Sie sind auf Seite 1von 21

Chemical Product and Process

Modeling
Volume 2, Issue 2 2007 Article 3
PAPERS FROM C HEMECA 2006

Simulation of Free Chlorine Decay and


Adaptive Chlorine Dosing by Discrete
Time-Space Model for Drinking Water
Distribution System
Abrar Muslim∗ Qin Li†
Moses O. Tadé‡


Curtin University of Technology, Perth, Western Australia, abr muslim@yahoo.com

Curtin University of Technology, Perth, Western Australia, q.li@curtin.edu.au

Curtin University of Technology, Perth, Western Australia, m.o.tade@curtin.edu.au

Copyright 2007
c The Berkeley Electronic Press. All rights reserved.
Simulation of Free Chlorine Decay and
Adaptive Chlorine Dosing by Discrete
Time-Space Model for Drinking Water
Distribution System∗
Abrar Muslim, Qin Li, and Moses O. Tadé

Abstract

An ideal drinking water distribution system (DWDS) must supply safe drinking water with
free chlorine residual (FCR) in the form of HOCl and OCl- at a required concentration level. The
FCR as a disinfectant decreases over time due to chemical reactions in the bulk phase and at the
pipe wall in the DWDS. In order to supply drinking water with the FCR concentration within
the safety range of 0.2-0.6 mg/l at the points of water consumption, it is important to develop
a dynamic model of the FCR using a discrete time-space model (DTSM) that accounts for free
chlorine transport in the axial direction by convection, diffusion and the decay kinetic. A DTSM
has been developed using Finite Difference Method (FDM) to predict the FCR in single pipes in
the DWDS. The DTSM has been computed using Matlab 7.0.1 and tested with step inputs and
rectangular pulse inputs to estimate the FCR at any point in the pipes over time. Data found in the
literature have been used to validate the DTSM. The modelling and simulation study shows that
the water velocity significantly affects the FCR concentration distribution along the pipe. Due to
the fluctuation of the drinking water demand, a model-based adaptive chlorine dosing scheme is
proposed to control the proper injection of chlorine.

KEYWORDS: free chlorine residual, discrete time-space model and simulation, model-based
adaptive chlorine dosing


The authors are very grateful to Curtin University of Technology for technical support and pro-
viding the Matlab 7.0.1 software used in this research. The authors would also like to thank the
anonymous referees of Journal of Chemical Product and Process Modeling (JCPPM) for their
valuable suggestions to improve the readability and quality of this paper.
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

INTRODUCTION
Drinking water should be free of pathogenic microorganisms that can cause
water-borne diseases. According to World Health Organization (2004), almost 3.3
million people (mostly children) die every year from these diseases. Thus,
reducing the level of pathogenic microorganisms in water that cause these
diseases is still a major objective, particularly in developing countries and
regions.
There are many water treatment processes that may contribute to pathogen
reduction, but according to Vasconcelos et al. (1997) chlorination is the most
widely used method for water disinfection. As a disinfectant, when gaseous
chlorine is dissolved in water, the hydrolysis reaction of chlorine occurs to form
hypochlorous acid (HOCl). The complete reaction occurs in a few tenths of a
second at 18oC and within a few seconds at 0oC at normal pH ranges (6.5 to 8.5).
Hypochlorous acid and the hypochlorite ion, known as the FCR, can kill pathogen
in several seconds (Connell, 1996).
On one hand, the FCR in the DWDS decays due to its reaction with
oxidisable compounds (Wei and Morris, 1974) and it rapidly decays with distance
from the treatment plant (Maul et al., 1985). On the other hand, the DWDS must
supply safe drinking water with the FCR concentration in the range of 0.2-0.6
mg/L and a maximum amount of disinfectant by-product of 0.1 mg/L that must be
presented at the points of water consumption (AWWA, 2003).
A number of researchers have developed chlorine decay models to predict
chlorine concentration in drinking water. A chlorine consumption model,
Ct = C1t n was developed by Feben and Taras (1951) in which Ct is the chlorine
consumed (parts/million) at time, t (hr); C1 is the amount of chlorine consumed
after one hour of contact; and n is the exponential constant characteristic of the
given water. The long-term chlorine decay model for distributed drinking water
and natural waters, Ct = C0 exp( kt ) , has been modeled by Johnson (1978) using
first-order kinetics which is the same as the model developed by Heraud et al.
(1997), where Ct is the chlorine concentration at time t; C0 is the initial
concentration; t is time (hr) and k represents a first-order reaction rate constant
(hr-1). Rosman et al. (1994) proposed a one dimensional mass conservation
equation for a dilute concentration of total FCR in water flowing through a
section of a pipe using the Discrete Volume Element Method. However, their
results on chlorine residual are only resolved with time.
This study presents the model development and simulation of the FCR
using a discrete time-space model (DTSM) resolving both temporal and spatial
distributions, and also proposes a model-based adaptive chlorine dosing to
maintain the FCR concentration along the single pipes of the DWDS.

Published by The Berkeley Electronic Press, 2007 1


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

MODEL DEVELOPMENT

Figure 1. A schematic drawing of the model for free chlorine residual transport
in a single pipe of the DWDS

A single input single output model, as shown in Figure 1 for the FCR
concentration in a pipe of the DWDS was developed using a mechanistic
approach. Here a space element x and time element t are chosen to develop the
overall mass balance by considering the FCR in the form of both HOCl and OCl-,
and assuming that the water flow in the axial direction only. The overall mass
balance equation can be written as below:

C C C
V = F1C1 F2 C 2 k C V + A D( |x x |x ) (1)
t x x

which can be discretized as:


t
( V )C t ( V )C t t = F1C1 x x F2C2 x dt +
t t
t t
C1 C2
( D1 A x x) ( D2 A x ) dt [ kC V ]dt (2)
t t
x x t t

where,
V = The volume of the imaginary control volume
A = The cross section area of water flow in the pipe (m2)
C = The FCR concentration in the control volume, CV (g/L)
C1 = The FCR concentration enters the CV, reaching the axial direction x -
x (g/L)
C2 = The FCR concentration enters the CV, reaching the axial direction x
(g/L)

http://www.bepress.com/cppm/vol2/iss2/3 2
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

D1 = The effective diffusivity rate constant of the FCR enters the CV,
reaching the axial direction x - x (m2/s)
D2 = The effective diffusivity rate constant of the FCR enters the CV,
reaching the axial direction x (m2/s)
F1 = The inflow water enters the CV, reaching the axial direction x - x
(L/s)
F2 = The outflow water enters the CV, reaching the axial direction x (L/s)
k = The overall chlorine decay rate constant in the CV (s-1)

Equation (2) can be simplified into a dynamic model for the FCR transport in the
axial direction by convection and diffusion:

2
C C C
= v +D 2 kC (3)
t x x

where k is the overall chlorine decay constant (s-1), v is the water flow velocity
(m/s) and D is chlorine effective diffusivity coefficient (m2/s). The Peclet number,
Pe, can be calculated as Pe = Lv/D, where L is the pipe length (m). When the Pe
>> 1 (turbulent condition), the governing equation for the FCR in pipes under this
condition is

C C
= v kC (4)
t x

In Equation (4), the FCR transport dynamic model in the axial direction by
convection is almost the same as the one dimensional conservation of mass
equation for a dilute concentration of total free chlorine proposed by Rosman et
al. (1994).

Solution Procedure
The DTSMs of chlorine decay with axial diffusion in a single pipe of the DWDS,
as the numerical solutions of Equations (3) and (4) can be obtained by adopting
the explicit scheme of the FDM of Euler method (Tveito and Winther, 1998) in
discretizing both time and space (pipe length). The concentration C at [i x,
(j+1) t] which is the point of time and space needed to predict Ci j +1 .

[C i
j +1
]
Ci j = v
t
x
Ci j 1 v
t
x
Ci j

Published by The Berkeley Electronic Press, 2007 3


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

+ D
t Ci j 1
x x
2D
t Ci j
x x
+D
t Ci j+1
x x
[k tC i j ] (5)

Bring the C i j in the LHS to the RHS, the discrete time-space model (DTSM) of
chlorine decay with axial diffusion in a single pipe of the DWDS, is

t D t t t
Ci j +1 = v+ Ci j 1 + 1 v 2D k t Ci j + D Ci +j 1
x x x x2 x
(6)
where,
C i j 1 = The FCR concentration at [(i-1) x, j t] (g/L)
Ci +j 1 = The FCR concentration at [(i+1) x, j t] (g/L)
Ci j = The FCR concentration at [i x, j t] (g/L)
j +1
Ci = The FCR concentration at [i x, (j+1) t] (g/L)
D = The effective diffusivity constant of chlorine in water (m2/s)
i = The time steps, i = 1,…,m-1
j = The space steps, j = 0, …, n-1
k = The overall chlorine decay rate constant in the V (s-1)
t = T/m, T is duration of time for chlorine transport in the pipe and m is
number of time steps in the time element
v = The water velocity in the pipe (m/s)
x = L/n, L is the length of single pipe and n is the number of space steps in
the space element.
The initial concentration of chlorine in the system is given by the equation below:
C i0 = a given value for i = 1, m – 1 (7)
The upstream concentration of chlorine in the pipe from the injection point is
given by Equations (8) and (9):
C 0j = a given value for j = 1,…, n (8)
C0j +1 = a given value for j = 1,…, n (9)
In the turbulent condition, where the axial diffusion term can be neglected, by
discretizing the time and space (pipe length), Equation (4) can be written as the
following:

http://www.bepress.com/cppm/vol2/iss2/3 4
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

[C i
j +1
]
Ci j = v
t
x
Ci j 1 v
t
x
Ci j [k tC i j ] (10)

This can be rearranged into:

t t
Ci j +1 = v Ci j 1 + 1 v k t Ci j (11)
x x

In this case, the boundary condition of upstream chlorine concentration for


Equation (10) is reduced into Equation (8) only.

Analytical Solution
For the DWDS, the axial Peclet number, Pe is typically found to be very large
(>>1). Thus, it is not necessary to develop the analytical solution for the dynamic
model for FCR transport in the axial direction diffusion. To obtain the analytical
solution, a coordinate transformation is needed, together with the Chain Formula
(O’Neil, 1987), to convert Equation (3) to the ordinary differential equation form
by substituting the coordinate transformation derivatives and using the initial
chlorine concentration C (t = 0) = Co . The analytical solutions are written as the
follows:
Ct = C0 exp( kt ) (12)

C x = C0 exp( kx / v) (13)

where Ct and Cx (mg/L) are the FCR concentrations after the residence time of t
(hr) and the distance of x (km), respectively, C0 is the initial concentration, v is the
water velocity (km/hr) and k represents the overall chlorine decay rate constant.

RESULTS AND DISCUSSION


The DTSM and the analytical solution
The DTSM that is the numerical solution of Equation (4) is given in Figure 2.b
with step input shown in Figure 2.a. The analytical solution of the DTSM
(DTSMAS) is shown by Figure 2.c. The k and v for the model simulation are
obtained from the paper of Rosman et al. (1994). The DTSM above shows the
FCR concentration along the space over time with the k of 0.208/hr, the v of 0.36
km/hr and the initial chlorine concentration (Co) of 1 mg/L, injected at the point
of 0 km where the first chlorine booster is installed.

Published by The Berkeley Electronic Press, 2007 5


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

(a)

(b)

http://www.bepress.com/cppm/vol2/iss2/3 6
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

(c)
Figure 2. Comparison of the DTSM for different injection time against the DTSM
analytical solution (DTSMAS) on the FCR decay in a single pipe of DWDS

In Figure 2.b, the first curve from left side is the FCR concentration after 1 hour
step input. On its right side, sequentially, the 2nd curve, 3rd curve and the
following curves represent the FCR concentration at 2, 3 up to 12 hours (injection
time, T=12hrs) respectively, until the last curve which represents the DTSM step
response at 50 hours. Figures 2.b and 2.c show that the numerical solution via
DTSM is able to capture the transient behaviour of FCR distribution, and the step
input is long enough for steady status to develop the numerical solution matches
the analytical solution. In this simulation, the DTSM needs at least 12 hours
injection time (step input mode) in order to fit the response with the analytical
from 0 to 4 km of the space length. Overall, the chlorine decay model follows
exponential decay trend as concluded by the DTSM simulation, as also reported
by Johnson (1978).

Considering the effective diffusivity coefficient of chlorine in the DWDS


So far we have not found literature that presents comparisons between the
chlorine decay model results in pipes of a DWDS with and without considering
the chlorine effective diffusivity coefficient (CEDC), apart from the work of
Tzatchokov et al. (2003), where they proposed a non-steady advection dispersion
reaction model with consideration of a fluoride effective diffusivity coefficient in

Published by The Berkeley Electronic Press, 2007 7


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

the simulation. Therefore, there is a need to simulate the DTSM when considering
the CEDC.

(a)

(b)

http://www.bepress.com/cppm/vol2/iss2/3 8
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

(c)
Figure 3. Comparison of the DTSM using the reference CEDC against the DTSM using
a simulated critical CEDC on the FCR decay by axial diffusion in a DWDS single pipe

Figure 3 shows the DTSM results with 12 hours (injection time, T=12hrs) of
rectangular pulse input (RPI) shown in Figure 3.a. The response during the RPI is
shown by the first curve from the left side and the responses after the RPI, at
every 6 hours, for the elapsed time of 6 to 36 hours (Te=6:6:36) are shown on the
right side. The simulation parameters are taken from the example of the Brushy
Plain-Cherry Hill network (Biswas et al., 1993). We used the data for the pipe
number 10 for the smallest CEDC in the water, Da = 1.8 x 10-5 m2/s or 6.48 x 10-8
km2/hr with the smallest water flow velocity of 0.0504 km/hr and the chlorine
decay constant of 0.02304/hr. The large CEDC, Db = 1.03 x 10-2 m2/s or 3.708 x
10-5 km2/hr, which is ten times the value normally used and is the largest CEDC
in the network (the pipe number 1). These parameters are reasonable because the
largest axial CEDC, at the smallest water flow velocity, results in the smallest
axial Peclet number that leads to the largest axial CEDC (U.S. EPA, 2002).
However, as revealed by the main plot in Figure 3.b, the FCR
concentration along the pipe over time, with the CEDC of 1.8 x 10-5 m2/s (Da),
appear the same as the FCR concentration without taking into account CEDC.
Moreover, when the largest CEDC of 1.03 x 10-2 m2/s (Db) assumed to occur in
the Brushy Plain-Cherry Hill network is applied, the response is almost the same
and even if the coefficient is neglected, the FCR concentration doesn’t change.

Published by The Berkeley Electronic Press, 2007 9


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

For the DTSM step input, the response is also the same as concluded in Figure
4.a. These could be true because under the condition of Da and Db, the Peclet
numbers, Pes are 237,066.6 and 414.3 respectively, which are significantly higher
than the value where the diffusion can be considered.
Using the formula of Peclet number, the CEDC of 0.002576 km2/hr (Dcr)
is obtained as the critical CEDC for this simulation based on the minimum Peclet
number of 6 proposed by US EPA (2002). The rectangular pulse response in this
condition is shown in Figure 3.c. As shown by the responses in Figure 3.c, the
diffusion of the CEDC of 0.002576 km2/hr affects the dispersion of the FCR
along the pipe over elapsed time, where it creates different responses and results
in the longer distance of the FCR existence and the lower FCR concentration at
the responses’ peak.

The effect of water flow rate velocity on the FCR


It is clear that, for the DTSM rectangular pulse input, the dispersion of the FCR in
single pipes of the DWDS is dominated by the axial convection, as can be seen in
Figure 4. This is reasonable because the water flow carries FCR distribution along
the pipe as a result of its straight movement along the pipe.

(a)

http://www.bepress.com/cppm/vol2/iss2/3 10
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

(b)

(c)

Published by The Berkeley Electronic Press, 2007 11


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

(d)
Figure 4. Comparison of the water flow rate effect against the CEDC effect on the
DTSM for step and rectangular pulse inputs

For example, for 12 hours rectangular pulse input, small changes in the velocity
from v = 0.0504 km/hr to v = 0.3024 km/hr result in a significant change in the
distance reached and a considerable change in the FCR concentration along the
pipe with the same k of 0.02304/hr and the same CEDC of 3.708 x 10-5 km2/hr.
For the same elapsed time, the maximum FCR concentrations for the velocity of
0.3024 km/hr are about 0.836, 0.718 and 0.533 mg/L at distances of about 2.3,
4.25 and 8.12 km, respectively (see Figure 4.c), and these are about 0.777, 0.649
and 0.452 mg/L at the distances of 0.5, 0.851 and 1.48 km, respectively (see
Figure 4.b).
For the DTSM step input, unlike the change of the CEDC shown in Figure
4.a, as can be seen in Figure 4.d, the change of water flow velocity from 0.014
m/s (0.0504 km/hr) to 0.084 m/s (0.3024 km/hr) can further move the point of the
required minimum FCR concentration of 0.2 mg/L from 0.71 km, 1.32 km and
2.42 km to 3.9 km, 7.42 km and 14.45 km for the step input of 12 hours, 24 hours
and 48 hours, respectively. The increase in the distance is in proportion to the
increase of the water flow velocity. Moreover, the higher water flow velocity
causes the higher FCR concentration at any point along the pipe. The change of
water flow velocity from 0.014 m/s to 0.084 m/s can raise the maximum FCR
concentration at 1 km from the booster from approximately 0.635 mg/L to

http://www.bepress.com/cppm/vol2/iss2/3 12
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

approximately 0.927 mg/L for 30, 36, 42 and 48 hours step input. At 2 km from
the booster, the maximum free chlorine residual increases from approximately
0.011, 0.32 and 0.39 mg/L to approximately 0.859 mg/L for 30, 42 and 48 hours
step input respectively.

MODEL VALIDATION
In order to validate the DTSM for the FCR, the data and the results from a
campaign on 11 July 1995 at the Cholet municipality DWDS (Heraud et al., 1997)
are taken into account, where most of the Cholet system pipes are cement-lined
case iron with k being 4.77 x 10-3 min-1 (0.2862 hr-1) for the pipe diameter of 100
mm and a low water flow of 5 m3/hr, and Co (at the plant outlet) of 0.97 mg/L.
Figure 5.a, the top curve is the DTSM of the FCR using the value of k
being 0.282/hr and v being 0.637 km/hr obtained from the calculation using the
pipe diameter and the flow in the Cholet System campaign. As can be seen from
Figure 5.a, the analytical solution of the DTSM (DTSMAS) matches extremely
well with the FCR measured data in the campaign. The DTSM in time domain is
the same as the model fitted by Heraud et al. (1997) with k being 4.7 x 10-3 min-1
(0.282 hr-1).

(a)

Published by The Berkeley Electronic Press, 2007 13


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

(b)
Figure 5. Comparison between the measured FCR, the fitted model and the DTSM

In fact, the distance for every single point of the measured FCR of the Cholet
System can be calculated using the formula s = v x t. The FCR concentration
based on the fitted model at certain distance for the random residence time of 1,
17, 3, 4, 5, 5.6, 8 and 10 hours is about 0.7316, 0.6006, 0.4163, 0.314, 0.2368,
0.2, 0.1016 and 0.0578 mg/L respectively. Meanwhile, the FCR concentration
based on the step input DTSM is shown in Figure 5.b. Hence, the DTSM in space
domain could be also validated using the data of the distance versus the FCR by
which the correlation coefficient R2 is 0.9999.

ADAPTIVE CHLORINE DOSING DESIGN


The significant time variation in water demands leading to time-varying chlorine
transport delays (which could be more than an hour) is one of the key difficulties
in controlling the FCR within the safe drinking water limit level of 0.2-0.6 mg/L
in the DWDS.
With regard to the low and high demand of drinking water and the
fluctuating volumetric water supply (VWS) which leaves the water plant over
time, this research proposes an Adaptive Chlorine Dosing Design (ACDD)
scheme, wherein the chlorine booster location logistic and the injection rates of
each chlorine booster can be obtained using Equations (14), (15) and (16), which
are based on the analytical solution of Equation (13). This is a model-based feed
forward approach in controlling the FCR concentration, where the correct

http://www.bepress.com/cppm/vol2/iss2/3 14
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

chlorine concentration injected by each booster is predicted by using the VWS


velocity (Vvws) entering the first booster.

CC1
CB1 = Cin (14)
exp ( kx1 / Vvws )

kxB12
CB 2 = CHL CB1 exp (15)
Vvws

0.5[ xB12 + 2( x i xi 1 xC1B 2 )] exp( xi )


CBi = CHL 1 exp k (16)
Vvws

Where CB1 is the initial chlorine concentration injected by the 1st booster; CC1 is
the FCR concentration at the first water consumer that should be equal to the high
limit required (CHL); Cin is the FCR concentration just before passing the 1st
booster; CBi is the additional chlorine concentration injected by the ith booster
(i=3,4, …); xB12 is the distance between the 1st and 2nd booster; xC1B2 is the
distance between the first consumer and the 2nd booster and x-i-1 and x-i is the
distance between the i-1 and i boosters, respectively.

(a)

Published by The Berkeley Electronic Press, 2007 15


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

(b)
Figure 6. The FCR concentrations for different volumetric water supply velocities after
the ACDD package of chlorine dosage applied

In this simulation, the first drinking water consumers are assumed at 0.83 km
from the 1st booster, the main pipe length is 8.5 km and the ACDD deals with the
required limit FCR of 0.2-0.6 mg/L. The value k, taken from Rosman (1994), is
related to the pipe diameter size of 12 inch and the pipe wall decay rate constant
of 0.45 m/day, which is also assumed to take place in the SDWDS. The maximum
VWS velocity is assumed to be 2.457 km/hr just leaving the first chlorine booster,
which is 25% higher than the maximum water flow rate of 0.546 m/s in the main
pipe of water distribution network of New Haven (Biswas et al., 1993).
As concluded in Figure 6.a which is the DTSM response for the 10 hours
step input, the proper initial chlorine concentration (CB1) of 0.4587 mg/L injected
by the 1st booster can be obtained using Equation (14) so that the first consumers
get the drinking water with the required maximum FCR concentration of 0.6
mg/L. By using Equations (15) and (16), the proper chlorine concentration of
0.0639, 0.1151, 0.1162 and 0.1163 mg/L (CB2, CB3, CB4 and CB5) is injected by the
2nd, 3rd, 4th and 5th boosters, respectively. This means that the ACDD is able to
carry the FCR concentration up to the high limit of 0.6 mg/L along the pipe with
the VWS velocity (Vvws) of 2.457 km/hr, as can bee seen by the top curve in
Figure 6.a. This ACDD package of chlorine dosage is also able to deal with the

http://www.bepress.com/cppm/vol2/iss2/3 16
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

total decrease of 0.5 times the Vvws and the lowest FCR concentration, which is
still approximately 0.19 mg/L higher than the lower limit of 0.2 mg/L, as shown
by the dashed curve in Figure 6.a.
In addition, the ACDD can also deal with the lower Vvws, 0.85 km/hr,
which is shown by the line curve in Figure 6.b. The simulation on the Vvws of
0.546 m/s, which is also normally found in the New Haven network, validates the
ACDD ability to lift the FCR concentration up to the high limit by injecting
chlorine with concentrations of 0.4419, 0.1355, 0.2304, 0.2321 and 0.2323 mg/L
at the 1st, 2nd, 3rd, 4th and 5th boosters, respectively, as can be seen by the dashed
curve in Figure 6.b. When we further lower the Vvws to 0.274 km/hr, the ACDD
scheme can still maintain the FCR within the limits, as shown by the dashed curve
in Figure 6.b.

CONCLUSION
Discrete Time-Space Models (DTSMs), which are developed using Finite
Difference Method and computed using Matlab 7.0.1, can be used to predict the
FCR concentration along the single pipes of the DWDS over time. In the DTSM
simulation, using step input and rectangular step input, the results have shown that
the dispersion of the FCR in the pipes could be dominated by axial convection
using the strong influence of water flow rate velocity on the FCR, but the FCR
axial diffusion in single pipes in DWDS should be considered, when the chlorine
effective diffusivity coefficient is exceedingly high in the DWDS. It was also
shown that the DTSM is valid to present FCR distribution in the DWDS from the
prospective of the FCR versus the elapsed time and space length. These modeling
studies resulted in an Adaptive Chlorine Dosing Design (ACDD) scheme in
which chlorine booster location logistics and proper injections of chlorine, based
on volumetric water supply velocity just before entering the 1st chlorine booster,
can be predicted in order to maintain the FCR concentration within the safe
drinking water limits.

REFERENCES
American Water Works Association Water (AWWA), Treatment - Principles and
Practices of Water Supply Operation, 2003, AWWA, 3rd ed., Denver, Colorado,
USA.

Biswas, P., Lu, C. and Clark, R. M., A Model for Chlorine Concentration Decay
in Pipes, Wat. Res., 1993, 27(12), 1715-1724.

Published by The Berkeley Electronic Press, 2007 17


Chemical Product and Process Modeling, Vol. 2 [2007], Iss. 2, Art. 3

Connell, G. F., The Chlorination/ Chloramination Handbook, 1996, AWWA,


Colorado, USA.

Feben, D. and Taras, M. J., Studies on Chlorine Demand Constant, Journal of


American Water Works Association, 1951, 43(11), 922-932.

Heraud, J., Kiene, L., Detay, M. and Levi, L., Optimized Modeling of Chlorine
Residual in a Drinking Water Distribution System with a Combination of on-line
Sensors, Journal of Water SRT-Aqua, 1997, 46(2), 59-70.

Johnson, J. D., Measurement and Persistence of Chlorine Residual in Natural


Waters, in Jolle W R. L. (Ed), Water chlorination: Environmental impact and
health effects, 1987, Ann Arbor Science, Ann Arbor, Michigan, 37-63.

Maul, A., El-Shaarawi, A. H. and Block, J. C., Heterotrophic Bacteria in Drinking


water distribution System: I. Spatial and Temporal Distribution, Sci. Total Envir.,
1985, 44, 201-214.

O’Neil, P. V., Advanced Engineering Mathematics, 1987, Wadsworth Publishing


Company, Belmont, California.

Rosman, L. A., Clark, R. M. and Grayman, W. M., Modelling Chlorine Residual


in Drinking Water Distribution System, Journal of Environmental Engineering,
1994, 120(4), 803-820.

Tveito, A. and Winther, R., Introduction to Partial Differential Equations: A


Computational Approach, 1998, Springer-Verlag Inc., New York.

Tzatchkov, V. G., Aldama, A. A., and Arreguin, F. I., A Domain Decomposition


Strategy for the Numerical Simulation of Contaminant Transport in Pipe
Networks, Fourteenth International Conference on Domain Decomposition
Methods, 2003, Ismael Herrera, David e. Keyes, Olof B. Widlund and Robert
Yates (Eds.).

U.S. Environment Protection Agency (EPA), The QTRACER2 Program for


Tracer-Breakthrough Curve Analysis for Tracer Tests in Karstic Aquifers and
Other Hydrologic Systems, 2002, EPA/600/R-02/001, EPA Office of Research
Development, Washington, D.C. 20460, 43.

http://www.bepress.com/cppm/vol2/iss2/3 18
DOI: 10.2202/1934-2659.1067
Muslim et al.: Simulation of Free Chlorine Decay and Adaptive Chlorine Dosing

Vasconcelos, J. J., Rossman, L. A., Grayman, W. M., Boulos, P. F. and Clark, R.


M.., Kinetic of Chlorine Decay, Journal of American Water Works Association,
1997, 89(7), 5-65.

Wei, I.W.T. and Morries, J.C., Dynamic of Break-point Chlorination, in Rubin,


A.J. (Ed), Chemistry of Water Supply, Treatment and Distribution, 1974, Ann
Arbor Science Publisher, Inc., 297-332.

World Health Organization, Water and Sanitation: Facts and Figures, 2004,
Retrieved: April 27, 2005, from <http://www.who.int/water_sanitation_ health/
publications/facts2004/en/>.

Published by The Berkeley Electronic Press, 2007 19

Das könnte Ihnen auch gefallen