Вы находитесь на странице: 1из 28

1

Lecture 4
Introduction to process analysis and selection

Objectives of the lecture

Recognize the different types of reactors used in wastewater treatment
Understand reaction kinetics
Design and conduct laboratory experiments to determine kinetics and
rate coefficients
Understand the concepts of mass balance in environmental modelling
Apply mass balance analysis and reaction kinetics to derive design
models for wastewater treatment processes.
Understand ideal and non ideal flow
Design and conduct experiments to determine time and flow
characteristics in reactors.

Summary
In this lecture we introduce the subject of process analysis in wastewater
treatment.

Basic concepts used in water quality modelling are initially presented to help you
familiarize with simple phenomena occurring in nature.

The different types of reactors used in wastewater treatment are introduced and
described. The problem of flow conditions in reactors and the difference between
ideal and non ideal flow is briefly presented.

The next section is devoted to common reactions involved in wastewater
treatment, their conceptual representation and the methods employed to
determine reaction rates and reaction coefficients.

Mass balance approach to study water quality and wastewater treatment
processes is presented. The modelling of completely mixed and plug flow reactors
types that are most commonly used in the field of waste water treatment under
ideal flow is presented and simple examples are given for the estimation of reactor
volumes.

Finally modelling flow in reactors using tracer studies is presented analytically. The
concept of non ideal flow conditions is treated more thoroughly and examples to
determine flow conditions in reactors, using the results of tracer studies, are
presented.

2
Fundamental quantities

In this chapter we will define some fundamental quantities used in water quality
modelling.

Mass and concentration
Mass is the amount of a pollutant in a system, i.e. water body, reaction tank, etc.
Mass is an extensive property, along with volume or heat.

Concentration is the amount of mass divided by the volume within which it is
contained:

V
m
C = (1)
where m = the mass and V = the volume, expressed in appropriate units.

Concentration is an intensive property, along with temperature, density and
pressure. Concentration is expressed in mg/L which is equivalent to g/m
3
. For
dilute aqueous solutions 1 mg/L is also expressed as ppm or parts per million.
For very dilute solutions concentration is expressed in ppb or parts per billion.
1ppb is equal to 1ug/L.

Rates
Rates are properties normalized to time. The most important types of rates in
water quality modelling are discussed next:

Mass loading rate, W
Mass loading rate, W is defined as the mass of pollutant, m, discharged in a water
body over time t:

t
m
W = (2)

Volumetric flow rate, Q
The volumetric flow rate is determined from the continuity equation:

c
UA Q = (3)

where U = the velocity of water in the conduit and A
c
= the cross sectional area of
the conduit.

Pollutant sources enter water bodies or reactors as point sources, such as pipes
or conduits. The mass loading rate is computed as:

C Q W = (4)

3
Mass flux rate, J
The term flux designates the rate of movement of mass or heat normalized to
area. The mass flux through a conduit is expressed as:

UC
A
W
tA
m
J
c c
= = = (5)

Reaction rate, r
Reaction rate designated the velocity with which a reaction occurs. Reaction types
and their rates will be dealt with extensively later on in this lecture

Conservation of mass and the mass balance
All models used to design wastewater treatment processes are based on the
conservation of mass principle, i.e. within a finite volume of water mass is neither
created or destroyed, it is simply transformed. In quantitative terms this principle is
expressed as a mass balance equation that accounts for all tranfers of mass
across the systems boundaries and all trasformations occurring within the system.
For a finite period of time this can be expressed as

Accumulation = loadings transport reactions (6)

Transport is the movement of mass trough the volume, along the flow of water.
Transpor of mass can be effected by advection, movement caused by the flow or
diffusion, movement cause by turbulent mixing or concentration gradients. These
notions will be elaborated further, later on in this lecture.

Reactions either add mass by changing another constituent into the substance
being modelled or remove mass by transforming the substance into another
constituent.

Finally the substance can by increase by external loadings.

By combining all the above factors in equation form we obtain the mass balance
equation which can be solved to model the change of the concentration of the
constituent over time or in general the response of a body of water to external
influences.

A system is said to be in steady state when the mass remains at a constant level
in the system.

Since water systems contain more than one interacting substances, separate
mass balance equations are written for each substance.

4
Example 1. Calculation of phosphorous loading and flux from bottom
sediments.

A pond with a constant volume and no outlet has a surface area of 10
4
m
2
and a
mean depth of 2 m. Initially the concentration of phosphorous in the pond is 0.8
ppm. Two days later a measurement indicates that the concentration has risen to
1.5 ppm. Calculate: (a) the mass loading of phosphorous during this time; (b)
assuming that the only possible source of this pollutant is bottom deposits
estimate the flux of the pollutant.

Solution
1. Mass loading
a. The volume of the pond is
3 4 2 4
10 2 2 10 m m m H A V
s
= = =
b. The mass of phosphorous initially present (t=0) is
g
m
g
m VC m
4
3
3 4
10 6 . 1 80 . 0 10 2 = = =
c. At time t = 2 days the mass is 3.0 x 10
4
g, i.e increase of 3.0 x 10
4
-1.6 x
10
4
= 1.4g x 10
4
.
d. The mass loading therefore is:
d g
d
g
t
m
W
4
4
10 70 . 0
2
10 4 . 1
=

= =
2. Flux of phosphorous
a. The flux of phosphorous across the bottom is:
d m g
m
d g
A
W
J
s
. 10 7 . 0
10 1
10 70 . 0
2 4
2 4
4
=

= =




5
Types of reactors

Processes used for the treatment of wastewater are physical, chemical or
biological. The individual methods are usually classified as unit processes. The
vessels and tanks, in which wastewater treatment processes are carried out, are
commonly known as reactors. In a reactor one or more reactions can be
contained or controlled at the same time.

Reactors used for the treatment of wastewater
The reactors used for the treatment of wastewater are
- the batch reactor
- the complete-mix reactor (also known as continuous-flow stirred-tank
reactor)
- the plug-flow reactor
- complete-mix reactors in series
- the packed-bed reactor
- the fluidized-bed reactor
The two reactor types that are used most commonly in the field of waste water
treatment are complete-mix and plug-flow reactors.

Batch Reactor
In the batch reactor the flow enters the reactor, is treated and then discharged.
There is no constant input or output. The liquid contents of the reactor are mixed
completely. Batch reactor is shown in Figure 1.

Figure 1. Batch reactor.

Batch reactors are often used to blend or dilute chemicals. In wastewater
treatment batch reactors are used especially for activated-sludge biological
treatment and mixing of concentrated solutions into working solutions. BOD tests
are carried out in batch reactors, although they are not mixed completely during
the incubation time.

Complete-Mix Reactor
In a complete-mix reactor (Figure 2), one or more fluid reagents are put into a
reactor and mixed completely while the reactor effluent is removed in the same
time. The actual time required to achieve completely mixed conditions will depend
on the reactor geometry and the power input. At steady-state, the flow rate in
6
equals the mass flow rate out, otherwise the tank will overflow or go empty
(transient state).

Figure 2. Complete-mix reactor.

In wastewater treatment aerated lagoons and aerobic sludge digestions can be
complete-mix reactors. A complete-mix reactor with recycle can be used for
activated-sludge biological treatment.

Plug-Flow Reactor
In a plug-flow reactor particles pass through the reactor with little or no longitudinal
mixing and exit the reactor in the same sequence in which they entered. Plug-flow
reactors can be e.g. long open tanks (Figure 3) or closed tubular reactors (shown
in Figure 4).


Figure 3. Plug-flow reactor (long open tank).




Figure 4. Plug-flow reactor (closed tubular reactor)

7
In wastewater treatment a possible application of a plug-flow reactor is e.g. a
chlorine contact basin.

Complete-Mix Reactors in Series
The series of complete-mix reactors (Figure 5) is used to model flow regime that
has patterns of both complete-mix and plug-flow reactors. If the series is
composed of one reactor, the complete-mix regime prevails. If the series consists
of two or more reactors in series, the plug-flow regime prevails.



Figure 5. Series of complete-mix reactors.

Packed-Bed Reactors
The packed-bed reactor is filled with some kind of packing material, e.g. rocks,
ceramic, or plastic. The packed-bed reactor can be operated in either the
downflow or upflow mode. Dosing can be continuous or intermittent. The packing
material can be continuous or arranged in stages. Packed-bed reactors are shown
in Figure 6.


Figure 6. Packed-bed reactors (downflow on the left, upflow in the right)
.

Fluidized-Bed Reactors
The fluidized-bed reactor (Figure 7) is quite similar to the packed-bed reactor, but
the packing material is expanded upwards by the movement of the fluid (air or
water) through the bed. The rate of the expansion can be varied by changing the
flowrate of the fluid.

8

Figure 7. Fluidized-bed reactors.

Hydraulic Characteristics of Reactors: Ideal Flow and Nonideal
Flow
As mentioned earlier, the two reactor types that are used most commonly in the
field of waste water treatment are complete-mix and plug-flow reactors.
The hydraulic flow characteristics of these reactors can be described as varying
from ideal and nonideal. The level of ideality depends on the relationships of the
incoming flow to outgoing flow. In practice the flow in complete-mix reactors and
plug-flow reactors is seldom ideal.
Nonideal flow occurs when a portion of the flow that enters the reactor arrives at
the outlet before the bulk of the flow.

Ideal flow in reactors has following requirements:
- Isometric, isobar, isothermic reactors.
- Incompressible matter.
- Stationaric state or steady state in the reactor.
- In complete-mix reactor the same concentration of the matter in every point
- A plug flow reactor can be frictional or frictionless. If it is frictional, the
friction between the flowing fluid and the wall of the reactor causes on the
interface a flow rate that equals zero.

This topic will be discussed further later in this lecture when we will explore the
use of tracer studies to determine the residence time and the degree of ideality in
reactors.

9

Reaction, Reaction Rates, and Reaction Rate Coefficients

In this section we will review the most important types of reactions, the governing
reaction formulations and the methods to calculate reaction rates, using laboratory
batch experiments.

The controlling stoichiometry and the rate of the reaction are very important factors
when selecting and designing wastewater treatment process. The number of
moles of a substance entering into a reaction and the number of moles of the
substances produced are defined by the stoichiometry of a reaction. The rate at
which a substance disappears or is formed in any given stoichiometric reaction is
defined as the rate of reaction.


Types of Reactions
There are two types of reactions that occur in wastewater treatment
homogeneous and heterogeneous (nonhomogeneous).

Homogeneous Reactions
Homogeneous reactions involve a single phase, i.e liquid, gas or solid, but liquid is
the most fundamental type in wastewater treatment. In homogeneous reactions,
the reactants are distributed uniformly throughout the fluid so that the potential for
reaction at any point within the fluid is the same. Homogeneous reactions may be
either irreversible or reversible.

A reversible reaction can proceed in either direction, depending on the relative
concentrations of the reactants and the products:

dD cC bB aA + + (7)

Irreversible reactions proceed in a single direction and continue until the reactants
are exhausted. A common exemple of an irreversible reaction is the
decomposition of organic matter which is generally represented by the equation:

O H CO O O H C
2 2 2 6 12 6
6 6 6 + + (8)

Irreversible reactions may proceed in consecutive steps, i.e. the oxidation of
ammonia to nitrates, presented in lecture 1.

Heterogeneous Reactions
Heterogeneous reactions occur between one or more constituents that can be
identified with specific sites, such as those on an ion-exchange resin in which one
of more ions is replaced by another ion. Reactions that require the presence of a
solid-phase catalyst are also classified as heterogeneous. Heterogeneous
reactions are usually carried out in packed and fluidized-bed reactors.
10

Rate of Reaction
The rate of reaction is the term used to describe the change in the number of
moles of a reactive substance per unit volume per unit time (for homogeneous
reactions) or per unit surface area or mass per unit time (for heterogeneous
reactions).

For homogeneous reactions the rate of reaction r is given by

| |
) )( (
1
time volume
moles
dt
N d
V
r = = (9)

If the number of moles, N, is replaced with the term VC, where V is the volume and
C the concentration ans assuming that the volume remains constant (e.g. no
evaporation), Eq. 9 can be expressed in terms of concentration as follows:
dt
dC
r = (10)
The plus sign indicates an increase or accumulation of the substance, and the
minus sign a decrease of the substance.

For heterogeneous reactions, where S is the surface area, the rate of the reaction
r is given by:

| |
) )( (
1
time area
moles
dt
N d
S
r = = (11)

The rate Law
The rate of reaction can generally be represented by a relationship:

( )
,
,
B A
A
C C f k
dt
dC
= (12)

This relationship is called a rate law. It specifies that the rate of reaction is
dependent on the product of a temperature-dependent constant k, and a function
of the concentrations of the reactants ( ) , ,
B A
C C f . A common form of Eq. 12 is:


B A
A
C C k
dt
dC
= (13)

The power to which the reactants are raised is referred to as the reaction order,
n. The reaction order of Eq. 13 is: + = n .

In this lecture we focus on a single reactant. For this case Eq. 13 is simplified as:

n
C k
dt
dC
= (14)

11
Types of Rate Expressions
The most common types of rate reactions occurring in wastewater treatment are
the following.

k
dt
dC
= (zero-order) (15)
kC
dt
dC
= (first-order) (16)
2
kC
dt
dC
= (second-order) (17)
C K
kC
dt
dC
+
= (saturation or mixed-order) (18)


(a) Zero-order
A zero-order reaction has a rate which is independent of the concentration of the
reactants.

Eq. 15 can be integrated by separation of variables to yield:

kt C C =
0
(19)

Where C
0
the concentration of the reactant at time t=0. Thus if a plot of
concentrations versus time yields a straight line, we can conclude that the reaction
is zero-order.

(b) First-order

First-order reactions proceed at a rate which is directly proportional to the
concentration of the reactant. Eq. 16 can be integrated by separation of variables
to yield:

kt
e C C
kt C C
kdt
C
dC

=
=
=
0
0
ln ln (20)

This model specifies an exponential decay, i.e. the concentration halves per unit
time. First order reactions are the most frequent type of reaction used in
wastewater treatment and environmental modelling.

(c) Second-order
Second-order reactions proceed at a rate proportional to the second power of a
single reactant. Eq. 17 can be integrated by separation of variables to yield:

12
t kC
C C
kt
C C
kC
dt
dC
0
0
0
2
1
1
1 1
+
=
+ =
=
(21)

(d) Saturation type
The saturation type of reaction (Eq. 18), has two rate coefficients, k and K. K is
also known as the half saturation coefficient because its value is equal to the
concentration of substrate when the reaction rates is half its maximum value. It is
most often used to model the growth of microorganisms and the substrate
utilization rate in biological systems.

A plot of this type of reaction, for two values of the saturation coefficient K is
shown in figure 8. At high C concentrations, the rate of reaction approaches zero-
order, and at low C concentration, the rate of the reaction approaches first-order.

Eq. 18 can be integrated to give:

( )
t
t
C C
C
C
K kt
dC
k C
dC
k
K
dt
k C k
K
dC
dt
C K
kC
dt
dC
=
. =
. =
.
+
=
0
0
ln
1
1 1
(22)

0
0,1
0,2
0,3
0,4
0,5
0,6
0,7
0,8
0,9
1
0 20 40 60 80 100
Concentration, C, mg/L
r
e
a
c
t
i
o
n

r
a
t
e
,

r
K=1
K=10
k = 1

Figure 8. Rate of reaction versus concentration for a saturation-type expression.


13
Analysis of rate data
Reaction rate coefficients are detrmined by using results obtained from batch
experiments. The coefficients can be determined by (a) the integral method and
(b) by the differential method.

(a) The integral method
In the integral method the integrated forms of equations 15-18 are used. Graphical
methods are then employed to determine whether the model fits the data
adequately. Table 1 summarizes the plotting data for these types of reactions:

Table 1. Summary of plotting characteristics with the integral method

Order Dependent
(y)
Independent
(x)
Intercept Slope
Zero (n=0) C t C
0
-k
First (n=1 lnC t lnC
0
-k
Second (n=2) 1/C t 1/C
0
k
Saturation C C t
0
ln 1 ( ) t C C
t

0

k/K 1/k

Example 2. Determination of reaction order and rate of reaction
You perform a series of batch experiments and come up with the following data.
Determine the order (n) of reaction, the rate coefficients.

t (hr) 0 2 4 6 8 10
C, g/L 10.5 5.1 3.1 2.8 2.2 1.9

Solution

1. Plots to evaluate the rate of reactions are shown in figure 9. The second
order reaction model gave the best regression coefficient, with the saturation
model also giving a satisfactory fit. The best-fit equation is:
t
C
0424 . 0 1135 . 0
1
+ =
k = 0.0424 L/g/hr
C
0
= 1/0.1135 = 8.81
t t kC
C C
37 . 0 1
1
81 . 8
1
1
0
0
+
=
+
=

2. Alternatively if we assume the saturation model, then the best fit equation is:

56 . 7 1323 . 0 1 = = k g/L.hr and
35 . 197 0383 . 0 56 . 7 0383 . 0 / 0383 . 0 = = = = k K K k g/L

14
Zero order
R
2
= 0,7298
0
2
4
6
8
10
12
0 2 4 6 8 10 12
Time, t , hr
C
,

g
/
L

First order
R
2
= 0,8908
0
0,5
1
1,5
2
2,5
0 2 4 6 8 10 12
Time, t , hr
l
n
C

Second order
y = 0,0424x + 0,1135
R
2
= 0,9832
0
0,1
0,2
0,3
0,4
0,5
0,6
0 2 4 6 8 10 12
Time, t, hr
1
/
C

Saturation
y = 0,1323x + 0,0383
R
2
= 0,9482
0
0,1
0,2
0,3
0,4
0,5
0 0,5 1 1,5 2 2,5 3
(C0 -Ct )/t
1
/
t

l
n
(
C
0
/
C
t
)


Figure 9. Plots to evaluate reaction rate model


(b) The differential method
The differential method applies a logarithmic transform of Eq. 14 to give:
C n k
dt
dC
log log log + =
|
.
|

\
|
(23)
Therefore if the general equation holds, a plot of ( ) dt dC log versus logC should
yield a straight line with a slope of n and an intercept of logk. The method has the
advantage that it automatically gives an estimate of the order of the reaction. The
difficulty is in estimating the derivative of C. The centred difference approach is
often employed to estimate the derivative:

1 1
1 1
+
+

= ~
i i
i i i
t t
C C
dt
dC
dt
dC
(24)

This method was already employed in Lecture 1 to estimate the kinetic parameters
of the BOD test.

(c) The method of initial rates
There are cases where when conducting batch experiments, complications may
arise over time, such as, reverse reactions, or very slow reactions. For such cases
15
the method of initial rates uses data from the initial stages of experiments to
determine the rate constant and the order of a reaction. A series of experiments is
carried out using different initial concentrations C
0
. For each experiment the initial
rate ( ) dt dC
0
is determined by differentiating the data and extrapolating to zero
time. A plot of ( ) dt dC
0
log versus log C
0
is used to estimate k and n.

Effects of Temperature on reaction Rate Coefficients
Most wastewater treatment operations and processes are carried out over a
relatively narrow temperature range. The temperature dependence of rate
constant is given by the vant Hoff-Arrhenius relationship.

) (
(
ln
1 2
2 1 2 1 2 1
) 1 2
1
2
T T
T RT
E
T RT
E
T RT
T T E
k
k
= =

= (25)
where:
k
2
= reaction rate constant at temperature T
2
(1/s, 1/min, etc.)
k
1
= reaction rate constant at temperature T
1
(1/s, 1/min, etc.)
T = absolute temperature, K = 273.15+TC
E = a constant characteristics of the reaction, J/mol
R = ideal gas constant, (8.314 J/mol.K)

Because the temperature differences are not great, the following relationship can
be used to adjust the value of the operative rate constant to reflect the effect of
temperature:

) (
1 2
1
2 T T
T
T
k
k

= (26)
Typical values for vary from about 1.020 to 1.10 for some biological treatment
systems. The temperature is expressed in C.

Example 3. Determination of temperature coefficient

From a laboratory experiment you are given the following results for a reaction.
Estimate for this reaction and determine the rate at 20C.
:

T, C k, d
-1

4 0.12
16 0.20
Solution
1. Taking the logarithm of Eq. 26 we get:

( ) 0435 . 1 4 16 log log log
4 16
= = k k

2. Equation 20 is then used to compute k at 20C.

( )
237 . 0 0435 . 1 20 . 0
16 20
20
= =

k d
-1

16
Mass-Balance Analysis

The fundamental approach to study the hydraulic flow characteristics of reactors is
the mass-balance analysis. With the mass-balance analysis the changes that take
place in reactor when a reaction is occurring, can be delineated. By accounting for
material entering and leaving a system, mass flows can be identified which might
have been unknown, or difficult to measure without this technique.

The general form quoted for a mass balance is The mass that enters a system
must, by conservation of mass, either leave the system or accumulate within
the system. Mathematically the mass balance for a system without a chemical
reaction is as follows:

Input = Output + Accumulation

The system boundary for mass balance is helping to identify all of the liquid and
constituent flows into and out of the system. The control volume boundary is used
to identify the actual volume in which the change is occurring. The system and the
control volume boundary are illustrated in Figure 10. In most cases, these
boundaries will coincide.


Figure 10. The system boundary and the control volume boundary for mass
balance.

With a chemical reaction the mass balance is as follows:

Rate of
accumulation of
reactant within
the system
boundary
=
rate of flow
of reactant
into the
system
boundary

rate of flow
of reactant
out of the
system
boundary
+
rate of
generation of
reactant within
the system
boundary

The corresponding statement is simplified as:
17

Accumulation = inflow outflow + generation

Depending on the flow regime or the treatment process, one or more of the terms
can be equal to zero. For example, in a batch reactor in which there is no inflow or
outflow the second and third terms will be equal to zero.

Application of the Mass-Balance Analysis
To apply a mass balance analysis for a complete-mix reactor the following
assumptions are made:

1. The volumetric flowrate into and out of the control volume is constant.
2. The liquid within the control volume is not subject to evaporation (constant
volume)
3. The liquid within the control volume is mixed completely.
4. A chemical reaction involving a reactant A is occurring within the reactor.
5. The rate of change in the concentration of the reactant A that is occurring
within the control volume is governed by a first-order reaction (r
c
= kC)

Using these assumptions the mass balance can be formulated as follows:

V r QC QC V
dt
dC
c
+ =
0
(27)

Substituting kC for r
c
yields

V kC QC QC V
dt
dC
) (
0
+ = (28)

where:
dC/dt = rate of change of reactant concentration within the control volume, kg/ m
3
.d
V = volume contained within control volume, m
3

Q = volumetric flowrate into and out of control, m
3
/d
C
0
= concentration of reactant entering the control volume, kg/m
3

C = concentration of reactant leaving the control volume, kg/m
3

r
c
= first-order reaction rate, kg/ m
3
.d
k = first-order reaction rate coefficient, d
-1


Steady-State Simplification
Fortunately, in most applications in the field of wastewater treatment, the solution
of mass-balance equations can be simplified, because the rate accumulation is
zero (dC/dt = 0). Thus Eq. 28 yields:

V r QC QC
c
+ =
0
0 (29)
and
( )
0
C C
V
Q
r
c
= (30)

18
The solution to the expression given by Eq. 30 will depend on the nature of the
rate expression (e.g. zero-, first-, or second-order).

Example 4. Calculation of reactor volume
A wastewater is to be treated in a continuous flow complete mix reactor. Assuming
that the reaction is irreversible and first order with a reaction rate coefficient k=0.15
d
-1
: (a) determine the flowrate that can be treated if the reactor has a volume of
2000 m
3
and 98% treatment efficiency is required. (b) What volume would be
required to treat the flow rate determined in (a) if the required treatment efficiency
was 92%?

Solution
1. Determine flow rate
a. Determine effluent concentration as a function of influent concentration
given treatment efficiency E:
0
0
0
02 . 0 98 . 0 C C
C
C C
E = =

=
b. Rearranging Eq. 29, replacing r
c
= kC and solving for Q yields:

( )
d m m d
C C
C kV
C C
kCV
Q
kCV C C Q
3 3 1
0 0
0
0
0
12 . 6 2000 15 . 0
98 . 0
02 . 0
02 . 0
02 . 0
= =

=
=


2. Determine the volume reduction for a decreased efficiency
a. Solving Eq. 29 for V and taking into assumption the new effluent efficiency
yields:
( )
3
1
3
0
4 . 469
08 . 0
92 . 0
15 . 0
12 . 6
m
d
d m
kC
C C Q
V = =

=


b. The volume reduction obtained when efficiency is reduced from 98% to
92% is 5 . 76 100
2000
469 2000
=

%
19

Modelling Flow in Reactors using tracer studies

Modelling of the hydraulic characteristics of reactors is important because the
results can be used to determinate the actual amount of time that a given volume
of water will remain in the reactor and its average age. Also the achieved degree
of treatment can be defined, based on applicable kinetics. Comparison of actual
hydraulic characteristics of a reactor can be measured using tracers. Actual to
theoretical response can be used to define how well the ideality has been
achieved.

Two types of tracer tests are most commonly used: (a) the pulse input, where an
amount of non-reactive tracer is injected into the reactor with a continuous inflow
of clear water and (b) the step input where a conservative tracer is continuously
injected in the reactor initially filled with clear water. Characteristic loading
functions versus time are shown in figure 11.


Figure 11. Loading function versus time (Adapted from Ref [2])


The mathematical analysis for these two tracer tests and for the two types of
reactors, using the equations 27-30 are described in the following paragraphs.

20
Complete-Mix Reactor
(a) Pulse input
The injection of a pulse dose of conservative tracer in the reactor, creates an
initial concentration of tracer in the reactor C
0
. The effluent tracer concentration, C
can be determined from Eq. 27 by noting that C
0
= 0 (continuous flow of clear
water) and r
c
= 0 (conservative tracer) .

Equation (1), yields:
C
V
Q
dt
dC
= . (31)

Integrating between the limits C= C
0
to C=C and t=0 to t=t yields

=
t C
C
dt
V
Q
C
dC
0
0
(32)
After integration

( )
= = = e C e C e C C
t V Q t
0 0 0
(33)
where:
C = concentration of the tracer in the reactor at time t, kg/m
3
C
0
= initial concentration of the tracer in the reactor, kg/m
3

t = time, sec
Q = volumetric flowrate, m
3
/sec
V = reactor volume, m
3

= theoretical detention time = V/Q, sec
= normalized detention time t/, unitless

(b) Step input

The corresponding response for a continuous step input of tracer which is mixed
instantaneously (Figure 8, a-2) is

( ) ( ) ( )

= = = e C e C e C C
t V Q t
1 1 1
0
/
0
) / (
0
(34)

Plug-Flow Reactor
(a) Ideal flow
The plug flow reactor can be treated as a continuum of completely mixed reactors
with differential volume. If an observer was standing at the outlet of the reactor we
would observe the appearance of the tracer in the effluent distributed over the
cross sectional area, after a time equal to the detention time has elapsed.

The materials balance for a nonreactive tracer is written for the differential volume
element (shown in Fig. 12) as follows:


x x
x
QC QC V
t
C
A +
= A
c
c
(35)
21

where:
t C c = rate of change of tracer concentration, kg/m
3

V = differential volume element, m
3

t = time, sec
Q = volumetric flow rate, m
3
/sec
x = some point along the reactor length, m
x = differential distance, m





Figure 12. The definition sketch for the hydraulic analysis of a plug-flow reactor
with (1) advection only and (2) with advection and axial dispersion.

The change in concentration with time term is written as partial differential
because the concentration is also changing with distance. Substituting the
differential form for the terms QC|
x
and QC|
x+x
yields:

|
.
|

\
|
A
A
A
+ = A
c
c
x
x
C
C Q QC V
t
C
(36)

Substituting V to its equal Ax, where A is the cross-sectional area in the x
direction, and simplifying yields:

x
x
C
QC x A
t
C
A
A
A
= A
c
c
(37)


Dividing by A and x yields

x
C
A
Q
t
C
A
A
=
c
c
(38)

Taking the limit as x approaches zero yields

x
C
u
x
C
A
Q
t
C
c
c
=
c
c
=
c
c
(39)

22
where u = velocity of flow, m/sec.

Because both ends of the equation are the same (note c t=c x/u, except for the
minus sign, the only way that the equation can be satisfied is if the change in
concentration with distance is equal to zero. Thus, concentrations of effluent and
influent must be equal.

(b) Nonideal flow
As mentioned earlier, ideal flow conditions never occur in practice and thus a
portion of the flow that enters the reactor arrives at the outlet before the bulk of the
flow. Factors causing nonideal flow include: (a) temperature differences in the bulk
of the liquid in the reactor; (b) wind driven circulation patterns; (c) inadequate
mixing and (d) poor design.

In plug flow reactors an additional cause of nonideality is axial dispersion.
Whereas the movement of the tracer with the current velocity described above (a)
for ideal flow conditions is termed advection, additional movement is caused by
dispersion brought about by velocity differences, turbulent eddies and molecular
diffusion. Remember that these terms were defined briefly in the first section of
this lecture.

(c) Distinction between molecular diffusion, turbulent diffusion and
dispersion

Molecular diffusion, D
m
= the mass transfer of material under no-flow conditions
due to gradients of concentration in the liquid.

Turbulent diffusion, D
e
or E: The mass transfer brought about by microscale
turbulent mixing without flow.

Dispersion, D: The longitudinal spreading of a tracer caused by advection due to
differences in the fluid velocity, turbulent eddies and molecular diffusion.

Table 2. Typical values for the three diffusion coefficients.

Coefficient Range of values
cm
2
/s
Molecular diffusion, D
m
10
-8
- 10
-4

Turbulent diffusion, D
e
10
-4
- 10
-2

Dispersion, D 10
2
10
6


Plug flow reactor with axial dispersion (one-dimensional problem)
The transfer of mass by diffusion is a function of the concentration gradient along
the reactor and can be represented by:

x
C
D r

= (40)
The material balance for the transport of tracer by advection and dispersion is
derived by adding a diffusion tern to equation (35):
23

x x x
x
C
AD uAC
x
C
AD uAC x A
t
C
A +
|
.
|

\
|
A
A

|
.
|

\
|
A
A
= A

(41)

where: D = coefficient of axial dispersion

Taking the limit of equation (41) results in the expression:

x
C
u
x
C
D
t
C

=
2
2
(42)

or by noting that t
u
x

=

t
C
x
C
D
t
C

=
2
2
(43)

Equation (43) has been solved for a unit pulse input and for two conditions:

(a) for small amounts of axial dispersion, resulting in a symmetrical output:

( )
( )
( )


=
uL D uL D
C
4
1
exp
2
1
2

(44)

The mean and variance of this equation which has the same general form as the
normal probability curve are:
1 = =

c
t
(45)

uL
D
c
2
2
2
2
= =

(46)

Where c t and
c
the mean detention time and variance calculated from
experimental measurements (see below), t = normalized time and Q V = the
theoretical detention time.

(b) for large amounts of axial dispersion, resulting in an asymmetrical output:

( )
( )
( )


=
uL D uL D
C

4
1
exp
2
1
2
(47)

where: L = characteristic length.


The corresponding mean and variance are:
24
uL
D t
c
2 1+ = =

(48)
2
2
2
2
8 2
|
.
|

\
|
+ = =
uL
D
uL
D
c

(49)

The term uL D d = is a unitless dispersion number and gives an estimate of the
magnitude of dispersion. The inverse of d is known as the Peclet number of
longitudinal dispersion, D uL P
e
= .

The ideal plug flow reactor is represented by d=0 whereas the ideal complete mix
reactor by d = .

Use of tracer curves to determine the flow characteristics of a
reactor
The flow characteristics of a reactor can be determined experimentally by the use
of tracer curves. A conservative tracer, a dye or salt, is introduced in the inlet of
the reactor, either as a slug dose pulse or as a continuous step input.
Grab samples are taken from the effluent at discrete times until no more tracer is
detected in the effluent (pulse dose) or the concentration in the effluent matches
the influent concentration (step dose).

The mean detention time, c t can be determined from the experimental
measurements of concentrations C
i
at times t
i
:

A
A
=
i i
i i i
c
t C
t C t
t (50)

The mean variance is defined as:

( )
2
2
t
t C
t C t
i i
i i i
c

A
A
=

(51)

The mean variance is then used to determine the dispersion number from Eq. 46
or 48 and 49.

Typical dispersion numbers for various wastewater treatment processes are
shown in table 3.

Table 3 Dispersion numbers for wastewater treatment processes(*).

Treatment process Dispersion
number
Rectangular sedimentation tanks 0.2 2.0
Activated sludge tanks
Long plug flow 0.1 1.0
Complete mix 3.0 - 4.0+
25
Oxidation ditch (Carrousel systems) 3.0 - 4.0+
Waste stabilization ponds
Single ponds 1.0 - 4.0+
Multiple cells in series 0.1 1.0
Chlorine contact basins 0.02 0.04
(*) Adapted from reference [1]

Example 5. Determination of mean residence time and dispersion
index fro tracer studies

Using the following data obtained from a dye tracer test to evaluate the hydraulic
performance of a chlorination tank determine (a) the mean residence time and
variance for the tracer response curve (b) the dispersion number and the
coefficient of dispersion (c) plot the tracer curve and the simulated curve. The flow
rate at the time the experiment was conducted was 15.000 m
3
/d. The chlorination
tank is 36.6 m long 3.0 m wide and has a water depth of 4.9 m.


Time
(min)
Concentration
(g/L)
Time
(min)
Concentration
(g/L)
0 0 54 11
6 0 60 9
12 0 66 4,3
18 0 72 3
24 0,3 78 1
30 1,8 84 0,2
36 4,5 90 0
42 8 96 0
48 11 102 0


Solution
1. Determination of mean residence time and variance using equations (50) and
(51).
a. Setup a computational table to solve the equations. The
t A
was omitted as it
is constant (6 min) for all measurements and it appears both on the
numerator and denominator of the equations:

Time t, min
Conc. C,
g/L t x C t
2
x C
0 0
6 0 0 0
12 0 0 0
18 0 0 0
24 0.3 7.2 172.8
30 1.8 54.0 1620.0
36 4.5 162.0 5832.0
42 8.0 336.0 14112.0
48 11.0 528.0 25344.0
26
54 11.0 594.0 32076.0
60 9.0 540.0 32400.0
66 4.3 283.8 18730.8
72 3.0 216.0 15552.0
78 1.0 78.0 6084.0
84 0.2 16.8 1411.2
90 0 0 0
96 0 0 0
102 0 0 0
Total 54.1 2815.8 153334.8

b. Determine the meat residence time and variance

h 87 . 0 min 05 . 52
1 . 54
8 . 2815
= = =
A
A
=

i i
i i i
c
t C
t C t
t

( )
2 2
2
2
min 28 . 125 05 . 52
8 . 2815
8 . 334 , 153
= =
A
A
=

t
t C
t C t
i i
i i i
c


2. Determination of the dispersion number and the dispersion coefficient.
a. Determine the theoretical detention time for the chlorination tank

min 51.65 h 0.86 d 0,036
000 , 15
9 . 4 0 . 3 6 . 36
= = =

= = Q V

b. Determine the normalized time using equation (19)
1
86 . 0
87 . 0
~ = =

c
t
(The curve is approximately symmetrical)
c. Determine the dispersion number using equation (20)
0,02348
65 . 51
28 . 125
2
1
2
1
2 2
2 2
2
2
2
2
= = =
= = =

c
c
d
d
uL
D

d. Determine the dispersion coefficient
m/sec 0,0118
60 60 24
1
9 . 4 0 . 3
15000
=

= =
A
Q
u

0,014 6 . 36 0118 . 0 02348 . 0 = = = L u d D

e. Based on the dispersion coefficient the chlronination tand can be defined
as low dispersion (d less than 0.02).

3. Plot the tracer curve and the simulation curve, using equation (18)

a. The two curves as shown in the following figure.
27


0
2
4
6
8
10
12
0 20 40 60 80 100 120
Time, min
C
o
n
c
e
n
t
r
a
t
i
o
n
,

g
/
L

Figure 13. Concentration versus time tracer curve for the chlorination tank


0
0,5
1
1,5
2
0 0,5 1 1,5 2
Normalized time, = t/
N
o
r
m
a
l
i
z
e
d

c
o
n
c
e
n
t
r
a
t
i
o
n
,

C


=

C
/
C
0
d = D/uL = 0,0235


Figure 14. Normalized concentration versus time tracer curve

28
Bibliography

1. Metcalf and Eddy, Wastewater Engineering: Treatment and Reuse, 4th
Edition, Mc Graw Hill, New York (2003).
2. S.C. Chapra Surface water quality modelling, McGraw-Hill (1997)

Вам также может понравиться