Sunteți pe pagina 1din 144

INTRODUCTION TO MATHEMATICAL MODELING OF CHEMICAL PROCESSES

Emad Ali, AbdelHamid Ajbar and Khalid Alhumaizi

King Saud University College of Engineering Chemical Engineering department

June 2006

Table of Contents 1. Chapter 1: Fundamentals 1.1. Introduction 1.2. Incentives for Process Modeling 1.3. Systems 1.3.1. Classification based on Thermodynamics 1.3.2. Classification based on number of phases 1.4. Classification of Models 1.5. State Variables and State Equations 1.6. Classification of theoretical models 1.6.1. Steady State vs Unsteady State 1.6.2. Lumped vs Distributed Parameters 1.6.3. Linear vs Non-Linear 1.6.4. Continuous vs Discrete 1.6.5. Deterministic vs Probabilistic 1.7. Building Steps for a Mathematical Model 1.8. Conservation Laws 1.8.1. Total Mass Balance 1.8.2. Component Balance 1.8.3. Momentum Balance 1.8.4. Energy Balance 1.9. Microscopic Balance 1.10. Macroscopic Balance 1.11. Transport rates 1.11.1. Mas Transport 1.11.2. Momentum Transport 1.11.3. Energy Transport 1.12. Thermodynamic relations 1.13. Phase Equilibrium 1.14. Chemical Kinetics 1.15. Control Laws 1.16. Degrees of Freedom 1.17. Model Solution Model Validation 1.18. 2. Chapter 2: Examples of Mathematical Models for Chemical processes 2.1. Examples of Lumped parameter Systems 2.1.1. Liquid Storage tank 2.1.2. Stirred Tank Heater 2.1.3. Isothermal CSTR 2.1.4. Gas-Phase Pressurized CSTR 2.1.5. Non-Isothermal CSTR 2.1.6. Mixing Processes 2.1.7. Heat Exchanger 2.1.8. Heat Exchanger with Steam 1 1 3 3 4 4 5 6 6 6 7 7 8 9 10 11 11 12 12 14 14 15 16 18 19 21 24 25 26 27 27 29 29 29 33 39 43 45 48 53 56

2.1.9. Single Stage Heterogeneous Systems 2.1.10. Two-Phase Reactor 2.1.11. Reaction with Mass Transfer 2.1.12. Multistage Heterogeneous Systems 2.1.13. Binary Absorption Column 2.1.14. Multi-Component Distillation Column 2.2. Examples of Distributed Parameter Systems 2.2.1. Liquid Flow in Pipe 2.2.2. Velocity profile in a Pipe 2.2.3. Diffusion with Chemical Reaction in a Slab catalyst 2.2.4. Temperature Profile in a heated Cylindrical Rod 2.2.5. Isothermal Plug Flow Reactor 2.2.6. Non-Isothermal Plug-Flow Reactor 2.2.7. Heat Exchanger: Distributed Parameter Model 2.2.8. Mass Exchange in a Packed Column 3. Chapter 3: Equations of Change 3.1. Total Mass Balance 3.2. Component Balance Equation 3.3. Momentum balance Equation 3.4. Energy balance 3.5. Conversion between the Coordinates 3.5.1. Balance Equations in Cartesian Coordinates 3.5.2. Balance equation in Cylindrical Coordinates 3.5.3. Balance Equation in Spherical Coordinates 3.6. Examples of Application of equations of Change 3.6.1. Liquid Flow in a Pipe 3.6.2. Diffusion with Chemical Reaction in a Slab Catalyst 3.6.3. Plug Flow Reactor 3.6.4. Energy Transport with Heat Generation 3.6.5. Energy Transport in a Circular Tube 3.6.6. Unsteady State Heat Generation 3.6.7. Laminar Flow Heat Transfer with Constant Wall Temperature 3.6.8. Laminar Flow and Mass Transfer Problems (Lumped Parameter Systems) Problems (Distributed Parameter Systems)

58 60 64 67 71 73 82 82 84 86 89 91 94 97 99 103 103 106 109 113 116 117 118 119 121 121 121 122 123 125 127 127 129 131 139

ii

Chapter 1 : Fundamentals

1.1 Introduction The objective of mathematical modeling is the development of sets of quantitative (mathematical) expressions that capture the essentials aspects of an existing system. A mathematical model can assist in understanding the complex physical interactions in the system and the causes and effects between the system variables .Mathematical models are valuable tools since they are abstract equations that can be solved and analyzed using computer calculations. It is therefore safer and cheaper to perform tests on the model using computer simulations rather than to carry out repetitive experimentations and observations on the real system. This becomes vital if the real system is new, hazardous, or expensive to operate . Modeling, thus prevails the field of science, engineering and business. It is used to assist in the design of equipment, to predict behavior, to interpret data, to optimize resources and to communicate information.

1.2 Incentives for process modeling In the chemical engineering field, models can be useful in all the phases, from research and development to plant operation. Models and their simulation are tools utilized by the chemical engineer to help him analyze the process in the following ways: Better understanding of the process Models can be used to study and investigate the effects of various process parameters and operating conditions on the process behavior. It can also be used to evaluate the interactions of different parts of the process. This analysis can be carried out easily on a computer simulation without interrupting the actual process, thus avoiding any delay or upsets for the process. Process synthesis and design

Model simulation can be utilized in the evaluation of equipment's size and arrangements and in the study of alternative process flow-sheeting and strategies. Furthermore, to verify the reliability and safety of the process design tests can be carried out even prior to plant commissioning. Plant operators training Models can be used to train plant personnel to simulate startup and shutdown procedures, to operate complex processes and to handle emergency situations and procedures . Controller design and tuning Models help in developing and evaluating better controller structure and configuration. Dynamic simulation of models is usually employed for testing and assessing the effectiveness of various controller algorithms. It is worthwhile to mention that models play a vital part in designing advanced model-based control algorithms such as model predictive and internal model controllers .Moreover it is a common practice of many control engineers to determine the optimum values of the controller settings through dynamic simulation . Process optimization It is desirable from economic standpoint to conduct process optimization before plant operation to determine the optimum values of the process key parameters or/and operating conditions that maximizes profit and reduces cost. Process optimization is also performed during process operation to account for variations in the feed-stock and utilities market and for changing environmental regulations. It is worth mentioning that despite all their usefulness, models at their best are no more than approximation of the real process since they do not necessarily incorporate all the features of the real system . Therefore modeling can not eliminate completely the need for some plant tests, especially to validate developed models or when some poorly known parameters in the process need to be experimentally evaluated. Models can be classified in a number of ways. But since mathematical models are developed from applying the fundamental physical and chemical laws on a specific system, 2

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

1.3.2 Classification based on number of phases Homogeneous system This is a system that involves only one phase such as gas-phase or liquid-phase chemical reaction processes. Heterogeneous system This is a system that involves more than one phase. This kind of systems exists in multi-phase reaction processes and in phase-based separation processes.

1.4 Classification of Models Models can be classified according to how they are derived: Theoretical models. These are models that are obtained from fundamental principles, such as the laws of conservation of mass, energy, and momentum along with other chemical principles such as chemical reaction kinetics and thermodynamic equilibrium, etc. This first-principle model is capable of explaining the underlying physics of the process and is often called `phenomenological model'. For this reason it is particularly suitable for process design and optimization . Theoretical models are, however, generally difficult to obtain and sometimes hard to solve.

Empirical models These models are based on experimental plant data. These models are developed using data fitting techniques such as linear and non-linear regression. Models obtained exclusively form experimental plant data are also known as black-box models . Such models do not provide detailed description of the underlying physics of the process. However, they do provide a description of the dynamic relationship between inputs and outputs. Thus they are sometimes more adequate for control design and implementation. Semi-empirical models These models are somehow between the two previous models where uncertain or poorly known process parameters are determined from plant data. This book focuses only on developing theoretical models. The interested reader in empirical modeling is referred to books listed in the references.

1.5 State variables and state equations Once the system has been classified, developing a theoretical model for it amounts to characterizing its behavior at any time and at any spatial position. For most processing systems a number fundamental quantities are used to describe the natural state of the system. These quantities are the mass, energy and momentum. Most often these fundamental quantities can not be measured directly thus they are usually represented by other variables that can be measured directly and conveniently. The most common variables are density, concentration, temperature pressure and flow rate. There are conveniently called 'state variables' since they characterize the state of the processing system. In order to describe the behavior of the system with time and position, the state (dependent) variables should be linked to the independent variables (time, spatial position) through sets of equations that are derived from writing mass, energy and momentum balances. The set of equations describing these variables are called 'state equations.'

1.6 Classification of theoretical models With this in mind, theoretical models may be further classified in more practical ways as discussed in the following.

1.6.1 Steady state Vs. unsteady state When the physical state of the processing system remains constant with the time, the system is said to be at steady state. Models that describe steady state situations are also called static, time-invariant or stationary models. Basically, almost all chemical process unit designs are carried out on static models. On the other hand unsteady state processes represent the situation when the process state (dependent variables) changes with time. Models that describe unsteady-state situations are also called dynamic and transient models .Such models are useful for process control design and development. Process dynamics are encountered in practice during startup, shutdown, and upsets (disturbances )

1.6.2 Lumped Vs. distributed parameters Lumped parameters models are those in which the state variables and other parameters have/or assumed to have no spatial dependence, i.e. they are considered to be uniform over the entire system. In this case the time (for unsteady state models) is the only independent variable. The chemical engineering examples for this case include the perfectly mixed CSTR, distillation columns. etc. Conceptually, these models are obtained through carrying out a macroscopic balance for the process as it will be discussed in chapter 2 . On the other hand, distributed parameters models are those in which states and other variables are function of both time and spatial position. In this case, modeling takes into account the variation of these variables with time and from point to point throughout the entire system. Some examples of such systems include plug flow reactor, heat exchangers, and packed columns. These models are essentially obtained through writing microscopic balances equations for the process as it will be discussed later in chapter 2. 6

1.6.3 Linear Vs non-linear Linear models have the important property of superposition whereas nonlinear models do not. Superposition means that the response of the system to a sum of inputs is the same as the sum of responses to the individual inputs .In linear models all the dependant variables or their derivatives appear in the model equations only to the first power. These properties do not hold for nonlinear models . In this respect, it is important to recognize the fact that most physical and chemical systems are nonlinear. Linear models are commonly obtained through linearization of the nonlinear model around a certain steady state. Linear models obtained this way are valid approximation of the original non linear model only in the neighborhood of the selected state .

1.6.4 Continuous Vs discrete When the dependant variables can assume any values within an interval the model is called continuous .When on the other hand one or several variables are assumed to take only discrete values, the model is called discrete. In chemical engineering discrete models arise for example when some variables are required to take only integer values, for example the number of stages in a distillation column, the number of heat exchanger in a plant, etc.

1.6.5 Deterministic Vs probabilistic Deterministic models are those in which each variable can be assigned a definite fixed number, for a given set of conditions. On the other hand in probabilistic or stochastic models some or all the variables used to describe the system are not precisely known. They are considered as random variables. Probabilistic models are often encountered when modeling systems that are subject to noise.

1.7 Building steps for a mathematical model Building a theoretical mathematical model of a processing system requires the knowledge of the physical and chemical interactions taking place within the boundaries of the system. With a given degree of fundamental knowledge of the system at a certain stage one can build different models with different degree of complexity depending on the purpose of the model building and the level of rigor and accuracy required .The choice of the level of rigor and degree of sophistication is in itself an art that requires much experience. Theoretical modeling of chemical processes may encounter difficulties that can be classified as follows: Problems arising from processes that exhibit complicated physical and chemical phenomena. Such problems appear for instance in multi-component interactive processes . Problems arising from imprecisely known process parameters. This situation can be handled by periodically estimating the unknown parameters from plant data. Problems arising from the size and complexity of the ensuing model. These can be overcomed by proper use of simplifying assumptions. Consequently a modeler should practice careful utilization of the simplifying assumptions based on engineering sense and experience. Failure to do so, the modeler may fall into one of the two extremes, i.e. creating rigorous but over complicated model or creating oversimplified model that does not capture all the critical features of the true process. The general procedure for building up a mathematical model includes the following steps:

Identification of the system configuration, its surrounding environment and the modes of interaction between them, the 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. identification of the relevant state variables that describe the system and identification of the process taking place

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:

Sin System

Sout

Figure 1.2

Total into system

flow the

Generation of within system

Total

flow

Accumulation of within system

amount of (S) the surrounding

rate of (S) + rate

(S) = rate of (S) out + rate of the system

(S) exchanged with (1.1)

The quantity S can be any one of the following quantities: Total Mass Component Mass (Mole) Total Energy Momentum

1.8.1 Total mass balance Since mass is always conserved, the balance equation for the total mass (m) of a given system is: Rate of mass in = rate of mass out + rate of mass accumulation (1.2)

10

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

1.8.2 Component balance The mass balance for a component A is generally written in terms of number of moles of A. Thus the component balance is flow of moles (A) in Rate of Generation + of moles of (A) Flow of moles = of (A) out Rate of Accumulation + of moles of (A)

(1.3)

The component balance has the unit of moles A/s. It should be noted that unlike the total mass, the number of moles of species A is not conserved. The species A can be generated or consumed by chemical reaction.

1.8.3 Momentum balance The linear momentum () of a mass (m) moving with velocity (v) is defined as: = 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 momentum in of Rate of Generation + of momentum Rate of Rate of Accumulation (1.5)

= momentum out + of momentum

The momentum balance has the unit of kg.m/s2. The momentum balance equation is usually written using the Newton's second law. The law states that the time rate of change of momentum of a system is equal to the sum of all forces F acting on the system,

11

d (mv) = F dt
1.8.4 Energy balance

(1.6)

The energy balance for a given system is: Rate of energy in Rate of + Generation of = energy Rate of energy out Rate of + accumulation of energy
amount of energy exchanged with the surrounding

(1.7)

The energy generated within a system includes the rate of heat and the rate of work. The rate of heat includes the heat of reaction (if a reaction occurs in the system), and the heat exchanged with the surroundings. For the rate of work we will distinguish between the work done against pressure forces (flow work) and the other work such as the work done against the gravity force, against viscous forces and shaft work. The reason for this distinction will appear clearly in the next chapter. For a given system the general conservation law (Eq. 1.1) can be carried out either on microscopic scale or macroscopic scale.

1.9 Microscopic balance

In the microscopic case, the balance equation is written over a differential element within the system to account for the variation of the state variables from point to point in the system, besides its variation with time. If we choose for example cartesian coordinates, the differential element is a cube as shown in Figure 1.3. Each state variable V of the system is assumed to depend on the three coordinates x,y and z plus the time. i.e. V = V(x,y,z,t). The microscopic balance can be also written in cylindrical coordinates (Figure 1.4) and in spherical coordinates (Figure 1.5). The selection of the appropriate coordinates depends on the geometry of the system under

12

study .It is possible to convert from one coordinate system to an other as it will be discussed in chapter 3.

y z y x x

Figure 1.3: Cartesian coordinates

Figure 1.4: Cylindrical coordinates

13

z
y

Figure 1.5: Spherical coordinates

1.10 Macroscopic balance

In some cases the process state variables are uniform over the entire system, that is each state variable does not depend on the spatial variables, i.e. x,y and z in cartesian coordinates but only on time t. In this case the balance equation is written over the whole system using macroscopic modeling . When modeling the process on microscopic scale the resulting models consists usually of partial differential equation (PDE) where time and one or more spatial position are the independent variables. At steady state, the PDE becomes independent of t and the spatial positions are the only independent variables .When, on the other hand, the modeling is based on macroscopic scale the resulting model consists of sets of ordinary differential equations (ODE) . In the next chapter we present examples of applying macroscopic and microscopic balances to model various chemical processes. The fundamental balance equations of mass, momentum and energy already discussed are usually supplemented with a number of transport rates and thermodynamic overview of some of these relations. equations associated with relationships. In the following we present an

1.11 Transport rates

14

Transport of the fundamental quantities, mass, energy and momentum occur by two mechanisms Transport due to convection or bulk flow Transport due to molecular diffusion or potential difference . In many cases the two transport mechanism occur together. Therefore, the flux due to the transport of any fundamental quantity is the sum of a flux due to convection and a flux due to diffusion.

1.11.1 Mass Transport

The total flux nAu (kg/m2s) of species A of density A (kg/m3) flowing with velocity vu (m/s) in the u-direction is the sum of the two terms: nAu = jAu + Avu total flux = diffusive flux + bulk flux The diffusive flux jAu (kg/m2s) for a binary mixture A-B is given by Fick's law:
dw A du

(1.8) (1.9)

j Au = D AB

(1.10)

where wA = A/ is the mass fraction of species , (kg/m3) the density of the mixture and DAB (m2/s) is the diffusivity coefficient of A in the mixture. In molar unit the flux
JAu (mol A/m2s) is given by:
dx A du

J Au = CD AB

(1.11)

15

where C is the total concentration of A and B (Kg(A+B)/m3) and xA = CA/C is the mole fraction of A in the mixture. For constant density the flux Eqs.( 1.10 and Eq. 1.11) become:
d A du

j Au = DAB

(1.12)

J Au = D AB

dC A du

(1.13)

1.11.2 Momentum transport

Momentum is also transported by convection and diffusion. But unlike the mass, the linear momentum = 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 xcomponent 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 (vx)vy (kg.m/s2) (1.14)

To determine the diffusion flux denoted by yx of the x-component of the momentum in the y-direction, consider a fluid flowing between two infinite parallel plates as shown in Figure 1.6. At a certain time the lower plate is moved by applying a constant force Fx while the upper plate is maintained constant. The force Fx is called a shear force since it is tangential to the area Ay on which it is applied (Fig. 1.6).

16

y y x F, force Vx

Figure 1.6: Momentum transfer between two parallel plates

The force per unit area


Fx Ay (1.15)

(kg.m / s 2 )

is called a stress and denoted yx (kg.m/s). It is also a shear stress since it is tangential. The force Fx imparts a constant velocity V = vx (y = 0 ) to the layer adjacent to the plate. Because of molecular transport, the layer above it has a slightly slower velocity vx(y) and so on as shown in Figure 1.6. Therefore, there is a transport by diffusion of the xcomponent of the momentum in the y-direction. The flux of this diffusive transport is in fact the shear stress yx. Therefore, the total flux yx of the x-component in the y-direction is the sum of the convection term (Eq. 1.14) and diffusive term yx

yx = yx + (vx)vy
momentum flux = diffusive flux + bulk flux For a Newtonian fluid the shear stress yx is proportional to the velocity gradient: v x y

(1.16) (1.17)

yx =

(1.18)

17

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

1.11.3 Energy transport


The total energy flux eu (J/s.m2) of a fluid at constant pressure flowing with a velocity vu in the u-direction can be expressed as: eu = qu + (CpT)vu energy flux = diffusive flux + bulk flux (1.19) (1.20)

The heat flux by molecular diffusion, i.e. conduction in the u-direction is given by Fourier's law: qu = k T u (1.21)

where k (J/s.m.K) is the thermal conductivity. The three relations (Eq. 1.10, 1.18, 1.21) show the analogy that exists between mass, momentum and energy transport. The diffusive flux in each case is given by the following form: Flux = transport property (gradient) The flux represents the rate of transfer per area, the potential difference indicates the driving force and the transport property is the proportionality constant . Table 1.1 summarizes the transport laws for molecular diffusion. Table 1.1: One-dimensional Transport laws for molecular diffusion Transport Type Law Flux Transport property gradient potential difference (1.22)

18

Mass Heat Momentum

Fick's Fourrier Newton

JAu qu

D k

dC A du

dT du
dv x du

ux

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
where K an overall mass transfer coefficient. The heat flux on the other hand is expressed as

(1.23)

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 2v 2 L

(1.25)

f is the Fanning friction factor, P is the pressure drop due to friction, D is the diameter
of the pipe, L the length and v is the velocity of the fluid.

1.12 Thermodynamic relations

19

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

(1.26)

= MwP/RT

(1.27)

where Mw is the molecular weight . As for liquids, tabulated values of density can be used and can be considered invariant unless large changes in composition or temperature occur.

Enthalpies :
Liquid and vapor enthalpies for pure component can be computed from simple formulas, based on neglecting the pressure effect as follows: ~ h = C p (T Tref )
~ H = C p (T Tref ) +

(1.28)

(1.29)

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

the average liquid heat capacity and is the latent heat of vaporization Note that the reference condition is taken to be liquid at temperature Tref. If heat of mixing is negligible, the enthalpy of a mixture can be taken as the sum of the specific enthalpies of the pure components multiplied by their corresponding

20

mole fractions Note that the above enthalpy functions are valid for small temperature variation and/or when the heat capacities of fluids are weak function of temperature. In general cases the heat capacity Cp can be taken as function of temperature such as:
Cp = a + bT + cT2 + dT3

(1.30)

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

~ h=

Tref

C p dT
T

(1.31)

~ H=

Tref

C p dT +

(1.32)

Internal energy : The internal energy U is a fundamental quantity that appears in energy balance equation. For liquids and solids the internal energy can be approximated by enthalpy .This can be also a good approximation for gases if the pressure change is small.

1.13 Phase Equilibrium

A large number of chemical processes involve more than one phase. In many cases the phases are brought in direct contact with each other such as in packed or tray towers. When the transfer of either mass or energy occurs from a fluid phase to another phase the interface between fluid phases is usually at equilibrium. In the case of heat transfer between phase I and Phase II, the equilibrium dictates that the temperature is the same at the interface. This is not the case for mass transfer. Figure 1.7 shows for instant the mass transfer of species A from a liquid to a gas phase. The concentration of

21

the bulk gas phase yAG decreases at the interface. The liquid concentration increases, on the other hand, from xAL to xAi. At the interface, an equilibrium exists and yAi and xAi are related by a relation of the form of yAi = F(xAi) (1.33)

The equilibrium relations are in general nonlinear. However, quite satisfactory results can be obtained through the use of simpler relations that are derived form assumptions of ideal behavior of the two phases:

Gas-phase mixture of A in gas G yAG NA yAi

Liquid -phase solution of A in liquid L

xAi

xAL

Interface

Figure 1.7: Equlibrium at the interface

For vapor phases at low concentration, Henry's law provides the following
simple equilibrium relation: pA = HxA 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: (1.34)

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

~ yA = H xA

(1.35)

22

~ 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, PAs is the vapor pressure of pure A at the temperature of the system, and P is the total pressure on gas-phase side. The dependence of vapor pressure PAs on temperature can be approximated by the Antoine equation

ln( PAS ) = A

B C +T

(1.37)

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

Raoults law can be modified to account for non-ideal liquid and vapor behavior
using the activity coefficients i and i of component (i) for liquid and vapor phase respectively. The following equilibrium relation, also known as the Gamma/Phi formulation, can be used:
y ii P = xi i Pi S

(1.38)

23

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. ( i = 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:
yA = KxA

(1.39)

where K is the equilibrium distribution coefficient that depends on pressure, temperature and concentration. Phase equilibrium relations are commonly used in the following calculations:
Bubble point calculations :

For a given molar liquid compositions (xi) and either T or P, the bubble point calculations consist in finding the molar vapor composition (yi) and either P or
T. Dew point calculations :

For a given molar vapor compositions (yi) and either T or P, the dew point calculations consist in finding the molar liquid compositions (xi) and either P or
T. Flash calculation :

For known mixture compositions (zi) and known (T) and (P), the calculations consist in simultaneous solution of component balance and energy balance equations.

flash

finding the liquid and vapor compositions via

1.14 Chemical kinetics

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

24

R=

1 dni viV dt

(1.40)

where
ni i V

the number of moles the stochiometric coefficient of component i 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
r1R1 + r2R2 + + rrRrr p1P1 + p2P2 + + ppRpp

(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 = kCR1 C R 2 LC R r
1 2

(1.42)

The powers i are determined experimentally and their values are not necessarily integers. The temperature dependence comes from the reaction rate constant k given by Arrhenius law
E e RT

k = ko

(1.43)

where

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

temperature and R is the ideal gas constant

1.15 Control Laws

Although the discussion of feedback control systems is beyond the scope of this book, the existence of control loops in any processing system provides extra relations

25

for the process. Specifically in many cases some key process variables are required to be controlled, i.e. maintained within desired range. This control objective can be achieved by closing the control loop, i.e., relating the controlled variables to inputs (manipulated variables). This introduces additional independent equations.

1.16 Degrees of Freedom

A key step in the model development and solution is checking its consistency or solvability, i.e. the existence of exact solution. This is done by checking the degrees of freedom of the model after the equations have written and before attempting to solve them . For a processing system described by a set of Ne independent equations and Nv variables, the degree of freedom f is
F = Nv Ne

(1.44)

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


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

balance equation has a finite number of solutions (one solution for linear systems)
f < 0. The system is

over-determined (over-specified)

by f equations.

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. 26

2. Determine other variables that can specified by the external world, for example, variables that are the outcome of an upstream processing units, and/or variables that can be used as forcing function or manipulated variables.

1.17 Model solution

After the solvability of the model has been checked, the next step is to solve the model. The purpose of the solution of the model is be able to obtain the variations of the state variables with the model independent variables (time, spatial positions..). The solution of the model permits also a parametric investigation of the model, that is a study of the effects of the changing the value of some parameters. It would be ideal to be able to solve the model analytically, that is to get closed forms of the state variables in term of the independent variables. Unfortunately this seldom occurs for chemical processes. The reason is that the vast majority of chemical processes are nonlinear. They may be a sets of nonlinear partial differential equations (PDE) as it is the case for distributed parameter models, or sets of nonlinear ordinary differential equations (ODE) or nonlinear algebraic equations as it is the case for lumped parameter models. However most non linear problems can not be solved analytically. In fact the only class of differential equations for which there is a well-developed framework are linear ODE . However linearizing the original nonlinear model and solving it is not always recommended (expect for control purposes) since the behavior of the linearized model matches the original nonlinear model only around the state chosen for linearization .For these reasons the solution of process models is usually carried out numerically, most often through a computer programs.

1.18 Model validation

Model verification (validation) is the last and the most important step of model building. Reliability of the obtained model depends heavily on faithfully passing this test. Implementation of the model without validation may lead to erroneous and 27

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.

28

Chapter 2: Examples of Mathematical Models for Chemical Processes


In this chapter we develop mathematical models for a number of elementary chemical processes that are commonly encountered in practice. We will apply the methodology discussed in the previous chapter to guide the reader through various examples. The goal is to give the reader a methodology to tackle more complicated processes that are not covered in this chapter and that can be found in books listed in the reference. The organization of this chapter includes examples of systems that can be described by ordinary differential equations (ODE), i.e. lumped parameter systems followed by examples of distributed parameters systems, i..e those described by partial differential equations (PDE). The examples cover both homogeneous and heterogeneous systems. Ordinary differential equations (ODE) are easier to solve and are reduced to simple algebraic equations at steady state. The solution of partial differential equations (PDE) on the other hand is a more difficult task. But we will be interested in the cases were PDE's are reduced to ODE's. This is naturally the case where under appropriate assumptions, the PDE's is a one-dimensional equation at steady state conditions. It is worth to recall, as noted in the previous chapters, that the distinction between lumped and distributed parameter models depends sometimes on the assumptions put forward by the modeler. Systems that are normally distributed parameter can be modeled under appropriate assumptions as lumped parameter systems. This chapter includes some examples of this situation.

29

2.1 Examples of Lumped Parameter Systems

2.1.1 Liquid Storage Tank

Consider the perfectly mixed storage tank shown in figure 2.1. Liquid stream with volumetric rate Ff (m3/s) and density f flow into the tank. The outlet stream has volumetric rate Fo and density . Our objective is to develop a model for the variations of the tank holdup, i.e. volume of the tank. The system is therefore the liquid in the tank. We will assume that it is perfectly mixed and that the density of the effluent is the same as that of tank content. We will also assume that the tank is isothermal, i.e. no variations in the temperature. To model the tank we need only to write a mass balance equation.

Figure 2-1 Liquid Storage Tank

Since the system is perfectly mixed, the system properties do not vary with position inside the tank. The only variations are with time. The mass balance equation can be written then on the whole system and not only on a differential element of it. This leads to therefore to a macroscopic model. We apply the general balance equation (Eq. 1.2), to the total mass m = V. This yields: Mass flow in:
f Ff

(2.1)

30

Mass flow out:


o Fo

(2.2)

Accumulation:
dm d (V ) = dt dt

(2.3)

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

f F f = o Fo +

d (V ) 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 ti, i.e.

V(ti) = Vi

(2.5)

Under isothermal conditions we can further assume that the density of the liquid is constant i.e. f = o=. In this case Eq. 2.4 is reduced to:

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)

31

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

dL = F f Fo dt

(2.8)

with initial condition:

L(ti) = Li

(2.9)

Degree of freedom analysis


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

Parameter of constant values: A Variables which values can be externally fixed (Forced variable): Ff Remaining variables: L and Fo Number of equations: 1 (Eq. 2.8)
Therefore the degree of freedom is: Number of remaining variables Number of equations = 2 1 = 1 For the system to be exactly specified we need therefore one more equations. This extra relation is obtained from practical engineering considerations. If the system is operated without control (at open loop) then the outlet flow rate Fo is a function of the liquid level L. Generally a relation of the form:
Fo = L

(2.10)

could be used, where is the discharge coefficient. If on the other hand the liquid level is under control, then its value is kept constant at certain desired value Ls. If Fo is used to control the height then a control law relates Fo to L and Ls:

32

Fo = Fo(L,Ls)
For instant, if a proportional controller Kc is used then the control law is given by:

(2.11)

Fo = Kc(L Ls) + Fob

(2.12)

Where Fob the bias, i.e. the constant value of Fo when the level is at the desired value i.e., L = Ls. Note that at steady state, the accumulation term is zero (height does not change with time), i.e., dL/dt = 0. The model of the tank is reduced to the simple algebraic equation:

F0 = Ff

(2.13)

2.1.2 Stirred Tank Heater


We consider the liquid tank of the last example but at non-isothermal conditions. The liquid enters the tank with a flow rate Ff (m3/s), density f (kg/m3) and temperature

Tf (K). It is heated with an external heat supply of temperature Tst (K), assumed
constant. The effluent stream is of flow rate Fo (m3/s), density o (kg/m3) and temperature T(K) (Fig. 2.2). Our objective is to model both the variation of liquid level and its temperature. As in the previous example we carry out a macroscopic model over the whole system. Assuming that the variations of temperature are not as large as to affect the density then the mass balance of Eq. 2.8 remains valid. To describe the variations of the temperature we need to write an energy balance equation. In the following we develop the energy balance for any macroscopic system (Fig. 2.3) and then we apply it to our example of stirred tank heater. The energy E(J) of any system of (Fig. 2.3) is the sum of its internal U(J), kinetic K(J) and potential energy (J):

E=U+K+
Consequently, the flow of energy into the system is:

(2.14)

33

~ ~ ~ f Ff ( U f + K f + f ) where the ( ~ ) denotes the specific energy (J/kg).

(2.15)

Figure 2-2 Stirred Tank Heater

Figure 2-3 General Macroscopic System


The flow of energy out of the system is:
~ ~ ~ o Fo( U o + K o + o )

(2.16)

The rate of accumulation of energy is:

34

~ ~ ~ d V (U + K + ) dt

(2.17)

As for the rate of generation of energy, it was mentioned in Section 1.8.4, that the energy exchanged between the system and the surroundings may include heat of reaction Qr (J/s), heat exchanged with surroundings Qe (J/s) and the rate of work done against pressure forces (flow work) Wpv (J/s), in addition to any other work Wo. The flow of work Wpv done by the system is given by:

W pv = Fo Po F f Pf

(2.18)

where Po and Pf are the inlet and outlet pressure, respectively. In this case, the rate of energy generation is:

Q e + Q r (W o + Fo Po F f P f )
Substituting all these terms in the general balance equation (Eq. 1.7) yields: ~ ~ ~ d V (U + K + ) ~ ~ ~ ~ ~ ~ = f F f U f + K f + f o Fo U o + K o + o dt + Qe + Qr (Wo + Fo Po F f Pf )

(2.19)

(2.20)

We can check that all terms of this equation have the SI unit of (J/s). Equation (2.20) can be also written as: ~ ~ ~ d V (U + K + ) ~ ~ ~ ~ ~ ~ = f F f U f + K f + f o Fo U o + K o + o dt Pf P + Qe + Qr Wo Foo o + F f f o f

)
(2.21)

35

~ The term V = 1 / is the specific volume (m3/kg). Thus Eq. 2.21 can be written as:
~ ~ ~ d V (U + K + ) ~ ~ ~ ~ ~ ~ ~ ~ = f F f U f + Pf V f + K f + f o Fo U o + PoVo + K o + o dt + Qe + Q r W o

(2.22)

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

the general energy balance equation for a macroscopic system can be written as: ~ ~ ~ ~ ~ ~ d V (U + K + ) ~ ~ ~ = f F f h f + K f + f o Fo ho + K o + o + Qe + Qr Wo dt

(2.23)

We return now to the liquid stirred tank heater. A number of simplifying assumptions can be introduced: We can neglect kinetic energy unless the flow velocities are high. We can neglect the potential energy unless the flow difference between the inlet and outlet elevation is large. All the work other than flow work is neglected, i.e. Wo = 0. There is no reaction involved, i.e. Qr = 0. The energy balance (Eq. 2.23) is reduced to: ~ d VU ~ ~ = f F f h f o Fo ho + Qe dt

(2.24)

Here Qe is the heat (J/s) supplied by the external source. Furthermore, as mentioned in ~ ~ Section 1.12, the internal energy U for liquids can be approximated by enthalpy, h . The enthalpy is generally a function of temperature, pressure and composition. However, it can be safely estimated from heat capacity relations as follows:

36

~ ~ h = Cp(T Tref ) ~ where Cp is the average heat capacity.

(2.25)

Furthermore since the tank is well mixed the effluent temperature To is equal to process temperature T. The energy balance equation can be written, assuming constant density f = o = , as follows: ~ d V (T Tref ) ~ ~ C p = F f C p (T f Tref ) FoC p (T Tref ) + Qe dt Taking Tref = 0 for simplicity and since V = AL result in: ~ d (LT ) ~ ~ C p A = F f C pT f FoC pT + Qe dt or equivalently: Qe d (LT ) = F f T f FoT + ~ dt C p (2.28) (2.27)

(2.26)

A Since

d (T ) d (L ) d (LT ) = AT + AL dt dt dt

(2.29)

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


Qe dT + T ( F f Fo ) = F f T f FoT + ~ dt C p (2.30)

AL

or equivalently:

37

AL

dT Qe = F f (T f T ) + ~ dt C p

(2.31)

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

dL = F f Fo dt

(2.32)

AL

dT Qe = F f (T f T ) + ~ dt C p

(2.33)

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

L(ti) = Li

and T(ti) = Ti

(2.34)

Degree of freedoms analysis


For this system we can make the following simple analysis:

Parameter of constant values: A, and Cp (Forced variable): Ff and Tf Remaining variables: L, Fo, T, Qe Number of equations: 2 (Eq. 2.32 and Eq. 2.33)
The degree of freedom is therefore, 4 2 = 2. We still need two relations for our problem to be exactly specified. Similarly to the previous example, if the system is operated without control then Fo is related to L through (Eq. 2.10). One additional relation is obtained from the heat transfer relation that specifies the amount of heat supplied:

Qe = UAH (TstT )

(2.35)

38

U and AH are heat transfer coefficient and heat transfer area. The source temperature Tst
was assumed to be known. If on the other hand both the height and temperature are under control, i.e. kept constant at desired values of Ls and Ts then there are two control laws that relate respectively Fo to L and Ls and Qe to T and Ts:

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

(2.36)

2.1.3 Isothermal CSTR


We revisit the perfectly mixed tank of the first example but where a liquid phase chemical reactions taking place:
k

A B

(2.37)

The reaction is assumed to be irreversible and of first order. As shown in figure 2.4, the feed enters the reactor with volumetric rate Ff (m3/s), density f (kg/m3) and concentration CAf (mole/m3). The output comes out of the reactor at volumetric rate Fo, density 0 and concentration CAo (mole/m3) and CBo (mole/m3). We assume isothermal conditions. Our objective is to develop a model for the variation of the volume of the reactor and the concentration of species A and B. The assumptions of example 2.1.1 still hold and the total mass balance equation (Eq. 2.6) is therefore unchanged

Figure 2.4 Isothermal CSTR

39

The component balance on species A is obtained by the application of (Eq. 1.3) to the number of moles (nA = CAV ). Since the system is well mixed the effluent concentration CAo and CBo are equal to the process concentration CA and CB. Flow of moles of A in:

Ff CAf
Flow of moles of A out:

(2.38)

Fo CAo
Rate of accumulation:

(2.39)

dn d (VC A ) = dt dt

(2.40)

Rate of generation:

-rV

where r (moles/m3s) is the rate of reaction. Substituting these terms in the general equation (Eq. 1.3) yields:

d (VC A ) = F f C Af Fo C A rV dt
We can check that all terms in the equation have the unit (mole/s).

(2.41)

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.

40

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

d (V ) d (C A ) d (VC A ) =V + CA = F f C Af Fo C A rV dt dt dt
Substituting Equation (2.6) into (2.42) and with some algebraic manipulations we obtain:

(2.42)

d (C A ) = F f ( C Af C A ) rV dt

(2.43)

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

r = k CA

(2.44)

Equations 2.6 and 2.43 define the dynamic behavior of the reactor. They can be solved if the system is exactly specified and if the initial conditions are given:

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


Degrees of freedom analysis

(2.45)

Parameter of constant values: A (Forced variable): Ff and CAf Remaining variables: V, Fo, and CA Number of equations: 2 (Eq. 2.6 and Eq. 2.43)
The degree of freedom is therefore 3 2 =1. The extra relation is obtained by the relation between the effluent flow Fo and the level in open loop operation (Eq. 2.10) or in closed loop operation (Eq. 2.11). The steady state behavior can be simply obtained by setting the accumulation terms to zero. Equation 2.6 and 2.43 become:

41

F0 = Ff F f ( C Af C A ) = rV
More complex situations can also be modeled in the same fashion. catalytic hydrogenation of ethylene:

(2.46) (2.47)

Consider the

A+B

(2.48)

where A represents hydrogen, B represents ethylene and P is the product (ethane). The reaction takes place in the CSTR shown in figure 2.5. Two streams are feeding the reactor. One concentrated feed with flow rate F1 (m3/s) and concentration CB1 (mole/m3) and another dilute stream with flow rate F2 (m3/s) and concentration CB2 (mole/m3). The effluent has flow rate Fo (m3/s) and concentration CB (mole/m3). The reactant A is assumed to be in excess.

F1, CB1

F2, CB2

V
Fo, CB
Figure 2-5 Reaction in a CSTR

The reaction rate is assumed to be:

r=

k1CB (1 + k2CB ) 2

( mole / m 3 .s )

(2.49)

42

where k1 is the reaction rate constant and k2 is the adsorption equilibrium constant. Assuming the operation to be isothermal and the density is constant, and following the same procedure of the previous example we get the following model: Total mass balance: dL = F1 + F2 Fo dt

(2.50)

Component B balance: d (C A ) = F1( CB1 CB ) + F2 ( CB 2 CB ) rV dt

(2.51)

Degrees of freedom analysis


Parameter of constant values: A, k1 and k1 (Forced variable): F1 F2 CB1 and CB2 Remaining variables: V, Fo, and CB Number of equations: 2 (Eq. 2.50 and Eq. 2.51) The degree of freedom is therefore 3 2 =1. The extra relation is between the effluent flow Fo and the level L as in the previous example.

2.1.4 Gas-Phase Pressurized CSTR


So far we have considered only liquid-phase reaction where density can be taken constant. To illustrate the effect of gas-phase chemical reaction on mass balance equation, we consider the following elementary reversible reaction:

A 2B

(2.52)

taking place in perfectly mixed vessel sketched in figure 2.6. The influent to the vessel has volumetric rate Ff (m3/s), density f (kg/m3), and mole fraction yf. Product comes out of the reactor with volumetric rate Fo, density o, and mole fraction yo. The temperature

43

and volume inside the vessel are constant. The reactor effluent passes through control valve which regulate the gas pressure at constant pressure Pg.

Ff,

f , yf

P, T, V ,y

F o,

yo

Pg

Figure 2-6 Gas Pressurized Reactor

Writing the macroscopic total mass balance around the vessel gives: d ( V ) = f F f o Fo dt

(2.53)

Since V is constant we have: d = f F f o Fo dt

(2.54)

Writing the component balance, for fixed V, results in: dC A = F f C Af FoC A0 r1V + r2V dt (2.55)

The reaction rates for the reversible reaction are assumed to be: r1 = k1 CA r2 = k2CB
2

(2.56) (2.57)

Equations (2.54) and (2.55) define the variations of density and molar concentration. One can also rewrite the equation to define the behavior of the pressure (P) and mole

44

fraction (y). The concentration can be expressed in term of the density through ideal gas law: CA = yP/RT CB = (1 y)P/RT Similarly, the density can be related to the pressure using ideal gas law: = MP/RT = [MAy + MB (1 y)]P/RT (2.60) (2.58) (2.59)

Where MA and MB are the molecular weight of A and B respectively. Therefore one can substitute equations (2.58) to (2.60) into equations (2.54 & 2.55) in order to explicitly write the latter two equations in terms of y and P. Or, alternatively, one can solve all equations simultaneously.

Degrees of freedom analysis:

Parameters: V, k1, k2, R, T, MA and MB Forcing function: Ff, CAf, yf Variables: CA, CB, y, P, , F Number of equations: 5 (Eqs. 2.54, 2.55, 2.58, 2.59, 2.60) The degree of freedom is therefore 6 5 =1. The extra relation relates the outlet flow to the pressure as follows:
P Pg

Fo = Cv

(2.61)

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

2.1.5 Non-Isothermal CSTR

We reconsider the previous CSTR example (Sec 2.1.3), but for non-isothermal conditions. The reaction A B is exothermic and the heat generated in the reactor is

45

removed via a cooling system as shown in figure 2.7. The effluent temperature is different from the inlet temperature due to heat generation by the exothermic reaction.

Figure 2-7 Non-isothermal CSTR

Assuming constant density, the macroscopic total mass balance (Eq. 2.6) and mass component balance (Eq. 2.43) remain the same as before. However, one more ODE will be produced from the applying the conservation law (equation 2.23) for total energy balance. The dependence of the rate constant on the temperature: k = koe-E/RT should be emphasized. The general energy balance (Eq. 2.23) for macroscopic systems applied to the CSTR yields, assuming constant density and average heat capacity: ~ d V (T Tref ) ~ ~ C p = F f C p (T f Tref ) FoC p (T Tref ) + Qr Qe dt (2.62)

(2.63)

where Qr (J/s) is the heat generated by the reaction, and Qe (J/s) the rate of heat removed by the cooling system. Assuming Tref = 0 for simplicity and using the differentiation principles, equation 2.63 can be written as follows: ~ dT ~ dV ~ ~ C pV + C pT = F f C pT f FoC pT + Qr Qe dt dt (2.64)

46

Substituting Equation 2.6 into the last equation and rearranging yields: ~ dT ~ C pV = F f C p (T f T ) + Qr Qe dt The rate of heat exchanged Qr due to reaction is given by: Qr = (Hr)Vr (2.66) (2.65)

where Hr (J/mole) is the heat of reaction (has negative value for exothermic reaction and positive value for endothermic reaction). The non-isothermal CSTR is therefore modeled by three ODE's: dV = F f Fo dt d (C A ) = F f ( C Af C A ) rV dt (2.67)

(2.68)

~ dT ~ C pV = F f C p (T f T ) + ( H r )Vr Qe dt where the rate (r) is given by: r = koe-E/RTCA

(2.69)

(2.70)

The system can be solved if the system is exactly specified and if the initial conditions are given: V(ti) = Vi T(ti) = Ti and CA(ti) = CAi (2.71)

Degrees of freedom analysis

Parameter of constant values: , E, R, Cp, Hr and ko (Forced variable): Ff , CAf and Tf 47

Remaining variables: V, Fo, T, CA and Qe Number of equations: 3 (Eq. 2.67. 2.68 and 2.69) The degree of freedom is 53 = 2. Following the analysis of example 2.1.3, the two extra relations are between the effluent stream (Fo) and the volume (V) on one hand and between the rate of heat exchanged (Qe) and temperature (T) on the other hand, in either open loop or closed loop operations. A more elaborate model of the CSTR would include the dynamic of the cooling jacket (Fig. 2.8). Assuming the jacket to be perfectly mixed with constant volume Vj, density j and constant average thermal capacity Cpj, the dynamic of the cooling jacket temperature can be modeled by simply applying the macroscopic energy balance on the whole jacket: dT j ~ ~ jC p j V j = j F jC p j (T jf T j ) + Qe dt (2.72)

Since Vj, j, Cpj and Tjf are constant or known, the addition of this equation introduces only one variable (Tj). The system is still exactly specified.

Fj , Tjf

Fo , CA , T

Fj , Tj Ff , CAf , Tf

Figure 2-8 Jacketed Non-isothermal CSTR

2.1.6 Mixing Process

Consider the tank of figure 2.9 where two solutions 1 and 2 containing materials A and B are being mixed. Stream 1 has flow rate F1 (m3/s), density 1 (kg/m3), T1 (K),

48

concentration CA1 (mole/m3) and CB1 (mole/m3) of material A and B. Similarly stream 2 has flow rate F2 (m3/s), density 2 (kg/m3), T2 (K), concentration CA2 (mole/m3) and CB2 (mole/m3) of material A and B. The effluent stream has flow rate Fo (m3/s), density o (kg/m3), To (K), concentration CAo (mole/m3) and CBo (mole/m3) of material A and B. We assume that the mixing releases heat of rate Q (J/s) which is absorbed by a cooling fluid flowing in a jacket or a coil. Our objective is to develop a model for the mixing process. We will assume that the tank is well mixed. In this case all the effluent properties are equal to the process properties. We also assume for simplicity that the densities and heat capacities of the streams are constant and equal: = 1 = 2 = o Cp = Cp1 = Cp2 = Cp3 (2.73) (2.74)

F1, T1, CA1, CB1

F2, T2, CA2, CB2

Fo, To, CAo, CBo


Figure 2-9 Mixing Process

Total mass balance The mass balance equation yields d ( V ) = (F11 + F2 2 ) Fo o dt

(2.75)

49

Since the densities are equals we have: dV = (F1 + F2 ) Fo dt

(2.76)

Component balance The component balance for species A for instant yields: d ( oC AoV ) = (F11C A1 + F2 2C A2 ) Fo oC Ao dt

(2.77)

Expanding Eq. 2.77 yields:

o V

dC Ao dV + C Ao = (F11C A1 + F2 2C A2 ) Fo oC Ao dt dt

(2.78)

Substituting Eq. 2.76 into 2.78 yields after some manipulation: dC Ao = F1 (C A1 C Ao ) + F2 (C A2 C Ao ) dt

(2.79)

A similar equation holds for component B,

d ( oCBoV ) = (F11CB1 + F2 2CB 2 ) Fo oCBo dt

(2.80)

Energy balance The general energy balance equation (Eq. 2.23) yields, assuming negligible kinetic, potential energy and since no reaction or shaft work occurs:

50

~ d oVho , mix ~ ~ ~ = 1F1h1, mix + 2 F2 h2, mix o Fo ho , mix Q dt ~ where hi ,mix (J/kg) is the specific enthalpy of mixture i.
~ The specific enthalpy hmix of n components can be written as: ~ ~ hmix (T , P) = hmix (Tref , P) + Cpmix (T Tref ) where ~
n

(2.81)

(2.82)

mix hmix (Tref , P ) = Ci hi (Ci , Tref , P ) + Ck hs (Ci , T , P )


i =1

(2.83)

with hi (J/mole) being the molar enthalpy of component i and hs is the heat of solution per mole of a key component k. Assume constant process pressure (P) then we can write the enthalpy of each stream as follows, taking component (A) as the key component,
1 h1, mix ( T1 ) = C A1 h A + C B 1 h B + C A1 h1, s + 1Cp 1, mix ( T1 T ref ) ~

(2.84)

2h2, mix (T2 ) = C A2hA + CB 2 hB + C A2 h2, s + 2Cp2, mix (T2 Tref ) o ho , mix (T ) = C Ao hA + CBo hB + C Ao ho , s + oCpo , mix (T Tref ) ~

(2.85)

(2.86)

Substituting Eq. 2.86 into the left hand side of Eq. 2.81 and expanding yields: ~ d oVho ,mix dt

(2.87)

d (VC Ao ) + h d (VC Bo ) + h d (VC Ao ) + CpV dT + CpT dV = hA B o ,s dt dt dt dt dt

51

The right hand side of equation (2.81) is equal after some manipulation to:
h A (C A1 F1 + C A2 F2 C Ao Fo ) + hB (C B1 F1 + C B 2 F2 C Bo Fo ) + (C F h + C F h C F h )
A1 1 1,s A2 2 2 ,s Ao o o ,s

(2.88)

+ Cp( F1T1 + F2T2 FoTo ) Q

Substituting into the right hand side of Eq. 2.87, the total mass balance equation (Eq. 2.76) and the component balance equation for species (A) (Eq. 2.77) and that of B (Eq. 2.80), and equating Eq. 2.87 to Eq. 2.88 yields, after some manipulations:

VCp

dT = C A1 F1 ( h1,s ho ,s ) + C A2 F2 ( h2,s ho ,s ) dt + Cp (F1 (T1 T ) + F2 (T2 T ) ) Q

(2.89)

The mixer is then described by three ODE's Equations (2.76, 2.79, 2.89). To these relations we should add the relations that give the heats of mixing: h1, s = f1 (C A1 , CB1 , T1 ) h2, s = f 2 (C A2 , CB 2 , T2 ) ho , s = f 3 (C Ao , CBo , To ) (2.90)

(2.91)

(2.92)

Degrees of freedom analysis

Parameter of constant values: , A, Tref, Cp, hA and hB

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


Remaining variables: V, Fo, To, CAo, Q, h1, s , h2, s , ho , s

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

52

The degree of freedoms is therefore 8 6 = 2. The two needed relations are the relation between effluent stream Fo and height L and the relation between the heat Q and temperature To in either open loop or closed-loop operations.

2.1.7 Heat Exchanger

Consider the shell and tube heat exchanger shown in figure 2.10. Liquid A of density A is flowing through the inner tube and is being heated from temperature TA1 to TA2 by liquid B of density B flowing counter-currently around the tube. Liquid B sees its temperature decreasing from TB1 to TB2. Clearly the temperature of both liquids varies not only with time but also along the tubes (i.e. axial direction) and possibly with the radial direction too. Tubular heat exchangers are therefore typical examples of distributed parameters systems. A rigorous model would require writing a microscopic balance around a differential element of the system. This would lead to a set of partial differential equations. However, in many practical situations we would like to model the tubular heat exchanger using simple ordinary differential equations. This can be possible if we think about the heat exchanger within the unit as being an exchanger between two perfect mixed tanks. Each one of them contains a liquid.
Liquid, B TB1

Tw

Liquid, A TA1

TA2

Liquid, B TB2

Figure 2-10 Heat Exchanger

For the time being we neglect the thermal capacity of the metal wall separating the two liquids. This means that the dynamics of the metal wall are not included in the model. We will also assume constant densities and constant average heat capacities.

53

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

TA2 and TB2 of each liquid. A better way would be to take as state variable not the exit
temperature but the average temperature between the inlet and outlet:

TA =

TA1 + TA2 2

(2.93) (2.94)

TB =

TB1 + TB 2 2

For liquid A, a macroscopic energy balance yields:

AC p VA
A

dTA = A FAC p A (TA1 TA2 ) + Q dt

(2.95)

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

BC p VB
B

dTB = B FBC p B (TB1 TB 2 ) Q dt

(2.96)

The amount of heat Q exchanged is:

Q = UAH (TB TA)


Or using the log mean temperature difference:

(2.97)

Q = UAH Tlm
where (TA2 TB1 ) (TA1 TB 2 ) (T TB1 ) ln A2 (TA1 TB 2 )

(2.98)

Tlm =

(2.99)

54

with U (J/m2s) and AH (m2) being respectively the overall heat transfer coefficient and heat transfer area. The heat exchanger is therefore describe by the two simple ODE's (Eq. 2.95) and (Eq. 2.96) and the algebraic equation (Eq. 2.97).

Degrees of freedom analysis

Parameter of constant values: , CpA, VA, , CpB, VB, U, AH (Forced variable): TA1, TB1, FA, FB Remaining variables: TA2, TB2, Q Number of equations: 3 (Eq. 2.95, 2.96, 2.97)
The degree of freedom is 5 3 = 2. The two extra relations are obtained by noting that the flows FA and FB are generally regulated through valves to avoid fluctuations in their values. So far we have neglected the thermal capacity of the metal wall separating the two liquids. A more elaborated model would include the energy balance on the metal wall as well. We assume that the metal wall is of volume Vw, density w and constant heat capacity Cpw. We also assume that the wall is at constant temperature Tw, not a bad assumption if the metal is assumed to have large conductivity and if the metal is not very thick. The heat transfer depends on the heat transfer coefficient ho,t on the outside and on the heat transfer coefficient hi,t on the inside. Writing the energy balance for liquid B yields:

BC p VB
B

dTB = B FBC p B (TB1 TB 2 ) ho ,t Ao , t (TB TW ) dt

(2.100)

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

wC p Vw
w

dTw = ho ,t Ao ,t (TB Tw ) hi , t Ai ,t (Tw TA ) dt

(2.101)

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

55

AC p VA
A

dTA = A FAC p A (TA1 TA2 ) + hi , t Ai ,t (Tw TA ) dt

(2.102)

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

2.1.8 Heat Exchanger with Steam

A common case in heat exchange is when a liquid L is heated with steam (Figure 2.11). If the pressure of the steam changes then we need to write both mass and energy balance equations on the steam side.

Tw

Steam Ts(t)

Liquid, L TL1

TL2

condensate, Ts

Figure 2-11 Heat Exchanger with Heating Steam

The energy balance on the tube side gives:

LC p LVL

dTL = L FLC pL (TL1 TL 2 ) + Qs dt

(2.103)

where

TL =

TL1 + TL 2 2

(2.104)

Qs = UAs (Ts TL)

(2.105)

56

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

Ts = Ts (P)
Assuming ideal gas law, then the mass flow of steam is:

(2.106)

ms =

M s PsVs RTs

(2.107)

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

M sVs dP = s Fs c Fc RTs dt

(2.108)

where Fc and c are the condensate flow rate and density. The heat losses at the steam side are related to the flow of the condensate by:

Qs = Fc s
Where s is the latent heat.

(2.109)

Degrees of freedom analysis

Parameter of constant values: L, CpL, Ms, As, U , Ms, R (Forced variable): TL1 Remaining variables: TL2, FL, Ts, Fs, Ps, Qs, Fc Number of equations: 5 (Eq. 2.103, 2.105, 2.106, 2.108, 2.109)
The degrees of freedom is therefore 7 5 = 2. The extra relations are given by the relation between the steam flow rate Fs with the pressure Ps either in open-loop or closed-loop operations. The liquid flow rate F1 is usually regulated by a valve.

57

2.1.9 Single Stage Heterogeneous Systems: Multi-component flash drum

The previous treated examples have discussed processes that occur in one single phase. There are several chemical unit operations that are characterized with more than one phase. These processes are known as heterogeneous systems. In the following we cover some examples of these processes. Under suitable simplifying assumptions, each phase can be modeled individually by a macroscopic balance. A multi-component liquid-vapor separator is shown in figure 2.12. The feed consists of Nc components with the molar fraction zi (i=1,2 Nc). The feed at high temperature and pressure passes through a throttling valve where its pressure is reduced substantially. As a result, part of the liquid feed vaporizes. The two phases are assumed to be in phase equilibrium. xi and yi represent the mole fraction of component i in the liquid and vapor phase respectively. The formed vapor is drawn off the top of the vessel while the liquid comes off the bottom of the tank. Taking the whole tank as our system of interest, a model of the system would consist in writing separate balances for vapor and liquid phase. However since the vapor volume is generally small we could neglect the dynamics of the vapor phase and concentrate only on the liquid phase.

Fv yi Fo zi To Po
P, T, Vv

VL

FL xi

Figure 2-12 Multicomponent Flash Drum

For liquid phase:

Total mass balance:

58

d ( LVL ) = f F f L FL v Fv dt

(2.110)

Component balance:

d ( LV L x i ) = f F f z i L FL x i v Fv y i (i=1,2,.,Nc-1) dt

(2.111)

Energy balance: ~ ~ ~ d ( LVL h ) ~ = f Ff h f L FL h v Fv H dt

(2.112)

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

In addition to the balance equations, the following supporting thermodynamic relations can be written:

Liquid-vapor Equilibrium:

Raoult's law can be assumed for the phase equilibrium

yi =

xi Pi s P
Nc

(i=1,2,.,Nc)

(2.113)

Together with the consistency relationships:

y
i =1

=1

(2.114)

x
i =1

Nc

=1

(2.115)

Physical Properties:

The densities and enthalpies are related to the mole fractions, temperature and pressure through the following relations:

59

L v

= f(xi,T,P) = f(yi,T,P) MvaveP/R T

(2.116) (2.117)
(2.118)

Mvave = h H m

y M
i =1 i

Nc

= f(xi,T) = f(yi,T) =

x Cp (T T
i =1 Nc i i i =1 i i

Nc

ref

) ) + m

(2.119) (2.120)

y Cp (T T

ref

y
i =1

Nc

(2.121)

i i

Degrees of freedom analysis:

Forcing variables: Ff, Tf, Pf , zi (i=1,2..Nc), Remaining variables:2Nc+5: VL, FL, FV, P, T, xi (i=1,2..Nc), yi(i=1,2,Nc) Number of equations: 2Nc+3: (Eq. 2.110, 2.111, 2.112, 2.113, 2.114, 2.115)
Note that physical properties are not included in the degrees of freedom since they are specified through given relations. The degrees of freedom is therefore (2Nc+5)-

(2Nc+3)=2. Generally the liquid holdup (VL) is controlled by the liquid outlet flow rate
(FL) while the pressure is controlled by FV. In this case, the problem becomes well defined for a solution.
2-1-10 Two-phase Reactor

Consider the two phase reactor shown in figure 2.13. Gaseous A and liquid B enters the reactor at molar flow rates FA and FB respectively. Reactant A diffuses into the liquid phase with molar flux (NA) where it reacts with B producing C. The latter diffuses into the vapor phase with molar flux (NC). Reactant B is nonvolatile. The product C is withdrawn with the vapor leaving the reactor. The objective is to write the mathematical equations that describe the dynamic behavior of the process. We consider all flows to be in molar rates.

60

Figure 2-4 Two Phase Reactor

Assumptions:

The individual phases are well mixed and they are in physical equilibrium at
pressure P and temperature T.

The physical properties such as molar heat capacity Cp, density , and latent heat
of vaporization are constant and equal for all the species.

The reaction mechanism is: A+B

C and its rate has the form: Rc = k CA CB VL

The two phases are in equilibrium and follows the Raoults law. Total enthalpy for the system is given as: H = NL HL + Nv Hv where HL and Hv are
molar enthalpies in the liquid and vapor phases respectively, and NL and Nv are their corresponding molar holdups. The assumption of well mixing allows writing the following macroscopic balances:

Vapor phase:

Total mass balance:

dN v = FA N A + N c Fv dt

(2.122)

61

Component balance for A:


d (Nv y A ) = FA N A Fv y A dt

(2.123)

Since d(NvyA)/dt = Nv dyA/dt + yA dNv/dt, and using equation (2.122), equation (2.123) can be written as follows:

Nv

dy A = FA (1 y A ) N A (1 y A ) N c y A dt

(2.124)

Liquid phase:

Total mass balance:

dN L = FB + N A N c FL Rc dt
Component balance for A:

(2.125)

d ( N L xA ) = N A FL x A Rc dt

(2.126)

Since d(NL xA)/dt = NL dxA/dt + xA dNL/dt, and using equation (2.125), equation (2.126) can be written as follows:

NL

dx A = N A (1 x A ) FB x A Rc (1 x A ) + N c x A dt

(2.127)

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

62

NL

dxB = N A (1 xB ) + FB xB Rc (1 xB ) + N c xB dt

(2.128)

Energy balance, assuming Tref = 0:

d ( N L H L + Nv H v ) = FBCpTB + FA (CpTA + ) FLCpT FvCpT Rc H r + Q dt


Note that:

(2.129)

d (NLH L ) d (H L ) d (NL ) d (T ) d (NL ) = NL + HL = N LCp + CpT dt dt dt dt dt d ( Nv ) d ( Nv H v ) d (Hv ) d ( Nv ) d (T ) = Nv + Hv = N vCp + (CpT + ) dt dt dt dt dt

(2.130)

(2.131)

Substituting the last two equations, and using the definition of dNL/dt and dNv/dt from equations (2.122) and (2.125), in equation (2.131) yields:

N L + Nv

H r d (T ) Q = FA (TA T ) + FB (TB T ) + Rc (T )+ ( N A Nc ) + dt Cp Cp Cp

(2.132)

The following additional equations are needed: Vapor-liquid equilibrium relations: yAP xA PAs = 0 yAP (1 xA xB) Pcs = 0 Total volume constraint: V = VL + Vv (2.135) (2.133) (2.134)

63

Or, using ideal gas law for vapor volume and total volume and knowing that VL = NL/ , we can write: nRT = NvRT + NLP/ or V = NvRT/P + NL/ (2.137) (2.136)

Degrees of freedom analysis:

Forcing variables: FA, FB, TA, TB, Q, P Physical properties and parameters: Hr, Cp, , R, , V, PAs , PCs Remaining variables: NA, Nc, NL, Nv, FL, T, xA , xB, yA Number of equations: (Eq. 2.122, 2.124, 2.125, 2.127, 2.128, 2.132, 2.133,2.134, 2.137) The degree is freedom is 9-9=0 and the problem is exactly specified. Note that the reaction rate Rc is defined and that the outlet flow Fv can be determined from the overall mass balance.

2-1-11 Reaction with Mass Transfer

Figure 2.14 shows

a chemical reaction that takes place in a gas-liquid

environment. The reactant A enters the reactor as a gas and the reactant B enters as a liquid. The gas dissolves in the liquid where it chemically reacts to produce a liquid C. The product is drawn off the reactor with the effluent FL. The un-reacted gas vents of the top of the vessel. The reaction mechanism is given as follows: A+B Assumptions: Perfectly mixed reactor Isothermal operation Constant pressure, density, and holdup. C (2.138)

64

Negligible vapor holdup.

Figure 2-14 Reaction with Mass Transfer

In such cases, when the two chemical phenomena, i.e., mass transfer and chemical reaction, occur together, the reaction process may become mass transfer dominant or reaction-rate dominant. If the mass transfer is slower reaction rate, then mass transfer prevail and vise versa. Due to the perfectly mixing assumption, macroscopic mass transfer of component A from the bulk gas to the bulk liquid is approximated by the following molar flux: NA = KL (C*A CA) where KL CA* CA is mass transfer coefficient is gas concentration at gas-liquid interface is gas concentration in bulk liquid (2.139)

To fully describe the process, we derive the macroscopic balance of the liquid phase where the chemical reaction takes place. This results in:

65

Liquid phase:

Total mass balance: dV = B FB + M A Am N A FL dt Component balance on A: dC A = Am N A FLC A rV dt (2.140)

(2.141)

Component balance on B: dCB = FBCBo FLCB rV dt

(2.142)

Vapor phase:

Here, since vapor holdup is negligible, we can write a steady state total continuity equation as follows: Fv = FA MA Am NA/A where Am MA V total mass transfer area of the gas bubble molecular weight of component A density liquid volume (2.143)

Degrees of freedom analysis:

Forcing variables: FA, FB, CB0, Parameters of constant values: KL, MA, Am, , A, B, 66

Remaining variables: CB, NA, CA, Fv, V Number of equations: (Eq. 2.139-2.143) Note that the liquid flow rate, FL can be determined from the overall mass balance and that the reaction rate r should be defined.

2.1.12 Multistage Heterogeneous Systems: Liquid-liquid extraction

There are many chemical processes which consist in a number of consecutive stages in series. In each stage two streams are brought in contact for separating materials due to mass transfer. The two streams could be flow in co-current or counter current patterns. Counter-current flow pattern is known to have higher separation performance Examples of these processes are distillation columns, absorption towers, extraction towers and multi-stage flash evaporator where distillate water is produced from brine by evaporation. The same modeling approach used for single stage processes will be used for the staged processes, where the conservation law will be written for one stage and then repeated for the next stage and so on. This procedure will result in large number of state equation depending on the number of stages and number of components. The separation process generally takes place in plate, packed or spray-type towers. In tray or spay-type columns the contact and the transfer between phases occur at the plates. Generally, we can always assume good mixing of phases at the plates, and therefore macroscopic balances can be carried out to model these type of towers. Packed towers on the other hand are used for continuous contacting of the two phases along the packing. The concentrations of the species in the phases vary obviously along the tower. Packed towers are therefore typical examples of distributed parameters systems that need to be modeled by microscopic balances. In the following we present some examples of mass separation units that can be modeled by simple ODE's, and we start with liquid-liquid extraction process.

67

Liquid-liquid extraction is used to move a solute from one liquid phase to another. Consider the single stage countercurrent extractor, shown in Figure 2.15, where it is desired to separate a solute (A) from a mixture (W) using a solvent (S). The stream mixture with flow rate W (kg/s) enters the stage containing XAf weight fraction of solute (A). The solvent with a flow rate (S) (kg/s) enters the stage containing YAf weight fraction of species (A). As the solvent flows through the stage it retains more of (A) thus extracting (A) from the stream (W). Our objective is to model the variations of the concentration of the solute. A number of simplifying assumptions can be used: The solvent is immiscible in the other phase. The concentration XA and YA are so small that they do not affect the mass flow rates. Therefore, we can assume that the flow rates W and S are constant. A total mass balance is therefore not needed. An equilibrium relationship exists between the weight fraction YA of the solute in the solvent (S) and its weight fraction XA in the mixture (W). The relationship can be of the form: YA = K XA (2.144)

Here K is assumed constant. Since both phases are assumed perfectly mixed a macroscopic balance can be carried out on the solute in each phase. A component balance on the solute in the solvent-free phase of volume V1 and density 1 gives: dX A = WX Af WX A N A dt

1V1

(2.145)

whereas NA (kg/s) is the flow rate due to transfer flow between the two phases. A similar component balance on the solvent phase of volume V2 and density 2 gives: dYA = SYAf SYA + N A dt

2V2

(2.146)

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

68

2V2 K

dX A = SY Af SKX A + N A dt

(2.147)

Adding Eq. 2.145 and 2.147 yields: dX A = WX Af + SY Af (W + KS ) X A dt

( 1V1 + 2V2 K )

(2.148)

The latter is a simple linear ODE with unknown XA. With the volume V1, V2 and flow rates W, S known the system is exactly specified and it can be solved if the initial concentration is known: XA(ti) = XAi Note that we did not have to express explicitly the transferred flux NA. (2.149)

Figure 2-15 Single Stage Liquid-Liquid Extraction Unit

The same analysis can be extended to the multistage liquid/liquid extraction units as shown in Figure 2.16. The assumptions of the previous example are kept and we also assume that all the units are identical, i.e. have the same volume. They are also assumed to operate at the same temperature.

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

69

A component balance in the ith stge (excluding the first and last stage) gives: Solvent-free phase, of volume V1i and density 1i dX Ai = WX Ai 1 WX Ai N Ai (i=2,N-1) dt (2.150)

1iV1i

where NAi is the flow rate due to transfer between the two phases at stage i. Solvent phase of volume V2i and density 2i: dY Ai = SYAi +1 SY Ai + N Ai (i = 2,N-1) dt (2.151)

2iV2i

Writing the equilibrium equation (Eq. 2.144) for each component YAi = K XAi (i=1,.,.N) and adding the last two equations yield: dX Ai = WX Ai 1 + SYAi +1 (W + KS ) X Ai (i=2,N-1) dt (2.152)

(1iV1i + 2iV2i K )

Since the volume and densities are equal, i.e.: V1i = V1 and V2i = V (2.153) (2.154)

1i = 1 and 2i = 2
Equation 2.152 is therefore equivalent to: dX Ai = WX Ai 1 + SYAi +1 (W + KS ) X Ai dt

(1V1 + 2V2 K )

(i=2,N-1)

(2.155)

The component balance in the first stage is:

70

(1V1 + 2V2 K )

dX A1 = WX Af + SYA1 (W + KS ) X A1 dt

(2.156)

And that for the last stage is:


dX AN = WX AN 1 + SYAf (W + KS ) X AN dt

(1V1 + 2V2 K )

(2.157)

The model is thus formed by a system of linear ODE's (Eq. 2.155, 2.156, 2.157) which can be integrated if the initial conditions are known:
XA(ti) = XAi

(i=1,2,N)

(2.158)

Degrees of freedom analysis

Parameter of constant values: 1, 2, K, V1, V2, W and S (Forced variable): XAf, YAf Remaining variables: XAi (2N variables): (i=1,2,N) and YAi (i=1,2,N) Number of equations: 2N [2.144 (N equations, one for each component), Eq.

2.155 (N-2 eqs), 2.156(1 eq), 2.157(1 eq)]. The problem is therefore is exactly specified.

2.1.13 Binary Absorption Column

Consider a N stages binary absorption tower as shown in figure 2.17. A Liquid stream flows downward with molar flow rate (L) and feed composition (xf). A Vapor stream flows upward with molar flow rate (G) and feed composition (yf). We are interested in deriving an unsteady state model for the absorber. A simple vapor-liquid equilibrium relation of the form of:
yi = a xi + b

(2.159)

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


Assumptions:

71

Isothermal Operation Negligible vapor holdup Constant liquid holdup in each stage Perfect mixing in each stage

According to the second and third assumptions, the molar rates can be considered constants, i.e. not changing from one stage to another, thus, total mass balance need not be written. The last assumption allows us writing a macroscopic balance on each stage as follows: Component balance on stage i:
dxi = G ( yi 1 yi ) + L( xi +1 xi ) (i=2,N-1) dt

(2.160)

where H is the liquid holdup, i.e., the mass of liquid in each stage. The last equation is repeated for each stage with the following exceptions for the last and the first stages:

Figure 2-17 N-stages Absorbtion Tower

In the last stage, xi+1 is replaced by xf In the first stage, yi-1 is replaced by yf 72

Degrees of freedom analysis

Parameter of constant values: , a, b (Forced variable): G, L, xf, yf Remaining variables: xi (i=1,2,N), yi (i=1,2,N) Number of equations:2N (Eqs.2.159, 2.160)

The problem is therefore is exactly specified.

2.1.14 Multi-component Distillation Column

Distillation columns are important units in petrochemical industries. These units process their feed, which is a mixture of many components, into two valuable fractions namely the top product which rich in the light components and bottom product which is rich in the heavier components. A typical distillation column is shown in Figure 2.18. The column consists of n trays excluding the re-boiler and the total condenser. The convention is to number the stages from the bottom upward starting with the re-boiler as the 0 stage and the condenser as the n+1 stage.
Description of the process:

The feed containing nc components is fed at specific location known as the feed tray (labeled f) where it mixes with the vapor and liquid in that tray. The vapor produced from the re-boiler flows upward. While flowing up, the vapor gains more fraction of the light component and loses fraction of the heavy components. The vapor leaves the column at the top where it condenses and is split into the product (distillate) and reflux which returned into the column as liquid. The liquid flows down gaining more fraction of the heavy component and loses fraction of the light components. The liquid leaves the column at the bottom where it is evaporated in the re-boiler. Part of the liquid is drawn as bottom product and the rest is recycled to the column. The loss and gain of materials occur at each stage where the two phases are brought into intimate phase equilibrium.

73

Cw

D xd

F z

steam

B xb

Figure 2-18 Distillation Column

Modeling the unit:

We are interested in developing the unsteady state model for the unit using the flowing assumptions:
100% tray efficiency Well mixed condenser drum and re-boiler. Liquids are well mixed in each tray. Negligible vapor holdups. liquid-vapor thermal equilibrium

Since the vapor-phase has negligible holdups, then conservation laws will only be written for the liquid phase as follows: Stage n+1 (Condenser), Figure 2.19a: Total mass balance:

74

dM D = Vn ( R + D) dt

(2.161)

Component balance:
d ( M D xD , j ) = Vn yn , j ( R + D ) x D , j dt

j = 1, nc 1

(2.162)

Energy balance:
d ( M D hD ) = Vn hn ( R + D )hD Qc dt

(2.163)

Note that R = Ln+1 and the subscript D denotes n+1 Stage n, Figure fig2.19b Total Mass balance:
dM n = Vn 1 Vn + R Ln dt

(2.164)

Component balance:
d (M n xn, j ) dt

= Vn 1 y n 1, j Vn y n , j + Rx D , j Ln x n , j

j = 1, nc 1

(2.165)

Energy balance:
d ( M n hn ) = Vn 1H n 1 Vn H n + RhD Ln hn dt

(2.166)

Stage i, Figure 2.19c Total Mass balance: 75

dM i = Vi 1 Vi + Li +1 Li dt

(2.167)

Component balance:
d ( M i xi , j ) = Vi 1 yi 1, j Vi yi , j + Li +1 xi +1, j Li xi , j dt

j = 1, nc 1

(2.168)

Energy balance:
d ( M i hi ) = Vi 1H i 1 Vi H i + Li +1hi +1 Li hi dt

(2.169)

Stage f (Feed stage), Figure 2.19d Total Mass balance:


dM f dt

= V f 1 (V f + (1 q ) F ) + L f +1 ( L f + qF )

(2.170)

Component balance:
d (M f x f , j ) = 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 (2.171)

Energy balance:
d (M f hf ) = 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

(2.172)

Stage 1, Figure 2.19e Total Mass balance:

76

dM 1 = VB V1 + L2 L1 dt Component balance:
d ( M 1 x1, j ) = VB y B , j V1 y1, j + L2 x2, j L1 x1, j dt

(2.173)

j = 1, nc 1

(2.174)

Energy balance: d ( M 1h1 ) = VB H B V1H 1 + L2 h2 L1h1 dt Stage 0 (Re-boiler), Figure 2.19f Total Mass balance: dM B = VB + L1 B dt Component balance:
d ( M B xB , j ) = VB y B , j + L1 x1, j Bx B , j dt

(2.175)

(2.176)

j = 1, nc 1

(2.177)

Energy balance: d ( M B hB ) = VB H B + L1h1 BhB + Qr dt Note that L0 = B and B denotes the subscript 0 Additional given relations: Phase equilibrium: yj = f (xj, T,P) Liquid holdup: Mi = f (Li) 77 (2.178)

Enthalpies: Hi = f (Ti, yi,j), hi = f (Ti, xi,j) Vapor rates: Vi = f (P) Notation: Li, Vi Hi, hi xi, yi Mi Q Z F Liquid and vapor molar rates Vapor and liquid specific enthalpies Liquid and vapor molar fractions Liquid holdup Liquid fraction of the feed Molar fractions of the feed Feed molar rate

Degrees of freedom analysis


Variables Mi MB, MD Li B,R,D xi,j xB,j,xD,j yi,j yB,j hi hB, hD Hi HB Vi VB Ti TD, TB Total n 2 n 3 n(nc 1) 2(nc 1) n(nc 1) nc 1 n 2 n 1 n 1 n 2 11+6n+2n(nc1)+3(nc1)

78

Equations: Total Mass Energy Component Equilibrium Liquid holdup Enthalpies Vapor rate hB = h1 yB = xB Total n+2 n+2 (n + 2)(nc 1) n(nc 1) n 2n+2 n 1 (nc 1) 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 reboiler. Use VB and R to control the end compositions i.e., xB, xD

79

Vn , y n

R, xd stage n (a) Qc (b) D, xd Vi, yi stage i (c) Ln, xn Lf+1, xf+1 stage f (d) Vi-1, yi-1 V1 , y 1 stage 1 (f) (e) Lf, xf VB, yB

V n , yn

R, xd

Vn-1, yn-1 Vf , yf

Li+1, xi+1

Li, xi L 2 , x2

Vf-1, yf-1

Qr

B, xB

L1, x1

VB, yB

L1, x1

Figure 2-19 Distillation Column Stages Simplified Model


One can further simplify the foregoing model by the following assumptions: (a) Equi-molar flow rates, i.e. whenever one mole of liquid vaporizes a tantamount of vapor condenses. This occur when the molar heat of vaporization of all components are about the same. This assumption leads to further idealization that implies constant temperature over the entire column, thus neglecting the energy balance. In addition, the vapor rate through the column is constant and equal to: VB = V1 = V2 = = Vn (2.179)

80

(b) Constant relative volatility, thus a simpler formula for the phase equilibrium can be used: yj = j xj/(1+(j 1) xj) (2.180)

Degrees of Freedom:
Variables: Mi, MB, MD Li, B,R,D xi,xB,xD yj, yB V Total Equations: Total Mass Component Equilibrium Liquid holdup yB = xB Total n+2 (n + 2)(nc 1) n(nc 1) n 1 2+2n+(2n+3)(nc-1) n+2 n+3 (n + 2)(nc 1) (n + 1)(nc 1) 1 2 + 2n + (2n + 3)(nc 1)

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

81

2.2 Examples of Distributed Parameter Systems 2.2.1 Liquid Flow in a Pipe


Consider a fluid flowing inside a pipe of constant cross sectional area (A) as shown in Figure 2.20. We would like to develop a mathematical model for the change in the fluid mass inside the pipe. Let v be the velocity of the fluid. Clearly the velocity changes with time (t), along the pipe length (z) and also with the radial direction (r). In order to simplify the problem, we assume that there are no changes in the radial direction. We also assume isothermal conditions, so only the mass balance is needed. Since the velocity changes with both time and space, the mass balance is to be carried out on microscopic scale. We consider therefore a shell element of width z and constant cross section area (A) as shown in Fig. 2.20.

Figure 2-20 Liquid flow in a pipe


Mass into the shell: vAt|z (2.181)

where the subscript (.|z) indicates that the quantity (.) is evaluated at the distance z. Mass out of the shell: vAt|z+z Accumulation: (2.182)

82

Az|t+t Az|t

(2.183)

Similarly the subscript (.|t) indicates that the quantity (.) is evaluated at the time t. The mass balance equation is therefore: vAt|z = vAt|z+z + Az|t+t Az|t (2.184)

We can check for consistency that the units in each term are in (kg). Dividing Eq. 2.184 by t z and rearranging yields: (A) (A) (vA) (vA) z (2.185)

t + t

t Taking the limit as t 0 and z

z + z

0 gives: (2.186)

(A) (vA) = t z Since the cross section area (A) is constant, Eq. 2.186 yields: ( v ) = t z or ( v ) + =0 t z

(2.187)

(2.188)

The ensuing equation is a partial differential equation (PDE) that defines the variation of 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:

83

v =0 z

(2.189)

This indicates that the velocity is independent of axial direction for one dimensional incompressible flow.

2.2.2 Velocity profile inside a pipe


We reconsider the flow inside the pipe of the previous example. Our objective is to find the velocity profile in the pipe at steady state. For this purpose a momentum balance is needed. To simplify the problem we also assume that the fluid is incompressible. We will carry out a microscopic momentum balance on a shell with radius r, thickness r and length z as shown in Fig 2.21.

Figure 2-21 Velocity profile for a laminar flow in a pipe

Momentum in: (rz2rz)|r (2.190)

where rz is the shear stress acting in the z-direction and perpendicular to the radius r. Momentum out:

84

(rz2rz)|r+r

(2.191)

As for the momentum generation we have mentioned earlier in section 1.8.3 that the generation term corresponds to the sum of forces acting on the volume which in this example are the pressure forces, i.e. (PA)|z (PA)|z+z = P(2rr)|z P(2rr)|z+z (2.192)

There is no accumulation term since the system is assumed at steady state. Substituting this term in the balance equation (Eq. 1.6) and rearranging yields, rrz |r + r rrz |r r ( P |z P |z + z ) = r z (2.193)

We can check for consistency that all the terms in this equation have the SI unit of (N/m2). Taking the limit of (Eq. 2.193) as z and r go to zero yields: d (rrz ) dP = r dr dz (2.194)

Here we will make the assumption that the flow is fully developed, i.e. it is not influenced by the entrance effects. In this case the term dP/dz is constant and we have: dP P2 P1 P = = dz L L (2.195)

where L is the length of the tube. Note that equation (2. 195) is a function of the shear stress rz, but shear stress is a function of velocity. We make here the assumption that the fluid is Newtonian, that is the shear stress is proportional to the velocity gradient: dv z dr (2.196)

rz =

85

Substituting this relation in Eq. 2.194 yields: d rdv z P ( )=r dr dr L (2.197)

or by expanding the derivative: d 2 v z 1 dv z P + )= 2 dr r dr L (2.198)

The system is described by the second order ODE (Eq. 2.198). This ODE can be integrated with the following conditions: The velocity is zero at the wall of the tube vz = 0 at r = R (2.199)

Due to symmetry, the velocity profile reaches a maximum at the center of the tube: dvz =0 dr (2.200)

at r = 0

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 chemical reaction A coupled with the following

B in a slab of catalyst shown in figure 2.22. Our objective is to

determine the variation of the concentration at steady state. The concentration inside the

86

slab varies with both the position z and time t. The differential element is a shell element of thickness z. Flow of moles A in: (SNA)|z where S (m2) is the surface area and NA (moles A/s m2) is the molar flux. Flow of moles A out: (SNA)|z+z Rate of generation of A: (Sz)r (2.203) (2.202) (2.201)

where r = kCA is the rate of reaction, assumed to be of first order. There is no accumulation term since the system is assumed at steady state. The mass balance equation is therefore, (SNA)|z (SNA)|z+z (Sz)kCA =0 (2.204)

Porous catalyst particle

Exterior surface

z z=0 z=L

Figure 2-22 diffusion with chemical reaction inside a slab catalyst


87

Dividing equation (2.204) by Sz results in: ( N A ) | z ( N A ) | z + z kC A = 0 z Taking the limit when z 0, the last equation becomes: dN A kC A = 0 dz The molar flux is given by Fick's law as follows:
dC A dz

(2.205)

(2.206)

N A = D A

(2.207)

where DA is diffusivity coefficient of (A) inside the catalyst particle. Equation (2.206) can be then written as follows: d 2C A dz
2

DA

kC A = 0

(2.208)

This is also another example where a one-dimensional distributed system is reduced to a lumped parameter system at steady state. In order to solve this second-order ODE, the following boundary conditions could be used: at z = L, CA = CAo at z = 0, dCA/dr = 0 (2.209) (2.210)

The first condition imposes the bulk flow concentration CAo at the end length of the slab. The second condition implies that the concentration is finite at the center of the slab.

88

2.2.4 Temperature profile in a heated cylindrical Rod


Consider a cylindrical metallic rod of radius R and length L, initially at a uniform temperature of To. Suppose that one end of the rod is brought to contact with a hot fluid of temperature Tm while the surface area of the rod is exposed to ambient temperature of Ta. We are interested in developing the mathematical equation that describes the variation of the rod temperature with the position. The metal has high thermal conductivity that makes the heat transfer by conduction significant. In addition, the rod diameter is assumed to be large enough such that thermal distribution in radial direction is not to be neglected. The system is depicted by figure 2.23. For modeling we take an annular ring of width z and radius r as shown in the figure. The following transport equation can be written:

Figure 2-23 Temperature Distribution In a cylindrical rod

Heat flow in by conduction at z: qz(2 r r)t Heat flow in by conduction at r: qr(2 r z)t 89 (2.212) (2.211)

where qz and qr are the heat flux by conduction in the z and r directions. Heat flow out by conduction at z+z: qz+z(2 r r)t Heat flow out by conduction at r+r: qr+r(2 (r+r)z)t Heat accumulation:
~ ~ (2 r r z)( h t+t h t)

(2.213)

(2.214)

(2.215)

~ where h is the specific enthalpy. Summing the above equation according to the

conservation law and dividing by (2 r z t), considering constant density, gives: ~ ~ ht + t ht q qz + z rqr rqr + r r =r z + t z r Taking the limit of t, z, and r go to zero yield: ~ h q ( rqr ) = r z t z r (2.217) (2.216)

~ Dividing by r and replacing h by C p (T Tref), where C p is the average heat

capacity, and substituting the heat fluxes q with their corresponding relations (Fourier law): T z (2.218)

qz = k z

90

and T r (2.219)

qr = k r

Equation (2.217) is then equivalent to: T 2T k T = kz 2 + r (r ) t z r r r (2.220)

C p

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 (r ) + z 2 r r r (2.221)

0 = kz

If in addition to steady state conditions, the radius of the rod is assumed to be small so that the radial temperature gradient can be neglected then the PDE (Eq. 2.221) can be further simplified. In this case, the differential element upon which the balance equation is derived is a disk of thickness z and radius R. The heat conduction in the radial direction is omitted and replaced by the heat transfer through the surface area which is defined as follows: Q = U(2RL)(Ta T) (2.222) the

Consequently, the above energy balance equation (Eq. 2.221) is reduced to following ODE d 2T U ( 2RL)(Ta T ) dz 2

0 = kz

(2.223)

2.2.5 Isothermal Plug Flow Reactor

91

Let consider a first-order reaction occurring in an isothermal tubular reactor as shown in figure 2.24. We assume plug flow conditions i.e. the density, concentration and velocity change with the axial direction only. Our aim is to develop a model for the reaction process in the tube.

z
v(t,z) CAo z (t,z), CA(t,z) z+ z z=L CA

Figure 2-5 Isothermal Plug flow reactor


In the following we derive the microscopic component balance for species (A) around differential slice of width z and constant cross-section area (S). Flow of moles of A in: As has been indicated in section 1.11.1 mass transfer occurs by two mechanism; convection and diffusion. The flow of moles of species A into the shell is therefore the sum of two terms: (vCA S t) |z + (NA S t)|z where NA is the diffusive flux of A ( moles of A/m2 s). Flow of moles of A out: (vCA S t) |z+z + (NA S t)|z+z Accumulation: (CA S z) |t+t (CA S z) |t (2.226) (2.225) (2.224)

92

Generation due to reaction inside the shell: r(Szt) where r = k CA is the rate of reaction. Substituting all the terms in the mass balance equation (Eq. 1.3) and dividing by t and z gives: (C A S ) |t + t (C A S ) |t (vC A S + N A S ) |z (vC A S + N A S ) |z + z = kC A S t z Taking the limit of t following PDE: vC A N A C A = kC A z z t where NA is the molar flux given by Ficks law as follows: dC A dz (2.230) (2.229) 0 and z (2.228) (2.227)

0 and omitting S from both sides give the

N A = DAB

where DAB is the binary diffusion coefficient. Equation 2.229 can be then written as follows: C A ( vC A ) 2C A = + DAb kC A t z z 2 Expanding the derivatives, the last equation can be reduced to: v 2C A C C A kC A = v A C A + DAb z z 2 z t (2.232) (2.231)

93

This equation can be further simplified by using the mass balance equation for incompressible fluids (Eq. 2.189). We get then: C 2C A C A = v A + D Ab kC A z z 2 t The equation is a PDE for which the state variable (CA) depends on both t and z. The PDE is reduced at steady state to the following second order ODE, dC A d 2C A + DAb kC A dz dz 2 (2.234) (2.233)

0 = v

The ODE can be solved with the following boundary conditions (BC): BC1: BC2: at z = 0 at z = L CA(0) = CA0 dC A ( z ) =0 dz (2.235) (2.236)

The first condition gives the concentration at the entrance of the reactor while the second condition indicates that there is no flux at the exit length of the reactor.

2.2.6 Non-Isothermal Plug-Flow reactor


The tubular reactor discussed earlier is revisited here to investigate its behavior under non-isothermal conditions. The heat of reaction is removed via a cooling jacket surrounding the reactor as shown in figure 2.25. Our objective is to develop a model for the temperature profile along the axial length of the tube. For this purpose we will need to write an energy balance around an element of the tubular reactor, as shown in Fig.2.25. The following assumptions are made for the energy balance:

94

Figure 2-25 Non-isothermal plug flow reactor


Assumptions: Kinetic and potential energies are neglected. No Shaft work Internal energy is approximated by enthalpy Energy flow will be due to bulk flow (convection) and conduction. Under these conditions, the microscopic balance around infinitesimal element of width z with fixed cross-section area is written as follows: Energy flow into the shell: As mentioned in Section 1.11.3 the flow of energy is composed of a term due to convection and another term due to molecular conduction with a flux qz.
~ (qzA + v A h )t|z

(2.237)

Energy flow out of the shell:


~ (qzA + v A h )t|z+z

(2.238)

Accumulation of energy:
~ ~ (A h z)|t+t (A h z)|t

(2.239)

95

Heat generation by reaction: (Hr )kCA A z t Heat transfer to the wall: ht(Dz)(T Tw)t where ht is film heat transfer coefficient. Substituting these equations in the conservation law (equation 1.7) and dividing by t z give: (2.241) (2.240)

~ ~ ~ ~ ( h ) |t + t ( h ) |t ( vh ) |z ( vh ) |z + z q z |z q z |z + z + = z z t D H r kC A ht ( )(T Tw ) A Taking the limit as t and z go to zero yields: ~ ~ ( h ) ( vh ) qz D = H r kC A ht ( )(T Tw ) t z z A The heat flux is defined by Fouriers law as follows: T z

(2.242)

(2.243)

q z = kt

(2.244)

~ where kt is the thermal conductivity. The specific enthalpy ( h ) can be approximated by:

~ h = C p(T Tref )

(2.245)

Since the fluid is incompressible it satisfies the equation of continuity (Eq. 2.189). Substituting these expressions in Eq. 2.243 and expanding gives:
96

C p

T T 2T D = C pv + kt 2 H r ko e E / RT C A ht ( )(T Tw ) t z z A

(2.246)

At steady state this PDE becomes the following ODE, dT d 2T D + kt 2 H r ko e E / RT C A ht ( )(T Tw ) dz dz A (2.247)

0 = C pv

Similarly to Eq. 2.234 we could impose the following boundary conditions: B.C1: B.C2: at z = 0 at z = L T(z) = To dT ( z ) =0 dz (2.248) (2.249)

The first condition gives the temperature at the entrance of the reactor and the second condition indicates that there is no flux at the exit length of the reactor.

2.2.7 Heat Exchanger: Distributed parameter model

We revisit the shell-and-tube heat exchanger already discussed in example 2.1.6. Steam of known temperature Ts flowing around the tube is heating a liquid L of density L and constant velocity v from temperature TL1 to TL2. The temperature in the tube varies obviously with axial direction z, radial direction r and time t. To simplify the problem we will assume that there are no change in the radial direction. This assumption is valid if the radius is small and no large amount of heat is transferred. The heat transfer from the steam to the liquid depends on the heat transfer coefficient on the steam side, hto and on the transfer on the liquid side hti. We also neglect the thermal capacity of the metal wall separating the steam and the liquid and assume that the exchange between the steam and liquid occurs with an overall heat transfer coefficient U. We also assume constant heat capacity for the liquid. An energy balance on a differential element of the exchanger of length z and cross-sectional area A, yields:

97

Flow of energy in: (2.250)

(vA C p )t|z

Flow of energy out: (2.251)

(vA C p )t|z+z

Energy accumulation:
~ ~ (A h z)|t+t (A h z)|t

(2.252)

Energy generated: U(Dz)(T Ts)t (2.253)

Using the expression for specific enthalpy (Eq. 2.245) and dividing by Atz, the energy balance yields:

(AC p(T Tref )z ) |t + t (AC p(T Tref )z ) | t = (vAC pT ) |z (vAC pT ) |z + z U ( D )(T T )


t z A
s

(2.254)

Taking the limit as t an z goes to zero gives:


( AC pT ) ( vAC pT ) D = U ( )(T Ts ) t z A

(2.255)

Since A = D2/4 and dividing by C p Eq. (2.255) is equivalent to: T T 4U ( = v )(T Ts ) z t C pD

(2.256)

98

At steady state the PDE becomes the following ODE, dT 4U ( )(T Ts ) dz C pD (2.257

0 = v

With the following condition: T ( z = 0) = T0 (2.258

2.2.8 Mass exchange in packed column

In previous section (section 2.1.12) we presented some examples of mass transfer units that can be described by simple ODE's. This includes all the operations that can occur in tray or spray-tray towers. In this section we present an example of modeling a mass transfer operation that occurs in packed tower. Absorption is a mass transfer process in which a vapor solute (A) in a gas mixture is absorbed by contact with a liquid phase in which the solute is more or less soluble. The gas phase consists usually of an inert gas and the solute. This process involves flow transfer of the solute A through a stagnant non diffusive gas B into a stagnant liquid C. The liquid is mainly immiscible in the gas phase. An example is the absorption of ammonia (A) from air (B) by liquid water (C). The operation can be carried out either in tray (plate) towers or in packed towers. The operation in tray towers can be modeled similarly to the liquidliquid extraction process in Example 2.1.12 and it is left as an exercise. We consider here the absorption taking place in a packed tower. Consider the binary absorption tower shown in Figure 2.26. A liquid stream flow downward with molar flow rate L and feed composition (XAf). Vapor stream flows upward with molar flow rate (G) and feed composition (YAf). A simple vapor-liquid equilibrium relation of the form of: YA = HXA (2.265)

is used, where H (mole fraction gas/mole fraction liquid) is the Henry's law constant. This assumption is valid for dilute streams. The molar rates can be considered constants,

99

i.e. not changing from one stage to another, thus the total mass balance need not be written. To establish the model equations we need to write equations for liquid and vapor phase. To simplify the problem we assume constant liquid and vapor holdup in each stage. We also assume isothermal conditions. An energy balance therefore is not needed. The flux NA transferred from bulk liquid to bulk gas is given by: NA = KY (YA YA*) (2.266)

Where KY is the overall mass transfer in the gas-phase (kgmole/m2s mole fraction) and YA* is the value that would be in equilibrium with XA. The flux can also be expressed as: NA = KX (XA XA*) (2.267)

Where KX is the overall mass transfer coefficient in the liquid-phase and XA* is the value that would be in equilibrium with YA. A mass balance on the liquid phase for a differential volume (Fig. 2.26) of the column length z and cross sectional area S yields: Flow of mole in: [(SLXA)t|z + (NASt) |z]z Flow of moles out: (SLXA)t|z+z Rate of accumulation: (SHLXAz)|t+t - (SHLXAz)|t where XA is the liquid fraction of A and HL the liquid holdup (mole/m3).
(2.270)

(2.268)

(2.269)

100

Figure 2-26 Packed column

The balance equation yields: (SHLXAz)|t+t - (SHLXAz)|t = (SLXA)t|z+z - [(SLXA)t|z + (NASt)t|z]z Dividing by Stz and taking the limits as z and t goes to zero yield:
X A X A =L + NA t z

(2.271)

HL

(2.272)

which is equivalent to X X A * = L A + KY (YA YA ) z t (2.273)

HL

We could also use the expression of flux (Eq. 2.267):

101

HL

X A X * = L A + KX (Xa X A) z t

(2.274)

We can develop material balances for the gas phase that are similar to Eq. 2.274. This gives: X A Y * = G A + KY (YA YA ) z t (2.275)

HG

or alternatively: Y YA * = G A + K X ( X a X A ) z t (2.276)

HG

It should be noted that the analysis carried here can be used for a number of operations where packed columns are used. This includes liquid-liquid extraction, gas-liquid absorption and gas-solid drying. In each of these operations an equilibrium relation of the type: YA = f (XA) is generally available. (2.277)

102

Chapter 3: Equations of Change


In the last chapter, we presented examples of microscopic balances in one or two dimensions for various elementary examples. In this chapter we present the general balance equations in multidimensional case. The balances, also called equations of changes can be written in cartesian, cylindrical or spherical coordinates. We will explicitly derive the balance equations in cartesian coordinates and present the corresponding equations in cylindrical and spherical coordinates. The reader can consult the books in reference for more details. Once the equations are presented we show through various examples how they can be used in a systematic way to model distributed parameter models.

3.1 Total Mass balance

Our control volume is the elementary volume xyz 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.

103

Figure 0-1 Total Mass balance in Cartesian coordinates

Mass in: The mass entering in the x-direction at the cross sectional area (yz) is (vx)|x yzt The mass entering in the y-direction at the cross sectional area (xz) is (vy)|y xzt The mass entering in the z-direction at the cross sectional area (xy) is (vz)|z x y t Mass out: The mass exiting in the x-direction is: (vx)|x+x yzt (3.4) (3.3) (3.2) (3.1)

104

The mass exiting in the y-direction is: (vy)|y+y xzt The mass exiting in the z-direction is: (vz)|z+z xy t Rate of accumulation: The rate of accumulation of mass in the elementary volume is: ()|t+t x y z - ()|t x y z rearranging gives: (|t+t |t) x y z = (vx|x vx|x+x)yzt + (vy|y vy|y+y)xzt + (vz|z vz|z+z) x y t Dividing the equation by x y zt results in: (3.8) (3.7) (3.6) (3.5)

Since there is no generation of mass, applying the general balance equation Eq. 1.2 and

|t +t |t vx | x vx |x+x v y | y v y | y +y vz | z vz | z +z = + + t x y z

(3.9)

By taking the limits as y,x,,z and t goes to zero, we obtain the following equation of change:
vx v y vz = t x y z

(3.10)

Expanding the partial derivative of each term yields after some rearrangement:

105

v v v = ( x + y + z ) + vz + vy + vx x y z z y x t

(3.11)

This is the general form of the mass balance in cartesian coordinates. The equation is also known as the continuity equation. If the fluid is incompressible then the density is assumed constant, both in time and position. That means the partial derivatives of are all zero. The total continuity equation (Eq. 3.11) is equivalent to: vx v y vz ) + + x y z (3.12)

0 = (

or simply:
vx v y vz + + x y z

0=

(3.13)

3.2 Component Balance Equation

We consider a fluid consisting of species A, B , and where a chemical reaction is generating the species A at a rate rA (kg/m3s). The fluid is in motion with mass-average velocity v = nt/ (m/s) where nt = nA + nB + (kg/m2s) is the total mass flux and (kg/m3s) is the density of the mixture. Our objective is to establish the component balance equation of A as it diffuses in all directions x,y,z (Figure 3.2).

106

nAy|y+ nAz|z

nAx|x y x z

nAx|x+

nAy|y

nAz|z+

Figure 0-2 Mass balance of component A

Mass of A in: The mass of species A entering the x-direction at the cross sectional (yz) is: (nAx)|x yzt where nAx kg/m2 is the flux transferred in the x-direction Similarly the mass of A entering the y and z direction are respectively: (nAy)|y xzt (nAz)|z x y t Mass of A out: The mass of species A exiting the x, y and z direction are respectively (3.15) (3.16) (3.14)

(nAx)|x+x yzt (nAy)|y+y xzt

(3.17) (3.18)

107

(nAz)|z+z x y t The rate of accumulation is: A|t+t x y z A|t x y z The rate of generation is: -rAx y zt Applying the general balance equation (Eq. 1.3) yields: (A|t+t A|t) x y z = (nAx|x+x nAx|x)yzt + (nAy|y+y nAy|y)xzt + (nAz|z+z nAz|z)xyt +rAx y zt Dividing each term by xyzt and letting each of these terms goes to zero yields: A n Ax n Ay n Az + + + = rA t t t t

(3.19)

(3.20)

(3.21)

(3.22)

(3.23)

We know from Section 1.11.1, that the flux nA is the sum of a term due to convection (Av) and a term due to diffusion jA (kg/m2s): nA = Av + jA Substituting the different flux in Eq. 3.23 gives: A ( A v x ) ( A v y ) ( A v z ) j Ax j Ay j Az + + = rA + + + + x y z x y z t (3.25) (3.24)

For a binary mixture (A,B), Ficks law gives the flux in the u-direction as : wA u (3.26)

j Au = DAB

108

where wA = A/. Expanding Eq. 3.25 and substituting for the fluxes yield:

v y vz A v A + vx + + v y A + vz A + A x + x z z y y x t DAB wA DAB wA DAB wA ( )+ ( )+ ( ) = rA x z z y y x (3.27)

This is the general component balance or equation of continuity for species A. This equation can be further reduced according to the nature of properties of the fluid involved. If the binary mixture is a dilute liquid and can be considered incompressible, then density and diffusivity DAB are constant. Substituting the continuity equation (Eq. 3.13) in the last equation gives:

2 A A 2 A 2 A ) = rA + vx + v y A + vz A DAB 2A + + x t x y z y 2 z 2

(3.28)

This equation can also be written in molar units by dividing it by the molecular weight MA to yield:
2C A 2C A 2C A C A C A C A = RA + vx + vy + vz + + D AB { 2 y z x 2 z 2 44x 44 2444444 4444y 44444 reaction 1 4 4 3 1 4 2 3
Convection Diffusion

accumulation

C A t 13 2

(3.29)

The component balance equation is composed then of a transient term, a convective term, a diffusive term and a reaction term.

3.3 Momentum Balance

We consider a fluid flowing with a velocity v(t,x,y,z) in the cube of Figure 3.3. The flow is assumed laminar. We know from Section 1.11.2 that the momentum is transferred through convection (bulk flow) and by molecular transfer (velocity gradient).

109

yx |y+ y

zx |z

xx |x

xx |x+ x

zx |z+ z

yx |y

y x z

Figure 0-3 Balance of the x-component of the momentum

Since, unlike the mass or the energy, the momentum is a vector that has three components, we will present the derivation of the equation for the conservation of the x-component of the momentum. The balance equations for the y-component and the z-component are obtained in a similar way. To establish the momentum balance for its x-component we need to consider its transfer in the x-direction, y-direction, and z-direction.

Momentum in:

The x-component of momentum entering the boundary at x-direction, by convection is: (vxvx)|x yzt (3.30)

The x-component of momentum entering the boundary at y-direction, by convection is: (vyvx)|y xzt and it enters the z-direction by convection with a momentum: (vzvx)|z xyt (3.32) (3.31)

110

The x-component of momentum entering the boundary at x-direction, by molecular diffusion is: (xx)|x yzt (3.33)

The x-component of momentum entering the boundary at y-direction, by molecular diffusion is: (yx)|y xzt and it enters the z-direction by molecular diffusion with a momentum: (zx)|z xyt Momentum out: The rate of momentum leaving the boundary at x+x, by convection is: (vxvx)|x+x yzt and at boundary y+y,: (vyvx)|y+y xzt and at boundary z+z: (vzvx)|z+z xyt (3.38) (3.37) (3.36) (3.35) (3.34)

The x-component of momentum exiting the boundary x+x, by molecular diffusion is: (xx)|x+x yzt and at boundary y+y: 111 (3.39)

(yx)|y+y xzt and at boundary z+z:: (zx)|z+z xyt 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) yzt The net gravitational force in the x-direction is: g|x x yz t Accumulation is: (vx|t+t vx|t) x yz

(3.40)

(3.41)

(3.42)

(3.43)

(3.44)

Substituting all these equations in Eq. 1.5, dividing by x y zt and taking the limit of each term goes zero gives: ( v x ) ( v x v x ) ( v x v y ) ( v x v z ) P + + + = ( xx + yx + zx ) + g x t x y y z z x x

(3.45)

Expanding the partial derivative and rearranging:


v y v x v z vx + + + t x y z ( v x v x v x v x + t + v x x + v y y + v z z =

(3.46)

yx xx zx P + + + g x ) x y z x

112

Using the equation of continuity (Eq. 3.10) for incompressible fluid, Equation (3.46) is reduced to: v x v v v P + v x x + v y x + v z x = ( xx + yx + zx ) + g x x y z x y z x t (3.47)

Using the assumption of Newtonian fluid, i.e. v x , x v x , y v x z

xx =

yx =

zx =

(3.48)

Equation 3.47 yields:


2v v 2v 2 v P v x v v + v x x + v y x + v z x = 2x + 2x + 2x + g x z 4 3 x 4 4 4 4 4y 1 t 2 3 144x 4 2y 444z 14x 4 24444 1 24 4 3 3 accumulation generation

(3.49)

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:
2v v y v v 2v 2 v P v + v x y + v y y + v z y = 2y + 2y + 2y + g y z y z 4 3 y 1 t 2 3 144x 4 2y 4444 14x 4 24444 1 24 4 4 44 4 3 3 accumulation

(3.50)

transport by bulk flow

transport by viscous forces

generation

2v v 2v 2 v P v z v v + g z + v x z + v y z + v z z = 2z + 2z + 2z z 4 3 z 4 4 4 4 4y 1 t 2 3 144x 4 2y 444z 14x 4 24444 1 24 4 3 3 generation accumulation

(3.51)

transport by bulk flow

transport by viscous forces

These equations constitute the Navier-Stocks equation.

3.4 Energy balance

113

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 + CpTv

(3.52)

Therefore, the energy coming by convection in the x-direction at boundary x is: (qx + CpTvx)yzt Similarly the energy entering the y and z directions are (qy + CpTvy)xzt (qz + CpTvz) x y t The energy leaving the x,y and z directions are: (qx + CpTvx)|x+x yzt (qy + CpTvy)|y+y xzt (qz + CpTvz)|z+z x y t The energy accumulated is approximated by:
(CpT|t+t CpT|t) x yz

(3.53)

(3.54) (3.55)

(3.56) (3.57) (3.58)

(3.59)

114

qy |y qz |z+
z

y x qx |x z qx |x+
x

qy |y+

qz |z

Figure 0-4 Energy Balance in Cartesian coordinates

The rate of generation is H where H includes all the sources of heat generation, i.e. reaction, pressure forces, gravity forces, fluid friction, etc. Substituting all these terms in the general energy equation (Eq. 1.7) and dividing the equation by the term xyzt and letting each of these terms approach zero yield:
( CpT ) ( CpTv x ) ( CpTv y ) ( CpTvz ) qx q y qz + = H + + + + + y z t x y z x

(3.60)

Expanding the partial derivative yields:


v vy vz qx qy qz T T T T + + vx + vy + vz + CpT + x + + + + = H t x x y z y z x y z t

Cp

(3.61)

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 Cp t + vx x + v y y + vz z + x + y + z = H

(3.62)

Using Fourier's law:

115

qu = k

dT du

(3.63)

into the last equation gives:


2T 2T 2T T T T T = k 2 + 2 + 2 + H Cp + Cp vx + vy + vz x z { 44 y 44z 3 generation 1 2t 1444x 4 4y 444 4 4 3 42 4 3 1 4 2 4
accumulation Transport by bulk flow Transport by thermal diffusion

(3.64)

The energy balance includes as before a transient term, a convection term, a diffusion term, and generation term. For solids, the density is constant and with no velocity, i.e. v = 0, the equation is reduced to:
2T 2T 2T T Cp = k 2 + 2 + 2 + H x { 44 y 44z 3 generation 1 2t 4 4 3 1 4 2 4 accumulation
Transport by thermal diffusion

(3.65)

3.5 Conversion between the coordinates

So far we have shown how to derive the equation of change in Cartesian coordinates. In the same way, the equations of change can be written other coordinate systems such as the cylindrical or spherical coordinates. Alternatively, one can transform the equation of change written in the Cartesian coordinates to the others through the following transformation expressions. The relations between Cartesian coordinates (x,y,z) and cylindrical coordinates (r,z,) (Figures 1.3 and 1.4) are the following:
x = rcos(), y = rsin(), z=z

(3.66)

116

Therefore;
y = tan 1 ( ) x

r = x2 + y 2 ,

(3.67)

The relations between Cartesian coordinates (x,y,z) and spherical coordinates (r,,) (Figures 1.3 and 1.5) are:
x = r sin() cos(), y = r sin() sin(), z = r cos()

(3.68)

Therefore; (3.69)

r= x +y +z ,
2 2 2

x2 + y2 = tan ( ), z
1

y = tan 1 ( ) x

Accordingly, we list the following general balance equations in the three coordinates. These equations are written under the assumptions mentioned in previous sections. For the more general case, where density is considered variable the reader can consult the books listed in the references.

3.5.1 Balance Equations in Cartesian Coordinates

Mass Balance
vx v y vz + + )=0 x y z

(3.70)

Component balance for component A in binary mixture with chemical reaction rate RA:
2C A 2C A 2C A C A C A C A C A = D AB + vx + vy + vz x 2 + y 2 + z 2 ) + R A t x y z

(3.71)

117

Energy balance
T T T T 2T 2T 2 T + Cp( v x + vy + vz ) = k( 2 + 2 + 2 ) + H t x y z x y z

Cp

(3.72)

Momentum balance
x component
P 2v 2v 2v v v v v x + g x + (vx x + v y x + vz x ) = ( 2x + 2x + 2x ) x z y x z y x t

(3.73)

y component:
v y t v y x v y y v y z 2v y x 2 2v y y 2 2v y z 2

+ (vx

+ vy

+ vz

) = (

P + g y y

(3.74)

z component
v z v v v 2v 2v 2v P + (vx z + v y z + v z z ) = ( 2z + 2z + 2z ) + g z t x y z x y z z

(3.75)

3.5.2 Balance Equations in Cylindrical Coordinates

Mass balance 1 rvr 1 v vz + + )=0 r r r z

(3.76)

Component balance for component A in binary mixture (A-B)with reaction rate RA:

118

1 C A C A C A C A 1 C A 1 2C A 2C A + + vz + v + vr (r )+ 2 ) + RA = D AB r r r z r r t r 2 z 2

(3.77)

Energy balance T 1 T 1 T 1 2T 2 T T T ) = k( (r )+ 2 ) + H + Cp( vr + v + vz + t r z r r r r r 2 z 2

Cp

(3.78)

Momentum balance r component


2 v v r v v r v v + (v r r + + vz r ) = t r r r z 2 2 vr 1 vr 2 v P 1 rv r ( )+ 2 ) + g r 2 + 2 r r r r r r z 2

(3.79)

component:

vv v v v v v + (vr + + r + vz ) = t r r z r 1 2v 2 v 1 rv 2v P ( )+ 2 + 2 r + 2 ) + g 2 r r r r r z

(3.80)

z component
v v v z v v 1 v z 1 2v z 2v z P + ( vr z + z + v z z ) = ( (r )+ 2 + 2 ) + g z r r r z t r z r r 2 z

(3.81)

3.5.3 Balance Equations in Spherical Coordinates

Mass Balance

119

v 1 (r 2 v r ) (v sin()) 1 1 =0 2 + + r r sin() r sin() r Component balance for component A in binary mixture (A-B) with reaction rate RA :
v C A C A v C A C A = + vr + + t r r r sin( ) 1 2 C A 1 1 C 2C A (r )+ 2 (sin( ) A ) + 2 2 ) + RA DAB 2 r r r r sin( ) r sin ( ) 2

(3.82)

(3.83)

Energy balance v T T v T T + Cp v r r + r + r sin( ) = t 1 T T 2T 1 1 + H )+ 2 (sin( ) ) + 2 2 k 2 (r 2 r r r r sin( ) r sin ( ) 2 Momentum balance r component


2 2 v v v v vr vr v + v = + vr r + r + r r t r r sin( ) 1 2 2 v 2vr P 1 1 (sin( ) r ) + 2 2 2 2 ( r vr ) + 2 + g r r r r sin( ) r sin ( ) 2 r

Cp

(3.84)

(3.85)

component:
2 v v v v v vr v v cot( ) v = + + vr + + r r r sin( ) r r t 1 2 v sin( )v 2v 1 1 2 (r )+ 2 (sin( ) )+ 2 2 r r r r r sin ( ) 2 v 1 P 2 v cos( ) + 2 r + g r sin 2 ( ) r

(3.86)

120

component
v v v v v vr v v v cot( ) = + vr + + + + r t r r sin( ) r r

1 2 v 2v 1 1 sin( )v 1 (r )+ 2 ( )+ 2 2 r 2 r r r sin( ) r sin ( ) 2 P 2 2 cos( ) v 1 vr + 2 r sin( ) + r 2 sin 2 ( ) r sin( ) + g

(3.87)

3.6 Examples of Application of Equations of change

Practically all the microscopic balance examples treated in the previous chapter can be treated using the equations of change presented in this chapter. In this section we review some of the previous examples and present additional applications.

3.6.1 Liquid flow in a Pipe

To model the one dimensional flow through the pipe of an incompressible fluid (Example 2.2.1) we may use the continuity balance. For constant density we have the continuity equation in cylindrical coordinates (Eq. 3.76). 1 rvr v vz + + =0 r r r z

(3.88)

The plug flow assumptions imply that vr = v = 0, and the continuity balance is reduced to: vz =0 z

(3.89)

3.6.2 Diffusion with Chemical Reaction in a Slab Catalyst

121

To model the steady state diffusion with chemical reaction of species A in a slab catalyst, (Example 2.2.3) we use the equation of change (Eq. 3.71). The fluid properties are assumed constant, C A C A C A C A 2C 2C A 2C A + vx + vy + vz D A ( 2A + + ) = RA t x y z x y 2 z 2

(3.90)

Since the system is at steady state we have

C A = 0 . If we assume t

that there is no bulk flow


= =0. x y

then vx = vy = vz = 0. For diffusion in the z-direction only, the following holds: The equation is then reduces to: d 2C A = RA dz 2

DA

(3.91)

3.6.3 Plug Flow Reactor

The isothermal plug flow reactor (Example 2.2.5) can be modeled using the component balance equation (3.77). The plug flow conditions imply that vr = v = 0 and
= =0. r

Equation (3.77) is reduced to:

C A C A 2C A = v z + DAB RA t z z 2

(3.92)

For the non-isothermal plug flow reactor (Example 2.2.6), the energy balance is obtained by using Eq. 3.78. For fluid with constant properties and at constant pressure, we have: T T v T T 2T 1 T 1 2T 2T + Cp (v r + + vz ) = k( 2 + + + ) + H z r r r 2 2 z 2 t r r r

Cp

(3.93)

122

Using the plug-flow assumptions, vr = v = 0 and

= = 0 , and neglecting the viscous r

forces, the term H includes the heat generation by reaction rate RA and heat exchanged with the cooling jacket, htA(T Tw). Equation (3.93) is reduced to:

Cp

T T 2T D = Cpv + k 2 H r ko e E / RT C A ht (T Tw ) t z z A

(3.94)

3.6.4 Energy Transport with Heat Generation

Consider the example of a solid cylinder of radius R in which heat is being generated due to some reaction at a uniform rate of H (J/m2s). A cooling system is used to remove heat from the system and maintain its surface temperature at the constant value Tw (Figure 3.5). Our objective is to derive the temperature variations in the cylinder. We assume that the solid is of constant density, thermal conductivity and heat capacity. Clearly this is a distributed parameter system since the temperature can vary with time and with all positions in the cylinder. We will use then the equation of change (Eq. 3.78) in cylindrical coordinates for a solid: T k 2T 1 T 1 2T 2T ( 2 + = + 2 2 + 2 )+ H t Cp r r r r z Cp (3.95)

Figure 0-5 Cylindrical solid rod

A number of assumptions can be made:

123

The system is at steady state i.e.

T =0 t

The variation of temperature is only allowed in radial directions. Therefore, the terms
2T z 2

and

2T 2

are zero.

The energy balance is reduced to:


k d 2T 1 dT ( 2 + )+ H Cp dr r dr Cp

0=

(3.96)

Or equivalently: d 2T 1 dT + )= H 2 dr r dr k

(3.97)

with the following boundary conditions:

The temperature at the wall is constant: T (r = R ) = Tw (3.98)

The maximum temperature will be reached at the center ( r = 0), therefore: dT =0 dr (3.99)

at r = 0

Note that Equation (3.97) can also be written as follows:

1 d dT (r )= H r dr dr k since qr = k dT/dr, this equation is equivalent to:

(3.100)

124

1 d (rqr ) = H r dr

(3.101)

The left hand side is the rate of diffusion of heat per unit volume while the right hand side is the rate of heat production per unit volume.

3.6.5 Momentum Transport in a Circular Tube

We revisit example 2.2.2 where we derived the steady state equations for the laminar flow inside a horizontal circular tube. We will see how the model can be obtained using the momentum equation of change. We assume as previously that the fluid is incompressible and Newtonian. The momentum equations of change in cylindrical coordinates are given by Eqs. (3.79-3.81). A number of simplifications are used: The flow has only the direction z, i.e. vr = v = 0. The flow is at steady state,
v z =0 t

The momentum equation in cylindrical coordinates Eq. 3.81 is reduced to: 2 v 1 vz 1 2vz 2vz P vz = 2z + + + + g z r z r r r 2 2 z 2 z

vz

(3.102)

Using the continuity equation (Eq. 3.76), and since vr = v = 0 gives: dv z =0 dz

(3.103)

We also note that because the flow is symmetrical around the z-axis we have necessarily no variation of the velocity with , i.e. 2vz 2 =0 Equation 3.102 is then reduced to 125

(3.104)

d 2 v z 1 dv z dP = + 2 r dr dz dr dP is constant. dz

(3.105)

Since the left hand side depends only on r, this equation suggests that Therefore: dp P = dz L

(3.106)

where P is the pressure drop across the tube. Equation 3.105 is equivalent to:

d 2 v z 1 dv z P = + 2 r dr L dr

(3.107)

with the following conditions identical to those in Example 2.2.2. Note also that Eq. 3.107 can also be written as: d rv z P ( )= dr dr L

(3.108)

which is also equivalent to: d ( r rz ) P = rdr L

(3.109)

where rz is the shear stress. The left term is the rate of momentum diffusion per unit volume and the right hand side is in fact the rate of production of momentum (due to pressure drop). Note then the similarity between Eq. 3.109 for momentum transfer with Eq. 3.101 for heat transfer.

126

3.6.6 Unsteady state Heat Generation

We reconsider Example 3.6.4 but we are interested in the variations of the temperature of the reactor with time as well. This may be needed to compute the heat transferred during start-up or shut-down operations. Keeping the same assumptions as Example 3.3.4 (except the steady state assumption), the energy balance in cylindrical coordinates yields: T 1 d dT (r )+ H =k r dr dr t k

Cp

(3.110)

with the initial and boundary conditions:


T ( R, t ) = Tw dT | r =0 = 0 dr T (r ,0) = Tw (3.111) (3.112) (3.113)

3.6.7 Laminar Flow Heat Transfer with Constant Wall Temperature

We consider a fluid flowing at constant velocity vz into a horizontal cylindrical tube. The fluid enters with uniform temperature Ti. The wall is assumed at constant temperature Tw. We would like to model the variations of the fluid temperature inside the tube. To apply the energy equation of change (Eq. 3.78) we will assume that the fluid is incompressible, Newtonian and of constant thermal conductivity. Since the system is at steady state d/dt = 0 and the flow is one-dimensional vr = v = 0, the energy equation in cylindrical coordinates Eq. 3.78 is reduced to: T 2T 1 T 1 2T 2T ) = k( 2 + + + t r r r 2 2 z 2 r

Cpv z

(3.114)

127

Figure 0-6 heat transfer with constant wall temperature

Since the temperature is symmetrical then conduction term


2T z 2

2T 2

= 0. In some cases we can neglect the


T z

compared to the convective term v z

. The system is then described

by the following energy and momentum equations: T 2T 1 T ) = k( 2 + t r r r d rv z P ( )= dr dr L (3.115)

Cpv z

(3.116)

These two equations are therefore coupled by vz, with the previous boundary conditions for vz, At r = R, At r = 0, At z = 0, At r = 0, At r = R, vz = 0 dvz/dr = 0 T = Ti dT/dr = 0 T = Tw (3.117) (3.118) (3.119) (3.120) (3.121)

128

3.6.8 Laminar Flow and Mass Transfer

We consider the example of a fluid flowing through a horizontal pipe with constant velocity vz. The pipe wall is made of a solute of constant concentration CAw that dissolved in the fluid. The concentration of the fluid at the entrance z = 0 is CAo. The regime is assumed laminar and at steady state. The fluid properties are assumed constant. We would like to model the variation of the concentration of A along the axis in the pipe. The component balance for A (Eq. 3.77) is:
C A ) 2 2 r + C A + C A ) rr r 3 2 z 2

C A C A C A = D AB ( + vz + v vr z r

(r

(3.122)

Since vr = v = 0, the balance equation becomes:

C A = D AB ( vz z
2C A z 2

(r

C A ) 2 r + C A ) rr z 2

(3.123)

If the diffusion term


vz

in the z-direction is negligible compared to convection term

C A then the last equation reduces to: z

C A ) (r C A r vz = D AB rr z

(3.124)

The equation (Eq. 3.116) is unchanged. The two equations are coupled through vz. The boundary conditions are analogous to the previous example:

129

At r = R, At r = 0, At z = 0, At r = 0, At r = R,

vz = 0 dvz/dr = 0 CA= CAo dCA/dr =0 CA= CAw

(3.125) (3.126) (3.127) (3.128) (3.129)

Note the similarity between this example and the heat transfer case of the previous example.

130

Problems: Lumped parameter systems


Problem 1

Consider the three storage tanks in series shown by Figure X-1. 1. Write the mathematical model that describes the dynamic behavior of the process. Assume the density of the fluid constant. 2. Identify the states of the process and the degrees of freedom.

F0

F1

Fr

F2

h1

h2
Figure X-1

h3

F3

Problem 2

Consider the three cooling storage tanks in series shown in Figure X-2.
Ff , Tf F0, T0 F1 , T 1 F3 , T 3

h1

h2

h3

F4, T4

wc1, Tc1

wc2, Tc2

wc3, Tc3

Figure X-2 1. Assuming constant fluid properties, write the mathematical model for the process. Identify the states and the degrees of freedom. 131

2. Assume now the following non-isothermal reversible first-order liquid-phase reaction: A B is taking place in the three above CSTR in series. Assuming all CSTRs are adiabatic, write down the unsteady state model for the process if the holdups is kept constant in all reactors.

Problem 3

Consider the well-stirred non-isothermal reactor shown in Figure X-3 below. Write down the mathematical equations that describe the dynamic behavior of the fundamental quantities of the process. Consider an exothermic liquid-phase second order reaction rate in the form of A B is taking place. The density is assumed to be constant, develop the necessary equations describing the process dynamic behavior.

F1, CA1

h1

F3, CA3

Figure X-3

Problem 4

Consider a two CSTRs in series with an intermediate mixer introducing a second feed as shown in Figure X-4. A first order irreversible exothermic reaction: A B is carrier out in the process. Water at ambient temperature (Tc1i and Tc2i) is used to cool the reactors. The densities and heat capacities are assumed to be constant and independent of temperature and concentration. Develop the necessary equations describing the process dynamic behavior. Note that the mixer has negligible dynamics and that the inlet feed to CSTR2 has the same temperature as that of the outlet of CSTR1.

132

Q1, C1f, T1f QC1, TC1 QC1, TC1i Q2, C2, T2 Mixer CSTR 1 QC2, TC2i Q4, C4, T4 CSTR 2 QC2, TC2 Q3, C3, T3

Figure X-4
Problem 5:

Consider a thermometer bulb with pocket is immersed in hot fluid of temperature T1 as shown in Figure X-5 below. The pocket temperature is T2 (assumed uniform through the thickness) and that for the bulb is T3. Write the model equation that describes how T2 and T3 varies with time.

T1

T2 T3

Figure X-5

Problem 6

Consider the single-effect steam evaporator shown in Figure X-6. A salty water (brine) with mass fraction Cb0 and mass flow rate B0 is fed to the evaporator where it is heated with saturated steam with mass flow rate W. The concentrated product comes out of the evaporator

133

with mass flow rate B1 and mass fraction Cb1 while the vapor is withdrawn from the top with mass flow rate V. Develop the dynamic model that describe the process behavior. Assume the heat supplied by the steam is mainly equal to the heat used for vaporizing the brine.

Vapor

Steam

Condensate Feed Concentrated liquid


Figure x-6

Problem 7

Consider the process shown in Figure X-7. A stream of pure component A is mixed with another stream of a mixture of component A and B in an adiabatic well stirred mixing tank. The effluent is fed into an adiabatic CSTR where the following reaction takes place: A + B C. Assume the process is isothermal. Develop the dynamic model for the process and determine the degrees of freedom assuming constant fluid properties.
Q2, CA2

Q1, CA1

Q3, CA3 Q4, CA4

Figure X-7

134

Problem 8:

Consider the following irreversible first-order liquid-phase reaction: A k1 B k2 C where k1 and k2 are the reaction rate constants in sec-1. 1. Write down the unsteady state model for the process if the reaction takes place in an isothermal well-mixed CSTR. 2. Write down the unsteady state model for the process if the reaction takes place in a non-isothermal well-stirred batch reactor, where the heat needed for the endothermic reaction is supplied through electrical coil. 3. Repeat part 2 but assuming that the reactor is cooled by cold fluid flowing in a jacket and that the reactor wall has high resistance to heat transfer.

Problem 9:

A perfectly mixed non-isothermal adiabatic reactor carries out a simple first-order exothermic reaction, A B in the liquid phase. The product from the reactor is cooled from the output temperature T to Tc and then introduced to a separation unit where the un-reacted A is separated from the product B. The feed to the separation unit is split into two equal parts top product and bottom product. The bottom product from the separation unit contains 95% of the un-reacted A in the effluent of the reactor and 1% of B in the same stream. The bottom product which is at Temperature Tc (since the separation unit is isothermal) is recycled and mixed with the fresh feed of the reactor and the mixed stream is heated to temperature Tf before being introduced to the reactor. Write the steady state mass and energy balances for the whole process assuming constant physical properties and heat of reaction. Discuss also the degree of freedom of the resulting model. The process is depicted in figure X-8

135

CSTR

V= 0.5 F

F, Tf

CA, CB , Tc F

Separator

Fo, To ,CAo L = 0.5 F

Figure X-8
Problem 10:

Consider a biological reactor with recycle usually used for wastewater treatment as depicted in figure X-9 below. Substrate and biomass are fed to the reactor with concentrations Sf and Xf. The effluent form the well-mixed reactor is settled in a clarifier and a portion of the concentrated sludge is returned to the reactor with flow rate Qr and Xr. If the reaction of substrate in the clarifier is negligible, the recycle stream would contain the same substrate concentration as the effluent from the reactor. Sludge is withdrawn directly from the reactor with a fraction W. Assuming constant holdup develop the dynamic model for the process. Assume the rate of disappearance of S is given by: r = mS/(K + S) and the rate of generation of X by r/Y where Y is constant.

W X S Q Xf Sf Reactor V X S Qr , Xr , S Q-W Xt S

Settler

Figure X-9

Problem 11

136

Consider the two phase reactor and a condenser with recycle. Gaseous A and liquid B enters the reactor at flow rates FA and FB respectively. Gas A diffuses into the liquid phase where it reacts with B producing C. The latter diffuses into the vapor phase where B is nonvolatile. The vapor phase is fed to the condenser where the un-reacted A is cooled and the condensate is recycled back to the reactor. The product C is withdrawn with the vapor leaving the condenser. For the given information develop the dynamic model for the process shown in Figure X-10. Consider all flows are in moles.

F1v, yA1, yc1 P2 T1, P1 FA, TA


NA NA

T2

FB, TB

NC

F2L,xA2 Q1

A + B = 2C

F1L, xA1, xB1, T1

Figure X-10
Problem 12:

Develop the mathematical model for the triple-effect evaporator system shown in figure X-11 below. Assume boiling point elevations are negligible and that the effect of composition on liquid enthalpy is neglected.

137

V1=F-L1
Vapor

V2=L1-L2
Vapor

V3=L2-L3
Vapor

F (feed) Vo Vo L1
Thick Liquor

V1 L2
Thick Liquor

V2 L3

Thick Liquor

Figure X-11

Problem 13

Liquid vaporizer as depicted in Figure X-12 below is one of the important processing units in a chemical plant. A liquid feed, enters the vaporizer at specific flow rate F0 density 0 and temperature T0. Inside the vessel, the liquid is vaporized by continuous heating using hot oil. The mass rate of vaporization is wn. The formed vapor is withdrawn continuously from the top of the vessel. We would like to develop a mathematical model to describe the process. Unlike the adiabatic flash operation, the temperature and pressure in the two phases are different. Correspondingly, the volume of the two phases varies with time.

Fv Pv , Vv , v Fo , To , o wv PL, VL, L Q

Figure X-12

138

Problems: Distributed Parameter Systems


Problem 14

At time t = 0 a billet of mass M, surface area A, and Temperature To is dropped into a tank of water at Temperature Tw. Assume that the heat-transfer coefficient between the billet and water is h. If the mass of water is large enough that its temperature is virtually constant, determine the equation describing the variation of billet temperature with time for the following two cases: 1. The thermal conductivity of the metal in the billet is sufficiently high that the billet is of uniform temperature. 2. The thermal conductivity of the metal is low and the billet radius is large enough that the temperature inside the billet is not uniform.

Problem 15

(a)Consider a wall made up of stacked layers of various materials each with different thermal conductivity usually used for insulation as shown by figure X-13a. Assume the inner side is at high temperature Tb and the outer side is at the ambient temperature Ta. Derive the temperature profile through the wall.

Tb

Ta

Figure X-13a (b) Repeat the above development for a composite cylindrical wall shown by figure X-13b.

139

r Tb Ta

Figure X-13b

Problem 16

A chemical reaction is being carried out in a fixed bed flow reactor. The reaction zone is filled with catalyst pellets. Assume plug flow and the reactor wall is well insulated such that the temperature is uniform in the radial direction. If the fluid enters the reactor with temperature T1 and superficial velocity v and that thermal energy per unit volume (S) is produced inside the reactor due to chemical reaction, derive the steady-state temperature axial distribution. The unit is depicted in figure X-14.

Insulating wall

Catalyst particles

Reactant

Product

z=0

Reaction zone

z=L

Figure X-14

Problem 17

A first order irreversible reaction is taking place in a tubular reactor. The reactor can be considered to be isothermal and radial dispersion is not to be neglected. Using Ficks law with effective diffusion coefficient to describe the axial and radial dispersion, derive the steady state mass balance equation and its boundary conditions.
Problem 18

140

(a) Derive the unsteady state mass balance equation for the diffusion of species A through a spherical porous pellet. (b) Derive the steady state mass balance equation for the diffusion of species A through a spherical porous pellet where it converts to B according to the first-order reaction: A B.

Problem 19

Species A dissolves in liquid B and diffuses into stagnant liquid phase contained in a cylindrical tank where it undergoes a homogeneous irreversible second-order chemical reaction: A+B C. Derive the mass balance equation for component A. Assume that the depth of liquid is so large compared to the radius of the tank, thus, the radial distribution can be neglected. The process is shown schematically in figure X-15.

Gas A

z=0 Liquid B z

z=L
Figure X-15

141

S-ar putea să vă placă și