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 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1 Why do we perform process and reactor calculations? 8
1.1 Processes and reactors . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.1.1 System boundaries and sub-systems . . . . . . . . . . . . . . . 11
1.1.2 Non-reaction and reaction systems . . . . . . . . . . . . . . . . 12
1.1.3 Element, components and inert substances . . . . . . . . . . . 12
1.1.4 Static and dynamic systems . . . . . . . . . . . . . . . . . . . . 13
1.2 The mass balance principle . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 Process calculations and reactor calculations . . . . . . . . . . . . . . . 15
1.4 Numerical methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4.1 Systems of linear algebraic equations . . . . . . . . . . . . . . . 15
1.4.2 Numerical solution of systems of algebraic equations . . . . . . 16
1.4.3 Solution of dierential equations . . . . . . . . . . . . . . . . . 17
2 Process calculations for non-reaction systems 22
2.1 Component mass balances . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2 Degrees of freedom analysis . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3 Process calculations for a separation process . . . . . . . . . . . . . . . 26
3 Process calculations with multiple units 32
3.1 Calculation methodology . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2 Common types of units . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2.1 Mixer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2.2 Splitter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2.3 Recirculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.2.4 Bypass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.2.5 Purge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4 Process calculations for reaction systems 41
4.1 Mass balances mass or mole? . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Important concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.2.1 Inert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.2.2 Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.2.3 Yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.2.4 Selectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3
Contents 4
4.3 The element mass balance method . . . . . . . . . . . . . . . . . . . . 44
4.3.1 The atomic matrix . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.4 The component mass balance method . . . . . . . . . . . . . . . . . . 47
4.4.1 The reaction parameter . . . . . . . . . . . . . . . . . . . . . 47
4.4.2 The reaction matrix . . . . . . . . . . . . . . . . . . . . . . . . 48
4.4.3 Degree of freedom analysis . . . . . . . . . . . . . . . . . . . . 49
5 Reactor calculations 52
5.1 Kinetic rate equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.1.1 First order kinetics . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.1.2 Second order kinetics . . . . . . . . . . . . . . . . . . . . . . . . 55
5.1.3 Overview of kinetic rate equation . . . . . . . . . . . . . . . . . 55
5.1.4 Reversible reactions . . . . . . . . . . . . . . . . . . . . . . . . 56
5.1.5 Consecutive reactions . . . . . . . . . . . . . . . . . . . . . . . 56
5.1.6 Parallel reactions . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.2 Mean residence time and reaction time . . . . . . . . . . . . . . . . . . 57
5.3 Reactor models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.3.1 The dierential mass balance . . . . . . . . . . . . . . . . . . . 60
5.3.2 Methodology for reactor calculations . . . . . . . . . . . . . . . 62
5.4 The ideal completely stirred tank reactor, CSTR . . . . . . . . . . . . 62
5.5 The ideal batch reactor . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.6 The ideal plug-ow reactor, PFR . . . . . . . . . . . . . . . . . . . . . 71
5.6.1 PFR reactor modeling in terms of conversion . . . . . . . . . . 72
6 Non-ideal reactors 76
6.1 Examples of non-ideal mixing . . . . . . . . . . . . . . . . . . . . . . . 76
6.2 Residence-time distributions . . . . . . . . . . . . . . . . . . . . . . . . 77
6.2.1 The normalized RTD E(t) . . . . . . . . . . . . . . . . . . . . 78
6.2.2 c(t) for an ideal CSTR . . . . . . . . . . . . . . . . . . . . . . . 79
6.2.3 E(t) for an ideal CSTR . . . . . . . . . . . . . . . . . . . . . . 81
6.2.4 Mean residence time from E(t) . . . . . . . . . . . . . . . . . . 81
6.2.5 The F(t)-distribution . . . . . . . . . . . . . . . . . . . . . . . . 82
6.3 Non-ideal reactor models . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.3.1 The CSTR in series model . . . . . . . . . . . . . . . . . . . . . 85
6.4 The segregation model . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.5 Other reactor models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7 Instationra CSTR 91
7.1 Svar p ndring i ingngskoncentration . . . . . . . . . . . . . . . . . . 92
7.2 Arbetsmetodik . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
8 Dispersion in porous media 99
8.1 Water ow- Darcys law . . . . . . . . . . . . . . . . . . . . . . . . . . 100
8.1.1 Flow rate distribution - reactor modeling . . . . . . . . . . . . 103
8.1.2 Advective transport . . . . . . . . . . . . . . . . . . . . . . . . 104
8.2 Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
8.2.1 Molecular diusion - Ficks law . . . . . . . . . . . . . . . . . . 104
8.2.2 Analogy dispersion - diusion . . . . . . . . . . . . . . . . . . . 105
8.2.3 Deduction of expression for spread . . . . . . . . . . . . . . . . 106
Contents 5
8.2.4 Interpretation of the standard deviation . . . . . . . . . . . . . 108
8.2.5 Application to a advection-dispersion system . . . . . . . . . . 109
8.3 Dispersion coecients in groundwater aquifers . . . . . . . . . . . . . 110
8.3.1 Dispersivitet . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
9 Transport in porous media 113
9.1 The continuity equation - advection and dispersion . . . . . . . . . . . 113
9.1.1 Continuous source . . . . . . . . . . . . . . . . . . . . . . . . . 114
9.2 Advection, dispersion and reaction . . . . . . . . . . . . . . . . . . . . 116
9.2.1 Continuous source . . . . . . . . . . . . . . . . . . . . . . . . . 117
9.2.2 Continuous addition during a limited time - step up and down 117
9.2.3 Steadystate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
9.3 Adsorption and transport . . . . . . . . . . . . . . . . . . . . . . . . . 119
9.3.1 Advection, dispersion, reaction and adsorption . . . . . . . . . 121
9.3.2 Determination of K
d
and R for organic substances . . . . . . . 123
9.4 The continuity equation in multiple dimensions . . . . . . . . . . . . . 123
10 Bioreaction engineering - biolms 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 Biolms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
10.3.1 General rate equations for biolms . . . . . . . . . . . . . . . . 135
10.3.2 Concentration prole and ux in a thin lm for 0th order reaction135
10.3.3 Concentration prole and ux in a deep lm 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 diusion/reaction
in a biolm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
Examples 6
Examples
11: Balance calculation for a bank account . . . . . . . . . . . . . . . . . . 14
12: Solving linear equation systems . . . . . . . . . . . . . . . . . . . . . . 16
13: Solving linear equations with fsolve . . . . . . . . . . . . . . . . . . . 17
14: Solving non-linear equations with fsolve . . . . . . . . . . . . . . . . 17
15: Numerical integration with Eulers method . . . . . . . . . . . . . . . 19
16: Numerical integration with ode45 . . . . . . . . . . . . . . . . . . . . 21
21: Separation of ethanol and water . . . . . . . . . . . . . . . . . . . . . 26
22: Separation of ethanol and water - 6 unknowns . . . . . . . . . . . . . 28
23: Ethanol and water concentration constraints . . . . . . . . . . . . . 29
24: Separation of ethanol and water moved constraint . . . . . . . . . . 30
31: Limitation of the number of splitter constraints . . . . . . . . . . . . . 35
32: Process calculation based on total system analysis . . . . . . . . . . . 37
33: Degree of freedom analysis based on sub-systems . . . . . . . . . . . . 40
41: Catalytic dehydrogenation . . . . . . . . . . . . . . . . . . . . . . . . . 44
42: Atomic matrix with linearly dependent rows . . . . . . . . . . . . . . 45
43: The rank of a reaction matrix . . . . . . . . . . . . . . . . . . . . . . . 48
44: Application of the reaction parameter method . . . . . . . . . . . . . 50
51: Average residence time of a lake . . . . . . . . . . . . . . . . . . . . . 59
52: CSTR with a 1st order irreversible reaction . . . . . . . . . . . . . . . 63
53: Example 52 with matrix notation . . . . . . . . . . . . . . . . . . . . 64
54: Numerical solution for non-unity reaction order . . . . . . . . . . . . . 65
55: Batch reactor with 1st order irreversible reaction . . . . . . . . . . . . 67
56: Dierential, numerical solution of Example 55 . . . . . . . . . . . . . 68
57: Integral, numerical solution of Example 55 . . . . . . . . . . . . . . . 69
58: Simulation of consecutive and parallel reactions . . . . . . . . . . . . . 70
59: Emission of a pollutant into a stream . . . . . . . . . . . . . . . . . . 73
61: Residence-time distribution from tracer experiments . . . . . . . . . . 83
62: Series with an innite number of CSTR . . . . . . . . . . . . . . . . . 87
63: The segregation model applied to example 61 . . . . . . . . . . . . . 89
71: Exempel p simulering av instationr tankreaktor . . . . . . . . . . . 95
81: Estimation of ow velocity . . . . . . . . . . . . . . . . . . . . . . . . 102
82: Diusive and dispersive spread . . . . . . . . . . . . . . . . . . . . . . 110
91: Inltration of a pollutant into a soil . . . . . . . . . . . . . . . . . . . 116
92: Complex pollutant transport in a soil . . . . . . . . . . . . . . . . . . 122
101: Simulation of a biochemical reaction in a batch reactor . . . . . . . . 130
102: Filmtjocklek vid exakt penetrerad lm . . . . . . . . . . . . . . . . . 139
103: Design of CSTR reactor volume . . . . . . . . . . . . . . . . . . . . . 140
104: Boundary layer thickness in a packed bed . . . . . . . . . . . . . . . 144
Examples 7
Notations
Note that the choice of mass unit (g, kg, mol) often is arbitrary
Properties introduced in Chapters 1-7
c
j
Mole or mass concentration of substance j mol L
1
, kg L
1
F
i,j
Molar ux of substance j in stream i mol or mol s
1
k Kinetic rate constant Varies
Q Volume ux m
3
s1
V Volume m
3
W
i,j
Mass ux of j in stream i kg or kg s
1
x
i,j
Mole fraction of j in liquid stream i
X Conversion with respect to substance j
y
i,j
Mole fraction of j in gas stream i

i,j
Stoichometric coecient for j in reaction i

i,j
Mass fraction of substance j in liquid stream i

i,j
Mass fraction of substance j in gas stream i
Residence time in a reactor time
1
Properties introduced in Chapters 8-10
B Penetration fraction
D Diusion coecient m s
1
E Dispersion coecient m s
1
J Diusive ux mass m
2
s
1
k
m
Mass transfer coecient m s
1
K
X
Hydraulic conductivity m s
1
K
d
Distribution coecient m
3
kg
soil
1
K
oc
Distribution coecient m
3
kg
SOM
1
L Length coordinate m
L
B
Thichness of biological lm m
L
P
Penetration depth m
R Retardation factor
S Sorbed amount g kg
1
u Linear ow velocity m s
1
v Darcy velocity m s
1
Dispersivity m
Boundary layer thickness m
Porosity

e
Eective porosity
Substrate turnover rate g s
1
Bulk density kg m
3
Volumetric water content
C H A P T E R 1
Why do we perform process and
reactor calculations?
There are several situations when an engineer needs to gain an understanding of how
dierent substances ow 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 pro-
cesses that man cannot inuence, making the design aspect less important for natural
systems.
Every environmental engineer needs to know how to perform mass balance cal-
culations for dierent 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 dierent compartments and pools relevant for the
global carbon cycle.
This compendium treats various techniques to carry apply mass balance calcula-
tions 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 calcula-
tions. 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.
8
Why do we perform process and reactor calculations? 9
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 prob-
lems 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 stoichiome-
try, element matrix, reaction matrix and reaction parameter. Other concepts include
conversion, yield and selectivity.
Reactor calculations treats what happens inside of chemical reactors. The chap-
ter addresses kinetic rate equations, ideal reactor models such as the tank reactor
CSTR, the batch reactor, and plug ow reactor. This chapter is important for gain-
ing an understanding of how a chemical reaction is inuenced by and aects 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 briey,
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 dierential mass balances and simulated with dierential equations.
Dispersion in porous media takes up irregular ow in groundwater systems,
among others. In such systems, mixing can be described by models analogous to
those for molecular diusion. 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 biolms with respect to the penetration rate,
reaction order and lm mass transport resistance.
1.1 Processes and reactors 10
solids
separation
denitrification nitrification P-precipitation
org-C
org-N
org-P
solids
org-C
org-N
org-P
solids
sludge sludge
sludge
CO
2
N
2

org-C
NH
4
+

org-P
org-C
NO
3
-

P
O
2
org-C
NO
3
-
P
Fe
3+

org-C
NO
3
-

P
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 identied 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 eective is the denitrication process?
Is enough carbon available in the rst biological treatment step (denitrication)?
How large should the recirculation of NO

3
from the nitrication process to the
denitrication process be?
How much will the purication eciency 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
solids
separation
denitrification nitrification P-precipitation
org-C
org-N
org-P
solids
org-C
org-N
org-P
solids
sludge sludge
sludge
CO
2
N
2

org-C
NH
4
+

org-P
org-C
NO
3
-

P
O
2
org-C
NO
3
-
P
Fe
3+

org-C
NO
3
-

P
Figur 1.2: Example of how system boundaries can be drawn around the dierent
units in a process.
1.1.1 System boundaries and sub-systems
All the above questions deal with how much material that ows at dierent points
in the treatment plant. When one calculates the magnitude of these ows, one must
rst dene the systems boundaries. If one is only interested in the total ows 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 ows in the streams crossing
the system boundary. Hence, one cannot say anything about the mass ows 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 dierent man-
ner. If we, for example, want to calculate how eective the last sub-system (the P-
precipitation) is, we must draw the system boundary as shown in Figure 1.2. There
the system has been conned so that all components that enter or leave this unit are
represented by streams across the system boundary.
1.1 Processes and reactors 12
1.1.2 Non-reaction and reaction systems
The sub-systems mentioned in the previous section, the sand trap and the biological
purication 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 owing
into the sub-system will thus be in exactly the same chemical form, but perhaps in
a dierent 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 purication 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 dicult to deter-
mine exactly which chemicals react and what products that are formed. One the
contrary, there are many examples of natural reaction systems that are very dicult
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 con-
servative 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 ow of an element that goes
into a system, both non-reaction and reaction systems, is equal to the mass ow out
of the system.
A component is a chemical compound of any kind, for example ions such as NO

3
and SO
2
4
, or compounds such as CO
2
or benzene. In a steady-state, non-reaction
system the mass ow of a component in is always equal to the mass ow out. In a
steady-state, reaction system, however, this precondition never applies. This makes
process balance calculations for reaction systems more dicult than those for non-
reaction 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 simplication
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
0 1 2 3 4 5 6 7 8 9 10
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
stationr fas
ny stationr fas
icke-statior, dynamisk fas
tiden
n

g
o
n
t
i
n
g

v
a
d
-
s
o
m
-
h
e
l
s
t
hr intrffar en strning
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 ow of a process, it always leads to changes
in the levels of reactants and/or products in the euent. 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 dicult to predict and understand
due to the dynamics of various processes, since many chemical processes can aect
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 ow of the substance into the system. The ow
can have the unit kg s
1
or similar. The term Output concerns, in the same way,
the ow 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 dicult. 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 denition. 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 dierence or dierential 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
year
45000
SEK
year
+ (2000 500)
SEK
year
= 6500
SEK
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:
Prod = 45000
SEK
year
50000
SEK
year
+ 5600
SEK
year
= 600
SEK
year
1.3 Process calculations and reactor calculations 15
1.3 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 uxes (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 dicult 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 equa-
tions. 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 = A
1
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:
A
1
=
_
1 1
2 3
_
, A
2
=
_
1 1
2 2
_
1.4 Numerical methods 16
only A
1
can be inverted, while A
2
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 A
1
=2, while the rank of
A
2
=1.
Example 12: Solving linear equation systems
Problem
Solve the following system of equations with Matlab:
3x
1
+ 4x
2
= 10
x
1
x
2
= 2
Solution
The system of equations can be written as a matrix in the form:
AX = Y
where:
A =
_
3 4
1 1
_
, X =
_
x
1
x
2
_
, 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 nding 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
dened the equations in matrix notation.
Solution
The system of equations can be written as a matrix on the form:
0 = AX Y
where
_
3 4
1 1
_
, Y =
_
10
2
_
Write an m-le: 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; x
1
= 2.5714, x
2
= 0.5714.
Example 14: Solving non-linear equations with fsolve
Problem
Solve the following system of equations with the routine fsolve in Matlab:
3x
1
+ 4x
2
2
= 10
x
3
1
x
2
= 2
Solution
The system of equations can be written as a matrix in the form:
0 = 3x
1
+ 4x
2
2
10
0 = x
3
1
x
2
2
The two functions are the dened in an m-file exfsolve1.m: which is executed
from workspace with: Solution: x
1
= 1.4708, x
2
= 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 dierential equations
Dynamics in all types of systems is described by dierential equations, ordinary or
partial. It is only for models of rather idealized systems that the resulting dierential
equations are analytically soluble. For many classes of problems it is possible to lin-
earize 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 dierential equations. In this
compendium we will almost exclusively solve dierential equations numerically. The
exceptions are the simplest cases.
How do we solve a dierential equation numerically? First we will take a quick
look at the simplest (and worst) method, Eulers solution method.
An ordinary dierential equation can be written in the form:
dX
dt
= f(t, X)
This dierential equation states that if X = X
t
at time t, then X = X
t+dt
at time
t +dt:
X
t+dt
= X
t
+dtf(t, X
t
)
We can interpret this graphically if we follow the derivative of the function X(t)
a rectilinear piece dt, starting at the point (t, X
t
). The approximately estimated
subsequent value of the function is thus a small distance away, in the direction of the
derivative. This value is dened as (t +dt, X
t+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 dier signicantly 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
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
-1
0
1
2
3
4
5
6
t+dt
t
X
t
X
t+dt
approximering av X
t+dt
X
dX/dt i (t,X
t
)
Figur 1.4: Illustration of Eulers method.
Example 15: Numerical integration with Eulers method
Problem
Integrate (simulate) the ordinary non-linear dierential equation:
dc
dt
= 0.5c Initial value:
_
t = 0
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-le exEuler.m: which is executed from workspace with exEuler. In
the m-le, 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!
0 0.5 1 1.5 2 2.5 3 3.5 4
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
exakt
euler
1.4 Numerical methods 21
Example 16: Numerical integration with ode45
Problem
Integrate (simulate) the ordinary non-linear dierential equation with ode45 in Matlab:
dX
1
dt
= k
1
X
1
k
1
= 0.5
dX
2
dt
= k
2
X
1
X
2
k
2
= 0.1
Solution
Write an m-le 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 X
1
and X
2
respectively. The solution is presented as a diagram:
0 1 2 3 4 5 6 7 8 9 10
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
X
1
X
2
V

r
d
e
t

p


t
i
d
s
v
a
r
i
a
b
e
l
t

t
i
l
l
s
t

n
d

X
tiden t
C H A P T E R 2
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 ows of various chemical substances.
The process calculations my be used to quantify uxes, 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 ows of mass and matter in any one of earths
ecosystems.
Mass balances are known under dierent 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, dierent types of equations and information
will be slowly introduced and will illustrate key concepts in the mass balance calcu-
lations.
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
W
1,A
W
1,B
W
2,A
W
3,B
W
4,A
W
4,B
1
2
3
4
Process
systemgrns
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 ow into the system at one end, and out
of the system at the other end. The mass ux of the substance is denoted as W
i,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 uxes:

in
W
i,j
=

out
W
i,j
Figure 2.1 shows a process where two physical streams, W
1
and W
2
are input
uxes, while W
3
and W
4
are output uxes. W
1
and W
4
both contain the chemical
components A and B, while W
2
only contains A and W
3
only contains B.
The units for mass ux W may be mass (kg) or mass ux (kg/time). Under these
conditions, the integral component mass balances for A and B become:
MB A : W
1,A
+W
2,A
= W
4,A
MB B : W
1,B
= W
3,B
+W
4,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 W
1,A
, etc., are called stream variables. A stream variable refers to
the ux of a specic substance in a specic 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 dened. This is
also true in the example above, although only 6 stream variables are explicitly written
out. The reason is that implicitly W
2,B
= W
3,A
= 0.
2.2 Degrees of freedom analysis 24
2.2 Degrees of freedom analysis
The dierence 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 dierent cases:
Degrees of freedom > 0 The system is not completely dened.
Degrees of freedom = 0 The system is dened.
Degrees of freedom < 0 The system denition contains to much information.
Thus, rst step when chemical systems are analyzed is to gure 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 infor-
mation that can be expressed as an equation! For example, in the following, we make
us of independent information given in terms of:
1. known uxes
2. known concentrations
3. total constraints
4. basis of calculations
With known uxes it is meant that the mass ux of a component in a stream, W
i,j
or
the molar ux, F
i,j
is given. As mentioned above, the ux may either be given as an
amount (kg, mole) or a ux which is amount per unit of time (kg/time, mol/time).
In the real world, it is quite dicult to make direct measurements of component
uxes. Instead, one measures the concentration of a component in a stream and
multiplies it by the total ux in the same stream. This will give W
i,j
and/or F
i,j
.
A known concentrations might mean the actual concentration c
i,j
(in kg/volume
or mole/volume) but just as often the mass fraction
i,j
, or the molar fraction x
i,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. Mathe-
matically, 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

j=1
x
i,j
= 1
One can also dene corresponding total constraints for mass uxes and molar uxes.
For a given physical stream W
i
, that contains n components j, we can state that the
2.2 Degrees of freedom analysis 25
1,A
W
1
1,B
W
1
1
2
3
4
Process
systemgrns
2,A
W
2
4,A
W
4
4,B
W
4
3,A
W
3
Figur 2.2: Steam variables expressed in terms of mass fractions
i,j
and the total
ux W
i
.
component uxes (quantied by the stream variables) must add up to the total ux.
Hence:
W
i
=
_

n
j=1
W
i,j
W
i

n
j=1

i,j
The corresponding relationships also hold true when the stream variables are ex-
pressed in molar units:
F
i
=
_

n
j=1
F
i,j
F
i

n
j=1
x
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 ux
W
i
.
A basis of calculation consists of an arbitrary denition of a ux. 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 uxes (i.e. no values of stream variables) are
known or given. This occurs when only concentrations or mass/molar fractions are
specied. 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 26
2.3 Process calculations for a separation process
In this section we shall see how we methodically can solve a process calculation prob-
lem for a non-reaction system at steady state. The starting point is a very simple
separation process where a mixture of ethanol, C
2
H
5
OH and H
2
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 dierent 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 rst examples dealing with separation of the two components is schematically
illustrated below. In the process, C
2
H
5
OH and H
2
O ow into a process with one
input stream and leave the system in the other end via two output streams. The
mass ux of each component is denoted as W
i,j
, where i denotes stream i, while j = 1
denotes C
2
H
5
OH and j = 2 denotes H
2
O:
W
1,1
kg C
2
H
5
OH
W
1,2
kg H
2
O
W
2,1
kg C
2
H
5
OH
W
2,2
kg H
2
O
W
3,1
kg C
2
H
5
OH
W
3,2
kg H
2
O
1
2
3
separations-
process
Example 21: Separation of ethanol and water
Problem:
The input stream (the feed) to a separation process contains 500 kg C
2
H
5
OH and
500 kg H
2
O. There are two output streams from the process, one containing 460 kg
C
2
H
5
OH and 60 kg H
2
O. How much C
2
H
5
OH and H
2
O leave the system in the other
stream?
Solution:
The rst 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 27
500 kg C
2
H
5
OH
500 kg H
2
O
460 kg C
2
H
5
OH
60 kg H
2
O
W
3,1
kg C
2
H
5
OH
W
3,2
kg H
2
O
1
2
3
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, W
3,1
and W
3,2
. 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 2
Independent component mass balances - 2
Degrees of freedom 0
Since the degrees of freedom = 0, the system is dened in an unambiguous manner!
It is thus adequate to dene the equstions of the system forming the mass balance
equations in the form Input=Output:
MB C
2
H
5
OH : 500 = 460 +W
3,1
MB H
2
O : 500 = 60 +W
3,2
or
MB C
2
H
5
OH : 1 W
3,1
+ 0 W
3,2
= 500 460
MB H
2
O : 0 W
3,1
+ 1 W
3,2
= 500 60
In matrix notation AX = Y this becomes:
A =
_
1 0
0 1
_
, X =
_
W
3,1
W
3,2
_
, Y =
_
500 460
500 60
_
We solve this quite trivial system with Matlab as: Hence, the solution to the problem
is: W
3,1
= 40 kg and W
3,2
= 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.
W
1,1
kg C
2
H
5
OH
W
1,2
kg H
2
O
W
2,1
kg C
2
H
5
OH
W
2,2
kg H
2
O
W
3,1
kg C
2
H
5
OH
W
3,2
kg H
2
O
1
2
3
separations-
process
There are six unknown stream variable, and the independent information are W
1,1
=
500, W
1,2
= 500, W
2,1
= 460 and W
2,2
= 60.
Unknown stream variables 6
Independent component mass balances -2
Independent information -4
Degrees of freedom 0
The equations are:
MB C
2
H
5
OH : 500 = 460 +W
3,1
MB H
2
O : 500 = 60 +W
3,2
Info 1 : 500 = W
1,1
Info 2 : 500 = W
1,2
Info 3 : 460 = W
2,1
Info 4 : 60 = W
2,2
Or, with matrix notation AX = Y :
A =
_

_
1 0 0 0 0 0
0 1 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
_

_
, X =
_

_
W
3,1
W
3,2
W
1,1
W
1,2
W
2,1
W
2,2
_

_
, Y =
_

_
500 460
500 60
500
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 C
2
H
5
OH and
500 kg H
2
O. The output consists of two streams. One amounts to a total of 400
kg and contains 96% C
2
H
5
OH. Calculate the mass fractions of C
2
H
5
OH and H
2
O
respectively in the other stream.
Solution:
First, the process chart is drawn with the streams numbered 1-3 and the components
1-2.
500 kg C
2
H
5
OH
500 kg H
2
O
96%, W
2,1
C
2
H
5
OH
W
2,2
H
2
O
W
3,1
kg C
2
H
5
OH
W
3,2
kg H
2
O
1
W
2
=400kg
3
The next step is to carry out a degree of freedom analysis. They way the process chart
is drawn, there are four unknowns, W
2,1
and W
2,2
as well as W
3,1
and W
3,2
. However,
we have two information: the concentration constraint 0.96 = W
2,1
/(W
2,1
+ W
2,2
)
and the total constraint 400 = W
2,1
+W
2,2
.
Unknown stream variables 4
Independent component mass balances -2
Independent information -2
Degrees of freedom 0
The four equations are:
MB C
2
H
5
OH : 500 = W
2,1
+W
3,1
MB H
2
O : 500 = W
2,2
+W
3,2
Info 1 : 400 = W
2,1
+W
2,2
Info 2 : 0.96 = W
2,1
/(W
2,1
+W
2,2
)
0 = 0.04W
2,1
+ 0.96W
2,2
In matrix notation, AX = Y we have:
A =
_

_
1 0 1 0
0 1 0 1
1 1 0 0
0.04 0.96 0 0
_

_
, X =
_

_
W
2,1
W
2,2
W
3,1
W
3,2
_

_
, Y =
_

_
500
500
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 C
2
H
5
OH and
500 kg H
2
O. The output consists of two streams. The rst of these output streams
contains 96% C
2
H
5
OH, while the second amounts to a total of 600 kg. Calculate the
mass fractions of C
2
H
5
OH and H
2
O respectively in the second stream.
Solution:
First, the process chart is drawn with the streams numbered 1-3 and the components
1-2.
500 kg C
2
H
5
OH
500 kg H
2
O
96%, W
2,1
C
2
H
5
OH
W
2,2
H
2
O
W
3,1
kg C
2
H
5
OH
W
3,2
kg H
2
O
1
2
W
3
=600kg
Just as in the previous example, there are four unknown stream variables. In addition,
the concentration constraint 0.96 = W
2,1
/(W
2,1
+ W
2,2
) is the same as above, while
the total constraint is moved so that 600 = W
3,1
+W
3,2
.
Unknown stream variables 4
Independent component mass balances -2
Independent information -2
Degrees of freedom 0
The four equations are:
MB C
2
H
5
OH : 500 = W
2,1
+W
3,1
MB H
2
O : 500 = W
2,2
+W
3,2
Info 1 : 600 = W
3,1
+W
3,2
Info 2 : 0.96 = W
2,1
/(W
2,1
+W
2,2
)
0 = 0.04W
2,1
+ 0.96W
2,2
In matrix notation, AX = Y we have:
A =
_

_
1 0 1 0
0 1 0 1
0 0 1 1
0.04 0.96 0 0
_

_
, X =
_

_
W
2,1
W
2,2
W
3,1
W
3,2
_

_
, Y =
_

_
500
500
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.
C H A P T E R 3
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 dierent 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 comprehen-
sive 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 specic 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 nds 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 dened
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 Common types of units
3.2.1 Mixer
In a mixer, one or more streams are merged:
Mixer
For a mixer, the total output ux must equal the sum of the input uxes. This holds
regardless of whether the uxes 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 uxes must
be equal to the input ux. 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 W
1
with two components A and B is split into 3 output streams, W
2
W
4
3.2 Common types of units 34
one splitter constraint is that:
W
1,A
W
1,A
+W
1,B
=
W
2,A
W
2,A
+W
2,B
If one composition in W
1
is known, one can use this information to re-write the above
expression as:

1,A
=
W
2,A
W
2,A
+W
2,B
However, it is not possible to dene 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 dene 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 dene (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
=
W
2,A
W
2,A
+W
2,B
where
1,A
= 0.2
one may directly write the linear equation:
0.2 =
W
2,A
W
2,A
+W
2,B
or 0 = 0.8W
2,A
0.2W
2,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.
Splitter 1
W
1,A
W
1,B
2
3
W
2,A
W
2,B
W
3,A
W
3,B
Solution:
We try to formulate 4 equations:
MB A W
1,A
= W
2,A
+W
3,A
MB B W
1,B
= W
2,B
+W
3,B
Splitter constraint 1
W
1,A
W
1,A
+W
1,B
=
W
2,A
W
2,A
+W
2,B
Splitter constraint 2
W
3,A
W
3,A
+W
3,B
=
W
2,A
W
2,A
+W
2,B
Re-write the mass balances:
MB A W
3,A
= W
2,A
W
1,A
MB B W
3,B
= W
2,B
W
1,B
and substitute W
3,A
and W
3,B
into splitter constraint 2, which is then simplies as:
W
3,A
W
3,A
+W
3,B
=
W
2,A
W
2,A
+W
2,B
(W
1,A
W
2,A
)(W
2,A
W
2,B
) = W
2,A
(W
1,A
W
2,A
+W
1,B
W
2,B
)
(W
1,A
W
2,A
W
2
2,A
+W
1,A
W
2,B
+W
2,A
W
2,B
) =
W
2,A
W
1,A
W
2
2,A
+W
2,A
W
1,B
W
2,A
W
2,B
W
1,A
W
2,B
= W
2,A
W
1,B
Now, we re-write and simplify splitter constraint 1:
Splitter constraint 1 W
1,A
(W
2,A
+W
2,B
) = W
2,A
(W
1,A
+W
1,B
)
W
1,A
W
2,B
= W
2,A
W
1,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 36
3.2.3 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 specic process step will pass through it once more:
Recirkulationsstrm
One quanties the recirculation by a quantity called the recirculation ratio, dened
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 treat-
ment plant illustrated in Figure 1.1 includes a recirculation stream from the nitrica-
tion to denitrication 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 o:
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 economi-
cally favourable to clean as small a part of the sewage ow as possible. Thus, a sewage
ow bypass is introduced.
Calculate all the ows in the system, and determine how large should the part of
the sewage ow 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:
Totalt
Stream variables 11
MB (A and B, Splitter) - 2
MB (A and B, Separation) - 2
MB (A and B, Mixer) - 2
Info conc. of A in W
1
- 1
Info conc. of A in W
5
- 1
Info separation of A -1
Splitter constraint -1
Basis of calc. -1
Degrees of freedom 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: W
1,A
= W
2,A
+W
3,A
MB B Splitter: W
1,B
= W
2,B
+W
3,B
MB A Separation: W
2,A
= W
4,A
+W
6,A
MB B Separation: W
2,B
= W
4,B
MB A Mixer: W
3,A
+W
4,A
= W
5,A
MB B Mixer: W
3,B
+W
4,B
= W
5,B
Conc. of A in stream 1: W
1,A
= 0.1(W
1,A
+W
1,B
)
Conc. of A in stream 5: W
5,A
= 0.03(W
5,A
+W
5,B
)
Separation of A: W
4,A
= 0.1W
2,A
Splitter constraint (trick): W
3,A
= 0.1(W
3,A
+W
3,B
)
Basis of Calc.: W
1,A
+W
1,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:
MB A Splitter: W
1,A
W
2,A
W
3,A
= 0
MB B Splitter: W
1,B
W
2,B
W
3,B
= 0
MB A Separation: W
2,A
W
4,A
W
6,A
= 0
MB B Separation: W
2,B
W
4,B
= 0
MB A Mixer: W
3,A
+W
4,A
W
5,A
= 0
MB B Mixer: W
3,B
+W
4,B
W
5,B
= 0
Conc. of A in stream 1: 0.9W
1,A
0.1(W
1,B
= 0
Conc. of A in stream 5: 0.97W
5,A
0.03W
5,B
= 0
Separation of A: W
4,A
0.1W
2,A
= 0
Splitter constraint (trick): 0.9W
3,A
0.1(W
3,B
= 0
Basis of Calc.: W
1,A
+W
1,B
= 1000
This system of equations can also be written in matrix form AX = Y where A is the
matrix of coecients, 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:
A =
_

_
1 0 1 0 1 0 0 0 0 0 0
0 1 0 1 0 1 0 0 0 0 0
0 0 1 0 0 0 1 0 0 0 1
0 0 0 1 0 0 0 1 0 0 0
0 0 0 0 1 0 1 0 1 0 0
0 0 0 0 0 1 0 1 0 1 0
0.9 0.1 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0.97 0.03 0
0 0 1 0 0 0 1 0 0 0 0
0 0 0.9 0.1 0 0 0 0 0 0 0
1 1 0 0 0 0 0 0 0 0 0
_

_
, X =
_

_
W
1,A
W
1,B
W
2,A
W
2,B
W
3,A
W
3,B
W
4,A
W
4,B
W
5,A
W
5,B
W
6,A
_

_
, Y =
_

_
0
0
0
0
0
0
0
0
0
0
1000
_

_
It is easy to solve this system of equations with Matlab. The Matlabsolution m-le
is: and the nal results are: We now recall that the initial aim was to calculate
the fraction of the stream W
1
that could bypass the separation process while still
preventing the concentration of A in the output stream W
5
from getting too high.
The result is that the ratio between the steam can be:
W
3,A
+W
3,B
W
2,A
+W
2,B
=
20.6979 + 186.2807
79.3021 + 713.7193
= 0.2610
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:
Process Splitter Separation Mixer
Stream variables 5 6 5 6
MB (A and B, Process) - 2
MB (A and B, Splitter) - 2
MB (A and B, Separation) - 2
MB (A and B, Mixer) - 2
Info conc. of A in W
1
- 1 - 1
Info conc. of A in W
5
- 1 - 1
Info Separation of A -1
Splitter constraint - 1
Degr. of freed.w/o BoC 1 2 2 3
Basis of Calc. - 1 -1 -1 -1
Degr. of freed. w BoC 0 1 1 2
Thus, the Process may be solved for, but this does not allow us to calculate all uxes
in the system. If the Process is solved for, W
1,A
, W
2,A
, W
2,B
, W
5,A
and W
5,B
are
quantied. 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 W
1
and W
5
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:
Splitter Separation Mixer
Stream variables 6 5 6
MB (A and B, Splitter) - 2
MB (A and B, Separation) - 2
MB (A and B, Mixer) - 2
Known stream variables -2 -1 -2
Splitter constraint - 1
Degrees of freedom 1 2 2
Conclusion:
We may now conclude that we cannot get any further. Thus, only the Total and the
Process may be solved for.
C H A P T E R 4
Process calculations for reaction
systems
In many natural and engineered systems chemical reactions take place in which reac-
tants 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 dierent techniques to make process calculations in the
system with one or more chemical reactions. Some important new concepts are dened
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:
N
2
+ 3H
2
2NH
3
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 N
2
and 3 kg
H
2
form 2 kg NH
3
.
We can also note two important things happening for the above reaction system:
1. N
2
och H
2
are consumed while NH
3
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 dene the system in which the reaction takes place. If one
formulates mass balances for the components N
2
, H
2
, NH
3
they will be in a dierent
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 N
2
, H
2
, NH
3
: Input + Prod = Output (4.2)
The dierence, 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:
For components:

in
W
i,j
=

out
W
i,j
and

in
F
i,j
=

out
F
i,j
For elements:

in
W
i,j
=

out
W
i,j
and

in
F
i,j
=

out
F
i,j
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 compo-
nents 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 vari-
ous 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 dierently in nature.
4.2 Important concepts
4.2.1 Inert
An inert substance is one that does not react chemically. In many processes the
synthesis of NH
3
described above is a rare exception N
2
is an inert. N
2
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 43
4.2.2 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 dened as:
X =
_

_
Converted amount
Input ux
Input ux - Output amount
Input ux
1

output
F

input
F
Hence, if the output ux (

output
F) is equal to the input ux (

input
F) the conver-
sion is obviously X = 0 (i.e., the reactant has not reacted at all within the system).
On the contrary, if the output ux (

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 quanties 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
denition, 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 dene the yield if the limiting reactant R is converted to the
main product P. The stoichiometry of the reaction is:

R
R
P
P
and the yield is thus dened as:
Yield =

output
F
i,P
/
P

input
F
i,R
/
R
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 uxes of all products
have been determined, the selectivity is calculated as:
Selectivity =

output
F
main product

output
F
bi-products
4.3 The element mass balance method 44
4.3 The element mass balance method
Element mass balances are based on the fact that the molar ux of elements can be
expressed in terms of molar uxes of components. For example, if:
F
H
2
O
= 1
then
F
H
= 2 and F
O
= 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 rst, 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:
C
2
H
6
catalysis
C
2
H
4
+ H
2
C
2
H
6
1 Process
systemgrns
C
2
H
6
2
C
2
H
4
H
2
Determine all uxes (i.e., stream variables) in the system if the feed contains 100 mol
C
2
H
6
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 4
- Independent element MB - 2
- Independent information - 2
Degrees of freedom 0
4.3 The element mass balance method 45
Hence the system is uniquely dened. Using F to denote molar ux, 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 ux of 1 mole of C
2
H
6
obviously
carries 2 moles of C and 6 moles of H. The system of equations then becomes:
MB C : 2F
1,C
2
H
6
= 2F
2,C
2
H
6
+ 2F
2,C
2
H
4
MB H : 6F
1,C
2
H
6
= 2F
2,H
2
+ 6F
2,C
2
H
6
+ 4F
2,C
2
H
4
Info 1 : F
1,C
2
H
6
= 100
Info 2 : F
2,C
2
H
6
= (1 0.6)F
1,C
2
H
6
In matrix notation AX = Y we get:
A =
_

_
2 0 2 2
6 2 6 4
1 0 0 0
(1 0.6) 0 1 0
_

_
, X =
_

_
F
1,C
2
H
6
F
2,H
2
F
2,C
2
H
6
F
2,C
2
H
4
_

_
, Y =
_

_
0
0
100
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 com-
position 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 C
2
H
6
, C
2
H
4
andH
2
:
C
2
H
6
H
2
C
2
H
4
C
H
_
2 0 2
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:
2C
2
H
4
catalysis
C
4
H
8
4.3 The element mass balance method 46
is carried out over a catalytic bed.
C
2
H
4
1 Process
systemgrns
C
4
H
8
2
C
2
H
4
Make a degree of freedom analysis provided that the feed contains 100 mol C
2
H
6
and
the conversion is 60%.
Solution
The atomic matrix is:
C
2
H
4
C
4
H
8
C
H
_
2 4
4 8
_
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 3
- Independent element MB - 1
- Independent information - 2
Degrees of freedom 0
We can thus conclude that the reaction may actually be viewed as if:
2A
catalysis
A
2
where A = C
2
H
4
. 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 47
4.4 The component mass balance method
In 4.3 the method to solve all uxes 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, quantication of
the chemical reactions are essential.
When applying a component mass balance method, the degree of freedom analysis
is somewhat dierent compared to the element mass balance method. The reason
is that the chemical reactions need to be quantied 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 dened 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 = Output
F
in,A
= F
out,A
F
in,B
+ = F
out,B
As we see, will always be a positive number since F
in,A
> F
out,A
and F
in,B
< F
out,B
.
In order to generalize, let us consider more complex situation where one chemical
reaction takes place:
N
2
+ 3H
2
2NH
3
For the two reactants N
2
and H
2
, has a negative sign in front of it since Prod<0.
In addition, there is a stoichiometric constraint since the number of moles of H
2
that
are consumed is always 3 times greater than the number of moles of N
2
produced.
If we let denote the reaction parameter for the reaction in question and while
the stoichiometric coecients are denoted
i
we get:
=
_

N
2

N
2
=

N
2
1

H
2

H
2
=

H
2
3

N
2

NH
3
=

NH
3
2
(4.3)
Consequently, it is also true that:

N
2
=
N
2
,
H
2
=
H
2
,
NH
3
=
NH
3

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 dene one reaction parameter for each reaction. This is indeed very
useful since the ratio between calculated values of the dierent 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:
F
input,j
+
all
recations

i,j

i
= F
output,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 inde-
pendent 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 equi-
librium 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 reac-
tion matrix. This is formed by the the chemical reactions (as rows) and the stoichio-
metric coecients associated with the dierent 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 fol-
lowing chemical reactions occur:
Reaction 1 A 2B
Reaction 2 B C
Reaction 3 A 2C
Solution
Form the reaction matrix by setting up the chemical reactions as rows and the chemical
components involved as columns:

A

B

C
Reaction 1
Reaction 2
Reaction 3
_
_
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 inde-
pendent. 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 quantied when reaction sys-
tems are treated by means of mass balances. One way to look at this is:
1. Input terms are quantied by stream variables representing uxes into the sys-
tem.
2. Output terms are quantied by stream variables representing uxes out of the
system.
3. Prod terms are quantied by means of reaction parameters, one for each reac-
tion.
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 of (unknown) stream variables
+ Number of reaction parameters
- Number of independent component mass balances
- Number of 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
(C
2
H
6
) is dehydrated to ethene (C
2
H
4
) and H
2
. In the other, ethane reacts with
H
2
to form the biproduct methane (CH
4
):
Reaction 1 C
2
H
6
C
2
H
4
+ H
2
Reaction 2 C
2
H
6
+ H
2
2CH
4
The feed (i.e., the input ow of reactant gas) consist of 85% C
2
H
6
while the rest is
inert gas (N
2
). Calculate the magnitude of all output uxes, if the total conversion
of C
2
H
6
is 50.1%, while the yield with respect to formed ethene C
2
H
4
is 47.1%. In
addition, calculate the selectivity with respect to ethene relative to formed CH
4
.
Solution
The process chart becomes::
F
1,C
2
H
6
F
1,N
2
1 2
dehydro-
genering
F
2,C
2
H
6
F
2,N
2
F
2,C
2
H
4
F
2,H
2
F
2,CH
4
The next step is to form a reaction matrix. Although it is obvious that the reactions
are linearly independent (CH
4
only appears in one of the reactions), this can be
analyzed formally by calculating the rank of:
Reaction 1
Reaction 2
_
_

C
2
H
6

C
2
H
4

CH
4

H
2
1 1 0 1
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 7
+ Reaction parameters 2
- Independent MB - 5
- Independent info - 3
Degrees of freedom 1
We now introduce the basis of calculation F
1
= 100 mole in order to eliminate the
last degree of freedom. This is allowed since no other ux is given. The equations
4.4 The component mass balance method 51
become:
MB C
2
H
6
: F
1,C
2
H
6

1

2
= F
2,C
2
H
6
MB C
2
H
4
:
1
= F
2,C
2
H
4
MB H
2
:
1

2
= F
2,H
2
MB CH
4
: 2
2
= F
2,CH
4
MB Inert : F
1,N
2
= F
2,N
2
Conc. inF
1
: F
1,C
2
H
6
= 0.85(F
1,N
2
+F
1,C
2
H
6
)
Conversion : F
2,C
2
H
6
= (1 0.501)F
1,C
2
H
6
Yield : F
2,C
2
H
4
= 0.471F
1,C
2
H
6
Basis of calc. : F
1,N
2
+F
1,C
2
H
6
= 100
As usual, the equations are written in matrix notation AX = Y :
A =
_

_
1 0 1 0 0 0 0 1 1
0 0 0 1 0 0 0 1 0
0 0 0 0 0 1 0 1 1
0 0 0 0 1 0 0 0 2
0 1 0 0 0 0 1 0 0
0.15 0.85 0 0 0 0 0 0 0
0.499 0 1 0 0 0 0 0 0
0.471 0 0 1 0 0 0 0 0
1 1 0 0 0 0 0 0 0
_

_
,X =
_

_
F
1,C
2
H
6
F
1,N
2
F
2,C
2
H
6
F
2,C
2
H
4
F
2,CH
4
F
2,H
2
F
2,N
2

2
_

_
,Y =
_

_
0
0
0
0
0
0
0
0
100
_

_
and the Matlab solution becomes: Thus, the selectivity is F
2,C
2
H
4
/F
2,CH
4
= 40.03/5.1 =
7.849.
C H A P T E R 5
Reactor calculations
In the previous chapters, we have mainly dealt with mass balance problems to calcu-
late certain material ows over system boundaries, utilizing other ows, information
and balances. For reaction systems, we have also dened 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 param-
eter 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 denitrication basin in a sewage treatment plant and
dough which is in the process of becoming bread in a bakery. Despite the dierences
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 dierent. The lake is largely
mixed, although it has periodically stable stratication between the epilimnion and
the hypolimnion. On the contrary, the denitrication basin is horizontally but not
vertically mixed. If it were, oxygen would be mixed into the water and the denitrica-
tion 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 inow and
outow 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 inows or outows.
The chemical process occurring in the dough systems is primarily a transformation
of carbohydrates (sugar) to CO
2
and H
2
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 con-
stants.
The concept of residence time which describes how much that that ows 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
per unit volume and time. It has units of the type
mol
volume time
.
Kinetic rate equations are not simply made up; they are deduced on theoretical
grounds, while the rate coecients are determined experimentally. Table 5.1 gives an
overview of dierent common rate equations for dierent reaction types.
5.1.1 First order kinetics
The most simple is the rst 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:
r = kc
A
mole
volumeunit time
where r is the reaction rate, k is the rate coecient and c
A
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 coecients into consideration. For A C, the
stoichiometric coecient 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 coecient:
r = kc
A
r
A
=
A
r = 1kc
A
= kc
A
r
C
=
B
r = 1kc
A
= kc
A
The reaction is called rst order because the exponent on c
A
is 1, as c
A
= c
1
A
.
1
Gas phase reaction systems dier from liquid phase reaction systems in two ways: as the total
number of moles change due to the reaction, the pressure and the ow 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
Type Example r
A
Units
order
0 irrev. A B r
A
= k k :
mol
Lhr
1 irrev. A B r
A
= kc
A
k :
1
hr
1 rev. A B r
A
= k
f
c
A
+k
b
c
B
k :
1
hr
Pseudo 1 irrev. A + H
2
O B r = kc
A
k :
1
hr
2 irrev. 2A B r
A
= kc
2
A
k :
L
molhr
2 irrev. A + B C r
A
= kc
A
c
B
k :
L
molhr
2 rev. A + B C r
A
= k
f
c
A
c
B
+k
b
c
C
k :
L
molhr
Monod irrev. A
Enzyme

catalysis
B r =
kc
A
K +c
A
k :
mol
Lhr
; K :
mol
L
5.1.2 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 = kc
A
c
B
Since the stoichiometric coecients are -1 or 1, we get the following rate equations
for A,B and C:
r
A
= kc
A
c
B
r
B
= kc
A
c
B
r
C
= kc
A
c
B
The reaction is called second order because the sum of exponents on c
A
and c
B
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 coecient will dier with
reaction order. The principle is that the units of the coecient must match the
reaction order (expressed as c
A
, c
2
A
, c
A
c
B
, etc.) so that the unit of r is correct.
A special case is the biological reactions described by Monod kinetics or Michaelis-
Menten 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-o is gradual,
but aected by the the value of the half-rate coecient, 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
dierence 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
dened as:
r = 0
_

_
0 = k
f
c
A
k
b
c
B
k
f
k
b
=
c
B
c
A
= K
Hence, chemical equilibrium coecients can be viewed as the ratio between the coef-
cients of the forward and the backward rate coecients.
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:
A
k
1
B
k
2
C
Here , B is the intermediate in the reaction AC. If we break this reaction down
into its two rst order reactions, given index 1 and index 2 respectively, we can
deduce the rate equation for each component. First we recognize that:
r
1
= k
1
c
A
r
2
= k
2
c
B
5.2 Mean residence time and reaction time 57
As we see, the stoichiometric coecients are either -1 (for reactants) or 1 (for prod-
ucts). Hence the rate equations for A, B and C become:
r
A
= r
1
= k
1
c
A
r
B
= r
1
r
2
= k
1
c
A
k
2
c
B
r
C
= r
2
= k
2
c
B
It is thus possible to express the equilibrium constant in terms of a ratio between two
rate coecients, or c
A
and c
B
.
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
k
1
B
A
k
2
C
Then, if:
r
1
= k
1
c
A
r
2
= k
2
c
A
we get:
r
A
= r
1
r
2
= (k
1
+k
2
)c
A
r
B
= r
1
= k
1
c
A
r
C
= r
2
= k
2
c
A
One example of parallel reactions in nature, is the reaction of nitrous oxide, NO
2

, in
the formation of nitric acid, HNO
3
, from reactions either with OH

or NO
3

:
NO
2

+ OH

HNO
3
NO
2

+ NO
3

N
2
O
5
+ H
2
O 2HNO
3
5.2 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 dier-
ences between them was that while two (the lake and the basin) both has inow and
outow, 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, m
3
, etc.)
The volumetric ow rate into or out of a reactor: Q (liters/s, m
3
/hour, etc.)
5.2 Mean residence time and reaction time 58
The average residence time for the reactor: =
V
Q
(sec, hours, etc.)
Technical systems, the ow rate is often constant. In natural systems however, the
inow to lakes or the ow in watercourses can vary greatly. Then the ow Q is not
a constant value, but may vary 1-2 orders of magnitude between high ow periods,
such as when snow melts in as compared to dry periods in the summer.
If the average volumetric ow rate is to be measured, we can distinguish three
dierent cases. First, the ow can be constant. Second, the ow may be variable
and the measurements may be evenly or unevenly spaced in time. Finally, the ow
may be variable but the measurements continuous. For these three cases, the average
volumetric ow rate is determined as:

Q =
_

_
Q constant ow
1
n
n

i=1
Q
i
n evenly spaced measurements
1
t
_
t
0
Q(t)dt continuous measurements
(5.1)
Based on the calculation of the average volumetric ow

Q, the residence time is
calculated from the volume as V/

Q. An example of a calculation with variable ow


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 ow is continuous. For a
closed system however, which has no volumetric inow or outow, 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 inow water from the
stream Aborrbcken. In the environmental monitor program, the ow rate is mea-
sured 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.
Month
Flow
(m
3
day
1
)
January 222
February 609
March 3327
April 2730
May 2870
June 693
July 33
August 93
September 441
October 1461
November 1147
December 435
Solution
The following calculations are made in Matlab, resulting in an average residence time
of water in the lake of = 1.2 years:
1 2 3 4 5 6 7 8 9 10 11 12
0
500
1000
1500
2000
2500
3000
3500
U
p
p
m

t
t

f
l

d
e

(
m
3
/
d
y
g
n
)
Mnad
Medelflde
5.3 Reactor models 60
5.3 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 aects the concentrations in the system, which in turn aect
the rate of chemical reactions. Here, we will only consider macroscopic mixing, not
diusive 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-ow reactor
As the drawings illustrate, these types of chemical reactor all dier 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-ow reactor is open and not mixed.
5.3.1 The dierential mass balance
All of the ideal reactor models can be characterized by the same mass balance equa-
tion, the dierential (component) mass balance
2
. As reaction systems are of prime
interest, the component mass balance will always be expressed in terms of molar uxes
of components, denoted F
A
, F
B
, etc. Furthermore, the uxes will be given in molar
uxes 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 = Output + Acc
F
in
+rV = F
out
+
d(cV )
dt
mole
unit time
where F is the molar ux (mole/unit time), c is the molar concentration (mole/unit vol-
ume), Q is the volumetric ow 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 dierences 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 uxes across the re-
actor system boundaries, F
in
and F
out
, are given by the product of the concentration
c and the volumetric ow rates Q
in
and Q
out
respectively. If the ow rate is constant,
Q
in
= Q
out
= Q.
The Prod-term: The Prod-term quanties how much of a substance that is pro-
duced 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 r
A
= kc
A
, we get:
Prod = kc
A
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:
Prod
j
=
all i

i=1

i,j
r
i
V
Thus, in terms of calculations there are three actions related to the Prod-term:
1. Identify all reactions.
2. Dene the corresponding kinetic rate equations.
3. Include the kinetic rate equations in the Prod-term with the right stoichiometric
coecients and signs (+ or -).
The Acc-term: The Acc-term quanties 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 dierential of cV , we get the Acc-term as:
Acc =
d
dt
cV =
_

_
V
dc
dt
+c
dV
dt
if V varies with time
V
dc
dt
if V is constant
In summary: For a component A which reacts in one single reaction in a reactor
system where the ow rate and the volume is constant, the dierential mass balance
becomes:
Q
in
c
in,A
+r
A
V = Q
in
c
out,A
+
d(c
A
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. Dene the ow and mixing conditions in terms of an ideal reactor model.
2. Determine the reactor-specic parameters.
3. Dene the kinetic equation(s) and rate constant(s).
4. Combine the reactor model and the kinetic equations(s) in the dierential 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:
c
ut
Q
ut
c
in
Q
in c

r V
In + Prod = Out + Acc
Q
in
c
in
+rV = Q
out
c
out
+
d
dt
cV
At steady state, when Acc=0 och Q
in
= Q
out
= Q the simplied model is:
In + Prod = Ut (5.3)
Qc
in
+rV = Qc
out
(5.4)
Final remark: Never forget that the concentration in the reactor c is the same as c
out
!
5.4 The ideal completely stirred tank reactor, CSTR 63
Example 52: CSTR with a 1st order irreversible reaction
Problem
Calculate output stream uxes from a steady-state CSTR in which a 1st order reaction
occurs:
AB r = kc
A
= 0.5c
A
mole m
3
min
provided that the volumetric ow rate is 1.25 m
3
min
1
, the reactor volume is 5 m
3
,
and the input stream concentration of the reactant A is 2 mole m
3
while the input
concentration of B is 0.2. mole m
3
.
Solution
First we select the appropriate model: the steady-state CSTR. For a CSTR the
dierential mass balances are given by:
Q
in
c
in,A
+r
A
V = Q
out
c
out,A
Q
in
c
in,B
+r
B
V = Q
out
c
out,B
Since Q
in
= Q
out
= Q, c
out,A
= c
A
and c
out,B
= c
B
the equations can be simplied
to:
Qc
in,A
+r
A
V = Qc
A
Qc
in,B
+r
B
V = Qc
B
All of the reactorspecic parameters are given in the problem denition. The kinetic
rate equations for rst order reaction with respect to the reactant A and the product
B are:
_
r
A
=
A
r = 1r = kc
A
r
B
=
B
r = 1r = kc
A
We can now combine the kinetic equations and the dierential mass balance model.
The combined equations are:
_
_
_
Qc
in,A
= +kc
A
V +Qc
A
Qc
in,B
= kc
A
V +Qc
B
or, if we introduce the residence time =
V
Q
:
_
_
_
c
in,A
= c
A
(1 +k) , c
A
=
c
in,A
1 +k
c
in,B
= kc
A
+c
B
, c
B
= c
in,B
+kc
A

(E52.5)
Equation E52.5 allows us to solve directly for c
A
. Before inserting numbers, we
recognize that =
5
1.25
= 4 minutes, so that:
c
A
=
2
1 + 0.54
= 0.667 mole m
3
and thus, c
B
= 0.2 + 0.5 4 0.667 = 1.533 mole m
3
.
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 rst step is to re-write E52.5 :
_
c
in,A
= c
A
(1 +k)
c
in,B
= kc
A
+c
B
with all terms including c
A
and c
B
on the left-hand side:
_
(1 +k)c
A
+ 0 c
B
= c
in,A
kc
A
+ 1 c
B
= c
in,B
This system of equations can be written in the form AX = Y :
A =
_
1 +k 0
k 1
_
, X =
_
c
A
c
B
_
, Y =
_
c
in,A
c
in,B
_
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:
A
1
2
B r = kc
A
1.2
= 0.5c
A
1.2
mole m
3
min
Solution
Here, the equation E52.5 will contain a non-linear term:
_
_
_
c
in,A
= c
A
+kc
1.2
A
c
in,A
= c
A
(1 +kc
0.2
A
)
c
in,B
=
1
2
kc
1.2
A
+c
B
Nevertheless, this equation can also be written in the form AX = Y :
A =
_
1 +kc
0.2
A
0

1
2
k 1
_
, X =
_
c
A
c
B
_
, Y =
_
c
in,A
c
in,B
_
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 dene the equation, one creates a Matlab m-le with the
function: and gets the solution: Thus c
A
= 0.6989 mole m
3
and c
B
= 0.8506
mole m
3
.
5.5 The ideal batch reactor 66
5.5 The ideal batch reactor
The ideal batch reactor consists of a closed system (i.e., a reactor, which has neither
inow or outow). In the reactor, the reactions cause the concentrations of reactant
and products to change with time. One cannot speak of a batch reactor at steady-
state.
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 dierent substances in the same
reactor. Another example of the use of batch reactors is biotechnological fermentation
reactors used to produce biomaterials using specic enzymes.
In the environment one can also nd 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 SO
2
to H
2
SO
4
. But the
most important example is when the input ow to a normally open system suddenly
decreases signicantly 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 inow 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:
c

r V
Prod = Acc
rV = V
dc
dt
or, if the volume V is eliminated:
dc
dt
= r (5.5)
At this point, one can see that there are two ways to use the above reactor model,
the dierential view and the integral view. In both cases, we must be aware that the
boundary conditions are not related to input and output uxes (i.e. Qc
in
or Qc
out
).
Instead, the boundary conditions relate to time. The relevant boundary condition is
thus the initial condition, or inititial concentration c
o
, which denotes c at time t=0.
With the dierental view we solve equation 5.5 to get:
dc
dt
= r c(t) from t=0 to t=t
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
dt
= r t required to change c from c
0
to c
5.5 The ideal batch reactor 67
This method is more useful if one asks, How long time does it take to reach concen-
tration c.
In more mathematical terms, the two cases can be expressed as:
Dierential :
dc
dt
= r
_
t = 0, c = c
0
t = t, c = c
(5.6)
Integral :
_
t
0
dt =
_
c
c
0
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 rst order reaction.
AB r = kc
A
= 0.5c
A
mole m
3
min
1
if the initial concentration, c
0,A
is 2 mole m
3
.
Solution
First we identify the correct mass balance equation for the reactor in question:
Acc = Prod
dc
A
dt
= r
A
and derive the correct kinetic rate equation for A:
r
A
=
A
r = kc
A
Combining the mass balance equation and the kinetic equation results in the reactor
model:
dc
A
kc
A
= dt
1
k
_
c
A
c
0,A
dc
A
c
A
=
_
t
0
dt
_
c
A
c
0,A
d lnc
A
= k
_
t
0
dt
[lnc
A
]
c
A
c
0,A
= k[t]
t
0
ln
c
A
c
0,A
= kt
c
A
= c
0,A
e
kt
(E55.8)
With numbers inserted we get c
A
= 2e
0.54
= 0.2707 mole m
3
.
5.5 The ideal batch reactor 68
Example 56: Dierential, numerical solution of Example 55
Problem
For the above reaction and reaction system, calculate by means of numerical integra-
tion 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
dt
= r
dc
A
dt
= kc
A
This dierential 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 c
0,A
= 2. First the function that should
be integrated is created: Thereafter the function is simulated. To nd the actual
nal value of c
A
one can make Matlab show the lowest value that cA has attained
during the calculation. The result becomes:
0 0.5 1 1.5 2 2.5 3 3.5 4
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
tid (minuter)
k
o
n
c
e
n
t
r
a
t
i
o
n
e
n

a
v

A

(
m
o
l
/
m
3
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 integra-
tion the time it takes for the concentration of A (c
A
) to decrease from 2 to 0.2707
mol m
3
. Use the integration method quad8 inMatlab.
Solution
The analysis above gave as intermediate that:
_
t
0
dt =
_
c
A
c
0,A
dc
A
kc
A
We can now form the function that should be integrated, and dene it in the m
le ex56: Thereafter the function in ex56 is integrated from c
A
= c
0,A
= 2 to
c
A
= 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:
A
r
1
B
r
2
C
A
r
3

r
4
C
Simulate the system between t = 0 to t = 10 with the initial conditions c
0,A
=
1, c
0,B
= 0andc
0,C
= 0. The stoichiometric coecients lead to the following rate
equations for A, B and C:
r
1
= 1c
A
r
2
= 0.4c
B
r
3
= 0.1c
A
r
4
= 0.2c
C
Model
The kinetic equations are combined with the mass balance equation for the batch
reactor to give reactor model equations:
MB A
dc
A
dt
= r
1
r
3
+r
4
= c
A
0.1c
A
+ 0.2c
C
MB B
dc
A
dt
= r
1
r
2
= c
A
0.4c
B
MB C
dc
C
dt
= r
2
+r
3
r
4
= 0.4c
B
+ 0.1c
A
0.2c
C
Numerical solution and answer
0 1 2 3 4 5 6 7 8 9 10
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
koncentrationen av A
koncentrationen av B
koncentrationen av C
5.6 The ideal plug-ow reactor, PFR 71
5.6 The ideal plug-ow 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-ow reactor is not mixed at all in the direction of ow. Perpendic-
ular to the direction of ow, however, it is perfectly mixed. This implies that all uid
that enters a plug-ow reactor moves forward at the same velocity. One can view the
reactor in several ways, but one useful conceptual model is that the uid 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 dierent 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-ow reactor, the smallest unit that is homo-
geneous is the individual slice. This volume element is sometimes called control
volume. The volume of the control volume is dierentially small, so the volume in
question is dV .
When a reaction occurs within dV , the molar ux will change by a dierential
amount. Therefore, if the input ux to dV is F, the output molar ux will be
F +dF. If the ow 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 dierential mass balance equation for a volume
dV within a (steady-state) plug-ow reactor as:
c
in
Q
dV
r
c+dc c
c
ut
Q
V
Input + Prod = Output
F +rdV = F +dF
Qc +rdV = Q(c +dc)
or, after dividing with Q:
rdV
1
Q
= dc , or rd = dc (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 ow.
We now realize that, just as for the batch reactor, the mass balance equation may
either be written in dierential or integral form:
Dierential :
dc
dV
=
r
Q
_
V = 0, c = c
0
V = V, c = c
(5.9)
Integral :
_
V
0
dV
Q
=
_
c
c
0
dc
r
(5.10)
5.6 The ideal plug-ow 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 innitely 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 correspond-
ing equations for the batch reactor, as shown in equation 5.6. This becomes even
clearer if we substitute V/Q with :
Dierential :
dc
d
= r
_
= 0, c = c
0
= , c = c
(5.11)
Integral :
_

0
d =
_
c
c
0
dc
r
(5.12)
If one, such in Example 55 solves the mass balance model analytically for the
rst order reaction AB, one gets:
r
A
= kc
A
c
A
= c
0,A
e
k
V
Q
= c
0,A
e
k
(5.13)
Obviously, the PFR and the batch reactor are modeled in exactly the same way,
the only dierence 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 gure 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 concen-
trations. By introducing the conversion, it is possible to express all concentrations of
reactants in terms of input concentrations, c
in
so that c = c
in
(1 X). If we dene
5.6 The ideal plug-ow reactor, PFR 73
the conversion of A as X
A
, the mass balance over the element dV in a PFR becomes:
In + Prod = Out
Qc
in,A
(1 X
A
) +r
A
dV = Qc
in,A
(1 X
A
dX
A
)
Combined with the kinetic equation for a rst order reaction r
A
= kc
in,A
(1 X
A
)
we get:
Qc
in,A
(1 X
A
) kc
in,A
(1 X
A
)dV = Qc
in,A
(1 X
A
dX
A
)
k(1 X
A
)
dV
Q
= dX
A
At this point one has to substitute variables, i.e. = (1 X
A
) so that:
d = 1dX
A
_
V
0
dV
Q
=
1
k
_

1
d

1
Q
[V ]
V
0
=
1
k
[ln]

1
Since ln( = 1) = 0 one gets, with = (1 X
A
) re-inserted:
=
ln(1 X
A
)
k
This equations can be seen as a normalized inverse of equation 5.8, which gave c
A
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 mole
0.5
m
1.5
min
1
Steam water ow Q : 10 m
3
s
1
Steam cross section area A r 50 m
2
Concentration of A at the plant: 0.02 mole m
3
Solution
The nature of the ow 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-ow 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
Q
dV = dc
rA
Q
dL = dc
Combined with the rate equation r
A
= kc
0.5
A
we get:
kc
0.5
A
A
Q
dL = dc
A
In this case, the conversion of A, X
A
should be 90%. We can then chose to write the
concentrations c
A
in terms of c
in,A
and the conversion so that dc
A
= c
in,A
dX
A
:
kc
0.5
in,A
(1 X
A
)
0.5
A
Q
dL = c
in,A
dX
A
kc
0.5
in,A
(1 X
A
)
0.5
A
Q
dL = dX
A
We will now solve the problem from a dierential as well as an integral perspective:
Dierential :
dX
A
dL
= kc
0.5
in,A
(1 X
A
)
0.5
A
Q
_
L = 0, X
A
= 0
L = L, X
A
= 0.9
Integral :
_
L
0
dL =
Qc
0.5
0,A
kA
_
0.9
0
dX
A
(1 X
A
)
0.5
5.6 The ideal plug-ow reactor, PFR 75
We let Matlab solve the integral form with quad8. The Matlab m-le 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 dierential form, as simulated with the
command:
ode45(ex58diff,[0 3000],[0]):
0 500 1000 1500 2000 2500 3000
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
o
m
s

t
t
n
i
n
g

a
v

f

r
o
r
e
n
i
n
g
e
n
strcka nedstrms utslppspunkten (m)
C H A P T E R 6
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 ow. In all cases,
except 0th order reactions, the mixing conditions aect 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 specic 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 sucient
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 ow
can occur along a stretch of river. Windings, connections and bay islands may aect
the water ow 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 dierent substances in many ways. As an exam-
ple, 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 sh, 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 non-
ideal 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 nds 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 inow 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 uid in the system
when added.
6.2 Residence-time distributions 78
6.2.1 The normalized RTD E(t)
If you add a certain amount of tracer at a given moment (instantaneously) dierent
tracer particles will leave the system at dierent 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 rst 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, M
S
, 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 reect 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 innitely long time, all the tracer will have left the system. Mathe-
matically, we can express it as the sum of the amount of substance that comes out
from t = 0 to t = :
M
s
=
_

0
Q(t) c(t)dt (6.1)
where
M
s
= amount of tracer mass
Q = volumetric ow rate volume time
1
c(t) = tracer concentration mass volume
1
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 mea-
sured concentration c(t). In this context, normalizing c(t) means that the integral of
c(t) with time becomes independent of M
S
.
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)
Q
M
s
=
c(t)
_

0
c(t)dt
1
time
(6.2)
From Equation 6.2 it follows that
_

0
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 79
c=0
Q
c
in=0
Q

c V
c=c
t=0
Q
c
in=hg
Q

c V
c < c
t=0
Q
c
in=0
Q

c V
c << c
t=0
Q
c
in=0
Q

c V
t < 0
t = 0
t > 0
t >> 0
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 =
_
t
0
E(t)dt (6.4)
The fraction that stays longer than time t in the system =
_

t
E(t)dt(6.5)
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 M
s
and the volume of the reactor is V , the
tracer concentration c(0) in the system will be:
c(0) =
M
s
V
To calculate E(t) we must rst 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 = Output + Acc
Qc
in
+rV = Qc +
d(cV )
dt
mass
unit time
6.2 Residence-time distributions 80
0.0
c(0)
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
tid/medeluppehllstid
c(t)
Figur 6.3: c(t) for a CSTR. The value of c(t) for t = is indicated.
where Q is the volumetric ow into the reactor. However, since there is no tracer
in the input ow (the tracer is added as a pulse) and since tracers do not react (by
denition), we get c
in
= 0 and r = 0. hence, the mass balance is reduced to:
d(cV )
dt
= Qc (6.6)
Since V/Q is the mean residence time we get:
dc =
1

dt (6.7)
_
c
c(0)
dc
c
=
1

_
t
0
dt (6.8)
ln
c
c(0)
=
t

(6.9)
c = c(0)e
t/
(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 ushed out of the reactor according to an
exponential process (Figure 6.3).
6.2 Residence-time distributions 81
0.0
0.2/
0.4/
0.6/
0.8/
1/
0.0 0.5 1.0 1.5 2.0 2.5 3.0
E(t) fr CSTR
t / tid/medelupphllstid
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) =
c(t)
_

0
c(t)dt
=
c(0) e
t/
c
t=0
_

0
e
t/
dt
=
e
t/
_

0
e
t/
dt
(6.11)
Since the primitive function to e
t/
is e
t/
, we simply get:
E(t) =
e
t/
[e
t/
]

0
=
e
t/

(6.12)
Figure 6.4 shows E(t) for the ideal CSTR. The gure indicates the value of E(t)
for t = , which is e
1
/ 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 time
1
. Doing this, we must get:
=
_

0
tE(t)dt (6.13)
1
In more advanced literature, one speaks about the "rst moment", the moment that the residence
time distribution creates relative to the y-axis.
6.2 Residence-time distributions 82
0.0
0.2
0.4
0.6
0.8
1.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
tid/medeluppehllstid
F(t)
E(t)
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:
_

0
tE(t)dt =
_

0
t
e
t/

dt =
1

_

0
te
t/
dt =
1

[
2
(
te
t/

e
t/
)]

0
=
=

2

(0 (1)) =
6.2.5 The F(t)-distribution
The equations 6.4 and 6.5 postulate how to calculate the proportion of uid elements
or a trace element that comes with them that have left the system after a specied
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) =
_
t
0
e
t/

dt =

[e
t/
]
t
0
= (e
t/
(1)) = 1 e
t/
(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 ow of water through an articial pond in a stream, an inert tracer is
added as a pulse in the inuent. In a natural system, this is a useful method, because
the ow often cannot be determined by direct measurement. The euent concentra-
tion is measured as follows:
Time (min) c(t)
0 0
2 1.1
4 2.9
6 4.8
8 5.9
10 6.2
12 5.8
14 5.1
16 4.3
18 3.7
24 2.2
30 1.2
40 0.4
50 0.2
0
1
2
3
4
5
6
7
0 10 20 30 40 50 60
Sprmneskoncentration
Tid (minuter)
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 measure-
ments. Compared to errors caused by eld measuring errors however, the integration
error is normally rather small. The time interval t around a measurement point i
has been dened as:
t
i
=
1
2
((t
i+1
t) (t i 1) =
1
2
(t
i+1
t
i1
)
Graphically, this means that:
6.2 Residence-time distributions 84
0
1
2
3
4
5
6
7
0 10 20 30 40 50 60
Sprmneskoncentration
Tid (minuter)
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 t (min) c
i
t
i
c
i
t
i
E
i
tE
i
t
i
0 0 0 2 - - -
1 2 1.1 2 2.2 0.0095 0.038
2 4 2.9 2 5.8 0.0250 0.200
3 6 4.8 2 9.6 0.0414 0.497
4 8 5.9 2 11.8 0.0509 0.815
5 10 6.2 2 12.4 0.0535 1.071
6 12 5.8 2 11.6 0.0500 1.202
7 14 5.1 2 10.2 0.0440 1.233
8 16 4.3 2 8.6 0.0371 1.188
9 18 3.7 4 14.8 0.0319 2.300
10 24 2.2 6 13.2 0.0190 2.736
11 30 1.2 8 9.6 0.0103 2.487
12 40 0.4 10 4 0.0034 1.382
13 50 0.2 10 2 0.0017 0.864

i
=115.8

i
= =16.0
The mean residence time is thus 16.0 minutes.
6.3 Non-ideal reactor models 85
6.3 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 condence.
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 uid 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 uid elements that enter the CSTR in series system
will rst mix into the rst CSTR, then ow 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 pro-
nounced. When the number reaches 30, almost all uid 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 exam-
ple, let us consider a 1st order reaction A B. For a CSTR, the output concentration
of THE reactant is (Equation E52.5):
c
out,A
= c
in,A
1
1 +k
Let us introduce the notation that the c
out,A
from the rst CSTR in a series is c
1,A
,
the output from the second CSTR is c
2,A
, etc. In that case, the input concentration
to the second CSTR will be c
1,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:
c
1,A
= c
in,A
1
1 +k/2
c
2,A
= c
1,A
1
1 +k/2
6.3 Non-ideal reactor models 86

c
1
V
tot
/2
c
in
Q

c
2
V
tot
/2
c
Q
0.0
0.5
1.0
1.5
2.0
2.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0
E(t)
tid/total medelupphllstid
ideal tank
ideal tub
n=2


c
1
V
tot
/5
c
in
Q
c
Q

c
2
V
tot
/5

c
3
V
tot
/5

c
4
V
tot
/5

c
5
V
tot
/5 0.0
0.5
1.0
1.5
2.0
2.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0
E(t)
tid/total medelupphllstid
ideal tank
ideal tub
n=5

c
1
V
tot
/10 c
2
V
tot
/10 c
3
V
tot
/10 c
4
V
tot
/10 c
5
V
tot
/10
c
in
Q
c
Q
c
6
V
tot
/19 c
7
V
tot
/10 c
8
V
tot
/10 c
9
V
tot
/10 c
10
V
tot
/10
0.0
0.5
1.0
1.5
2.0
2.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0
E(t)
tid/total medelupphllstid
ideal tank
ideal tub
n=10

c
1
V
tot
/30 c
2
V
tot
/30 c
3
V
tot
/30 c
4
V
tot
/30 c
5
V
tot
/30
c
in
Q
c
Q
c
26
V
tot
/30 c
27
V
tot
/30 c
28
V
tot
/30 c
29
V
tot
/30 c
30
V
tot
/30
0.0
0.5
1.0
1.5
2.0
2.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0
E(t)
tid/total medelupphllstid
ideal tank
ideal tub
n=30

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


6.4 The segregation model 87
or, since c
2,A
= c
out,A
;
c
out,A
= c
in,A
(
1
1 +k/2
)
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:
c
out
= c
in
_
1
1 +
k
n
_
n
(6.15)
Example 62: Series with an innite number of CSTR
From Equation 6.15, it follows that the output concentration of the reactant is:
c
out
= c
in
_
1
1 +
k

From this expression, it is not obvious how c


out
will be expressed. However, one
dimensional mathematics teaches us that:
lim
x
_
1
1 +
k
x
_
x
= e
k
If we apply this boundary value to an innite number of CSTRs in series we get:
c

= c
in
lim
n
_
1
1 +
k
n
_
n
= c
in
e
k
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 uid 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 dierent 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 con-
centrations 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 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 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 c
out,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 rst 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
c
A
= c
0,A
e
kt
We must now recall that each tiny, segregated batch reactor has the concentration
c
in,A
at the moment that they enter the reactor. Thus, c
0,A
= c
in,A
. According to
the segregation model, the output from the actual, non-ideal reactor is calculated as:
c
out,A
= c
in,A
_

0
e
kt
E(t)dt (6.16)
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 minutes
1
.
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:
c
A
c
in,A
=
_

0
e
kt
E(t)dt
where E(t) is known. We can then make the following calculations.
i t (min) E
i
c
A,t
c
in
c
A,t
c
in
E
i

i
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

i
= 0.283
For the CSTR, the conversion is:
c
c
in
=
1
1 +k
=
1
1 + 0.116
= 0.385
For the PFR, the conversion is:
c
c
in
= e
k
= e
0.116
= 0.202
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 ow 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
ow patterns, chemical reactions and other phenomena such as adsorption.
In the following chapter, the dispersion model will be used as a basis calculating
how dierent substances are transported in porous media.
C H A P T E R 7
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 reaktormod-
elellerna. Vad betrar de dynamiska egenskaperna kan man notera att satsreaktorn
aldrig r stationra, med mindre n att ingen kemisk reaktion sker i den. Fr tubreak-
torn gller att den kan vara instationra, men att det medfr att viktiga tillstns s-
som koncentration av reaktant och produkt varierar i bde tid och rum. Instationra
tubreaktorer mste drfr modelleras med partiella dierentialekvationer. 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 era 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 exempli-
eras i Exempel 71, och i vningsexempel.
Tabell 7.1: Sammanstllning ver viktiga karaktristika fr de ideala reaktor-
modellerna sats, tank och tubreaktorerna.
Satsreaktorn Tankreaktorn Tubreaktorn
Transport ver
systemgrns
Sluten ppen ppen
Makroskopisk
omblandning
Omblandad Omblandad Ej omblandad
Minsta homogena
volymselement
Hela volymen V Hela volymen V
Dierentiellt
volymselement dV
Stationr
materialbalans
Aldrig stationr In+Prod = Ut In+Prod = Ut
Instationr
materialbalans
Prod = Ack In+Prod = Ut+Ack In+Prod = Ut+Ack
Matematisk form
p instationr
materialbalans
Kopplade ordinra
di.ekvationer
Kopplade ordinra
di.ekvationer
Kopplade partiella
di.ekvationer
91
7.1 Svar p ndring i ingngskoncentration 92
7.1 Svar p ndring i ingngskoncentration
Fr att illustrera hur CSTR kan svara p ndringar i nyckelstorheter som ingngskon-
centrationen av reaktant, skall vi betrakta en CSTR dr det sker en kemisk reaktion:
AB r = kc
A
, k = 1
Systemet kan beskrivas med fljande gur:
[A]
in,t0
Q
in
[A]
t0
[B]
t0
V
[A]
t0
[B]
t0
Q
ut
[A]
in,t0
Q
in
[A]
[B]
V
[A]
[B]
Q
ut
Som beskrivits i ekvation 52 och ekvation E52.5 kan koncentrationen c
A
i en sta-
tionr CSTR med en irreversibel 1:a ordningens reaktion berknas med ekvationen:
c
A
c
A,in
=
1
1 +k
I ovanstende system k blir allts c
A
= 0.5, eftersom c
A,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 ingngskon-
centrationen ndras till c
A,in
= 1.9 vid t = 2. Som vntat stiger c
A
till 0.95 efter en
tid. Man kallar strningen fr en stegstrningen och reaktorns svar p detta fr ett
stegsvar.
Nsta bild visar c
A
om ingngskoncentrationen tillts ka under en begrnsad tid.
Hr ser man hur koncentrationen i reaktorn (och i utdet) 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 in-
gngskoncentration. Man kan se att svaret blir att ven c
A
uppvisar ett sinusformat
beteende, men att det nns en amplitudskillnad och en fasfrskjutning mellan c
A,in
och c
A
.
Dessa exempel r i detta fall bara illustrationer av olika dynamiska frlopp och
beteenden. Det nns 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
0 1 2 3 4 5 6 7 8 9 10
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
tid
k
o
n
c
e
n
t
r
a
t
i
o
n
inkoncentration
utkoncentration
0 1 2 3 4 5 6 7 8 9 10
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
tid
k
o
n
c
e
n
t
r
a
t
i
o
n
inkoncentration
utkoncentration
0 1 2 3 4 5 6 7 8 9 10
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
tid
k
o
n
c
e
n
t
r
a
t
i
o
n
inkoncentration
utkoncentration
Figur 7.1: Exempel p hur en tankreaktor kan svara p en ndring i ingende
koncentration av reaktant.
7.2 Arbetsmetodik 94
7.2 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 dierentialekvationer som uppkommer eftersom Acktermen nns
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 dierentialekvationerna.
De olika momententen r:
1. Beskriv blandningsfrhllandena i reaktorn med en reaktormodell och system-
grnser
2. Stll upp materialbalanser fr alla (oberoende) komponenter, och bestm de
reaktorspecika parametrarna
3. Beskriv de kemiska reaktionerna med en kinetisk modell och hastighetskonstanter
4. Kombinera reaktormodell och kinetisk modell i form av en dierentiell materi-
albalans
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 dierentialekvationerna, 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 dierentialekvationer p rtt stt. Simuleringen skall givetvis ge som resultat att
de inga koncentrationer frndras relativt de stationra betingelserna. Om kurvorna
blir krokiga nns 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 ver-
raskningar, 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 = kc
A
Normala driftsbetingelser r V = 5, Q = 1, k = 2 och c
in,A
= 100 (antag att enheterna
r harmoniserade....). Simulera vad som hnder om koncentrationen av A i indet
till reaktorn frdubblas. Simulera fram till tiden 10.
Lsning
1: Reaktormodell och systemgrnser
Systemet beskrivs som en ideal tankreaktor, dr inoch utstrmmar passerar system-
grnsen:
[A]
in,t0
Q
in
[A]
t0
[B]
t0
V
[A]
t0
[B]
t0
Q
ut
[A]
in,t0
Q
in
[A]
[B]
V
[A]
[B]
Q
ut
2: Materialbalanser
Materialbalanserna tecknar vi som Ut + Prod = In + Ack, och eftersom volymen r
konstant s gller:
In + Prod = Ut + Ack
A : Qc
in,A
+r
A
V = Qc
A
+V
dc
A
dt
B : 0 +r
B
V = Qc
B
+V
dc
B
dt
3: Kinetisk modell och hastighetsekvationer
Hastighetsuttrycken tecknas som:
r = kc
A
r
A
= kc
A
r
B
= kc
A
7.2 Arbetsmetodik 96
4: Kombination av materialbalanser och hastighetsekvationer
A :
dc
A
dt
=
Qc
in,A
V
c
A
(
Q
V
+k)
B :
dc
B
dt
= kc
A
c
B
Q
V
eller, i matrisform der = Ac +Y :
der =
_
dc
A
dt
dc
B
dt
_
A =
_
(
Q
V
+k) 0
k
Q
V
_
c =
_
c
A
c
B
_
Y =
_
Qc
in,A
V
0
_
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
der = Ac +Y t = 0
_
c
A
= 9.0909
c
B
= 90.9091
Funktionslen blir nstan identisk med den som anvndes fr att ta fram initialvr-
dena. Man kollar frsta att man verkligen har stationra betingelser, och kr med
ode45(DynExempel,[0 10],[9.0909;90.9091]). Man fr fljande gur, som visar
att man har stationritet:
1 2 3 4 5 6 7 8 9 10
0
10
20
30
40
50
60
70
80
90
100
k
o
n
c
e
n
t
r
a
t
i
o
n
tid
cb
ca
7.2 Arbetsmetodik 98
7: Dynamisk simulering av strningen
Efter ndring av inkoncentration i funktionslen: krs denna med ode45(DynExempel,[0
10],[9.0909;90.9091]). Det dynamiska frloppet blir:
1 2 3 4 5 6 7 8 9 10
0
20
40
60
80
100
120
140
160
180
ca
cb
tid
k
o
n
c
e
n
t
r
a
t
i
o
n
C H A P T E R 8
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 eld. One concept is dis-
persion 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 diu-
sion. Accordingly, we hereby combine together the basic molecular mass transport
theories with environmental transport phenomena in the eld scale. Applications in-
clude, 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 (ow) are linked. In mathe-
matical model, the chemical and physical processes involved can be linked and treated
quantitatively.
This section begins with the fundamentals of ow in porous media such as ground-
water 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 aected
by degradation and adsorption, introducing and using the important retardation fac-
tor.
It is inevitable to use partial dierential 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 ow- Darcys law 100
Grundvatten-
frorening
lipofilicitet
MNET
hast.koeff
nedbrytning
hast.
ekv
K
ow
MARKEN
organisk halt
adsorption
R-vrde
BERKNING
Femlab
kontinuitets
ekvation
porositet
vattenhalt
VATTNET
Darcy
Darcy-hast.
dispersion
advektion
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 trans-
port in a porous soil system.
8.1 Water ow- Darcys law
If possible, water ows 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 aected by gravitational forces, the potential energy of
the water increases with height but the energy is readily lost or transformed as the
water ows 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 dierence between two
points is called hydraulic head.
In practice, the driving force for ow between two points is often proportional
to the dierence in height. However, the force acting on a package of water in the
gravitational eld 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 (diusion) and heat transport, we can assume that
the ow velocity is proportional to the hydraulic gradient:
v
x

y
x
where v is the ow velocity in the x direction (e.g. m s
1
), as shown in gure 8.2.
The proportionality constant is called the hydraulic conductivity, K
x
, which has
the same unit as the ow velocity. In 1856 the french engineer Darcy stated that:
v
x
= K
x
y
x
(8.1)
8.1 Water ow- Darcys law 101
x
-y
g
ru
n
d
v
a
tte
n
y
ta

vattendrag
eller brunn
Figur 8.2: The principles behind Darcys law - gravity makes water ow from a
state with higher potential energy to a point with lower energy.
Tabell 8.1: Exemples of hydrogeologic parameters
Typical Total Eective Hydraulic
Fraction particle- porosity porosity konductivity
storlek , %
e
, % K, cm s
1
Gravel 1 mm 25 35 25 35 1 - 100
Sand 0.1 mm 30 45 25 40 110
4
110
1
Silt 0.02 mm 35 45 20 35 110
6
110
4
Clay 0.002 mm 40 55 2 10 110
9
110
6
In practice, K is dicult to predict with reasonable accuracy. I table 8.1 gives
a summary of typical values for dierent types of soils. As we see is the hydraulic
conductivity is much higher for coarse grained than ne-grained soil. The ranges are
large, so in practice, the best values in elds are deterimend by pumping.
The hydraulic gradient can be measured by installing pressure sensors in the sat-
urated 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 ow velocity
The velocity obtained with Darcys law must be interpreted as the ow Q through a
surface A perpendicular to the direction x during a certain time. Hence:
v
x
= K
x
y
x
=
Q
x
A
It is common to call the rate determined in this way for Darcy-velocity, or supercial
velocity. However, in a porous medium, a signicant 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 eective porocity is very low, the ow velocity in the fraction of the soil where
the transport takes place must exceed the average velocity in the soil.
The actual ow velocity u
x
is calculated as:
u
x
=
K
x

y
x
(8.2)
8.1 Water ow- Darcys law 102
where
e
is the eective porosity. As shown in table 8.1. it may dier signicantly
between the total and eective porosity. The dierence is greatest for very ne-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
ne pores, soil capillarity can be much larger than the gravitational force.
Example 81: Estimation of ow velocity
Problem
To determine the ow in an open groundwater aquifer, 3 observation wells were in-
stalled along a trans-sect of 400 m. The wells were thus at 200 m intervals. Ground-
water 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 con-
ductivity of 210
3
cm s
1
. The eective 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
x
=
88.5 90
400 0
=
1.5
400
= 3.7510
3
m m
1
The unit of K is changed, so that:
210
3
cm s
1
= 210
3
cm
s
10
2
m
cm
606024365
s
year
= 630
m
year
The Darcyvelocity can then be calculated as:
v
x
= K
x
y
x
= 6303.7510
3
= 2.36
m
year
The real ow velocity is:
u
x
=
v
x

e
=
2.36
0.3
= 7.87
m
year
and the time it takes for water to be transported 400 m becomes:
t =
x
u
x
=
400
7.87
= 58 year!
8.1 Water ow- Darcys law 103
lng vg
kort vg
friktion mot partikelytor
Figur 8.3: Illustration of mechanisms that causes a distribution of ow 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 simplication. Each water parcel does in fact move at dierent velocity
at dierent 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 dierent paths around a particle meet in a por?
If one is only interested in total water ux, as represented by the Darcy-velocity,
the distribution of ow velocities in groundwater aquifer is of no importance. In
that case, the fact that the water packages have spent dierent amounts of time in
the ground, and founds their way forward in dierent 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 ow 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 gure 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 ow 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 conned 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 ow, 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 s
1
). In these terms Q the convective ow of the bulk,
while Qc represents the advective transport of dissolved substances.
8.2 Dispersion
8.2.1 Molecular diusion - Ficks law
The molecular diusion 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 ow of molecules from higher concentrations to lower concentrations. This means
that the net ow 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
ux J will be proportional to the concentration gradient:
J = D
dc
dL
mass m
2
s
1
(8.3)
For molecular diusion, the proportionality constant is the diusion rate coecient
D (e.g. m
2
s).
It is obvious that the ux into a control volume can aect the concentration of a
substance in the volume. If the input ux equals the output ux, the concentration
in the volume will not change . We can, of course, conclude that the change in con-
centration per unit time in a control volume must depends on the dierence between
the input and the output uxes.
8.2 Dispersion 105
L-riktning
diffusions-
riktning
dL
c c+dc
A
koncentrations-
gradient
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 ux is specied per unit area, and volume of the control volume can be
written dV = AdL we get:
Ack = In Ut
V
c
t
= AJ
in
AJ
ut
V
c
t
= AD
c
L
A(D
(c + c)
L
)
AL
c
t
= AD

2
c
L
c
t
= D

2
c
L
2
(8.4)
8.2.2 Analogy dispersion - diusion
The irregular ow 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 Darcy-
velocity 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 ow in a porous medium, one nds 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 specic route, or go faster.
8.2 Dispersion 106
Dispersion is thus a process that occurs in the macroscopic scale, cm and meters,
while diusion 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 diusion. Thus, we can apply Ficks law also
for dispersion. The only dierence is that we replace the physical diusion constant
D (e.g. m
2
s
1
) with the empirical dispersion coecient E.
8.2.3 Deduction of expression for spread
Diusion (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 diusion 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 diusion 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 innitesimal
slice in space.
The diusion equation is a partial dierential 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:
c
t
= D

2
c
L
2
_

_
Boundary condition I c(, t) = 0
Boundary condition II
c
L
(0, t) = 0
Initial value I c(0, 0) = M/
(8.5)
The condition
c
L
(0, t) = 0 expresses the fact that the curve is symmetrical around
the maximum value. The solution is:
c(L, t) =
M
(4Dt)
1
2
e
L
2
/4Dt
(8.6)
The spread around the symmetrical axis at x = 0 is given by the variance,
2
. It is
dened as the weighted deviation from x = 0. As the spread changes with time, the
variance is also time dependent:

2
(t) =
_
+

L
2
c(L, t)dL
_
+

c(L, t)dL
(8.7)
We shall now consider the numerator and denominator separately.
The numerator We can now develop the numerator by inserting the solution of
the dierential equation: equation 8.7. This gives:
Numerator =
_
+

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

L
2
M
(4Dt)
1
2
e
L
2
/4Dt
dL
8.2 Dispersion 107
We can solve this integral identifying that it is on the form:
b
_
+

y
2
e
ay
2
dy where
_

_
y = L
a =
1
4Dt
b =
M

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

y
2
e
ay
2
dy = 2b

4a

a
so that:
Numerator = 2b

4a

a
= 2M

4a

a
= 2M
1
4a
= 2M
4Dt
4
= 2MDt
Denominator The denominator can be treated in about the same way. The integral
fromto +may, because of symmetry, be divided into the two segments 0
and 0 . Inserting the solution to the diusion equation into the expression for
the denomimator we get:
Denominator =
_
+

c(L, t)dL =
_
+

M
(4Dt)
1
2
e
L
2
/4Dt
dL = 2
_
+
0
M
(4Dt)
1
2
e
L
2
/4Dt
dL
In this case too, we can nd a standard integral:
b
_
+
0
e
ay
2
dy where
_

_
y = L
a =
1
4Dt
b = M
_
a

The solution is
b
2
_

a
and for the whole interval + we get:
Denominator =
2b
2
_

a
= M
_
a

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

2
=
2MDT
M
= 2Dt
and the standard deviation:
=

2Dt
8.2 Dispersion 108
-2 -1,5 -1 -0,5 0 0,5 1 1,5 2
Lngdskala (t ex meter)
= 0.50
= 0.25
= 1
= 0
Koncentration av mne
Figur 8.5: Development of the concentration prole as a substance spreads by
means of diusion in a stagnant medium.
8.2.4 Interpretation of the standard deviation
Spreading of a substance through diusive and dispersive processes can thus be ex-
pressed in terms of standard deviation .
Figure 8.5 shows how the concentration prole develops in one (1) dimension as
caused by diusion, i.e. in a system without any convective ow. For example,
if a tracer at the time t = 0 is contained in a point in space, it will diuse out
symmetrically from the point. The concentration distribution of the substance will
always be always correspond to a normal distribution
1
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 =

2
2D
1
The distribution is given by
1

2e
L/
8.2 Dispersion 109
0 L 4L 16L
Transportstrcka L=ut (t ex meter)
= 2EL/u
L
= 0
= 2E4L/u=2
L

= 2E16L/u=4
L

Koncentration av mne
Figur 8.6: Illustration of how the concentration prole develops during coupled
advection and dispersion.
8.2.5 Application to a advection-dispersion system
During ow in a porous medium, the substance moves both by advective transport,
and by random diusion and dispersion. Some molecules of the substance move faster
than the average ow velocity, and just as many move slower than the average ow
velocity. Therefore, around the top of concentration prole, the concentration devel-
ops in principle in the same manner during advective as molecules that during diuse
from a point. Thus, the standard deviation does characterize the concentration
prole in a advection-dispersion system just as in a stagnant system with diusion.
The dierence is that dispersion creates much greater spread than diusion, i.e.
grows faster because of dispersion than because of diusion. 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 prole 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 = u
L
t (8.9)
Combining equations 8.9 and 8.8 we see that when the peak is at L,
L
is:

L
=
_
2EL/u
L
That means that as time (and thus the transport distance) doubles, the standard devi-
ation increases by a factor

2. Since the concentration follows a normal distribution,


68% of the substance will be found within the distance represented by 2
_
2EL/u
L
when the peak is at L. When the peak is at 16L, 95% of the substance will be within
2
_
162EL/u
L
from L, i.e. a 4 times wider distance.
8.3 Dispersion coecients in groundwater aquifers 110
Example 82: Diusive 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
diusion and/or dispersion a pulse of a tracer introduced into the ow 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 dened as the distance that contains 95% of the pulse.
1. The spread is diusive with D = 110
9
m
2
s
1
2. The spread is dispersive with E = 110
6
m
2
s
1
Solution
The spread can be quantied 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:
4
diff
= 4

2Dt = 4
_
2(110
9
360024365)58 = 7.6 m (E82.10)
4
disp
= 4

2Et = 4
_
2(110
6
360024365)58 = 242 m (E82.11)
We can thus conclude that dispersion is the process responsible for the spread, simply
because the dispersion coecient is greater than the diusion coecient.
8.3 Dispersion coecients in groundwater aquifers
It is dicult to obtain precise values of the dispersion coecient under eld 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 coecients are highly variable.
In many situations though, one does not need to know the dispersion coecient
exactly. If one for example should carry out a risk assessments, it may be sucient
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 ow velocity. The higher the
ow velocity is, the higher the E-value becomes. If one assumes that E is proportional
to u
L
, one can dene a proportionality constant, the dispersivity :
E
L
= u
L
(8.10)
The dispersivity has the unit meter, or any other unit of length.
8.3 Dispersion coecients in groundwater aquifers 111
Tabell 8.2: Empirical values of dispersivity in soils (after Schnoor 1996)
Scale (m)
Longitudinal dispersivity (m)
Typical Interval
Lab studies < 1 0.0010.01 0.00010.01
Field studies 1-10 0.11.0 0.0011.0
Field studies 10-100 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 hetero-
geneous 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 signicant,
and one must not forget that the axes are in log - scale.
8.3 Dispersion coecients in groundwater aquifers 112
Empirical correlation In the university text book Applied Hydrogeology one nds
the following correlation:
= 0.0175L
1.45
(8.11)
where L should be given in meter. This correlation gives similar results as the empir-
ical values described above.
Pe-correlation Yet another possibility is to use correlations that uses a dimension-
less property, the Pe-number (after Pclet).
The Pe-number is dened as the ratio between a term that reects the advective
transport, divided by the dispersive mixing.
Pe =
u
L
L
E
In the literature, one nds empirical equations such as:
E
D
= 8.8Pe
1.17
, 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
coecient. That is, provided that the systems from which they have been derived are
similar to the system of application.
C H A P T E R 9
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 nding the right parameter
values. It may be kinetic rate and chemical equilibrium constants, values of hydraulic
conductivity and dispersion coecients.
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 disper-
sion
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
Processes considered Time-
Equation
Advection Dispersion Reaction Adsorption variability
Continuous - - Dynamic 9.5
Pulse - - Dynamic 9.7
Continuous - Dynamic 9.11
Step - Dynamic 9.16
Continuous - Stead-state 9.18
Continuous I Dynamic E92.29
Pulse Dynamic 9.28
113
9.1 The continuity equation - advection and dispersion 114
strmnings-
riktning
L = 0
c = c
0
L =
c = 0
dL
A
t = nnu strre
t = strre
t = litet
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 left-
hand boundary L = 0 the concentration is constant, c
in
. 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 dierence in ux J
between the two boundaries:
AdL
c
t
= AdJ (9.1)
In terms of the two processes advection and dispersion, the ux J becomes:
AJ = advection term + dispersion term = A(u
L
c E
c
L
) (9.2)
This equation directly gives:
c
t
=
A
A
J
L
=

L
(u
L
c E
c
L
) = u
L
c
L
+E

2
c
L
2
(9.3)
The negative sign in front of the advection term makes sense; if the concentration
decreases in the L-direction,
c
L
will be negative, and u
L
c
L
will be positive. This
9.1 The continuity equation - advection and dispersion 115
corresponds to a situation where the advective transport adds substance to the control
volume, causing a positive
c
t
.
Equation 9.3 should be solved with three conditions:
c
t
= u
L
c
L
+E

2
c
L
2
_

_
Boundary condition I c(0, t) = c
in
Boundary condition II
c
L
(, t) = 0
Initial value I c(L, 0) = 0, L 0
(9.4)
The equation has the complete solution:
c(L, t) =
c
in
2
_
erfc(
L u
L
t

4Et
) +e
u
L
L/E
erfc(
L +u
L
t

4Et
)
_
(9.5)
However, with an error of a few percent, it can be simplied to:
c(L, t) =
c
in
2
erfc(
L u
L
t

4Et
) (9.6)
The simplied solution makes it obvious that at L = u
L
t (i.e. at the mean convective
travel distance), the concentration is always c = c
in
/2! Often the development of the
concentration, as illustrated in example 91, is seen as a front that moves forward in
the ow 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 gure 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 equa-
tion 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 u
L
t
from the starting point at L = 0. We can now modify equation 8.6 by replacing the
spatial coordinate L with L u
L
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
e
(L u
L
t)
2
/4Et
(9.7)
as a solution to:
c
t
= u
L
c
L
+E

2
c
L
2
_

_
Boundary condition I c(, t) = 0
Boundary condition II
c
L
(u
L
t, t) = 0
Initial value I c(0, 0) = M/
Note that (L u
L
t)
2
always is positive, hence the c is symmetrical around L = u
L
t.
How should then the amount M be interpreted? M is the amount that is added
per m
2
at a certain instant. The follows from the fact that the left hand side has the
unit mass per unit volume (e.g. g m
3
) while the denominator (4Et)
1
2
has the unit
length (e.g. m).
9.2 Advection, dispersion and reaction 116
Example 91: Inltration of a pollutant into a soil
Problem
A non-reactive, water soluble substance, dissolved in 0.5 m
3
per m
2
and year, is
continuously added onto a water-saturated soil. If the eective 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 dierent times. The
parameters become:
u
L
=
0.5
0.35
= 1.43
m
years
E = u
L
= 1.430.6 = 0.86 m
The equation is dened in a MATLAB le: which is run with:
>>fplot(exempel11_1,[0 10])
and so on. The result becomes:
0 1 2 3 4 5 6 7 8 9 10
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Lngdkoordinat
K
o
n
c
e
n
t
r
a
t
i
o
n

(
a
n
d
e
l

a
v

i
n
k
o
n
c
e
n
t
r
a
t
i
o
n
)
tid=0.5 r
tid=1.0 r
tid=2.0 r
tid=5.0 r
9.2 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 exam-
ple, equation 5.13 was used to show that for a closed system, the time-derivative of
the concentration can be expressed as:
c
t
= r (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
t
= u
L
c
L
+E

2
c
L
2
+r (9.9)
In a case where the substance is degraded according to a 1st order reaction, for which
r = kc, the continuity equation becomes:
c
t
= u
L
c
L
+E

2
c
L
2
kc
_

_
Boundary condition I c(0, t) = c
in
Boundary condition II
c
L
(, t) = 0
Initial value I c(L, 0) = 0, L 0
(9.10)
The complete, quite complicated, analytic solution is:
c(L, t) =
c
in
2
_
e
u
L
L(1)
2E
erfc(
L u
L
t

4Et
) +e
u
L
L(1+)
2E
erfc(
L +u
L
t

4Et
)
_
(9.11)
where:
=
_
(1 + 4kE/u
2
L
) (9.12)
Figure 9.2 shows a calculation corresponding to example 91, with the dierence 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:
c(L, t) =
c
in
2
e
u
L
L(1 )
2E
erfc(
L u
L
t

4Et
) (9.13)
If 4kE/u
2
L
< 0.0025 then (1 + 2kE/u
2
L
), and equation 9.13 becomes:
c(L, t) =
c
in
2
e
kL/u
L
erfc(
L u
L
t(1 + 2kE/u
2
L
)

4Et
) (9.14)
9.2.2 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 conned period of time. In practice, this could occur if a spill take place from a
certain time (t = 0) and is prevented at t = t
f
. In this case the model becomes:
c
t
= u
L
c
L
+E

2
c
L
2
kc
_

_
Boundary condition I c(0, 0 t t
f
) = c
in
Boundary condition II c(0, t > t
f
) = 0
Boundary condition III
c
L
(, t) = 0
Initial value I c(L, 0) = 0, L 0
9.2 Advection, dispersion and reaction 118
0 1 2 3 4 5 6 7 8 9 10
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Lngdkoordinat
K
o
n
c
e
n
t
r
a
t
i
o
n

(
a
n
d
e
l

a
v

i
n
k
o
n
c
e
n
t
r
a
t
i
o
n
)
k=0.0 /r k=0.1 /r
k=0.2 /r
k=0.5 /r
Figur 9.2: Calculation with equation 9.11 for the time t = 5 years.
(9.15)
A simplied [sic!] form of the analytical mathematical solution is:
c(L, t) = (9.16)
c
in
2
e
kL
u
L

_
erfc(
L u
L
t(1 +
2kE
u
2
L
)

4Et
) erfc(
L u
L
(t t
f
)(1 +
2kE
u
2
L
)

4Et
)
_
Now, let us concider the extreme cases. If the pulse is innitely short, i.e. t
f
0,
then the two erfc-terms will be equal and cancel out, causing c = 0 for all t
f
. On the
other hand, if t
f
than we have a continuous source. In that case the second
term is erfc() = 0. That leaves us with the rst term, and the nal 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 c
in
.
Figure 9.3 shows how the concentration prole develops for a reference case (E =
1.43, u
L
= 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 coecient 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 1 2 3 4 5 6 7 8 9 10
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Lngdkoordinat
K
o
n
c
e
n
t
r
a
t
i
o
n

(
a
n
d
e
l

a
v

i
n
k
o
n
c
e
n
t
r
a
t
i
o
n
)
tiden = 1 r
tiden = 2 r
tiden = 5 r
tiden = ondlig=stationritet
Figur 9.3: Calculation for dierent equation 9.11 values on t.
of ow we get:
lim
t
e
u
L
L(1

(1+4kE/u
2
L
)
2E
= e
kL/u
L
for E 0 (9.17)
lim
t
erfc(
Lu
L
t

(1+4kE/u
2
L
)

4E
) = erfc() = 2 for E 0
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/u
L
. If t = L/u
L
is substituted into equation 9.13 we can conclude that:
c(L, t)
t
= 2
c
in
2
e
kt
= c
in
e
kt
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 dierent 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 reac-
tions
1
, involving the substance in dissolved form and adsorbed to some type of solid
compound.
Classic physical chemistry suggests a number of adsorption isotherms, with dif-
ferent 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 = K
d
c (9.19)
where S is the amount of sorbing solid substance (sorbate) in the soil, expressed in
unit such as mg kg
1
. Since the concentration is given in mass of substance per
unit mass of soil, the partitioning constant K
d
will have a unit of the type m
3
kg
1
.
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:
1. soil bulk density (
kg
soil
m
3
soil
)
2. soil water content, e.g. the eective porosity (
m
3
water
m
3
soil
) in the saturated zone
We can now form the relationships:
amount sorbed
volume of water
=
S

=
K
d
c

(9.20)
This means that the eect of a change in sorbed amount within a unit volume of soil
water can be expressed as:
d
dt
S

=
d
dt
K
d
c

=
K
d

dc
dt
(9.21)
1
http://en.wikipedia.org/wiki/Partition coecient
9.3 Adsorption and transport 121
9.3.1 Advection, dispersion, reaction and adsorption
Since adsorption aects 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
t
= u
L
c
L
+E

2
c
L
2
kc

t
S

(9.22)
or, withe the help of equation 9.21:
c
t
= u
L
c
L
+E

2
c
L
2
kc
K
d

c
t
(9.23)
(1 +
K
d

)
c
t
= u
L
c
L
+E

2
c
L
2
kc (9.24)
(9.25)
We have now deduced the so called retardation factor, R. This important quantity is
dened as:
R = (1 +
K
d

) (9.26)
R can be substituted into the continuity equation as:
c
t
=
u
L
R
c
L
+
E
R

2
c
L
2

k
R
c (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
(4Et/R)
1
2
e

(L u
L
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 m
3
per m
2
and
year, is continuously added onto a water-saturated soil. If the eective soil bulk
density 1.5 kg L
1
, 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
k = 0.1
1
years
. The substance is adsorbed to the soil organic matter (K
d
= 0.6 L kg
1
),
and other relevant data can be found in example 91.
Solution
We can use equation 9.11, where the parameters u
L
, E and k are scaled with R.
c(L, t) = (E92.29)
c
in
2
_
e
u
L
L(1)
2E
erfc(
L u
L
t/R
_
4Et/R
) +e
u
L
L(1+)
2E
erfc(
L +u
L
t/R
_
4Et/R
)
_
Note that in the exponent
u
L
L(1)
2E
as well as in =
_
(1 + 4kE/u
2
L
) R disappears
from both the numerator and the denominator.
We calculate R as:
R = (1 +
K
d

) = (1 +
0.61.5
0.35
) = (1 + 2.53) = 3.43
Applying the above equations and data gives:
0 1 2 3 4 5 6 7 8 9 10
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Lngdkoordinat
K
o
n
c
e
n
t
r
a
t
i
o
n

(
a
n
d
e
l

a
v

i
n
k
o
n
c
e
n
t
r
a
t
i
o
n
)
lite dispersion
dispersion
dispersion+reaktion
dispersion+reaktion+adsorption
Something strange to think about: shouldnt the curves for only dispersion cross at
c/c
in
=0.5?
9.4 The continuity equation in multiple dimensions 123
9.3.2 Determination of K
d
and R for organic substances
The tendency of dierent substances to partition between water and organic sub-
stances vary over a very wide range, as does the partitioning coecient K
d
. As
indicated earlier, K
d
quanties the partitioning between an organic substance partly
in a water phase, partly sorbed to soil organic matter. K
d
is dened as:
K
d
=
S
c
m
3
water
kg soil
It is quite logical that K
d
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 f
oc
, and the coecient for partitioning of an organic substance
between water and the SOM is K
oc
. This allows us to write:
K
d
= K
oc
m
3
water
kg SOM
f
oc
kg SOM
kg soil
= K
oc
f
oc
m
3
water
kg soil
The magnitude of the partitioning coecient K
oc
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 parti-
tion between n-octanol and water. The partitioning coecient for this three compo-
nent system (substance, n-octanol, water) is normally denoted K
ow
. Reliable K
ow
-values can be found in the literature
2
, often as log K
ow
A lipophilic substance like
benzene has a log K
ow
around 3 (K
ow
= 1000), while extremely lipophilic substances
such has DDT, PCB and PAH all have log K
ow
6.
Table 9.2 shows useful correlations that can be used to estimate K
oc
from K
ow
together with K
o
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, ow 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:
R
c
t
= u
x
c
L
u
y
c
L
+E
x

2
c
L
2
+E
y

2
c
L
2
+E
z

2
c
L
2
r (9.29)
Thus can be written in terms of operators on the general form:
R
c
t
= uc +

Ec r (9.30)
Referring to the notation in gure 8.2, it is always the case that:
E
x
>> E
y
> E
z
2
http://logkow.cisti.nrc.ca/logkow/
9.4 The continuity equation in multiple dimensions 124
Tabell 9.2: The following correlations have been used:
Chlorinated substances: log K
oc
= 0.72log K
ow
+0.5
Aromatics: log K
oc
= log K
ow
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 K
ow
log K
oc
Solubility (mg/L)
Chlorinated substances:
Bromoform 2.30 2.16
Carbontetrachloride 2.64 2.40 800
Chlorethane 1.49 1.57 5 740
Chloroform 1.97 1.92 8 200
Chloromethane 0.95 1.18 6 450
Dichloromethane 1.6 1.41 20 000
Hexachloroethane 3.34 2.81 50
Tetrachloroethylene 2.88 2.57 200
Trichloroethylene 2.29 2.15 1 100
Vinylchlorid 0.60 0.93 90
1,1-Dihlorethane 1.80 1.80 400
1,2-Dichloroethylene 1.48 1.57 8 000
1,1,1-Dichlorethane 2.51 2.30 4 400
1,1,2-Dichlorethane 2.07 1.99 4 500
Aromatics:
Benzene 2.13 1.92 1 780
Benso(a)pyrene 6.06 5.85 0.0038
Chlorobenzene 2.84 2.63 500
Ethylbenzene 3.34 3.13 152
Hexachlorobenzene 6.41 6.20 0.006
Naftalene 3.29 3.08 31
Nitrobenzene 1.87 1.66 1 900
Pentachlorophenole 5.04 4.83 14
Phenole 1.48 1.27 93 000
Toluene 2.69 2.48 535
Dichlorbenzene 3.56 3.35 100
2,4,5,2,4,5-PCB 6.72 6.51
Pesticides:
Atrazine 2.69 2.55 33
Dieldrine 2.17 1.96 28 500
DDT 6.91 6.70 0.0055
2,4-D (fenoxiacid) 1.78 1.65 900
C H A P T E R 10
Bioreaction engineering -
biolms
An especially exciting group of reaction systems are biological systems. Biosystems
is often dicult 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 biolm 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 pur-
poses (gure 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 con-
centration 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
tillvxtreglerande
substrat
metaboliter
biomassa
energi
cell
vriga
substrat
metabolism
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 dened in several ways, such as per unit volume reactor per
unit volume of cells or per unit mass of cells. This section denes:
Reaction rate per unit mass of cells
g
substrate
g
cells
tidsenhet
r
S
Reaction rate per unit volume of reactor
g
substrate
m
3
reactor
unit time
r
B
Reaction rate per unit volume of cells
g
substrate
m
3
cells
unit time
10.2.1 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 specic 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 SE
Conversion of enzyme-bound substrate to product: SE P + E
One can formulate two rate equations for these two reactions:
r
1
= k
1
[S][E] k

1
[SE]
r
2
= k
2
[SE]
One can now see three things:
The total amount of enzyme E
tot
is limited, i.e. [E] + [SE] = E
tot
At steady-state r
1
= r
2
since [SE] is constant
Conversion of substrate , must be = r
1
= r
2
One can now deduce a general expression for . Fits we eliminate [E] by introducing
[E] = E
tot
[SE] in both rate equations:
r
1
= k
1
[S](E
tot
[SE]) k

1
[SE]
r
2
= k
2
[SE]
Then, one uses that r
1
= r
2
:
k
1
[S](E
tot
[SE]) k

1
[SE] = k
2
[SE]
k
1
[S]E
tot
[SE](k
1
[S] +k

1
+k
2
) = 0
From this, one may solve for and simplify [SE]:
[SE] =
k
1
E
tot
[S]
(k
1
[S] +k

1
+k
2
)
10.2 Kinetic equations 128
Finally, one can insert the expression for [SE] in the most simple of the rate equations,
together with = r
2
= k
2
[SE], to get:
=
k
2
k
1
E
tot
[S]
(k
1
[S] +k

1
+k
2
)
This expression can be compressed by lumping together various coecients to the
classic rate equation:
=

max
[S]
[S] +K
1/2
g
substrate
g
cells
unit time
(10.1)
As we see, the nal rate equations does only include two parameters ,
max
and
K
1/2
. The interpretation of these parameters is shown in gure 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
coecient.
K
1/2
is the substrate concentration at which is 50% of
max
. Therefore, K
1/2
is named the half rate constant. As clear from equation 10.1 and gure 10.3, the rate
equation becomes ater as K
1/2
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

max
[S]
K
1/2
High substrate concentration
max
One should be aware that the experimentally determined equations and parame-
ters can not be extrapolated to other system. Although micro-organisms of the same
type and strain can behave dierently in dierent environments, e.g. by forming
aggregates aecting the mass transport.
10.2 Kinetic equations 129
0.0
0.5
1.0
0 2 4 6 8 10
Reaktionshastighet
Substratkoncentration [S]
K
1/2
=1
K
1/2
=4
= 1*[S] / (1+[S])
= 1*[S] / (4+[S])
Figur 10.3: Plot of the kinetic equation for two cases. Note that the parameter
K
1/2
corresponds to the substrate concentration where r
S
is 50% of
max
.
0.0
0.5
1.0
0 2 4 6 8 10
=
max
[S] / (K
1/2
+[S])
Reaktionshastighet
Substratkoncentration [S]
1.a ordningen:

max
[S]/K
1/2
0:e ordningen:

max
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 sub-
strate concentration.
10.2 Kinetic equations 130
10.2.2 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
has the unit
g
substrate
m
3
reactor
.
Then, the volume-based rate equation is written:
r
S
= [X] =

max
[S]
[S] +K
1/2
[X] (10.2)
10.2.3 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 uxes 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 dierent degrees by
the substrate concentration, the mortality can be assumed to be directly proportional
to the concentration of microorganisms, [X]. The proportionality constant is k
d
which
has the unit per unit time.
Net production of micro-organisms per unit volume in the system denoted by
r
X
, and is the dierence between growth and mortality. The unit of r
X
becomes
g
cells
m
3
reactor
unit time
:
r
X
= (Y k
d
) [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:

max
3 g
substrate
g
1
cells
day
1
K 3 g m
3
Y 3 g
cells
g
1
substrate
[S] at t = 0 100 g
substrate
m
3
[X] at t = 0 1 g
cells
m
3
10.2 Kinetic equations 131
Solution
For a batch reactor, Prod = Acc, which means that:
d[X]
dt
= r
X
;
d[S]
dt
= r
S
where
r
X
= (Y k
d
)[X] = (

max
[S]
[S] +K
1/2
Y k
d
)[X] (E101.4)
r
S
= [X] = (

max
[S]
[S] +K
1/2
)[X] (E101.5)
These equations are put in a MATLAB-le:
0 1 2 3 4 5 6 7 8 9 10
0
10
20
30
40
50
60
70
80
90
100
Cellkoncentration X
Substratkoncentration S
Tid (dygn)
K
o
n
c
e
n
t
r
a
t
i
o
n

(
g

p
e
r

m
3
)
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 Biolms 132
Figur 10.5: Biolm of denitrication bacteria on carriers of plastic from
AnoxKaldnes. The lm is about 150 (m) thick and comes from the Klagshamn
water treatment plant. The carrier has a specic surface area of 570 m
2
m
3
.
10.3 Biolms
In many technical applications the reaction takes place in biological lms. It is very
common in water treatment contexts. These biolms may grow on carriers made out
of plastics or beds of stones in a reactor.
Biolms oer some practical advantages. They counter that the level of suspended
solids becomes too high in the water, they are surprisingly resistant to harmful pol-
lutants and they are physically stable.
A basis for biolm process is that the substrate is transported from the surrounding
liquid bulk into the lm. Here it is converted into products, e.g. biomass. Thereby
, a oncentration gradient is created from the biolm surface in towards the carriers
wall. This creates a driving force for a ux from the liquid to the biolm (Figure
10.7). This ux in each point is proportional to the gradient magnitude.
Intuitively, we may also nd that the faster the substrate is transported into the
biolm, the stronger the gradient and the ux become
Figure 10.7 shows schematically how the concentration prole in a biolm can
look. The main principle is that substrates diuse from the biolm surface, through
the lm toward the carriers surface. The driving force is the concentration gradient
as incurred by the substrate consumed in the biolm.
The part of the biolm that is supplied by substrate is called the penetration depth,
L
P
. The fraction of the lm that is supplied is called the degree of penetration and is
dened as the ratio between the penetration depth and the thickness of the biolm,
L
P
/L
B
.
10.3 Biolms 133
Figur 10.6: Biolm on a plastic carrier.
The degree of penetration is dened as:
B =
L
P
L
B
The 4 dierent cases that are shown in gure 10.7 are:
a: Deep lm The substrate can be consumed so rapidly that the concentration at
the carrier surface becomes 0. Deep refers to the fact that the lm is unnec-
essary 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 lm Substrate is converted everywhere the biolm, and
the concentration of substrate becomes 0, exactly at the carriers surface. This
is the optimal situation, the whole lm is used and and substrate conversion is
complete. The degree of penetration B = 1.
c: Thin lm The whole lm 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 lm is high
and/if the mass transfer from the mixed liquid bulk to the lm surface is slow,
the external mass transfer becomes important. It means that the concentration
at the biolm surface is lower than the bulk concentration. The result is that the
driving force for both the chemical reaction in the lm as well as the diusion
of substrate in the lm is adversely aected.
10.3 Biolms 134
L=0 L
B
L
P
brare
substratflux
biofilm omblandad vtskebulk
C=C
0
C=0
(a) Deep lm, non-penetrated. Parameter B < 1.
L=0 L
P
=L
B
brare
substratflux
biofilm omblandad vtskebulk
C=C
0
C=0
(b) Exactly penetrated. Parameter B = 1.
L=0 L
B
brare
substratflux
biofilm omblandad vtskebulk
C=C
0
C>0
(c) Thin lm, penetrated. Parameter B > 1.
brare
substratflux
biofilm omblandad
vtskebulk
C=C
0
laminrt
grnsskikt
L=0 L
B
C=C
bulk
(d) A laminar boundary layer creates an external mass transfer
resistance between the liquid bulk and the surface if the biolm.
Figur 10.7: Dierent types of concentration proles in biolms. B was dened
in equation 10.6.
10.3 Biolms 135
10.3.1 General rate equations for biolms
In all biolm processes, the biolm should ideally have constant thickness, and fairly
constant activity. For a certain amount of substrates, there is a certain volume of
cells. For biolms we can therefore postulate that:
1. The reaction rate should be dened in terms of units of volume biolm. Hence,
we use r
B
(
g
substrate
m
3
biofilm
unit time
)
2. Cell mass is constant, so that
d[X]
dt
= 0 everywhere in the reactor.
As previously mentioned, the kintetiska parameters and K
1/2
specic to a par-
ticular system. In this section, we will simply assume that have been determined
for the current biolm system.
If we denote cell mass per unit volume of biolm with X
B
the rate equation for a
biolm becomes:
r
B
= X
B
=

max
[S]
[S] +K
1/2
X
B
(10.4)
This is analogous with equation 10.2. In practice, however, one determines the prod-
uct X
B
, and the rate equation can be written:
r
B
=
k
B
[S]
[S] +K
1/2
(10.5)
where the rate constant k
B
, has the unit g of substrate per unit volume of biolm
and time.
As mentioned earlier, the conversion of substrate depends on substrate concentra-
tion only at low values of mconc (S). We may therefore identify two special cases
with respect to kinetics:
Low substrate concentration: r
B

k
B
[S]
K
1/2
= k
1
[S]
High substrate concentration: r
B
k
B
= k
0
10.3.2 Concentration prole and ux in a thin lm for 0th
order reaction
Concentration prole
In this case the rate expression is reduced to:
r
B
= k
0
For each point in the lm, the continuity equation gives:
D

2
c
L
2
= k
0
10.3 Biolms 136
Now dimensionless quantities are introduced for concentration and for length. If the
concentration at the surface of the lm is c
0
and the biolm thickness is L
B
, then:
c

=
c
c
0
; c = c

c
0
L

=
L
L
B
; L = L

L
B
If we combine all these equations, the continuity equation becomes:
D

2
c

c
0
(L

L
B
)
2
= k
0
; D

2
c

L
2
= k
0
L
2
B
c
0
;

2
c

L
2
=
k
0
L
2
B
Dc
0
It is now possible to clean up the right hand side by introducing the variable:
B
2
=
2Dc
0
k
0
L
2
B
(10.6)
and we get:

2
c

L
2
=
2
B
2
(10.7)
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 biolm


L

= 1
c

= 0 Flux at the carriers surface is zero


Two integrations of equation 10.7 gives:
c

=
2
B
2
L

+H
1
(10.8)
c

=
L
2
B
2
+H
1
L

+H
2
(10.9)
We use the second boundary condition with equation 10.8, and thereafter the second
boundary condition with equation 10.9 to get:
0 =
2
B
2
+H
1
H
1
=
2
B
2
1 = H
2
This is put into equation 10.9 and we get at complete expression for the concentration
prole in the biolm:
c

=
L
2
B
2

2L

B
2
+ 1 (10.10)
Obviously, the concentration prole 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 lm, so that the substrate concentration c

= 0 exactly when
L

= 1. In that case, equation 10.10 gives:


c

=
L
2
B
2

2L

B
2
+ 1 ; 0 =
1
B
2
+ 1 ; B = 1
This means that exactly when B
2
=
2Dc
0
k
0
L
2
B
= 1 (i.e. when B=1) the substrate
concentration is 0 at the surface of the carrier.
10.3 Biolms 137
Flux into the lm
The ux into the lm, expressed as g
substrate
per unit surface area of biolm and unit
time, is obtained by dierentiation of the non-dimensionless form of equation 10.10,
taking the concentration gradient at L = 0. First, we re-introduce:
c

=
c
c
0
; L

=
L
L
B
to get:
c
c
0
=
L
2
L
2
B

1
B
2

L
L
B

2
B
2
+ 1
c
L
= c
0
(
2L
L
2
B
B
2

2
L
B
B
2
)
c
L
=
2c
0
L
B
B
2
i L = 0
If we insert this and the denition of B into the diusion equation we get the nal
expression for the ux of substrate into a thin biolm:
J = k
0
L
B
(10.11)
The ux depends only by the rate constant and the lm thickness. Hence the ux, i.e.
the diusion rate into the lm is independent of the diusivity of the substrate in the
biolm itself, provided that the reaction is independent of the substrate concentration.
10.3.3 Concentration prole and ux in a deep lm for 0th
order reaction
Concentration prole
If the biolm is deep, the substrate to be consumed completely in the lm, and ow
towards the carrier becomes 0. As shown in Figure 10.7 the substrate penetrates the
lm to a depth of L
P
.
As shown above, the general equation describing simultaneous, steadystate, dif-
fusive mass transfer and chemical reaction is described by:

2
c

L
2
=
2
B
2
(10.12)
For a deep lm, this transfer equation should be solve with three conditions:
_

_
L

= 0 c

= 1 Concentration known at the surface of the biolm


L

= L
P
c

= 0 Flux at the carriers surface is zero


L

= L
P
c

= 0 Flux at the penetration depth is zero


One integration gives:
c

=
2
B
2
L

+H
1
10.3 Biolms 138
Tabell 10.1: Equations for uxes into biolms for dierent cases.
0:e ordningen
B 1 Tunn lm J = k
0
L
B
B 1 Tjock lm J =

2Dk
0
c
0
1:a ordningen Generellt J = c
0

Dk
1
Then, the second boundary condition gives that:
0 =
2
B
2
L
P
+H
1
H
1
=
2L
P
B
2
Another integration gives:
c

=
L
2
B
2
+H
1
L

+H
2
If we rst insert the rst boundary condition and then the second boundary condition
in the above equation we get:
1 = 0 + 0 +H
2
H
2
= 1
0 =
2L
P
B
2

2L
P
B
2
+ 1 0 =
L
P
B
2
+ 1 L
P
= B H
1
=
2
B
This can be summarized in the polynomial equation describing the concentration
prole in a deep lm:
c

=
L
2
B
2

2L

B
+ 1 (10.13)
Flux into the lm
The ux into the lm is obtained by dierentiation of the non-dimensionless form of
equation 10.13, taking the concentration gradient at L = 0.
c
c
0
=
L
2
L
2
B

1
B
2

L
L
B

2
B
+ 1 (10.14)
c
L
= c
0
(
2L
L
2
B
B
2

2
L
B
B
) (10.15)
c
L
=
2c
0
L
B
B
i L = 0 (10.16)
If we combine this, and the denition of B, with the continuity equation obtain the
nal expression for the ux into a deep biolm at 0th order reaction. The expression
is:
J =
_
2Dk
0
c
0
(10.17)
The ux depends on all possible parameters, except the thickness of the lm, which
is logical.
10.3 Biolms 139
Example 102: Filmtjocklek vid exakt penetrerad lm
Problem
Calculate the minimum biolm thickness if it should be fully penetrated under the
following conditions:
k
B
0.3 g m
3
s
1
K1
2
1 mg L
1
D 310
10
m s
1
c
0
25 mg L
1
Solution
The unit mg L
1
is equivalent to g m
3
. Since c
0
>> K1
2
we can assume 0th order
reaction in the entire biolm. and thus k
0
= k
b
. This is, however, an approximation
as clear from the gure below, showing the concentration prole: c = c
0
L
2
2L

+1
(always valid when B = 1):
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
5
10
15
20
25
Dimensionsls lngd L
*
= L/L
B
S
u
b
s
t
r
a
t
k
o
n
c
e
n
t
r
a
t
i
o
n

[
S
]

(
m
g
/
L
)
For a 0th order reaction, the rate coecient k
0
= k
b
., and since the lm is fully
penetrated, B 1. Since:
B
2
=
2Dc
0
k
0
L
2
B
1
the biolm thickness when B = 1 may be written:
L
B

_
2Dc
0
k
0
=
_
2 310
10
25
0.3
= 223 10
6
(E102.18)
If the biolm thickness is less than ca 220 m, it will thus be fully penetrated, i.e. a
thin lm.
10.3 Biolms 140
10.3.4 Reactor calculations
For typical steadystate CSTR, the mass balance equation is:
In + Prod = Ut
Q
in
c
in
+rV = Q
out
c
out
In a biolm reactor, the reaction occurs because the reactant (substrate) is removed
by means of a ux into biolm surfaces in the reactor. The Prod term can therefore
be written:
r = J A
where A is the number of m
2
biolm surface per m
3
reactor. Thereby, we can easily
develop a relevant mass balance equation:
Q
in
c
in
JAV = Q
ut
c
ut
Example 103: Design of CSTR reactor volume
Problem
Calculate the volume of a biolm reactor design to reach 80% conversion in a biolm
system characterized by the data in example 102. Reactor data:
L
B
100 m
Q 10 m
3
hour
1
c
in
50 mg L
1
A 150 m
2
m
3
Solution
The concentration in the reactor will be 20% of the input concentration, i.e. 10
mg m
3
. In the absence of external mass transfer resistance, this is also the concen-
tration at the surface of the biolm. c
0
. First, one must check if the lm is penetrated
or not:
B =

2Dc
0
k
0
L
2
B
=

2 310
10
10
0.3(10010
6
)
2
= 1.41
Thus, the lm 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 = c
0
(
1
B
2

2
B
2
+ 1) = 10(
1
1.41
2

2
1.41
2
+ 1) = 5 mg/L
Since c >> K1
2
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
c
in
c
ut
JA
(E103.18)
10.4 External mass transfer 141
Since J = k
0
L
B
(equation 10.11) we nally get:
V = Q
c
in
c
out
k
0
L
B
A
=
10
3600

50 10
0.310010
6
150
= 37 m
3
(E103.19)
10.4 External mass transfer
In many systems the microbial turnover of substrate is limited by how fast the sub-
strate can be transported to the biolm surface. This situation is outlined in the
lower panel in Figure 10.7. Applying the lm theory, this means that there is a lam-
inar boundary layer. Through this boundary layer, substrate is transported toward
the surface only by diusion. 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:
J
L
=
D

(c
bulk
c
0
) = k
m
(c
bulk
c
0
) (10.18)
The ux is thus dependent on the diusivity of the substrate in water (D), as well as
the thickness of the laminar boundary layer (). The ratio between D and , k
m
, is
called the mass transfer coecient.
There are many methods available to calculate the mass transfer coecient. Com-
mon 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 biolms in packed bed.
An example of correlation is:
Sh =
_
_
_
1.09

Sc
1/3
for 0.0016 < Re < 55 och 165 < Sc < 70 000
0.25

Re
0.69
Sc
1/3
for 55 < Re < 1500 och 165 < Sc < 11 000
(10.19)
When you count on other than pipes, one must be careful to dene the Re in the right
way. In a packed bed the characteristic length of Re is dened as the particle radius.
If the particles in the bed are of dierent sizes one base the equivalent particle radius
on the material surface area:
L =
1
2
_
A
p

(10.20)
10.4 External mass transfer 142
Tabell 10.2: Relationships needed to calculate mass transfer rates.
Dimension-less number Denition
Re Reynolds number
vL
D
Sc Schmidts number

D
Sh Sherwoods number
k
m
L
D
Entity: Unit
Eective porosity
v Darcyvelocity m s
1
L Characteristic length m
D Diusivity m
2
s
1
Kinematic viscosity m
2
s
1
k
m
Mass transfer coecient m s
1
where A
p
is the average particle area express in m
2
.
The thickness of the laminar boundary layer can be calculated by combining the
denition of the mass transfer coecient into the denition of the of the Sherwood
number:
Sh =
k
m
L
D
=
D

L
D
=
L

(10.21)
so that:
=
L
Sh
(10.22)
10.4.2 Coupled external mass transfer resistance and diu-
sion/reaction in a biolm
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 biolm
10.4 External mass transfer 143
With the relationships developed earlier, we can thus state that:
0th order, thin lm:
D

(c
bulk
c
0
) = k
0
L
B
0th order, deep lm:
D

(c
bulk
c
0
) =
_
Dk
0
c
0
1st order, general:
D

(c
bulk
c
0
) = J = c
0
_
Dk
1
In qualitative terms, we can se that:
The ux into the biolm will be less if the the external mass transfer resistance
increases
On can calculate c
0
from the relationships above, after having checked B
Although there are countless possibilities to perform calculations on this type of sys-
tems, 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 minute
1
. The bed consists of sherical particles
with a diameter of 2 cm, and the eective porosity is 0.40.
Solution:
In SI-unigts, the kinematic viscosity of water is 1 10
6
and the diusivity in water
is typically 1 10
9
. The calculation has been summarized in the following MATLAB
m-le: The results indicate that the boundary layer thickness depends on the ow
condition in almost the entire interval. The maximum boundary layer thickness is 0.8
mm.
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0
1
2
3
4
5
6
7
8
x 10
-4
Darcy-hastighet (m/minut) genom bdden
G
r

n
s
f
i
l
m
s
t
j
o
c
k
l
e
k

(
m
)
Berkning av grnsfilmstjocklek i packad bdd
Hr r Re < 55
Hr r Re > 55

You might also like