You are on page 1of 80

Chemical Engineering Department

University of Technology

Mathematical
Modeling and
Numerical Analysis

Post graduate M. Sc.

Assoc. Prof. Dr. Zaidoon Mohsin shakor


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 test
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
Model simulation can be utilized in the evaluation of equipment's size and
arrangements and in the study of alternative process flow-sheeting and

1
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 controller.
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

2
system, 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).

Boundary

System
Suroundings

Figure 1.1: system and its boundary

3
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. Theoretical models are, however, generally difficult to obtain
and sometimes hard to solve.
• 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.
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.

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

4
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.'

1.6 Classification of theoretical models


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
5
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 . 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.

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.

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.
Consequently a modeler should practice careful utilization of the simplifying
assumptions based on engineering sense and experience. Failure to do so, the modeler
6
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:
• 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:

7
S

Sin
System Sout

Figure 1.2

Total flow Generation Total flow rate Accumulation amount of (S)


rate of (S) + rate of (S) = of (S) out of + rate of (S) ± exchanged (1.1)
into the within the system within system with the
system system surrounding

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)

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 :

8
Rate of Flow of Rate of
Flow of
Generation of moles of (A) Accumulation of (1.3)
moles (A) in + = +
moles of (A) out moles of (A)

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 Rate of Rate of Rate of


momentum in + Generation of = momentum + Accumulation of (1.5)
momentum out momentum

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

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,
d (mv)
= ∑F
(1.6)
dt

1.8.4 Energy balance


The energy balance for a given system is:

9
Rate of Rate of Rate of Rate of ± Amount of
energy + Generation = energy + accumulation energy (1.7)
in of energy out of energy exchanged
with the
surrounding

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.

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
study.

10
z

y
∆z
∆y

∆x

Figure 1.3: Cartesian coordinates

y
r
θ

Figure 1.4: Cylindrical coordinates

θ
r z

Figure 1.5: Spherical coordinates

11
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. On the other hand, when the modeling is based on
macroscopic scale the resulting model consists of sets of ordinary differential
equations (ODE) .
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


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 occurs 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 n Au (kg/m2s) of species A of density ρ A (kg/m3) flowing with velocity
R R P P R R P P

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


R R

n Au = j Au + ρ A v u
R R R R R R R
(1.8)

12
total flux = diffusive flux + bulk flux (1.9)

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

dw A (1.10)
j Au = − ρD AB
du

where w A = ρ A /ρ is the mass fraction of species Α, ρ (kg/m3) the density of the


R R R R P P

mixture and D AB (m2/s) is the diffusivity coefficient of A in the mixture. In molar unit
R R P P

the flux J Au (mol A/m2s) is given by:


R R P P

dx A (1.11)
J Au = −CD AB
du

where C is the total concentration of A and B (Kg(A+B)/m3) and x A = C A /C is the P P R R R R

mole fraction of A in the mixture. For constant density the flux Eqs. ( 1.10 and Eq.
1.11) become:
dρ A (1.12)
j Au = − DAB
du
dC A (1.13)
J Au = − D AB
du

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:
(ρv x )v y
R R R R (kg.m/s2)
P P
(1.14)

To determine the diffusion flux denoted by τ yx of the x-component of the momentum R R

in the y-direction, consider a fluid flowing between two infinite parallel plates as
13
shown in Figure 1.6. At a certain time the lower plate is moved by applying a
constant force F x while the upper plate is maintained constant. The force F x is called
R R R R

a shear force since it is tangential to the area A y on which it is applied (Fig. 1.6). R R

y
∆y F, force
∆ Vx

Figure 1.6: Momentum transfer between two parallel plates

The force per unit area


Fx
(kg.m / s 2 )
(1.15)
Ay

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

The force F x imparts a constant velocity V = v x (y = 0 ) to the layer adjacent to the


R R R R

plate. Because of molecular transport, the layer above it has a slightly slower velocity
v x (y) and so on as shown in Figure 1.6. Therefore, there is a transport by diffusion of
R R

the x-component of the momentum in the y-direction. The flux of this diffusive
transport is in fact the shear stress τyx . R R

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

convection term (Eq. 1.14) and diffusive term τ yx R

π yx = τ yx + (ρv x )v y
R R R R R R R
(1.16)
momentum flux = diffusive flux + bulk flux (1.17)

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

∂v x (1.18)
τ yx = − µ
∂y

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


14
1.11.3 Energy transport
The total energy flux e u (J/s.m2) of a fluid at constant pressure flowing with a
R R P P

velocity v u in the u-direction can be expressed as:


R R

e u = q u + (ρC p T)v u
R R R R R R R
(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:
∂T (1.21)
qu = −k
∂u

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 Law Flux Transport gradient
Type property
dC A
Mass Fick's J Au
R D
du
dT
Heat Fourrier qu R K
du
dv x
Momentum Newton τ ux R µ
du

15
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:
J A = K × ∆C A (1.23)
where K an overall mass transfer coefficient.

The heat flux on the other hand is expressed as


q = U × ∆T (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.

f =
D
∆P
(1.25)
2v 2 Lρ

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


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
16
ρ = M w P/RT R R
(1.27)
where M w is the molecular weight. As for liquids, tabulated values of density
R R

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:
~
h = C p (T − Tref ) (1.28)
~
H = C p (T − Tref ) + λ (1.29)
~
Where h is the liquid specific enthalpy, H~ is its vapor specific enthalpy, C p

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 T ref . If


R R

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 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 C p can be
R R

taken as function of temperature such as:


C p = a + bT + cT2 + dT3
R R P P P (1.30)
The specific enthalpies for vapors and liquids are computed as integrals as
follows:
~
T (1.31)
h= ∫ C p dT
Tref

~
T (1.32)
H= ∫ C p dT + λ
Tref

• Internal energy :

17
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 the bulk gas phase y AG decreases at the interface. The liquid
R R

concentration increases, on the other hand, from x AL to x Ai . At the interface, an


R R R R

equilibrium exists and y Ai and x Ai are related by a relation of the form of


R R R R

y Ai = F(x Ai )
R R R R (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 mixture Liquid -phase solution


of A in gas G of A in liquid L

yAG yAi
xAi
xAL
NA

Interface

Figure 1.7: Equlibrium at the interface

18
• For vapor phases at low concentration, Henry's law provides the following
simple equilibrium relation:
p A = Hx A
R R R (1.34)
Where p A (atm) is the partial pressure of species A in the vapor, x A is the mole
R R R R

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
R R R
(1.35)
~
The constant ( H ) depends on temperature and pressure.

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


mixtures
y A P = x A PAS (1.36)

Where x A is the liquid-phase mole fraction, y A is the vapor-phase mole


fraction, P A s is the vapor pressure of pure A at the temperature of the system,
R RP P

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

The dependence of vapor pressure P A s on temperature can be approximated


R RP P

by the Antoine equation

ln( PAS ) = A −
B (1.37)
C +T
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 γ i and φi of component (i) for liquid R R R R

and vapor phase respectively. The following equilibrium relation, also known
as the Gamma/Phi formulation, can be used:
y iφi P = xi γ i Pi S (1.38)

19
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. ( φi = γ i =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:
y A = Kx A
R R R (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 (x i ) and either T or P, the bubble point
R R

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


R R

T.
• Dew point calculations :
For a given molar vapor compositions (y i ) and either T or P, the dew point
R R

calculations consist in finding the molar liquid compositions (x i ) and either P


R R

or T.
• Flash calculation :
For known mixture compositions (z i ) and known (T) and (P), the flash
R R

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:
P P

R=
1 dni (1.40)
viV dt

20
where
ni
R R the number of moles
νi
R R 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


ν r1 R 1 + ν r2 R 2 + …+ ν rr R rr  ν p1 P 1 + ν p2 P 2 + …+ ν pp R pp
R R R R R R R R R R R R R R R R R R R R R R R
(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.
R = kCRζ1 C Rζ 2 C Rζ r (1.42)
1 2 r

The powers ξ i are determined experimentally and their values are not necessarily
R R

integers. The temperature dependence comes from the reaction rate constant k given
by Arrhenius law
−E (1.43)
k = ko e RT

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


R R

temperature and R is the ideal gas constant

1.15 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 N e independent equations and N v R R R R

variables, the degree of freedom f is


F = Nv − Ne R R R (1.44)

21
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.
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.

1.16 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, which 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.

22
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
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.

23
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. 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.

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 F f (m3/s) and density ρ f flow into the tank. The outlet stream has
R R P P R R

volumetric rate F o and density ρ ο . Our objective is to develop a model for the
R R R R

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
24
isothermal, i.e. no variations in the temperature. To model the tank we need only to
write a mass balance equation.

Ff
ρf V
Fo
ρo

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 FfR R R
(2.1)
Mass flow out:
ρo Fo
R R R
(2.2)
Accumulation:
dm d (ρV )
=
dt dt (2.3)
The generation term is zero since the mass is conserved. The balance equation yields:
d (ρV )
ρ f F f = ρo Fo +
dt (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 t i , i.e.
R R

25
V(t i ) = V i
R R R (2.5)
Under isothermal conditions we can further assume that the density of the liquid is
constant i.e. ρ f = ρ o =ρ. In this case Eq. 2.4 is reduced to:
R R R R

dV
= F f − Fo
dt (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)
Since (A) is constant then we obtain the equation in terms of the state variable L:
dL
A = F f − Fo
dt (2.8)
with initial condition:
L(t i ) = L i
R R R (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): F f R

• Remaining variables: L and F o R

• 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 F o is a function of
R R

the liquid level L. Generally a relation of the form:


Fo = α L (2.10)

Could be used, where α is the discharge coefficient.

26
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
R R R R (2.11)

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 F f (m3/s), density ρ f (kg/m3) and
R R P P R R P P

temperature T f (K). It is heated with an external heat supply of temperature T st (K),


R R R R

assumed constant. The effluent stream is of flow rate F o (m3/s), density ρ o (kg/m3) R R P P R R P P

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.
Consequently, the flow of energy into the system is:
~
ρf Ff h f R R R R
(2.12)

Where the ( ~• ) denotes the specific enthalpy (J/kg).


Ff , Tf , ρf

Q
L

Fo , To , ρo

Tst
Heat supply

Figure 2.2 Stirred Tank Heater

27
Figure 2.3 General Macroscopic System

The flow of energy out of the system is:


~
ρ o F o ho
R R R R
(2.13)

The rate of accumulation of energy is:


(
d ρVh
~
) (2.14)
dt
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 Q r (J/s), heat exchanged with surroundings Q e (J/s).
R R R R

Substituting all these terms in the general balance equation (Eq. 1.7) yields:
(
d ρVh
~
) ~ ~
= ρ f F f h f − ρ o Fo ho + Qe + Qr (2.15)
dt
We can check that all terms of this equation have the SI unit of (J/s). We return now
to the liquid stirred tank heater. A simplifying assumption can be introduced, that
there is no reaction involved, i.e. Q r = 0.R R

The energy balance (Eq. 2.15) is reduced to:


d ρVh
~
( ) ~ ~
= ρ f F f h f − ρ o Fo ho + Qe (2.16)
dt
Here Q e is the heat (J/s) supplied by the external source. The enthalpy is generally a
R R

function of temperature, pressure and composition. However, it can be safely


estimated from heat capacity relations as follows:

28
~ ~
h = Cp(T − Tref ) (2.17)
~
Where Cp is the average heat capacity.

Furthermore since the tank is well mixed the effluent temperature T o is equal to R R

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


constant density ρ f = ρ o = ρ , as follows:
R R R R

ρC p
(
~ d V (T − Tref ) ~ ) ~
= ρF f C p (T f − Tref ) − ρFoC p (T − Tref ) + Qe (2.18)
dt
Taking T ref = 0 for simplicity and since V = AL result in:
R R

~ d (LT ) ~ ~
ρC p A = ρF f C pT f − ρFoC pT + Qe (2.19)
dt
or equivalently:
d (LT ) Q
A = F f T f − FoT + ~e (2.20)
dt ρC p

Since
d (LT ) d (L ) d (T )
A = AT + AL (2.21)
dt dt dt
And using the mass balance (Eq. 2.8) we get:
dT Q
AL + T ( F f − Fo ) = F f T f − FoT + ~e (2.22)
dt ρC p

or equivalently:
dT Q
AL = F f (T f − T ) + ~e (2.23)
dt ρC p

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

A
dL
= F f − Fo
(2.24)
dt

AL
dT Q
= F f (T f − T ) + ~e
(2.25)
dt ρC p

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


initial time are known,
L(t i ) = L i
R R R R and T(t i ) = T i
R R R (2.26)

29
Degree of freedoms analysis
For this system we can make the following simple analysis:
• Parameter of constant values: A, ρ and C p R

• (Forced variable): F f and T fR R R

• Remaining variables: L, F o , T, Q e R R R

• Number of equations: 2 (Eq. 2.24 and Eq. 2.25)

The degree of freedom is therefore, 4− 2 = 2. We s till need two relations for our
problem to be exactly specified. Similarly to the previous example, if the system is
operated without control then F o is related to L through (Eq. 2.10). One additional
R R

relation is obtained from the heat transfer relation that specifies the amount of heat
supplied:
Q e = UA H (T st −T )
R R R R R R (2.27)
U and A H are heat transfer coefficient and heat transfer area. The source temperature
R R

T st was assumed to be known.


R R

2.1.3 Isothermal CSTR

We revisit the perfectly mixed tank of the first example but where a liquid phase
chemical reactions taking place:
k
A
→ B (2.28)
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 F f (m3/s), density ρ f (kg/m3) and R R P P R R P P

concentration C Af (mole/m3). The output comes out of the reactor at volumetric rate
R R P P

F o , density ρ 0 and concentration C Ao (mole/m3) and C Bo (mole/m3). We assume


R R R R R R P P R R P P

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

30
Ff
ρf
CAf V Fo
CBf ρo
CAo
CBo
Figure 2.4 Isothermal CSTR

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


the number of moles (n A = C A V ). Since the system is well mixed the effluent
R R R R

concentration C Ao and C Bo are equal to the process concentration C A and C B .


R R R R R R R R

Flow of moles of A in:


F f C AfR R R (2.29)
Flow of moles of A out:
F o C Ao
R R R (2.30)
Rate of accumulation:
dn d (VC A )
= (2.31)
dt dt
Rate of generation: -rV
where r (moles/m3s) is the rate of reaction.
P P

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


d (VC A )
= F f C Af − Fo C A − rV (2.32)
dt
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.

31
Using the differential principles, equation (2.32) can be written as follows:
d (VC A )
=V
d (C A )
+ CA
d (V )
= F f C Af − Fo C A − rV
(2.33)
dt dt dt
Substituting Equation (2.6) into (2.33) and with some algebraic manipulations we
obtain:
d (C A )
V + C A ( F f − Fo ) = F f C Af − Fo C A − rV (2.34)
dt
or equivalently:
d (C A )
V = F f ( C Af − C A ) − rV (2.35)
dt
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 R R (2.36)
Equations 2.6 and 2.35 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(t i ) = V i and C A (t i ) = C Ai
R R R R R R R R R (2.37)

Degrees of freedom analysis


• Parameter of constant values: A
• (Forced variable): F f and C Af
R R R

• Remaining variables: V, F o , and C A R R R

• Number of equations: 2 (Eq. 2.6 and Eq. 2.35)

The degree of freedom is therefore 3− 2 =1. The extra relation is obtained by the
relation between the effluent flow F o and the level in open loop operation (Eq. 2.10).
R R

The steady state behavior can be simply obtained by setting the accumulation terms
to zero. Equation 2.6 and 2.35 become:
F0 = Ff
R R R (2.38)
F f ( C Af − C A ) = rV (2.39)

32
2.1.4 Isothermal CSTR of Two feed streams
More complex situations can also be modeled in the same fashion. Consider the
catalytic hydrogenation of ethylene:
A+BP (2.40)
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 F 1 (m3/s) and concentration C B1 R R P P R R

(mole/m3) and another dilute stream with flow rate F 2 (m3/s) and concentration C B2
P P R R P P R R

(mole/m3). The effluent has flow rate F o (m3/s) and concentration C B (mole/m3). The
P P R R P P R R P P

reactant A is assumed to be in excess.


F1, CB1 F2, CB2

V
Fo, CB

Figure 2.5 Reaction in a CSTR

The reaction rate is assumed to be:

r=
k1CB
( mole / m 3 .s )
(2.41)
(1 + k2CB ) 2

Where k 1 is the reaction rate constant and k 2 is the adsorption equilibrium constant.
R R R R

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:
dL
A = F1 + F2 − Fo (2.42)
dt
Component B balance:
d (C B )
V = F1 ( C B1 − C B ) + F2 ( C B 2 − C B ) − rV (2.43)
dt

33
Degrees of freedom analysis
• Parameter of constant values: A, k 1 and k 2 R R R

• (Forced variable): F 1 F 2 C B1 and C B2


R R R R R R R

• Remaining variables: V, F o , and C B R R R

• Number of equations: 2 (Eq. 2.42 and Eq. 2.43)


The degree of freedom is therefore−3 2 =1. The extra relation is between the
effluent flow F o and the level L as in the previous example.
R R

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
removed via a cooling system as shown in figure 2.6. The effluent temperature is
different from the inlet temperature due to heat generation by the exothermic
reaction.

Ff , CAf , Tf

Qe V Fo, CA, T

Figure 2.6 Non-isothermal CSTR

Assuming constant density, the macroscopic total mass balance (Eq. 2.6) and mass
component balance (Eq. 2.35) remain the same as before. However, one more ODE
will be produced from the applying the conservation law (equation 2.20) for total
energy balance. The dependence of the rate constant on the temperature:
k = k o e-E/RT R R P (2.44)
The general energy balance (Eq. 2.20) for macroscopic systems applied to the CSTR
yields, assuming constant density and average heat capacity:

34
ρC p
(
~ d V (T − Tref ) ~ ) ~
= ρF f C p (T f − Tref ) − ρFoC p (T − Tref ) + Qr − Qe (2.45)
dt
Where Q r (J/s) is the heat generated by the reaction, and Q e (J/s) the rate of heat
R R R R

removed by the cooling system. Assuming T ref = 0 for simplicity and using the R R

differentiation principles, equation 2.45 can be written as follows:


~ dT
ρC pV
~ dV
+ ρC pT
~ ~
= ρF f C pT f − ρFoC pT + Qr − Qe
(2.46)
dt dt
Substituting Equation 2.6 into the last equation and rearranging yields:
~ dT
ρC pV
~
= ρF f C p (T f − T ) + Qr − Qe
(2.47)
dt
The rate of heat exchanged Q r due to reaction is given by: R R

Q r = −(∆H r )Vr
R R R R
(2.48)
Where ∆H r (J/mole) is the heat of reaction (has negative value for exothermic
R R

reaction and positive value for endothermic reaction). The non-isothermal CSTR is
therefore modeled by three ODE's:
dV
= F f − Fo
(2.49)
dt

V
d (C A )
= F f ( C Af − C A ) − rV
(2.50)
dt
~ dT
ρC pV
~
= ρF f C p (T f − T ) + (−∆H r )Vr − Qe
(2.51)
dt
Where the rate (r) is given by:
r = k o e-E/RTC A R R P P R (2.52)
The system can be solved if the system is exactly specified and if the initial
conditions are given:
V(t i ) = V i
R R R R T(t i ) = T i R R R R and C A (t i ) = C Ai
R R R R R (2.53)

Degrees of freedom analysis


• Parameter of constant values: ρ, E, R, C p , ∆H r and k o R R R R R

• (Forced variable): F f , C Af and T f R R R R R

• Remaining variables: V, F o , T, C A and Q e R R R R R

• Number of equations: 3 (Eq. 2.49, 2.50 and 2.51)

35
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 (F o ) and the volume (V) on one hand
R R

and between the rate of heat exchanged (Q e ) and temperature (T) on the other hand,
R R

in either open loop or closed loop operations.

2.1.6 Heat Exchanger

Consider the shell and tube heat exchanger shown in figure 2.7. Liquid A of
density ρ A is flowing through the inner tube and is being heated from temperature
R R

T A1 to T A2 by liquid B of density ρ B flowing counter-currently around the tube.


R R R R R R

Liquid B sees its temperature decreasing from T B1 to T B2 . Clearly the temperature of


R R R R

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.
Tw Liquid, B
TB1

Liquid, A
TA1 TA2

Liquid, B
TB2

Figure 2.7 Heat Exchanger

36
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.
One way to model the heat exchanger is to take as state variable the exit temperatures
T A 2 and T B 2 of each liquid. A better way would be to take as state variable not the
R RR R R RR R

exit temperature but the average temperature between the inlet and outlet:
TA1 + TA2
TA = (2.54)
2
TB1 + TB 2
TB = (2.55)
2
For liquid A, a macroscopic energy balance yields:
dTA
ρ AC p VA = ρ A FAC p A (TA1 − TA2 ) + Q (2.56)
A
dt
where Q (J/s) is the rate of heat gained by liquid A. Similarly for liquid B:
dTB
ρ BC p VB = ρ B FBC p B (TB1 − TB 2 ) − Q (2.57)
B
dt
The amount of heat Q exchanged is:
Q = UA H (T B – T A )
R R R R R R (2.58)
Or using the log mean temperature difference:
Q = UA H ∆T lm R R R
(2.59)
Where
(TA2 − TB1 ) − (TA1 − TB 2 )
∆Tlm =
(T − TB1 )
ln A2 (2.60)
(TA1 − TB 2 )

with U (J/m2s) and A H (m2) being respectively the overall heat transfer coefficient and
P P R R P P

heat transfer area. The heat exchanger is therefore describe by the two simple ODE's
(Eq. 2.56) and (Eq. 2.57) and the algebraic equation (Eq. 2.58).

Degrees of freedom analysis


• Parameter of constant values: ρ Α , Cp A , V A , ρ Β , Cp B , V B , U, A H R R R R R R R R R R R R R

• (Forced variable): T A1, T B1 , F A , F B R R R R R R R

37
• Remaining variables: T A2 , T B2 , Q R R R R

• Number of equations: 3 (Eq. 2.56, 2.57, 2.58)

The degree of freedom is 5− 3 = 2. The two extra relations are obtained by noting
that the flows F A and F B are generally regulated through valves to avoid fluctuations
R R R R

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 V w , density ρ w and constant heat R R R R

capacity Cp w . We also assume that the wall is at constant temperature T w , not a bad
R R R R

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 h o,t on the R R

outside and on the heat transfer coefficient h i,t on the inside. Writing the energy R R

balance for liquid B yields:

ρ BC p VB
dTB
= ρ B FBC p B (TB1 − TB 2 ) − ho , t Ao , t (TB − TW )
(2.61)
B
dt
Where A o,t is the outside heat transfer area. The energy balance for the metal yields:
R R

ρ wC p Vw
dTw
= ho , t Ao , t (TB − Tw ) − hi , t Ai , t (Tw − TA )
(2.62)
w
dt
Where A i,t is the inside heat transfer area. The energy balance for liquid A yields:
R R

ρ AC p VA
dTA
= ρ A FAC p A (TA1 − TA2 ) + hi , t Ai , t (Tw − TA )
(2.63)
A
dt
Note that the introduction of equation (Eq. 2.62) does not change the degree of
freedom of the system.

2.7 Heat Exchanger with Steam


A common case in heat exchange is when a liquid L is heated with steam (Figure
2.8). If the pressure of the steam changes then we need to write both mass and energy
balance equations on the steam side.

38
Tw Steam
Ts(t)

Liquid, L
TL1 TL2

condensate, Ts

Figure 2.8 Heat Exchanger with Heating Steam

The energy balance on the tube side gives:


dTL
ρ LC p L VL = ρ L FLC p L (TL1 − TL 2 ) + Qs (2.64)
dt
where
TL1 + TL 2
TL = (2.65)
2
Q s = UA s (T s – T L )
R R R R R R R R (2.66)
The steam saturated temperature T s is also related to the pressure P s : R R R R

T s = T s (P)
R R R R (2.67)
Assuming ideal gas law, then the mass flow of steam is:
M s PsVs
ms = (2.68)
RTs

Where M s is the molecular weight and R is the ideal gas constant. The mass balance
R R

for the steam yields:


M sVs dP
= ρ s Fs − ρc Fc (2.69)
RTs dt

Where F c and ρ c are the condensate flow rate and density. The heat losses at the
R R R R

steam side are related to the flow of the condensate by:


Qs = Fc λs
R R R R R (2.70)
Where λ s is the latent heat.
R R

Degrees of freedom analysis


• Parameter of constant values: ρ L , Cp L , M s , A s , U , M s , R R R R R R R R R R R

39
• (Forced variable): T L1 R

• Remaining variables: T L2 , F L , T s , F s , P s , Q s , F c
R R R R R R R R R R R R R R

• Number of equations: 5 (Eq. 2.64, 2.66, 2.67, 2.69, 2.70)

The degrees of freedom is therefore 7 – 5 = 2. The extra relations are given by the
relation between the steam flow rate F s with the pressure P s either in open-loop or
R R R R

closed-loop operations. The liquid flow rate F 1 is usually regulated by a valve. R R

2.1.8 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 2.9. The feed consists
of N c components with the molar fraction z i (i=1,2… N c ). The feed at high
R R R R R R

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. x i and y i represent the mole fraction of
R R R R

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.

40
Fv
yi
P, T, Vv
Fo
zi
To
Po VL ρL FL
xi

Figure 2.9 Multicomponent Flash Drum

For liquid phase:


Total mass balance:
d ( ρ LVL )
= ρ f F f − ρ L FL − ρ v Fv (2.71)
dt
Component balance:
d ( ρ LV L x i )
= ρ f F f z i − ρ L FL x i − ρ v Fv y i (i=1,2,….,N c -1)
R R
(2.72)
dt
Energy balance:
~
d ( ρ LVL h ) ~ ~ ~
= ρ f Ff h f − ρ L FL h − ρv Fv H
dt (2.73)
~ ~
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
xi Pi s
yi = (i=1,2,….,N c )
(2.74)
R R

P
Together with the consistency relationships:
Nc

∑y
i =1
i =1
(2.75)
Nc

∑x
i =1
i =1
(2.76)

41
• Physical Properties:
The densities and enthalpies are related to the mole fractions, temperature and
pressure through the following relations:
ρL R R = f(x i ,T,P)
R R (2.77)
ρv R R = f(y i ,T,P) ≈ M v aveP/R T
R R R RP P (2.78)
Nc
(2.79)
M v ave =
R RP P

∑y M
i =1
i i

Nc
(2.80)
h = f(x i ,T) ≈
R R

∑ x Cp (T − T
i =1
i i ref )

Nc
(2.81)
H = f(y i ,T) ≈
R R

∑ y Cp (T − T
i =1
i i ref ) + λm

Nc
(2.82)
λm R R = ∑yλ
i =1
i i

Degrees of freedom analysis:


• Forcing variables: F f , T f , P f , z i (i=1,2..N c ), R R R R R R R R R R

• Remaining variables:2N c +5: V L , F L , F V , P, T, x i (i=1,2..N c ), y i (i=1,2,…N c ) R R R R R R R R R R R R R R R R

• Number of equations: 2N c +3: (Eq. 2.71, 2.72, 2.73, 2.74, 2.75, 2.76) R R

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 (2N c +5)- R R

(2N c +3)=2. Generally the liquid holdup (V L ) is controlled by the liquid outlet flow
R R R R

rate (F L ) while the pressure is controlled by F V . In this case, the problem becomes
R R R R

well defined for a solution.

2.1.9 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,

42
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.

Liquid-liquid extraction is used to move a solute from one liquid phase to another.
Consider the single stage countercurrent extractor, shown in Figure 2.10, 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 X Af weight fraction of
R R

solute (A). The solvent with a flow rate (S) (kg/s) enters the stage containing Y AfR R

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 X A and Y A are so small that they do not affect the mass flow
R R R R

rates. Therefore, we can assume that the flow rates W and S are constant. A total
mass balance is therefore not needed.

43
• An equilibrium relationship exists between the weight fraction Y A of the solute R R

in the solvent (S) and its weight fraction X A in the mixture (W). The relationship R R

can be of the form:


Y A = K XA
R R R (2.83)
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 V 1 and density ρ 1 gives: R R R R

dX A
ρ1V1 = WX Af − WX A − N A (2.84)
dt
whereas N A (kg/s) is the flow rate due to transfer flow between the two phases. A
R R

similar component balance on the solvent phase of volume V 2 and density ρ 2 gives: R R R R

dYA
ρ2V2 = SYAf − SYA + N A (2.85)
dt
Since Y A = K X A and K is constant, the last equation is equivalent to:
R R R R

dX A
ρ 2V2 K = SY Af − SKX A + N A
dt (2.86)
Adding Eq. 2.84 and 2.86 yields:
dX A
( ρ1V1 + ρ 2V2 K ) = WX Af + SY Af − (W + KS ) X A
dt (2.87)
The latter is a simple linear ODE with unknown X A . With the volume V 1 , V 2 and flow R R R R R R

rates W, S known the system is exactly specified and it can be solved if the initial
concentration is known:
X A (t i ) = X Ai
R R R R R (2.88)
Note that we did not have to express explicitly the transferred flux N A . R R

S, YA S, YAf

W, XAf W, xA

Figure 2.10 Single Stage Liquid-Liquid Extraction Unit

The same analysis can be extended to the multistage liquid/liquid extraction units
as shown in Figure 2.11. The assumptions of the previous example are kept and we

44
also assume that all the units are identical, i.e. have the same volume. They are also
assumed to operate at the same temperature.

V1, y1 V2, y2 V3, y3 Vi, yi Vi+1, yi+1 VN, yN VN+1, yf

1 2 i N

L0, xf L1, x1 L2, x2 Li-1, xi-1 Li, xi LN-1, xN-1 LN, xN

Figure 2.11 Multi-Stage Liquid-Liquid Extraction Unit

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

• Solvent-free phase, of volume V 1 i and density ρ 1 i R RR R R RR

ρ1iV1i
dX Ai
= WX Ai −1 − WX Ai − N Ai (i=2…,N-1)
(2.89)
dt
where N Ai is the flow rate due to transfer between the two phases at stage i.
R R

• Solvent phase of volume V 2 i and density ρ 2 i : R RR R R RR R

ρ 2iV2i
dY Ai
= SY Ai +1 − SY Ai + N Ai … (i = 2…,N-1)
(2.90)
dt
Writing the equilibrium equation (Eq. 2.83) for each component Y Ai = K X Ai R R R

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

(ρ1iV1i + ρ2iV2i K )
dX Ai
= WX Ai −1 + SYAi +1 − (W + KS ) X Ai (i=2…,N-1)
(2.91)
dt
Since the volume and densities are equal, i.e.:
V 1i = V 1 and V 2i = V
R R R R R R (2.92)
ρ 1i = ρ 1 and ρ 2i = ρ 2
R R R R R R R
(2.93)
Equation 2.91 is therefore equivalent to:
dX Ai
(ρ1V1 + ρ2V2 K ) = WX Ai −1 + SYAi +1 − (W + KS ) X Ai (i=2…,N-1)
dt (2.94)
The component balance in the first stage is:
dX A1
(ρ1V1 + ρ2V2 K ) = WX Af + SYA1 − (W + KS ) X A1
dt (2.95)
And that for the last stage is:

45
dX AN
(ρ1V1 + ρ2V2 K ) = WX AN −1 + SYAf − (W + KS ) X AN
dt (2.96)
The model is thus formed by a system of linear ODE's (Eq. 2.94, 2.95, 2.96) which
can be integrated if the initial conditions are known:
X A (t i ) = X Ai (i=1,2…,N)
R R R R R R (2.97)

Degrees of freedom analysis


• Parameter of constant values: ρ 1 , ρ 2, K, V 1 , V 2 , W and S R R R R R R R R

• (Forced variable): X Af, Y Af


R R R

• Remaining variables: X Ai (2N variables): (i=1,2…,N) and Y Ai (i=1,2…,N)


R R R R

• Number of equations: 2N [2.83 (N equations, one for each component), Eq.


2.94 (N-2 eqs), 2.95(1 eq), 2.96(1 eq)].
The problem is therefore is exactly specified.

2.10 Binary Absorption Column


Consider a N stages binary absorption tower as shown in figure 2.12. A Liquid
stream flows downward with molar flow rate (L) and feed composition (x f ). A Vapor R R

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

interested in deriving an unsteady state model for the absorber. A simple vapor-liquid
equilibrium relation of the form of:
yi = a xi + b
R R R R (2.98)
Can be used for each stage i (i=1,2,…,N).

Assumptions:
• 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

46
not be written. The last assumption allows us writing a macroscopic balance on each
stage as follows:
Component balance on stage i:
dxi
H = G ( yi −1 − yi ) + L( xi +1 − xi ) (i=2…,N-1) (2.99)
dt
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:
L, xf G, yN

stage N

xN yN-1

xi+1 yi

stage i
xi yi-1

x3 y2

stage 2

x2 y1

stage 1

L, x1 G, yf

Figure 2.12 N-stages Absorbtion Tower

In the last stage, x i+1 is replaced by x f


R R R

In the first stage, y i-1 is replaced by y f


R R R

Degrees of freedom analysis


• Parameter of constant values: Η, a, b
• (Forced variable): G, L, x f , y f R R R

• Remaining variables: x i (i=1,2…,N), y i (i=1,2…,N)


R R R R

• Number of equations:2N (Eqs.2.98, 2.99)


The problem is therefore is exactly specified.
47
2.11 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 2.13. The column consists of n trays excluding the reboiler and the
total condenser. The convention is to number the stages from the bottom upward
starting with the reboiler 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 reboiler 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 reboiler. 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.

48
Cw

D
xd

F
z

steam

B
xb

Figure 2.13 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 reboiler.
• 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 2.14a:


Total mass balance:
dM D
= Vn − ( R + D)
dt (2.100)
Component balance:
49
d ( M D xD , j ) (2.101)
= Vn yn , j − ( R + D ) x D , j j = 1, nc − 1
dt
Energy balance:
d ( M D hD )
= Vn hn − ( R + D )hD − Qc
(2.102)
dt
Note that R = L n+1 and the subscript D denotes n+1
R R

Stage n, Figure fig2.19b


Total Mass balance:
dM n
= Vn −1 − Vn + R − Ln
(2.103)
dt
Component balance:
d (M n xn, j ) (2.104)
= Vn −1 y n −1, j − Vn y n , j + Rx D , j − Ln x n , j j = 1, nc − 1
dt
Energy balance:
d ( M n hn )
= Vn −1H n −1 − Vn H n + RhD − Ln hn
(2.105)
dt

Stage i, Figure 2.14c


Total Mass balance:
dM i
= Vi −1 − Vi + Li +1 − Li
(2.106)
dt
Component balance:
d ( M i xi , j ) (2.107)
= Vi −1 yi −1, j − Vi yi , j + Li +1 xi +1, j − Li xi , j j = 1, nc − 1
dt
Energy balance:
d ( M i hi )
= Vi −1H i −1 − Vi H i + Li +1hi +1 − Li hi
(2.108)
dt

Stage f (Feed stage), Figure 2.14d


Total Mass balance:
dM f (2.109)
= V f −1 − (V f + (1 − q ) F ) + L f +1 − ( L f + qF )
dt
Component balance:
50
d (M f x f , j ) (2.110)
= V f −1 y f −1, j − (V f y f , j + (1 − q) Fz j ) + L f +1 x f +1, j − ( L f x f , j + qFz j )
dt
j = 1, nc − 1

Energy balance:
d (M f hf ) (2.111)
= V f −1H f −1 − (V f H f + (1 − q) Fh f ) + L f +1h f +1 − ( L f h f + qFh f )
dt

Stage 1, Figure 2.14e


Total Mass balance:
dM 1
= VB − V1 + L2 − L1 (2.112)
dt
Component balance:
d ( M 1 x1, j )
= VB y B , j − V1 y1, j + L2 x2, j − L1 x1, j j = 1, nc − 1 (2.113)
dt

Energy balance:
d ( M 1h1 )
= VB H B − V1H1 + L2 h2 − L1h1 (2.114)
dt

Stage 0 (Reboiler), Figure 2.14f


Total Mass balance:
dM B
= −VB + L1 − B (2.115)
dt
Component balance:
d ( M B xB , j )
= −VB y B , j + L1 x1, j − Bx B , j j = 1, nc − 1 (2.116)
dt
Energy balance:
d ( M B hB )
= −VB H B + L1h1 − BhB + Qr (2.117)
dt
Note that L 0 = B and B denotes the subscript 0
R R

Additional given relations:


Phase equilibrium: y j = f (x j , T,P)
R R R R

51
Liquid holdup: M i = f (L i ) R R R R

Enthalpies: H i = f (T i , y i,j ), h i = f (T i , x i,j ) R R R R R R R R R R R R

Vapor rates: V i = f (P) R R

Notation:
Li, Vi
R R R R Liquid and vapor molar rates
Hi, hi R R R R Vapor and liquid specific enthalpies
xi, yi
R R R R Liquid and vapor molar fractions
Mi R R Liquid holdup
Q Liquid fraction of the feed
Z Molar fractions of the feed
F Feed molar rate

Degrees of freedom analysis


Variables

Mi R R n
MB, MD R R R R 2
Li R R n
B,R,D 3
x i,j
R R n(nc − 1)
x B,j ,x D,j
R R R 2(nc − 1)
y i,j
R R n(nc − 1)
y B,j
R R nc − 1
hi R R n
hB, hDR R R R 2
Hi R R n
HB R R 1
Vi R R n
VB R R 1
Ti R R n
52
TD, TB R R R R 2

Total 11+6n+2n(nc−1)+3(nc−1)

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 R R R 1
yB = xB
R R R (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 V B and R to control the end compositions i.e., x B , x D
R R R R R

53
R, xd Vn, yn
Vn, yn

stage n
(a)
Qc (b)

R, xd D, xd
Ln, xn Vn-1, yn-1

Li+1, xi+1 Vi, yi Lf+1, xf+1 Vf , yf

stage i stage f

(c) (d)

Li, xi Vi-1, yi-1 Lf, xf Vf-1, yf-1

L2, x2 V1, y1 VB, yB


Qr

stage 1
(f)
B, xB
(e)

L1, x1 V B , yB
L1, x1

Figure 2.14 Distillation Column Stages

2.12 Multi-component Batch Distillation Column


For batch distillation column of constant trays holdup, all equation applied for
conventional distillation column could be used to model batch distillation except the
rebolier equations. Figure (2.15) show the reboiler of batch distillation column.

Section 3

Section 2
L2 Figure (2.15) Reboiler of batch distillation column
V1
Reboiler M1

54
Total mass balance.
dM B
= L1 − VB (2.118)
dt
Component balance
d ( M B ⋅ x B ,i )
= L1 x1,i − VB y B ,i (2.119)
dt
Energy balance
d ( M B hB )
= L1h1 − VB H B + Qr (2.120)
dt
Where Q r is heat supplied by reboiler
R R

2.13 Continuous Reactive Distillation Column


Figure (2.16) represents the continuous packed reactive distillation column. Reactive
distillation column can be divided into three different zones: rectifying, reactive and
stripping. The rectifying section and the stripping section serve as the reactant
recovery from product. However, they may not be needed depending on the reaction
and desired separation. In the reactive zone, chemical reaction and distillation occur
simultaneously. Reactive zone also is the place where catalysts, either homogeneous
or heterogeneous are packed. The feed input can be either single or double. Double
feed input can improve the contact between the reactants, resulting more uniformly
distributed reaction in the column.

55
Figure (2.16) continuous packed reactive distillation column.

2.13.1 Model Assumptions


The packed reactive distillation column is vertically divided into a number of
segments. The reboiler and condenser stages are numbered 1 and N, respectively. The
following assumptions were made to simplify the model of continuous reactive
distillation column:-
1. Neglect of vapor holdup and assume total condensation.
2. Perfect mixing on all stages and in all vessels (condenser and reboiler), and the
condenser and the reboiler are treated as equilibrium stages.
3. Ideal vapor phase for all components in mixture.
4. Liquid and vapor phases in thermodynamic phase equilibrium.

1.13.2 Model Equations


All equations applied for conventional distillation could be used to model reactive
distillation except component and heat balance equations in reaction stages. Figure
(2.17) represent reactive stage.

56
Lm+1 Vm

Stage Figure (2.17) Reaction Stage

Lm Vm-1

Total mass balance


dM m
= Lm+1 + Vm−1 − Lm − Vm (2.121)
dt
Component balance
d ( M m x m ,i )
= Lm+1 xm+1,i + Vm−1 y m−1,i − Lm xm ,i − Vm y m ,i + ε iWm Re m ,i (2.122)
dt
Energy balance
d ( M m hm )
= Lm+1hm+1 + Vm−1 H m−1 − Lm hm − Vm H m + Wm Re m ,i ∆H R (2.123)
dt
Where
M : Molar holdup mol
Re : Forward and reverse reaction rate mol/(s. gm catalyst)
W : Catalyst weight gm
ε : Void fraction of the packing -
ΔHR : Heat of reaction j/mol

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 2.18. 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

57
out on microscopic scale. We consider therefore a shell element of width ∆z and
constant cross section area (A) as shown in Fig. 2.18.
ρ(t,z)
∆z

vo v(t,z)

z=0 z z+∆z z=L

Figure 2.18 Liquid flow in a pipe

Mass into the shell:


ρvA∆t| z R
(2.124)
Where the subscript ( .| z ) indicates that the quantity (.) is evaluated at the distance z.
R R

Mass out of the shell:


ρvA∆t| z+∆z R
(2.125)
Accumulation:
ρA∆z| t+∆t − ρA∆z| t
R R R
(2.126)
Similarly the subscript (.| t ) indicates that the quantity (.) is evaluated at the time t.
R R

The mass balance equation is therefore:


ρvA∆t| z = ρvA∆t| z+∆z + ρA∆z| t+∆t − ρA∆z| t
R R R R R R R
(2.127)

We can check for consistency that the units in each term are in (kg). Dividing Eq.
2.127 by ∆t ∆z and rearranging yields:
(ρA) − (ρA) (ρvA) − (ρvA) (2.128)
t + ∆t t
= z z + ∆z
∆t ∆z

Taking the limit as ∆t  0 and ∆z  0 gives:


∂ (ρA) ∂ (ρvA) (2.129)
=−
∂t ∂z
Since the cross section area (A) is constant, Eq. 2.129 yields:
∂ρ ∂ ( ρv ) (2.130)
=−
∂t ∂z
or

58
∂ρ ∂ ( ρv ) (2.131)
+ =0
∂t ∂z

The equation (2.131) 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:
∂v (2.132)
=0
∂z
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 2.19.

r
∆r
R
r
∆r z
∆z
P|z P|z+∆z

Figure 2.19 Velocity profile for a laminar flow in a pipe


Momentum in:
(τrz 2πr∆z)| r
R R R
(2.133)
where τ rz is the shear stress acting in the z-direction and perpendicular to the radius r.
R R

Momentum out:
59
(τrz 2πr∆z)| r+∆r
R R R
(2.134)
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
R R R R R R R
(2.135)
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,
rτrz |r + ∆r −rτrz |r r ( P |z − P |z + ∆z ) (2.136)
=
∆r ∆z
We can check for consistency that all the terms in this equation have the SI unit of
(N/m2). Taking the limit of (Eq. 2.136) as ∆z and ∆r go to zero yields:
P P

d (rτrz ) dP (2.137)
= −r
dr dz
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:
dP P2 − P1 ∆P (2.138)
= =
dz L L
where L is the length of the tube. Note that equation (2.138) is a function of the shear
stress τrz , but shear stress is a function of velocity. We make here the assumption that
R R

the fluid is Newtonian, that is the shear stress is proportional to the velocity gradient:

τ rz = − µ
dv z (2.139)
dr
Substituting this relation in Eq. 2.137 yields:
d rdv z ∆P (2.140)
µ ( )=r
dr dr L
or by expanding the derivative:
d 2 v z 1 dv z ∆P (2.141)
µ( 2
+ )=
dr r dr L
The system is described by the second order ODE (Eq. 2.141). This ODE can be
integrated with the following conditions:
• The velocity is zero at the wall of the tube

60
vz = 0
R R at r = R (2.142)
• Due to symmetry, the velocity profile reaches a maximum at the center of the
tube:
dvz
=0 at r = 0 (2.143)
dr
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 2.20. Our objective is to
determine the variation of the concentration at steady state. The concentration inside
the 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:


(SN A )| z R R R (2.144)
where S (m2) is the surface area and N A (moles A/s m2) is the molar flux.
P P R R P P

Flow of moles A out:


(SN A )| z+∆zR R R
(2.145)
Rate of generation of A:
−(S∆z)r (2.146)
where r = kC A is the rate of reaction, assumed to be of first order. There is no
R R

accumulation term since the system is assumed at steady state. The mass balance
equation is therefore,
(SN A )| z − (SN A )| z+∆z − (S∆z)kC A =0
R R R R R R R R R R
(2.147)

61
Porous
catalyst
particle
Exterior
surface

∆z
z=0 z=L
Figure 2.20 diffusion with chemical reaction inside a slab catalyst

Dividing equation (2.147) by S∆z results in:


( N A ) | z − ( N A ) | z + ∆z (2.148)
− kC A = 0
∆z

Taking the limit when ∆z  0, the last equation becomes:


dN A
− kC A = 0
(2.149)
dz
The molar flux is given by Fick's law as follows:
dC A (2.150)
N A = −D A
dz
where D A is diffusivity coefficient of (A) inside the catalyst particle. Equation (2.149)
R R

can be then written as follows:


d 2C A (2.151)
DA 2
− kC A = 0
dz
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, C A = C Ao
R R R R (2.152)
at z = 0, dC A /dr = 0 R R (2.153)
The first condition imposes the bulk flow concentration C Ao at the end length of the R R

slab. The second condition implies that the concentration is finite at the center of the
slab.

62
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 T o . Suppose that one end of the rod is brought to contact with a hot
R R

fluid of temperature T m while the surface area of the rod is exposed to ambient
R R

temperature of T a . We are interested in developing the mathematical equation that


R R

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
2.21. 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:

R
∆r
r
∆r

∆z

Figure 2.21 Temperature Distribution In a cylindrical rod

Heat flow in by conduction at z:


q z (2π r ∆r)∆t
R R
(2.154)
Heat flow in by conduction at r:
q r (2π r ∆z)∆t
R R
(2.155)
where q z and q r are the heat flux by conduction in the z and r directions.
R R R R

Heat flow out by conduction at z+∆z:


q z+∆z (2π r ∆r)∆t
R R
(2.156)
Heat flow out by conduction at r+∆r:

63
q r+∆r (2π (r+∆r)∆z)∆t
R R
(2.157)
Heat accumulation:
~ ~
ρ(2π r ∆r ∆z)( h t+∆t − h t ) R R R R
(2.158)
~
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:
~ ~
ht + ∆t − ht q − qz + ∆z rqr − rqr + ∆r (2.159)
ρr =r z +
∆t ∆z ∆r

Taking the limit of ∆t, ∆z, and ∆r go to zero yield:


~
∂h ∂q ∂ ( rqr ) (2.160)
ρr
= −r z −
∂t ∂z ∂r
~
Dividing by r and replacing h by C p (T − T ref ), where C p R R is the average heat
capacity, and substituting the heat fluxes q with their corresponding relations (Fourier
law):
∂T (2.161)
qz = −k z
∂z
and
∂T (2.162)
qr = −kr
∂r
Equation (2.160) is then equivalent to:
∂T ∂ 2T k ∂ ∂T (2.163)
ρC p = kz 2 + r (r )
∂t ∂z r ∂r ∂r
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:
∂ 2T k r ∂ ∂T (2.164)
0 = kz + (r )
∂z 2 r ∂r ∂r
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. 2.164) 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)(T a – T) R R
(2.165)

64
Consequently, the above energy balance equation (Eq. 2.165) is reduced to the
following ODE
d 2T (2.166)
0 = kz − U ( 2πRL)(Ta − T )
dz 2

2.2.5 Isothermal Plug Flow Reactor


Let consider a first-order reaction occurring in an isothermal tubular reactor as
shown in figure 2.22. 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

v(t,z)
CAo CA

z z+∆z z=L

ρ(t,z), CA(t,z)

Figure 2.22 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:
(vC A S ∆t) | z + (N A S ∆t)| z
R R R R R R R R
(2.167)
Where N A is the diffusive flux of A ( moles of A/m2 s).
R R P P

Flow of moles of A out:


(vC A S ∆t) | z+∆z + (N A S ∆t)| z+∆z
R R R R R R R
(2.168)
Accumulation:

65
(C A S ∆z) | t+∆t − (C A S ∆z) | t
R R R R R R R
(2.169)
Generation due to reaction inside the shell:
− r(S∆z∆t) (2.170)
Where r = k C A is the rate of reaction.
R R

Substituting all the terms in the mass balance equation (Eq. 1.3) and dividing by ∆t
and ∆z gives:
(C A S ) |t + ∆t −(C A S ) |t (vC A S + N A S ) |z −(vC A S + N A S ) |z + ∆z (2.171)
= − kC A S
∆t ∆z

Taking the limit of ∆t  0 and ∆z  0 and omitting S from both sides give the
following PDE:
∂C A ∂vC A ∂N A (2.172)
=− − − kC A
∂t ∂z ∂z

Where N A is the molar flux given by Fick’s law as follows:


R R

N A = − DAB
dC A (2.173)
dz
Where D AB is the binary diffusion coefficient. Equation 2.172 can be then written as
R R

follows:
∂C A ∂ ( vC A ) ∂ 2C A (2.174)
=− + DAb − kC A
∂t ∂z ∂z 2
Expanding the derivatives, the last equation can be reduced to:
∂C A ∂C ∂v ∂ 2C A (2.175)
= −v A − C A + DAb − kC A
∂t ∂z ∂z ∂z 2
This equation can be further simplified by using the mass balance equation for
incompressible fluids (Eq. 2.132). We get then:
∂C A ∂C ∂ 2C A (2.176)
= −v A + DAb − kC A
∂t ∂z ∂z 2
The equation is a PDE for which the state variable (C A ) depends on both t and z.R R

The PDE is reduced at steady state to the following second order ODE,
dC A d 2C A (2.177)
0 = −v + DAb − kC A
dz dz 2
The ODE can be solved with the following boundary conditions (BC):

66
BC1: at z = 0 C A (0) = C A0
R R R (2.178)
BC2: at z = L dC A ( z )
=0
(2.179)
dz
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 2.23. 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.2.23. The following assumptions are made for the energy balance:
Water in

To T
CAo v(t,z) v(t,z)
CA

Water out T(t,z)


CA(t,z)
ρ(t,z)

Figure 2.23 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:
67
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 q z . R R

~
(q z A + v Aρ h )∆t| z
R R R R
(2.180)
Energy flow out of the shell:
~
(q z A + v Aρ h )∆t| z+∆z
R R R R
(2.181)
Accumulation of energy:
~ ~
(ρA h ∆z)| t+∆t − (ρA h ∆z)| t R R R R
(2.182)
Heat generation by reaction:
(−∆H r )kC A A ∆z ∆t R R R R
(2.183)
Heat transfer to the wall:
h t (πD∆z)(T − T w )∆t
R R R R
(2.184)
Where h t is film heat transfer coefficient.
R R

Substituting these equations in the conservation law (equation 1.7) and dividing by
Α∆t ∆z give:
~ ~ ~ ~
( ρh ) |t + ∆t −( ρh ) |t ( ρvh ) | z −( ρvh ) | z + ∆z q z | z −q z | z + ∆z πD
= + − ∆H r kC A − ht ( )(T − Tw ) (2.185)
∆t ∆z ∆z A

Taking the limit as ∆t and ∆z go to zero yields:


~ ~
∂ ( ρh ) ∂ ( ρvh ) ∂qz πD (2.186)
=− − − ∆H r kC A − ht ( )(T − Tw )
∂t ∂z ∂z A
The heat flux is defined by Fourier’s law as follows:
∂T (2.187)
q z = − kt
∂z
~
Where k t is the thermal conductivity. The specific enthalpy ( h ) can be approximated
R R

by:
~
h = C p(T − Tref ) (2.188)

Since the fluid is incompressible it satisfies the equation of continuity (Eq. 2.132).
Substituting these expressions in Eq. 2.186 and expanding gives:
∂T ∂T ∂ 2T πD
ρC p = − ρC pv + kt 2 − ∆H r ko e − E / RT C A − ht ( )(T − Tw ) (2.189)
∂t ∂z ∂z A

68
At steady state this PDE becomes the following ODE,
dT d 2T πD
0 = − ρC pv + kt 2 − ∆H r ko e − E / RT C A − ht ( )(T − Tw ) (2.190)
dz dz A
Similarly to Eq. 2.177 we could impose the following boundary conditions:
B.C1: at z = 0 T(z) = T oR (2.191)
B.C2: at z = L dT ( z )
=0
(2.192)
dz
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.

69
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.

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.
(ρvy)y+∆y
(ρvz)z+∆z

∆y

∆x
(ρvx)x+∆x
(ρvx)x

∆z

z (ρvy)y
y
(ρvz)z

Figure 3.1 Total Mass balance in Cartesian coordinates

Mass in:
The mass entering in the x-direction at the cross sectional area (∆y∆z) is
(ρv x )| x ∆y∆z∆t
R R R R (3.1)
The mass entering in the y-direction at the cross sectional area (∆x∆z) is
(ρv y )| y ∆x∆z∆t R R R R (3.2)
The mass entering in the z-direction at the cross sectional area (∆x∆y) is
(ρv z )| z ∆x ∆y ∆t
R R R R
(3.3)

Mass out:
The mass exiting in the x-direction is:
(ρv x )| x+∆x ∆y∆z∆t
R R R R
(3.4)
The mass exiting in the y-direction is:
(ρv y )| y+∆y ∆x∆z∆t
R R R R
(3.5)
The mass exiting in the z-direction is:
(ρv z )| z+∆z ∆x∆y ∆t
R R R R
(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
R R R R
(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 = (ρv x | x − ρv x | x+∆x )∆y∆z∆t + (ρv y | y
R R R R R R R R R R R R R R R R

(3.8)
− ρv y | y+∆y )∆x∆z∆t + (ρv z | z − ρv z | z+∆z ) ∆x ∆y ∆t
R R R R R R R R R R R R

Dividing the equation by ∆x ∆y ∆z∆t results in:


ρ |t +∆t −ρ |t ρvx | x −ρvx |x+∆x ρv y | y −ρv y | y +∆y ρvz | z −ρvz | z +∆z
= + + (3.9)
∆t ∆x ∆y ∆z

By taking the limits as ∆y,∆x,∆,z and ∆t goes to zero, we obtain the following
equation of change:
∂ρ ∂ρvx ∂ρv y ∂ρvz
=− − − (3.10)
∂t ∂x ∂y ∂z

Expanding the partial derivative of each term yields after some rearrangement:
∂ρ ∂ρ ∂ρ ∂ρ ∂v ∂v ∂v
+ vx + vy + vz = −ρ( x + y + z ) (3.11)
∂t ∂x ∂y ∂z ∂x ∂y ∂z
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:
∂vx ∂v y ∂vz
0 = −ρ( + + ) (3.12)
∂x ∂y ∂z

or simply:
∂vx ∂v y ∂vz
0= + + (3.13)
∂x ∂y ∂z

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 r A (kg/m3s). The fluid is in motion with mass-
R R P P

average velocity v = n t /ρ (m/s) where n t = n A + n B + … (kg/m2s) is the total mass flux


R R R R R R R R P P

and ρ (kg/m3s) is the density of the mixture. Our objective is to establish the
P P

component balance equation of A as it diffuses in all directions x,y,z (Figure 3.2).


z nAy|y+∆y
nAz|z

∆y

nAx|x+∆x
nAx|x
∆z
y

∆x

nAy|y nAz|z+∆z
x

Figure 3.2 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:
(n Ax )| x ∆y∆z∆t
R R R R
(3.14)
where n Ax kg/m2 is the flux transferred in the x-direction
R R P P
Similarly the mass of A entering the y and z direction are respectively:
(n Ay )| y ∆x∆z∆t
R R R R
(3.15)
(n Az )| z ∆x∆y∆t
R R R R
(3.16)
Mass of A out:
The mass of species A exiting the x, y and z direction are respectively
(n Ax )| x+∆x ∆y∆z∆t
R R R R
(3.17)
(n Ay )| y+∆y ∆x∆z∆t
R R R R
(3.18)
(n Az )| z+∆z ∆x∆y∆t
R R R R
(3.19)
The rate of accumulation is:
ρ A | t+∆t ∆x ∆y ∆z − ρ A | t ∆x ∆y ∆z
R R R R R R R R
(3.20)

The rate of generation is:


-r A ∆x ∆y ∆z ∆t
R R
(3.21)

Applying the general balance equation (Eq. 1.3) yields:


(ρ A | t+∆t − ρ A | t ) ∆x ∆y ∆z = (n Ax | x+∆x − n Ax | x )∆y∆z∆t + (n Ay | y+∆y −
R R R R R R R R R R R R R R R R R R R R

(3.22)
n Ay | y )∆x∆z∆t + (n Az | z+∆z − n Az | z )∆x∆y∆t +r A ∆x ∆y ∆z∆t
R R R R R R R R R R R R R R

Dividing each term by ∆x∆y∆z∆t and letting each of these terms goes to zero yields:
∂ρ A ∂n Ax ∂n Ay ∂n Az
+ + + = rA (3.23)
∂t ∂t ∂t ∂t
We know from Section 1.11.1, that the flux n A is the sum of a term due to convection R R

(ρ A v) and a term due to diffusion j A (kg/m2s):


R R R R P P

n A = ρA v + j A
R R R R R (3.24)
Substituting the different flux in Eq. 3.23 gives:
∂ρ A ∂ ( ρ A v x ) ∂ ( ρ A v y ) ∂ ( ρ A v z ) ∂j Ax ∂j Ay ∂j Az
+ + + + + + = rA (3.25)
∂t ∂x ∂y ∂z ∂x ∂y ∂z

For a binary mixture (A,B), Fick’s law gives the flux in the u-direction as :
∂wA
j Au = −ρDAB (3.26)
∂u

where w A = ρ A /ρ. Expanding Eq. 3.25 and substituting for the fluxes yield:
R R R R
∂ρ A  ∂v ∂v y ∂v z   ∂ρ A ∂ρ ∂ρ 
+ ρ A  x + +  +  v x + v y A + v z A 
∂t  ∂x ∂y ∂z   ∂x ∂y ∂z 
(3.27)
 ∂ ∂ρD AB w A ∂ ∂ρD AB w A ∂ ∂ρD AB w A 
−  ( )+ ( )+ ( )  = rA
 ∂x ∂x ∂y ∂y ∂z ∂z 

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 D AB are constant. Substituting the
R R

continuity equation (Eq. 3.13) in the last equation gives:


∂ρ A  ∂ρ A ∂ρ ∂ρ   ∂ 2ρ A ∂ 2ρ A ∂ 2ρ A 
+  vx + v y A + vz A  − DAB  + + )  = rA (3.28)
∂t  ∂x ∂y ∂z   ∂x 2 ∂y 2 ∂z 2 

This equation can also be written in molar units by dividing it by the molecular
weight M A to yield:
R R

∂C A  ∂C A ∂C A ∂C A   ∂ 2C A ∂ 2C A ∂ 2C A 
+  v x + vy + vz  − D AB  + +  = RA
∂
t  ∂x ∂y ∂z   ∂x 2 ∂ 2
∂ 2   (3.29)
     y

z
 reaction
accumulation Convection Diffusion

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).
−σyx |y+∆y σzx |z

−σxx |x+∆x
σxx |x

−σzx |z+∆z
σyx |y

x
z

Figure 3.3 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:
(ρv x v x )| x ∆y∆z∆t
R R R R R R
(3.30)
The x-component of momentum entering the boundary at y-direction, by
convection is:
(ρv y v x )| y ∆x∆z∆t
R R R R R R
(3.31)
and it enters the z-direction by convection with a momentum:
(ρv z v x )| z ∆x∆y∆t
R R R R R R
(3.32)
The x-component of momentum entering the boundary at x-direction, by molecular
diffusion is:
(τxx )| x ∆y∆z∆t
R R R R
(3.33)
The x-component of momentum entering the boundary at y-direction, by molecular
diffusion is:
(τyx )| y ∆x∆z∆t R R R R
(3.34)
and it enters the z-direction by molecular diffusion with a momentum:
(τzx )| z ∆x∆y∆t R R R R
(3.35)

Momentum out:
The rate of momentum leaving the boundary at x+∆x, by convection is:
(ρv x v x )| x+∆x ∆y∆z∆t R R R R R R
(3.36)
and at boundary y+∆y,:
(ρv y v x )| y+∆y ∆x∆z∆t
R R R R R R
(3.37)
and at boundary z+∆z:
(ρv z v x )| z+∆z ∆x∆y∆t R R R R R R
(3.38)

The x-component of momentum exiting the boundary x+∆x, by molecular


diffusion is:
(τxx )| x+∆x ∆y∆z∆t
R R R R
(3.39)
and at boundary y+∆y:
(τyx )| y+∆y ∆x∆z∆t
R R R R
(3.40)
and at boundary z+∆z::
(τzx )| z+∆z ∆x∆y∆t
R R R R (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
R R R R (3.42)
The net gravitational force in the x-direction is:
ρg| x ∆x ∆y∆z ∆t R R (3.43)
Accumulation is:
(ρv x | t+∆t − ρv x | t ) ∆x ∆y∆z
R R R R R R R R (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:
∂ ( ρv x ) ∂ ( ρv x v x ) ∂ ( ρv x v y ) ∂ ( ρv x v z ) ∂τ ∂τ ∂τ ∂P
+ + + = −( xx + yx + zx ) − + ρg x (3.45)
∂t ∂x ∂y ∂z ∂x ∂y ∂z ∂x

Expanding the partial derivative and rearranging:


 ∂ρ ∂ρv x ∂ρv y ∂ρv z   ∂v ∂v x ∂v x ∂v x 
vx  + + +  + ρ  x + v x + vy + vz  =
 ∂t ∂x ∂y ∂z   ∂t ∂x ∂y ∂z 
∂τ ∂τ yx ∂τ zx ∂P (3.46)
− ( xx + + )− + ρg
∂x ∂y ∂z ∂x x

Using the equation of continuity (Eq. 3.10) for incompressible fluid, Equation (3.46)
is reduced to:
 ∂v x ∂v ∂v ∂v  ∂τ ∂τ ∂τ ∂P
ρ + v x x + v y x + v z x  = −( xx + yx + zx ) − + ρg x (3.47)
 ∂t ∂x ∂y ∂z  ∂x ∂y ∂z ∂x

Using the assumption of Newtonian fluid, i.e.


∂v x ∂v x ∂v x
τ xx = − µ , τ yx = − µ , τ zx = − µ (3.48)
∂x ∂y ∂z

Equation 3.47 yields:


∂v x  ∂v ∂v ∂v   ∂ 2v ∂ 2v ∂ 2 v  ∂P
ρ + ρ  v x x + v y x + v z x  = µ  2x + 2x + 2x  − + ρg x
∂ t
   ∂x ∂y ∂z   ∂ x ∂ y ∂ z  ∂x (3.49)
accumulation
       
generation
transport by bulk flow transport by viscous forces

The momentum balances in the y-direction and z-direction can be obtained in a


similar fashion:
∂v y  ∂v y ∂v y ∂v y   ∂ 2 v y ∂ 2 v y ∂ 2 v y  ∂P
ρ + ρ  v x + vy + vz  = µ  2 + 2 + 2  − + ρg y

 
t ∂x ∂y ∂ z   ∂x ∂y ∂z   ∂
y  
(3.50)
accumulation
    
transport by bulk flow transport by viscous forces generation

∂v z  ∂v ∂v ∂v   ∂ 2v ∂ 2v ∂ 2 v  ∂P
ρ + ρ  v x z + v y z + v z z  = µ  2z + 2z + 2z  − + ρg z
 ∂
t  ∂x ∂y ∂z   ∂x ∂y ∂z   ∂
z   (3.51)
accumulation
      generation
transport by bulk flow transport by viscous forces

These equations constitute the Navier-Stock’s equation.

3.4 Energy balance


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:
(q x + ρCpTv x )∆y∆z∆t
R R R R
(3.53)
Similarly the energy entering the y and z directions are
(q y + ρCpTv y )∆x∆z∆t
R R R R
(3.54)
(q z + ρCpTv z ) ∆x ∆y ∆t
R R R R
(3.55)
The energy leaving the x,y and z directions are:
(q x + ρCpTv x )| x+∆x ∆y∆z∆t
R R R R R R
(3.56)
(q y + ρCpTv y )| y+∆y ∆x∆z∆t
R R R R R R
(3.57)
(q z + ρCpTv z )| z+∆z ∆x ∆y ∆t
R R R R R R
(3.58)
The energy accumulated is approximated by:
(ρCpT| t+∆t − ρCpT| t ) ∆x ∆y∆z R R R R
(3.59)
qy | y
qz |z+∆z

∆y

∆x
qx |x+∆x
qx |x
∆z

qy |y+∆y
qz |z

Figure 3.4 Energy Balance in Cartesian coordinates

The rate of generation is Φ H where Φ H includes all the sources of heat generation, i.e.
R R R R

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:
∂ ( ρCpT ) ∂ ( ρCpTv x ) ∂ ( ρCpTv y ) ∂ ( ρCpTvz ) ∂qx ∂q y ∂qz
+ + + + + + = ΦH (3.60)
∂t ∂x ∂y ∂z ∂x ∂y ∂z

Expanding the partial derivative yields:


 ∂T ∂T ∂T ∂T   ∂ρ ∂ρv x ∂ρv y ∂ρvz  ∂q x ∂q y ∂q z
ρCp + vx + vy + vz  + ρCpT  + + + +
 ∂x + ∂y + ∂z = Φ H (3.61)
 ∂t ∂x ∂y ∂z   ∂t ∂x ∂y ∂z 

Using the equation of continuity (Eq. 3.10) for incompressible fluids the equation is
reduced to:
 ∂T ∂T ∂T ∂T  ∂q x ∂q y ∂q z
ρCp + vx + vy + vz + + + = ΦH (3.62)
 ∂t ∂x ∂y ∂z  ∂x ∂y ∂z

Using Fourier's law:


dT
qu = −k (3.63)
du
into the last equation gives:
∂T  ∂T ∂T ∂T   ∂ 2T ∂ 2T ∂ 2T 
ρCp + ρCp vx + vy + vz  = k  2 + 2 + 2  + Φ H
∂ ∂x ∂y ∂z  
 t
   ∂x ∂y

∂z  generation

(3.64)
accumulation Transport by bulk flow Transport by thermal diffusion

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:
∂T  ∂ 2T ∂ 2T ∂ 2T 
ρCp = k  2 + 2 + 2  + Φ 
∂ ∂ ∂y ∂z  generation
H
 t x (3.65)
accumulation
  
Transport by thermal diffusion

You might also like