kg/s.

The resulting model (Eq. 2.4) is an ordinary differential equation (ODE) of first order

where time (t) is the only independent variable. This is therefore a lumped parameter

model. To solve it we need one initial condition that gives the value of the volume at

initial time ti, i.e.

V(ti) = Vi (2.5)

Under isothermal conditions we can further assume that the density of the liquid is

constant i.e. ρf = ρo=ρ. In this case Eq. 2.4 is reduced to:

of FFdtdV

−=

(2.6)

The volume V is related to the height of the tank L and to the cross sectional area A by:

V = AL (2.7)

Page 35: Modeling Notes Complete

32

Since (A) is constant then we obtain the equation in terms of the state variable L:

of FFdtdLA −=

(2.8)

with initial condition:

L(ti) = Li (2.9)

Degree of freedom analysis

For the system described by Eq. 2.8 we have the following information:

• Parameter of constant values: A

• Variables which values can be externally fixed (Forced variable): Ff

• Remaining variables: L and Fo

• Number of equations: 1 (Eq. 2.8)

Therefore the degree of freedom is:

Number of remaining variables – Number of equations = 2 – 1 = 1

For the system to be exactly specified we need therefore one more equations. This extra

relation is obtained from practical engineering considerations. If the system is operated

without control (at open loop) then the outlet flow rate Fo is a function of the liquid

level L. Generally a relation of the form:

LFo α= (2.10)

could be used, where α is the discharge coefficient.

If on the other hand the liquid level is under control, then its value is kept constant at

certain desired value Ls. If Fo is used to control the height then a control law relates Fo

to L and Ls:

Page 36: Modeling Notes Complete

33

Fo = Fo(L,Ls) (2.11)

For instant, if a proportional controller Kc is used then the control law is given by:

Fo = Kc(L − Ls) + Fob (2.12)

Where Fob the bias, i.e. the constant value of Fo when the level is at the desired value

i.e., L = Ls.

Note that at steady state, the accumulation term is zero (height does not change with

time), i.e., dL/dt = 0. The model of the tank is reduced to the simple algebraic equation:

F0 = Ff (2.13)

2.1.2 Stirred Tank Heater

We consider the liquid tank of the last example but at non-isothermal conditions.

The liquid enters the tank with a flow rate Ff (m3/s), density ρf (kg/m3) and temperature

Tf (K). It is heated with an external heat supply of temperature Tst (K), assumed

constant. The effluent stream is of flow rate Fo (m3/s), density ρo (kg/m3) and

temperature T(K) (Fig. 2.2). Our objective is to model both the variation of liquid level

and its temperature. As in the previous example we carry out a macroscopic model over

the whole system. Assuming that the variations of temperature are not as large as to

affect the density then the mass balance of Eq. 2.8 remains valid.

To describe the variations of the temperature we need to write an energy balance

equation. In the following we develop the energy balance for any macroscopic system

(Fig. 2.3) and then we apply it to our example of stirred tank heater.

The energy E(J) of any system of (Fig. 2.3) is the sum of its internal U(J), kinetic K(J)

and potential energy φ(J):

E = U + K + φ (2.14)

Consequently, the flow of energy into the system is:

Page 37: Modeling Notes Complete

34

ρf Ff ( fff K U φ++~ ~~ ) (2.15)

where the ( •~ ) denotes the specific energy (J/kg).

Figure 2-2 Stirred Tank Heater

Figure 2-3 General Macroscopic System

The flow of energy out of the system is:

ρo Fo( ooo K U φ++~ ~~ ) (2.16)

The rate of accumulation of energy is:

Page 38: Modeling Notes Complete

35

( )dt

K UVd )~ ~~( φ++ρ (2.17)

As for the rate of generation of energy, it was mentioned in Section 1.8.4, that the

energy exchanged between the system and the surroundings may include heat of

reaction Qr (J/s), heat exchanged with surroundings Qe (J/s) and the rate of work done

against pressure forces (flow work) Wpv (J/s), in addition to any other work Wo.

The flow of work Wpv done by the system is given by:

ffoopv PFPFW −= (2.18)

where Po and Pf are the inlet and outlet pressure, respectively.

In this case, the rate of energy generation is:

)( ffooore PFPFWQQ −+−+ (2.19)

Substituting all these terms in the general balance equation (Eq. 1.7) yields:

( ) ( ) ( ))(

~ ~~ ~ ~~ )~ ~~(

ffooore

ooooofffff

PFPFWQQ

K UF K UFdt

K UVd

−+−++

φ++ρ−φ++ρ=φ++ρ

(2.20)

We can check that all terms of this equation have the SI unit of (J/s). Equation (2.20)

can be also written as:

( ) ( ) ( )

f

fff

o

oooore

ooooofffff

PFPFWQQ

K UF K UFdt

K UVd

ρρ+

ρρ−−++

φ++ρ−φ++ρ=φ++ρ ~ ~~ ~ ~~ )~ ~~(

(2.21)

Page 39: Modeling Notes Complete

36

The term ρ= /1~V is the specific volume (m3/kg). Thus Eq. 2.21 can be written as:

( ) ( ) ( )ore

ooooooofffffff

WQQ

KV P UF KV P UFdt

K UVd

−++

+++−+++=++ φρφρφρ ~ ~~~ ~ ~~~ )~ ~~(

(2.22)

The term V P U ~~ + that appears in the equation is the specific enthalpy h~ . Therefore,

the general energy balance equation for a macroscopic system can be written as:

( ) ( ) ( ) oreooooofffff WQQ KhF K hFdt

K UVd−++φ++ρ−φ++ρ=

φ++ρ ~ ~~ ~ ~~ )~ ~~( (2.23)

We return now to the liquid stirred tank heater. A number of simplifying assumptions

can be introduced:

• We can neglect kinetic energy unless the flow velocities are high.

• We can neglect the potential energy unless the flow difference between the inlet

and outlet elevation is large.

• All the work other than flow work is neglected, i.e. Wo = 0.

• There is no reaction involved, i.e. Qr = 0.

The energy balance (Eq. 2.23) is reduced to:

( )eooofff QhFhF

dtUVd

+ρ−ρ=ρ ~ ~

~

(2.24)

Here Qe is the heat (J/s) supplied by the external source. Furthermore, as mentioned in

Section 1.12, the internal energy U~ for liquids can be approximated by enthalpy, h~ .

The enthalpy is generally a function of temperature, pressure and composition.

However, it can be safely estimated from heat capacity relations as follows:

Page 40: Modeling Notes Complete

37

)(~~refTTpCh −= (2.25)

where pC~ is the average heat capacity.

Furthermore since the tank is well mixed the effluent temperature To is equal to process

temperature T. The energy balance equation can be written, assuming constant density

ρf = ρo = ρ , as follows:

( )erefporeffpf

refp QTTCFTTCF

dtTTVd

C +−ρ−−ρ=−

ρ )(~ )(~ )(~

(2.26)

Taking Tref = 0 for simplicity and since V = AL result in:

( )epofpfp QTCFTCF

dtLTdAC +ρ−ρ=ρ

~ ~ ~ (2.27)

or equivalently:

( )p

eoff C

QTFTFdtLTdA ~

ρ+−=

(2.28)

Since

( ) ( ) ( )dtTdAL

dtLdAT

dtLTdA +=

(2.29)

and using the mass balance (Eq. 2.8) we get:

p

eoffof C

QTFTFFFT

dtdTAL ~ )(

ρ+−=−+

(2.30)

or equivalently:

Page 41: Modeling Notes Complete

38

p

eff C

QTTFdtdTAL ~)(

ρ+−=

(2.31)

The stirred tank heater is modeled, then by the following coupled ODE's:

of FFdtdLA −=

(2.32)

p

eff C

QTTFdtdTAL ~)(

ρ+−=

(2.33)

This system of ODE's can be solved if it is exactly specified and if conditions at initial

time are known,

L(ti) = Li and T(ti) = Ti (2.34)

Degree of freedoms analysis

For this system we can make the following simple analysis:

• Parameter of constant values: A, ρ and Cp

• (Forced variable): Ff and Tf

• Remaining variables: L, Fo, T, Qe

• Number of equations: 2 (Eq. 2.32 and Eq. 2.33)

The degree of freedom is therefore, 4 − 2 = 2. We still need two relations for our

problem to be exactly specified. Similarly to the previous example, if the system is

operated without control then Fo is related to L through (Eq. 2.10). One additional

relation is obtained from the heat transfer relation that specifies the amount of heat

supplied:

Qe = UAH (Tst−T ) (2.35)

Page 42: Modeling Notes Complete

39

U and AH are heat transfer coefficient and heat transfer area. The source temperature Tst

was assumed to be known. If on the other hand both the height and temperature are

under control, i.e. kept constant at desired values of Ls and Ts then there are two control

laws that relate respectively Fo to L and Ls and Qe to T and Ts:

Fo = Fo(L, Ls), and Qe = Qe (T, Ts) (2.36)

2.1.3 Isothermal CSTR

We revisit the perfectly mixed tank of the first example but where a liquid phase

chemical reactions taking place:

BA k⎯→⎯ (2.37)

The reaction is assumed to be irreversible and of first order. As shown in figure 2.4, the

feed enters the reactor with volumetric rate Ff (m3/s), density ρf (kg/m3) and

concentration CAf (mole/m3). The output comes out of the reactor at volumetric rate Fo,

density ρ0 and concentration CAo (mole/m3) and CBo (mole/m3). We assume isothermal

conditions.

Our objective is to develop a model for the variation of the volume of the reactor

and the concentration of species A and B. The assumptions of example 2.1.1 still hold

and the total mass balance equation (Eq. 2.6) is therefore unchanged

Figure 2.4 Isothermal CSTR

Page 43: Modeling Notes Complete

40

The component balance on species A is obtained by the application of (Eq. 1.3)

to the number of moles (nA = CAV ). Since the system is well mixed the effluent

concentration CAo and CBo are equal to the process concentration CA and CB.

Flow of moles of A in:

Ff CAf (2.38)

Flow of moles of A out:

Fo CAo (2.39)

Rate of accumulation:

dtVCd

dtdn A)(

=

(2.40)

Rate of generation: -rV

where r (moles/m3s) is the rate of reaction.

Substituting these terms in the general equation (Eq. 1.3) yields:

rV CF CFdtVCd

AoAffA −−=

)( (2.41)

We can check that all terms in the equation have the unit (mole/s).

We could write a similar component balance on species B but it is not needed

since it will not represent an independent equation. In fact, as a general rule, a system of

n species is exactly specified by n independent equations. We can write either the total

mass balance along with (n −1) component balance equations, or we can write n

component balance equations.

Page 44: Modeling Notes Complete

41

Using the differential principles, equation (2.41) can be written as follows:

rV CF CFdtVdC

dtCdV

dtVCd

AoAffAAA −−=+=

)()()( (2.42)

Substituting Equation (2.6) into (2.42) and with some algebraic manipulations we

obtain:

rVC CFdtCdV AAff

A −−= )()( (2.43)

In order to fully define the model, we need to define the reaction rate which is for a

first-order irreversible reaction:

r = k CA (2.44)

Equations 2.6 and 2.43 define the dynamic behavior of the reactor. They can be solved

if the system is exactly specified and if the initial conditions are given:

V(ti) = Vi and CA(ti) = CAi (2.45)

Degrees of freedom analysis

• Parameter of constant values: A

• (Forced variable): Ff and CAf

• Remaining variables: V, Fo, and CA

• Number of equations: 2 (Eq. 2.6 and Eq. 2.43)

The degree of freedom is therefore 3 − 2 =1. The extra relation is obtained by the

relation between the effluent flow Fo and the level in open loop operation (Eq. 2.10) or

in closed loop operation (Eq. 2.11).

The steady state behavior can be simply obtained by setting the accumulation terms to

zero. Equation 2.6 and 2.43 become:

Page 45: Modeling Notes Complete

42

F0 = Ff (2.46)

rVC CF AAff =− )( (2.47)

More complex situations can also be modeled in the same fashion. Consider the

catalytic hydrogenation of ethylene:

A + B P (2.48)

where A represents hydrogen, B represents ethylene and P is the product (ethane). The

reaction takes place in the CSTR shown in figure 2.5. Two streams are feeding the

reactor. One concentrated feed with flow rate F1 (m3/s) and concentration CB1 (mole/m3)

and another dilute stream with flow rate F2 (m3/s) and concentration CB2 (mole/m3). The

effluent has flow rate Fo (m3/s) and concentration CB (mole/m3). The reactant A is

assumed to be in excess.

VFo, CB

F1, CB1 F2, CB2

Figure 2-5 Reaction in a CSTR

The reaction rate is assumed to be:

)./()1(

32

2

1 smmoleCk

CkrB

B

+=

(2.49)

Page 46: Modeling Notes Complete

43

where k1 is the reaction rate constant and k2 is the adsorption equilibrium constant.

Assuming the operation to be isothermal and the density is constant, and following the

same procedure of the previous example we get the following model:

Total mass balance:

oFFFdtdLA −+= 21

(2.50)

Component B balance:

rVC CFC CFdtCdV BBBB

A −−+−= )()()(2211

(2.51)

Degrees of freedom analysis

• Parameter of constant values: A, k1 and k1

• (Forced variable): F1 F2 CB1 and CB2

• Remaining variables: V, Fo, and CB

• Number of equations: 2 (Eq. 2.50 and Eq. 2.51)

The degree of freedom is therefore 3 − 2 =1. The extra relation is between the effluent

flow Fo and the level L as in the previous example.

2.1.4 Gas-Phase Pressurized CSTR

So far we have considered only liquid-phase reaction where density can be taken

constant. To illustrate the effect of gas-phase chemical reaction on mass balance

equation, we consider the following elementary reversible reaction:

BA 2↔ (2.52)

taking place in perfectly mixed vessel sketched in figure 2.6. The influent to the vessel

has volumetric rate Ff (m3/s), density ρf (kg/m3), and mole fraction yf. Product comes out

of the reactor with volumetric rate Fo, density ρo, and mole fraction yo. The temperature

Page 47: Modeling Notes Complete

44

and volume inside the vessel are constant. The reactor effluent passes through control

valve which regulate the gas pressure at constant pressure Pg.

P, T, V ,y Fo, yoFf, f , yf

Pg

Figure 2-6 Gas Pressurized Reactor

Writing the macroscopic total mass balance around the vessel gives:

ooff FFdt

Vd ρρρ−=

)(

(2.53)

Since V is constant we have:

ooff FFdtdV ρρρ

−=

(2.54)

Writing the component balance, for fixed V, results in:

VrVrCFCFdt

dCV AoAffA

210+−−=

(2.55)

The reaction rates for the reversible reaction are assumed to be:

r1 = k1 CA (2.56)

2

22 BCkr = (2.57)

Equations (2.54) and (2.55) define the variations of density and molar concentration.

One can also rewrite the equation to define the behavior of the pressure (P) and mole

Page 48: Modeling Notes Complete

45

fraction (y). The concentration can be expressed in term of the density through ideal gas

law:

CA = yP/RT (2.58)

CB = (1 − y)P/RT (2.59)

Similarly, the density can be related to the pressure using ideal gas law:

ρ = MP/RT = [MAy + MB (1 − y)]P/RT (2.60)

Where MA and MB are the molecular weight of A and B respectively. Therefore one can

substitute equations (2.58) to (2.60) into equations (2.54 & 2.55) in order to explicitly

write the latter two equations in terms of y and P. Or, alternatively, one can solve all

equations simultaneously.

Degrees of freedom analysis:

• Parameters: V, k1, k2, R, T, MA and MB

• Forcing function: Ff, CAf, yf

• Variables: CA, CB, y, P, ρ, F

• Number of equations: 5 (Eqs. 2.54, 2.55, 2.58, 2.59, 2.60)

The degree of freedom is therefore 6 − 5 =1. The extra relation relates the outlet flow to

the pressure as follows:

ρg

vo

PPCF

−=

(2.61)

where Cv is the valve-sizing coefficient. Recall also that Pg is assumed to be constant.

2.1.5 Non-Isothermal CSTR

We reconsider the previous CSTR example (Sec 2.1.3), but for non-isothermal

conditions. The reaction A B is exothermic and the heat generated in the reactor is

Page 49: Modeling Notes Complete

46

removed via a cooling system as shown in figure 2.7. The effluent temperature is

different from the inlet temperature due to heat generation by the exothermic reaction.

Figure 2-7 Non-isothermal CSTR

Assuming constant density, the macroscopic total mass balance (Eq. 2.6) and mass

component balance (Eq. 2.43) remain the same as before. However, one more ODE will

be produced from the applying the conservation law (equation 2.23) for total energy

balance. The dependence of the rate constant on the temperature:

k = koe-E/RT (2.62)

should be emphasized.

The general energy balance (Eq. 2.23) for macroscopic systems applied to the CSTR

yields, assuming constant density and average heat capacity:

( )errefporeffpf

refp QQTTCFTTCF

dtTTVd

C −+−ρ−−ρ=−

ρ )(~ )(~ )(~

(2.63)

where Qr (J/s) is the heat generated by the reaction, and Qe (J/s) the rate of heat

removed by the cooling system. Assuming Tref = 0 for simplicity and using the

differentiation principles, equation 2.63 can be written as follows:

erpofpfpp QQTCFTCFdtdVTC

dtdTVC −+ρ−ρ=ρ+ρ

~ ~ ~~ (2.64)

Page 50: Modeling Notes Complete

47

Substituting Equation 2.6 into the last equation and rearranging yields:

erfpfp QQTTCFdtdTVC −+−ρ=ρ )(~ ~

(2.65)

The rate of heat exchanged Qr due to reaction is given by:

Qr = −(∆Hr)Vr (2.66)

where ∆Hr (J/mole) is the heat of reaction (has negative value for exothermic reaction

and positive value for endothermic reaction). The non-isothermal CSTR is therefore

modeled by three ODE's:

of FFdtdV

−= (2.67)

rVC CFdtCdV AAff

A −−= )()( (2.68)

erfpfp QVrHTTCFdtdTVC −−+−ρ=ρ )()(~ ~

∆ (2.69)

where the rate (r) is given by:

r = koe-E/RTCA (2.70)

The system can be solved if the system is exactly specified and if the initial conditions

are given:

V(ti) = Vi T(ti) = Ti and CA(ti) = CAi (2.71)

Degrees of freedom analysis

• Parameter of constant values: ρ, E, R, Cp, ∆Hr and ko

• (Forced variable): Ff , CAf and Tf

Page 51: Modeling Notes Complete

48

• Remaining variables: V, Fo, T, CA and Qe

• Number of equations: 3 (Eq. 2.67. 2.68 and 2.69)

The degree of freedom is 5−3 = 2. Following the analysis of example 2.1.3, the two

extra relations are between the effluent stream (Fo) and the volume (V) on one hand and

between the rate of heat exchanged (Qe) and temperature (T) on the other hand, in either

open loop or closed loop operations.

A more elaborate model of the CSTR would include the dynamic of the cooling

jacket (Fig. 2.8). Assuming the jacket to be perfectly mixed with constant volume Vj,

density ρj and constant average thermal capacity Cpj, the dynamic of the cooling jacket

temperature can be modeled by simply applying the macroscopic energy balance on the

whole jacket:

ejjfpjjj

jpj QTTCFdt

dTVC

jj+−ρ=ρ )(~ ~

(2.72)

Since Vj, ρj, Cpj and Tjf are constant or known, the addition of this equation introduces

only one variable (Tj). The system is still exactly specified.

V

Ff , CAf , Tf

Fo , CA , T

Fj , Tjf

Fj , Tj

Figure 2-8 Jacketed Non-isothermal CSTR

2.1.6 Mixing Process

Consider the tank of figure 2.9 where two solutions 1 and 2 containing materials

A and B are being mixed. Stream 1 has flow rate F1 (m3/s), density ρ1 (kg/m3), T1 (K),

Page 52: Modeling Notes Complete

49

concentration CA1 (mole/m3) and CB1 (mole/m3) of material A and B. Similarly stream 2

has flow rate F2 (m3/s), density ρ2 (kg/m3), T2 (K), concentration CA2 (mole/m3) and CB2

(mole/m3) of material A and B. The effluent stream has flow rate Fo (m3/s), density ρo

(kg/m3), To (K), concentration CAo (mole/m3) and CBo (mole/m3) of material A and B. We

assume that the mixing releases heat of rate Q (J/s) which is absorbed by a cooling fluid

flowing in a jacket or a coil.

Our objective is to develop a model for the mixing process. We will assume that

the tank is well mixed. In this case all the effluent properties are equal to the process

properties. We also assume for simplicity that the densities and heat capacities of the

streams are constant and equal:

ρ = ρ1 = ρ2 = ρo (2.73)

Cp = Cp1 = Cp2 = Cp3 (2.74)

Fo, To, CAo, CBo

F1, T1, CA1, CB1 F2, T2, CA2, CB2

Q

Figure 2-9 Mixing Process

Total mass balance

The mass balance equation yields

( ) ooFFFdt

Vd ρρρρ−+= 2211

)(

(2.75)

Page 53: Modeling Notes Complete

50

Since the densities are equals we have:

( ) oFFFdtdV

−+= 21

(2.76)

Component balance

The component balance for species A for instant yields:

( ) AoooAAAoo CFCFCF

dtVCd ρρρρ

−+= 222111)(

(2.77)

Expanding Eq. 2.77 yields:

( ) AoooAAAoAo

o CFCFCFdtdVC

dtdCV ρρρρ −+=⎟

⎠⎞

⎜⎝⎛ + 222111

(2.78)

Substituting Eq. 2.76 into 2.78 yields after some manipulation:

( ) ( )AoAAoAAo CCFCCF

dtdCV −+−= 2211

(2.79)

A similar equation holds for component B,

( ) BoooBBBoo CFCFCF

dtVCd ρρρρ

−+= 222111)(

(2.80)

Energy balance

The general energy balance equation (Eq. 2.23) yields, assuming negligible kinetic,

potential energy and since no reaction or shaft work occurs:

Page 54: Modeling Notes Complete

51

( )QhFhFhF

dthVd

mixooomixmixmixoo −−+= ,,222,111

, ~~ ~ ~

ρρρρ

(2.81)

where mixih ,~ (J/kg) is the specific enthalpy of mixture i.

The specific enthalpy mixh~ of n components can be written as:

)()(~),(~, refmixrefmixmix TTCpPThPTh −+= (2.82)

where

),,(ˆ),,(ˆ),(~1

PTChCPTChCPTh iskrefi

n

iiirefmixmix ∆+= ∑

=

ρ

(2.83)

with ih (J/mole) being the molar enthalpy of component i and sh∆ is the heat of

solution per mole of a key component k. Assume constant process pressure (P) then

we can write the enthalpy of each stream as follows, taking component (A) as the key

component,

)(ˆˆˆ)(~1,11,11111,11 refmixsABBAAmix TTCphChChCTh −+∆++= ρρ (2.84)

)(ˆˆˆ)(~2,22,22222,22 refmixsABBAAmix TTCphChChCTh −+∆++= ρρ (2.85)

)(ˆˆˆ)(~,,, refmixoosoAoBBoAAomixoo TTCphChChCTh −+∆++= ρρ (2.86)

Substituting Eq. 2.86 into the left hand side of Eq. 2.81 and expanding yields:

( )dt

hVd mixoo ,~ρ

=dtdVCpT

dtdTCpV

dtVCdh

dtVCdh

dtVCdh Ao

soBo

BAo

A ρρ ++∆++)(ˆ)(ˆ)(ˆ

,

(2.87)

Page 55: Modeling Notes Complete

52

The right hand side of equation (2.81) is equal after some manipulation to:

QTFTFTFCphFChFChFC

FCFCFChFCFCFCh

oo

sooAosAsA

oBoBBBoAoAAA

−−++

∆−∆+∆+

−++−+

)()ˆˆˆ(

)(ˆ)(ˆ

2211

,,222,111

22112211

ρ

(2.88)

Substituting into the right hand side of Eq. 2.87, the total mass balance equation (Eq.

2.76) and the component balance equation for species (A) (Eq. 2.77) and that of B (Eq.

2.80), and equating Eq. 2.87 to Eq. 2.88 yields, after some manipulations:

( ) QTTFTTFCp

hhFChhFCdtdTVCp sosAsosA

−−+−+

∆−∆+∆−∆=

)()(

)ˆˆ()ˆˆ(

2211

,,222,,111

ρ

ρ

(2.89)

The mixer is then described by three ODE's Equations (2.76, 2.79, 2.89). To these

relations we should add the relations that give the heats of mixing:

),,(ˆ1111,1 TCCfh BAs =∆ (2.90)

),,(ˆ2222,2 TCCfh BAs =∆ (2.91)

),,(ˆ3, oBoAoso TCCfh =∆ (2.92)

Degrees of freedom analysis

• Parameter of constant values: ρ, A, Tref, Cp, Ah and Bh

• (Forced variable): F1 , F2, CA1, CA2, T1 and T2

• Remaining variables: V, Fo, To, CAo, Q, sh ,1∆ , sh ,2ˆ∆ , soh ,

ˆ∆

• Number of equations: 6 (Eq. 2.76, 2.79, 2.89 and 2.90-2.92)

Page 56: Modeling Notes Complete

53

The degree of freedoms is therefore 8 – 6 = 2. The two needed relations are the relation

between effluent stream Fo and height L and the relation between the heat Q and

temperature To in either open loop or closed-loop operations.

2.1.7 Heat Exchanger

Consider the shell and tube heat exchanger shown in figure 2.10. Liquid A of

density ρA is flowing through the inner tube and is being heated from temperature TA1

to TA2 by liquid B of density ρB flowing counter-currently around the tube. Liquid B

sees its temperature decreasing from TB1 to TB2. Clearly the temperature of both liquids

varies not only with time but also along the tubes (i.e. axial direction) and possibly with

the radial direction too. Tubular heat exchangers are therefore typical examples of

distributed parameters systems. A rigorous model would require writing a microscopic

balance around a differential element of the system. This would lead to a set of partial

differential equations. However, in many practical situations we would like to model the

tubular heat exchanger using simple ordinary differential equations. This can be

possible if we think about the heat exchanger within the unit as being an exchanger

between two perfect mixed tanks. Each one of them contains a liquid.

Liquid, ATA1 TA2

Liquid, BTB1

Liquid, BTB2

Tw

Figure 2-10 Heat Exchanger

For the time being we neglect the thermal capacity of the metal wall separating the two

liquids. This means that the dynamics of the metal wall are not included in the model.

We will also assume constant densities and constant average heat capacities.

Page 57: Modeling Notes Complete

54

One way to model the heat exchanger is to take as state variable the exit temperatures

TA2 and TB2 of each liquid. A better way would be to take as state variable not the exit

temperature but the average temperature between the inlet and outlet:

221 AA

ATTT +

=

(2.93)

221 BB

BTTT +

= (2.94)

For liquid A, a macroscopic energy balance yields:

QTTCFdt

dTVC AApAAA

ApA AA+−= )( 21ρρ

(2.95)

where Q (J/s) is the rate of heat gained by liquid A. Similarly for liquid B:

QTTCFdt

dTVC BBpBBB

BpB BB−−= )( 21ρρ

(2.96)

The amount of heat Q exchanged is:

Q = UAH (TB – TA) (2.97)

Or using the log mean temperature difference:

Q = UAH ∆Tlm (2.98)

where

)()(ln

)()(

21

12

2112

BA

BA

BABAlm

TTTT

TTTTT

−−

−−−=∆

(2.99)

Page 58: Modeling Notes Complete

55

with U (J/m2s) and AH (m2) being respectively the overall heat transfer coefficient and

heat transfer area. The heat exchanger is therefore describe by the two simple ODE's

(Eq. 2.95) and (Eq. 2.96) and the algebraic equation (Eq. 2.97).

Degrees of freedom analysis

• Parameter of constant values: ρΑ, CpA, VA, ρΒ, CpB, VB, U, AH

• (Forced variable): TA1, TB1, FA, FB

• Remaining variables: TA2, TB2, Q

• Number of equations: 3 (Eq. 2.95, 2.96, 2.97)

The degree of freedom is 5 − 3 = 2. The two extra relations are obtained by noting that

the flows FA and FB are generally regulated through valves to avoid fluctuations in their

values.

So far we have neglected the thermal capacity of the metal wall separating the

two liquids. A more elaborated model would include the energy balance on the metal

wall as well. We assume that the metal wall is of volume Vw, density ρw and constant

heat capacity Cpw. We also assume that the wall is at constant temperature Tw, not a bad

assumption if the metal is assumed to have large conductivity and if the metal is not

very thick. The heat transfer depends on the heat transfer coefficient ho,t on the outside

and on the heat transfer coefficient hi,t on the inside. Writing the energy balance for

liquid B yields:

)()( ,,21 WBtotoBBpBBB

BpB TTAhTTCFdt

dTVCBB

−−−= ρρ (2.100)

where Ao,t is the outside heat transfer area. The energy balance for the metal yields:

)()( ,,,, AwtitiwBtotow

wpw TTAhTTAhdt

dTVCw

−−−=ρ (2.101)

where Ai,t is the inside heat transfer area. . The energy balance for liquid A yields:

Page 59: Modeling Notes Complete

56

)()( ,,21 AwtitiAApAAA

ApA TTAhTTCFdt

dTVCAA

−+−= ρρ (2.102)

Note that the introduction of equation (Eq. 2.101) does not change the degree of

freedom of the system.

2.1.8 Heat Exchanger with Steam

A common case in heat exchange is when a liquid L is heated with steam (Figure 2.11).

If the pressure of the steam changes then we need to write both mass and energy balance

equations on the steam side.

Liquid, LTL1 TL2

SteamTs(t)

condensate, Ts

Tw

Figure 2-11 Heat Exchanger with Heating Steam

The energy balance on the tube side gives:

sLLpLLL

LpL QTTCFdt

dTVCLL

+−ρ=ρ )( 21 (2.103)

where

221 LL

LTTT +

= (2.104)

Qs = UAs (Ts – TL) (2.105)

Page 60: Modeling Notes Complete

57

The steam saturated temperature Ts is also related to the pressure Ps:

Ts = Ts (P) (2.106)

Assuming ideal gas law, then the mass flow of steam is:

s

ssss RT

VPMm = (2.107)

where Ms is the molecular weight and R is the ideal gas constant. The mass balance for

the steam yields:

ccsss

ss FFdtdP

RTVM

ρ−ρ= (2.108)

where Fc and ρc are the condensate flow rate and density. The heat losses at the steam

side are related to the flow of the condensate by:

Qs = Fc λs (2.109)

Where λs is the latent heat.

Degrees of freedom analysis

• Parameter of constant values: ρL, CpL, Ms, As, U , Ms, R

• (Forced variable): TL1

• Remaining variables: TL2, FL, Ts, Fs, Ps, Qs, Fc

• Number of equations: 5 (Eq. 2.103, 2.105, 2.106, 2.108, 2.109)

The degrees of freedom is therefore 7 – 5 = 2. The extra relations are given by the

relation between the steam flow rate Fs with the pressure Ps either in open-loop or

closed-loop operations. The liquid flow rate F1 is usually regulated by a valve.

Page 61: Modeling Notes Complete

58

2.1.9 Single Stage Heterogeneous Systems: Multi-component flash drum

The previous treated examples have discussed processes that occur in one single

phase. There are several chemical unit operations that are characterized with more than

one phase. These processes are known as heterogeneous systems. In the following we

cover some examples of these processes. Under suitable simplifying assumptions, each

phase can be modeled individually by a macroscopic balance.

A multi-component liquid-vapor separator is shown in figure 2.12. The feed

consists of Nc components with the molar fraction zi (i=1,2… Nc). The feed at high

temperature and pressure passes through a throttling valve where its pressure is reduced

substantially. As a result, part of the liquid feed vaporizes. The two phases are assumed

to be in phase equilibrium. xi and yi represent the mole fraction of component i in the

liquid and vapor phase respectively. The formed vapor is drawn off the top of the

vessel while the liquid comes off the bottom of the tank. Taking the whole tank as our

system of interest, a model of the system would consist in writing separate balances for

vapor and liquid phase. However since the vapor volume is generally small we could

neglect the dynamics of the vapor phase and concentrate only on the liquid phase.

Fvyi

P, T, Vv

VL LFLxi

FoziToPo

Figure 2-12 Multicomponent Flash Drum

For liquid phase:

Total mass balance:

Page 62: Modeling Notes Complete

59

vvLLffLL FFF

dtVd ρρρρ

−−=)(

(2.110)

Component balance:

ivviLLiffiLL yFxFzF

dtxVd

ρρρρ

−−=)(

(i=1,2,….,Nc-1)

(2.111)

Energy balance:

HFhFhFdt

hVdvvLLfff

LL ~~~)~( ρρρρ−−=

(2.112)

where h~ and H~ are the specific enthalpies of liquid and vapor phase respectively.

In addition to the balance equations, the following supporting thermodynamic

relations can be written:

• Liquid-vapor Equilibrium:

Raoult's law can be assumed for the phase equilibrium

PPx

ys

iii = (i=1,2,….,Nc)

(2.113)

Together with the consistency relationships:

11

=∑=

Nc

iiy

(2.114)

11

=∑=

Nc

iix

(2.115)

• Physical Properties:

The densities and enthalpies are related to the mole fractions, temperature and pressure

through the following relations:

Page 63: Modeling Notes Complete

60

ρL = f(xi,T,P) (2.116)

ρv = f(yi,T,P) ≈ MvaveP/R T (2.117)

Mvave = ∑

=

Nc

iiiMy

1

(2.118)

h = f(xi,T) ≈ )(1

ref

Nc

iii TTCpx −∑

=

(2.119)

H = f(yi,T) ≈ mref

Nc

iii TTCpy λ+−∑

=

)(1

(2.120)

λm = ∑=

Nc

iiiy

1

λ (2.121)

Degrees of freedom analysis:

• Forcing variables: Ff, Tf, Pf , zi (i=1,2..Nc),

• Remaining variables:2Nc+5: VL, FL, FV, P, T, xi (i=1,2..Nc), yi(i=1,2,…Nc)

• Number of equations: 2Nc+3: (Eq. 2.110, 2.111, 2.112, 2.113, 2.114, 2.115)

Note that physical properties are not included in the degrees of freedom since they are

specified through given relations. The degrees of freedom is therefore (2Nc+5)-

(2Nc+3)=2. Generally the liquid holdup (VL) is controlled by the liquid outlet flow rate

(FL) while the pressure is controlled by FV. In this case, the problem becomes well

defined for a solution.

2-1-10 Two-phase Reactor

Consider the two phase reactor shown in figure 2.13. Gaseous A and liquid B

enters the reactor at molar flow rates FA and FB respectively. Reactant A diffuses into

the liquid phase with molar flux (NA) where it reacts with B producing C. The latter

diffuses into the vapor phase with molar flux (NC). Reactant B is nonvolatile. The

product C is withdrawn with the vapor leaving the reactor. The objective is to write the

mathematical equations that describe the dynamic behavior of the process. We consider

all flows to be in molar rates.

Page 64: Modeling Notes Complete

61

Figure 2-4 Two Phase Reactor

Assumptions:

• The individual phases are well mixed and they are in physical equilibrium at

pressure P and temperature T.

• The physical properties such as molar heat capacity Cp, density ρ , and latent heat

of vaporization λ are constant and equal for all the species.

• The reaction mechanism is: A+B C and its rate has the form: Rc = k CA CB VL

• The two phases are in equilibrium and follows the Raoult’s law.

• Total enthalpy for the system is given as: H = NL HL + Nv Hv where HL and Hv are

molar enthalpies in the liquid and vapor phases respectively, and NL and Nv are

their corresponding molar holdups.

The assumption of well mixing allows writing the following macroscopic balances:

Vapor phase:

Total mass balance:

vcAAv FNNF

dtdN

−+−=

(2.122)

Page 65: Modeling Notes Complete

62

Component balance for A:

AvAAAv yFNF

dtyNd

−−=)(

(2.123)

Since d(NvyA)/dt = Nv dyA/dt + yA dNv/dt, and using equation (2.122), equation (2.123)

can be written as follows:

AcAAAAA

v yNyNyFdt

dyN −−−−= )1()1(

(2.124)

Liquid phase:

Total mass balance:

cLcABL RFNNF

dtdN

−−−+=

(2.125)

Component balance for A:

cALAAL RxFN

dtxNd

−−=)(

(2.126)

Since d(NL xA)/dt = NL dxA/dt + xA dNL/dt, and using equation (2.125), equation (2.126)

can be written as follows:

AcAcABAAA

L xNxRxFxNdt

dxN +−−−−= )1()1(

(2.127)

Component balance for B:

Repeating the same reasoning used for component A, we can write:

Page 66: Modeling Notes Complete

63

BcBcBBBAB

L xNxRxFxNdt

dxN +−−+−= )1()1(

(2.128)

Energy balance, assuming Tref = 0:

QHRCpTFCpTFCpTFCpTFdt

HNHNdrcvLAABB

vvLL +−−−λ++=+

∆)()(

(2.129)

Note that:

dtNdCpT

dtTdCpN

dtNdH

dtHdN

dtHNd L

LL

LL

LLL )()()()()(

+=+=

(2.130)

dtNdCpT

dtTdCpN

dtNdH

dtHdN

dtHNd v

vv

vv

vvv )()()()()()(

λ++=+=

(2.131)

Substituting the last two equations, and using the definition of dNL/dt and dNv/dt from

equations (2.122) and (2.125), in equation (2.131) yields:

CpQNN

CpCpHTRTTFTTF

dtTdNN cA

rcBBAAvL +−

λ+−+−+−=+ )()()()()( ∆

(2.132)

The following additional equations are needed:

Vapor-liquid equilibrium relations:

yAP − xA PAs = 0 (2.133)

yAP − (1 − xA − xB) Pcs = 0 (2.134)

Total volume constraint:

V = VL + Vv (2.135)

Page 67: Modeling Notes Complete

64

Or, using ideal gas law for vapor volume and total volume and knowing that VL = NL/ρ ,

we can write:

nRT = NvRT + NLP/ρ (2.136)

or

V = NvRT/P + NL/ρ (2.137)

Degrees of freedom analysis:

• Forcing variables: FA, FB, TA, TB, Q, P

• Physical properties and parameters: ∆Hr, Cp, λ, R, ρ, V, sAP , s

CP

• Remaining variables: NA, Nc, NL, Nv, FL, T, xA , xB, yA

• Number of equations: (Eq. 2.122, 2.124, 2.125, 2.127, 2.128, 2.132, 2.133,2.134,

2.137)

The degree is freedom is 9-9=0 and the problem is exactly specified. Note that the

reaction rate Rc is defined and that the outlet flow Fv can be determined from the overall

mass balance.

2-1-11 Reaction with Mass Transfer

Figure 2.14 shows a chemical reaction that takes place in a gas-liquid

environment. The reactant A enters the reactor as a gas and the reactant B enters as a

liquid. The gas dissolves in the liquid where it chemically reacts to produce a liquid C.

The product is drawn off the reactor with the effluent FL. The un-reacted gas vents of

the top of the vessel. The reaction mechanism is given as follows:

A + B C (2.138)

Assumptions:

• Perfectly mixed reactor

• Isothermal operation

• Constant pressure, density, and holdup.

Page 68: Modeling Notes Complete

65

• Negligible vapor holdup.

Figure 2-14 Reaction with Mass Transfer

In such cases, when the two chemical phenomena, i.e., mass transfer and chemical

reaction, occur together, the reaction process may become mass transfer dominant or

reaction-rate dominant. If the mass transfer is slower reaction rate, then mass transfer

prevail and vise versa.

Due to the perfectly mixing assumption, macroscopic mass transfer of

component A from the bulk gas to the bulk liquid is approximated by the following

molar flux:

NA = KL (C*A − CA) (2.139)

where

KL is mass transfer coefficient

CA* is gas concentration at gas-liquid interface

CA is gas concentration in bulk liquid

To fully describe the process, we derive the macroscopic balance of the liquid phase

where the chemical reaction takes place. This results in:

Page 69: Modeling Notes Complete

66

Liquid phase:

Total mass balance:

LAmABB FNAMFdt

Vdρ−+ρ=

ρ (2.140)

Component balance on A:

rVCFNAdt

dCV ALAmA −−=

(2.141)

Component balance on B:

rVCFCFdt

dCV BLBoBB −−=

(2.142)

Vapor phase:

Here, since vapor holdup is negligible, we can write a steady state total continuity

equation as follows:

Fv = FA − MA Am NA/ρA (2.143)

where

Am total mass transfer area of the gas bubble

MA molecular weight of component A

ρ density

V liquid volume

Degrees of freedom analysis:

• Forcing variables: FA, FB, CB0,

• Parameters of constant values: KL, MA, Am, ρ, ρA, ρB,

Page 70: Modeling Notes Complete

67

• Remaining variables: CB, NA, CA, Fv, V

• Number of equations: (Eq. 2.139-2.143)

Note that the liquid flow rate, FL can be determined from the overall mass balance and

that the reaction rate r should be defined.

2.1.12 Multistage Heterogeneous Systems: Liquid-liquid extraction

There are many chemical processes which consist in a number of consecutive

stages in series. In each stage two streams are brought in contact for separating

materials due to mass transfer. The two streams could be flow in co-current or counter

current patterns. Counter-current flow pattern is known to have higher separation

performance Examples of these processes are distillation columns, absorption towers,

extraction towers and multi-stage flash evaporator where distillate water is produced

from brine by evaporation.

The same modeling approach used for single stage processes will be used for the

staged processes, where the conservation law will be written for one stage and then

repeated for the next stage and so on. This procedure will result in large number of state

equation depending on the number of stages and number of components.

The separation process generally takes place in plate, packed or spray-type

towers. In tray or spay-type columns the contact and the transfer between phases occur

at the plates. Generally, we can always assume good mixing of phases at the plates, and

therefore macroscopic balances can be carried out to model these type of towers.

Packed towers on the other hand are used for continuous contacting of the two phases

along the packing. The concentrations of the species in the phases vary obviously along

the tower. Packed towers are therefore typical examples of distributed parameters

systems that need to be modeled by microscopic balances.

In the following we present some examples of mass separation units that can be

modeled by simple ODE's, and we start with liquid-liquid extraction process.

Page 71: Modeling Notes Complete

68

Liquid-liquid extraction is used to move a solute from one liquid phase to another.

Consider the single stage countercurrent extractor, shown in Figure 2.15, where it is

desired to separate a solute (A) from a mixture (W) using a solvent (S). The stream

mixture with flow rate W (kg/s) enters the stage containing XAf weight fraction of solute

(A). The solvent with a flow rate (S) (kg/s) enters the stage containing YAf weight

fraction of species (A). As the solvent flows through the stage it retains more of (A) thus

extracting (A) from the stream (W). Our objective is to model the variations of the

concentration of the solute. A number of simplifying assumptions can be used:

• The solvent is immiscible in the other phase.

• The concentration XA and YA are so small that they do not affect the mass flow

rates. Therefore, we can assume that the flow rates W and S are constant. A total

mass balance is therefore not needed.

• An equilibrium relationship exists between the weight fraction YA of the solute in

the solvent (S) and its weight fraction XA in the mixture (W). The relationship can

be of the form:

YA = K XA (2.144)

Here K is assumed constant. Since both phases are assumed perfectly mixed a

macroscopic balance can be carried out on the solute in each phase. A component

balance on the solute in the solvent-free phase of volume V1 and density ρ1 gives:

AAAfA NWXWX

dtdXV −−=ρ 11

(2.145)

whereas NA (kg/s) is the flow rate due to transfer flow between the two phases. A similar

component balance on the solvent phase of volume V2 and density ρ2 gives:

AAAfA NSYSY

dtdYV +−=ρ 22

(2.146)

Since YA = K XA and K is constant, the last equation is equivalent to:

Page 72: Modeling Notes Complete

69

AAAfA NSKXSY

dtdXKV +−=22ρ

(2.147)

Adding Eq. 2.145 and 2.147 yields:

AAfAfA XKSWSYWX

dtdXKVV )()( 2211 +−+=+ ρρ

(2.148)

The latter is a simple linear ODE with unknown XA. With the volume V1, V2 and flow

rates W, S known the system is exactly specified and it can be solved if the initial

concentration is known:

XA(ti) = XAi (2.149)

Note that we did not have to express explicitly the transferred flux NA.

Figure 2-15 Single Stage Liquid-Liquid Extraction Unit

The same analysis can be extended to the multistage liquid/liquid extraction

units as shown in Figure 2.16. The assumptions of the previous example are kept and

we also assume that all the units are identical, i.e. have the same volume. They are also

assumed to operate at the same temperature.

Figure 2-16 Multi-Stage Liquid-Liquid Extraction Unit

Page 73: Modeling Notes Complete

70

A component balance in the ith stge (excluding the first and last stage) gives:

• Solvent-free phase, of volume V1i and density ρ1i

AiAiAiAi

ii NWXWXdt

dXV −−=ρ −111 (i=2…,N-1) (2.150)

where NAi is the flow rate due to transfer between the two phases at stage i.

• Solvent phase of volume V2i and density ρ2i:

AiAiAiAi

ii NSYSYdt

dYV +−= +122ρ … (i = 2…,N-1) (2.151)

Writing the equilibrium equation (Eq. 2.144) for each component YAi = K XAi

(i=1,.,.N) and adding the last two equations yield:

AiAiAiAi

iiii XKSWSYWXdt

dXKVV )()( 112211 +−+=ρ+ρ +− (i=2…,N-1) (2.152)

Since the volume and densities are equal, i.e.:

V1i = V1 and V2i = V (2.153)

ρ1i = ρ1 and ρ2i = ρ2 (2.154)

Equation 2.152 is therefore equivalent to:

AiAiAiAi XKSWSYWX

dtdXKVV )()( 112211 +−+=ρ+ρ +− (i=2…,N-1)

(2.155)

The component balance in the first stage is:

Page 74: Modeling Notes Complete

71

111

2211 )()( AAAfA XKSWSYWX

dtdXKVV +−+=ρ+ρ

(2.156)

And that for the last stage is:

ANAfANAN XKSWSYWX

dtdXKVV )()( 12211 +−+=ρ+ρ −

(2.157)

The model is thus formed by a system of linear ODE's (Eq. 2.155, 2.156, 2.157)

which can be integrated if the initial conditions are known:

XA(ti) = XAi (i=1,2…,N) (2.158)

Degrees of freedom analysis

• Parameter of constant values: ρ1, ρ2, K, V1, V2, W and S

• (Forced variable): XAf, YAf

• Remaining variables: XAi (2N variables): (i=1,2…,N) and YAi (i=1,2…,N)

• Number of equations: 2N [2.144 (N equations, one for each component), Eq.

2.155 (N-2 eqs), 2.156(1 eq), 2.157(1 eq)].

The problem is therefore is exactly specified.

2.1.13 Binary Absorption Column

Consider a N stages binary absorption tower as shown in figure 2.17. A Liquid

stream flows downward with molar flow rate (L) and feed composition (xf). A Vapor

stream flows upward with molar flow rate (G) and feed composition (yf). We are

interested in deriving an unsteady state model for the absorber. A simple vapor-liquid

equilibrium relation of the form of:

yi = a xi + b (2.159)

can be used for each stage i (i=1,2,…,N).

Assumptions:

Page 75: Modeling Notes Complete

72

• Isothermal Operation

• Negligible vapor holdup

• Constant liquid holdup in each stage

• Perfect mixing in each stage

According to the second and third assumptions, the molar rates can be considered

constants, i.e. not changing from one stage to another, thus, total mass balance need not

be written. The last assumption allows us writing a macroscopic balance on each stage

as follows:

Component balance on stage i:

)()( 11 iiiii xxLyyG

dtdxH −+−= +− (i=2…,N-1)

(2.160)

where H is the liquid holdup, i.e., the mass of liquid in each stage. The last equation is

repeated for each stage with the following exceptions for the last and the first stages:

Figure 2-17 N-stages Absorbtion Tower

In the last stage, xi+1 is replaced by xf

In the first stage, yi-1 is replaced by yf

Page 76: Modeling Notes Complete

73

Degrees of freedom analysis

• Parameter of constant values: Η, a, b

• (Forced variable): G, L, xf, yf

• Remaining variables: xi (i=1,2…,N), yi (i=1,2…,N)

• Number of equations:2N (Eqs.2.159, 2.160)

The problem is therefore is exactly specified.

2.1.14 Multi-component Distillation Column

Distillation columns are important units in petrochemical industries. These units

process their feed, which is a mixture of many components, into two valuable fractions

namely the top product which rich in the light components and bottom product which is

rich in the heavier components. A typical distillation column is shown in Figure 2.18.

The column consists of n trays excluding the re-boiler and the total condenser. The

convention is to number the stages from the bottom upward starting with the re-boiler as

the 0 stage and the condenser as the n+1 stage.

Description of the process:

The feed containing nc components is fed at specific location known as the feed tray

(labeled f) where it mixes with the vapor and liquid in that tray. The vapor produced

from the re-boiler flows upward. While flowing up, the vapor gains more fraction of the

light component and loses fraction of the heavy components. The vapor leaves the

column at the top where it condenses and is split into the product (distillate) and reflux

which returned into the column as liquid. The liquid flows down gaining more fraction

of the heavy component and loses fraction of the light components. The liquid leaves

the column at the bottom where it is evaporated in the re-boiler. Part of the liquid is

drawn as bottom product and the rest is recycled to the column. The loss and gain of

materials occur at each stage where the two phases are brought into intimate phase

equilibrium.

Page 77: Modeling Notes Complete

74

Bxb

Fz

Dxd

Cw

steam

Figure 2-18 Distillation Column

Modeling the unit:

We are interested in developing the unsteady state model for the unit using the flowing

assumptions:

• 100% tray efficiency

• Well mixed condenser drum and re-boiler.

• Liquids are well mixed in each tray.

• Negligible vapor holdups.

• liquid-vapor thermal equilibrium

Since the vapor-phase has negligible holdups, then conservation laws will only be

written for the liquid phase as follows:

Stage n+1 (Condenser), Figure 2.19a:

Total mass balance:

Page 78: Modeling Notes Complete

75

)( DRVdt

dMn

D +−=

(2.161)

Component balance:

1,1)()(

,,, −=+−= ncjxDRyV

dtxMd

jDjnnjDD

(2.162)

Energy balance:

cDnnDD QhDRhV

dthMd

−+−= )()( (2.163)

Note that R = Ln+1 and the subscript D denotes n+1

Stage n, Figure fig2.19b

Total Mass balance:

nnnn LRVV

dtdM

−+−= −1 (2.164)

Component balance:

1,1)(

,,,,11, −=−+−= −− ncjxLRxyVyV

dtxMd

jnnjDjnnjnnjnn

(2.165)

Energy balance:

nnDnnnnnn hLRhHVHV

dthMd

−+−= −− 11)(

(2.166)

Stage i, Figure 2.19c

Total Mass balance:

Page 79: Modeling Notes Complete

76

iiiii LLVV

dtdM

−+−= +− 11 (2.167)

Component balance:

1,1)(

,,11,,11, −=−+−= ++−− ncjxLxLyVyV

dtxMd

jiijiijiijiijii

(2.168)

Energy balance:

iiiiiiiiii hLhLHVHV

dthMd

−+−= ++−− 1111)(

(2.169)

Stage f (Feed stage), Figure 2.19d

Total Mass balance:

)())1(( 11 qFLLFqVVdt

dMffff

f +−+−+−= +− (2.170)

Component balance:

1,1

)())1(()(

,,11,,11,

−=

+−+−+−= ++−−

ncj

qFzxLxLFzqyVyVdt

xMdjjffjffjjffjff

jff

(2.171)

Energy balance:

)())1(()(

1111 ffffffffffff qFhhLhLFhqHVHV

dthMd

+−+−+−= ++−− (2.172)

Stage 1, Figure 2.19e

Total Mass balance:

Page 80: Modeling Notes Complete

77

1211 LLVV

dtdM

B −+−= (2.173)

Component balance:

1,1)(

,11,22,11,,11 −=−+−= ncjxLxLyVyV

dtxMd

jjjjBBj

(2.174)

Energy balance:

11221111 )( hLhLHVHV

dthMd

BB −+−= (2.175)

Stage 0 (Re-boiler), Figure 2.19f

Total Mass balance:

BLVdt

dMB

B −+−= 1 (2.176)

Component balance:

1,1)(

,,11,, −=−+−= ncjBxxLyV

dtxMd

jBjjBBjBB

(2.177)

Energy balance:

rBBBBB QBhhLHV

dthMd

+−+−= 11)(

(2.178)

Note that L0 = B and B denotes the subscript 0

Additional given relations:

Phase equilibrium: yj = f (xj, T,P)

Liquid holdup: Mi = f (Li)

Page 81: Modeling Notes Complete

78

Enthalpies: Hi = f (Ti, yi,j), hi = f (Ti, xi,j)

Vapor rates: Vi = f (P)

Notation:

Li, Vi Liquid and vapor molar rates

Hi, hi Vapor and liquid specific enthalpies

xi, yi Liquid and vapor molar fractions

Mi Liquid holdup

Q Liquid fraction of the feed

Z Molar fractions of the feed

F Feed molar rate

Degrees of freedom analysis

Variables

Mi n

MB, MD 2

Li n

B,R,D 3

xi,j n(nc − 1)

xB,j,xD,j 2(nc − 1)

yi,j n(nc − 1)

yB,j nc − 1

hi n

hB, hD 2

Hi n

HB 1

Vi n

VB 1

Ti n

TD, TB 2

Total 11+6n+2n(nc−1)+3(nc−1)

Page 82: Modeling Notes Complete

79

Equations:

Total Mass n + 2

Energy n + 2

Component (n + 2)(nc − 1)

Equilibrium n(nc − 1)

Liquid holdup n

Enthalpies 2n+2

Vapor rate n

hB = h1 1

yB = xB (nc − 1)

Total 7+6n+2n(nc-1)+3(nc-1)

Constants: P, F, Z

Therefore; the degree of freedom is 4

To well define the model for solution we include four relations imported from inclusion

of four feedback control loops as follows:

• Use B, and D to control the liquid level in the condenser drum and in the re-

boiler.

• Use VB and R to control the end compositions i.e., xB, xD

Page 83: Modeling Notes Complete

80

R, xd

Ln, xn

Vn, yn

Vn-1, yn-1

stage n

Li, xi Vi-1, yi-1

stage i

Li+1, xi+1 Vi, yi

(a)(b)

Lf+1, xf+1

Lf, xf Vf-1, yf-1

stage f

Vf , yf

(c)

L2, x2

L1, x1

V1, y1

stage 1

(d)

VB, yB

VB, yB

L1, x1

B, xB

(e)

Vn, yn

R, xd D, xd

Qc

Qr

(f)

Figure 2-19 Distillation Column Stages

Simplified Model

One can further simplify the foregoing model by the following assumptions:

(a) Equi-molar flow rates, i.e. whenever one mole of liquid vaporizes a tantamount

of vapor condenses. This occur when the molar heat of vaporization of all

components are about the same. This assumption leads to further idealization

that implies constant temperature over the entire column, thus neglecting the

energy balance. In addition, the vapor rate through the column is constant and

equal to:

VB = V1 = V2 =… = Vn (2.179)

Page 84: Modeling Notes Complete

81

(b) Constant relative volatility, thus a simpler formula for the phase equilibrium can be

used:

yj = αj xj/(1+(αj − 1) xj) (2.180)

Degrees of Freedom:

Variables:

Mi, MB, MD n + 2

Li, B,R,D n + 3

xi,xB,xD (n + 2)(nc − 1)

yj, yB (n + 1)(nc − 1)

V 1

Total 2 + 2n + (2n + 3)(nc − 1)

Equations:

Total Mass n + 2

Component (n + 2)(nc − 1)

Equilibrium n(nc − 1)

Liquid holdup n

yB = xB 1

Total 2+2n+(2n+3)(nc-1)

It is obvious that the degrees of freedom is still 4.

Page 85: Modeling Notes Complete

82

2.2 Examples of Distributed Parameter Systems

2.2.1 Liquid Flow in a Pipe

Consider a fluid flowing inside a pipe of constant cross sectional area (A) as

shown in Figure 2.20. We would like to develop a mathematical model for the change

in the fluid mass inside the pipe. Let v be the velocity of the fluid. Clearly the velocity

changes with time (t), along the pipe length (z) and also with the radial direction (r). In

order to simplify the problem, we assume that there are no changes in the radial

direction. We also assume isothermal conditions, so only the mass balance is needed.

Since the velocity changes with both time and space, the mass balance is to be carried

out on microscopic scale. We consider therefore a shell element of width ∆z and

constant cross section area (A) as shown in Fig. 2.20.

Figure 2-20 Liquid flow in a pipe

Mass into the shell:

ρvA∆t|z (2.181)

where the subscript (.|z) indicates that the quantity (.) is evaluated at the distance z.

Mass out of the shell:

ρvA∆t|z+∆z (2.182)

Accumulation:

Page 86: Modeling Notes Complete

83

ρA∆z|t+∆t − ρA∆z|t (2.183)

Similarly the subscript (.|t) indicates that the quantity (.) is evaluated at the time t.

The mass balance equation is therefore:

ρvA∆t|z = ρvA∆t|z+∆z + ρA∆z|t+∆t − ρA∆z|t (2.184)

We can check for consistency that the units in each term are in (kg). Dividing Eq. 2.184

by ∆t ∆z and rearranging yields:

zvAvA

tAA

zzzttt

∆∆∆∆ ++

ρ−ρ=

ρ−ρ )()()()(

(2.185)

Taking the limit as ∆t 0 and ∆z 0 gives:

zvA

tA

∂ρ∂

−=∂ρ∂ )()(

(2.186)

Since the cross section area (A) is constant, Eq. 2.186 yields:

zv

t ∂∂

−=∂∂ )(ρρ

(2.187)

or

0)(=

∂∂

+∂∂

zv

tρρ

(2.188)

The ensuing equation is a partial differential equation (PDE) that defines the variation

of ρ and v with the two independent variables t and z. This equation is known as the

one-dimensional continuity equation. For incompressible fluids for which the density is

constant, the last equation can also be written as:

Page 87: Modeling Notes Complete

84

0=∂∂

zv

(2.189)

This indicates that the velocity is independent of axial direction for one dimensional

incompressible flow.

2.2.2 Velocity profile inside a pipe

We reconsider the flow inside the pipe of the previous example. Our objective is

to find the velocity profile in the pipe at steady state. For this purpose a momentum

balance is needed. To simplify the problem we also assume that the fluid is

incompressible. We will carry out a microscopic momentum balance on a shell with

radius r, thickness ∆r and length ∆z as shown in Fig 2.21.

Figure 2-21 Velocity profile for a laminar flow in a pipe

Momentum in:

(τrz2πr∆z)|r (2.190)

where τrz is the shear stress acting in the z-direction and perpendicular to the radius r.

Momentum out:

Page 88: Modeling Notes Complete

85

(τrz2πr∆z)|r+∆r (2.191)

As for the momentum generation we have mentioned earlier in section 1.8.3 that the

generation term corresponds to the sum of forces acting on the volume which in this

example are the pressure forces, i.e.

(PA)|z – (PA)|z+∆z = P(2πr∆r)|z − P(2πr∆r)|z+∆z (2.192)

There is no accumulation term since the system is assumed at steady state. Substituting

this term in the balance equation (Eq. 1.6) and rearranging yields,

zPPr

rrr zzzrrzrrrz

∆∆∆∆ )||(|| ++ −

=τ−τ

(2.193)

We can check for consistency that all the terms in this equation have the SI unit of

(N/m2). Taking the limit of (Eq. 2.193) as ∆z and ∆r go to zero yields:

dzdPr

drrd rz −=τ )(

(2.194)

Here we will make the assumption that the flow is fully developed, i.e. it is not

influenced by the entrance effects. In this case the term dP/dz is constant and we have:

LP

LPP

dzdP ∆

=−

= 12 (2.195)

where L is the length of the tube. Note that equation (2. 195) is a function of the shear

stress τrz, but shear stress is a function of velocity. We make here the assumption that

the fluid is Newtonian, that is the shear stress is proportional to the velocity gradient:

drdvz

rz µτ −= (2.196)

Page 89: Modeling Notes Complete

86

Substituting this relation in Eq. 2.194 yields:

LPr

drrdv

drd z ∆

=)(µ (2.197)

or by expanding the derivative:

LP

drdv

rdrvd zz ∆

=+ )1( 2

2

µ (2.198)

The system is described by the second order ODE (Eq. 2.198). This ODE can be

integrated with the following conditions:

• The velocity is zero at the wall of the tube

vz = 0 at r = R (2.199)

• Due to symmetry, the velocity profile reaches a maximum at the center of the

tube:

0=drdvz at r = 0

(2.200)

Note that the one-dimensional distributed parameter system has been reduced to a

lumped parameter system at steady state.

2.2.3 Diffusion with chemical reaction in a slab catalyst

We consider the diffusion of a component A coupled with the following

chemical reaction A B in a slab of catalyst shown in figure 2.22. Our objective is to

determine the variation of the concentration at steady state. The concentration inside the

Page 90: Modeling Notes Complete

87

slab varies with both the position z and time t. The differential element is a shell

element of thickness ∆z.

Flow of moles A in:

(SNA)|z (2.201)

where S (m2) is the surface area and NA (moles A/s m2) is the molar flux.

Flow of moles A out:

(SNA)|z+∆z (2.202)

Rate of generation of A:

−(S∆z)r (2.203)

where r = kCA is the rate of reaction, assumed to be of first order. There is no

accumulation term since the system is assumed at steady state. The mass balance

equation is therefore,

(SNA)|z − (SNA)|z+∆z − (S∆z)kCA =0 (2.204)

z = 0 z = L

Porous catalyst particle

Exterior surface

z

Figure 2-22 diffusion with chemical reaction inside a slab catalyst

Page 91: Modeling Notes Complete

88

Dividing equation (2.204) by S∆z results in:

0|)(|)(=−

− +A

zzAzA kCzNN

∆∆

(2.205)

Taking the limit when ∆z 0, the last equation becomes:

0=− AA kC

dzdN

(2.206)

The molar flux is given by Fick's law as follows:

dzdC

DN AAA −=

(2.207)

where DA is diffusivity coefficient of (A) inside the catalyst particle. Equation (2.206)

can be then written as follows:

02

2

=− AA

A kCdz

CdD

(2.208)

This is also another example where a one-dimensional distributed system is reduced to a

lumped parameter system at steady state. In order to solve this second-order ODE, the

following boundary conditions could be used:

at z = L, CA = CAo (2.209)

at z = 0, dCA/dr = 0 (2.210)

The first condition imposes the bulk flow concentration CAo at the end length of the

slab. The second condition implies that the concentration is finite at the center of the

slab.

Page 92: Modeling Notes Complete

89

2.2.4 Temperature profile in a heated cylindrical Rod

Consider a cylindrical metallic rod of radius R and length L, initially at a

uniform temperature of To. Suppose that one end of the rod is brought to contact with a

hot fluid of temperature Tm while the surface area of the rod is exposed to ambient

temperature of Ta. We are interested in developing the mathematical equation that

describes the variation of the rod temperature with the position. The metal has high

thermal conductivity that makes the heat transfer by conduction significant. In addition,

the rod diameter is assumed to be large enough such that thermal distribution in radial

direction is not to be neglected. The system is depicted by figure 2.23. For modeling

we take an annular ring of width ∆z and radius ∆r as shown in the figure. The following

transport equation can be written:

Figure 2-23 Temperature Distribution In a cylindrical rod

Heat flow in by conduction at z:

qz(2π r ∆r)∆t (2.211)

Heat flow in by conduction at r:

qr(2π r ∆z)∆t (2.212)

Page 93: Modeling Notes Complete

90

where qz and qr are the heat flux by conduction in the z and r directions.

Heat flow out by conduction at z+∆z:

qz+∆z(2π r ∆r)∆t (2.213)

Heat flow out by conduction at r+∆r:

qr+∆r(2π (r+∆r)∆z)∆t (2.214)

Heat accumulation:

ρ(2π r ∆r ∆z)( h~ t+∆t − h~ t) (2.215)

where h~ is the specific enthalpy. Summing the above equation according to the

conservation law and dividing by (2π ∆r ∆z ∆t), considering constant density, gives:

rrqrq

zqqr

thhr rrrzzzttt

∆−

+∆

−=

∆− ∆+∆+∆+

~~ρ

(2.216)

Taking the limit of ∆t, ∆z, and ∆r go to zero yield:

rrq

zqr

thr rz

∂∂

−∂∂

−=∂∂ )(~

ρ (2.217)

Dividing by r and replacing h~ by pC (T − Tref), where pC is the average heat

capacity, and substituting the heat fluxes q with their corresponding relations (Fourier

law):

zTkq zz ∂

∂−=

(2.218)

Page 94: Modeling Notes Complete

91

and

rTkq rr ∂

∂−=

(2.219)

Equation (2.217) is then equivalent to:

)(2

2

rTr

rrk

zTk

tTpC r

z ∂∂

∂∂

+∂∂

=∂∂ρ

(2.220)

This is a PDE where the temperature depends on three variables: t, z, and r. If we

assume steady state conditions then the PDE becomes:

)(0 2

2

rTr

rrk

zTk r

z ∂∂

∂∂

+∂∂

= (2.221)

If in addition to steady state conditions, the radius of the rod is assumed to be small

so that the radial temperature gradient can be neglected then the PDE (Eq. 2.221) can

be further simplified. In this case, the differential element upon which the balance

equation is derived is a disk of thickness ∆z and radius R. The heat conduction in the

radial direction is omitted and replaced by the heat transfer through the surface area

which is defined as follows:

Q = U(2πRL)(Ta – T) (2.222)

Consequently, the above energy balance equation (Eq. 2.221) is reduced to the

following ODE

))(2(0 2

2

TTRLUdz

Tdk az −−= π (2.223)

2.2.5 Isothermal Plug Flow Reactor

Page 95: Modeling Notes Complete

92

Let consider a first-order reaction occurring in an isothermal tubular reactor as

shown in figure 2.24. We assume plug flow conditions i.e. the density, concentration

and velocity change with the axial direction only. Our aim is to develop a model for the

reaction process in the tube.

z

z = L z z+ z

v(t,z)

(t,z), CA(t,z)

CAo CA

Figure 2-5 Isothermal Plug flow reactor

In the following we derive the microscopic component balance for species (A) around

differential slice of width ∆z and constant cross-section area (S).

Flow of moles of A in:

As has been indicated in section 1.11.1 mass transfer occurs by two mechanism;

convection and diffusion. The flow of moles of species A into the shell is therefore the

sum of two terms:

(vCA S ∆t) |z + (NA S ∆t)|z (2.224)

where NA is the diffusive flux of A ( moles of A/m2 s).

Flow of moles of A out:

(vCA S ∆t) |z+∆z + (NA S ∆t)|z+∆z (2.225)

Accumulation:

(CA S ∆z) |t+∆t − (CA S ∆z) |t (2.226)

Page 96: Modeling Notes Complete

93

Generation due to reaction inside the shell:

− r(S∆z∆t) (2.227)

where r = k CA is the rate of reaction.

Substituting all the terms in the mass balance equation (Eq. 1.3) and dividing by ∆t and

∆z gives:

SkCz

SNSvCSNSvCt

SCSCA

zzAAzAAtAttA −+−+

=− ++

∆∆∆∆ |)(|)(|)(|)(

(2.228)

Taking the limit of ∆t 0 and ∆z 0 and omitting S from both sides give the

following PDE:

AAAA kC

zN

zvC

tC

−∂

∂−

∂∂

−=∂

∂ (2.229)

where NA is the molar flux given by Fick’s law as follows:

dzdCDN A

ABA −= (2.230)

where DAB is the binary diffusion coefficient. Equation 2.229 can be then written as

follows:

AA

AbAA kC

zCD

zvC

tC

−∂

∂+

∂∂

−=∂

∂2

2)( (2.231)

Expanding the derivatives, the last equation can be reduced to:

AA

AbAAA kC

zCD

zvC

zCv

tC

−∂

∂+

∂∂

−∂

∂−=

∂∂

2

2

(2.232)

Page 97: Modeling Notes Complete

94

This equation can be further simplified by using the mass balance equation for

incompressible fluids (Eq. 2.189). We get then:

AA

AbAA kC

zCD

zCv

tC

−∂

∂+

∂∂

−=∂

∂2

2

(2.233)

The equation is a PDE for which the state variable (CA) depends on both t and z.

The PDE is reduced at steady state to the following second order ODE,

AA

AbA kC

dzCdD

dzdCv −+−= 2

2

0 (2.234)

The ODE can be solved with the following boundary conditions (BC):

BC1: at z = 0 CA(0) = CA0 (2.235)

BC2: at z = L 0)(

=dz

zdCA (2.236)

The first condition gives the concentration at the entrance of the reactor while the

second condition indicates that there is no flux at the exit length of the reactor.

2.2.6 Non-Isothermal Plug-Flow reactor

The tubular reactor discussed earlier is revisited here to investigate its behavior

under non-isothermal conditions. The heat of reaction is removed via a cooling jacket

surrounding the reactor as shown in figure 2.25. Our objective is to develop a model

for the temperature profile along the axial length of the tube. For this purpose we will

need to write an energy balance around an element of the tubular reactor, as shown in

Fig.2.25. The following assumptions are made for the energy balance:

Page 98: Modeling Notes Complete

95

Figure 2-25 Non-isothermal plug flow reactor

Assumptions:

• Kinetic and potential energies are neglected.

• No Shaft work

• Internal energy is approximated by enthalpy

• Energy flow will be due to bulk flow (convection) and conduction.

Under these conditions, the microscopic balance around infinitesimal element of width

∆z with fixed cross-section area is written as follows:

Energy flow into the shell:

As mentioned in Section 1.11.3 the flow of energy is composed of a term due to

convection and another term due to molecular conduction with a flux qz.

(qzA + v Aρ h~ )∆t|z (2.237)

Energy flow out of the shell:

(qzA + v Aρ h~ )∆t|z+∆z (2.238)

Accumulation of energy:

(ρA h~ ∆z)|t+∆t − (ρA h~ ∆z)|t (2.239)

Page 99: Modeling Notes Complete

96

Heat generation by reaction:

(−∆Hr )kCA A ∆z ∆t (2.240)

Heat transfer to the wall:

ht(πD∆z)(T − Tw)∆t (2.241)

where ht is film heat transfer coefficient.

Substituting these equations in the conservation law (equation 1.7) and dividing by

Α∆t ∆z give:

))((

|||)~(|)~(|)~(|)~(

wtAr

zzzzzzzzttt

TTADhkCH

zqq

zhvhv

thh

−−∆−

∆−

+∆−

=∆

− ∆+∆+∆+

π

ρρρρ

(2.242)

Taking the limit as ∆t and ∆z go to zero yields:

))(()~()~(wtAr

z TTADhkCH

zq

zhv

th

−−∆−∂∂

−∂

∂−=

∂∂ πρρ

(2.243)

The heat flux is defined by Fourier’s law as follows:

zTkq tz ∂

∂−=

(2.244)

where kt is the thermal conductivity. The specific enthalpy ( h~ ) can be approximated by:

)(~refTTpCh −= (2.245)

Since the fluid is incompressible it satisfies the equation of continuity (Eq. 2.189).

Substituting these expressions in Eq. 2.243 and expanding gives:

Page 100: Modeling Notes Complete

97

))((/2

2

wtARTE

ort TTADhCekH

zTk

zTpvC

tTpC −−∆−

∂∂

+∂∂

−=∂∂ − πρρ

(2.246)

At steady state this PDE becomes the following ODE,

))((0 /2

2

wtARTE

ort TTADhCekH

dzTdk

dzdTpvC −−∆−+−= − πρ

(2.247)

Similarly to Eq. 2.234 we could impose the following boundary conditions:

B.C1:

at z = 0 T(z) = To (2.248)

B.C2:

at z = L 0)(

=dz

zdT (2.249)

The first condition gives the temperature at the entrance of the reactor and the second

condition indicates that there is no flux at the exit length of the reactor.

2.2.7 Heat Exchanger: Distributed parameter model

We revisit the shell-and-tube heat exchanger already discussed in example 2.1.6.

Steam of known temperature Ts flowing around the tube is heating a liquid L of density

ρL and constant velocity v from temperature TL1 to TL2. The temperature in the tube

varies obviously with axial direction z, radial direction r and time t. To simplify the

problem we will assume that there are no change in the radial direction. This

assumption is valid if the radius is small and no large amount of heat is transferred. The

heat transfer from the steam to the liquid depends on the heat transfer coefficient on the

steam side, hto and on the transfer on the liquid side hti. We also neglect the thermal

capacity of the metal wall separating the steam and the liquid and assume that the

exchange between the steam and liquid occurs with an overall heat transfer coefficient

U. We also assume constant heat capacity for the liquid. An energy balance on a

differential element of the exchanger of length ∆z and cross-sectional area A, yields:

Page 101: Modeling Notes Complete

98

Flow of energy in:

(vAρ pC )∆t|z (2.250)

Flow of energy out:

(vAρ pC )∆t|z+∆z (2.251)

Energy accumulation:

(Aρ h~ ∆z)|t+∆t − (Aρ h~ ∆z)|t (2.252)

Energy generated:

U(πD∆z)(T – Ts)∆t (2.253)

Using the expression for specific enthalpy (Eq. 2.245) and dividing by A∆t∆z, the

energy balance yields:

( ) ( )))((|)(|)(|)(|)(

szzzrefttref TT

ADU

zpTCvApTCvA

ttzTTpCAzTTpCA

−−∆

−=

∆−−∆− ∆+∆+ πρρρρ (2.254)

Taking the limit as ∆t an ∆z goes to zero gives:

))(()()(sTT

ADU

zpTCvA

tpTCA

−−∂

∂−=

∂∂ πρρ

(2.255)

Since A = πD2/4 and dividing by ρ pC Eq. (2.255) is equivalent to:

))(4( sTTpDCU

zTv

tT

−−∂∂

−=∂∂

ρ

(2.256)

Page 102: Modeling Notes Complete

99

At steady state the PDE becomes the following ODE,

))(4(0 sTTpDCU

dzdTv −−−=

ρ

(2.257

With the following condition:

0)0( TzT == (2.258

2.2.8 Mass exchange in packed column

In previous section (section 2.1.12) we presented some examples of mass

transfer units that can be described by simple ODE's. This includes all the operations

that can occur in tray or spray-tray towers. In this section we present an example of

modeling a mass transfer operation that occurs in packed tower. Absorption is a mass

transfer process in which a vapor solute (A) in a gas mixture is absorbed by contact with

a liquid phase in which the solute is more or less soluble. The gas phase consists usually

of an inert gas and the solute. This process involves flow transfer of the solute A

through a stagnant non diffusive gas B into a stagnant liquid C. The liquid is mainly

immiscible in the gas phase. An example is the absorption of ammonia (A) from air (B)

by liquid water (C). The operation can be carried out either in tray (plate) towers or in

packed towers. The operation in tray towers can be modeled similarly to the liquid-

liquid extraction process in Example 2.1.12 and it is left as an exercise. We consider

here the absorption taking place in a packed tower.

Consider the binary absorption tower shown in Figure 2.26. A liquid stream flow

downward with molar flow rate L and feed composition (XAf). Vapor stream flows

upward with molar flow rate (G) and feed composition (YAf). A simple vapor-liquid

equilibrium relation of the form of:

YA = HXA (2.265)

is used, where H (mole fraction gas/mole fraction liquid) is the Henry's law constant.

This assumption is valid for dilute streams. The molar rates can be considered constants,

Page 103: Modeling Notes Complete

100

i.e. not changing from one stage to another, thus the total mass balance need not be

written. To establish the model equations we need to write equations for liquid and

vapor phase. To simplify the problem we assume constant liquid and vapor holdup in

each stage. We also assume isothermal conditions. An energy balance therefore is not

needed.

The flux NA transferred from bulk liquid to bulk gas is given by:

NA = KY (YA – YA*) (2.266)

Where KY is the overall mass transfer in the gas-phase (kgmole/m2s mole fraction) and

YA* is the value that would be in equilibrium with XA. The flux can also be expressed as:

NA = KX (XA – XA*) (2.267)

Where KX is the overall mass transfer coefficient in the liquid-phase and XA* is the value

that would be in equilibrium with YA.

A mass balance on the liquid phase for a differential volume (Fig. 2.26) of the column

length z and cross sectional area S yields:

Flow of mole in:

[(SLXA)∆t|z + (NAS∆t) |z]∆z (2.268)

Flow of moles out:

(SLXA)∆t|z+∆z (2.269)

Rate of accumulation:

(SHLXA∆z)|t+∆t - (SHLXA∆z)|t (2.270)

where XA is the liquid fraction of A and HL the liquid holdup (mole/m3).

Page 104: Modeling Notes Complete

101

Figure 2-26 Packed column

The balance equation yields:

(SHLXA∆z)|t+∆t - (SHLXA∆z)|t = (SLXA)∆t|z+∆z - [(SLXA)∆t|z + (NAS∆t)∆t|z]∆z (2.271)

Dividing by S∆t∆z and taking the limits as ∆z and ∆t goes to zero yield:

AAA

L Nz

XL

tX

H +∂

∂=

∂∂

(2.272)

which is equivalent to

)( *AAY

AAL YYK

zXL

tXH −+

∂∂

=∂

∂ (2.273)

We could also use the expression of flux (Eq. 2.267):

Page 105: Modeling Notes Complete

102

)( *AaX

AAL XXK

zXL

tXH −+

∂∂

=∂

∂ (2.274)

We can develop material balances for the gas phase that are similar to Eq. 2.274. This

gives:

)( *AAY

AAG YYK

zYG

tXH −+

∂∂

−=∂

∂ (2.275)

or alternatively:

)( *AaX

AAG XXK

zYG

tYH −+

∂∂

−=∂

∂ (2.276)

It should be noted that the analysis carried here can be used for a number of operations

where packed columns are used. This includes liquid-liquid extraction, gas-liquid

absorption and gas-solid drying. In each of these operations an equilibrium relation of

the type:

YA = f (XA) (2.277)

is generally available.

Page 106: Modeling Notes Complete

103

Chapter 3: Equations of Change

In the last chapter, we presented examples of microscopic balances in one or two

dimensions for various elementary examples. In this chapter we present the general balance

equations in multidimensional case. The balances, also called equations of changes can be

written in cartesian, cylindrical or spherical coordinates. We will explicitly derive the balance

equations in cartesian coordinates and present the corresponding equations in cylindrical and

spherical coordinates. The reader can consult the books in reference for more details. Once

the equations are presented we show through various examples how they can be used in a

systematic way to model distributed parameter models.

3.1 Total Mass balance

Our control volume is the elementary volume ∆x∆y∆z shown in Figure 3.1. The

volume is assumed to be fixed in space. To write the mass balance around the volume we

need to consider the mass entering in the three directions x,y, and z.

Page 107: Modeling Notes Complete

104

Figure 0-1 Total Mass balance in Cartesian coordinates Mass in:

The mass entering in the x-direction at the cross sectional area (∆y∆z) is

(ρvx)|x ∆y∆z∆t (3.1)

The mass entering in the y-direction at the cross sectional area (∆x∆z) is

(ρvy)|y ∆x∆z∆t (3.2)

The mass entering in the z-direction at the cross sectional area (∆x∆y) is

(ρvz)|z ∆x ∆y ∆t (3.3)

Mass out:

The mass exiting in the x-direction is:

(ρvx)|x+∆x ∆y∆z∆t (3.4)

Page 108: Modeling Notes Complete

105

The mass exiting in the y-direction is:

(ρvy)|y+∆y ∆x∆z∆t (3.5)

The mass exiting in the z-direction is:

(ρvz)|z+∆z ∆x∆y ∆t (3.6)

Rate of accumulation:

The rate of accumulation of mass in the elementary volume is:

(ρ)|t+∆t ∆x ∆y ∆z - (ρ)|t ∆x ∆y ∆z (3.7)

Since there is no generation of mass, applying the general balance equation Eq. 1.2 and

rearranging gives:

(ρ|t+∆t − ρ|t) ∆x ∆y ∆z = (ρvx|x − ρvx|x+∆x)∆y∆z∆t + (ρvy|y − ρvy|y+∆y)∆x∆z∆t

+ (ρvz|z − ρvz|z+∆z) ∆x ∆y ∆t (3.8)

Dividing the equation by ∆x ∆y ∆z∆t results in:

zvv

yvv

xvv

tzzzzzyyyyyxxxxxttt

∆∆∆∆∆∆∆∆ ++++ ρ−ρ

+ρ−ρ

+ρ−ρ

=ρ−ρ ||||||||

(3.9)

By taking the limits as ∆y,∆x,∆,z and ∆t goes to zero, we obtain the following equation of

change:

zv

yv

xv

tzyx

∂ρ∂

−∂ρ∂

−∂ρ∂

−=∂ρ∂

(3.10)

Expanding the partial derivative of each term yields after some rearrangement:

Page 109: Modeling Notes Complete

106

)(zv

yv

xv

zv

yv

xv

tzyx

zyx ∂∂

+∂∂

+∂∂

ρ−=∂ρ∂

+∂ρ∂

+∂ρ∂

+∂ρ∂

(3.11)

This is the general form of the mass balance in cartesian coordinates. The equation is also

known as the continuity equation. If the fluid is incompressible then the density is assumed

constant, both in time and position. That means the partial derivatives of ρ are all zero. The

total continuity equation (Eq. 3.11) is equivalent to:

)(0zv

yv

xv zyx

∂∂

+∂∂

+∂∂

ρ−= (3.12)

or simply:

zv

yv

xv zyx

∂∂

+∂∂

+∂∂

=0 (3.13)

3.2 Component Balance Equation

We consider a fluid consisting of species A, B … , and where a chemical reaction is

generating the species A at a rate rA (kg/m3s). The fluid is in motion with mass-average

velocity v = nt/ρ (m/s) where nt = nA + nB + … (kg/m2s) is the total mass flux and ρ (kg/m3s) is

the density of the mixture. Our objective is to establish the component balance equation of A

as it diffuses in all directions x,y,z (Figure 3.2).

Page 110: Modeling Notes Complete

107

x

z

y

nAx|xnAx|x+ x

x

y

znAz|z

nAz|z+ znAy|y

nAy|y+ y

Figure 0-2 Mass balance of component A

Mass of A in:

The mass of species A entering the x-direction at the cross sectional (∆y∆z) is:

(nAx)|x ∆y∆z∆t (3.14)

where nAx kg/m2 is the flux transferred in the x-direction

Similarly the mass of A entering the y and z direction are respectively:

(nAy)|y ∆x∆z∆t (3.15)

(nAz)|z ∆x ∆y ∆t (3.16)

Mass of A out:

The mass of species A exiting the x, y and z direction are respectively

(nAx)|x+∆x ∆y∆z∆t (3.17)

(nAy)|y+∆y ∆x∆z∆t (3.18)

Page 111: Modeling Notes Complete

108

(nAz)|z+∆z ∆x ∆y ∆t (3.19)

The rate of accumulation is:

ρA|t+∆t ∆x ∆y ∆z − ρA|t ∆x ∆y ∆z (3.20)

The rate of generation is:

-rA∆x ∆y ∆z∆t (3.21)

Applying the general balance equation (Eq. 1.3) yields:

(ρA|t+∆t − ρA|t) ∆x ∆y ∆z = (nAx|x+∆x − nAx|x)∆y∆z∆t + (nAy|y+∆y − nAy|y)∆x∆z∆t

+ (nAz|z+∆z − nAz|z)∆x∆y∆t +rA∆x ∆y ∆z∆t (3.22)

Dividing each term by ∆x∆y∆z∆t and letting each of these terms goes to zero yields:

AAzAyAxA rt

nt

nt

nt

=∂

∂+

∂+

∂∂

+∂ρ∂

(3.23)

We know from Section 1.11.1, that the flux nA is the sum of a term due to convection (ρAv)

and a term due to diffusion jA (kg/m2s):

nA = ρAv + jA (3.24)

Substituting the different flux in Eq. 3.23 gives:

AAzAyAxzAyAxAA rz

jy

jx

jzv

yv

xv

t=

∂∂

+∂

∂+

∂∂

+∂

∂+

∂+

∂∂

+∂

∂ )()()( ρρρρ

(3.25)

For a binary mixture (A,B), Fick’s law gives the flux in the u-direction as :

uwDj A

ABAu ∂∂

ρ−= (3.26)

Page 112: Modeling Notes Complete

109

where wA = ρA/ρ. Expanding Eq. 3.25 and substituting for the fluxes yield:

AAABAABAAB

Az

Ay

Ax

zyxA

A

rz

wDzy

wDyx

wDx

zv

yv

xv

zv

yv

xv

t

=⎟⎟⎠

⎞⎜⎜⎝

⎛∂

ρ∂∂∂

+∂

ρ∂∂∂

+∂

ρ∂∂∂

⎟⎟⎠

⎞⎜⎜⎝

⎛∂ρ∂

+∂ρ∂

+∂ρ∂

+⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂

∂+

∂∂

ρ+∂ρ∂

)()()(

(3.27)

This is the general component balance or equation of continuity for species A. This equation

can be further reduced according to the nature of properties of the fluid involved. If the binary

mixture is a dilute liquid and can be considered incompressible, then density ρ and diffusivity

DAB are constant. Substituting the continuity equation (Eq. 3.13) in the last equation gives:

AAAA

ABA

zA

yA

xA r

zyxD

zv

yv

xv

t=⎟

⎟⎠

⎞⎜⎜⎝

∂ρ∂

+∂

ρ∂+

∂ρ∂

−⎟⎟⎠

⎞⎜⎜⎝

⎛∂ρ∂

+∂ρ∂

+∂ρ∂

+∂ρ∂ )2

2

2

2

2

2

(3.28)

This equation can also be written in molar units by dividing it by the molecular weight MA to

yield:

{reaction

A

Diffusion

AAAAB

Convection

Az

Ay

Ax

onaccumulati

A RzC

yC

xC

Dz

Cv

yC

vx

Cv

tC

=⎟⎟⎠

⎞⎜⎜⎝

∂+

∂+

∂−⎟⎟

⎞⎜⎜⎝

⎛∂

∂+

∂∂

+∂

∂+

∂∂

44444 344444 21444444 3444444 213212

2

2

2

2

2

(3.29)

The component balance equation is composed then of a transient term, a convective term, a

diffusive term and a reaction term.

3.3 Momentum Balance

We consider a fluid flowing with a velocity v(t,x,y,z) in the cube of Figure 3.3. The

flow is assumed laminar. We know from Section 1.11.2 that the momentum is transferred

through convection (bulk flow) and by molecular transfer (velocity gradient).

Page 113: Modeling Notes Complete

110

xx |x+ xxx |x

zx |z+ zyx |y

yx |y+ y zx |z

y

x

z

Figure 0-3 Balance of the x-component of the momentum

Since, unlike the mass or the energy, the momentum is a vector that has three components,

we will present the derivation of the equation for the conservation of the x-component of the

momentum. The balance equations for the y-component and the z-component are obtained in

a similar way. To establish the momentum balance for its x-component we need to consider

its transfer in the x-direction, y-direction, and z-direction. Momentum in:

The x-component of momentum entering the boundary at x-direction, by convection

is:

(ρvxvx)|x ∆y∆z∆t (3.30)

The x-component of momentum entering the boundary at y-direction, by convection

is:

(ρvyvx)|y ∆x∆z∆t (3.31)

and it enters the z-direction by convection with a momentum:

(ρvzvx)|z ∆x∆y∆t (3.32)

Page 114: Modeling Notes Complete

111

The x-component of momentum entering the boundary at x-direction, by molecular

diffusion is:

(τxx)|x ∆y∆z∆t (3.33)

The x-component of momentum entering the boundary at y-direction, by molecular

diffusion is:

(τyx)|y ∆x∆z∆t (3.34)

and it enters the z-direction by molecular diffusion with a momentum:

(τzx)|z ∆x∆y∆t (3.35)

Momentum out:

The rate of momentum leaving the boundary at x+∆x, by convection is:

(ρvxvx)|x+∆x ∆y∆z∆t (3.36)

and at boundary y+∆y,:

(ρvyvx)|y+∆y ∆x∆z∆t (3.37)

and at boundary z+∆z:

(ρvzvx)|z+∆z ∆x∆y∆t (3.38)

The x-component of momentum exiting the boundary x+∆x, by molecular diffusion is:

(τxx)|x+∆x ∆y∆z∆t (3.39)

and at boundary y+∆y:

Page 115: Modeling Notes Complete

112

(τyx)|y+∆y ∆x∆z∆t (3.40)

and at boundary z+∆z::

(τzx)|z+∆z ∆x∆y∆t (3.41)

Forces acting on the volume:

The net fluid pressure force acting on the volume element in the x-direction is:

(P|x – P|x+∆x) ∆y∆z∆t (3.42)

The net gravitational force in the x-direction is:

ρg|x ∆x ∆y∆z ∆t (3.43)

Accumulation is:

(ρvx|t+∆t − ρvx|t) ∆x ∆y∆z (3.44)

Substituting all these equations in Eq. 1.5, dividing by ∆x ∆y ∆z∆t and taking the limit of each

term goes zero gives:

xzxyxxxzxyxxxx g

xP

zyxzvv

yvv

xvv

tv ρτττρρρρ

+∂∂

−∂

∂+

∂∂

+∂

∂−=

∂∂

+∂

∂+

∂∂

+∂

∂ )()()()()(

(3.45)

Expanding the partial derivative and rearranging:

xgxP

zyx

zv

vy

vv

xv

vt

vzv

yv

xv

tv

zxyxxx

xz

xy

xx

xzyxx

ρτττ

ρρρρρ

+∂∂

−∂

∂+

∂+

∂∂

=⎟⎟⎠

⎞⎜⎜⎝

⎛∂

∂+

∂∂

+∂

∂+

∂∂

+⎟⎟⎠

⎞⎜⎜⎝

∂∂

+∂

∂+

∂∂

+∂∂

)(

(3.46)

Page 116: Modeling Notes Complete

113

Using the equation of continuity (Eq. 3.10) for incompressible fluid, Equation (3.46) is

reduced to:

xzxyxxxx

zx

yx

xx g

xP

zyxzvv

yvv

xvv

tv ρτττρ +

∂∂

−∂

∂+

∂∂

+∂

∂−=⎟

⎞⎜⎝

⎛∂∂

+∂∂

+∂∂

+∂

∂ )( (3.47)

Using the assumption of Newtonian fluid, i.e.

zv

yv

xv x

zxx

yxx

xx ∂∂

−=∂∂

−=∂∂

−= µτµτµτ ,,

(3.48)

Equation 3.47 yields:

434214444 34444 2144444 344444 21321generationforces by viscoustransport

2

2

2

2

2

2

flowbulk by transport onaccumulati

xxxxx

zx

yx

xx g

xP

zv

yv

xv

zvv

yvv

xvv

tv ρµρρ +

∂∂

−⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

=⎟⎠

⎞⎜⎝

⎛∂∂

+∂∂

+∂∂

+∂

(3.49)

The momentum balances in the y-direction and z-direction can be obtained in a similar

fashion:

434214444 34444 2144444 344444 21321generationforces by viscoustransport

2

2

2

2

2

2

flowbulk by transport onaccumulati

yyyyy

zy

yy

xy g

yP

zv

yv

xv

zv

vyv

vxv

vt

vρµρρ +

∂∂

−⎟⎟⎠

⎞⎜⎜⎝

∂∂

+∂∂

+∂∂

=⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

+∂

(3.50)

434214444 34444 2144444 344444 21321generationforces by viscoustransport

2

2

2

2

2

2

flowbulk by transport onaccumulati

zzzzz

zz

yz

xz g

zP

zv

yv

xv

zvv

yvv

xvv

tv ρµρρ +

∂∂

−⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

=⎟⎠

⎞⎜⎝

⎛∂∂

+∂∂

+∂∂

+∂∂

(3.51)

These equations constitute the Navier-Stock’s equation.

3.4 Energy balance

Page 117: Modeling Notes Complete

114

In deriving the equation for energy balance we will be guided by the analogy that

exists between mass and energy transport mentioned in Section 1.11.3 We will assume

constant density, heat capacity and thermal conductivity for the incompressible fluid. The

fluid is assumed at constant pressure (Fig 3.4).

The total energy flux is the sum of heat flux and bulk flux:

e = q + ρCpTv (3.52)

Therefore, the energy coming by convection in the x-direction at boundary x is:

(qx + ρCpTvx)∆y∆z∆t (3.53)

Similarly the energy entering the y and z directions are

(qy + ρCpTvy)∆x∆z∆t (3.54)

(qz + ρCpTvz) ∆x ∆y ∆t (3.55)

The energy leaving the x,y and z directions are:

(qx + ρCpTvx)|x+∆x ∆y∆z∆t (3.56)

(qy + ρCpTvy)|y+∆y ∆x∆z∆t (3.57)

(qz + ρCpTvz)|z+∆z ∆x ∆y ∆t (3.58)

The energy accumulated is approximated by:

(ρCpT|t+∆t − ρCpT|t) ∆x ∆y∆z (3.59)

Page 118: Modeling Notes Complete

115

y

x

z

qz |z

qz |z+ z

qx |x+ x

qy |y+ y

qx |x

qy |y

Figure 0-4 Energy Balance in Cartesian coordinates

The rate of generation is ΦH where ΦH includes all the sources of heat generation, i.e.

reaction, pressure forces, gravity forces, fluid friction, etc. Substituting all these terms in the

general energy equation (Eq. 1.7) and dividing the equation by the term ∆x∆y∆z∆t and letting

each of these terms approach zero yield:

Hzyxzyx

zq

yq

xq

zCpTv

yCpTv

xCpTv

tCpT

Φ=∂∂

+∂∂

+∂∂

+∂

∂+

∂∂

+∂

∂+

∂∂ )()()()( ρρρρ

(3.60)

Expanding the partial derivative yields:

Hzyxzyx

zyx zq

yq

xq

zv

yv

xv

tCpT

zTv

yTv

xTv

tTCp Φ=

∂∂

+∂

∂+

∂∂

+⎟⎟⎠

⎞⎜⎜⎝

⎛∂

∂+

∂+

∂∂

+∂∂

+⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

+∂∂ ρρρρρρ

(3.61)

Using the equation of continuity (Eq. 3.10) for incompressible fluids the equation is reduced

to:

Hzyx

zyx zq

yq

xq

zTv

yTv

xTv

tTCp Φ=

∂∂

+∂∂

+∂∂

+⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

+∂∂

ρ (3.62)

Using Fourier's law:

Page 119: Modeling Notes Complete

116

dudTkqu −=

(3.63)

into the last equation gives:

{generation

Hzyx

onaccumulati

zT

yT

xTk

zTv

yTv

xTvCp

tTCp Φ+⎟⎟

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

=⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

ρ+∂∂

ρ444 3444 2144444 344444 2143421

diffusion by thermalTransport

2

2

2

2

2

2

flowbulk by Transport

(3.64)

The energy balance includes as before a transient term, a convection term, a diffusion term,

and generation term. For solids, the density is constant and with no velocity, i.e. v = 0, the

equation is reduced to:

{generation

H

onaccumulati

zT

yT

xTk

tTCp Φ+⎟⎟

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

=∂∂

ρ444 3444 2143421

diffusion by thermalTransport

2

2

2

2

2

2

(3.65)

3.5 Conversion between the coordinates

So far we have shown how to derive the equation of change in Cartesian coordinates.

In the same way, the equations of change can be written other coordinate systems such as the

cylindrical or spherical coordinates. Alternatively, one can transform the equation of change

written in the Cartesian coordinates to the others through the following transformation

expressions.

The relations between Cartesian coordinates (x,y,z) and cylindrical coordinates (r,z,θ)

(Figures 1.3 and 1.4) are the following:

x = rcos(θ), y = rsin(θ), z = z (3.66)

Page 120: Modeling Notes Complete

117

Therefore;

)(tan, 122

xyyxr −=θ+=

(3.67)

The relations between Cartesian coordinates (x,y,z) and spherical coordinates (r,θ,φ) (Figures

1.3 and 1.5) are:

)cos(),sin()sin(),cos()sin( θ=φθ=φθ= rzryrx (3.68)

Therefore;

)(tan),(tan, 122

1222

xy

zyx

zyxr −− =φ+

=θ++= (3.69)

Accordingly, we list the following general balance equations in the three coordinates. These

equations are written under the assumptions mentioned in previous sections. For the more

general case, where density is considered variable the reader can consult the books listed in

the references.

3.5.1 Balance Equations in Cartesian Coordinates

Mass Balance

0)( =∂∂

+∂∂

+∂∂

ρzv

yv

xv zyx

(3.70)

Component balance for component A in binary mixture with chemical reaction rate RA:

AAAA

ABA

zA

yA

xA R

zC

yC

xCD

zCv

yCv

xCv

tC

+⎟⎟⎠

⎞⎜⎜⎝

⎛∂

∂+

∂∂

+∂

∂=⎟⎟

⎞⎜⎜⎝

⎛∂

∂+

∂∂

+∂

∂+

∂∂ )2

2

2

2

2

2

(3.71)

Page 121: Modeling Notes Complete

118

Energy balance

Hzyx zT

yT

xTk

zTv

yTv

xTvCp

tTCp Φ+

∂∂

+∂∂

+∂∂

=∂∂

+∂∂

+∂∂

+∂∂ )()( 2

2

2

2

2

2

ρρ

(3.72)

Momentum balance

• x component

xxxxx

zx

yx

xx g

xP

zv

yv

xv

zvv

yvv

xvv

tv

ρ+∂∂

−∂∂

+∂∂

+∂∂

µ=∂∂

+∂∂

+∂∂

ρ+∂

∂ρ )()( 2

2

2

2

2

2

(3.73)

• y component:

yyyyy

zy

yy

xy g

yP

zv

yv

xv

zv

vyv

vxv

vt

vρ+

∂∂

−∂∂

+∂∂

+∂∂

µ=∂∂

+∂∂

+∂∂

ρ+∂

∂ρ )()( 2

2

2

2

2

2

(3.74)

• z component

zzzzz

zz

yz

xz g

zP

zv

yv

xv

zvv

yvv

xvv

tv

ρ+∂∂

−∂∂

+∂∂

+∂∂

µ=∂∂

+∂∂

+∂∂

ρ+∂∂

ρ )()( 2

2

2

2

2

2

(3.75)

3.5.2 Balance Equations in Cylindrical Coordinates

Mass balance

0)11( =∂∂

+θ∂

∂+

∂∂

ρ θ

zvv

rrrv

rzr

(3.76)

Component balance for component A in binary mixture (A-B)with reaction rate RA:

Page 122: Modeling Notes Complete

119

AAAA

ABA

zAA

rA R

zCC

rrC

rrr

Dz

Cv

Cr

vr

Cv

tC

+⎟⎟⎠

⎞⎜⎜⎝

∂+

∂+

∂∂

∂∂

=⎟⎠

⎞⎜⎝

⎛∂

∂+

∂∂

+∂

∂+

∂∂

)1)(112

2

2

2

2 θθθ

(3.77)

Energy balance

Hzr zTT

rrTr

rrk

zTvT

rv

rTvCp

tTCp Φ+

∂∂

+∂∂

+∂∂

∂∂

=∂∂

+∂∂

+∂∂

+∂∂ )1)(1()1( 2

2

2

2

2 θθρρ θ

(3.78)

Momentum balance

• r component

rrrr

rz

rrr

r

grP

zvv

rv

rrrv

rr

zv

vr

vvr

vr

vv

tv

ρθθ

µ

θρρ

θ

θθ

+∂∂

−∂

∂+

∂∂

−∂

∂+

∂∂

∂∂

=∂

∂+−

∂∂

+∂

∂+

∂∂

)21)1((

)(

2

2

22

2

2

2

(3.79)

• θ component:

θθθθ

θθθθθθ

ρθθθ

µ

θρρ

gPzvv

rv

rrrv

rr

zvv

rvvv

rv

rvv

tv

r

zr

r

+∂∂

−∂∂

+∂∂

+∂∂

+∂

∂∂∂

=∂∂

++∂∂

+∂∂

+∂

)21)1((

)(

2

2

22

2

2

(3.80)

• z component

zzzzz

zzz

rz g

zP

zvv

rrvr

rrzvvv

rv

rvv

tv

ρθ

µθ

ρρ θ +∂∂

−∂∂

+∂∂

+∂∂

∂∂

=∂∂

+∂∂

+∂∂

+∂

∂ )1)(1()( 2

2

2

2

2 (3.81)

3.5.3 Balance Equations in Spherical Coordinates

Mass Balance

Page 123: Modeling Notes Complete

120

0)sin(

1))sin(()sin(

1)(1 2

2 =⎟⎟⎠

⎞⎜⎜⎝

⎛φ∂

θ+

θ∂θ∂

θ+

∂∂

ρ φθv

rv

rrvr

rr

(3.82)

Component balance for component A in binary mixture (A-B) with reaction rate RA :

AAAA

AB

AAAr

A

RCr

Crr

Crrr

D

Cr

vCr

vr

Cvt

C

+⎟⎟⎠

⎞⎜⎜⎝

∂∂

+∂

∂∂∂

+∂

∂∂∂

=⎟⎟⎠

⎞⎜⎜⎝

⎛∂

∂+

∂∂

+∂

∂+

∂∂

))(sin

1))(sin()sin(

1)(1

)sin(

2

2

2222

2 φθθθ

θθ

φθθφθ

(3.83)

Energy balance

H

r

Tr

Trr

Trrr

k

Tr

vTr

vrTvCp

tTCp

Φ+⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

∂∂

+∂∂

∂∂

=⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

+∂∂

+∂∂

2

2

2222

2 )(sin1))(sin(

)sin(1)(1

)sin(

φθθθ

θθ

φθθρρ φθ

(3.84)

Momentum balance

• r component

rrr

r

rrrr

r

grPv

rv

rvr

rr

rvvv

rvv

rv

rvv

tv

ρφθθ

θθθ

µ

φθθρρ φθφθ

+∂∂

−⎟⎟⎠

⎞⎜⎜⎝

∂∂

+∂∂

∂∂

+∂∂

=⎟⎟

⎜⎜

⎛ +−

∂∂

+∂∂

+∂∂

+∂

2

2

2222

2

2

2

22

)(sin1))(sin(

)sin(1)(1

)sin(

(3.85)

• θ component:

θφ

θθθ

φθθφθθθθ

ρθφθ

θθ

µ

φθθθ

θθ

µ

θφθθ

ρρ

gPr

vvr

vr

vrr

vrrr

rv

rvvv

rvv

rv

rvv

tv

r

rr

+∂∂

−⎟⎟⎠

⎞⎜⎜⎝

∂−

∂∂

+

⎟⎟⎠

⎞⎜⎜⎝

∂∂

+∂

∂∂∂

+∂∂

∂∂

=⎟⎟

⎜⎜

⎛−+

∂∂

+∂∂

+∂∂

+∂

1)(sin)cos(2

)(sin1))sin()(sin(1)(1

)cot()sin(

22

2

2

2222

2

2

(3.86)

Page 124: Modeling Notes Complete

121

• Φ component

φθ

φφφ

φθφφφφθφφ

ρφθφθ

θφθ

µ

φθθθ

θµ

θφθθ

ρρ

gPr

vr

vr

v

r

vθrr

vr

rr

rvv

rvvv

rvv

rv

rv

vt

v

r

rr

+∂∂

−⎟⎟⎠

⎞⎜⎜⎝

∂∂

+∂∂

+

⎟⎟

⎜⎜

∂+

∂∂

+∂

∂∂

=⎟⎟

⎜⎜

⎛++

∂+

∂+

∂+

)sin(1

)(sin)cos(2

)sin(2

)(sin1)

)sin()sin(

1(1)(1

)cot()sin(

222

2

2

2222

2

(3.87)

3.6 Examples of Application of Equations of change

Practically all the microscopic balance examples treated in the previous chapter can be

treated using the equations of change presented in this chapter. In this section we review some

of the previous examples and present additional applications.

3.6.1 Liquid flow in a Pipe

To model the one dimensional flow through the pipe of an incompressible fluid (Example

2.2.1) we may use the continuity balance. For constant density we have the continuity

equation in cylindrical coordinates (Eq. 3.76).

01=

∂∂

+θ∂

∂+

∂∂ θ

zv

rv

rrv

rzr

(3.88)

The plug flow assumptions imply that vr = vθ = 0, and the continuity balance is reduced to:

0=∂∂

zvz

(3.89)

3.6.2 Diffusion with Chemical Reaction in a Slab Catalyst

Page 125: Modeling Notes Complete

122

To model the steady state diffusion with chemical reaction of species A in a slab

catalyst, (Example 2.2.3) we use the equation of change (Eq. 3.71). The fluid properties are

assumed constant,

AAAA

AA

zA

yA

xA R

zC

yC

xCD

zCv

yCv

xCv

tC

=∂

∂+

∂∂

+∂

∂−

∂∂

+∂

∂+

∂∂

+∂

∂ )( 2

2

2

2

2

2

(3.90)

Since the system is at steady state we have 0=∂

∂t

CA . If we assume that there is no bulk flow

then vx = vy = vz = 0. For diffusion in the z-direction only, the following holds: =∂∂x

0=∂∂y

.

The equation is then reduces to:

AA

A Rdz

CdD =− 2

2

(3.91)

3.6.3 Plug Flow Reactor

The isothermal plug flow reactor (Example 2.2.5) can be modeled using the

component balance equation (3.77). The plug flow conditions imply that vr = vθ = 0 and

=∂∂r

0=θ∂∂ . Equation (3.77) is reduced to:

AA

ABA

zA R

zCD

zCv

tC

−∂

∂+

∂∂

−=∂

∂2

2

(3.92)

For the non-isothermal plug flow reactor (Example 2.2.6), the energy balance is obtained by

using Eq. 3.78. For fluid with constant properties and at constant pressure, we have:

Hzr zTT

rrT

rrTk

zTvT

rv

rTvCp

tTCp Φ+

∂∂

+∂∂

+∂∂

+∂∂

=∂∂

+∂∂

+∂∂

+∂∂ )11()( 2

2

2

2

22

2

θθρρ θ

(3.93)

Page 126: Modeling Notes Complete

123

Using the plug-flow assumptions, vr = vθ = 0 and =∂∂r

0=θ∂∂ , and neglecting the viscous

forces, the term ΦH includes the heat generation by reaction rate RA and heat exchanged with

the cooling jacket, htA(T –Tw). Equation (3.93) is reduced to:

)(/2

2

wtARTE

or TTADhCekH

zTk

zTCpv

tTCp −

π−−

∂∂

+∂∂

ρ−=∂∂

ρ −∆

(3.94)

3.6.4 Energy Transport with Heat Generation

Consider the example of a solid cylinder of radius R in which heat is being generated

due to some reaction at a uniform rate of ΦH (J/m2s). A cooling system is used to remove heat

from the system and maintain its surface temperature at the constant value Tw (Figure 3.5).

Our objective is to derive the temperature variations in the cylinder. We assume that the solid

is of constant density, thermal conductivity and heat capacity. Clearly this is a distributed

parameter system since the temperature can vary with time and with all positions in the

cylinder. We will use then the equation of change (Eq. 3.78) in cylindrical coordinates for a

solid:

CpzTT

rrT

rrT

Cpk

tT H

ρ+

∂∂

+θ∂

∂+

∂∂

+∂∂

ρ=

∂∂ Φ)11( 2

2

2

2

22

2

(3.95)

Figure 0-5 Cylindrical solid rod

A number of assumptions can be made:

Page 127: Modeling Notes Complete

124

• The system is at steady state i.e. 0=∂∂

tT

• The variation of temperature is only allowed in radial directions. Therefore, the terms

2

2

zT

∂∂ and 2

2

θ∂∂ T are zero.

The energy balance is reduced to:

CpdrdT

rdrTd

Cpk H

ρρΦ

++= )1(0 2

2

(3.96)

Or equivalently:

kdrdT

rdrTd HΦ

−=+ )1( 2

2

(3.97)

with the following boundary conditions:

• The temperature at the wall is constant:

wTRrT == )( (3.98)

• The maximum temperature will be reached at the center ( r = 0), therefore:

0 0 == ratdrdT

(3.99)

Note that Equation (3.97) can also be written as follows:

kdrdTr

drd

rHΦ

−=)(1

(3.100)

since qr = −k dT/dr, this equation is equivalent to:

Page 128: Modeling Notes Complete

125

Hrrqdrd

rΦ−=)(1

(3.101)

The left hand side is the rate of diffusion of heat per unit volume while the right hand side is

the rate of heat production per unit volume.

3.6.5 Momentum Transport in a Circular Tube

We revisit example 2.2.2 where we derived the steady state equations for the laminar

flow inside a horizontal circular tube. We will see how the model can be obtained using the

momentum equation of change. We assume as previously that the fluid is incompressible and

Newtonian. The momentum equations of change in cylindrical coordinates are given by Eqs.

(3.79-3.81). A number of simplifications are used:

• The flow has only the direction z, i.e. vr = vθ = 0.

• The flow is at steady state, 0=∂

∂t

vz

The momentum equation in cylindrical coordinates Eq. 3.81 is reduced to:

zzzzzz

z gzP

zvv

rrv

rrv

zvv ρ+

∂∂

−⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+θ∂

∂+

∂∂

+∂∂

µ=∂∂

ρ 2

2

2

2

22

2 11

(3.102)

Using the continuity equation (Eq. 3.76), and since vr = vθ = 0 gives:

0=dzdvz

(3.103)

We also note that because the flow is symmetrical around the z-axis we have necessarily no

variation of the velocity with θ, i.e.

02

2

=⎟⎟⎠

⎞⎜⎜⎝

⎛θ∂

∂ zv

(3.104)

Equation 3.102 is then reduced to

Page 129: Modeling Notes Complete

126

dzdP

drdv

rdrvd zz =⎟⎟

⎞⎜⎜⎝

⎛+

12

2

µ

(3.105)

Since the left hand side depends only on r, this equation suggests that dzdP is constant.

Therefore:

LP

dzdp ∆

=

(3.106)

where ∆P is the pressure drop across the tube. Equation 3.105 is equivalent to:

LP

drdv

rdrvd zz ∆

=⎟⎟⎠

⎞⎜⎜⎝

⎛+

12

2

µ

(3.107)

with the following conditions identical to those in Example 2.2.2. Note also that Eq. 3.107 can

also be written as:

LP

drrv

drd z ∆

=)(µ

(3.108)

which is also equivalent to:

LP

rdrrd rz ∆

=)( τ

(3.109)

where τrz is the shear stress. The left term is the rate of momentum diffusion per unit volume

and the right hand side is in fact the rate of production of momentum (due to pressure drop).

Note then the similarity between Eq. 3.109 for momentum transfer with Eq. 3.101 for heat

transfer.

Page 130: Modeling Notes Complete

127

3.6.6 Unsteady state Heat Generation

We reconsider Example 3.6.4 but we are interested in the variations of the temperature

of the reactor with time as well. This may be needed to compute the heat transferred during

start-up or shut-down operations. Keeping the same assumptions as Example 3.3.4 (except the

steady state assumption), the energy balance in cylindrical coordinates yields:

kdrdTr

drd

rk

tTCp HΦ

+=∂∂

ρ )(1

(3.110)

with the initial and boundary conditions:

w

r

w

TrTdrdT

TtRT

=

=

=

=

)0,(

0|

),(

0

(3.111)

(3.112)

(3.113)

3.6.7 Laminar Flow Heat Transfer with Constant Wall Temperature

We consider a fluid flowing at constant velocity vz into a horizontal cylindrical tube.

The fluid enters with uniform temperature Ti. The wall is assumed at constant temperature Tw.

We would like to model the variations of the fluid temperature inside the tube. To apply the

energy equation of change (Eq. 3.78) we will assume that the fluid is incompressible,

Newtonian and of constant thermal conductivity. Since the system is at steady state d/dt = 0

and the flow is one-dimensional vr = vθ = 0, the energy equation in cylindrical coordinates

Eq. 3.78 is reduced to:

)11( 2

2

2

2

22

2

zTT

rrT

rrTk

tTCpvz ∂

∂+

θ∂∂

+∂∂

+∂∂

=∂∂

ρ

(3.114)

Page 131: Modeling Notes Complete

128

Figure 0-6 heat transfer with constant wall temperature

Since the temperature is symmetrical then θ∂

∂T = 2

2

θ∂

∂ T = 0. In some cases we can neglect the

conduction term 2

2

zT

∂ compared to the convective term zTvz ∂

∂ . The system is then described

by the following energy and momentum equations:

)1( 2

2

rT

rrTk

tTCpvz ∂

∂+

∂∂

=∂∂

ρ (3.115)

LP

drrv

drd z ∆

=)(µ

(3.116)

These two equations are therefore coupled by vz, with the previous boundary conditions for vz,

At r = R, vz = 0 (3.117)

At r = 0, dvz/dr = 0 (3.118)

At z = 0, T = Ti (3.119)

At r = 0, dT/dr = 0 (3.120)

At r = R, T = Tw (3.121)

Page 132: Modeling Notes Complete

129

3.6.8 Laminar Flow and Mass Transfer

We consider the example of a fluid flowing through a horizontal pipe with constant

velocity vz. The pipe wall is made of a solute of constant concentration CAw that dissolved in

the fluid. The concentration of the fluid at the entrance z = 0 is CAo. The regime is assumed

laminar and at steady state. The fluid properties are assumed constant. We would like to

model the variation of the concentration of A along the axis in the pipe. The component

balance for A (Eq. 3.77) is:

))(

( 2

2

23

2

zC

rC

rrr

Cr

Dz

CvCvr

Cv AA

A

ABA

zAA

r ∂∂

+θ∂

∂+

∂∂

∂∂

=∂

∂+

θ∂∂

+∂

∂θ

(3.122)

Since vr = vθ = 0, the balance equation becomes:

))(

( 2

2

zC

rrr

Cr

Dz

Cv A

A

ABA

z ∂∂

+∂∂

∂∂

=∂

(3.123)

If the diffusion term 2

2

zCA

∂ in the z-direction is negligible compared to convection term

zC

v Az ∂

∂ then the last equation reduces to:

⎟⎟⎟⎟

⎜⎜⎜⎜

∂∂

∂∂

=∂

∂rrr

Cr

Dz

Cv

A

ABA

z

)(

(3.124)

The equation (Eq. 3.116) is unchanged.

The two equations are coupled through vz. The boundary conditions are analogous to the

previous example:

Page 133: Modeling Notes Complete

130

At r = R, vz = 0 (3.125)

At r = 0, dvz/dr = 0 (3.126)

At z = 0, CA= CAo (3.127)

At r = 0, dCA/dr =0 (3.128)

At r = R, CA= CAw (3.129)

Note the similarity between this example and the heat transfer case of the previous example.

Page 134: Modeling Notes Complete

131

Problems: Lumped parameter systems

Problem 1

Consider the three storage tanks in series shown by Figure X-1.

1. Write the mathematical model that describes the dynamic behavior of the process.

Assume the density of the fluid constant.

2. Identify the states of the process and the degrees of freedom.

h1

F0

h2

F1 F2

F3h3

Fr

Figure X-1

Problem 2

Consider the three cooling storage tanks in series shown in Figure X-2.

h1

F0, T0

h2 h3

F1, T1 F3, T3

F4, T4

Ff , Tf

wc1, Tc1 wc2, Tc2 wc3, Tc3 Figure X-2

1. Assuming constant fluid properties, write the mathematical model for the process.

Identify the states and the degrees of freedom.

Page 135: Modeling Notes Complete

132

2. Assume now the following non-isothermal reversible first-order liquid-phase reaction:

A B is taking place in the three above CSTR in series. Assuming all CSTRs are

adiabatic, write down the unsteady state model for the process if the holdups is kept

constant in all reactors.

Problem 3

Consider the well-stirred non-isothermal reactor shown in Figure X-3 below. Write down the

mathematical equations that describe the dynamic behavior of the fundamental quantities of

the process. Consider an exothermic liquid-phase second order reaction rate in the form of A

B is taking place. The density is assumed to be constant, develop the necessary equations

describing the process dynamic behavior.

h1

F1, CA1

F3, CA3

Figure X-3

Problem 4

Consider a two CSTRs in series with an intermediate mixer introducing a second feed as

shown in Figure X-4. A first order irreversible exothermic reaction: A B is carrier out in

the process. Water at ambient temperature (Tc1i and Tc2i) is used to cool the reactors. The

densities and heat capacities are assumed to be constant and independent of temperature and

concentration. Develop the necessary equations describing the process dynamic behavior.

Note that the mixer has negligible dynamics and that the inlet feed to CSTR2 has the same

temperature as that of the outlet of CSTR1.

Page 136: Modeling Notes Complete

133

Mixer

Q1, C1f, T1f

QC1, TC1i

Q4, C4, T4

QC2, TC2i

Q2, C2, T2

Q3, C3, T3

CSTR 1

CSTR 2

QC1, TC1

QC2, TC2

Figure X-4

Problem 5:

Consider a thermometer bulb with pocket is immersed in hot fluid of temperature T1 as shown

in Figure X-5 below. The pocket temperature is T2 (assumed uniform through the thickness)

and that for the bulb is T3. Write the model equation that describes how T2 and T3 varies with

time.

T1

T2

T3

Figure X-5

Problem 6

Consider the single-effect steam evaporator shown in Figure X-6. A salty water (brine) with

mass fraction Cb0 and mass flow rate B0 is fed to the evaporator where it is heated with

saturated steam with mass flow rate W. The concentrated product comes out of the evaporator

Page 137: Modeling Notes Complete

134

with mass flow rate B1 and mass fraction Cb1 while the vapor is withdrawn from the top with

mass flow rate V. Develop the dynamic model that describe the process behavior. Assume the

heat supplied by the steam is mainly equal to the heat used for vaporizing the brine.

Steam

Feed

Vapor

Condensate

Concentratedliquid

Figure x-6

Problem 7

Consider the process shown in Figure X-7. A stream of pure component A is mixed with

another stream of a mixture of component A and B in an adiabatic well stirred mixing tank.

The effluent is fed into an adiabatic CSTR where the following reaction takes place: A + B

C. Assume the process is isothermal. Develop the dynamic model for the process and

determine the degrees of freedom assuming constant fluid properties.

Q1, CA1

Q3, CA3

Q2, CA2

Q4, CA4

Figure X-7

Page 138: Modeling Notes Complete

135

Problem 8:

Consider the following irreversible first-order liquid-phase reaction:

CBA kk ⎯→⎯⎯→⎯ 21

where k1 and k2 are the reaction rate constants in sec-1.

1. Write down the unsteady state model for the process if the reaction takes place in an

isothermal well-mixed CSTR.

2. Write down the unsteady state model for the process if the reaction takes place in a

non-isothermal well-stirred batch reactor, where the heat needed for the endothermic

reaction is supplied through electrical coil.

3. Repeat part 2 but assuming that the reactor is cooled by cold fluid flowing in a jacket

and that the reactor wall has high resistance to heat transfer.

Problem 9:

A perfectly mixed non-isothermal adiabatic reactor carries out a simple first-order exothermic

reaction, A B in the liquid phase. The product from the reactor is cooled from the output

temperature T to Tc and then introduced to a separation unit where the un-reacted A is

separated from the product B. The feed to the separation unit is split into two equal parts top

product and bottom product. The bottom product from the separation unit contains 95% of the

un-reacted A in the effluent of the reactor and 1% of B in the same stream. The bottom

product which is at Temperature Tc (since the separation unit is isothermal) is recycled and

mixed with the fresh feed of the reactor and the mixed stream is heated to temperature Tf

before being introduced to the reactor. Write the steady state mass and energy balances for the

whole process assuming constant physical properties and heat of reaction. Discuss also the

degree of freedom of the resulting model. The process is depicted in figure X-8

Page 139: Modeling Notes Complete

136

CA, CB , Tc

Fo, To ,CAo

F, Tf

F

L = 0.5 F

V= 0.5 FCSTR

Separator

Figure X-8

Problem 10:

Consider a biological reactor with recycle usually used for wastewater treatment as depicted

in figure X-9 below. Substrate and biomass are fed to the reactor with concentrations Sf and

Xf. The effluent form the well-mixed reactor is settled in a clarifier and a portion of the

concentrated sludge is returned to the reactor with flow rate Qr and Xr. If the reaction of

substrate in the clarifier is negligible, the recycle stream would contain the same substrate

concentration as the effluent from the reactor. Sludge is withdrawn directly from the reactor

with a fraction W. Assuming constant holdup develop the dynamic model for the process.

Assume the rate of disappearance of S is given by: r = mS/(K + S) and the rate of generation

of X by r/Y where Y is constant.

ReactorV

WXS

XS

QXfSf

Q-WXtS

Qr , Xr , S

Settler

Figure X-9

Problem 11

Page 140: Modeling Notes Complete

137

Consider the two phase reactor and a condenser with recycle. Gaseous A and liquid B enters

the reactor at flow rates FA and FB respectively. Gas A diffuses into the liquid phase where it

reacts with B producing C. The latter diffuses into the vapor phase where B is nonvolatile.

The vapor phase is fed to the condenser where the un-reacted A is cooled and the condensate

is recycled back to the reactor. The product C is withdrawn with the vapor leaving the

condenser. For the given information develop the dynamic model for the process shown in

Figure X-10. Consider all flows are in moles.

FB, TB

FA, TA

F1L, xA1, xB1, T1

F2L,xA2

F1v, yA1, yc1

T1, P1

P2

T2

NA

NA

NC

Q1

A + B = 2C

Figure X-10

Problem 12:

Develop the mathematical model for the triple-effect evaporator system shown in figure X-11

below. Assume boiling point elevations are negligible and that the effect of composition on

liquid enthalpy is neglected.

Page 141: Modeling Notes Complete

138

Vo

V1=F-L1

L1

V2=L1-L2

V1

L2

V2

L3

V3=L2-L3

Vo

Thick Liquor Thick Liquor Thick Liquor

F (feed)

Vapor Vapor Vapor

Figure X-11

Problem 13

Liquid vaporizer as depicted in Figure X-12 below is one of the important processing units in

a chemical plant. A liquid feed, enters the vaporizer at specific flow rate F0 density ρ0 and

temperature T0. Inside the vessel, the liquid is vaporized by continuous heating using hot oil.

The mass rate of vaporization is wn. The formed vapor is withdrawn continuously from the

top of the vessel. We would like to develop a mathematical model to describe the process.

Unlike the adiabatic flash operation, the temperature and pressure in the two phases are

different. Correspondingly, the volume of the two phases varies with time.

Fo , To , ρo

Fv

Q

PL, VL, ρL

Pv , Vv , ρv

wv

Figure X-12

Page 142: Modeling Notes Complete

139

Problems: Distributed Parameter Systems

Problem 14

At time t = 0 a billet of mass M, surface area A, and Temperature To is dropped into a tank of

water at Temperature Tw. Assume that the heat-transfer coefficient between the billet and

water is h. If the mass of water is large enough that its temperature is virtually constant,

determine the equation describing the variation of billet temperature with time for the

following two cases:

1. The thermal conductivity of the metal in the billet is sufficiently high that the billet is

of uniform temperature.

2. The thermal conductivity of the metal is low and the billet radius is large enough that

the temperature inside the billet is not uniform.

Problem 15

(a)Consider a wall made up of stacked layers of various materials each with different thermal

conductivity usually used for insulation as shown by figure X-13a. Assume the inner side is at

high temperature Tb and the outer side is at the ambient temperature Ta. Derive the

temperature profile through the wall.

x

Ta

Tb

Figure X-13a

(b) Repeat the above development for a composite cylindrical wall shown by figure X-13b.

Page 143: Modeling Notes Complete

140

Ta

Tb

r

Figure X-13b

Problem 16

A chemical reaction is being carried out in a fixed bed flow reactor. The reaction zone is

filled with catalyst pellets. Assume plug flow and the reactor wall is well insulated such that

the temperature is uniform in the radial direction. If the fluid enters the reactor with

temperature T1 and superficial velocity v and that thermal energy per unit volume (S) is

produced inside the reactor due to chemical reaction, derive the steady-state temperature axial

distribution. The unit is depicted in figure X-14.

z = 0 z = LReaction zone

Reactant Product

Insulating wall Catalyst particles

zr

Figure X-14

Problem 17

A first order irreversible reaction is taking place in a tubular reactor. The reactor can be

considered to be isothermal and radial dispersion is not to be neglected. Using Fick’s law with

effective diffusion coefficient to describe the axial and radial dispersion, derive the steady

state mass balance equation and its boundary conditions.

Problem 18

Page 144: Modeling Notes Complete

141

(a) Derive the unsteady state mass balance equation for the diffusion of species A through a

spherical porous pellet.

(b) Derive the steady state mass balance equation for the diffusion of species A through a

spherical porous pellet where it converts to B according to the first-order reaction: A B.

Problem 19

Species A dissolves in liquid B and diffuses into stagnant liquid phase contained in a

cylindrical tank where it undergoes a homogeneous irreversible second-order chemical

reaction: A+B C. Derive the mass balance equation for component A. Assume that the

depth of liquid is so large compared to the radius of the tank, thus, the radial distribution can

be neglected. The process is shown schematically in figure X-15.

Gas A

Liquid B∆z

z = 0

z = L

Figure X-15