You are on page 1of 144

Process Calculations and

Reactor Calculations
For Environmental Engineering

Per Warfvinge

c
!Per
Warfvinge
Department of Chemical Engineering
P. O. Box 124
221 00 Lund
Edition for academic year 2009/10

Contents
Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1 Why do we perform process and reactor calculations?
1.1 Processes and reactors . . . . . . . . . . . . . . . . . . . . .
1.1.1 System boundaries and sub-systems . . . . . . . . .
1.1.2 Non-reaction and reaction systems . . . . . . . . . .
1.1.3 Element, components and inert substances . . . . .
1.1.4 Static and dynamic systems . . . . . . . . . . . . . .
1.2 The mass balance principle . . . . . . . . . . . . . . . . . .
1.3 Process calculations and reactor calculations . . . . . . . . .
1.4 Numerical methods . . . . . . . . . . . . . . . . . . . . . . .
1.4.1 Systems of linear algebraic equations . . . . . . . . .
1.4.2 Numerical solution of systems of algebraic equations
1.4.3 Solution of differential equations . . . . . . . . . . .

3
6
7

.
.
.
.
.
.
.
.
.
.
.

8
10
11
12
12
13
13
15
15
15
16
17

2 Process calculations for non-reaction systems


2.1 Component mass balances . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Degrees of freedom analysis . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Process calculations for a separation process . . . . . . . . . . . . . . .

22
23
24
26

3 Process calculations with multiple units


3.1 Calculation methodology . . . . . . . . .
3.2 Common types of units . . . . . . . . .
3.2.1 Mixer . . . . . . . . . . . . . . .
3.2.2 Splitter . . . . . . . . . . . . . .
3.2.3 Recirculation . . . . . . . . . . .
3.2.4 Bypass . . . . . . . . . . . . . . .
3.2.5 Purge . . . . . . . . . . . . . . .
4 Process calculations for reaction
4.1 Mass balances mass or mole?
4.2 Important concepts . . . . . . .
4.2.1 Inert . . . . . . . . . . .
4.2.2 Conversion . . . . . . .
4.2.3 Yield . . . . . . . . . . .
4.2.4 Selectivity . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

32
32
33
33
33
36
36
36

systems
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

41
41
42
42
43
43
43

Contents

4.3

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

44
45
47
47
48
49

5 Reactor calculations
5.1 Kinetic rate equations . . . . . . . . . . . . . . . . .
5.1.1 First order kinetics . . . . . . . . . . . . . . .
5.1.2 Second order kinetics . . . . . . . . . . . . . .
5.1.3 Overview of kinetic rate equation . . . . . . .
5.1.4 Reversible reactions . . . . . . . . . . . . . .
5.1.5 Consecutive reactions . . . . . . . . . . . . .
5.1.6 Parallel reactions . . . . . . . . . . . . . . . .
5.2 Mean residence time and reaction time . . . . . . . .
5.3 Reactor models . . . . . . . . . . . . . . . . . . . . .
5.3.1 The differential mass balance . . . . . . . . .
5.3.2 Methodology for reactor calculations . . . . .
5.4 The ideal completely stirred tank reactor, CSTR . .
5.5 The ideal batch reactor . . . . . . . . . . . . . . . .
5.6 The ideal plug-flow reactor, PFR . . . . . . . . . . .
5.6.1 PFR reactor modeling in terms of conversion

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

52
54
54
55
55
56
56
57
57
60
60
62
62
66
71
72

6 Non-ideal reactors
6.1 Examples of non-ideal mixing . . . . .
6.2 Residence-time distributions . . . . . .
6.2.1 The normalized RTD E(t) . .
6.2.2 c(t) for an ideal CSTR . . . . .
6.2.3 E(t) for an ideal CSTR . . . .
6.2.4 Mean residence time from E(t)
6.2.5 The F(t)-distribution . . . . . .
6.3 Non-ideal reactor models . . . . . . . .
6.3.1 The CSTR in series model . . .
6.4 The segregation model . . . . . . . . .
6.5 Other reactor models . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

76
76
77
78
79
81
81
82
85
85
87
90

7 Instationra CSTR
7.1 Svar p ndring i ingngskoncentration . . . . . . . . . . . . . . . . . .
7.2 Arbetsmetodik . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

91
92
94

4.4

The element mass balance method . .


4.3.1 The atomic matrix . . . . . . .
The component mass balance method
4.4.1 The reaction parameter . . .
4.4.2 The reaction matrix . . . . . .
4.4.3 Degree of freedom analysis . .

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

8 Dispersion in porous media


8.1 Water flow- Darcys law . . . . . . . . . . . . . .
8.1.1 Flow rate distribution - reactor modeling
8.1.2 Advective transport . . . . . . . . . . . .
8.2 Dispersion . . . . . . . . . . . . . . . . . . . . . .
8.2.1 Molecular diffusion - Ficks law . . . . . .
8.2.2 Analogy dispersion - diffusion . . . . . . .
8.2.3 Deduction of expression for spread . . . .

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

99
100
103
104
104
104
105
106

Contents

8.3

8.2.4 Interpretation of the standard deviation . . .


8.2.5 Application to a advection-dispersion system
Dispersion coefficients in groundwater aquifers . . .
8.3.1 Dispersivitet . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

9 Transport in porous media


9.1 The continuity equation - advection and dispersion . . . . . . .
9.1.1 Continuous source . . . . . . . . . . . . . . . . . . . . .
9.2 Advection, dispersion and reaction . . . . . . . . . . . . . . . .
9.2.1 Continuous source . . . . . . . . . . . . . . . . . . . . .
9.2.2 Continuous addition during a limited time - step up and
9.2.3 Steadystate . . . . . . . . . . . . . . . . . . . . . . . .
9.3 Adsorption and transport . . . . . . . . . . . . . . . . . . . . .
9.3.1 Advection, dispersion, reaction and adsorption . . . . .
9.3.2 Determination of Kd and R for organic substances . . .
9.4 The continuity equation in multiple dimensions . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

108
109
110
110

. . . .
. . . .
. . . .
. . . .
down
. . . .
. . . .
. . . .
. . . .
. . . .

113
113
114
116
117
117
118
119
121
123
123

10 Bioreaction engineering - biofilms


125
10.1 Basic concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
10.2 Kinetic equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
10.2.1 Kinetic equations for conversion on cellular level . . . . . . . . 127
10.2.2 Rate equations for conversion in reactors . . . . . . . . . . . . 130
10.2.3 Rate equation for the microbial population . . . . . . . . . . . 130
10.3 Biofilms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
10.3.1 General rate equations for biofilms . . . . . . . . . . . . . . . . 135
10.3.2 Concentration profile and flux in a thin film for 0th order reaction135
10.3.3 Concentration profile and flux in a deep film for 0th order reaction137
10.3.4 Reactor calculations . . . . . . . . . . . . . . . . . . . . . . . . 140
10.4 External mass transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
10.4.1 Mass transfer in packed beds . . . . . . . . . . . . . . . . . . . 141
10.4.2 Coupled external mass transfer resistance and diffusion/reaction
in a biofilm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

Examples

Examples
11: Balance calculation for a bank account . . . . . . . . . .
12: Solving linear equation systems . . . . . . . . . . . . . .
13: Solving linear equations with fsolve . . . . . . . . . . .
14: Solving non-linear equations with fsolve . . . . . . . .
15: Numerical integration with Eulers method . . . . . . .
16: Numerical integration with ode45 . . . . . . . . . . . .
21: Separation of ethanol and water . . . . . . . . . . . . .
22: Separation of ethanol and water - 6 unknowns . . . . .
23: Ethanol and water concentration constraints . . . . .
24: Separation of ethanol and water moved constraint . .
31: Limitation of the number of splitter constraints . . . . .
32: Process calculation based on total system analysis . . .
33: Degree of freedom analysis based on sub-systems . . . .
41: Catalytic dehydrogenation . . . . . . . . . . . . . . . . .
42: Atomic matrix with linearly dependent rows . . . . . .
43: The rank of a reaction matrix . . . . . . . . . . . . . . .
44: Application of the reaction parameter method . . . . .
51: Average residence time of a lake . . . . . . . . . . . . .
52: CSTR with a 1st order irreversible reaction . . . . . . .
53: Example 52 with matrix notation . . . . . . . . . . . .
54: Numerical solution for non-unity reaction order . . . . .
55: Batch reactor with 1st order irreversible reaction . . . .
56: Differential, numerical solution of Example 55 . . . . .
57: Integral, numerical solution of Example 55 . . . . . . .
58: Simulation of consecutive and parallel reactions . . . . .
59: Emission of a pollutant into a stream . . . . . . . . . .
61: Residence-time distribution from tracer experiments . .
62: Series with an infinite number of CSTR . . . . . . . . .
63: The segregation model applied to example 61 . . . . .
71: Exempel p simulering av instationr tankreaktor . . .
81: Estimation of flow velocity . . . . . . . . . . . . . . . .
82: Diffusive and dispersive spread . . . . . . . . . . . . . .
91: Infiltration of a pollutant into a soil . . . . . . . . . . .
92: Complex pollutant transport in a soil . . . . . . . . . .
101: Simulation of a biochemical reaction in a batch reactor
102: Filmtjocklek vid exakt penetrerad film . . . . . . . . .
103: Design of CSTR reactor volume . . . . . . . . . . . . .
104: Boundary layer thickness in a packed bed . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

14
16
17
17
19
21
26
28
29
30
35
37
40
44
45
48
50
59
63
64
65
67
68
69
70
73
83
87
89
95
102
110
116
122
130
139
140
144

Examples

Notations
Note that the choice of mass unit (g, kg, mol) often is arbitrary

Properties introduced in Chapters 1-7


cj
Fi,j
k
Q
V
Wi,j
xi,j
X
yi,j

Mole or mass concentration of substance j


Molar flux of substance j in stream i
Kinetic rate constant
Volume flux
Volume
Mass flux of j in stream i
Mole fraction of j in liquid stream i
Conversion with respect to substance j
Mole fraction of j in gas stream i

i,j
i,j
i,j

Stoichometric coeffficient for j in reaction i


Mass fraction of substance j in liquid stream i
Mass fraction of substance j in gas stream i
Residence time in a reactor

mol L1 , kg L1
mol or mol s1
Varies
m3 s1
m3
kg or kg s1

time1

Properties introduced in Chapters 8-10


B
D
E
J
km
KX
Kd
Koc
L
LB
LP
R
S
u
v

Penetration fraction
Diffusion coefficient
Dispersion coefficient
Diffusive flux
Mass transfer coefficient
Hydraulic conductivity
Distribution coefficient
Distribution coefficient
Length coordinate
Thichness of biological film
Penetration depth
Retardation factor
Sorbed amount
Linear flow velocity
Darcy velocity

Dispersivity
Boundary layer thickness
Porosity
Effective porosity
Substrate turnover rate
Bulk density
Volumetric water content

m s1
m s1
mass m2 s1
m s1
m s1
m3 kgsoil 1
m3 kgSOM 1
m
m
m
g kg1
m s1
m s1
m
m
g s1
kg m3

CHAPTER

Why do we perform process and


reactor calculations?
There are several situations when an engineer needs to gain an understanding of how
different substances flow and react in a system. The system in question may exist in
an industry where a substance should be produced, destroyed or redistributed. The
analyses of such processes is based on the principle of conservation of matter, and
the methods are built upon mass balance. In the chemical industry, and in chemical
engineering, process and reactor calculations based on mass balances are the single
most important tool for the analysis and design of chemical processes.
For natural systems such as lakes, streams, groundwater aquifers as well as entire
ecosystems, mass balances are used to describe and predict how matter is transported
and transformed. In nature, transformations of matter are often controlled by processes that man cannot influence, making the design aspect less important for natural
systems.
Every environmental engineer needs to know how to perform mass balance calculations for different systems. Just as importantly, they also have great use of a
profound ability to think in terms of mass conservation and mass balances. Indeed,
the ability to think abstractly in terms of the principles of conservation is a hallmark
of engineers.
Mass balances play an important role in environmental studies and environmental
science. For example, atmospheric transport models, groundwater models and models
to predict the water quality of lakes and streams are all based on mass balances for
chemical components. The integrated models used to predict climate change also
include mass balance equations for different compartments and pools relevant for the
global carbon cycle.
This compendium treats various techniques to carry apply mass balance calculations to natural and engineered systems. The notations used throughout the text are
given on the previous pages. The text is structured as follows:
This Introductory chapter provides a brief introduction to mass balance calculations. The key concepts introduced are systems, system boundaries, the mass balance
principle, process and reactors, as well as the entities mass fraction and molar fraction,
and examples of numerical methods.

Why do we perform process and reactor calculations?

Process calculations for non-reaction systems introduces component balances


and element balances, the concept of degrees of freedom, and how to solve systems
of equations by introducing a basis of calculation, independent information and total
constraint.
Process calculations for multiple units shows how to solve mass balances problems for multiple, connected units or subsystems. The chapter shows how to develop
the degree of freedom analysis. A vast number of examples are presented.
Process calculations for reaction systems addresses how the calculations must
be changed when chemical reactions occur in the system. It discusses stoichiometry, element matrix, reaction matrix and reaction parameter. Other concepts include
conversion, yield and selectivity.
Reactor calculations treats what happens inside of chemical reactors. The chapter addresses kinetic rate equations, ideal reactor models such as the tank reactor
CSTR, the batch reactor, and plug flow reactor. This chapter is important for gaining an understanding of how a chemical reaction is influenced by and affects its
surroundings, the reactor.
Non-ideal reactors describes how to manage systems that are not perfectly mixed.
It introduces the concepts of residence time distribution and the CSTR-in-series model
and the segregation model.
Chemical reactors under non-steady-state conditions are only discussed briefly,
but rather illustrated by examples and calculation tasks. This is because the theory
has already been introduced in previous chapters. Exercises aim primarily to provide
skills in modeling and simulation methods. Here, the systems are described in terms
of differential mass balances and simulated with differential equations.
Dispersion in porous media takes up irregular flow in groundwater systems,
among others. In such systems, mixing can be described by models analogous to
those for molecular diffusion. Key concepts are advection, dispersion and dispersivity.
Transport in porous media treats mass transport especially in the soil and
groundwater. Models describing the physical processes advection and dispersion as
well as the chemical processes reaction and adsorption are discussed and applied. New
key concepts are distribution factor and retardation factor.
Bioreaction engineering treats the principles and equations for biological reaction
systems. Special attention is given to biofilms with respect to the penetration rate,
reaction order and film mass transport resistance.

1.1 Processes and reactors

10

CO2
N2

Fe3+
org-C
NO3P

org-C
org-N
org-P
solids

org-C
org-N
org-P

org-C
NH4+
org-P

org-C
NO3P

sludge

sludge

solids

solids
separation

O2

denitrification

org-C
NO3P

nitrification

sludge

P-precipitation

Figur 1.1: Examples of a wastewater treatment process which includes physical,


chemical and biological sub-processes.

1.1

Processes and reactors

Figure 1.1 is a schematic illustration of a treatment plant for municipal wastewater.


The goal of treatment is to remove contaminants so that the water may be displaced
to a recipient in a safe and environmentally friendly way. The most importance
polluting substances are carbon (C), nitrogen (N) and phosphorus (P). In sewage,
however, these are not present as pure elements or as easily identified ions, but in the
form of organic substances, such as carbohydrates and proteins.
A water treatment plant contains many parts where various physical, biological
and chemical processes take place. Many issues related to the design and the operation
of such plants involve quantitative estimates. Process and reactor calculations based
on mass balance calculations provide powerful tools to address questions such as:
How much sludge is produced, and how high is the P content in the sludge?
How effective is the denitrification process?
Is enough carbon available in the first biological treatment step (denitrification)?
How large should the recirculation of NO
3 from the nitrification process to the
denitrification process be?
How much will the purification efficiency decrease if 25 % more households are
connected to the sewage treatment plant?
How will the system react when the temperature drops in winter?
Some of the above questions are about how the whole treatment plant operates
under steady-state conditions. Other questions arise when the system is exposed to a
disturbance (e.g., increased load).

1.1 Processes and reactors

11

CO2
N2

Fe3+
org-C
NO3P

org-C
org-N
org-P
solids

org-C
org-N
org-P

org-C
NH4+
org-P

org-C
NO3P

sludge
solids

solids
separation

sludge
O2

denitrification

org-C
NO3P

nitrification

sludge

P-precipitation

Figur 1.2: Example of how system boundaries can be drawn around the different
units in a process.

1.1.1

System boundaries and sub-systems

All the above questions deal with how much material that flows at different points
in the treatment plant. When one calculates the magnitude of these flows, one must
first define the systems boundaries. If one is only interested in the total flows in and
out, the whole plant is chosen when the system boundaries are drawn, as in Figure
1.1. Then it is possible to calculate how the process works as a whole: How much
material goes into treatment plant (as liquid), and how much leaves either as a gas,
liquid or solid (sludge).
A general rule is that one can only calculate the mass flows in the streams crossing
the system boundary. Hence, one cannot say anything about the mass flows between
the various units within waste water treatment systems, in cases where the system
boundary is drawn outside all minor units.
If we want to be able to say something about the individual sub-systems, we
must develop our systems analysis and draw the system boundary in a different manner. If we, for example, want to calculate how effective the last sub-system (the Pprecipitation) is, we must draw the system boundary as shown in Figure 1.2. There
the system has been confined so that all components that enter or leave this unit are
represented by streams across the system boundary.

1.1 Processes and reactors

1.1.2

12

Non-reaction and reaction systems

The sub-systems mentioned in the previous section, the sand trap and the biological
purification step were selected to illustrate two fundamentally disparate systems. The
sand trap is an example of a non-reaction system which characterized by the fact that
no chemical reaction takes place in the system. Input and output streams flowing
into the sub-system will thus be in exactly the same chemical form, but perhaps in
a different phase or separated from other substances. All separation processes are
examples of non-reaction systems.
Process calculations based on mass balances for non-reaction system is usually
relatively simple. Chapters 2 and 3 deal with mass balances for non-reaction systems.
The biological purification step is an example of a reaction system. This is a
typical topic for chemical engineering, which deals with how to design and operate
reaction systems where a raw material is transformed into a chemical product of any
kind. In our example, it is microorganisms using C, N and P in wastewater for their
biochemical metabolism.
In chemical plants, it is obvious that one knows the exact chemical composition
of both the reactant and products. In ecosystems, however, it is difficult to determine exactly which chemicals react and what products that are formed. One the
contrary, there are many examples of natural reaction systems that are very difficult
to characterize, such as the atmosphere.

1.1.3

Element, components and inert substances

In the following chapters, the concepts of elements, components and inert substances
are used frequently. A characteristics of elements is that they are neither consumed
nor produced in a physical or chemical process. This means that elements are conservative within a system. If the system is at steady state (i.e., nothing in the system
changes with time) one can always assume that the mass flow of an element that goes
into a system, both non-reaction and reaction systems, is equal to the mass flow out
of the system.
A component is a chemical compound of any kind, for example ions such as NO
3
and SO2
4 , or compounds such as CO2 or benzene. In a steady-state, non-reaction
system the mass flow of a component in is always equal to the mass flow out. In a
steady-state, reaction system, however, this precondition never applies. This makes
process balance calculations for reaction systems more difficult than those for nonreaction system.
An inert substance is one that does not react, even if it is present in a system where
other substances react. In many systems, water is an inert substance, especially if it
only serves as a solvent. In other cases, one can comfortably make the simplification
that a substance is inert. This applies, for example, if it only reacts to a small
extent, or is present in such large surpluses that the change in the total amount of
the substance in the system can be neglected.

1.2 The mass balance principle

13

1.8

1.6
stationr fas

hr intrffar en strning

ngonting vad-som-helst

1.4

1.2

icke-statior, dynamisk fas

0.8

0.6
ny stationr fas

0.4

5
tiden

10

Figur 1.3: Illustration of steady-state and non-steady-state conditions.

1.1.4

Static and dynamic systems

A fundamental characteristic of every system is if varies over time or if the conditions


in the system are time constant. A time constant system is static or at steady state.
It is relatively easy to make calculations on steady-state systems as they are normally
described by algebraic equations.
One must also distinguish between steady state and systems at equilibrium. In
a system at equilibrium, the various states (pressure, concentration etc.) are at the
levels dictated by the chemical equilibrium constants. Thus, a system may be at
steady state but not at equilibrium!
Time-varying systems are called dynamic systems and are at non-steady state.
Figure 1.3 illustrates how the dynamics can arise in a system. If one, for example,
changes the content of a reactant in the flow of a process, it always leads to changes
in the levels of reactants and/or products in the effluent. All of the changes will
not happen immediately: it will always take a while until the system reaches a new
steady-state. Often, the path to steady state is very difficult to predict and understand
due to the dynamics of various processes, since many chemical processes can affect
each other. However, it is an important engineering capability to understand how
to describe and model such dynamic systems and how to develop practical skills to
perform such calculations.

1.2

The mass balance principle

This entire compendium is based on the application of one single equation: the mass
balance. The mass balance equation is written as:
Input + Prod = Output + Acc

1.2 The mass balance principle

14

The term Input refers to the flow of the substance into the system. The flow
can have the unit kg s1 or similar. The term Output concerns, in the same way,
the flow of the substance out of the system.
The term Prod refers to the amounts of the substance that is produced inside
the system. It must however be noted that Prod can be either positive or negative.
If Prod is negative, it means that the substance is consumed inside the system. In
the chemical context, it is obvious that reactants are consumed (Prod < 0), while the
desired products are produced (Prod > 0).
The last term, Acc is conceptually the most difficult. Acc stands for the amount
of a substance accumulated per unit time inside the system. That means that in a
steady-state system Acc = 0 by definition. In a non-steady state system, Acc > 0 if
the amount of the substance increases in the system. If it decreases, Acc < 0. As we
shall see later, the term Acc introduces a time derivative in system models. That is
why the mass balance model for non-steady-state systems always is composed of one
or more difference or differential equations.

Example 11: Balance calculation for a bank account


Problem 1
Each year, you deposit 50 000 SEK in your bank account, while you withdraw a total
of 45 000 SEK. The bank pays you 2 000 SEK in interest, but draws 500 SEK annually
in incomprehensible fees. How much will your bank balance change during the year?
Solution 1
We apply the equation
Input + Prod = Output + Acc
Since the question dealt with the change during a time-interval, we are seeking the
Acc-term. The equation i re-written as:
Acc = Input Output + Prod

With the information given the solution becomes:


Acc = 50000

SEK
SEK
SEK
SEK
45000
+ (2000 500)
= 6500
year
year
year
year

Problem 2
Given the information about Input and Output, how large must the annual interest
(i.e., the Prod term) be if the account balance should increase by 5 600 SEK a year?
Solution 2
We seek the Prod term and rewrite the equation as:
Prod = Output Input + Acc

and calculate the interest as:


SEK
SEK
SEK
SEK
Prod = 45000
50000
+ 5600
= 600
year
year
year
year

1.3 Process calculations and reactor calculations

1.3

15

Process calculations and reactor calculations

In chemical engineering, one distinguishes between Process calculations and Reactor


calculations. Both largely rely on the use of mass balance equations. However, process
calculations are used to establish an overview of a system, often without taking into
account the kinetic laws that determine consumption and production of substances
in the process.
Process calculations may, for example, allow calculation of all mass fluxes (i.e.,
expressed in kg or moles) that cross the boundaries of a system. This is typically
done in terms of Inputs and Outputs. However, one can also use process calculations
for design of chemical plants, especially on a a level of low detail. In this compendium,
however, process calculations are mainly used to describe system, much as if they were
black boxes.
Reactor calculations, however, involves the creation of models capable of answering
why a chemical reaction occurs to a certain extent, or why the the system changes with
time. Reactor calculations builds on two types of equstions: mass balance equations
and kinetic equations. Reactor calculations can thus be used for the design of chemical
reactors but also to gain a deeper understanding of what happens within a reaction
system. They are thus commonly used to create mathematical models that allow us
to understand black boxes.

1.4

Numerical methods

1.4.1

Systems of linear algebraic equations

Process calculations and reactor calculations often give rise to models in terms of
systems of equations. Many of the equation systems created have analytical solutions.
In some cases, it is fairly easy to solve the unknowns from the equation system. In
other cases, it is difficult to solve equation systems if you do not accidentally start
in the right end.
There are various computer programs that can be used to solve systems of equations. Maple does it analytically, while Matlab is used for numerical calculations. In
the examples that follow Matlab is used.
If we want to solve a system of linear algebraic equations we can use the standard
functions of Matlab. In principle Matlab solves the equation system:
AX = Y
by calculating
X = A1 Y
However, the inverse of a matrix can only be calculated if all the rows of the (square)
matrix are linearly independent. Linear independence means that no rows can be
created by additions and/or subtractions involving the other rows. For example, of
the two matrixes:
!
"
!
"
1 1
1 1
A1 =
, A2 =
2 3
2 2

1.4 Numerical methods

16

only A1 can be inverted, while A2 cannot. Furthermore, linear algebra has taught
us that the dimension of the largest (square) sub-matrix that is possible to invert is
called the rank of the original matrix. Hence, the rank of A1 =2, while the rank of
A2 =1.

Example 12: Solving linear equation systems


Problem
Solve the following system of equations with Matlab:
3x1 + 4x2

= 10

x1 x2

=2

Solution
The system of equations can be written as a matrix in the form:
AX = Y
where:
A=

!
3
1

"
4
,
1

X=

! "
x1
,
x2

Y =

! "
10
2

With Matlab the system of equations is solved as:

1.4.2

Numerical solution of systems of algebraic equations

Matlab can also solve the system of equations AX = Y by finding the roots to the
equation 0 = AX Y . In contrast to the above method, this calls for some more
work by the computer. However, in this case one is not limited to linear systems of
equations, non-linear systems can also be solved easily. What Matlab actually does
is to seek F as follows:
F = AX Y, so that F < *
where * is so small that it, in practice, equals 0. The only thing one has remember
as a user is to write the equations in a correct form and then submit a guess of X is
that is not completely out of line.

1.4 Numerical methods

17

Example 13: Solving linear equations with fsolve


Problem
Solve the above systems of equations in 12 using fsolve in Matlab after having
defined the equations in matrix notation.
Solution
The system of equations can be written as a matrix on the form:
0 = AX Y
where
!
3
1

"
! "
4
10
, Y =
1
2

Write an m-file: where f0 is a guess at the solution needed to initialize fsolve. The
Matlab function fsolve is run from workspace with: Of course, the solution is the
same as in the above example; x1 = 2.5714, x2 = 0.5714.

Example 14: Solving non-linear equations with fsolve


Problem
Solve the following system of equations with the routine fsolve in Matlab:
3x1 + 4x22
x31

x2

= 10
=2

Solution
The system of equations can be written as a matrix in the form:
0
0

= 3x1 + 4x22 10
= x31 x2 2

The two functions are the defined in an m-file exfsolve1.m: which is executed
from workspace with: Solution: x1 = 1.4708, x2 = 1.1819. The vector [1 ; 1], above
called f0, is an initial guess of what the solution vector X is.

1.4.3

Solution of differential equations

Dynamics in all types of systems is described by differential equations, ordinary or


partial. It is only for models of rather idealized systems that the resulting differential
equations are analytically soluble. For many classes of problems it is possible to linearize the equations, and thereby obtain analytical mathematical solution. However,
this is a trick, which is most often used in Automatic Control and it will not be
further discussed here.

1.4 Numerical methods

18

In making the mass balance calculations for non-steady state systems, the starting
point is normally one or more ordinary non-linear differential equations. In this
compendium we will almost exclusively solve differential equations numerically. The
exceptions are the simplest cases.
How do we solve a differential equation numerically? First we will take a quick
look at the simplest (and worst) method, Eulers solution method.
An ordinary differential equation can be written in the form:
dX
= f (t, X)
dt
This differential equation states that if X = Xt at time t, then X = Xt+dt at time
t + dt:
Xt+dt = Xt + dtf (t, Xt )
We can interpret this graphically if we follow the derivative of the function X(t)
a rectilinear piece dt, starting at the point (t, Xt ). The approximately estimated
subsequent value of the function is thus a small distance away, in the direction of the
derivative. This value is defined as (t + dt, Xt+dt ). After this value of the function has
been calculated, another value is calculated in the same way. The whole procedure is
repeated to create a vector of estimates function value.
As we see in Figure 1.4 the approximation may differ significantly from the real
function X. With Eulers method, the calculation error is relatively large, but if one
reduces the dt, the calculation error is also reduced. However, the smaller the dt,
the longer the calculations take. For all practical purposes it is better to use more
sophisticated solution methods such as the ode45 routine in Matlab.

1.4 Numerical methods

19

5
X

t+dt

approximering av Xt+dt

dX/dt i (t,Xt)

Xt

0
t
-1

1.2

1.4

1.6

t+dt
1.8

2.2

2.4

2.6

2.8

Figur 1.4: Illustration of Eulers method.

Example 15: Numerical integration with Eulers method


Problem
Integrate (simulate) the ordinary non-linear differential equation:
#
t=0
dc
= 0.5c
Initial value:
dt
c=1
using Eulers method. The simulations shall be done over the range t = 0 to t = 4 in
steps (dt) of 0.1 units of time.

1.4 Numerical methods

20

Solution
Write an m-file exEuler.m: which is executed from workspace with exEuler. In
the m-file, one can see how the time interval (from 0 to 4) has been divided into 40
pieces, each representing 0.1 units of time. The results are given below, together with
the exact solution. The conclusion is: As long as the solution is monotonous, this
simple integration method works quite well!
2

1.8

1.6

1.4

1.2

0.8

0.6

exakt
euler

0.4

0.2

0.5

1.5

2.5

3.5

1.4 Numerical methods

21

Example 16: Numerical integration with ode45


Problem
Integrate (simulate) the ordinary non-linear differential equation with ode45 in Matlab:
dX1
dt
dX2
dt

= k1 X1

k1 = 0.5

= k2 X1 X2

k2 = 0.1

Solution
Write an m-file exODE45.m: and run it from workspace with: where [0 10] denotes
the time interval that should be covered, and the vector [1 1] represents the initial
values for r X1 and X2 respectively. The solution is presented as a diagram:
1

0.9
X2

Vrdet p tidsvariabelt tillstnd X

0.8

0.7

0.6
X1

0.5

0.4

0.3

0.2

0.1

5
tiden t

10

CHAPTER

Process calculations for


non-reaction systems
In this chapter we shall go through the methods for resolving integral mass balances
for the very simplest type of system. As mentioned in the previous chapter, process
calculations allow us to describe systems with flows of various chemical substances.
The process calculations my be used to quantify fluxes, but may also be extended to
process design and, partly, to explain what is happening within a process system.
Process calculations have played an enormously important role in industry, but
also in environmental research. Without mass balances, we would not have had
any quantitative understanding of flows of mass and matter in any one of earths
ecosystems.
Mass balances are known under different names. Often one uses the terms budget
calculations or material balances.
Here we will use a somewhat formal approach to process calculations. They will
involve models with several equations. Solving these equations involves the collection
of as many equations as that there are unknown variables. By working with degree of
freedom analyses as a tool we will systematically examine wether it is possible or not
to solve a certain system of equations. We will also see how they can become solved
by doing certain tricks.
In a series of very simple examples, different types of equations and information
will be slowly introduced and will illustrate key concepts in the mass balance calculations.
The methodology will probably seem pretty tedious, and give the impression of
being overly formal. If so, the impression is correct: the idea is to show a formal
and systematic approach that always leads to a solution if a solution exists, and that
reveals why a solutions sometimes cannot be provided.
The mass balance for element/component A will be denoted as MB A etc.

22

2.1 Component mass balances

23

systemgrns
1

W1,A W1,B

W3,B

Process
2

W2,A

W4,A W4,B

Figur 2.1: Illustration of the concepts of process, system and stream variable.

2.1

Component mass balances

Let us consider a process where substances flow into the system at one end, and out
of the system at the other end. The mass flux of the substance is denoted as Wi,j ,
where i refers to stream i and j to the substances. For each substance, we can always
write the general mass balance:
Input + Prod = Output + Acc
If the system is a non-reaction system and is at steady state, we recognize that
Prod = Acc = 0 and we get:
Input = Output
or, expressed in terms of mass fluxes:
$
$
Wi,j =
Wi,j
in

out

Figure 2.1 shows a process where two physical streams, W1 and W2 are input
fluxes, while W3 and W4 are output fluxes. W1 and W4 both contain the chemical
components A and B, while W2 only contains A and W3 only contains B.
The units for mass flux W may be mass (kg) or mass flux (kg/time). Under these
conditions, the integral component mass balances for A and B become:
MB A : W1,A + W2,A

= W4,A

MB B :

= W3,B + W4,B

W1,B

We can also see that these component mass balances show linear independence (i.e.,
they are not identical). If they were independent, one of the balances would be
worthless. The solution would be trivial, of the type 0=0, or 1=1.
The variables W1,A , etc., are called stream variables. A stream variable refers to
the flux of a specific substance in a specific physical stream.
The maximum number of stream variables in a system always equals the number
of streams multiplied by the number of components. If there are 4 streams with 2
components in each, there is a maximum of 8 stream variables. This means that we
need 8 equations in order to allow the system to be unambiguously defined. This is
also true in the example above, although only 6 stream variables are explicitly written
out. The reason is that implicitly W2,B = W3,A = 0.

2.2 Degrees of freedom analysis

2.2

24

Degrees of freedom analysis

The difference between the number of stream variables and the number of equations
available is called the number of degrees of freedom. We can now distinguish three
fundamentally different cases:
Degrees of freedom
Degrees of freedom
Degrees of freedom

>0
=0
<0

The system is not completely defined.


The system is defined.
The system definition contains to much information.

Thus, first step when chemical systems are analyzed is to figure out the number of
degrees of freedom.
So, how is the the number of degrees of freedom determined? One can use the
following :
-

Number of (unknown) stream variables


number of independent component mass balances
number of available independent information

number of degrees of freedom

So, what is independent information? In fact, it is a piece of quantitative information that can be expressed as an equation! For example, in the following, we make
us of independent information given in terms of:
1. known fluxes
2. known concentrations
3. total constraints
4. basis of calculations
With known fluxes it is meant that the mass flux of a component in a stream, Wi,j or
the molar flux, Fi,j is given. As mentioned above, the flux may either be given as an
amount (kg, mole) or a flux which is amount per unit of time (kg/time, mol/time).
In the real world, it is quite difficult to make direct measurements of component
fluxes. Instead, one measures the concentration of a component in a stream and
multiplies it by the total flux in the same stream. This will give Wi,j and/or Fi,j .
A known concentrations might mean the actual concentration ci,j (in kg/volume
or mole/volume) but just as often the mass fraction i,j , or the molar fraction xi,j .
The total constraint is often very useful. This is in spite of the fact that the total
constraint merely states that the whole must equal the sum of the parts. Mathematically, the total constraint implies that the sum of the mass fractions and/or the
molar fractions in a stream equals 1. For example, for a stream i that contains n
components j we can state that:
n
$
j=1

i,j = 1 ,

n
$

xi,j = 1

j=1

One can also define corresponding total constraints for mass fluxes and molar fluxes.
For a given physical stream Wi , that contains n components j, we can state that the

2.2 Degrees of freedom analysis

25

systemgrns
1

1,A W1
1,B W1

2,A W2

3,A W3
Process

4,A W4
4,B W4

Figur 2.2: Steam variables expressed in terms of mass fractions i,j and the total
flux Wi .

component fluxes (quantified by the stream variables) must add up to the total flux.
Hence:
)
n

j=1 Wi,j
Wi =

W )n
i
j=1 i,j
The corresponding relationships also hold true when the stream variables are expressed in molar units:

Fi =

)
n

j=1 Fi,j

F )n x
i
j=1 i,j

The equations are illustrated in Figure 2.2 which shows how the stream variables
in Figure 2.1 may be expressed in terms of the mass fraction i,j and the total flux
Wi .
A basis of calculation consists of an arbitrary definition of a flux. This number is
used as an equation in the calculations. A basis of calculation may be (and has to
be!) introduced if no other values of fluxes (i.e. no values of stream variables) are
known or given. This occurs when only concentrations or mass/molar fractions are
specified. For example, a basis of calculation can be set to 1 kg, 100 kg/hour, etc.,
but may only be introduced of it does not contradict other information.

2.3 Process calculations for a separation process

2.3

26

Process calculations for a separation process

In this section we shall see how we methodically can solve a process calculation problem for a non-reaction system at steady state. The starting point is a very simple
separation process where a mixture of ethanol, C2 H5 OH and H2 O are separated into
two streams with another composition than the input stream. By varying the way
we look at the problem, we will see how different types of information can be used
for problem solution. All examples are solved with Matlab, but it is valuable to solve
equation systems by hand now and then, for sake of training.
The solution methodology is:
Draw a picture, a process chart with the system boundary and all the physical
streams.
Enter all stream variables and information in the process chart.
Make a degree of freedom analysis.
Set up and solve the equation system.
Answer the question(s) asked.
The first examples dealing with separation of the two components is schematically
illustrated below. In the process, C2 H5 OH and H2 O flow into a process with one
input stream and leave the system in the other end via two output streams. The
mass flux of each component is denoted as Wi,j , where i denotes stream i, while j = 1
denotes C2 H5 OH and j = 2 denotes H2 O:
W2,1 kg C2H5OH

W2,2 kg H2O

W1,1 kg C2H5OH
W1,2 kg H2O

separationsprocess
W3,1 kg C2H5OH

W3,2 kg H2O

Example 21: Separation of ethanol and water


Problem:
The input stream (the feed) to a separation process contains 500 kg C2 H5 OH and
500 kg H2 O. There are two output streams from the process, one containing 460 kg
C2 H5 OH and 60 kg H2 O. How much C2 H5 OH and H2 O leave the system in the other
stream?
Solution:
The first thing to do is to draw a process chart, where the streams are numbered
1-3 and the components 1-2. The stream variables and the information given are all
introduced into the process chart:

2.3 Process calculations for a separation process

460 kg C2H5OH

27

2
60 kg H2O

500 kg C2H5OH
500 kg H2O
W3,1 kg C2H5OH

W3,2 kg H2O

The next step is to carry out a degree of freedom analysis. The way the process chart
is drawn, there are only two unknown stream variables, W3,1 and W3,2 . Now, for each
component, it is possible to make a separate mass balance, called MB A and MB B.
Thus, the degree of freedom analysis becomes:
Unknown stream variables
Independent component mass balances
Degrees of freedom

2
-2
0

Since the degrees of freedom = 0, the system is defined in an unambiguous manner!


It is thus adequate to define the equstions of the system forming the mass balance
equations in the form Input=Output:
MB C2 H5 OH :

500

= 460 + W3,1

MB H2 O :

500

= 60 + W3,2

or
MB C2 H5 OH :
MB H2 O :

1 W3,1 + 0 W3,2
0 W3,1 + 1 W3,2

= 500 460
= 500 60

In matrix notation AX = Y this becomes:


!
"
!
"
!
"
1 0
W3,1
500 460
A=
, X=
, Y =
500 60
0 1
W3,2
We solve this quite trivial system with Matlab as: Hence, the solution to the problem
is: W3,1 = 40 kg and W3,2 = 440 kg.

2.3 Process calculations for a separation process

28

Example 22: Separation of ethanol and water - 6 unknowns


Problem:
Solve the problem above while considering every stream variable as an unknown, and
all of the information given as indepenent information.
Solution:
First, the process chart is drawn with the streams numbered 1-3 and the components
1-2.
W2,1 kg C2H5OH
2
W2,2 kg H2O

W1,1 kg C2H5OH
W1,2 kg H2O

separationsprocess
W3,1 kg C2H5OH

W3,2 kg H2O

There are six unknown stream variable, and the independent information are W1,1 =
500, W1,2 = 500, W2,1 = 460 and W2,2 = 60.
Unknown stream variables
Independent component mass balances
Independent information

6
-2
-4

Degrees of freedom

The equations are:


MB C2 H5 OH :

500

= 460 + W3,1

MB H2 O :

500

= 60 + W3,2

Info 1 :

500

= W1,1

Info 2 :

500

= W1,2

Info 3 :

460

= W2,1

Info 4 :

60

= W2,2

Or, with matrix notation AX = Y :

1 0 0 0 0 0
W3,1
0 1 0 0 0 0
W3,2

0 0 1 0 0 0
W1,1

A=
, X = W1,2 ,
0 0 0 1 0 0

0 0 0 0 1 0
W2,1
0 0 0 0 0 1
W2,2

500 460
500 60

500

Y =

500
460
60

A is a unity matrix, which can be created with the Matlab command eye(6):

2.3 Process calculations for a separation process

29

Example 23: Ethanol and water concentration constraints


Problem:
A separation process is fed with an input stream consisting of 500 kg C2 H5 OH and
500 kg H2 O. The output consists of two streams. One amounts to a total of 400
kg and contains 96% C2 H5 OH. Calculate the mass fractions of C2 H5 OH and H2 O
respectively in the other stream.
Solution:
First, the process chart is drawn with the streams numbered 1-3 and the components
1-2.
96%, W2,1 C2H5OH
W2,2 H2O

W2=400kg

500 kg C2H5OH
500 kg H2O
W3,1 kg C2H5OH

W3,2 kg H2O

The next step is to carry out a degree of freedom analysis. They way the process chart
is drawn, there are four unknowns, W2,1 and W2,2 as well as W3,1 and W3,2 . However,
we have two information: the concentration constraint 0.96 = W2,1 /(W2,1 + W2,2 )
and the total constraint 400 = W2,1 + W2,2 .
Unknown stream variables
Independent component mass balances
Independent information

4
-2
-2

Degrees of freedom

The four equations are:


MB C2 H5 OH :

500

= W2,1 + W3,1

MB H2 O :

500

= W2,2 + W3,2

Info 1 :

400

= W2,1 + W2,2

Info 2 :

0.96

= W2,1 /(W2,1 + W2,2 )

0
In matrix notation,

1
0
A=
1
0.04

= 0.04W2,1 + 0.96W2,2

AX = Y we have:

0
1 0
W2,1

1
0 1
, X = W2,2 ,
W3,1
1
0 0
0.96 0 0
W3,2


500
500

Y =
400
0

2.3 Process calculations for a separation process

30

and the solution becomes: The mass fractions in stream 3 are thus 3,1 = 116/(116+
484) = 0.1933, and 3,2 = 484/(116 + 484) = 0.8067.

Example 24: Separation of ethanol and water moved constraint


Problem:
A separation process is fed with an input stream consisting of 500 kg C2 H5 OH and
500 kg H2 O. The output consists of two streams. The first of these output streams
contains 96% C2 H5 OH, while the second amounts to a total of 600 kg. Calculate the
mass fractions of C2 H5 OH and H2 O respectively in the second stream.
Solution:
First, the process chart is drawn with the streams numbered 1-3 and the components
1-2.
96%, W2,1 C2H5OH
2
W2,2 H2O

500 kg C2H5OH
500 kg H2O
W3,1 kg C2H5OH
W3,2 kg H2O

W3=600kg

Just as in the previous example, there are four unknown stream variables. In addition,
the concentration constraint 0.96 = W2,1 /(W2,1 + W2,2 ) is the same as above, while
the total constraint is moved so that 600 = W3,1 + W3,2 .
Unknown stream variables
Independent component mass balances
Independent information

4
-2
-2

Degrees of freedom

The four equations are:


MB C2 H5 OH :

500

= W2,1 + W3,1

MB H2 O :

500

= W2,2 + W3,2

Info 1 :

600

= W3,1 + W3,2

Info 2 :

0.96

= W2,1 /(W2,1 + W2,2 )

0
In matrix notation,

1
0
A=
0
0.04

= 0.04W2,1 + 0.96W2,2

AX = Y we have:

0
1 0
W2,1

1
0 1
, X = W2,2 ,

0
1 1
W3,1
0.96 0 0
W3,2


500
500

Y =
600
0

2.3 Process calculations for a separation process

31

and the solution is: The mass fractions of all constituents are, of course, the same as in
the previous example: 3,1 = 116/(116+484) = 0.1933, and 3,2 = 484/(116+484) =
0.8067.

CHAPTER

Process calculations with


multiple units
In practice, it is rare for natural or engineered systems to be described with only
one unit. As examples of this we have . Instead, it is normally several streams and
units with different characteristics, which together form a complex process, such as
the treatment plant in Figure 1.1.
In this chapter, we will see how to manage process calculations for non-reaction
system with multiple units. An important part is to implement the degree of freedom
analysis of multiple-unit systems. The chapter also introduces typical units that are
important to recognize, especially the mixer and the splitter. Some fairly comprehensive examples, such as 32, are included.

3.1

Calculation methodology

One important aspect that distinguishes multiple unit systems from those discussed
in the previous chapter is that the system boundaries can be drawn in several ways.
For example, consider the system:

Here, we can distinguish three levels:


1. Degree of freedom analysis and equation system for a specific unit.
2. Degree of freedom analysis and equation system relating to the outer system
boundaries, from now on referred to as the Process.

32

3.2 Common types of units

33

3. Degree of freedom analysis and equation system relating to all units and the
streams that connect them, from now on referred to as the Total.
What one often finds is that the degrees of freedom are > 0 for the process and
for any of the individual units. For other entities the number of degrees of freedom
= 0. If the problem is solvable, the number of degrees of freedom of the Total is 0.
In case any of teh individual units have degrees of freedom < 0, the system is defined
in a unique manner, and all stream variables can not be calculated.
If you solve the equation system by hand, you have to begin with a unit for
which the number of degrees of freedom is 0. The information thereby obtained is
used stepwise as independent information for other, non-solvable units for which the
number of degrees of freedom > 0. However, if a numerical equation solver is used,
the easiest way to handle to problem is by solving all equations simultaneously.

3.2
3.2.1

Common types of units


Mixer

In a mixer, one or more streams are merged:

Mixer

For a mixer, the total output flux must equal the sum of the input fluxes. This holds
regardless of whether the fluxes are given in mass units or in molar units.

3.2.2

Splitter

A splitter divides one input stream into several output streams.

Splitter

For a splitter, it is evident that the sum of all output component fluxes must
be equal to the input flux. But in addition, the composition, as given by mass
fractions or molar fractions, identical in all output streams. This may be utilized in
the calculations by means of splitter constraints.
A splitter constraint states that the concentration of a component in any of the
output streams is equal to the concentration in the input stream. Thus, if an input
stream W1 with two components A and B is split into 3 output streams, W2 W4

3.2 Common types of units

34

one splitter constraint is that:


W1,A
W2,A
=
W1,A + W1,B
W2,A + W2,B
If one composition in W1 is known, one can use this information to re-write the above
expression as:
1,A =

W2,A
W2,A + W2,B

However, it is not possible to define any number of splitter constraints. Instead,


the number of independent splitter constraints is limited. This fact origins from the
nature of the splitter: All stream do have the same composition! If too many splitter
constraints are added to each other, the result will be the component mass balance.
Thus, one can only define a limited number of independent splitter constraints.
If the input stream to a splitter is divided into S output streams, and there
are N components in the input, one can define (N 1)(S 1) independent splitter
constraints. This holds true if the components mass balances are used in the process
calculation.
For example, if a stream with two components is split into two streams, we may
form exactly (2-1)(2-1)=1 independent splitter constraint. However, for a splitter
receiving four components, split into four output streams as many as (4-1)(4-1)=9
splitter constraints can be included in the equation system.
The splitter constraints are not only a whole lot of fun; they give rise to non-linear
algebraic equations that cannot be solved by simple numerical methods. Therefore,
it is advised to make a short-cut to overcome this obstacle. If one knows that:
1,A =

W2,A
W2,A + W2,B

where 1,A = 0.2

one may directly write the linear equation:


0.2 =

W2,A
W2,A + W2,B

or 0 = 0.8W2,A 0.2W2,B

3.2 Common types of units

35

Example 31: Limitation of the number of splitter constraints


Problem:
Show that only 1 independent splitter constraint exists if one at the same time sets
up the component mass balances of A and B.
W2,A
2
W2,B

W1,A
1

Splitter
W1,B

W3,A

W3,B

Solution:
We try to formulate 4 equations:
MB A

W1,A = W2,A + W3,A

MB B

W1,B = W2,B + W3,B

Splitter constraint 1

W1,A
W1,A +W1,B

W2,A
W2,A +W2,B

Splitter constraint 2

W3,A
W3,A +W3,B

W2,A
W2,A +W2,B

Re-write the mass balances:


MB A

W3,A

MB B

W3,B

= W2,A W1,A

= W2,B W1,B

and substitute W3,A and W3,B into splitter constraint 2, which is then simplifies as:
W3,A
W3,A +W3,B

W2,A
W2,A +W2,B

(W1,A W2,A )(W2,A W2,B ) = W2,A (W1,A W2,A + W1,B W2,B )


2
(W1,A W2,A W2,A
+ W1,A W2,B + W2,A W2,B ) =

2
W2,A W1,A W2,A
+ W2,A W1,B W2,A W2,B

W1,A W2,B = W2,A W1,B


Now, we re-write and simplify splitter constraint 1:
Splitter constraint 1 W1,A (W2,A + W2,B ) = W2,A (W1,A + W1,B )
W1,A W2,B = W2,A W1,B
Splitter constraint 1 is identical to a combination of splitter constraint 2 and the two
component mass balances. Hence, only 3 out of 4 equations are independent!

3.2 Common types of units

3.2.3

36

Recirculation

A recirculation stream returns material from a downstream part of a process to a


mixer upstream point. Through recirculation, a portion of the material that has
passed through a specific process step will pass through it once more:

Recirkulationsstrm

One quantifies the recirculation by a quantity called the recirculation ratio, defined
as the ratio between the recirculated stream and the feed. If the feed is larger than
the recirculation stream the recirculation ratio <1.
Recirculation is very common in complex processes. For example, the water treatment plant illustrated in Figure 1.1 includes a recirculation stream from the nitrification to denitrification stage in the process. Another example is municipal incineration
plants, where particles are recycled from the gas cleaning units to the furnace.

3.2.4

Bypass

Bypass is simply a unit that allows a stream to bypass a process unit:

Frbiledning

However, a bypass stream is very easy to ignore by mistake. In fact, bypass always
introduces two additional units: a splitter and a mixer. This must be accounted for
when making a complete degree of freedom analysis. In industry, bypass is used as a
tool to manage limited capacity in separation process units.

3.2.5

Purge

Purge means that a fraction of a recycle stream is led off:

Recirkulationsstrm

Avtappning

Purging can be seen as a combination of a process unit, at least one splitter


and a mixer. Purging is used to avoid the accumulation of inert components in
process systems. For example, if nitrogen is present in a feed and the system involves
recirculation, nitrogen would build up unless purging is included.

3.2 Common types of units

37

Example 32: Process calculation based on total system analysis


Problem:
From a manufacturing process, a sewage output stream contains 10 weight-% of the
environmentally harmful component A. The substance A exists in a liquid stream
which mainly contains water, here referred to as B.
A government requirement exists that says the concentration of A may not be
more than 3 % in the outgoing stream to the recipient. To meet this requirement a
separation plant in which A is separated to 90% has been introduced.
The separation requires expensive additional chemicals. Therefore, it is economically favourable to clean as small a part of the sewage flow as possible. Thus, a sewage
flow bypass is introduced.
Calculate all the flows in the system, and determine how large should the part of
the sewage flow stream that is led through the process should be as compared to the
fraction that is bypassed? Base the calculation on simultaneous solution of the total
system.
The process can be described by the following chart:

Process chart and stream variables:


First we draw a process chart where the streams have been numbered 16 and the
stream variables necessary are introduced.

I fact, the problem only states that the ratio between stream 2 and 3 needs to be
calculated. But to do this, all other stream variables have to be calculated.
Degree of freedom analysis:
First, a degree of freedom analysis is carried out. For an analysis based on the total
system, including all sub-systems, the system boundaries are drawn as follows:

3.2 Common types of units

38

We can now make the degree of freedom analysis to investigate if the problem is
possible to solve. Calling the basis of calculation BoC we get:
Stream variables
MB (A and B, Splitter)
MB (A and B, Separation)
MB (A and B, Mixer)
Info conc. of A in W1
Info conc. of A in W5
Info separation of A
Splitter constraint
Basis of calc.
Degrees of freedom

Totalt
11
-2
-2
-2
-1
-1
-1
-1
-1
0

Equations:
Here, we will utilize the fact that the system can be solved by solving for all the 11
equations at once.
MB A Splitter:

W1,A = W2,A + W3,A

MB B Splitter:

W1,B = W2,B + W3,B

MB A Separation:

W2,A = W4,A + W6,A

MB B Separation:

W2,B = W4,B

MB A Mixer:

W3,A + W4,A = W5,A

MB B Mixer:

W3,B + W4,B = W5,B

Conc. of A in stream 1:

W1,A = 0.1(W1,A + W1,B )

Conc. of A in stream 5:

W5,A = 0.03(W5,A + W5,B )

Separation of A:
Splitter constraint (trick):
Basis of Calc.:

W4,A = 0.1W2,A
W3,A = 0.1(W3,A + W3,B )
W1,A + W1,B = 1000

3.2 Common types of units

39

Then, all equations are re-written so that all the stream variables end up on the
left side:
W1,A W2,A W3,A

=0

W2,A W4,A W6,A

=0

W3,A + W4,A W5,A


W3,B + W4,B W5,B

=0
=0

Conc. of A in stream 5: 0.97W5,A 0.03W5,B


Separation of A:
W4,A 0.1W2,A

=0
=0

MB A Splitter:
MB B Splitter:
MB A Separation:
MB B Separation:
MB A Mixer:
MB B Mixer:
Conc. of A in stream 1:

Splitter constraint (trick):


Basis of Calc.:

W1,B W2,B W3,B

=0

W2,B W4,B

=0

0.9W1,A 0.1(W1,B

=0

0.9W3,A 0.1(W3,B

=0

W1,A + W1,B

= 1000

This system of equations can also be written in matrix form AX = Y where A is the
matrix of coefficients, X is a column vector containing all the stream variables, while
Y is the right hand side of the equations, only containing numbers. We then get the
following matrixes:

W1,A
1
0
1
0
1 0
0
0
0
0
0
W1,B
0
1
0
1
0 1 0
0
0
0
0

W2,A

0
0
1
0
0
0 1 0
0
0 1

W2,B
0
0
0
1
0
0
0 1
0
0
0

W3,A
0
0
0
0
1
0
1
0
1
0
0

0
0
0
0
1
0
1
0
1 0
A=
, X = W3,B ,
0
W4,A

0.9 0.1 0
0
0
0
0
0
0
0
0

W4,B
0
0
0
0
0
0
0
0 0.97 0.03 0

W5,A
0
0
1
0
0
0
1
0
0
0
0

W5,B
0
0
0.9 0.1 0
0
0
0
0
0
0
W6,A
1
1
0
0
0
0
0
0
0
0
0

It is easy to solve this system of equations with Matlab. The Matlabsolution m-file
is: and the final results are: We now recall that the initial aim was to calculate
the fraction of the stream W1 that could bypass the separation process while still
preventing the concentration of A in the output stream W5 from getting too high.
The result is that the ratio between the steam can be:
W3,A + W3,B
20.6979 + 186.2807
=
= 0.2610
W2,A + W2,B
79.3021 + 713.7193

0
0

Y =
0
0

0
1000

3.2 Common types of units

40

Example 33: Degree of freedom analysis based on sub-systems


Problem:
Make a degree of freedom analysis for Example 32 for the Process and the three
sub-system (units) that form the system.
Solution:
In this case, the degree of freedom analysis is:
Stream variables
MB (A and B, Process)
MB (A and B, Splitter)
MB (A and B, Separation)
MB (A and B, Mixer)
Info conc. of A in W1
Info conc. of A in W5
Info Separation of A
Splitter constraint
Degr. of freed.w/o BoC
Basis of Calc.
Degr. of freed. w BoC

Process
5
-2

-1
-1
1
-1
0

Splitter
6

Separation
5

-2

-2

-1
-1
2
-1
1

-1
2
-1
1

Mixer
6

-2
-1
3
-1
2

Thus, the Process may be solved for, but this does not allow us to calculate all fluxes
in the system. If the Process is solved for, W1,A , W2,A , W2,B , W5,A and W5,B are
quantified. If we should continue to with the other sub-systems, we must take into
consideration that:
We may not introduce an additional basis of calculation, this has already been
done.
The information regarding W1 and W5 have already been used.
The information regarding the separation has already been used.

Since no additional basis of calculation may be introduced at this stage, the degree
of freedom analysis for the remaining units becomes:
Stream variables
MB (A and B, Splitter)
MB (A and B, Separation)
MB (A and B, Mixer)
Known stream variables
Splitter constraint
Degrees of freedom

Splitter
6
-2

Separation
5

-2
-1
1

-1

-2
-2

-2

Mixer
6

Conclusion:
We may now conclude that we cannot get any further. Thus, only the Total and the
Process may be solved for.

CHAPTER

Process calculations for reaction


systems
In many natural and engineered systems chemical reactions take place in which reactants are converted into products. Products and reactants are the usual nomenclature
in chemistry and chemical engineering. Regarding biological systems one says that
a substrate is metabolized to a metabolite. The treatment plant in Figure 1.1 is an
example of a technological system that is dominated by metabolic chemical processes.
It is not as easy to describe a system with chemical reactions as it is to describe
non-reaction systems. One reason is that new components are created in the system;
another reason is that we may not know exactly which components are actually in
the system.
This chapter describes different techniques to make process calculations in the
system with one or more chemical reactions. Some important new concepts are defined
to help when independent information is to be interpreted.

4.1

Mass balances mass or mole?

Suppose that the following chemical reaction occurs in a system:


N2 + 3H2 2NH3
Here, it is important to note that a chemical reaction formula is always written in
terms of molar units. Indeed, we all know that it is not true that 1 kg N2 and 3 kg
H2 form 2 kg NH3 .
We can also note two important things happening for the above reaction system:
1. N2 och H2 are consumed while NH3 is produced.
2. The chemical form of the elements N and H change.
These simple observations have two implications with respect to the mass balances
that can be used to define the system in which the reaction takes place. If one
formulates mass balances for the components N2 , H2 , NH3 they will be in a different

41

4.2 Important concepts

42

form compared to if mass balance equations are set up for the elements N and H:
Element mass balance or N, H

Input = Output

(4.1)

Component mass balance for N2 , H2 , NH3

Input + Prod = Output

(4.2)

The difference, obviously, is caused by the fact that elements are indivisible while
components may react and change their nature in chemical reactions. In terms of
constraints related to stream variables, this can be expressed more formally as:
$
$
$
$
Fi,j
Wi,j and
Fi,j $=
Wi,j $=
For components:
For elements:

in

out

in

out

in

Wi,j =

out

Wi,j and

in

Fi,j =

Fi,j

out

When one solves mass balance problems for reaction systems one should therefore:
1. Never make mass balances in terms of mass units, but always use molar units.
2. Only use component mass balances if there are good reasons, otherwise element
mass balances should be used (still in molar units).
Unfortunately, in most natural systems one can never know exactly which components are present, how they react and what they form. However, by use of chemical
analytical methods it is quite possible to track the elements that are present in various phases, and how they are transported within the system. In conclusion, systems
where biological transformations are important should normally be handled be means
of element mass balances.
Another exciting possibility, not further discussed here, is the use of isotopes.
For example 18 Oand 16 O both have element characteristics, but do in fact behave
somewhat differently in nature.

4.2
4.2.1

Important concepts
Inert

An inert substance is one that does not react chemically. In many processes the
synthesis of NH3 described above is a rare exception N2 is an inert. N2 stays inert
as long as the temperature is below 1200 o C. An inert substance can be considered
as an element, and handled just like any other element.

4.2 Important concepts

4.2.2

43

Conversion

The conversion is a number that represents the extent to which a reactant takes part
in a chemical process. The conversion X for a certain reactant is defined as:

Converted amount

Input flux

Input flux - Output amount

Input flux
X=

output

1 $

input

)
)
Hence, if the output flux ( output F ) is equal to the input flux ( input F ) the conversion is obviously X = 0 (i.e., the reactant
has not reacted at all within the system).
)
On the contrary, if the output flux ( output F ) is equal to 0, all of the reactant has
been converted (i.e., X = 1, or as percentage in the range 0-100%).

4.2.3

Yield

The yield quantifies the amount of a substance formed, in relation to what could have
been formed if the limiting reactant were converted to a desired product. With this
definition, the yield is 100% if:
1. All of the reactant converted is converted to the desired product (i.e. no-side
reactions take place).
2. The conversion of the reactant in question is 100%.
For example, let us define the yield if the limiting reactant R is converted to the
main product P. The stoichiometry of the reaction is:
and the yield is thus defined as:
$
Fi,P /P

R R P P

output

Yield = $

Fi,R /R

input

4.2.4

Selectivity

The selectivity tells the amount of the main product that is formed in relation to the
amount of (undesired) bi-products that are formed. If the molar fluxes of all products
have been determined, the selectivity is calculated as:
$
Fmain product
Selectivity =

output

output

Fbi-products

4.3 The element mass balance method

4.3

44

The element mass balance method

Element mass balances are based on the fact that the molar flux of elements can be
expressed in terms of molar fluxes of components. For example, if:
FH 2 O = 1
then
FH = 2 and FO = 1
This section shows how to set up element mass balances for reaction systems. The
degree of freedom analysis includes the analysis of the atomic matrix that determines
how many independent element mass balances can be set up.
The degree of freedom analysis for element mass balances is:
-

Number of (unknown) stream variables


Number of independent element mass balances
Number of independent information

Number of degrees of freedom

In a first, trivial example, the methodology for process calculations for reaction
system and use of mass balances will be introduced.

Example 41: Catalytic dehydrogenation


Problem
In a process, the following reaction takes place:
catalysis

C2 H6 C2 H4 + H2
systemgrns

C2H6

Process

C2H6

C2H4 H2

Determine all fluxes (i.e., stream variables) in the system if the feed contains 100 mol
C2 H6 and the conversion is 60%.
Solution
If we assume that the element mass balances for C and H are independent, the degree
of freedom analysis becomes:
Stream variables
- Independent element MB
- Independent information
Degrees of freedom

4
-2
-2
0

4.3 The element mass balance method

45

Hence the system is uniquely defined. Using F to denote molar flux, we can set
up mass balances that describe the number of moles of C and H that are associated
with the input and output streams. For example, a flux of 1 mole of C2 H6 obviously
carries 2 moles of C and 6 moles of H. The system of equations then becomes:
MB C :

2F1,C2 H6

= 2F2,C2 H6 + 2F2,C2 H4

MB H :

6F1,C2 H6

= 2F2,H2 + 6F2,C2 H6 + 4F2,C2 H4

Info 1 :

F1,C2 H6

= 100

Info 2 :

F2,C2 H6

= (1 0.6)F1,C2 H6

In matrix notation AX = Y we get:


2
0 2 2
F1,C2 H6
0

F2,H2
0
6
2
6
4


, X =
A=

F2,C2 H6 , Y = 100
1
0
0
0
(1 0.6) 0 1 0
F2,C2 H4
0

and the Matlab solution becomes:

4.3.1

The atomic matrix

In example 41 it was simply assumed that the element mass balances for C and
H were independent. Normally, this assumption is valid but there are reasons to be
cautious.
The number of independent element mass balances is given by the rank of the
atomic matrix. The columns of the atomic matrix contains the stoichiometric composition of all components in the system, while the rows refer to each element. This
means that if any rows are linearly dependent, one of the rows is redundant. The
consequence is that two elements are dependent.
For example, let us consider the process above (example 41), which includes the
components C2 H6 , C2 H4 andH2 :
C 2 H6 H2 C 2 H4
0
1
C
2 0 2
H
6 2 4

We can now calculate the rank of this matrix. If the rank is 2, then the elements
C and H are independent. And if they are independent, it is possible to form two
independent element mass balances, one for C and one for H. In Matlab the rank is
calculated as: As we see, the rank of the atomic matrix is 2 which means that we can
form the corresponding independent element mass balances which both add unique
information.

Example 42: Atomic matrix with linearly dependent rows


Problem
In a process, the conversion:
catalysis

2C2 H4 C4 H8

4.3 The element mass balance method

46

is carried out over a catalytic bed.


systemgrns

C2H4

Process

C4H8
C2H4

Make a degree of freedom analysis provided that the feed contains 100 mol C2 H6 and
the conversion is 60%.
Solution
The atomic matrix is:
C 2 H4 C 4 H8
1
0
2 4
C
4 8
H
Obviously, the rank for the atomic matrix is 1, which means that only one of the
element mass balances for C or H is dependent on the other. Hence, the degree of
freedom analysis is:
Stream variables
- Independent element MB
- Independent information
Degrees of freedom

3
-1
-2
0

We can thus conclude that the reaction may actually be viewed as if:
catalysis

2AA2
where A = C2 H4 . Obviously, the elements C and H will always appear in the same
proportions. They are thus not independent, but dependent on each other.

4.4 The component mass balance method

4.4

47

The component mass balance method

In 4.3 the method to solve all fluxes in a system was based on i) element mass balances
and ii) stream variables for components.
In this section, we will introduce a method to base the calculations on component
mass balances instead. We then have to take into consideration that components
react, while elements do not.
The reason why an alternative to element mass balance calculation can be useful
is that sometimes, one needs to know the extent to which chemical reactions occur
in a system. With the element balances method, the chemical reactions were not
considered at all. With the component mass balance calculations, quantification of
the chemical reactions are essential.
When applying a component mass balance method, the degree of freedom analysis
is somewhat different compared to the element mass balance method. The reason
is that the chemical reactions need to be quantified by some kind of Prod term as
outlined in equation 4.2, which introduces mathematical constraints in the calculation.

4.4.1

The reaction parameter

The reaction parameter is a calculation helper. It is defined as:


= The number of moles converted (per unit time) in a system via a certain reaction.
For the most simple example, where A B and there is only one input and one
output stream, the component mass balances become:
Input + Prod
Fin,A

Fin,B +

Output

= Fout,A
= Fout,B

As we see, will always be a positive number since Fin,A > Fout,A and Fin,B < Fout,B .
In order to generalize, let us consider more complex situation where one chemical
reaction takes place:
N2 + 3H2 2NH3

For the two reactants N2 and H2 , has a negative sign in front of it since Prod<0.
In addition, there is a stoichiometric constraint since the number of moles of H2 that
are consumed is always 3 times greater than the number of moles of N2 produced.
If we let denote the reaction parameter for the reaction in question and while
the stoichiometric coefficients are denoted i we get:

N
N

2 = 2

N2

H2
H2
=
=
(4.3)

H2
3

N2 = N H3
N H3
2
Consequently, it is also true that:
N2 = N2

, H2 = H2

, N H3 = N H3

4.4 The component mass balance method

48

In many (most) systems, more than one chemical reaction takes place. In that case
it is necessary to define one reaction parameter for each reaction. This is indeed very
useful since the ratio between calculated values of the different reaction parameters
is the same as the ratio between the extent to which the chemical reactions occur.
If there are many chemical reactions i in the system, the general component mass
balance for component j will thus become:
all

Finput,j +

recations
$

i,j i = Foutput,j

4.4.2

The reaction matrix

One cannot introduce any number of reaction parameters when component mass
balance problems are to be solved. The maximum number of reaction parameters
corresponds to the maximum number of independent chemical reactions. An independent chemical reaction is one that cannot be expressed in terms of other chemical
reactions used to characterize the system.
The concept of independent chemical reactions is also important in chemical equilibrium calculations; if too many equilibrium equations are formed for a chemical
system, the calculations will not result in an unambiguous solution.
The number of independent chemical reactions is determined by means of a reaction matrix. This is formed by the the chemical reactions (as rows) and the stoichiometric coefficients associated with the different chemical reactions (as columns).
The number of independent chemical reactions is equal to the rank of the reaction
matrix. If the rank is lower than the number of rows, one of the rows should be
removed as it is not linearly independent of the others.

Example 43: The rank of a reaction matrix


Problem
Investigate the number of independent chemical reactions in a system where the following chemical reactions occur:
Reaction 1

Reaction 2

Reaction 3

2B
C

2C

Solution
Form the reaction matrix by setting up the chemical reactions as rows and the chemical
components involved as columns:
Reaction 1
Reaction 2
Reaction 3

A B C

1 2 0
0 1 1
1 0 2

4.4 The component mass balance method

49

We now use Matlab to calculate the rank of the reaction matrix: Obviously, the rank
of the matrix is 2, not 3. Hence only two of the rows and reactions are linearly independent. These two can, in this case, be chosen freely among the three reactions.

4.4.3

Degree of freedom analysis

Example 43 indicates how the Prod terms have to be quantified when reaction systems are treated by means of mass balances. One way to look at this is:
1. Input terms are quantified by stream variables representing fluxes into the system.
2. Output terms are quantified by stream variables representing fluxes out of the
system.
3. Prod terms are quantified by means of reaction parameters, one for each reaction.
This means that the total number of unknowns in the equation system is increased
by the number of reaction parameters. Consequently, the degree of freedom analysis
becomes:
+
-

Number
Number
Number
Number

of
of
of
of

(unknown) stream variables


reaction parameters
independent component mass balances
independent information

Number of degrees of freedom

4.4 The component mass balance method

50

Example 44: Application of the reaction parameter method


Problem
In a reactor, two chemical reactions occur simultainuously. In one of those, ethane
(C2 H6 ) is dehydrated to ethene (C2 H4 ) and H2 . In the other, ethane reacts with
H2 to form the biproduct methane (CH4 ):
Reaction 1

C 2 H6

Reaction 2

C 2 H6 + H 2

C2 H4 + H2
2CH4

The feed (i.e., the input flow of reactant gas) consist of 85% C2 H6 while the rest is
inert gas (N2 ). Calculate the magnitude of all output fluxes, if the total conversion
of C2 H6 is 50.1%, while the yield with respect to formed ethene C2 H4 is 47.1%. In
addition, calculate the selectivity with respect to ethene relative to formed CH4 .
Solution
The process chart becomes::

F1,C2H6

dehydrogenering

F1,N2

F2,C2H6 F2,N2

F2,C2H4 F2,H2
F2,CH4

The next step is to form a reaction matrix. Although it is obvious that the reactions
are linearly independent (CH4 only appears in one of the reactions), this can be
analyzed formally by calculating the rank of:

C2 H6 C2 H4 CH4 H2
Reaction 1 1
1
0
1
Reaction 2
1
0
2
1

Obviously, the rank of the reaction matrix is 2, there are 2 independent reactions
and we need to introduce 2 reaction parameters in our calculations. The degree of
freedom analysis, combined with the information given results in:
Stream variables
+ Reaction parameters
- Independent MB
- Independent info

7
2
-5
-3

Degrees of freedom
1
We now introduce the basis of calculation F1 = 100 mole in order to eliminate the
last degree of freedom. This is allowed since no other flux is given. The equations

4.4 The component mass balance method

51

become:
MB C2 H6 : F1,C2 H6 1 2
MB C2 H4 :

MB H2 :
MB CH4 :
MB Inert :
Conc. inF1 :
Conversion :
Yield :
Basis of calc. :

1 2
22

F1,N2
F1,C2 H6
F2,C2 H6
F2,C2 H4
F1,N2 + F1,C2 H6

= F2,C2 H6
= F2,C2 H4
= F2,H2
= F2,CH4
= F2,N2
= 0.85(F1,N2 + F1,C2 H6 )
= (1 0.501)F1,C2 H6
= 0.471F1,C2 H6
= 100

As usual, the equations are written in matrix notation AX = Y :


0
F1,C2 H6
1
0 1 0 0 0 0 1 1
0
F1,N2

0
0
0
1
0
0
0
1
0

0
F2,C2 H6
0
0
0 0 0 1 0 1 1

F2,C2 H4
0
0
0
0
1
0
0
0
2

1
0 0 0 0 1 0 0 ,X = F2,CH4 ,Y =
A= 0
0
0
F2,H2
0.15 0.85 0 0 0 0 0 0 0

0
F2,N2
0.499 0 1 0 0 0 0 0 0

0
1
0.471 0
0 1 0 0 0 0 0
100
2
1
1
0 0 0 0 0 0 0

and the Matlab solution becomes: Thus, the selectivity is F2,C2 H4 /F2,CH4 = 40.03/5.1 =
7.849.

CHAPTER

Reactor calculations
In the previous chapters, we have mainly dealt with mass balance problems to calculate certain material flows over system boundaries, utilizing other flows, information
and balances. For reaction systems, we have also defined valuable quantities such as
conversion, yield and selectivity.
However, we have not given the chemical reactions much attention. The integral
mass balances can, at the most, provide a possibility to calculate the reaction parameter which is a measure of the total production/consumption of a substance in a
system. The integral mass balances cannot, however, explain why a reaction occurs
to a certain extent, and certainly cannot be used to describe the dynamics of the
reaction system.
In this chapter, we will study the chemical reactor by which we mean simply a
volume in which occur one or several chemical reactions. Figure 5.1 shows some
chemical reactors: a lake, a denitrification basin in a sewage treatment plant and
dough which is in the process of becoming bread in a bakery. Despite the differences
between reactors, they have two important features in common: they have a clear
system boundary, and they contain reactants that participate in chemical reactions.
In several aspects, the three reactors shown are very different. The lake is largely
mixed, although it has periodically stable stratification between the epilimnion and
the hypolimnion. On the contrary, the denitrification basin is horizontally but not
vertically mixed. If it were, oxygen would be mixed into the water and the denitrification process, which occurs under anaerobic, reducing conditions would slow down or
even stop. In terms of system boundaries, both the lake and the basin have inflow and
outflow streams. Flow rates vary with time, as well as the concentrations of dissolved
substances in the inputs and outputs as well as within the system itself.
The bread dough undergoing fermentation, however, has no inflows or outflows.
The chemical process occurring in the dough systems is primarily a transformation
of carbohydrates (sugar) to CO2 and H2 O. However, these products do not leave
the dough; they stay within the system. We see this as the dough rises. When the
fermentation has progressed to a certain level, the dough is placed on a moving belt
that will convey it into a continuous baking oven.
In this chapter, we limit ourselves to treating chemical reactions in liquid systems.
These are called homogeneous liquid phase systems. We will not treat gas phase

52

Reactor calculations

53

Figur 5.1: Three examples of reactors; a lake, a basin in a treatment plant and
a piece of dough at Lockarps in Malm.

5.1 Kinetic rate equations

54

reaction systems.1
We will focus on three fundamental concepts that are very important for chemical
reactor calculations:
Kinetic models used to calculate the rate of chemical reactions, and rate constants.
The concept of residence time which describes how much that that flows in and
out of the system.
The mixing models or reactor models that are used to describe how the reactants
come in contact with each other within the reactor.

5.1

Kinetic rate equations

The kinetic rate equation describes how fast a reaction proceeds at a given moment
mol
per unit volume and time. It has units of the type
.
volume time
Kinetic rate equations are not simply made up; they are deduced on theoretical
grounds, while the rate coefficients are determined experimentally. Table 5.1 gives an
overview of different common rate equations for different reaction types.

5.1.1

First order kinetics

The most simple is the first order rate equation. It is valid for irreversible reactions
including only one reactant:
A C
The rate equation describing how fast the reaction proceeds from left to right is:

mole
volumeunit time
where r is the reaction rate, k is the rate coefficient and cA is the concentration of the
reactant A.
The theoretical basis is simple: the probability that a certain molecule of A will
react within the system is directly proportional to the concentration of A!
One must note that the rate equation refers to the rate by which the reaction
proceeds from left to right. For the rate of consumption of A and the production of
C, one must take the stoichiometric coefficients into consideration. For A C, the
stoichiometric coefficient of A, A is -1 (it is negative since A is consumed) while C
equals 1.
To get the rate equations for A and C, one should multiply the general rate
equation r by the stoichiometric coefficient:
r = kcA

kcA

rA

rC

A r = 1kcA = kcA
B r = 1kcA = kcA

The reaction is called first order because the exponent on cA is 1, as cA = c1A .


1 Gas phase reaction systems differ from liquid phase reaction systems in two ways: as the total
number of moles change due to the reaction, the pressure and the flow rate will also change.

5.1 Kinetic rate equations

55

Tabell 5.1: Examples of kinetic rate equations. The abbreviations rev. and
irrev. denote reversible and irreversible reactions respectively. The units
are based on the use of mole, liters (L) and hours (hr) but any measure of volume
and time may be used.

Reaction
order
0

Type

Example

rA

irrev.

A B

rA = k

k:

A B

rA = kcA

A !B

rA = kf cA + kb cB

k:

A + H2 O B

r = kcA

irrev.

rev.

Pseudo 1

irrev.

irrev.

2
2
Monod

5.1.2

Units
mol
Lhr
1
: hr
1
hr

1
k : hr

2A B

rA = kc2A

k:

L
molhr

irrev.

A + B C

rA = kcA cB

k:

L
molhr

rev.

A+B !C

rA = kf cA cB + kb cC

k:

L
molhr

irrev.

Enzyme

A B
catalysis

r=

kcA
K + cA

k:

mol
Lhr ;

K:

mol
L

Second order kinetics

An example of a simple second order reaction is:


A + B C
The theoretical basis is that, in order for A and B to react they must meet, and
the probability that a certain molecule of A meets B within the system is directly
proportional to the product of the concentrations of A and B. Thus, the general rate
equation becomes:
r = kcA cB
Since the stoichiometric coefficients are -1 or 1, we get the following rate equations
for A,B and C:
rA

rB

rC

kcA cB
kcA cB
kcA cB

The reaction is called second order because the sum of exponents on cA and cB in the
rate equations is 1+1=2.

5.1.3

Overview of kinetic rate equation

Table 5.1 gives an overview of common rate equations for common stoichiometries.

5.1 Kinetic rate equations

56

The table is based on the reaction order of the reaction in question. One should
note that if the reaction rate is independent of the concentrations of reactants, the
reaction is said to be a zero order reaction
One should also be aware that the the unit of the rate coefficient will differ with
reaction order. The principle is that the units of the coefficient must match the
reaction order (expressed as cA , c2A , cA cB , etc.) so that the unit of r is correct.
A special case is the biological reactions described by Monod kinetics or MichaelisMenten kinetics. Many enzyme catalyzed reactions follow this type of kinetics, which
can be inferred theoretically. Such reactions are of 1st order at low concentrations of
reactants, but at 0 order at high concentrations of reactants. The cut-off is gradual,
but affected by the the value of the half-rate coefficient, K. More on this in Chapter
10.

5.1.4

Reversible reactions

As shown in Table 5.1, rate equations for reversible reactions have the form of the
difference between the rate of the backward reaction and and the rate of the forward
reaction. This means that the net reaction may go in either direction, depending on
the concentrations of the reactants and products.
A special case is when the net reaction rate is 0. This implies that the forward
and the backward reactions proceed at the same rate. This state is called equilibrium.
Thus, we can state that for the reversible reaction A ! B, chemical equilibrium is
defined as:

0 = kf cA kb cB
r = 0 " kf
cB

=
=K
kb
cA
Hence, chemical equilibrium coefficients can be viewed as the ratio between the coefficients of the forward and the backward rate coefficients.

5.1.5

Consecutive reactions

In practice, it is rare that reaction products are perfectly stable. One example is
alcohol in wine; it is formed by sugar but will react to acetic acid (vinegar) if it is
exposed to oxygen. In this case, alcohol is an intermediate product between sugar and
acetic acid. The reaction system is an example of a system where consecutive reactions
occur.
More formally stated, one example of a consecutive reaction is:
k

1
2
AB
C

Here , B is the intermediate in the reaction AC. If we break this reaction down
into its two first order reactions, given index 1 and index 2 respectively, we can
deduce the rate equation for each component. First we recognize that:
r1 = k1 cA
r2 = k2 cB

5.2 Mean residence time and reaction time

57

As we see, the stoichiometric coefficients are either -1 (for reactants) or 1 (for products). Hence the rate equations for A, B and C become:
rA

= r1

rB

= r 1 r2
= r2

rC

= k1 cA

= k1 cA k2 cB
= k2 cB

It is thus possible to express the equilibrium constant in terms of a ratio between two
rate coefficients, or cA and cB .

5.1.6

Parallel reactions

In parallel reactions, a reactant may participate in several reactions. For example, if


the reactant A may react either to B or to C, the reaction system includes:
A
A

k2

Then, if:
r1 = k1 cA
r2 = k2 cA
we get:
rA
rB
rC

= r1 r2
= r1
= r2

= (k1 + k2 )cA
= k1 cA
= k2 cA

One example of parallel reactions in nature, is the reaction of nitrous oxide, NO2 , in
the formation of nitric acid, HNO3 , from reactions either with OH or NO3 :
NO2 + OH
NO2 + NO3

5.2

HNO3
N2 O5 + H2 O 2HNO3

Mean residence time and reaction time

In all systems in which chemical reactions occur, the conversion depends on how long
the reaction takes place inside the system. If the chemical reaction proceeds for a
short time, the conversion from reactants to products will be lower than if it takes
longer.
In Figure 5.1, three examples of chemical reactors were shown. One of the differences between them was that while two (the lake and the basin) both has inflow and
outflow, the third (the dough) has neither. Thus, the lake and the basin are examples
of open systems while the dough is an example of a closed system.
We can now introduce three very important quantities relevant to open systems:
The volume of a reactor: V (liters, m3 , etc.)
The volumetric flow rate into or out of a reactor: Q (liters/s, m3 /hour, etc.)

5.2 Mean residence time and reaction time

The average residence time for the reactor: =

58
V
(sec, hours, etc.)
Q

Technical systems, the flow rate is often constant. In natural systems however, the
inflow to lakes or the flow in watercourses can vary greatly. Then the flow Q is not
a constant value, but may vary 1-2 orders of magnitude between high flow periods,
such as when snow melts in as compared to dry periods in the summer.
If the average volumetric flow rate is to be measured, we can distinguish three
different cases. First, the flow can be constant. Second, the flow may be variable
and the measurements may be evenly or unevenly spaced in time. Finally, the flow
may be variable but the measurements continuous. For these three cases, the average
volumetric flow rate is determined as:

Q
constant flow

n
$

1
Qi
n evenly spaced measurements
= n
(5.1)
Q

i=1

2 t

1
Q(t)dt continuous measurements
t
0

the residence time is


Based on the calculation of the average volumetric flow Q,

calculated from the volume as V /Q. An example of a calculation with variable flow
and monthly measurements (case 2) is given in Example 51 on the following page.
For an open system there is no time=0, since the flow is continuous. For a
closed system however, which has no volumetric inflow or outflow, one can say that
time=0 is when the doors are closed. In terms of chemical reactions, t = 0 is when
the chemical reaction starts in the closed system. In later sections in this chapter, we
will look into this situation in conjunction with the batch chemical reactor.

5.2 Mean residence time and reaction time

59

Example 51: Average residence time of a lake


Problem
Lake Kvarnsjn in the province of Vsterbotten receives its inflow water from the
stream Aborrbcken. In the environmental monitor program, the flow rate is measured the 15th of each month. Calculate the average residence time of the water in
Kvarnsjn, if the lake area is A = 15 ha and the average lake depth is h = 3.5 m,
expressed in years. Use the data for 1999 given below.
Flow
Month
(m3 day1 )
January
222
February
609
March
3327
April
2730
May
2870
June
693
July
33
August
93
September
441
October
1461
November
1147
December
435
3500

Solution
The following calculations are made in Matlab, resulting in an average residence time
3000
of water in the lake of = 1.2 years:
Uppmtt flde (m3/dygn)

2500

2000

1500
Medelflde
1000

500

7
Mnad

10

11

12

5.3 Reactor models

5.3

60

Reactor models

In this section we deal with the basics of reactor calculations by studying ideal reactor
models. They always are the starting point when trying to describe the conditions
in a reactor system, especially with respect to mixing. Mixing is of fundamental
importance because it affects the concentrations in the system, which in turn affect
the rate of chemical reactions. Here, we will only consider macroscopic mixing, not
diffusive transport at the molecular level.
It is also of fundamental importance whether the reactor is an open or closed
system.
Given these to dimension, mixing and open/closed, there are three ideal reactor
models to consider:

1. The ideal tank reactor

2. The ideal batch reactor

3. The ideal plug-flow reactor


As the drawings illustrate, these types of chemical reactor all differ since:
1. The ideal tank reactor is open and perfectly mixed.
2. The ideal batch reactor is closed and perfectly mixed.
3. The ideal plug-flow reactor is open and not mixed.

5.3.1

The differential mass balance

All of the ideal reactor models can be characterized by the same mass balance equation, the differential (component) mass balance2 . As reaction systems are of prime
interest, the component mass balance will always be expressed in terms of molar fluxes
of components, denoted FA , FB , etc. Furthermore, the fluxes will be given in molar
fluxes per unit time (i.e., mol/hour etc.).
The general mass balance for a given component, valid for all ideal reactor models
is:
Input + Prod

Fin + rV

Output + Acc
d(cV )
Fout +
dt

mole
unit time

where F is the molar flux (mole/unit time), c is the molar concentration (mole/unit volume), Q is the volumetric flow rate (volume/unit time) and V is the reactor volume
(volume).
2 From

now on it will be understood that mass balances refer to components, not elements.

5.3 Reactor models

61

Although based on Input + Prod = Output + Acc, there are differences relative to
how the mass balance principle was used in Chapters 2-4: the Prod-term is expressed
in terms of kinetic rate equations, and the Acc-term deals with possible changes with
time.
The Input and Output-terms: The terms representing the fluxes across the reactor system boundaries, Fin and Fout , are given by the product of the concentration
c and the volumetric flow rates Qin and Qout respectively. If the flow rate is constant,
Qin = Qout = Q.
The Prod-term: The Prod-term quantifies how much of a substance that is produced in the reactor per unit of time. The Prod-term is formed by multiplying the
kinetic rate equation (mole/unit time and unit volume) by the volume. This is logical
since scaling up will lead to a higher Prod while scaling down will decrease the
amount converted in the chemical reaction.
If, for example, only the reaction A B occurs with rA = kcA , we get:
Prod = kcA V
In more general terms, we can express the Prod-term for component j reacting in
several chemical reactions i, as the sum of the contributions for all reactions:
Prodj =

all i
$

i,j ri V

i=1

Thus, in terms of calculations there are three actions related to the Prod-term:
1. Identify all reactions.
2. Define the corresponding kinetic rate equations.
3. Include the kinetic rate equations in the Prod-term with the right stoichiometric
coefficients and signs (+ or -).
The Acc-term: The Acc-term quantifies the change in the number of moles per
unit of time (i.e., the timederivative of the molar amount). The molar amount in the
reactor is given by the product of the concentration c and the volume V (i.e., c V ).
By taking the differential of cV , we get the Acc-term as:

dc
dV

+c
if V varies with time
V
d
dt
dt
Acc = cV =

dt
dc

V
if V is constant
dt

In summary: For a component A which reacts in one single reaction in a reactor


system where the flow rate and the volume is constant, the differential mass balance
becomes:
Qin cin,A + rA V = Qin cout,A +

d(cA V )
dt

mole
unit time

(5.2)

5.4 The ideal completely stirred tank reactor, CSTR

62

Figur 5.2: Many environmental systems can be modeled by means of a CSTR.

5.3.2

Methodology for reactor calculations

Regardless of rector model, the chemical reactor calculations will follow the procedure:
1. Define the flow and mixing conditions in terms of an ideal reactor model.
2. Determine the reactor-specific parameters.
3. Define the kinetic equation(s) and rate constant(s).
4. Combine the reactor model and the kinetic equations(s) in the differential mass
balance.
5. Do the calculations, analytically or numerically.

5.4

The ideal completely stirred tank reactor, CSTR

Almost all industrial and natural, open systems can be described in terms of one or
more ideal tank reactors, often called Continuously Stirred Tank Reactor, CSTR.
Since the CSTR is perfectly mixed, the concentrations (and temperature) are the
same everywhere in the entire reactor. Thus, there are no gradients whatsoever in the
reactor. The important implication of perfect mixing is that the reaction rates are
the same everywhere in the reactor. The general mass balance for a CSTR becomes:
cin
Qin

crV

cut
Qut

In + Prod

Out + Acc

Qin cin + rV

Qout cout +

At steady state, when Acc=0 och Qin

d
cV
dt
= Qout = Q the simplified model is:

In + Prod

Ut

(5.3)

Qcin + rV

Qcout

(5.4)

Final remark: Never forget that the concentration in the reactor c is the same as cout !

5.4 The ideal completely stirred tank reactor, CSTR

63

Example 52: CSTR with a 1st order irreversible reaction


Problem
Calculate output stream fluxes from a steady-state CSTR in which a 1st order reaction
occurs:
AB
r = kcA = 0.5cA mole m3 min
provided that the volumetric flow rate is 1.25 m3 min1 , the reactor volume is 5 m3 ,
and the input stream concentration of the reactant A is 2 mole m3 while the input
concentration of B is 0.2. mole m3 .
Solution
First we select the appropriate model: the steady-state CSTR. For a CSTR the
differential mass balances are given by:
Qin cin,A + rA V

Qout cout,A

Qin cin,B + rB V

Qout cout,B

Since Qin = Qout = Q, cout,A = cA and cout,B = cB the equations can be simplified
to:
Qcin,A + rA V

QcA

Qcin,B + rB V

QcB

All of the reactorspecific parameters are given in the problem definition. The kinetic
rate equations for first order reaction with respect to the reactant A and the product
B are:
#
rA = A r = 1r = kcA
rB = B r = 1r
= kcA
We can now combine the kinetic equations and the differential mass balance model.
The combined equations are:

Qcin,A = +kcA V + QcA


Qc
in,B

= kcA V + QcB

or, if we introduce the residence time =

cin,A = cA (1 + k)

c
in,B = kcA + cB

V
:
Q

cin,A
1 + k
, cB = cin,B + kcA
, cA =

(E52.5)

Equation E52.5 allows us to solve directly for cA . Before inserting numbers, we


5
recognize that =
= 4 minutes, so that:
1.25
2
cA =
= 0.667 mole m3
1 + 0.54
and thus, cB = 0.2 + 0.5 4 0.667 = 1.533 mole m3 .

5.4 The ideal completely stirred tank reactor, CSTR

64

Example 53: Example 52 with matrix notation


Problem
Utilize the fact that the above reaction calculations can be seen as a system of linear
equations, and solve this system with Matlab.
Solution
The first step is to re-write E52.5 :
#
cin,A = cA (1 + k)
cin,B = kcA + cB
with all terms including cA and cB on the left-hand side:
#
(1 + k)cA + 0 cB = cin,A
kcA + 1 cB = cin,B
This system of equations can be written in the form AX = Y :
!
"
!
"
! "
cin,A
1 + k
0
cA
,
Y =
A=
,
X=
cin,B
k
1
cB
Thus the Matlab solution becomes:

5.4 The ideal completely stirred tank reactor, CSTR

65

Example 54: Numerical solution for non-unity reaction order


Problem
Repeat the calculation in Example 52 given that the chemical reaction and the rate
equation are:
1
A B
2

r = kcA 1.2 = 0.5cA 1.2 mole m3 min

Solution
Here, the equation E52.5 will contain a non-linear term:

cin,A = cA + kc1.2
cin,A = cA (1 + kc0.2
A
A )
1
c
1.2
in,B = 2 kcA + cB

Nevertheless, this equation can also be written in the form AX = Y :


!
"
!
"
! "
0.2
1 + kcA

0
cin,A
cA
A=
,
X
=
,
Y
=
cin,B
cB
12 k
1
In Matlab this class of problems are solved as 0 = AX Y with fsolve. The
methodology, outlined in Example 14, means that fsolve determines X so that
F = AX Y 0. To define the equation, one creates a Matlab m-file with the
function: and gets the solution: Thus cA = 0.6989 mole m3 and cB = 0.8506
mole m3 .

5.5 The ideal batch reactor

5.5

66

The ideal batch reactor

The ideal batch reactor consists of a closed system (i.e., a reactor, which has neither
inflow or outflow). In the reactor, the reactions cause the concentrations of reactant
and products to change with time. One cannot speak of a batch reactor at steadystate.
In industry batch reactors are very common. Production of active ingredients in
the pharmaceutical industry is always done in batch to meet the quality requirements.
Multi-purpose batch reactors are used to produce different substances in the same
reactor. Another example of the use of batch reactors is biotechnological fermentation
reactors used to produce biomaterials using specific enzymes.
In the environment one can also find systems that are best modeled as a batch
reactor. An example is a drop of water suspended in the atmosphere. In such droplets,
many chemical reactions may occur, for example oxidation of SO2 to H2 SO4 . But the
most important example is when the input flow to a normally open system suddenly
decreases significantly or stops. In nature, this might be a lake during the dry season
when it hardly receives any discharge from the catchment. Another example is when
the inflow pump to a basin within a sewage treatment plant suddenly breaks down.
Then, a system that normally acts as a steady-state CSTR changes into a batch
reactor.
Since In=Out=0 for a closed reaction system, the general component mass balance
for a batch reactor becomes:
crV

Prod

Acc
dc
rV
=
V
dt
or, if the volume V is eliminated:
dc
=r
dt

(5.5)

At this point, one can see that there are two ways to use the above reactor model,
the differential view and the integral view. In both cases, we must be aware that the
boundary conditions are not related to input and output fluxes (i.e. Qcin or Qcout ).
Instead, the boundary conditions relate to time. The relevant boundary condition is
thus the initial condition, or inititial concentration co , which denotes c at time t=0.
With the differental view we solve equation 5.5 to get:
dc
= r " c(t) from t=0 to t=t
dt
In this case we get an expression that allows us to evaluate c for every t between 0
and t. It provides an answer to a question such as, What is the concentration after
a certain time.
With the integral view, we do the opposite:
dc
= r " t required to change c from c0 to c
dt

5.5 The ideal batch reactor

67

This method is more useful if one asks, How long time does it take to reach concentration c.
In more mathematical terms, the two cases can be expressed as:
#
t = 0, c = c0
dc
Differential :
=r
(5.6)
dt
t = t, c = c
2

Integral :

dt =

c0

dc
r

(5.7)

Example 55: Batch reactor with 1st order irreversible reaction


Problem
Calculate the concentration of the reactant A if it reacts in a batch reactor for 4
minutes according to a first order reaction.
AB

r = kcA = 0.5cA mole m3 min1

if the initial concentration, c0,A is 2 mole m3 .


Solution
First we identify the correct mass balance equation for the reactor in question:
Acc
dcA
dt

Prod

rA

and derive the correct kinetic rate equation for A:


rA = A r = kcA
Combining the mass balance equation and the kinetic equation results in the reactor
model:
dcA
kcA
2
1 cA dcA
k c0,A cA
2 cA
d ln cA

=
=

dt
2 t
0

c0,A

dt
2

dt

0
k[t]t0

[ln cA ]ccA
0,A
cA
ln
c0,A

=
=

kt

cA

c0,A ekt

With numbers inserted we get cA = 2e0.54 = 0.2707 mole m3 .

(E55.8)

5.5 The ideal batch reactor

68

Example 56: Differential, numerical solution of Example 55


Problem
For the above reaction and reaction system, calculate by means of numerical integration the concentration of A after 4 minutes. Use the integration method ode45 in
Matlab.
Solution
As shown above, the mass balance for the batch reactor and kinetic equation combined
give:
dc
=r
dt

"

dcA
= kcA
dt

This differential equation should be integrated in a way that gives us c(t) for t = 0 to
t = 4. The initial value with respect to A is c0,A = 2. First the function that should
be integrated is created: Thereafter the function is simulated. To find the actual
final value of cA one can make Matlab show the lowest value that cA has attained
during the calculation. The result becomes:

1.8

koncentrationen av A (mol/m3

1.6

1.4

1.2

0.8

0.6

0.4

0.2

0.5

1.5

2
tid (minuter)

2.5

3.5

5.5 The ideal batch reactor

69

Example 57: Integral, numerical solution of Example 55


Problem
For the above reaction and reaction system, calculate by means of numerical integration the time it takes for the concentration of A (cA ) to decrease from 2 to 0.2707
mol m3 . Use the integration method quad8 inMatlab.
Solution
The analysis above gave as intermediate that:
2 t
2 cA
dcA
dt =
kc
A
0
c0,A
We can now form the function that should be integrated, and define it in the m
file ex56: Thereafter the function in ex56 is integrated from cA = c0,A = 2 to
cA = 0.2707. In Matlab this is done as: The result is t 4, which of course
corresponds perfectly with the result obtained in Example 55.

5.5 The ideal batch reactor

70

Example 58: Simulation of consecutive and parallel reactions


Problem
Suppose that the following 4 reactions, involving 3 components, occur at the same
time in a batch reactor system:
r

1
A

2
B

r3

A!

r4

Simulate the system between t = 0 to t = 10 with the initial conditions c0,A =


1, c0,B = 0andc0,C = 0. The stoichiometric coefficients lead to the following rate
equations for A, B and C:
r1
r2
r3
r4

= 1cA
= 0.4cB
= 0.1cA
= 0.2cC

Model
The kinetic equations are combined with the mass balance equation for the batch
reactor to give reactor model equations:
MB A
MB B
MB C

dcA
dt
dcA
dt
dcC
dt

= r1 r3 + r4 = cA 0.1cA + 0.2cC
= r1 r2 = cA 0.4cB

= r2 + r3 r4 = 0.4cB + 0.1cA 0.2cC


1

0.9

0.8

0.7
koncentrationen av C
0.6

0.5

0.4
koncentrationen av B
0.3

0.2
koncentrationen av A
0.1

Numerical solution and answer

10

5.6 The ideal plug-flow reactor, PFR

5.6

71

The ideal plug-flow reactor, PFR

Above, we discussed the CSTR and the batch reactor. These two ideal reactor models
share the same feature: they are all perfectly mixed.
The ideal plug-flow reactor is not mixed at all in the direction of flow. Perpendicular to the direction of flow, however, it is perfectly mixed. This implies that all fluid
that enters a plug-flow reactor moves forward at the same velocity. One can view the
reactor in several ways, but one useful conceptual model is that the fluid moves as a
series of thin slices along the reactor.
One consequence of the lack of mixing is that each slice does not know what the
concentration of different components is in the proceeding slice, nor in the trailing
slice.
When making a mass balance model of any kind, it is important the the volume
considered is homogeneous. In a plug-flow reactor, the smallest unit that is homogeneous is the individual slice. This volume element is sometimes called control
volume. The volume of the control volume is differentially small, so the volume in
question is dV .
When a reaction occurs within dV , the molar flux will change by a differential
amount. Therefore, if the input flux to dV is F , the output molar flux will be
F + dF . If the flow Q is constant, the input concentration to dV is c while the output
concentration is c + dc.
With this in mind we can create the differential mass balance equation for a volume
dV within a (steady-state) plug-flow reactor as:

cin

cut

c+dc

Q
V

dV
r
Input + Prod

Output

F + rdV

F + dF

Qc + rdV

Q(c + dc)

or, after dividing with Q:


rdV

1
= dc , or rd = dc
Q

(5.8)

Note that this is only true for the steady-state PFR, for which the Acc-term = 0.
In this case, steady state means that the concentration in each point in the reactor is
the same, despite the fact that the concentrations vary along the direction of flow.
We now realize that, just as for the batch reactor, the mass balance equation may
either be written in differential or integral form:
#
V = 0, c = c0
dc
r
Differential :
=
(5.9)
dV
Q
V = V, c = c
Integral :

dV
=
Q

c0

dc
r

(5.10)

5.6 The ideal plug-flow reactor, PFR

72

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

dV

Figur 5.3: Illustration of how a PFR can be seen as a train of infinitely small
batch reactors with the volumes dV that together add up to the volume V.

The mass balance equations for the PFR are virtually identical to the corresponding equations for the batch reactor, as shown in equation 5.6. This becomes even
clearer if we substitute V /Q with :
#
= 0, c = c0
dc
Differential :
(5.11)
=r
d
= , c = c
Integral :

d =

c0

dc
r

(5.12)

If one, such in Example 55 solves the mass balance model analytically for the
first order reaction AB, one gets:
V

rA = kcA " cA = c0,A ek Q = c0,A ek

(5.13)

Obviously, the PFR and the batch reactor are modeled in exactly the same way,
the only difference being that the reaction time t for the batch reactor is substituted
by the residence time for a PFR. This coincidence has been illustrated in Figure
5.3. The figure shows that a PFR can be seen as a train moving through a tunnel
where each wagon is independent of all the others. The time that a wagon stays
in the tunnel corresponds to the residence time, while the time the reactants stay
in the wagon corresponds to the reaction time,

5.6.1

PFR reactor modeling in terms of conversion

In chemical engineering, it is common that the mass balance equation and the kinetic
rate equation are expressed in terms of the conversion, rather than in local concentrations. By introducing the conversion, it is possible to express all concentrations of
reactants in terms of input concentrations, cin so that c = cin (1 X). If we define

5.6 The ideal plug-flow reactor, PFR

73

the conversion of A as XA , the mass balance over the element dV in a PFR becomes:
In + Prod

Out

Qcin,A (1 XA ) + rA dV

Qcin,A (1 XA dXA )

Combined with the kinetic equation for a first order reaction rA = kcin,A (1 XA )
we get:
Qcin,A (1 XA ) kcin,A (1 XA )dV

Qcin,A (1 XA dXA )

dV
Q

dXA

k(1 XA )

At this point one has to substitute variables, i.e. = (1 XA ) so that:


d

dV
Q

1
[V ]V
Q 0

1dXA
2
1 d

k 1
1

[ln ]1
k

Since ln( = 1) = 0 one gets, with = (1 XA ) re-inserted:


=

ln(1 XA )
k

This equations can be seen as a normalized inverse of equation 5.8, which gave cA as
a function of the residence time .

Example 59: Emission of a pollutant into a stream


Problem
A pollutant A is, by mistake, emitted from a chemical plant to a recipient stream.
Calculate how far downstream from the plant the concentration of pollutant has
decreased by 90%, provided that the pollutant is degraded irreversibly by a 0.5 order
chemical reaction:
Data:
Kinetic rate constant k: 0.0008 mole0.5 m1.5 min1
Steam water flow Q : 10 m3 s1
Steam cross section area A r 50 m2
Concentration of A at the plant: 0.02 mole m3
Solution
The nature of the flow system (and the fact that the example is in this section)
suggests that the most appropriate analogy to the natural system is the PFR. In
order to solve the problem, the mass balance equation has to be written in a form
where is is possible to solve for the distance L downstream, of the source point. The
source point is thus at L = 0.

5.6 The ideal plug-flow reactor, PFR

74

The substitution is made by recognizing that the control volume dV can be written
as the product AdL. We can then re-write equation 5.8 as:
r
dV
Q

dc

rA
dL
Q

dc

Combined with the rate equation rA = kc0.5


A we get:
kc0.5
A

A
dL = dcA
Q

In this case, the conversion of A, XA should be 90%. We can then chose to write the
concentrations cA in terms of cin,A and the conversion so that dcA = cin,A dXA :
A
dL = cin,A dXA
Q
0.5 A
kc0.5
dL = dXA
in,A (1 XA )
Q

0.5
kcin,A
(1 XA )0.5

We will now solve the problem from a differential as well as an integral perspective:
#
L = 0, XA = 0
dXA
0.5
0.5 A
Differential :
= kcin,A (1 XA )
dL
Q
L = L, XA = 0.9
Integral :

Qc0.5
0,A
dL =
kA

0.9

dXA
(1 XA )0.5

5.6 The ideal plug-flow reactor, PFR

75

We let Matlab solve the integral form with quad8. The Matlab m-file is: and it is
executed with: The integration shows that the concentration of A will have decreased
by 90% (or to 10%) of the initial concentration ca 2900 meter downstream the source
point. The same result is provided by the differential form, as simulated with the
command:
1

0.9

0.8

omsttning av froreningen

0.7

0.6

0.5

0.4

0.3

0.2

0.1

ode45(ex58diff,[0 3000],[0]):

500

1000
1500
2000
strcka nedstrms utslppspunkten (m)

2500

3000

CHAPTER

Non-ideal reactors
Chapter 5 presented the basis for chemical reactor calculations. It included theory
and examples on how to calculate the extent to which a chemical reaction occurs as a
function of the reaction kinetics, the concentration of reactants and products, as well
as mixing conditions. The mixing conditions were incorporated by using the ideal
reactor models CSTR, batch and PFR. These reactor types can be seen as extremes,
where the tank reactor and batch reactor are based on complete mixing, while the
PFR is stirred in two dimensions but not along the direction of flow. In all cases,
except 0th order reactions, the mixing conditions affect conversion of reactants in the
system.
In many cases, however, both the ideal CSTR and the ideal PFR are quite bad
approximations of the real mixing conditions in a specific open systems. Thus, in
those cases our models and reactor calculations can be misleading if we use on of the
ideal reactor models. This chapter describes the causes of non-ideal behavior, general
methods to describe the mixing conditions, as well as helpful non-ideal reactor models.
The most important new concept is residence time distribution.

6.1

Examples of non-ideal mixing

In engineered, man-made systems, such as equipment for unit processes and chemical
reactors, one design criteria is to get as close to ideal mixing as possible. For example,
the design of a distillation tower aims at obtaining complete mixing at each step, and
in a biological bed one wants to get maximum mixing in order to obtain sufficient
oxygenation in all parts of the bed.
In nature, it is obvious that the variability of the mixing in time and space is large,
and the mixture conditions are more complicated. Figure 6.1 illustrates how the flow
can occur along a stretch of river. Windings, connections and bay islands may affect
the water flow so that a certain proportion of water mass will be lagging behind,
moving at lower speeds than average.
The presence of biota complicates the picture further; organisms can of course
take up, store, destroy and redirect different substances in many ways. As an example, consider the radionuclides that spread all over Sweden in conjunction with the
Chernobyl accident in 1986. Much of the cesium ended up in lakes and is still present

76

6.2 Residence-time distributions

77

Figur 6.1: Illustration of how the water in a river or stream may be irregular,
so that some packages of water move faster while others move slower than
average.

there, tied up in the tissue of fish, though it would otherwise have been washed out
since long.
These are examples of situations in which a reaction system cannot be described
well with an ideal reactor model (CSTR or PFR). In summary, one can say that nonideal mixing conditions arise due to short-cutting, channel formation, the presence of
stagnant zones, or back-mixing.

6.2

Residence-time distributions

The most useful and general way to characterize the mixing conditions is to determine
how long each small volume element resides in system. One finds that some volume
elements pass through a system quickly, while others stay longer. This creates a
unique residence time distribution for the system. This is sometimes abbreviated
RTD.
An RTD can be determined experimentally by adding a tracer to the inflow to a
system. As a tracer, one should use a substance that is neither consumed, produced,
or delayed in the system. Non-toxic anions such as Cl and Br can be used in
natural systems. Under controlled conditions one may also use radioactive isotopes
such as 15 N or 3 H. One can also make use of natural tracers, such as the stable
isotopes 13 C and 18 O.
The conceptual starting point for the discussion and calculations regarding RTD
is that added particles of tracers behave exactly as elements of the fluid in the system
when added.

6.2 Residence-time distributions

6.2.1

78

The normalized RTD E(t)

If you add a certain amount of tracer at a given moment (instantaneously) different


tracer particles will leave the system at different times, some sooner and some later.
It is measured as a variation in the concentration of tracer in the output stream. If
the input point (injection point) and the output point are close, a fair share of the
tracer may quickly leave the system. If the distance is long, or the mixing is poor, it
will take much longer until the first portion of tracer leaves the system.
From now on, we will call the injection point the upper boundary and the output
point the lower boundary.
Let us assume that a certain mass of tracer, MS , is added to the system at the
upper boundary at time t = 0. At the lower boundary, the concentration can be
measured as a function of time, and will be c(t). Obvisously, c(t) will in some way
directly reflect how many tracer particles have stayed during t in the system, since
they were all added at t = 0.
Regardless of how the tracer moves through the system, we can be sure of one
thing: after an infinitely long time, all the tracer will have left the system. Mathematically, we can express it as the sum of the amount of substance that comes out
from t = 0 to t = :
Ms =

where
Ms =
Q=
c(t) =

(6.1)

Q(t) c(t)dt

amount of tracer
volumetric flow rate
tracer concentration

mass
volume time1
mass volume1

Note that if Q(t) is constant, the factor Q can be placed outside the integral.
To get the residence time distribution, one should simply normalize the the measured concentration c(t). In this context, normalizing c(t) means that the integral of
c(t) with time becomes independent of MS .
The normalized c(t), the RTD, is called the E(t)-distribution. If Q is constant,
E(t) is calculated as:
E(t) = c(t)

c(t)
Q
=2
Ms
c(t)dt

1
time

(6.2)

From Equation 6.2 it follows that


2
E(t)dt = 1

(6.3)

This is really quite obvious, since the similarity implies that all trace elements which
are introduced into a system must have a residence time between 0 and .
Another way to interpret E(t) is to see what fraction of the injected tracer remains
after a certain time in the system. We may then split the integral Equation 6.3 into

6.2 Residence-time distributions

cin=0
Q

79

c=0
c

t<0

cin=hg
Q

c=ct=0
c

t=0

V
Q

cin=0
Q

c < ct=0
c

t>0

V
Q

cin=0
Q

c << ct=0
c

t >> 0

V
Q

Figur 6.2: Illustration of a tracer experiments in a CSTR.

two parts so that:


The fraction that stays shorter than time t in the system
The fraction that stays longer than time t in the system

E(t)dt (6.4)
2 0
=
E(t)dt(6.5)
t

6.2.2

c(t) for an ideal CSTR

Conceptually, the CSTR is the most simple ideal reactor model. So, which E(t) does
this reactor have?
If one instantaneously adds a pulse of tracer at the upper boundary at t = 0, the
tracer will immediately be distributed perfectly evenly in the entire volume of the
CSTR. In fact, this is the property that makes the CSTR ideal (see Figure 6.2).
If the added pulse of tracer amounts to Ms and the volume of the reactor is V , the
tracer concentration c(0) in the system will be:
c(0) =

Ms
V

To calculate E(t) we must first calculate c(t) (i.e., the concentration of the tracer
in the output as a function of time). We derive this by starting at the general mass
balance for the CSTR:
Input + Prod

Qcin + rV

Output + Acc
d(cV )
Qc +
dt

mass
unit time

6.2 Residence-time distributions

80

c(0)

c(t)
0.0
0.0

0.5

1.0

1.5
2.0
2.5
tid/medeluppehllstid

3.0

3.5

Figur 6.3: c(t) for a CSTR. The value of c(t) for t = is indicated.

where Q is the volumetric flow into the reactor. However, since there is no tracer
in the input flow (the tracer is added as a pulse) and since tracers do not react (by
definition), we get cin = 0 and r = 0. hence, the mass balance is reduced to:
d(cV )
= Qc
dt

(6.6)

Since V /Q is the mean residence time we get:


dc

dc
c(0) c
c
ln
c(0)

1
dt

2
1 t
dt
0
t

c(0)et/

(6.7)
(6.8)
(6.9)
(6.10)

The obtained function c(t) thus shows that a tracer which is added to an ideal (i.
e., perfectly stirred) tank reactor, will be flushed out of the reactor according to an
exponential process (Figure 6.3).

6.2 Residence-time distributions

81

E(t) fr CSTR

1/
0.8/
0.6/
0.4/
0.2/
0.0
0.0

0.5

1.0
t /

1.5
2.0
tid/medelupphllstid

2.5

3.0

Figur 6.4: E(t) for a CSTR. The value of E(t) for t = is indicated.

6.2.3

E(t) for an ideal CSTR

To get E(t) we apply Equation 6.2 to the c(t) obtained above:


E(t) = 2

c(t)

c(t)dt

c(0) et/
et/
2
=2
ct=0
et/ dt
et/ dt
0

(6.11)

Since the primitive function to et/ is et/ , we simply get:


E(t) =

et/
et/
=

[et/ ]
0

(6.12)

Figure 6.4 shows E(t) for the ideal CSTR. The figure indicates the value of E(t)
for t = , which is e1 / 0.37/. This means that when one mean residence
time () has passed, the concentration of tracer will amount to 37% of the original
concentration. This also means that 63% of the tracer has left the reactor.

6.2.4

Mean residence time from E(t)

For reactor calculations, the mean residence time is crucial quantity. We will now
go through how the value of can be determined from tracer experiments by using
E(t).
The mean residence time is calculated by simply weighting together the time
the tracer particles stay in the system and the proportion of elements that stays the
corresponding time1 . Doing this, we must get:
2
=
tE(t)dt
(6.13)
0

1 In

more advanced literature, one speaks about the "first moment", the moment that the residence
time distribution creates relative to the y-axis.

6.2 Residence-time distributions

82

1.0

F(t)

0.8
0.6
0.4

E(t)

0.2
0.0
0.0

0.5

1.0

1.5
2.0
2.5
tid/medeluppehllstid

3.0

3.5

Figur 6.5: F (t) or a CSTR, which shows the proportion of the volume elements,
or tracer, which has left the reactor after a certain time.

We can now show that the hydraulic residence time = V /Q of a CSTR is identical
to the average residence time of the added tracer. We show that by inserting the E(t)
of the CSTR (Equation 6.12) into Equation 6.13:
2

tE(t)dt =

6.2.5

et/
1
t
dt =

tet/ dt =

1 2 tet/
[ (
et/ )]
0 =

2
= (0 (1)) =

The F(t)-distribution

The equations 6.4 and 6.5 postulate how to calculate the proportion of fluid elements
or a trace element that comes with them that have left the system after a specified
time. The accumulated amount that has left the system is called the F (t)-distribution.
For a reactor with a known E(t) we get F (t) by integrating E(t) with respect to
time. Thus, for a CSTR, F (t) becomes:
F (t) =

et/

dt = [et/ ]t0 = (et/ (1)) = 1 et/

(6.14)

Figure 6.5 shows how F (t) increases, and F (t) =1 when all of the tracer has crossed
the lower boundary.2

2 One can also show that the trace element is added continuously from t = 0. The resulting E(t)
will look exactly like the F (t) obtained if the tracer is added as a pulse to the reactor.

6.2 Residence-time distributions

83

Example 61: Residence-time distribution from tracer experiments


Problem
To study the flow of water through an artificial pond in a stream, an inert tracer is
added as a pulse in the influent. In a natural system, this is a useful method, because
the flow often cannot be determined by direct measurement. The effluent concentration is measured as follows:
Time (min)
0
2
4
6
8
10
12
14
16
18
24
30
40
50

c(t)
0
1.1
2.9
4.8
5.9
6.2
5.8
5.1
4.3
3.7
2.2
1.2
0.4
0.2

Sprmneskoncentration

6
5
4
3
2
1
0
0

10

20
30
Tid (minuter)

40

50

60

Calculate the mean residence time of the reactor from the resulting data.
Solution
The average residence time is obtained by going through the steps:
1. Make an E(t)-distribution from the resulting data.
2. Calculate the mean residence time by numerical integration of Equation 6.13.
The numerical integration can be performed in a variety of ways. Here a crude
method is used. One measuring point represents the interval between to measurements. Compared to errors caused by field measuring errors however, the integration
error is normally rather small. The time interval t around a measurement point i
has been defined as:
ti =

1
1
((ti+1 t) (t i 1) = (ti+1 ti1 )
2
2

Graphically, this means that:

6.2 Residence-time distributions

84

Sprmneskoncentration

6
5
4
3
2
1
0
0

10

20
30
Tid (minuter)

40

50

60

One can also directly see that the mean residence time must be greater than the time
at which the concentration of tracer in the output is greatest. This is because the
tail of the tracer gives a high contribution to the mean residence time.
The necessary calculations are compiled in the following table:
i
0
1
2
3
4
5
6
7
8
9
10
11
12
13

t (min)
0
2
4
6
8
10
12
14
16
18
24
30
40
50

ci
0
1.1
2.9
4.8
5.9
6.2
5.8
5.1
4.3
3.7
2.2
1.2
0.4
0.2

ti
2
2
2
2
2
2
2
2
2
4
6
8
10
10

ci ti
2.2
5.8
9.6
11.8
12.4
11.6
10.2
8.6
14.8
13.2
9.6
4
2
$
=115.8

Ei
0.0095
0.0250
0.0414
0.0509
0.0535
0.0500
0.0440
0.0371
0.0319
0.0190
0.0103
0.0034
0.0017

The mean residence time is thus 16.0 minutes.

tEi ti
0.038
0.200
0.497
0.815
1.071
1.202
1.233
1.188
2.300
2.736
2.487
1.382
0.864
$
= =16.0
i

6.3 Non-ideal reactor models

6.3

85

Non-ideal reactor models

The practical use of THE concept of residence time distribution, and thus of the the
E(t), is that it shows which reactor model should be used for a given reactor. For
example, if a tracer experiment results in an E(t) similar that of the CSTR, the CSTR
model can obviously be use with great confidence.
However, since most real-world open reactors fall between the CSTR and the
PFR with respect to mixing, there is a need for reactor models with other E(t)distributions. OnE such simple and useful model is the CSTR in series reactor model,
another is the segregation model.

6.3.1

The CSTR in series model

The E(t)-distribution of many real-worlD reactors may be modeled as a series of


connected CSTRS. The advantage of the CSTR in series model is that it leads to
simple calculations.
Figure 6.6 shows the E(t)-distributions for several examples of CSTR in series,
ranging from 2 to 30 CSTRs.
We recall that the E(t) for one ideal CSTR was a monotonous, declining function.
The interpretation of this is that the most common residence time for fluid elements
that enter the CSTR is, in fact, 0. This is quite easy to understand if one recalls that
a tracer added to a CSTR attains its highest concentration at t = 0, just after the
addition.
However, as soon as one has more than one CSTR in a series, the E(t)-distribution
will be bell shaped. Thus, the fluid elements that enter the CSTR in series system
will first mix into the first CSTR, then flow to the second CSTR before they can leave
the system at the lower boundary. Since this process takes time, the bell-shaped E(t)
is created.
As the number of CSTR in a series increases, the bell-shape becomes more pronounced. When the number reaches 30, almost all fluid elements have the same
residence time in the system. The extreme is, of course, the PFR which is made up
of a CSTR in series where each CSTR has a volume dV .
With respect to reactor calculations, the CSTR in series is convenient. For example, let us consider a 1st order reaction A B. For a CSTR, the output concentration
of THE reactant is (Equation E52.5):
cout,A = cin,A

1
1 + k

Let us introduce the notation that the cout,A from the first CSTR in a series is c1,A ,
the output from the second CSTR is c2,A , etc. In that case, the input concentration
to the second CSTR will be c1,A .
If we model a system with that a given has 2 CSTRs in a series, each CSTR will
have a mean residence time of /2. We then get:
c1,A

cin,A

c2,A

c1,A

1
1 + k/2

1
1 + k/2

6.3 Non-ideal reactor models

cin
Q

2.5

c1
Vtot/2
c
Q

1.0 ideal tank


n=2

0.5
0.0
0.0

2.5
c1
Vtot/5

0.5

1.0
1.5
2.0
2.5
tid/total medelupphllstid

ideal tub

1.5
1.0 ideal tank
c4
Vtot/5

c5
Vtot/5

c
Q

n=5

0.5
0.0
0.0

2.5
c2
Vtot/10

3.0

E(t)

2.0

c3
Vtot/5

c1
Vtot/10

ideal tub

1.5

c2
Vtot/5

cin
Q

E(t)

2.0

c2
Vtot/2

cin
Q

86

0.5

1.0
1.5
2.0
2.5
tid/total medelupphllstid

E(t)
ideal tub

2.0
c3
Vtot/10

c4
Vtot/10

3.0

1.5

c5
Vtot/10

1.0 ideal tank


c6
Vtot/19

cin
Q

c7
Vtot/10

c8
Vtot/10

c9
Vtot/10

c10
Vtot/10

c
Q

0.0
0.0

2.5
c1
Vtot/30

c2
Vtot/30

n=10

0.5
0.5

1.0
1.5
2.0
2.5
tid/total medelupphllstid

E(t)
ideal tub

2.0
c3
Vtot/30

c4
Vtot/30

3.0

1.5

c5
Vtot/30

1.0 ideal tank


c26
Vtot/30

c27
Vtot/30

c28
Vtot/30

c29
Vtot/30

c30
Vtot/30

c
Q

n=30

0.5
0.0
0.0

0.5

1.0
1.5
2.0
2.5
tid/total medelupphllstid

Figur 6.6: Example of E(t) for CSTR in series models.

3.0

6.4 The segregation model

87

or, since c2,A = cout,A ;


cout,A = cin,A (

1
)2
1 + k/2

In general terms, for a 1st order reaction that takes place in a reactor modeled as n
CSTR in series, the output concentration of reactant is:
3
4n
1
cout = cin
(6.15)
1 + k
n

Example 62: Series with an infinite number of CSTR

From Equation 6.15, it follows that the output concentration of the reactant is:
3
4
1
cout = cin
1 + k

From this expression, it is not obvious how cout will be expressed. However, one
dimensional mathematics teaches us that:
3
4x
1
lim
= ek
x
1 + xk
If we apply this boundary value to an infinite number of CSTRs in series we get:
3
4n
1
c = cin lim
= cin ek
n
1 + k
n
The output concentration from this CSTR in series is thus exactly the same as for a
PFR with the residence time . This is of course in full accordance with Equation
E55.8.

6.4

The segregation model

The segregation model is a method to calculate the properties and performance of


a reactor directly from the E(t)-distribution and the kinetic rate equation. The
name stems from the fact that one can view each package of input fluid as a small
volume that does not interact at all with the content of the reactor, nor with other
elements entering the reactor. These segregated packages will stay in the reactor
during different lengths of time. Hence, each package has a unique concentration of
reactants and products, dependent on their unique residence time.
Consequently the output concentrations equal the weighted average value of concentrations of the packages. The E(t)-distribution is the weighting factor used to
calculate this average.
Instead of talking about segregated elements, we can use the equivalent concept:
the batch reactor!

6.4 The segregation model

dV

dV

88

dV
dV
dV
dV

dV

dV
dV

dV

dV

dV
dV

dV

dV
dV

dV

dV

dV
dV
dV

dV

dV

dV

dV

dV

dV

dV

dV
dV

dV

dV
dV

dV

Figur 6.7: Illustration of the segregation model: a constant stream of small


batch reactors enter a larger volume where they stay according to a residence
time distribution given by E(t).

Figure 6.7 illustrates how the segregated elements, batches, add up to form the
reactor.
To put the segregation model into practice to calculate cout,A one needs to combine
three elements:
1. Kinetic equation for the reaction in which A is a reactant or product.
2. The mass balance for the segregated system.
3. The E(t)-the distribution of the reactor in question.
The first step is to combine the kinetic equation of mass balance for the batch
reactor. This gives, as before,
dc = rdt
For example, we have shown that for a 1st order reaction A B we get
cA = c0,A ekt
We must now recall that each tiny, segregated batch reactor has the concentration
cin,A at the moment that they enter the reactor. Thus, c0,A = cin,A . According to
the segregation model, the output from the actual, non-ideal reactor is calculated as:
2
cout,A = cin,A
ekt E(t)dt
(6.16)
0

6.4 The segregation model

89

Example 63: The segregation model applied to example 61


Problem
In the example 61 the E(t) for a pond was derived. In the pond, an organic pollutant
A is degraded according to an irreversible, 1st order reaction. The rate constant is
0.4 minutes1 .
Calculate:
1. The conversion of A in the pond given the E(t).
2. The conversion of A if the pond had worked as a CSTR.
3. The conversion of A if the pond had worked as a PFR.
Solution
According to Equation 6.16 we know that:
2
cA
=
ekt E(t)dt
cin,A
0
where E(t) is known. We can then make the following calculations.
cA,t
cA,t
i
t (min)
Ei
Ei i
cin
cin
0
0
1
1
2
0.0095 0.819
0.0156
2
4
0.0250 0.670
0.0336
3
6
0.0414 0.549
0.0454
4
8
0.0509 0.449
0.0457
5
10
0.0535 0.368
0.0394
6
12
0.0500 0.301
0.0301
7
14
0.0440 0.247
0.0217
8
16
0.0371 0.202
0.0149
9
18
0.0319 0.165
0.0211
10
24
0.0190 0.091
0.0103
11
30
0.0103 0.050
0.00413
12
40
0.0034 0.018
0.00063
13
50
0.0017 0.007
0.00011
$
= 0.283
i

For the CSTR, the conversion is:

c
1
1
=
=
= 0.385
cin
1 + k
1 + 0.116
For the PFR, the conversion is:
c
= ek = e0.116 = 0.202
cin
We see therefore that our non-ideal reactor falls in between the perfectly mixed CSTR
and thenon-mixed PFR. The answers are:

6.5 Other reactor models

90

1. The conversion of A is 1-0.283 = 71.7%.


2. The conversion of A would be 1-0.385 = 61.5% if the reactor were a CSTR.
3. The conversion of A would be 1-0.202 = 79.8% if the reactor were a PFR.

6.5

Other reactor models

In simple cases, one can often apply the CSTR in series model and the segregation
model. In natural systems, such as groundwater aquifers, however, one cannot simply
assume that the flow is one-dimensional. One then needs to use more sophisticated
reactor models.
One of the most general models is the dispersion model. In this model it is assumed
that mixing is analogous to a random motion. This allows us to take into account
flow patterns, chemical reactions and other phenomena such as adsorption.
In the following chapter, the dispersion model will be used as a basis calculating
how different substances are transported in porous media.

CHAPTER

Instationra CSTR
I Kapitel 5 behandlades de ideala reaktorerna sats, tank och tubreaktorn. Som vi
d sg representerade dessa reaktormodeller olika ytterligheter vad gller transport
av material ver systemgrnsen och omblandingsfrhllanden.
Tabell 7.1 ger en sammanstllning av ngra egenskaper hos de ideala reaktormodelellerna. Vad betrffar de dynamiska egenskaperna kan man notera att satsreaktorn
aldrig r stationra, med mindre n att ingen kemisk reaktion sker i den. Fr tubreaktorn gller att den kan vara instationra, men att det medfr att viktiga tillstns ssom koncentration av reaktant och produkt varierar i bde tid och rum. Instationra
tubreaktorer mste drfr modelleras med partiella differentialekvationer. Till detta
terkommer vi nsta kapitel.
Eftersom i stort sett alla tekniska och naturliga system r ppna och dynamiska
kan mnga intressanta frgestllningar belysas med hjlp av en eller flera instationra
CSTR. I detta avsnitt skall vi ta upp hur man kan anvnda instationra tankreaktorn
fr att beskriva ngra dynamiska system. Den arbetsmetodik som fresls exemplifieras i Exempel 71, och i vningsexempel.
Tabell 7.1: Sammanstllning ver viktiga karaktristika fr de ideala reaktormodellerna sats, tank och tubreaktorerna.

Transport ver
systemgrns
Makroskopisk
omblandning
Minsta homogena
volymselement
Stationr
materialbalans
Instationr
materialbalans
Matematisk form
p instationr
materialbalans

Satsreaktorn

Tankreaktorn

Tubreaktorn

Sluten

ppen

ppen

Omblandad

Omblandad

Ej omblandad

Hela volymen V

Hela volymen V

Differentiellt
volymselement dV

Aldrig stationr

In+Prod = Ut

In+Prod = Ut

Prod = Ack

In+Prod = Ut+Ack In+Prod = Ut+Ack

Kopplade ordinra
diff.ekvationer

Kopplade ordinra
diff.ekvationer

91

Kopplade partiella
diff.ekvationer

7.1 Svar p ndring i ingngskoncentration

7.1

92

Svar p ndring i ingngskoncentration

Fr att illustrera hur CSTR kan svara p ndringar i nyckelstorheter som ingngskoncentrationen av reaktant, skall vi betrakta en CSTR dr det sker en kemisk reaktion:
AB

r = kcA , k = 1

Systemet kan beskrivas med fljande figur:


[A]in,t0
Qin

[A] in,t0
[A]t0
[B]t0
V

Qin
[A]t0
[B]t0
Qut

[A]
[B]
V

[A]
[B ]
Qut

Som beskrivits i ekvation 52 och ekvation E52.5 kan koncentrationen cA i en stationr CSTR med en irreversibel 1:a ordningens reaktion berknas med ekvationen:
1
cA
=
cA,in
1 + k
I ovanstende system k blir allts cA = 0.5, eftersom cA,in = 1 och k = 1.
Hur reagerar d systemet om ingngskoncentrationen ndras? Figur 7.1 visar
ngra exempel p vad som kan hnda. Den vre bilden visar svaret d ingngskoncentrationen ndras till cA,in = 1.9 vid t = 2. Som vntat stiger cA till 0.95 efter en
tid. Man kallar strningen fr en stegstrningen och reaktorns svar p detta fr ett
stegsvar.
Nsta bild visar cA om ingngskoncentrationen tillts ka under en begrnsad tid.
Hr ser man hur koncentrationen i reaktorn (och i utfldet) kar fr att sedan sjunka
till ursprungsnivn. Detta r exempel p pulsstrning och pulssvar.
Det sista exemplet visar hur systemet reagerar p en sinusformad variation i ingngskoncentration. Man kan se att svaret blir att ven cA uppvisar ett sinusformat
beteende, men att det finns en amplitudskillnad och en fasfrskjutning mellan cA,in
och cA .
Dessa exempel r i detta fall bara illustrationer av olika dynamiska frlopp och
beteenden. Det finns dock mycket viktiga teorier fr hur systemens svar p olika
strningar kan anvndas fr att karaktrisera olika typer av system matematiskt.
Inom reglertekniken anvnds dessa teorier bland annat fr att underska systems
stabilitet, och fr att konstruera regulatorer som gr att variationerna i olika tillstnd
hller sig inom frutbestmda grnser.

7.1 Svar p ndring i ingngskoncentration

93

1.8
inkoncentration
1.6

koncentration

1.4

1.2

0.8
utkoncentration
0.6

0.4

0.2

5
tid

10

5
tid

10

5
tid

10

1.8

inkoncentration

1.6

koncentration

1.4

1.2

0.8
utkoncentration

0.6

0.4

0.2

1.8

1.6

koncentration

1.4

1.2

1
inkoncentration
0.8

0.6

0.4

utkoncentration

0.2

Figur 7.1: Exempel p hur en tankreaktor kan svara p en ndring i ingende


koncentration av reaktant.

7.2 Arbetsmetodik

7.2

94

Arbetsmetodik

Arbetsgngen fr reaktorberkningarna fr dynamiska system skiljer sig inte mycket


frn metodiken fr att hantera stationra system. Skillnaden ligger naturligtvis i att
man mste lsa de differentialekvationer som uppkommer eftersom Acktermen finns
med. Metodiken fr stationra reaktorsystem beskrivs p sidan 62.
Eftersom vi i detta avsnitt uteslutande skall anvnda numeriska lsningsmetoder
kommer begreppet simulera anvndas synonymt med ls differentialekvationerna.
De olika momententen r:
1. Beskriv blandningsfrhllandena i reaktorn med en reaktormodell och systemgrnser
2. Stll upp materialbalanser fr alla (oberoende) komponenter, och bestm de
reaktorspecifika parametrarna
3. Beskriv de kemiska reaktionerna med en kinetisk modell och hastighetskonstanter
4. Kombinera reaktormodell och kinetisk modell i form av en differentiell materialbalans
5. Ta fram initialvrden genom att lsa de stationra materialbalanserna, d v s
koncentrationerna av alla mnen innan strningen sker.
6. Som en kontroll av att du har satt upp alla ekvationerna korrekt, simulera de
instationra materialbalanserna med ingngskoncentrationerna som gller fre
strningen
7. Simulera det instationra systemet fr ndringen/strningen, d v s fr de nya
ingngskoncentrationerna
Det frsta nya momentet r att bestmma initialvrdena i systemet. I praktiken
r det dock exakt samma sak som att lsa den typ av stationra problem som beskrevs
i Kapitel 5. Innan man har rutin r det dock ltt att blanda samman begreppen
ingngskoncentration och initialvrde. Med initialvrde menas tillstnden i reaktorn
precis innan det gonblick d en strning sker.
Sedan man har bestmt initialvrdena kan man simulera systemet, det vill sga
lsa differentialekvationerna, med den urspungliga ingngskoncentrationen. Orsaken
till att man vill gra det r att det ger en koll p att man verkligen har stllt upp
sina differentialekvationer p rtt stt. Simuleringen skall givetvis ge som resultat att
de inga koncentrationer frndras relativt de stationra betingelserna. Om kurvorna
blir krokiga finns det helt enkelt ett tankefel eller ett programmeringsfel ngonstans!
Det sista momentet r att infra den strning som man vill studera och simulera
systemet. Om fregende moment har utfrts skall detta inte innebra ngra verraskningar, frutom de som ges av systemets svar p strningen.

7.2 Arbetsmetodik

95

Example 71: Exempel p simulering av instationr tankreaktor


Problemstllning
I en tankreaktor sker den frsta ordningens reaktionen:
AB

r = kcA

Normala driftsbetingelser r V = 5, Q = 1, k = 2 och cin,A = 100 (antag att enheterna


r harmoniserade....). Simulera vad som hnder om koncentrationen av A i infldet
till reaktorn frdubblas. Simulera fram till tiden 10.

Lsning
1: Reaktormodell och systemgrnser
Systemet beskrivs som en ideal tankreaktor, dr inoch utstrmmar passerar systemgrnsen:
[A]in,t0

[A] in,t0

Qin

Qin

[A]t0
[B]t0

[A]
[B]
V

[A]t0
[B]t0
Qut

[A]
[B ]
Qut

2: Materialbalanser
Materialbalanserna tecknar vi som Ut + Prod = In + Ack, och eftersom volymen r
konstant s gller:
In + Prod
A:

Qcin,A + rA V

B:

0 + rB V

= Ut + Ack
dcA
= QcA + V
dt
= QcB + V

dcB
dt

3: Kinetisk modell och hastighetsekvationer


Hastighetsuttrycken tecknas som:
r
rA
rB

= kcA
= kcA
= kcA

7.2 Arbetsmetodik

96

4: Kombination av materialbalanser och hastighetsekvationer


A:

dcA
Qcin,A
Q
=
cA ( + k)
dt
V
V

B:

dcB
Q
= kcA cB
dt
V

eller, i matrisform der = Ac + Y :


5 dc 6
A

der =
A=
c=
Y =

dt

dcB
dt
! Q
( V

+ k)
k

"
cA
cB
5 Qcin,A 6

0
VQ

"

5: Initialvrden
Initialvrdena fr man genom att lsa ekvationen
0 = Ac + Y
med fsolve. Funktionen: krs med:

7.2 Arbetsmetodik

97

6: Dynamisk simulering av stationra frhllanden


Det dynamiska systemet simuleras som
#
cA = 9.0909
der = Ac + Y
t=0
cB = 90.9091
Funktionsfilen blir nstan identisk med den som anvndes fr att ta fram initialvrdena. Man kollar frsta att man verkligen har stationra betingelser, och kr med
ode45(DynExempel,[0 10],[9.0909;90.9091]). Man fr fljande figur, som visar
att man har stationritet:
100

90
cb
80

koncentration

70

60

50

40

30

20
ca
10

6
tid

10

7.2 Arbetsmetodik

98

7: Dynamisk simulering av strningen


Efter ndring av inkoncentration i funktionsfilen: krs denna med ode45(DynExempel,[0
10],[9.0909;90.9091]). Det dynamiska frloppet blir:
180

160
cb

140

koncentration

120

100

80

60

40
ca
20

6
tid

10

CHAPTER

Dispersion in porous media


In previous sections we have treated both ideal and non-ideal reactor models. By using
the CSTR, the PFR as well as the tanks-in-series model, we were able to describe,
analyze and perform calculations for a variety of open systems.
The above reactor models are just - models! They have evolved to be useful in
practice, i.e. decent descriptions of reality suited for mathematical calculations. We
have thus been able to handle relatively complex chemical system with quite simple
mathematics.
In the following sections, we shall focus on another important class of concepts
and models that often are utilized in the environmental field. One concept is dispersion which we shall use to calculate solute transport in porous media. Dispersion
is a mixing model that mathematically is described in the same manner as diffusion. Accordingly, we hereby combine together the basic molecular mass transport
theories with environmental transport phenomena in the field scale. Applications include, among others, how organic pollutants move, sorb and degrade in groundwater
aquifers. This is a very important issue in society, and knowledge is needed in relation
to waste management, soil remediation and water conservation.
Figure 8.1 illustrates, in a simplistic fashion, how features related to substances
(pollution), soils (the reactor) and transport processes (flow) are linked. In mathematical model, the chemical and physical processes involved can be linked and treated
quantitatively.
This section begins with the fundamentals of flow in porous media such as groundwater aquifers. This leads in to the concept of dispersion. Then we take up how the
solutes, i.e. a dissolved substance and the solvent (in this case water) move together in
process called advection. The next section describes how solute transport is affected
by degradation and adsorption, introducing and using the important retardation factor.
It is inevitable to use partial differential equations in this context. The important
thing this context, however, is to understand what the equations represent, and what
boundary conditions account for. If one understands the fundamental equations and
boundary conditions, can easily perform amazing numerical model calculations using
software such as COMSOL (earlier called Femlab).

99

8.1 Water flow- Darcys law

100

organisk halt

R-vrde

adsorption
porositet

MARKEN

MNET

vattenhalt
Darcy-hast.

Grundvattenfrorening

VATTNET
Darcy
dispersion

lipofilicitet
Kow
nedbrytning
hast.koeff

advektion

BERKNING
kontinuitets
Femlab
ekvation

hast.
ekv

randvillkor
Figur 8.1: A simple mind map that illustrates how the properties of substances,
the soil and the hydraulic system interact to determine the rate of solute transport in a porous soil system.

8.1

Water flow- Darcys law

If possible, water flows from a point with higher pressure to a point of lower pressure.
This reduces water energy, since the point with the higher pressure represents a greater
energy content.
In systems that are only affected by gravitational forces, the potential energy of
the water increases with height but the energy is readily lost or transformed as the
water flows downwards. Porous soil systems largely follows this behavior, although
capillary forces, created by surface tension between soil particles and friction forces
complicate the picture. The measure of the actual energy difference between two
points is called hydraulic head.
In practice, the driving force for flow between two points is often proportional
to the difference in height. However, the force acting on a package of water in the
gravitational field depends on the distance over which it acts. The larger the distance,
the samller the driving force is to set the water in motion. Completely analogous to
the principles of mass transport (diffusion) and heat transport, we can assume that
the flow velocity is proportional to the hydraulic gradient:
vx

y
x

where v is the flow velocity in the x direction (e.g. m s1 ), as shown in figure 8.2.
The proportionality constant is called the hydraulic conductivity, Kx , which has
the same unit as the flow velocity. In 1856 the french engineer Darcy stated that:
vx = Kx

y
x

(8.1)

8.1 Water flow- Darcys law

101

grun
dvat
te

-y

nyta

vattendrag
eller brunn

x
Figur 8.2: The principles behind Darcys law - gravity makes water flow from a
state with higher potential energy to a point with lower energy.
Tabell 8.1: Exemples of hydrogeologic parameters

Fraction
Gravel
Sand
Silt
Clay

Typical
particlestorlek
1 mm
0.1 mm
0.02 mm
0.002 mm

Total
porosity
, %
25 35
30 45
35 45
40 55

Effective
porosity
e , %
25 35
25 40
20 35
2 10

Hydraulic
konductivity
K, cm s1
1 - 100
1104 1101
1106 1104
1109 1106

In practice, K is difficult to predict with reasonable accuracy. I table 8.1 gives


a summary of typical values for different types of soils. As we see is the hydraulic
conductivity is much higher for coarse grained than fine-grained soil. The ranges are
large, so in practice, the best values in fields are deterimend by pumping.
The hydraulic gradient can be measured by installing pressure sensors in the saturated groundwater zone. Another way is to install observation wells which gives a
direct measure of the ground water level along a stretch (trans-sect).

Darcy-velocity and real flow velocity


The velocity obtained with Darcys law must be interpreted as the flow Q through a
surface A perpendicular to the direction x during a certain time. Hence:
y
Qx
=
x
A
It is common to call the rate determined in this way for Darcy-velocity, or superficial
velocity. However, in a porous medium, a significant part of area does not carry any
water as it consists of solid material. This means that the real rate by which the water
moves in the porous medium must be greater than the Darcy-velocity. For example,
if the effective porocity is very low, the flow velocity in the fraction of the soil where
the transport takes place must exceed the average velocity in the soil.
The actual flow velocity ux is calculated as:
vx = Kx

ux =

Kx y

e x

(8.2)

8.1 Water flow- Darcys law

102

where e is the effective porosity. As shown in table 8.1. it may differ significantly
between the total and effective porosity. The difference is greatest for very fine-grained
materials, such as clay. This is because a large proportion of total porosity occupied
by immobile water that is very strongly tied to the earth by capillary forces. In the
fine pores, soil capillarity can be much larger than the gravitational force.

Example 81: Estimation of flow velocity


Problem
To determine the flow in an open groundwater aquifer, 3 observation wells were installed along a trans-sect of 400 m. The wells were thus at 200 m intervals. Groundwater levels in the test point was measured at 90, 89.25 and 88.5 m relative to the
sea level, respectively. Aquifer consisted of sandy silty moraine with a hydraulic conductivity of 2103 cm s1 . The effective porosity was 0.30.
Calculate:
1. Darcyvelocity
2. Flow velocity
3. The time it takes for water to be transported 400 m
Solution
The hydraulic gradient is constant along the trans-sect and amounts to:
y
88.5 90
1.5
=
=
= 3.75103 m m1
x
400 0
400

The unit of K is changed, so that:


2103 cm s1 = 2103

cm
m
s
m
102
606024365
= 630
s
cm
year
year

The Darcyvelocity can then be calculated as:


vx = Kx

y
m
= 6303.75103 = 2.36
x
year

The real flow velocity is:


ux =

vx
2.36
m
=
= 7.87
e
0.3
year

and the time it takes for water to be transported 400 m becomes:


t=

x
400
=
= 58 year!
ux
7.87

8.1 Water flow- Darcys law

103

lng vg
kort vg

friktion mot partikelytor

Figur 8.3: Illustration of mechanisms that causes a distribution of flow velocities


in porous media.

8.1.1

Flow rate distribution - reactor modeling

Darcys Law is a general equation that applies to every water package in the ground
at each time. But in practice it operates only in the macroscopic scale. Down on
microscopic scale,i.e. a scale equivalent to the size of soil particles and pores is a
very rough simplification. Each water parcel does in fact move at different velocity
at different times.
Figure 8.3 illustrates some reasons why the water in the ground is moving with
varying velocity.
Flow in the ground, as in all porous media, is a stochastic process. It means that
one cannot describe exactly how each parcel of water moves in detail. Will a drop turn
to the right or to the left around a soil particle? What happens when two packages
which have taken different paths around a particle meet in a por?
If one is only interested in total water flux, as represented by the Darcy-velocity,
the distribution of flow velocities in groundwater aquifer is of no importance. In
that case, the fact that the water packages have spent different amounts of time in
the ground, and founds their way forward in different ways does not matter if the
water, for example, is used for irrigation purposes. But if you are interested in water
chemistry, i.e. its quality, the distribution of flow velocities is of prime importance.
The reason is obviously to speed the distribution gives rise to a residence-time
distribution in the groundwater aquifer. As we previously discussed is the distribution
of residence times crucial for how the water chemistry develops in a system.
If we view at the groundwater aquifer in figure 8.2 as a reactor, then we realize
that the chemistry of the water leaving the groundwater aquifer is dependent on the
velocity distribution. One may also realize that none of the ideal reactor models, i.e.
the CSTR or the PFR constitute appropriate descriptions of a groundwater aquifer
as a reactor.

8.2 Dispersion

104

In the following sections, we will therefore develop a general mathematical model


for simultaneous flow and transport of chemicals in groundwater moving in a porous
medium.

8.1.2

Advective transport

In previous sections we have dealt with two types of open chemical reactors; the
CSTR and the PFR reactors. We have confined the discussion to situations where
the transport of substances into and from the reactors have been a part of a liquid
stream. in the following, this liquid stream will be called the bulk.
The transport of the bulk into the reactors is caused by a pressure that is exerted on
the liquid. We call this process convective transport, normally distinguishing between
forced convection and natural convection. The forced convection can be driven by a
pump or fan. The natural convection can be driven by gravitational forces and/or
density-induced boyant forces.
For the transport of substances that follow a convective flow, we will use the term
advection. The concept is almost exclusively used for natural systems. It applies to
substance transport in groundwater, but also to the transport of such as water vapor
and pollutants in the atmosphere.
The mass balances for the CSTR and the PFR includes Input and Output terms
on the form Qc (e.g., mol s1 ). In these terms Q the convective flow of the bulk,
while Qc represents the advective transport of dissolved substances.

8.2

Dispersion

8.2.1

Molecular diffusion - Ficks law

The molecular diffusion of a a substance follows Ficks law. In its original form, Ficks
law is based on the assumption that the probability that an individual molecule
of a substance is identical in all directions. Each molecule does is therefore move
completely random, independent of the concentration, in any directions.
However, this reasoning leads to the conclusion that the probability is greater that
more molecules will move from a volume with many molecules to a volume with fewer
molecules, than in the opposite direction. We then conclude that there must be a
net flow of molecules from higher concentrations to lower concentrations. This means
that the net flow goes in the direction opposite to the concentration gradient.
This can be formulated as Ficks Law, which says that, in the direction of L the
flux J will be proportional to the concentration gradient:
J = D

dc
dL

mass m2 s1

(8.3)

For molecular diffusion, the proportionality constant is the diffusion rate coefficient
D (e.g. m2 s).
It is obvious that the flux into a control volume can affect the concentration of a
substance in the volume. If the input flux equals the output flux, the concentration
in the volume will not change . We can, of course, conclude that the change in concentration per unit time in a control volume must depends on the difference between
the input and the output fluxes.

8.2 Dispersion

105

koncentrationsgradient

dL

diffusionsriktning

c+dc
L-riktning

Figur 8.4: Application of Ficks law on a control volume.

Let us now establish a mass balance for the control volume shown in Figure 8.4.
Since the flux is specified per unit area, and volume of the control volume can be
written dV = AdL we get:
Ack

In Ut

c
t

AJin AJut

c
t

AD

AL

c
t

AD

c
t

8.2.2

c
(c + c)
A(D
)
L
L

2c
L

2c
L2

(8.4)

Analogy dispersion - diffusion

The irregular flow in porous media (Figure 8.3 leads, as said, to a situation where
certain parcel of water move faster than others. In practice it means that Darcyvelocity is an average of a distribution of velocities.
Also, substances that are transported with the bulk through advective transport
will move according to a distribution of velocities. Therefore, if one adds a pulse of
tracer into a bulk flow in a porous medium, one finds that the trace element spreads
out during the transport. This phenomenon is called dispersion.
Dispersion is a mixing phenomenon, which occurs in all porous media. It can not
be described exactly, just because dispersion is a random phenomenon. A given water
parcel can slow down more times than others along a specific route, or go faster.

8.2 Dispersion

106

Dispersion is thus a process that occurs in the macroscopic scale, cm and meters,
while diffusion takes place in a scale several orders of magnitude smaller.
Since dispersion is a stochastic process, we can describe it with the same equations
and mathematics as we use to describe diffusion. Thus, we can apply Ficks law also
for dispersion. The only difference is that we replace the physical diffusion constant
D (e.g. m2 s1 ) with the empirical dispersion coefficient E.

8.2.3

Deduction of expression for spread

Diffusion (and dispersion) give rise to spread of substances from a point with a high
concentration to points with lower concentration. The concentration distributions
caused by these processes can that diffusion can be described in terms of variance and
standard deviation realtive a mean value.
In the following sections, general measures of the spread of a substance in bulk
as a result of diffusion are derived. Thereafter, the same results and reasoning are
applied to dispersion.
Let us the apply equation 8.4 to calculate the spread when a tracer is added as
a pulse at a point in space. If the amount M is added in the point L = 0 , the
concentration in that point becomes c(L = 0, t = 0) = M/ where infinitesimal
slice in space.
The diffusion equation is a partial differential equation with respect to time and
space, and must be solved applying an appropriate set of boundary conditions. Since
the equation is of 1st order with respect of time, and 2nd order with respect of space,
3 conditions are necessary:

Boundary condition I c(, t) = 0

2c
c
c
=D 2
(8.5)
(0, t) = 0
Boundary condition II

t
L
L

Initial value I c(0, 0) = M/


c
The condition L
(0, t) = 0 expresses the fact that the curve is symmetrical around
the maximum value. The solution is:
M
L2 /4Dt
c(L, t) =
(8.6)
1 e
(4Dt) 2

The spread around the symmetrical axis at x = 0 is given by the variance, 2 . It is


defined as the weighted deviation from x = 0. As the spread changes with time, the
variance is also time dependent:
2 +
L2 c(L, t)dL

2
(t) = 2 +
(8.7)
c(L, t)dL

We shall now consider the numerator and denominator separately.

The numerator We can now develop the numerator by inserting the solution of
the differential equation: equation 8.7. This gives:
2 +
2 +
M
2
L2 /4Dt dL
Numerator =
L c(L, t)dL =
L2
1 e
(4Dt) 2

8.2 Dispersion

107

We can solve this integral identifying that it is on the form:

y=L

2 +
1

2
a=
b
y 2 eay dy where
4Dt

M a

b =

This is a standard integral which (for positive values of a) has the solution:

2 +
2

b
y 2 eay dy = 2b
4a
a

so that:

1
4Dt
Numerator = 2b = 2M = 2M
= 2M
= 2M Dt
4a
4
4a a
4a a
Denominator The denominator can be treated in about the same way. The integral
from to + may, because of symmetry, be divided into the two segments 0
and 0 . Inserting the solution to the diffusion equation into the expression for
the denomimator we get:
Denominator =
2 +
2
c(L, t)dL =

M
(4Dt)

1
2

2
eL /4Dt dL = 2

M
(4Dt)

1
2

2
eL /4Dt dL

In this case too, we can find a standard integral:

y=L

2 +
1

2
a=
b
eay dy where
4Dt

7
0

b = M

7
b
The solution is
and for the whole interval + we get:
2 a
7
7 7
2b
a

Denominator =
=M

=M
2 a

a
Variance and standard deviation Finally, we can simply divide the numerator
by the denominator to get:
2 =

2M DT
= 2Dt
M

and the standard deviation:

= 2Dt

8.2 Dispersion

108

Koncentration av mne

=0

= 0.25
= 0.50
=1
-2

-1,5

-1

-0,5

0,5

1,5

Lngdskala (t ex meter)

Figur 8.5: Development of the concentration profile as a substance spreads by


means of diffusion in a stagnant medium.

8.2.4

Interpretation of the standard deviation

Spreading of a substance through diffusive and dispersive processes can thus be expressed in terms of standard deviation .
Figure 8.5 shows how the concentration profile develops in one (1) dimension as
caused by diffusion, i.e. in a system without any convective flow. For example,
if a tracer at the time t = 0 is contained in a point in space, it will diffuse out
symmetrically from the point. The concentration distribution of the substance will
always be always correspond to a
normal distribution1 relative to L = 0, and the
standard deviation is given by = 2Dt.
Since the substance concentration follows a normal distribution, it is true that:
68% of the substance is contained within the distance on both sides of L = 0,
i.e. the width 2
95% of the substance is contained within the distance 2, i.e. the width 4
98% of the substance is contained within the distance 4, i.e. the width 8

I Figure 8.5, we see how a substance spreads out from a point, i.e. how the
substance is distributed after a certain time.
From the other point of view, because:

= 2Dt
(8.8)
the time it takes for a certain standard deviation to develop is:
t=
1 The

2
2D

distribution is given by

1
2eL/

8.2 Dispersion

109

Koncentration av mne

=0

L= 2EL/u
=2E4L/u=2L

0 L

=2E16L/u=4L

4L

16L
Transportstrcka L=ut (t ex meter)

Figur 8.6: Illustration of how the concentration profile develops during coupled
advection and dispersion.

8.2.5

Application to a advection-dispersion system

During flow in a porous medium, the substance moves both by advective transport,
and by random diffusion and dispersion. Some molecules of the substance move faster
than the average flow velocity, and just as many move slower than the average flow
velocity. Therefore, around the top of concentration profile, the concentration develops in principle in the same manner during advective as molecules that during diffuse
from a point. Thus, the standard deviation does characterize the concentration
profile in a advection-dispersion system just as in a stagnant system with diffusion.
The difference is that dispersion creates much greater spread than diffusion, i.e.
grows faster because of dispersion than because of diffusion. Figure 8.6 shows how a
substance spreads by means of dispersion as it is transported by means of advection.
It is vital to realize that the the concentration profile is centered (i.e. the peak
is localized) at the mean transport distance covered during a certain period of time.
Thus, at a certain time t the peak will always be at:
L = uL t
Combining equations 8.9 and 8.8 we see that when the peak is at L, L is:
8
L = 2EL/uL

(8.9)

That means that as time (and


thus the transport distance) doubles, the standard deviation increases by a factor 2. Since the concentration follows a normal distribution,
8
68% of the substance will be found within the distance represented by 2 2EL/uL
when
8 the peak is at L. When the peak is at 16L, 95% of the substance will be within
2 162EL/uL from L, i.e. a 4 times wider distance.

8.3 Dispersion coefficients in groundwater aquifers

110

Example 82: Diffusive and dispersive spread


Problem
In the example ?? we considered water traveling along a 400 m long stretch in
a groundwater aquifer. The trip was estimated to take 58 years. Also, because of
diffusion and/or dispersion a pulse of a tracer introduced into the flow will be spread
out during transport.
Calculate the width of the pulse after 8 years (i.e. that it has travelled 400 m) if
the width is defined as the distance that contains 95% of the pulse.
1. The spread is diffusive with D = 1109 m2 s1
2. The spread is dispersive with E = 1106 m2 s1
Solution
The spread can be
quantified in
terms of the standard deviation , which can be
calculated as = 2Dt or = 2Et. 95% of the pulse is contained within 22.
We can then calculate the width of the pulse (in meters) as:
8

4dif f
= 4 2Dt
= 4 2(1109 360024365)58 = 7.6 m
8

4disp
= 4 2Et
= 4 2(1106 360024365)58 = 242 m

(E82.10)
(E82.11)

We can thus conclude that dispersion is the process responsible for the spread, simply
because the dispersion coefficient is greater than the diffusion coefficient.

8.3

Dispersion coefficients in groundwater aquifers

It is difficult to obtain precise values of the dispersion coefficient under field conditions.
An indication of this can already be found in table 8.1. There we can see that the
hydraulic conductivity is a parameter with high variability, e.g. due to the fact
that the soil is heterogeneous. Since dispersion is caused by irregularities in the soil
structure, it should also be expected that dispersion coefficients are highly variable.
In many situations though, one does not need to know the dispersion coefficient
exactly. If one for example should carry out a risk assessments, it may be sufficient
to know what is worst case. If one has reliable groundwater data it is, however, in
many cases possible to use these to estimate the E.

8.3.1

Dispersivitet

The dispersion phenomenon is strongly linked to the flow velocity. The higher the
flow velocity is, the higher the E-value becomes. If one assumes that E is proportional
to uL , one can define a proportionality constant, the dispersivity :
EL = uL
The dispersivity has the unit meter, or any other unit of length.

(8.10)

8.3 Dispersion coefficients in groundwater aquifers

111

Tabell 8.2: Empirical values of dispersivity in soils (after Schnoor 1996)

Scale (m)
Lab studies
Field studies
Field studies

<1
1-10
10-100

Longitudinal dispersivity (m)


Typical
Interval
0.0010.01
0.00010.01
0.11.0
0.0011.0
25
1-100

Figur 8.7: Empirically determined values of the dispersivity in groundwater


aquifers (Logan, 2000).

Empirical values 1 Table 8.2 shows example values of the dispersivity. As the
table shows, the dispersivity is scale dependent. The greater the distance of travel,
the higher value does the dispersivity attain. The reason for this is that the larger
the distance of transport is, the greater the chance that the water encounters heterogeneous zones. It may be impervious lerlinser, permeable deposits and similar.
Empirical values 2 Another data set has been published by Logan (2000). here as
well, the dispersivity is dependent on the spatial scale. Uncertainties are significant,
and one must not forget that the axes are in log - scale.

8.3 Dispersion coefficients in groundwater aquifers

112

Empirical correlation In the university text book Applied Hydrogeology one finds
the following correlation:
= 0.0175L1.45

(8.11)

where L should be given in meter. This correlation gives similar results as the empirical values described above.
Pe-correlation Yet another possibility is to use correlations that uses a dimensionless property, the Pe-number (after Pclet).
The Pe-number is defined as the ratio between a term that reflects the advective
transport, divided by the dispersive mixing.
Pe =

uL L
E

In the literature, one finds empirical equations such as:


E
= 8.8Pe1.17 ,
D

for Pe > 0.5

from which it is possible to calculate E by iteration. The usefulness of such equations


varies, but in any case it is possible to get an idea of the magnitude of the dispersion
coefficient. That is, provided that the systems from which they have been derived are
similar to the system of application.

CHAPTER

Transport in porous media


Modeling and calculation of transport in porous media is challenging. There are a
variety of phases, compounds and processes to keep track of, and no one can be
described precisely. But even if you have good knowledge of individual processes
and the system as a whole, the problem often remains of finding the right parameter
values. It may be kinetic rate and chemical equilibrium constants, values of hydraulic
conductivity and dispersion coefficients.
Despite all the limitations, one often has to be able to estimate rates of transport
and concentrations of chemical substances in soil. In this chapter, we go through the
basics of coupled advection, dispersion, adsorption and chemical reaction in the soil
and in the water phase, based on the continuity equation. Table 9.1 gives an overview
over the cases treated in this chapter.

9.1

The continuity equation - advection and dispersion

In the previous chapter, we used the continuity equation to study the dispersion. We
regarded dispersion as a fairly isolated incident and did not link it mathematically
with the bulk transport, i.e. the advection. here, this will be done, producing an
expression for how the concentration in an aquifer with time increases if a substance
ss released from a continuous source, a pulse or a step-increase.
Tabell 9.1: Summary of cases treated in this chapter.

Source
Continuous
Pulse
Continuous
Step
Continuous
Continuous
Pulse

Processes considered
TimeEquation
Advection Dispersion Reaction Adsorption variability

Dynamic
9.5

Dynamic
9.7

Dynamic
9.11

Dynamic
9.16

Stead-state
9.18

I Dynamic E92.29

Dynamic
9.28
113

9.1 The continuity equation - advection and dispersion

dL

114

t = litet

t = strre

t = nnu strre

L=0
c = c0

L=
c=0

strmningsriktning

Figur 9.1: Illustration of transport by means of advection and dispersion.

9.1.1

Continuous source

Figure 9.1 illustrates a case where a substance moves from left to right. At the lefthand boundary L = 0 the concentration is constant, cin . At the right-hand boundary
L = the concentration is 0 as long as t < .
Equation 8.4 showed that, unless there a chemical reaction takes place, the change
in concentration within the control volume AdL depends on the difference in flux J
between the two boundaries:
AdL

c
= AdJ
t

(9.1)

In terms of the two processes advection and dispersion, the flux J becomes:
AJ = advection term + dispersion term = A(uL c E

c
)
L

(9.2)

This equation directly gives:


c
A J

c
c
2c
=
=
(uL c E
) = uL
+E 2
t
A L
L
L
L
L

(9.3)

The negative sign in front of the advection term makes sense; if the concentration
c
c
decreases in the L-direction,
will be negative, and uL
will be positive. This
L
L

9.1 The continuity equation - advection and dispersion

115

corresponds to a situation where the advective transport adds substance to the control
c
volume, causing a positive
.
t
Equation 9.3 should be solved with three conditions:

Boundary condition I c(0, t) = cin

c
c
2c
c
= uL
+E 2
(9.4)
Boundary condition II
(, t) = 0

t
L
L
L

Initial value I c(L, 0) = 0, L 0


The equation has the complete solution:
!
"
cin
L uL t
L + uL t
u
L/E
L
c(L, t) =
erfc(
)+e
erfc(
)
2
4Et
4Et

(9.5)

However, with an error of a few percent, it can be simplified to:


c(L, t) =

cin
L uL t
erfc(
)
2
4Et

(9.6)

The simplified solution makes it obvious that at L = uL t (i.e. at the mean convective
travel distance), the concentration is always c = cin /2! Often the development of the
concentration, as illustrated in example 91, is seen as a front that moves forward in
the flow direction. The curve is often referred to as a breakthrough curve.

Pulse
What happens then if a substance added to a system as a single event, a pulse. In
this case, we have the situation described in figure 8.6. The results is that the pulse
is transported while it is spreading due to dispersion.
One can easily produce an equation for the process, from equation ??. The equation gives c(L, t) for a system without advection. With advection, the shape of the
concentration curve will be the same after a certain time, but the whole curve has
moved along the direction of advective transport. We can view this as if the entire
coordinate system is moved. After time t, the whole pulse has moved a distance uL t
from the starting point at L = 0. We can now modify equation 8.6 by replacing the
spatial coordinate L with L uL t to obtain an expression for the concentration c at
any point L and at any time t:
c(L, t) =

M
(4Et)

1
2

2
e(L uL t) /4Et

as a solution to:
c
c
2c
= uL
+E 2
t
L
L

Boundary condition I c(, t) = 0

c
Boundary condition II
(uL t, t) = 0
L

Initial value I c(0, 0) = M/

(9.7)

Note that (L uL t)2 always is positive, hence the c is symmetrical around L = uL t.


How should then the amount M be interpreted? M is the amount that is added
per m2 at a certain instant. The follows from the fact that the left hand side has the
1
unit mass per unit volume (e.g. g m3 ) while the denominator (4Et) 2 has the unit
length (e.g. m).

9.2 Advection, dispersion and reaction

116

Example 91: Infiltration of a pollutant into a soil


Problem
A non-reactive, water soluble substance, dissolved in 0.5 m3 per m2 and year, is
continuously added onto a water-saturated soil. If the effective soil porosity 0.35, and
the dispersivity 0.6 m, how will the front move down towards 10 meter depth during
5 years.
Solution
A solution can be obtained by evaluation of equation 9.5 for different times. The
parameters become:
=

uL

0.5
m
= 1.43
0.35
years

= uL = 1.430.6 = 0.86 m

The equation is defined in a MATLAB file: which is run with:


>>fplot(exempel11_1,[0 10])
and so on. The result becomes:
1

0.9

Koncentration (andel av inkoncentration)

0.8
tid=5.0 r

0.7

0.6

0.5
tid=2.0 r
0.4
tid=1.0 r
0.3
tid=0.5 r
0.2

0.1

9.2

5
6
Lngdkoordinat

10

Advection, dispersion and reaction

In earlier chapters, the fundamentals of time-variant (dynamic) and steady-state


chemical reactors have been treated, based on mass balance equations. For example, equation 5.13 was used to show that for a closed system, the time-derivative of
the concentration can be expressed as:
c
=r
t

(9.8)

9.2 Advection, dispersion and reaction

117

where r is one (or more) kinetic rate expression(s).

9.2.1

Continuous source

In order to model transport of a substance coming from a continuous source, equation


9.3 just need be adjusted to comprise the chemical reaction:
c
c
2c
= uL
+E 2 +r
t
L
L

(9.9)

In a case where the substance is degraded according to a 1st order reaction, for which
r = kc, the continuity equation becomes:

Boundary condition I c(0, t) = cin

c
c
2c
c
= uL
+ E 2 kc
(9.10)
Boundary condition II
(, t) = 0

t
L
L
L

Initial value I c(L, 0) = 0, L 0


The complete, quite complicated, analytic solution is:
!
"
uL L(1+)
cin uL L(1)
L uL t
L + uL t
2E
2E
c(L, t) =
e
erfc(
)+e
erfc(
)
2
4Et
4Et
where:

9
(1 + 4kE/u2L )

(9.11)

(9.12)

Figure 9.2 shows a calculation corresponding to example 91, with the difference that
the chemical reaction is included.. The time considered is 5 years. Note that for
k = 0, the results are identical to the correspond calculation in example 91.
Normally, one can disregard the second term. Then c is:
cin
c(L, t) =
e
2

uL L(1 )
L uL t
2E
erfc(
)
4Et

(9.13)

If 4kE/u2L < 0.0025 then (1 + 2kE/u2L ), and equation 9.13 becomes:


c(L, t) =

9.2.2

L uL t(1 + 2kE/u2L )
cin kL/uL

erfc(
e
)
2
4Et

(9.14)

Continuous addition during a limited time - step up and


down

It is also interesting to consider a case where a pollutant is added to a system during


a confined period of time. In practice, this could occur if a spill take place from a
certain time (t = 0) and is prevented at t = tf . In this case the model becomes:

c
c
2c
= uL
+ E 2 kc
t
L
L

Boundary condition I c(0, 0 t tf ) = cin

Boundary condition II c(0, t > tf ) = 0


c

Boundary condition III


(, t) = 0

Initial value I c(L, 0) = 0, L 0

9.2 Advection, dispersion and reaction

118

0.9

Koncentration (andel av inkoncentration)

0.8

0.7

0.6

k=0.2 /r

k=0.1 /r

k=0.0 /r

k=0.5 /r
0.5

0.4

0.3

0.2

0.1

5
6
Lngdkoordinat

10

Figur 9.2: Calculation with equation 9.11 for the time t = 5 years.

(9.15)
A simplified [sic!] form of the analytical mathematical solution is:
c(L, t) =
kL !
L uL t(1 +
cin

e uL erfc(
2
4Et

(9.16)
2kE
)
u2L

L uL (t tf )(1 +

) erfc(
4Et

2kE
)
u2L

"
)

Now, let us concider the extreme cases. If the pulse is infinitely short, i.e. tf 0,
then the two erfc-terms will be equal and cancel out, causing c = 0 for all tf . On the
other hand, if tf than we have a continuous source. In that case the second
term is erfc() = 0. That leaves us with the first term, and the final expression for
c(L, t) is identical to equation 9.14!

9.2.3

Steadystate

If there is no chemical reaction in a system that is continuously replenished, the


concentration will eventually be equal to the input concentration to the system. In a
system of chemical reaction, however, a steadystate situation will develop, where c
is lower than cin .
Figure 9.3 shows how the concentration profile develops for a reference case (E =
1.43, uL = 0.86, k = 0.2). we see that when t , the concentration will approach
a steadyvalue along the entire reactor, just as in the ideal PFR.
Now let us consider what happens if the dispersion coefficient E is large or small.
The most simple case is obviously E = 0.If we do not have any mixing in the direction

9.3 Adsorption and transport

119

0.9

Koncentration (andel av inkoncentration)

0.8

0.7

0.6

0.5

tiden = ondlig=stationritet

0.4

0.3

tiden = 5 r
tiden = 2 r

0.2
tiden = 1 r
0.1

5
6
Lngdkoordinat

10

Figur 9.3: Calculation for different equation 9.11 values on t.

of flow we get:
lim e

uL L(1

(1+4kE/u2 )
L

2E

lim erfc(

LuL t

(1+4kE/u2L )

)
4E

= ekL/uL

for E 0

= erfc() = 2

for E 0

(9.17)

Furthermore, if we compare the spatial dimension (L) with the temporal dimension
(t), med tidsskalan, we can again use that the time it takes for the water to move L
is t = L/uL . If t = L/uL is substituted into equation 9.13 we can conclude that:
c(L, t) = 2
t

cin kt
e
= cin ekt
2

for E 0

(9.18)

In other words: when the dispersion is negligible, this rather complicated model is
reduced to the simple PFR model. This result had also been possible to obtain by
removing the dispersion term from equation 9.9.

9.3

Adsorption and transport

In many cases, the substances that are transported through the soil are sorbed by

soil particles of different kinds. As a rule, sorbtion anions (e.g., NO


3 , Cl ) is very
weak, while metal as well as neutral organic molecules can bind very hard. Metal
ions sorb mainly on negatively charged surfaces of organic carbon in the soil and in
crystal lattice of clay minerals. Organic molecules, however, also binds to soil organic

9.3 Adsorption and transport

120

matter. This soil organic carbon, SOM, found in soils is a non-degradable residue of
litter and root biomass.
The content of organic material can vary from a few percent in a sandy mineral
soil to nearly 100 % in an organogenic soils such as peat soils.
The sorption reactions are quite fast in relation to the transport processes. It
is therefore reasonable to model sorption processes as partitioning equilibrium reactions1 , involving the substance in dissolved form and adsorbed to some type of solid
compound.
Classic physical chemistry suggests a number of adsorption isotherms, with different underlying mechanistic assumptions. In practice, however, one has to assess
adsorption isotherms empirically.
A common form of adsorption isotherm is the Freundlich isotherm. It is based
on the assumption that the amount of adsorbed substances is proportional to the
concentration of the substance in the soil solution. If the relationship involves direct
proportionality, we get:
S = Kd c

(9.19)

where S is the amount of sorbing solid substance (sorbate) in the soil, expressed in
unit such as mg kg1 . Since the concentration is given in mass of substance per
unit mass of soil, the partitioning constant Kd will have a unit of the type m3 kg1 .
However, if we want to express how much sorbed substance that is present in the soil
in relation to the amount of water in the soil, we have to take into account:
soil
)
1. soil bulk density ( kg
m3
soil

2. soil water content, e.g. the effective porosity (

m3water
)
m3soil

in the saturated zone

We can now form the relationships:


amount sorbed
S
Kd c
=
=
volume of water

(9.20)

This means that the effect of a change in sorbed amount within a unit volume of soil
water can be expressed as:
d S
d Kd c
Kd dc
=
=
dt
dt
dt

1 http://en.wikipedia.org/wiki/Partition

(9.21)

coefficient

9.3 Adsorption and transport

9.3.1

121

Advection, dispersion, reaction and adsorption

Since adsorption affects the transport of a substance in an advective system, this


must be represented in continuity equation. If the adsorbed amount of increases this
will cause a decrease in the content of substance in the water. And, conversely, if the
substance sorbed to the soil particles are released, this will increase the content of
substance in the water accordingly.
A model which takes into account this must be:
c
c
2c
S
= uL
+ E 2 kc
t
L
L
t

(9.22)

or, withe the help of equation 9.21:


c
c
2c
Kd c
= uL
+ E 2 kc
t
L
L
t

(9.23)

Kd c
c
2c
)
= uL
+ E 2 kc
t
L
L

(9.24)

(1 +

(9.25)
We have now deduced the so called retardation factor, R. This important quantity is
defined as:
R = (1 +

Kd
)

(9.26)

R can be substituted into the continuity equation as:


c
uL c
E 2c
k
=
+
c
t
R L R L2
R

(9.27)

The reason why R is called retardation factor is obvious. If there is an adsorption


process, then R > 1 and the adsorption process will slow down all other processes,
including the advection, the dispersion as well as the reaction!
Pulse
In the case a pulse of substance is added we get, in analogy with equation 9.7, that:

c(L, t) =

M
1

(4Et/R) 2

(L uL t/R)2
4Et/R

(9.28)

9.3 Adsorption and transport

122

Example 92: Complex pollutant transport in a soil


Problem
A reactive, partly water soluble organic substance, dissolved in 0.5 m3 per m2 and
year, is continuously added onto a water-saturated soil. If the effective soil bulk
density 1.5 kg L1 , how will the front move down towards 10 meter depth during
5 years, provided that the substance is degraded according to a 1st order reaction
1
k = 0.1 years
. The substance is adsorbed to the soil organic matter (Kd = 0.6 L kg1 ),
and other relevant data can be found in example 91.
Solution
We can use equation 9.11, where the parameters uL , E and k are scaled with R.
c(L, t) =
!
"
uL L(1+)
cin uL L(1)
L uL t/R
L + uL t/R
e 2E erfc( 8
) + e 2E erfc( 8
)
2
4Et/R
4Et/R

Note that in the exponent uL L(1)


as well as in =
2E
from both the numerator and the denominator.
We calculate R as:
R = (1 +

(E92.29)

8
(1 + 4kE/u2L ) R disappears

Kd
0.61.5
) = (1 +
) = (1 + 2.53) = 3.43

0.35

Applying the above equations and data gives:


1

0.9
dispersion

lite dispersion

Koncentration (andel av inkoncentration)

0.8

0.7
dispersion+reaktion
0.6

0.5

0.4

0.3
dispersion+reaktion+adsorption

0.2

0.1

5
6
Lngdkoordinat

10

Something strange to think about: shouldnt the curves for only dispersion cross at
c/cin =0.5?

9.4 The continuity equation in multiple dimensions

9.3.2

123

Determination of Kd and R for organic substances

The tendency of different substances to partition between water and organic substances vary over a very wide range, as does the partitioning coefficient Kd . As
indicated earlier, Kd quantifies the partitioning between an organic substance partly
in a water phase, partly sorbed to soil organic matter. Kd is defined as:
Kd =

S m3 water
c kg soil

It is quite logical that Kd depends on (at least) to factors; the capacity of the soil
organic matter to sorb organic substance, as well of the sorption properties of the
substance itself.
If we limit the discussion to organic substances, then the adsorption is soil organic
carbon, not the crystalline soil material, SOM. The carbon content in the soil is
normally denoted foc , and the coefficient for partitioning of an organic substance
between water and the SOM is Koc . This allows us to write:
Kd = Koc

kg SOM
m3 water
m3 water
foc
= Koc foc
kg SOM
kg soil
kg soil

The magnitude of the partitioning coefficient Koc depends on the properties of


substance itself, especially charge, molecular weight and polarity. These properties
combined determine the lipophilicity of the substance in question. The lipophilicity
can be determined experimentally by measuring how the substance in question partition between n-octanol and water. The partitioning coefficient for this three component system (substance, n-octanol, water) is normally denoted Kow . Reliable Kow
-values can be found in the literature2 , often as log Kow A lipophilic substance like
benzene has a log Kow around 3 (Kow = 1000), while extremely lipophilic substances
such has DDT, PCB and PAH all have log Kow 6.
Table 9.2 shows useful correlations that can be used to estimate Koc from Kow
together with Ko w values.

9.4

The continuity equation in multiple dimensions

In this chapter, the discussion and calculations have been limited to transport and
dispersion in one dimension, denoted L. In real system, flow and dispersion occur
in 3 dimensions. Under the conditions that advective transport takes place in the
dimension x and y, and dispersion takes place in all direction x, y and z, equation
9.27 may be generalized as:
c
c
c
2c
2c
2c
= ux
uy
+ E x 2 + E y 2 + Ez 2 r
t
L
L
L
L
L
Thus can be written in terms of operators on the general form:
R

=u
c + Ec
r
t
Referring to the notation in figure 8.2, it is always the case that:
R

Ex >> Ey > Ez
2 http://logkow.cisti.nrc.ca/logkow/

(9.29)

(9.30)

9.4 The continuity equation in multiple dimensions

124

Tabell 9.2: The following correlations have been used:


Chlorinated substances: log Koc = 0.72 log Kow + 0.5
Aromatics: log Koc = log Kow 0.21
N B: When these values are used to calculate the retardation factor R, the soil
bulk density must be given in kg/L.

mne
log Kow
Chlorinated substances:
Bromoform
2.30
Carbontetrachloride
2.64
Chlorethane
1.49
Chloroform
1.97
Chloromethane
0.95
Dichloromethane
1.6
Hexachloroethane
3.34
Tetrachloroethylene
2.88
Trichloroethylene
2.29
Vinylchlorid
0.60
1,1-Dihlorethane
1.80
1,2-Dichloroethylene
1.48
1,1,1-Dichlorethane
2.51
1,1,2-Dichlorethane
2.07
Aromatics:
Benzene
Benso(a)pyrene
Chlorobenzene
Ethylbenzene
Hexachlorobenzene
Naftalene
Nitrobenzene
Pentachlorophenole
Phenole
Toluene
Dichlorbenzene
2,4,5,2,4,5-PCB
Pesticides:
Atrazine
Dieldrine
DDT
2,4-D (fenoxiacid)

log Koc

Solubility (mg/L)

2.16
2.40
1.57
1.92
1.18
1.41
2.81
2.57
2.15
0.93
1.80
1.57
2.30
1.99

800
5 740
8 200
6 450
20 000
50
200
1 100
90
400
8 000
4 400
4 500

2.13
6.06
2.84
3.34
6.41
3.29
1.87
5.04
1.48
2.69
3.56
6.72

1.92
5.85
2.63
3.13
6.20
3.08
1.66
4.83
1.27
2.48
3.35
6.51

1 780
0.0038
500
152
0.006
31
1 900
14
93 000
535
100

2.69
2.17
6.91
1.78

2.55
1.96
6.70
1.65

33
28 500
0.0055
900

CHAPTER

10

Bioreaction engineering biofilms


An especially exciting group of reaction systems are biological systems. Biosystems
is often difficult to characterize and model, not least because they are time-variable
in nature. The knowledge of biological processes, both at the cellular level and at
the population level increases tremendously rapidly with the commercial interest in
biosystems increases. But the quantitative understanding and concept formation is
still in its infancy.
Yet it may be exciting to make any attempt to illuminate how the reaction kinetics,
coupled with the mass transport can be used for analysis the design of biological
reaction systems. We should go through the basics of reaction kinetics, and then select
the biofilm as an example of system with strong coupling between mass transport and
chemical reactions.

10.1

Basic concepts

Since microorganisms convert a substrate used energy and materials for several purposes (figure 10.1). A certain amount is spent to maintain and support existing living
cells. The substrate is utilized to form metabolites with a lower energy content than
the substrate. For example, a carbohydrate such as sugar converted to metabolites
such as carbon dioxide and water, whereby energy can be extracted and made useful
in the organism.
A part of the substrate is required for the formation of new cells. If this formation
is larger than mortality and removal, the number of active micro-organisms in the
system will increase. In many cases, it may mean that the total capacity to do the
job increases.
Figure 10.2 illustrates how the substrate concentration in a closed system may
decrease as a result of a reaction driven by suspended microorganisms. The concentration of microorganisms increases, while the substrate concentration decreases.
This will continue until the availability of substrate is so low that it limits microbial
growth.

125

10.2 Kinetic equations

126

cell
tillvxtreglerande
substrat

metabolism

biomassa
metaboliter
energi

vriga
substrat
Figur 10.1: Principle drawing of how a substrate is metabolized in a cell.

..
..

..
..
..
..
..
..
..
..

t=0

t>0

..
..
..

t >> 0

Figur 10.2: Illustration of how the population of micro-organisms can change in


a closed biological reaction system.

One can therefore conclude that calculations of biological systems requires models
of both populatationens dynamic that substrate dynamics.
In the following sections, we will develop general models for the sub-systems, and
then combine the equations representing these parts to describe the whole. Mass units
used will always be in grams or kg, not molar quantities.

10.2

Kinetic equations

For organisms to be able to translate the substrate, several processes must take place
in the series. The overall reaction includes several processes in sequence:
1. Transport of substrate to the cell surface
2. Transport of substrates through the cell membrane
3. Conversion of the substrate by biochemical reactions in the cell
In engineering situations, there is no real reason to distinguish the two last steps.
Thus, it is enough to note that the overall reaction requires both the transport of the
substance to micro-organism, and a turnover in the cell. We shall temporarily ignore
the fact that transport to the cell - the external mass transfer may be limiting for
overall progress.

10.2 Kinetic equations

127

Reaction rate can be defined in several ways, such as per unit volume reactor per
unit volume of cells or per unit mass of cells. This section defines:

Reaction rate per unit mass of cells

rS

Reaction rate per unit volume of reactor

rB

Reaction rate per unit volume of cells

10.2.1

gsubstrate
gcells tidsenhet
gsubstrate
3
mreactor unit time
gsubstrate
3
mcells unit time

Kinetic equations for conversion on cellular level

Living organisms have a limit on how much substrate (S) they can concert per unit
time. One reason for this is that it often is highly specific enzymes that handle a
particular type of substrate.
Assuming that the is a limited amount of enzyme (E) that is active, you can see
in an enzymatic process as a chain of two reactions:
Reversibel binding of substrate to enzyme:

S + E

Conversion of enzyme-bound substrate to product:

SE

! SE
P + E

One can formulate two rate equations for these two reactions:
r1

r2

k1 [S][E] k1% [SE]


k2 [SE]

One can now see three things:


The total amount of enzyme Etot is limited, i.e. [E] + [SE] = Etot
At steady-state r1 = r2 since [SE] is constant
Conversion of substrate , must be = r1 = r2
One can now deduce a general expression for . Fits we eliminate [E] by introducing
[E] = Etot [SE] in both rate equations:
r1
r2

=
=

k1 [S](Etot [SE]) k1% [SE]


k2 [SE]

Then, one uses that r1 = r2 :


k1 [S](Etot [SE]) k1% [SE]

k1 [S]Etot [SE](k1 [S] +

k1%

+ k2 )

k2 [SE]

From this, one may solve for and simplify [SE]:


[SE] =

k1 Etot [S]
(k1 [S] + k1% + k2 )

10.2 Kinetic equations

128

Finally, one can insert the expression for [SE] in the most simple of the rate equations,
together with = r2 = k2 [SE], to get:
=

k2 k1 Etot [S]
(k1 [S] + k1% + k2 )

This expression can be compressed by lumping together various coefficients to the


classic rate equation:
=

max [S]
[S] + K1/2

gsubstrate
gcells unit time

(10.1)

As we see, the final rate equations does only include two parameters , max and
K1/2 . The interpretation of these parameters is shown in figure 10.3.
max is the maximum reaction rate. Regardless of the substrate concentration
can be never be greater than max . It is convention to use the max to denote this
coefficient.
K1/2 is the substrate concentration at which is 50% of max . Therefore, K1/2
is named the half rate constant. As clear from equation 10.1 and figure 10.3, the rate
equation becomes flater as K1/2 increases.
Enzyme catalyzed reactions therefore follows very special kinetics. Nevertheless,
we can distinguish two special cases. At low substrate concentrations, depends on
[S], i.e. it is 1st order with respect to [S]. At high values of [S], the enzymatic systems
of the cells work at full capacity. Thus the conversion rate is independent of the
substrate concentration, i.e. the reaction is of 0 order. This is illustrated in Figure
10.4.
From equation 10.1 we can identify the two extremes:
Low substrate concentration

"

High substrate concentration

"

max [S]
K1/2
max

One should be aware that the experimentally determined equations and parameters can not be extrapolated to other system. Although micro-organisms of the same
type and strain can behave differently in different environments, e.g. by forming
aggregates affecting the mass transport.

10.2 Kinetic equations

129

Reaktionshastighet
1.0

= 1*[S] / (1+[S])
= 1*[S] / (4+[S])
0.5

K1/2=4

K1/2=1
0.0

10

Substratkoncentration [S]
Figur 10.3: Plot of the kinetic equation for two cases. Note that the parameter
K1/2 corresponds to the substrate concentration where rS is 50% of max .

Reaktionshastighet
1.0

0:e ordningen:
max

0.5

1.a ordningen:
max[S]/K1/2

0.0

= max[S] / (K1/2+[S])
4

10

Substratkoncentration [S]
Figur 10.4: Illustration of the rate equation assumes the form of a 1st-order
expression at low substrate concentration, and a 0th-order reaction at high substrate concentration.

10.2 Kinetic equations

10.2.2

130

Rate equations for conversion in reactors

Reactor calculations for mixed batch and tank reactors need to be the based on
reaction rate per volume - and time. Above, we have found that the conversion of
substrate per unit cell mass is . If one wants to make calculation per unit volume
of reactor, one needs to scale the rate equation with respect to cell density [X], which
gsubstrate
.
has the unit
m3reactor
Then, the volume-based rate equation is written:
rS = [X] =

10.2.3

max [S]
[X]
[S] + K1/2

(10.2)

Rate equation for the microbial population

The concentration of microorganisms in a closed system increases if the cell growth


is greater than cell mortality. In an open system, however, we must also take into
account the input and output fluxes to the system.
As mentioned earlier, it is only a certain fraction of the substrate that is being
converted into new micro-organisms. This fraction is called cell yield factor and
denoted Y . Y indicates the number of grams of biomass produced per gram substrate
consumed. Y can very well be > 1 if the limiting substrate itself only constitutes a
minor part of the biomass.
We can assume that both the growth and mortality is proportional to the number
of microorganisms in the system. While the growth is governed in different degrees by
the substrate concentration, the mortality can be assumed to be directly proportional
to the concentration of microorganisms, [X]. The proportionality constant is kd which
has the unit per unit time.
Net production of micro-organisms per unit volume in the system denoted by
rX , and is the difference between growth and mortality. The unit of rX becomes
gcells
:
3
mreactor unit time
rX = (Y kd ) [X]

(10.3)

Example 101: Simulation of a biochemical reaction in a batch reactor


Problem
Assume that a microbial reaction takes place in a batch reactor, where substrate is
converted according to MichaelisMenten kinetics. Parameters:
1
max
3
gsubstrate gcells
day1
3
K
3
gm
1
Y
3
gcells gsubstrate
[S] at t = 0

100

[X] at t = 0

gsubstrate m3
gcells m3

10.2 Kinetic equations

131

Solution
For a batch reactor, Prod = Acc, which means that:
d[X]
d[S]
= rX ;
= rS
dt
dt
where
rX =

(Y kd )[X]

=(

max [S]
Y kd )[X]
[S] + K1/2

(E101.4)

rS =

[X]

=(

max [S]
)[X]
[S] + K1/2

(E101.5)

These equations are put in a MATLAB-file:


100

90

80
Substratkoncentration S

Koncentration (g per m3)

70

60

50

40
Cellkoncentration X
30

20

10

5
Tid (dygn)

10

You can clearly see how the microorganisms grow as long as the supply of substrate
is high. As [S] is decreases, the growth becomes lower than the mortality, and [X] is
decreases as well. Note that the initial value of X must be> 0 for the reaction to get
started.

10.3 Biofilms

132

Figur 10.5: Biofilm of denitrification bacteria on carriers of plastic from


AnoxKaldnes. The film is about 150 (m) thick and comes from the Klagshamn
water treatment plant. The carrier has a specific surface area of 570 m2 m3 .

10.3

Biofilms

In many technical applications the reaction takes place in biological films. It is very
common in water treatment contexts. These biofilms may grow on carriers made out
of plastics or beds of stones in a reactor.
Biofilms offer some practical advantages. They counter that the level of suspended
solids becomes too high in the water, they are surprisingly resistant to harmful pollutants and they are physically stable.
A basis for biofilm process is that the substrate is transported from the surrounding
liquid bulk into the film. Here it is converted into products, e.g. biomass. Thereby
, a oncentration gradient is created from the biofilm surface in towards the carriers
wall. This creates a driving force for a flux from the liquid to the biofilm (Figure
10.7). This flux in each point is proportional to the gradient magnitude.
Intuitively, we may also find that the faster the substrate is transported into the
biofilm, the stronger the gradient and the flux become
Figure 10.7 shows schematically how the concentration profile in a biofilm can
look. The main principle is that substrates diffuse from the biofilm surface, through
the film toward the carriers surface. The driving force is the concentration gradient
as incurred by the substrate consumed in the biofilm.
The part of the biofilm that is supplied by substrate is called the penetration depth,
LP . The fraction of the film that is supplied is called the degree of penetration and is
defined as the ratio between the penetration depth and the thickness of the biofilm,
LP /LB .

10.3 Biofilms

133

Figur 10.6: Biofilm on a plastic carrier.

The degree of penetration is defined as:


B=

LP
LB

The 4 different cases that are shown in figure 10.7 are:


a: Deep film The substrate can be consumed so rapidly that the concentration at
the carrier surface becomes 0. Deep refers to the fact that the film is unnecessary deep in relation to what is needed to consume all the substrate. This
implies that microorganisms closest to the carrier surface starve and die. The
degree of penetration B < 1.
b: Exactly penetrated film Substrate is converted everywhere the biofilm, and
the concentration of substrate becomes 0, exactly at the carriers surface. This
is the optimal situation, the whole film is used and and substrate conversion is
complete. The degree of penetration B = 1.
c: Thin film The whole film is utilized but the microorganisms can not convert all
of the substrate. The degree of penetration B > 1.
d: External mass transfer limitation If the reaction rate in the bio film is high
and/if the mass transfer from the mixed liquid bulk to the film surface is slow,
the external mass transfer becomes important. It means that the concentration
at the biofilm surface is lower than the bulk concentration. The result is that the
driving force for both the chemical reaction in the film as well as the diffusion
of substrate in the film is adversely affected.

10.3 Biofilms

134

omblandad vtskebulk

brare

biofilm
substratflux

C=C0

C=0
L=0

LP

LB

(a) Deep film, non-penetrated. Parameter B < 1.


omblandad vtskebulk

brare

biofilm
substratflux

C=C0

C=0
L=0

LP=LB

(b) Exactly penetrated. Parameter B = 1.


omblandad vtskebulk

brare

biofilm
substratflux

C=C0

C>0

L=0

LB

(c) Thin film, penetrated. Parameter B > 1.


omblandad
vtskebulk

laminrt
grnsskikt

brare

biofilm
substratflux

C=Cbulk
C=C0

L=0

LB

(d) A laminar boundary layer creates an external mass transfer


resistance between the liquid bulk and the surface if the biofilm.

Figur 10.7: Different types of concentration profiles in biofilms. B was defined


in equation 10.6.

10.3 Biofilms

10.3.1

135

General rate equations for biofilms

In all biofilm processes, the biofilm should ideally have constant thickness, and fairly
constant activity. For a certain amount of substrates, there is a certain volume of
cells. For biofilms we can therefore postulate that:
1. The reaction rate should be defined in terms of units of volume biofilm. Hence,
we use rB ( m3 gsubstrate
)
unit time
biofilm

2. Cell mass is constant, so that

d[X]
= 0 everywhere in the reactor.
dt

As previously mentioned, the kintetiska parameters and K1/2 specific to a particular system. In this section, we will simply assume that have been determined
for the current biofilm system.
If we denote cell mass per unit volume of biofilm with XB the rate equation for a
biofilm becomes:
rB = XB =

max [S]
XB
[S] + K1/2

(10.4)

This is analogous with equation 10.2. In practice, however, one determines the product XB , and the rate equation can be written:
rB =

kB [S]
[S] + K1/2

(10.5)

where the rate constant kB , has the unit g of substrate per unit volume of biofilm
and time.
As mentioned earlier, the conversion of substrate depends on substrate concentration only at low values of mconc (S). We may therefore identify two special cases
with respect to kinetics:
Low substrate concentration:

rB "

High substrate concentration:

rB "

10.3.2

kB [S]
= k1 [S]
K1/2
kB = k0

Concentration profile and flux in a thin film for 0th


order reaction

Concentration profile
In this case the rate expression is reduced to:
rB = k0
For each point in the film, the continuity equation gives:
D

2c
= k0
L2

10.3 Biofilms

136

Now dimensionless quantities are introduced for concentration and for length. If the
concentration at the surface of the film is c0 and the biofilm thickness is LB , then:
c
c =
;
c = c c0
c0
L
;
L = L LB
LB
If we combine all these equations, the continuity equation becomes:
L =

2 c c0
= k0
(L LB )2

2 c
L2B
=
k
0
L 2
c0

2 c
k0 L2B
=
L 2
Dc0

It is now possible to clean up the right hand side by introducing the variable:
B2 =
and we get:

2Dc0
k0 L2B

(10.6)

2 c
2
= 2
(10.7)
L 2
B
We will now solve equation 10.7 with two boundary condition. We utilize that:

L = 0 c = 1
Concentration known at the surface of the biofilm

c
L = 1
= 0 Flux at the carriers surface is zero
L
Two integrations of equation 10.7 gives:
c
2
=
L + H1
(10.8)
L
B2
L2
c
=
+ H1 L + H2
(10.9)
B2
We use the second boundary condition with equation 10.8, and thereafter the second
boundary condition with equation 10.9 to get:
2
2
0 = 2 + H1
H1 = 2
B
B
1 = H2
This is put into equation 10.9 and we get at complete expression for the concentration
profile in the biofilm:
L2
2L
2 +1
(10.10)
2
B
B
Obviously, the concentration profile is a relatively simple 2nd order equation.
There is one very important special case. In a situation where the substrate just
is consumed in the film, so that the substrate concentration c = 0 exactly when
L = 1. In that case, equation 10.10 gives:
c =

L2
2L
1
2 +1 ; 0= 2 +1 ; B =1
2
B
B
B
2Dc0
This means that exactly when B 2 =
= 1 (i.e. when B=1) the substrate
k0 L2B
concentration is 0 at the surface of the carrier.
c =

10.3 Biofilms

137

Flux into the film


The flux into the film, expressed as gsubstrate per unit surface area of biofilm and unit
time, is obtained by differentiation of the non-dimensionless form of equation 10.10,
taking the concentration gradient at L = 0. First, we re-introduce:
c =

c
L
; L =
c0
LB

to get:
c
c0

L2 1
L 2
2

+1
2
LB B
LB B 2

c
L

c0 (

c
L

2L
2

)
L2B B 2
LB B 2

2c0
i L=0
LB B 2

If we insert this and the definition of B into the diffusion equation we get the final
expression for the flux of substrate into a thin biofilm:
J = k0 LB

(10.11)

The flux depends only by the rate constant and the film thickness. Hence the flux, i.e.
the diffusion rate into the film is independent of the diffusivity of the substrate in the
biofilm itself, provided that the reaction is independent of the substrate concentration.

10.3.3

Concentration profile and flux in a deep film for 0th


order reaction

Concentration profile
If the biofilm is deep, the substrate to be consumed completely in the film, and flow
towards the carrier becomes 0. As shown in Figure 10.7 the substrate penetrates the
film to a depth of LP .
As shown above, the general equation describing simultaneous, steadystate, diffusive mass transfer and chemical reaction is described by:
2 c
2
= 2
L 2
B
For a deep film, this transfer

L = 0 c = 1

L = LP c = 0

L = LP c = 0
L
One integration gives:

c
2
= 2 L + H1
L
B

(10.12)
equation should be solve with three conditions:
Concentration known at the surface of the biofilm
Flux at the carriers surface is zero
Flux at the penetration depth is zero

10.3 Biofilms

138

Tabell 10.1: Equations for fluxes into biofilms for different cases.

0:e ordningen

B1

1:a ordningen

Generellt

Tunn film

B1

Tjock film

J = k0 LB

J = 2Dk0 c0

J = c0 Dk1

Then, the second boundary condition gives that:


0=

2
LP + H1
B2

H1 =

2LP
B2

Another integration gives:


c =

L2
+ H1 L + H2
B2

If we first insert the first boundary condition and then the second boundary condition
in the above equation we get:

0=

1 = 0 + 0 + H2

H2 = 1

2LP
2LP
2 +1
2
B
B

0=

LP
+1
B2

LP = B

H1 =

2
B

This can be summarized in the polynomial equation describing the concentration


profile in a deep film:
c =

L2
2L

+1
B2
B

(10.13)

Flux into the film


The flux into the film is obtained by differentiation of the non-dimensionless form of
equation 10.13, taking the concentration gradient at L = 0.
c
c0

L2 1
L 2
2
+1
2
LB B
LB B

c
L

c0 (

c
L

2L
L2B B 2

2
LB B

2c0
i L=0
LB B

(10.14)
(10.15)
(10.16)

If we combine this, and the definition of B, with the continuity equation obtain the
final expression for the flux into a deep biofilm at 0th order reaction. The expression
is:
8
J = 2Dk0 c0
(10.17)
The flux depends on all possible parameters, except the thickness of the film, which
is logical.

10.3 Biofilms

139

Example 102: Filmtjocklek vid exakt penetrerad film


Problem
Calculate the minimum biofilm thickness if it should be fully penetrated under the
following conditions:
kB
K 12
D
c0

0.3
1
31010
25

g m3 s1
mg L1
m s1
mg L1

Solution
The unit mg L1 is equivalent to g m3 . Since c0 >> K 12 we can assume 0th order
reaction in the entire biofilm. and thus k0 = kb . This is, however, an approximation
as clear from the figure below, showing the concentration profile: c = c0L 2 2L + 1
(always valid when B = 1):
25

Substratkoncentration [S] (mg/L)

20

15

10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Dimensionsls lngd L* = L/LB

For a 0th order reaction, the rate coefficient k0 = kb ., and since the film is fully
penetrated, B 1. Since:
B2 =

2Dc0
1
k0 L2B

the biofilm thickness when B = 1 may be written:


7
7
2Dc0
2 31010 25
LB
=
= 223 106
k0
0.3

(E102.18)

If the biofilm thickness is less than ca 220 m, it will thus be fully penetrated, i.e. a
thin film.

10.3 Biofilms

10.3.4

140

Reactor calculations

For typical steadystate CSTR, the mass balance equation is:


In + Prod
Qin cin + rV

=
=

Ut
Qout cout

In a biofilm reactor, the reaction occurs because the reactant (substrate) is removed
by means of a flux into biofilm surfaces in the reactor. The Prod term can therefore
be written:
r = J A
where A is the number of m2 biofilm surface per m3 reactor. Thereby, we can easily
develop a relevant mass balance equation:
Qin cin JAV = Qut cut

Example 103: Design of CSTR reactor volume


Problem
Calculate the volume of a biofilm reactor design to reach 80% conversion in a biofilm
system characterized by the data in example 102. Reactor data:
LB 100 m
Q
10 m3 hour1
cin
50 mg L1
A
150 m2 m3
Solution
The concentration in the reactor will be 20% of the input concentration, i.e. 10
mg m3 . In the absence of external mass transfer resistance, this is also the concentration at the surface of the biofilm. c0 . First, one must check if the film is penetrated
or not:
:
:
2Dc0
2 31010 10
B=
=
= 1.41
k0 L2B
0.3(100106 )2
Thus, the film is fully penetrated.
We can also check the assumption that the reaction is of 0th order. At the carrier
surface (L = 1), equation 10.10 states that the concentration is:
c = c0 (

1
2
1
2
2 + 1) = 10(

+ 1) = 5 mg/L
2
2
B
B
1.41
1.412

Since c >> K 21 it is clear that assumption regarding 0:e order is reasonable. We can
now solve for the reactor volume V by means of the reactor mass balance:
V =Q

cin cut
J A

(E103.18)

10.4 External mass transfer

141

Since J = k0 LB (equation 10.11) we finally get:


V =Q

10.4

cin cout
10
50 10
=

= 37 m3
k0 LB A
3600 0.3100106 150

(E103.19)

External mass transfer

In many systems the microbial turnover of substrate is limited by how fast the substrate can be transported to the biofilm surface. This situation is outlined in the
lower panel in Figure 10.7. Applying the film theory, this means that there is a laminar boundary layer. Through this boundary layer, substrate is transported toward
the surface only by diffusion. We may see this external and internal mass transport
processes as two resistors in series.
In general terms, the mass transfer through the laminar boundary layer can be
described by:
JL =

D
(cbulk c0 ) = km (cbulk c0 )

(10.18)

The flux is thus dependent on the diffusivity of the substrate in water (D), as well as
the thickness of the laminar boundary layer (). The ratio between D and , km , is
called the mass transfer coefficient.
There are many methods available to calculate the mass transfer coefficient. Common features are:
1. They are empirically developed for a particular type of system
2. They are based on dimension-less numbers such as Sh, Re and Sc
3. They are associated with uncertainty and should be used very critically and
carefully

10.4.1

Mass transfer in packed beds

Table 10.2 shows the dimension-less numbers and quantities one needs as minimum
to be able to assess external mass transport for biofilms in packed bed.
An example of correlation is:

1/3
1.09
for 0.0016 < Re < 55 och 165 < Sc < 70 000
Sc
Sh =
(10.19)
0.25 0.69 1/3
Sc
for 55 < Re < 1500 och 165 < Sc < 11 000
Re
When you count on other than pipes, one must be careful to define the Re in the right
way. In a packed bed the characteristic length of Re is defined as the particle radius.
If the particles in the bed are of different sizes one base the equivalent particle radius
on the material surface area:
7
1 Ap
L=
(10.20)
2

10.4 External mass transfer

142

Tabell 10.2: Relationships needed to calculate mass transfer rates.

Dimension-less number
Re

Reynolds number

Sc

Schmidts number

Sh

Sherwoods number

Definition
vL
D

D
km L
D

Entity:

Effective porosity
v
Darcyvelocity
L
Characteristic length
D
Diffusivity

Kinematic viscosity
km Mass transfer coefficient

Unit

m s1
m
m2 s1
m2 s1
m s1

where Ap is the average particle area express in m2 .


The thickness of the laminar boundary layer can be calculated by combining the
definition of the mass transfer coefficient into the definition of the of the Sherwood
number:
Sh =

km L
DL
L
=
=
D
D

(10.21)

so that:
=

10.4.2

L
Sh

(10.22)

Coupled external mass transfer resistance and diffusion/reaction in a biofilm

Since there is no chemical reaction in a laminar boundary layer, it must be true that:
J=
= Flux into the laminar boundary layer
= Flux out from the laminar boundary layer
= Flux into the biofilm

10.4 External mass transfer

143

With the relationships developed earlier, we can thus state that:


0th order, thin film:
0th order, deep film:
1st order, general:

D
(cbulk c0 ) = k0 LB

8
D
(cbulk c0 ) = Dk0 c0

8
D
(cbulk c0 ) = J = c0 Dk1

In qualitative terms, we can se that:


The flux into the biofilm will be less if the the external mass transfer resistance
increases
On can calculate c0 from the relationships above, after having checked B
Although there are countless possibilities to perform calculations on this type of systems, the discussion is left at this point.

10.4 External mass transfer

144

Example 104: Boundary layer thickness in a packed bed


Problem:
Calculate the laminar boundary layer, as a function of the Darcyvelocity in a packed
bed, if v varies between 0.01 and 2 m minute1 . The bed consists of sherical particles
with a diameter of 2 cm, and the effective porosity is 0.40.
Solution:
In SI-unigts, the kinematic viscosity of water is 1 106 and the diffusivity in water
is typically 1 109 . The calculation has been summarized in the following MATLAB
m-file: The results indicate that the boundary layer thickness depends on the flow
condition in almost the entire interval. The maximum boundary layer thickness is 0.8
mm.
-4

Berkning av grnsfilmstjocklek i packad bdd

x 10

Hr r Re < 55
7

Grnsfilmstjocklek (m)

3
Hr r Re > 55
2

0.2

0.4

0.6
0.8
1
1.2
1.4
Darcy-hastighet (m/minut) genom bdden

1.6

1.8

You might also like