Sei sulla pagina 1di 15

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Available online at www.sciencedirect.com

ScienceDirect
journal homepage: www.elsevier.com/locate/watres

Three-dimensional three-phase model for


simulation of hydrodynamics, oxygen mass
transfer, carbon oxidation, nitrification and
denitrification in an oxidation ditch
Li Lei a,b, Jinren Ni a,b,*
a

Department of Environmental Engineering, Peking University, Beijing 100871, China


The Key Laboratory of Water and Sediment Sciences, Ministry of Education, Peking University, Beijing 100871,
China

article info

abstract

Article history:

A three-dimensional three-phase fluid model, supplemented by laboratory data, was

Received 7 June 2013

developed to simulate the hydrodynamics, oxygen mass transfer, carbon oxidation, nitri-

Received in revised form

fication and denitrification processes in an oxidation ditch. The model provided detailed

6 October 2013

phase information on the liquid flow field, gas hold-up distribution and sludge sedimen-

Accepted 15 January 2014

tation. The three-phase model described water-gas, water-sludge and gasesludge in-

Available online 23 January 2014

teractions. Activated sludge was taken to be in a pseudo-solid phase, comprising an


initially separated solid phase that was transported and later underwent biological re-

Keywords:

actions with the surrounding liquidmedia. Floc parameters were modified to improve the

Three-dimensional three-phase

sludge viscosity, sludge density, oxygen mass transfer rate, and carbon substrate uptake

model

due to adsorption onto the activated sludge. The validation test results were in very

Pseudo-solid phase

satisfactory agreement with laboratory data on the behavior of activated sludge in an

Sedimentation

oxidation ditch. By coupling species transport and biological process models, reasonable

Mass transfer

predictions are made of: (1) the biochemical kinetics of dissolved oxygen, chemical oxygen

Biochemical kinetics

demand (COD) and nitrogen variation, and (2) the physical kinematics of sludge

Oxidation ditch

sedimentation.
2014 Elsevier Ltd. All rights reserved.

1.

Introduction

Oxidation ditches (ODs) are widely used in wastewater treatment due to their simple construction, low capital and
maintenance costs, high and flexible capacity, and low sludge
production (Hong et al., 2003). More than 10,000 oxidation
ditches are to be found in China and the USA alone. However,

oxidation ditches occupy large areas of land, consume substantial energy, and produce uneven deposits of sludge (Yang
et al., 2011). Much work is presently being undertaken to
mand and Carlsson,
optimize the treatment process (see e.g. A
2012), to improve sludge deposition and reduce energy consumption (see e.g. Zhou et al., 2012).
Mathematical models offer an effective means of simulating the physical, chemical and biological processes in ODs

* Corresponding author. Department of Environmental Engineering, Peking University, Beijing 100871, China. Tel.: 86 10 62751185; fax:
86 10 62756526.
E-mail address: nijinren@iee.pku.edu.cn (J. Ni).
0043-1354/$ e see front matter 2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.watres.2014.01.021

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Nomenclature
aj,i
bA
bH
Cs
dG
DL
dO
!
F lift; q
fP
!
Fq
!
F vm; q
!
g
hKs
iXB
iXP
!
J q; i
ka
kh
KLaL
KNH
KNO
KO,A
KO,H
KS
KX
mpq
O2
OF
p
Q
!
R pq
Rq, i

S
SI
SND
SNH
SNO
SO

stoichiometric number of species i in the j-th


process, dimensionless
decay coefficient for XB,A, d1
decay coefficient for XB,H, d1
wall roughness constant, dimensionless
bubble diameter, m
diffusivity of oxygen in the liquid phase, m2 s1
diameter of aeration orifice, m
lift force, kg m2 s2
fraction of biomass leading to particulate
products, dimensionless
external body force, kg m2 s2
virtual mass force, kg m2 s2
gravitational acceleration, m s2
wall roughness height, m
mass of nitrogen per mass of COD in biomass,
g g1
mass of nitrogen per mass of COD in products
from biomass, g g1
diffusion flux of species i in phase q, kg m2 s1
ammonification rate, m3 g1 d1
maximum specific hydrolysis rate, d1
mass transfer coefficient, s1
ammonia half-saturation coefficient for XB,A,
g m3
nitrate half-saturation coefficient for XB,H, g m3
oxygen half-saturation coefficient for XB,A, g m3
oxygen half-saturation coefficient for XB,H, g m3
half-saturation coefficient for XB,H, g m3
half-saturation coefficient for hydrolysis of slowly
biodegradable substrate, g g1
mass transfer from phase p to q, kg m3 s1
oxygen concentration in air, g m3
normalized standard error, dimensionless
pressure, N m2
flow rate, m3 s1
interaction force between phase p and q,
kg m2 s2
source term representing mass transfer of species
i from other phases to phase q, and the
production/consumption rate of the species i for
biochemical reactions, kg m3 s1
soluble constituent concentration, g m3
soluble inert pollution concentration, g m3
soluble organic nitrogen concentration, g m3
ammonium concentration, g m3
nitrate and nitrite concentration, g m3
oxygen concentration in liquid, g m3

which experience complicated alternating aerobic and anoxic


conditions for nitrification and denitrification. For example,
the Activated Sludge Model (ASM) predicts the effluent water
quality and biomass production of wastewater treatment
plants (Henze et al., 2000). In ODs, the wastewater treatment
efficiency is influenced not only by the bio-reaction of activated sludge, but also by the dynamics of liquid-bubble flows

SO(S)
Sq
SRT
SS
Uslip
!
v pq
!
v
q

X
XB,A
XB,H
xci
XI
xmi
XND
XP
XS
YA
YH
Yq, i

201

saturated dissolved oxygen, g m3


source term of phase q, kg m3 s1
sludge age, d
soluble biodegradable pollution concentration,
g m3
slip velocity between a gas bubble and water,
m s1
interphase velocity from phase p to q phase, m s1
velocity vector of phase q, m s1
particulate component concentration, g m3
autotrophic biomass, g m3
heterotrophic biomass, g m3
calculated result of the i-th parameter
particulate inert pollution concentration, g m3
measured result of the i-th parameter
particulate organic nitrogen concentration, g m3
inert biomass, g m3
particulate biodegradable pollution
concentration, g m3
yield for XB,A, g g1
yield for XB,H, g g1
mass fraction of species i in phase q,
dimensionless

Greek letters
a
modification coefficient for SO(S), dimensionless
modification coefficient for dG, dimensionless
adG
volume fraction of phase q, dimensionless
aq
b
modification coefficient for KLaL, dimensionless
g
modification coefficient for KLaL, dimensionless
3
dissipation rate of turbulent kinetic energy, m2 s3
hg
correction factor for mH under anoxic conditions,
dimensionless
correction factor for hydrolysis under anoxic
hh
conditions, dimensionless
maximum specific growth rate for XB,A, d1
mA
mH
maximum specific growth rate for XB,H, d1
mq
shear viscosity of phase q, kg m1 s1
rj
process rate of the j-th bioreaction, kg m3 s1
rq
density of phase q, kg m3
s
surface tension, kg s2
sq
stress-strain tensor of phase q, kg m1 s2
Subscripts
G
gas phase
in
inflow
L
liquid phase
out
outflow
rec
recirculation flow
S
pseudo-solid phase

(Insel et al., 2005). Carbon oxidation process was firstly


coupled in one-dimensional (1D) convectionedispersion
equation by Stamou (1994), followed by coupling more processes such as nitrification and denitrification (Stamou, 1997;
Stamou et al., 1999; Mantziaras et al., 2011) to reveal the effects of local hydrodynamics on water quality in ODs. For
more detailed understanding of the effects of hydrodynamics,

202

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

two-dimensional (2D) or three-dimensional (3D) models


would be preferred (Yang et al., 2011). Littleton et al. (2007)
introduced the Activated Sludge Model No. 2 (ASM2) to a 3D
fluid dynamics model for elucidating the role of the bioreactor
macro-environment in simultaneous biological nutrient
removal. Other investigators also attempted to describe the
complex phenomena in OD (Zhang et al., 2010). On the other
hand, effects of hydrodynamics would be reflected by matters
in different phases. However, most previous models have
been limited to one or two phases, although multiple-phase

phenomena and processes have been observed (e.g. Pipes,


1969; Schmid et al., 2003; Fayolle et al., 2007) in ODs.
Understanding of coupled physicalechemicalebiological
processes relies on accurate assessment of the transport
processes and phase interactions. There are three basic phases of the primary medias needed to be fully considered in
ODs. First, oxygen must be supplied in order to maintain the
level of dissolved oxygen (DO) during the aerobic process and
so directly affects the effluent water quality (Fayolle et al.,
2007), which implies the model that does not take gas phase

Fig. 1 e Framework of the three-dimensional three-phase model for an oxidation ditch (AGH denotes aerobic growth of
heterotrophs; AGA denotes aerobic growth of autotrophs; OMT denotes oxygen mass transfer; ANGH denotes anoxic growth
of heterotrophs; HEO denotes hydrolysis of entrapped organics; ASON denotes ammonification of soluble organic nitrogen;
HEON denotes hydrolysis of entrapped organics nitrogen; DH denotes decay of heterotrophs; DA denotes decay of
autotrophs).

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

into consideration cannot reasonably simulate the oxygen


mass transfer between gas and sewage water. Second, activated sludge comprises a strongly hydrated solid phase, and
has different physical properties to those of pure water
(Schmid et al., 2003). Furthermore, sludge settling can lead to
septic sludge formation at the dead angle in OD, with associated odor (Pipes, 1969). The complicated phase interactions
and transformations in an OD system include transfer of
dissolved oxygen from the gas phase, and carbon oxidation,
nitrification and denitrification in the liquid and solid phases.
In this paper, a 3D three-phase model was developed by
taking the sewage water, air bubbles and activated sludge to
be in liquid, gas and pseudo-solid phases, respectively. The
proposed model could not only vividly describe local hydrodynamic structures with 3D fluid velocities but also reasonably simulate the interactions of sewage water, air bubbles
and activated sludge treated as pseudo-solid phase. Fig. 1
shows the proposed framework in which the relationship
between the phases is outlined. Coupled gas transport and
modified oxygen mass transfer models simulate the DO distribution. Coupled species-transport and modified biological
kinetic models simulate the sludge distribution and pollutant
degradation, and include the biochemical kinetics of chemical
oxygen demand (COD) and nitrogen removal in an OD. More
specifically, the 3D and three-phase model is able to simultaneously describe the 3D transport of sewage water (e.g.
secondary flow, see Yang et al., 2011), bubbles and activated
sludge (settling process), in addition to the phase interactions
and the sedimentation processes related to the activated
sludge. During calibration, the input parameters are determined by iteration for target values of sludge viscosity,
settling capacity, oxygen mass transfer rate, and carbon substrate uptake due to adsorption onto the pseudo-solid phase.
The sludge transport process and the effect of inclusion of the
pseudo-solid phase on mass transfer and pollutant transformation are investigated through simulations of variations
in concentration of activated sludge, DO, COD, ammonia nitrogen and nitrate in a pilot-scale OD.

liquid, and is termed a pseudo-phase. To approximate the


pseudo-phase behavior in the model, the activated sludge is
first represented as in a separated solid phase regarding
transport and sedimentation processes, and later as in a solideliquid phase after biological changes have taken place (see
Fig. 2). The 3D three-phase model quantifies the phasedependent behavior of the sewage, activated sludge, and
gas, and their complicated interactions in an OD (Fig. 1). To
permit analogy between the behavior of activated sludge and
that of granular particles, the model parameters require
adjustment to account for differences between floc sludgeand
granular sludge. Floc sludge has much higher water content,
contains extracellular polymeric substances, and has a
negatively-charged surface, whereas granular sludge is more
permeable (Wang et al., 2012). To describe the transport and
evolution of the pollutants and biomass, an advectiondiffusion species transport model is coupled with the modified oxygen mass transfer model and modified biological kinetic process models. The governing equations are based on
mass and momentum conservation laws. The model assumptions are as follows:
(1) sewage water, activated sludge, and air are in liquid,
pseudo-solid, and gas phase, respectively;
(2) pollutants are divided into soluble and particulate
components regarded as species of liquid and pseudosolid phases, respectively;
(3) heterotrophic and autotrophic biomass are species in
the pseudo-solid phase;
(4) DO is a species in the liquid phase;
(5) oxygen mass transfer is a biological rather than a
physical process;
(6) accumulation of biomass with ammonia may be
neglected;
(7) alkalinity is not a limiting parameter;
(8) no biological reaction occurs in the secondary settling
tank.

2.1.1.

2.

Methodology

2.1.

Model development

A hydrated activated sludge floc is essentially a gelatinous,


coagulated material whose phase lies between solid and

203

Multiphase hydrodynamics model

A standard 3D steady-state multi-phase flow model, described


in detail by Fluent Corporation (2006), is used to describe the
complicated hydrodynamic behavior of the sewage water,
activated sludge, and gas transportation behavior in an OD.
The steady-state equilibrium mass conservation equation is
given by:

Fig. 2 e Dual roles of the pseudo-solid phase.

204

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Table 1 e Quantifications of the major chemical and biological processes.


No.

Process

Bioreaction

SO
S
r1 mH KSSS
XB; H
S KO; H SO

Oxygen mass transfer


Anoxic growth of heterotrophs

O2/SO

Hydrolysis of particulate organics


Ammonification
Hydrolysis of particulate organics nitrogen
Decay of heterotrophs
Decay of autotrophs

XS/SS
SND/SNH
XND/SND
XB, H/(1fP)XSfPXP(iXBfPi)XPXND
XB, A/(1fP)XSfPXP(iXBfPi)XPXND

r3(KLaL)wastewater(aSO(S)SO)
KO; H
SNO
S
r4 mH KSSS
h XB; H
S KO; H S
h O KNO SNO g
i
K H
XS =XB; H
SO
SNO
XB; H
r5 kh KX XS =XB; H KO; H SO hh KO; O;H S
O KNO SNO
r6kaSNDXB, H
r7 r5 XND =XS
r8 bHXB,H
r9 bAXB,A

Aerobic growth of heterotrophs

Aerobic growth of autotrophs

3
4
5
6
7
8
9

1
YH

1
YH

(1)

p1

where aq is the volume fraction, rq is density, !


v q is velocity
vector, Sq is the source term, p and q are the different phases,
and mpq characterizes the mass transfer from phase p to q.
Subscripts q W, G and S denote water, gas and active sludge,
respectively. The steady-state equilibrium momentum equation is:
n


X
!
v q!
v q aq Vp V$sq aq rq !
g
v pq
R pq mpq !
V$ aq rq !
p1



 ! !
!
Fq F lift; q F vm; q
 mqp !
v qp

(2)

where p is the pressure shared by all phases, !


g is gravitational
!
v pq
acceleration, R pq is the interaction force between phases, !
is the inter-phase velocity from phase p to q, mpq !
v pq  mqp !
v qp
denotes the momentum change due to mass transfer between
!
the p-th and the q-th phases, F q is an external body force,
!
!
F lift;q is a lift force, F vm;q is a virtual mass force and sq is the
stress-strain tensor for the q-th phase. The multi-phase flow
model is closed using a steady-state k-3 turbulence model (see
e.g. Fluent Corporation, 2006).

Species transport model

The following 3D steady-state advection-dispersion species


transport equation with source term is used to describe the
transport and growth/decay of activated biomass and contaminants in the ditch:


 ! 
v q Yq; i V$ aq J q; i Rq; i
V$ aq rq !

(3)

where Yq,i is the mass fraction of species i in phase q, Rq, i is a


source term representing mass transfer of species i from other
phases to phase q, and the production/consumption rate of
!
the species i for biochemical reactions, J q; i is the diffusion
flux of species i in phase q.

2.1.3.

SO
NH
r2 mA KNHSS
XB; A
NH KO; A SO

1YH
SS 2:86Y
SNO iXB SNH /XB; H
H

3
 X



V$ aq rq !
vq
mpq  mqp Sq

2.1.2.

Process rate

H
SS 1Y
YH SO iXB SNH /XB; H


4:57YA
SO iXB Y1A SNH /XB; A Y1A SNO
YA

Biochemical kinetics model

The source term Rq, i in Equation (3) is given by the modified


Activated Sludge Model No.1 (ASM1) (Henze et al., 2000).
Biochemical processes, such as carbon oxidation, nitrification
and denitrification, are described by considering thirteen
components (Fig. 1). Autotrophic, heterotrophic and inert
biomass is regarded as particulate matter. Fig. 1 and Table 1
summarize the detailed biological interactions among the

three phases and corresponding kinetic processes. The source


term Rq, i is expressed by
Rq; i

9 
X

 aj; i rj

(4)

j1

where j is the number of the process in Table 1, aj,i is the


stoichiometric number of species i in the j-th process, and rj is
the process rate. In Equation (4), if species i is produced,  is
taken positive; whereas if species i is consumed,  is taken
negative.
Carbon substrate removal from storage always occurs in an
activated sludge system, because activated sludge is in the
separated phase (Carucci et al., 2001). COD then accumulates
in sludge (Beccari et al., 2002), leading to the heterotrophs
 gn et al., 2011).
having a competitive growth advantage (Cg
Hence, certain stoichiometric and kinetic parameters (such as
mH, YH, KS and bH) are modified herein, given that activated
sludge is treated as in the separated pseudo-solid phase. Table
2 lists the proposed parameters used by ASM1 and the modified input parameters used in the present work.

2.1.4.

Oxygen mass transfer model

The oxygen mass transfer rate r3 (see Table 1) from the gas to
the liquid phase is determined by Kulkarni (2007)
r3 KL aL


wastewater

aSOS  SO

(5)

in which KLaL wasterwater is the mass transfer coefficient, SO(S) is


saturated DO concentration in clean water, SO is oxygen concentration in the liquid phase, and a expresses the proportionality between saturated DO in wastewater and its clean
water value. For bottom aeration, the oxygen mass transfer
coefficient for sewage water is expressed as:
KL aL

wastewater

gb

12aG
dG

s
DL Uslip
pdG

(6)

where g is introduced to consider the pseudo solid-effect, b is


the ratio of wastewater to clean water mass transfer coefficients, aG is the volume fraction occupied by the gas phase,
dG is the Sauter mean diameter of the bubbles, DL is diffusivity
of oxygen in the liquid phase, and Uslip is the slip velocity
between a gas bubble and water, and can be estimated by
evaluating jvL-vGj. For surface aeration, the oxygen mass
transfer correlates directly with the dissipation rate of energy
(Kumar and Rao, 2009) such that

205

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Table 2 e Summary of parameters used in modeling of biological processes.


Parameter
Stoichiometric

Kinetic

YA
YH
fP
iXB
iXP
mH
KS
KO,H
KNO
bH
hg
hh
kh
KX
mA
KNH
KO,A
ka
bA

Unit

ASM1

Su and Yu (2006)

Present

g(COD) g(COD)1
g(COD) g(COD)1
Dimensionless
g(N) g(COD)1
g(N) g(COD)1
d1
g(COD) m3
g(O2) m3
g(N) m3
d1
Dimensionless
Dimensionless
d1
g(COD) g(COD)1
d1
g(N) m3
g(O2) m3
m3 g(COD)1 d1
d1

0.24
0.67
0.08
0.086
0.06
6.00
20.0
0.20
0.50
0.62
0.8
0.40
3.0
0.03
0.80
1.0
0.4
0.08
0.15

0.24
0.58
0.08
0.086
0.06
4.98
26.1
0.20
0.50
0.92
0.8
0.44
3.0
0.03
0.80
1.0
0.4
0.08
0.15

0.24
0.63e0.67
0.08
0.086
0.06
5.50e6.00
20.0e23.0
0.20
0.50
0.62e0.77
0.8
0.44
3.0
0.03
0.80
1.0
0.4
0.08
0.15

i
h


KL aL
p 0:64 exp 0:293 0:98 1:6  106 exp 0:723  3:912:0
3

effect of the latter incorporated via the source term in Equation (3).

(7)
where 3 is the dissipation rate of the turbulent kinetic energy.
Hence, the oxygen mass transfers due to the surface impellers
and bottom aeration system can be evaluated by using Equations (5)e(7).

2.2.

Phase interaction

2.2.1.

Liquidegas interaction

Air bubbles introduced into OD experience a drag force due to


the different velocities in the liquid and gas phases. The
!
equations governing the interaction force R pq are listed by
Fluent Corporation (2006). Liquid-gas interaction directly influences both physical and mass transfer processes, with the

2.2.2.

Liquid-solid interaction

The density of activated sludge density ranges from 1010 to


1060 kg/m3 (Dammel and Schroeder, 1991), and so sludge
settling always occurs. Flow turbulence influences sludge
settling, resulting in intensive interaction between sewage
!
water and activated sludge floc. The interaction force R pq
between pseudo-solid and liquid phases is also described
using interaction equations listed by Fluent Corporation
(2006).
In activated sludge systems, intensive interaction occurs
between different species in the sewage and activated sludge.
To sustain growth, heterotrophs and autotrophs biomass
utilize ammonium and nitrate species in sewage (Table 1).

Fig. 3 e Plan view sketch of the pilot-scale oxidation ditch (a) and monitoring points of liquid velocity (b) (Unit: mm).

206

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Table 3 e Operation modes of impellers and stirrers.


Moving
part
Impeller 1
Impeller 2
Stirrer 1
Stirrer 2
Stirrer 3
Stirrer 4
a
b

Case I

Case II

Speed (rpm) Direction Speed (rpm) Direction


a

b

80
80
90
90
90
70

40
40
70
70
70
50

rotation in clockwise direction.


- rotation in anticlockwise direction.

Meanwhile, heterotrophs and autotrophs decompose to particulate biodegradable organic nitrogen and slowly biodegradable substrate, which are further hydrolyzed respectively
to soluble biodegradable organic nitrogen and readily biodegradable substrate. The present model describes the interactions between species in liquid and pseudo-solid phases,
unlike traditional models that neglect the pseudo-solid
phase.

2.2.3.

Gasesolid interaction

Many models, such as ASM-series models, consider activated


sludge to be perfectly soluble in the liquid phase. Given that an
activated sludge floc has properties different from pure water,
it is better to treat the floc as in a separated pseudo-solid
phase, with oxygen mass transfer between gas and liquid
phases modified accordingly in terms of solid concentration
(Mena et al., 2011). For this purpose, the parameter g is
introduced in Equation (6).

2.3.

Boundary conditions

For the OD (Fig. 3), open boundary conditions are applied at


the influent inlet and effluent outlet. The walls are treated as
fixed, impermeable boundaries. At the inlet, the velocity is
prescribed (according to the inflow rate), and the inlet sludge
concentration (Xin, the sum of the concentrations of all

particulate components Xi in Table 1) is initially also prescribed. At steady state, when the internal condition of the OD
is balanced after sludge return, Xin is assumed to be the same
as the sludge concentration at the outlet (Xout) provided the
average sludge concentration (Xaverage) in OD is less than a
pre-set value (Xset). For Xaverage > Xset, the surplus sludge
(Xsurplus) is removed from the OD by pumping in order to
maintain the sludge concentration at a desired level. Full details about the surplus sludge pumping model and the inlet
sludge concentration condition are given by Stamou (1997).
The recirculation flow recycles water to the inlet. If Sin and Srec
denote the soluble constituents (e.g. ammonium and nitrate)
of wastewater and recirculation flow, the inlet value of the
soluble constituent variable is:
S Qin Sin Qrec Srec =Qin Qrec

(8)

where Qout is the outflow rate, Qin is the inflow rate, and Qrec is
the recirculation flow rate. The wastewater constituents are
determined from experiments, following Henze et al. (2000).
At the OD outlet, the boundary pressure is atmospheric.
For bottom aeration, the introduction of gas can be treated
as source term in the Equations (1)e(3), and the oxygen concentration in the pumped air is evaluated using an equation
provided by Fayolle et al. (2007). Surface aeration refers to
rotation of the impellers which aerated the water in their vicinity. The oxygen mass transfer rate due to surface aeration
is obtained by solving Equations (5) and (7).
A rigid-lid, slip wall boundary condition (see e.g. Yang
et al., 2011) is applied to the liquid and pseudo-solid phases
at the water surface. Injected air from bottom aeration escapes the OD at the gaseliquid surface, and so a degasification
condition (Le Moullec et al., 2011) is applied to the gas phase at
the water surface.
No-slip boundary conditions are assigned for all other
walls, including the bottom surface, the side and central walls
of the ditch. The roughness constant and the roughness
height at the no-slip boundaries are calibrated to the
measured data, using the Fluent values of 1 and 0.02 m,
respectively following Yang et al. (2011).

Table 4 e Experimental conditions and range of primary variables for model calibration and verification.
Variables

Operation conditions

Range of variables
3

Rotation mode

Aeration rate (m /h)

Model Calibration

Liquid velocity

Case I

2.2

Model Verification

MLSS
DO
COD
Ammonium
Nitrate
Liquid velocity

Case
Case
Case
Case
Case
Case

I
I
I
I
I
II

2.2
2.2
2.2
2.2
2.2
2.2

MLSS
DO
COD
Ammonium
Nitrate

Case
Case
Case
Case
Case

II
II
II
II
II

2.2
2.2
2.2
2.2
2.2

ux: 0.000e0.155 m/s


uy: 0.000e0.044 m/s
uz: 0.000e0.116 m/s
3.80e4.16 g/L
0.32e1.66 mg/L
12.8e14.0 mg/L
0.33e0.90 mg/L
16.6e17.5 mg/L
ux: 0.000e0.132 m/s
uy: 0.000e0.031 m/s
uz: 0.000e0.101 m/s
3.42e4.55 g/L
0.28e1.07 mg/L
14.7e15.8 mg/L
3.04e3.91 mg/L
11.78e14.01 mg/L

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Table 5 e Description of the parameters used in the


model.
Parameter
Sewage water density
Sewage water viscosity
Dissolved oxygen of
synthetic wastewater
Soluble inert organic
substrate of synthetic
wastewater
Readily biodegradable
substrate
synthetic wastewater
Particulate inert organic of
synthetic wastewater
Slowly biodegradable
substrate of synthetic
wastewater
Nitrate and nitrite
of synthetic wastewater
Ammonium of synthetic
wastewater
Soluble biodegradable
organic nitrogen of
synthetic
wastewater
Particulate biodegradable
organic nitrogen of
synthetic wastewater
Sludge floc density
Sludge floc viscosity
Sludge floc diameter
Pre-set average sludge
concentration
Sludge age
Air density
Air viscosity
Air bubble diameter
Air bubble diameter
modification coefficient
Wall roughness constant
Wall roughness height
Modification coefficient a
Modification coefficient b
Modification coefficient g
Yield for heterotrophic
biomass
Maximum specific growth
rate for heterotrophic
biomass
Half-saturation coefficient
for heterotrophic
biomass
Decay coefficient for
heterotrophic biomass

Notation

Unit

Value

rL
mL
SO

kg/m3
kg/m/s
mg/L

1000
0.0010
0.20

SI

mg/L

10.0

SS

mg/L

80.0

XI

mg/L

0.0

XS

mg/L

160.0

SNO

mg/L

0.0

SNH

mg/L

50.0

SND

mg/L

0.0

XND

mg/L

0.0

rs
ms
ds
Xset

kg/m3
kg/m/s
mm
g/L

1010
0.0046
0.40
3.8

SRT
rG
mG
dG
adG

d
kg/m3
kg/m/s
mm
Dimensionless

25
1.225
1.8  105
2.60
0.58

Cs
hKs
a
b
g
YH

Dimensionless
m
Dimensionless
Dimensionless
Dimensionless
g/g

1
0.02
0.92
0.44
0.75
0.64

mH

1/d

5.80

KS

g/m3

22.0

bH

1/d

0.70

The multiphase flow in OD is agitated by the moving parts


such as rotating blades, disc aerators and submerged impellers.
Here, the rotation of the rotating blades and disc aerators is
simulated by a moving wall model and the submerged impeller
described by a fan model. For further details see Yang et al. (2011).

2.4.

Parameter estimation

Activated sludge flocs are considered to be in a pseudo-solid


phase, whose density ranges from 1010 to 1060 g/mL

207

(Dammel and Schroeder, 1991) and dynamic viscosity ranges


from 3.8 to 11.0 mPa S (Jin et al., 2004), and mean floc diameter
from 0.05 to 0.5 mm (Grijspeerdt and Verstraete, 1997).
The interaction force between the gas and liquid phases is
closely related to the bubble diameter, and is given by:
dG 2:9adG

 1=3
sdo
grL

(9)

where do is diameter of the aeration orifice, s is surface tension


of liquid phase, g is the acceleration due to gravity, rL is density of the liquid phase, and adG is a coefficient allowing for the
aeration orifice configuration and the presence of pseudosolid phase.
The mass transfer rate corresponding to the pseudo solidphase is greatly dependent on the coefficients a[0.92,1.0]
(Stenstrom and Gilbert, 1981), b[0.44,0.98] (Gillot and Heduit,
2008), and g (to be calibrated experimentally).
Table 2 lists the major stoichiometric and kinetic parameters (after eliminating the pseudo-solid phase effect). Parameters related to heterotrophs growth and decay processes
(such as mH, YH, KS and bH) are modified in the present model
compared to ASM1 reflecting an improved understanding of
biodegradable material storage (see e.g. Gujer et al., 1999;
Krishna and Van Loosdrecht, 1999; Su and Yu, 2006). The
maximum specific storage rate and size of floc-like sludge
particles are both less than for granular sludge, and their
physical properties (e.g. moisture content and negative surface charge) are also different (Grijspeerdt and Verstraete,
1997; Liu et al., 2005; Wang et al., 2012). In general, the parameters in the present model lie between those of ASM1 and
the granular sludge system (Su and Yu, 2006).

3.
Experimental measurements in pilotscale oxidation ditch
Fig. 3 depicts a plan view of the pilot-scale carrousel-type
oxidation ditch. The ditch was fabricated from plexiglass, and
comprised four straight 1.15 m lengths of channel each with
semi-circular end channels, the smaller semi-circles having
radius 0.35 m, the larger semi-circles having radius 0.7 m. The
total working volume was 1.4 m3. The channels had rectangular cross-section of width 0.35 m and still depth 0.5 m. Two
surface impellers (Impeller 1 and Impeller 2) and four submerged stirrers (Stirrer 1, Stirrer 2, Stirrer 3 and Stirrer 4)
located in the curved channels drove the recirculating flow in
the oxidation ditch. Each spindle-like impeller consisted of 18
steel strips; each stirrer comprised an S-shape blade of
diameter 0.2 m Table 3 lists the operating rotational speeds
and angular directions of the impellers and stirrers for the
calibration Case I and validation Case II. Air was introduced
from a series of 1 m long gas distributors located at the base of
the second and third channels, and also from the surface
entrainment effect of the impellers. Each orifice of the gas
distributors was of diameter 0.1 mm. The overall aeration rate,
controlled by a rotameter, was 2.2 m3/h. Synthetic wastewater, originally stored in a tank of volume 1.8 m3, was
pumped into the ditch at a flow rate of 0.1 m3/h, and this flow
rate then maintained for a hydraulic residence time of 14 h.

208

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

The composition of synthetic wastewater was as follows:


1000 L of tap water, 250.0 g of sugar (approximately 250 mg/L
COD), 107.2 g of ammonium chloride (50 mg/L total nitrogen
concentration), 61.3 g of Na3PO4,12H2O, 500.0 g of NaHCO3,
3.0 g of FeSO4,7H2O, 10.0 g of CaCl2, 12.0 g of MgSO4, and 50 mL
of trace element solution. The trace element solution contained (per liter of tap water): 3.5 g of ethylene diamine tetraacetic acid, 2.0 g of ZnSO4,7H2O, 1.0 g of CuSO4,5H2O, 1.0 g of
MnSO4,7H2O, 1.0 g of Na2MoO4,2H2O, 1.0 g of H3BO3 and 0.2 g
of CoCl2,6H2O. A settling tank of volume 0.15 m3 separated the
sludge which is recycled at the ditch inlet. The sludge recycle
ratio was 100%, and the average activated sludge concentration in the ditch was maintained at about 3.8 g/L. The sludge
retention time was 25 d. All experiments were performed at
atmospheric pressure and room temperature.
A High-resolution Acoustic Doppler Velocimeter for the
Laboratory (Vectrino II, developed by Nortek AS, Vangkroken, Norway) was used to measure 3D liquid velocities at the
Sections 1-1, 2-2 and 3-3 as shown in Fig. 3. Twice a week,
activated sludge was sampled 0.1 m from the bed at M1, M2,
M3, M4, M5 and M6 (Fig. 3). The mixed liquor suspended
solid (MLSS) of each sample was determined using Method
2540 D (APHA, 1998). Twice daily, water samples for were
obtained 25 cm above the bed at W1, W2, W3, outlet, W4 and
W5 (Fig. 3). Each water sample was immediately filtered
using a 0.45 mm filter, and then stored at 4  C in refrigerator.
Soluble COD was determined by the closed reflux titrimetric
method 5220C (APHA, 1998). Ammonia nitrogen and nitrate

concentrations were measured by Nesslers reagent spectrophotometry HJ 535-2009 (Environmental Protection


Agency of China, 2009) and ultraviolet spectrophotometric
screening 4500-NO
3 B (APHA, 1998), respectively. DO concentrations were also monitored twice daily by an oxygen
probe (YSI-550A, YSI, Yellow Springs, Ohio, USA) located
0.25 m above the bed at W1, W2, W3, outlet, W4 and W5.
Ranges of primary variables experimental conditions are
listed in Table 4.

4.

Results and discussion

4.1.

Numerical model calibration

A 3D unstructured tetrahedral mesh was created for the OD


system using GAMBIT (Fluent Corporation, 2005) with mesh
refinement in the rotating and aerator zones. The governing
equations at steady state were solved using the finite volume
computer code, FLUENT. The grid independent analysis was
done, in which the liquid velocity was selected to conduct the
mesh test. Three different sets of meshes with cells of 55946,
161987 and 367881 respectively, were chosen to simulate the
liquid velocity field in the OD. As a result, the second set was
selected for all computations, considering the low difference
(less than 5%) of liquid velocities simulated with the optimal
mesh (161987 cells) and the refined mesh (36788 cells). The
segregated solver of Fluent 6.3 was used, with default

Fig. 4 e Comparison between measured and simulated concentrations of: (a) MLSS, (b) DO, (c) soluble COD, (d) ammonia
nitrogen, and (e) nitrate, at the sampling locations for Case I.

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

209

Fig. 5 e Comparisons of simulated and measured liquid velocity components at different layers over cross-section 1-1:
surface (a), top (b), middle (c) and bottom (d) for Case II.

210

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

parameter settings applied. All simulations were computed on


a workstation equipped with two Intel Xeon 2.93 GHz processors and 24 GB RAM. Each run took about 24 h of CPU time
to reach steady state condition.
The inlet boundary conditions in the numerical model of
the pilot-scale OD utilized measured parameters relating to
the liquid and pseudo-solid phases. The inlet had a rectangular cross-section of breadth 0.15 m, depth 0.05 m, and area
0.0075 m2. At the inlet, the flow speeds of the combined liquid
and pseudo-solid phases were kept at 0.0074 m/s. The synthetic wastewater was composed of SNO 0 mg/L, SNH 50 mg/
L, SND 0 mg/L and XND 0 mg/L. The quantity of slowly
biodegradable substrate was about twice that of readily
biodegradable substrate (Henze et al., 2000; Orhon et al., 1997).
Sugar provided the carbon source for the synthetic wastewater. The particulate inert organic concentration was
assumed zero, following Le Moullec et al. (2011). Hence, the
COD composition was: SI 10 mg/L, SS 80 mg/L, XS 160 mg/
L and XI 0 mg/L.
In order to calibrate the numerical model, the input parameters were adjusted iteratively in order to minimize the
following objective function (Squires, 2001):
q

Pn
2
1
i1 xci  xmi
nn1
Pn
OF
1
i1 xci
n

(10)

where OF was the normalized standard error; xmi and xci were
the measured and the calculated results of the i-th parameter,
and n was the number of monitored samples. Table 5 lists the
input parameters determined by minimizing the objective
function for MLSS, DO, COD, ammonia nitrogen and nitrate
concentration, where the corresponding errors were 2.7%,
12.7%, 2.8%, 6.7% and 6.6%, respectively. Fig. 4 shows the close
agreement between the simulated and measured variables
after calibration, Case I.

4.2.

Activated sludge distribution

The model was calibrated using one group of data at Case I


and further verified using another group of data obtained
under Cases II. In Fig. 5, comparison of the simulated and
measured 3D liquid velocities at Section 1-1 demonstrated
reasonable agreements between them with normalized standard error less than 9.6% at Case II.
Fig. 6 shows the numerically predicted horizontal velocity
component field 0.1 m from the bed and the simulated
streamwise-vertical velocity component field at longitudinal
sections taken along the and third channel in the carrousel
oxidation ditch for Case II. In Fig. 6a, vectors of the horizontal
velocity components are superimposed on contours of
magnitude of the horizontal velocity components. Strong

Fig. 6 e Case II: (a) predicted horizontal velocity component distribution, 0.1 m above bed; (b) predicted stream-wise-vertical
velocity component distributions along the third channel.

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

clockwise-rotating forced vortices are evident at Stirrers 2, 3,


and 4, and a weaker anti-clockwise flow deflection can be seen
at Stirrer 1. The surface impellers help direct flow from the wall
to the interior of the OD. The gas distributors cause an upwelling effect driving a ring vortex (which appears like two
counter-rotating vortices in the two-dimensional plane view)
through which a vertical flow passes, later radiating away from
a stagnation point close to the free surface, as shown in Fig. 6b.
From the inlet onwards, the overall horizontal plane flow
conditions in Fig. 6a are as follows: the flow spreads out to
move along Channel 1; it then runs into the opposing flow from
Impeller 1 and enters a strong mixing zone towards the end of
the straight portion of Channel 1. It appears that some material
can remain trapped for considerable time in Channel 1. On
entering Channel 2, the flow is essentially uniform across the
breadth of the ditch, but runs into the opposing flow generated
by Stirrer 1 and is driven upwards by the air bubbles at the
bottom aerator, causing a persistent horizontal eddy-like
feature to form about halfway along Channel 2, perhaps
linked to the ring vortex generated by the bottom aeration
system. Stirrer 1 rotates in the anti-clockwise direction, and
this causes a weakly deflected flow into Channel 3. This flow
meets the rotating flow from Impeller 2 and the vertical upflow
from the bottom aerators (Fig. 6b), and again an eddy-like
feature forms in the mixing region towards the middle of
Channel 3. The flow in Channel 4 is partly driven by Impeller 2

211

and is initially fairly uniform, but appears to separate at a


stagnation point about 0.25 m along the internal wall, after
which a recirculation zone develops. The flow then meets the
strongly rotating vortex associated with Stirrer 2, and stagnates. A free anti-clockwise rotating eddy occupies the final
quarter of Channel 4, and is divided from the forced vortex of
Stirrer 2 by a transverse flow towards another stagnation point
at the external wall. Strong forced clockwise rotating vortices
can be seen at Stirrers 1, 2, and 3. Mixing regions and stagnation
points are evident between the stirrers.
Figs. 7 and 8 present the results obtained for the water
quality parameters for Case II. Fig. 7a shows the predicted
MLSS distribution, where the liquid sludge was taken to be in
the pseudo-solid phase closely coupled with liquid and gas
phases. Superimposed on Fig. 7a are six experimental measurements of MLSS taken from the pilot-scale laboratory test
samples. It can be seen that the 3D three-phase model successfully predicted the observed processes, i.e. activated
sludge tended to settle out where the liquid velocity was
small, especially in the nearly stagnant zones and slowly
rotating free eddies. Fig. 8a presents a comparison of the
modeled and measured MLSS concentrations. The normalized
standard error between the numerically predicted and laboratory measured MLSS concentration is less than 2.8%,
demonstrating that the proposed model usefully described
the interaction between liquid and pseudo-solid phases in the

Fig. 7 e Case II: (a) MLSS concentration distribution, 0.1 m above the bed; (b), (c), (d) and (e) DO, COD, ammonia nitrogen and
nitrate distribution, 0.25 above the bed (where D indicates measured data values at the sampling points in Fig. 3).

212

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Fig. 8 e Comparison between measured and simulated concentrations of: (a) MLSS, (b) DO, (c) soluble COD, (d) ammonia
nitrogen, and (e) nitrate, at the sampling locations for Case II.

OD in reproducing the transport and sedimentation of activated sludge.

4.3.

Modified oxygen mass transfer

Fig. 7b shows the predicted DO distribution, on which are


superimposed five experimental measurements of DO taken
from the pilot-scale laboratory test samples. The DO content
rises in a fairly uniform fashion from a minimum at Impeller 1
to a maximum in Channel 3, after which DO content reduces
along Channel 4. Fig. 8b also shows the predicted DO concentrations for the validation Case II without and with the
pseudo-solid phase (based on the mass transfer model preand post-modification) for Case II. It is obvious that the
modified model (with pseudo-solid phase) provides a much
closer fit to the measurements. This is further verified by
calculating the normalized standard errors obtained by
comparing the two sets of predicted values against the
experimental data. Here, the normalized standard error is
24.6% using the unmodified model for a 0.92 and b 0.44.
The normalized standard error is reduced to 9.9% by modifying the pseudo-solid phase effect through the parameter g
in Equation (6) (Mena et al., 2011; Su and Yu, 2006). The very
satisfactory agreement between the simulated and measured
DO concentrations confirms that the proposed model is
capable of providing an accurate description of the liquidegasesolid interaction in terms of oxygen mass transfer.

4.4.

Water quality

In predicting the water quality, the stoichiometric and kinetic


parameters in ASM1 were modified along with the oxygen
mass transfer model. Fig. 7cee show the predicted soluble
COD, ammonia nitrogen, and nitrate distributions. In each
plot, the five experimental measurements are superimposed.
For the COD and ammonia nitrogen, both contour plots tell a
similar story: a contaminant hot spot can be seen immediately
downstream of the inlet, and a further high concentration
zone at the end of Channel 1. The water quality then consistently improves throughout Channels 2, 3, and 4. Fig. 7e shows
that the nitrate distribution follows an almost inverse trend to
the ammonia nitrogen, as would be expected. Fig. 8cee show
the close agreement between the numerically predicted and
measured values of soluble COD, ammonia nitrogen, and nitrate respectively at different locations in the OD for Case II. In
general, both soluble COD (Figs. 7c and 8c) and ammonia nitrogen (Figs. 7d and 8d) decline in the OD from a peak at the
inlet to a low values of about 16 mg/L and 4 mg/L respectively
at W2, after which there is little further change. The nitrate
concentration (Figs. 7e and 8e) increases substantially from
the inlet to W2, and seems to saturate at about 13 mg/L
beyond. The normalized standard errors between the simulated and measured soluble COD, ammonia nitrogen and nitrate concentrations are 1.5%, 1.5% and 3.3%, respectively.
Hence, it can be concluded that the present model provides a

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

reasonable description of the interactions between species in


liquid and pseudo-solid phases. The modified stoichiometric
and kinetic parameters in presence of solid phase lie between
those of ASM and granular sludge system (Su and Yu, 2006).

5.

Conclusions

A 3D three-phase fluid model based on computational fluid


dynamics was developed that satisfactorily represents phase
motion and phase interactions in an oxidation ditch. Activated sludge flocs were interpreted as having pseudo-solid
phase, and the related parameters modified by calibrating
the sludge viscosity, settling capacity, oxygen mass transfer
rate, and carbon substrate uptake due to adsorption on activated sludge. The assumption that activated sludge was in a
pseudo-solid phase made it possible to describe its transport
and sedimentation. By modifying the oxygen mass transfer
model, and the stoichiometric and kinetic parameters, the
numerical model was able to represent biochemical transformation of sludge. Experimental data on mixed liquor suspended solid (MLSS), DO concentration, soluble COD,
ammonia nitrogen, and nitrate were obtained from a pilotscale carrousel oxidation ditch. The numerical predictions of
flow field in the OD showed the great importance of the impellers and stirrers in promoting mixing. The excellent
agreement obtained between the numerical simulations and
sampled data measurements for water quality parameters
indicated that the numerical model accurately simulated the
kinematics of the multi-phase flow and the carbon oxidation,
nitrification and denitrification processes in the OD. The present paper has focused on model calibration and verification,
and used the results to provide some insights into the
behavior of an oxidation ditch. In future, it is recommended
that the model be used for parameter studies as a design tool
in helping to select optimal arrangements of impellers, stirrers, and aeration zones in planned oxidation ditches.

Acknowledgments
Financial support from National Natural Science Foundation
of China (Grant No. 21261140336/B070302) is very much
appreciated. Sincere thanks are also to Professor Alistair G.L.
Borthwick at Department of Engineering Science, Oxford
University for his careful editing on the manuscript.

references

mand, L., Carlsson, B., 2012. Optimal aeration control in a


A
nitrifying activated sludge process. Water Res. 46 (7),
2101e2110.
APHA, 1998. Standard Methods for the Examination of Water and
Wastewater, 20th ed. American Public Health Association,
Washington, DC.
Beccari, M., Dionisi, D., Giuliani, A., Majone, M., Ramadori, R.,
2002. Effect of different carbon sources on aerobic storage by
activated sludge. Water Sci. Technol. 45 (6), 157e168.

213

Carucci, A., Dionisi, D., Majone, M., Rolle, E., Smurra, P., 2001.
Aerobic storage by activated sludge on real wastewater. Water
Res. 35 (16), 3833e3844.
 gn, A.S., Orhon, D., Rossetti, S., Majone, M., 2011. Short-term
Cg
and long-term effects on carbon storage of pulse feeding on
acclimated or unacclimated activated sludge. Water Res. 45
(10), 3119e3128.
Dammel, E.E., Schroeder, E.D., 1991. Density of activated sludge
solids. Water Res. 25 (7), 841e846.
Environmental Protection Agency of China, 2009. Water qualityDetermination of ammonia nitrogen-Nesslers reagent
spectrophotometry (HJ 535-2009). China Environmental
Science Press, Beijing.
Fayolle, Y., Cockx, A., Gillot, S., Roustan, M., Heduit, A., 2007.
Oxygen transfer prediction in aeration tanks using CFD.
Chem. Eng. Sci. 62 (24), 7163e7171.
Fluent Corporation, 2005. Gambit 2.2 Users Guide, Lebanon, NH,
USA.
Fluent Corporation, 2006. Fluent 6.3 Users Guide, Lebanon, NH, USA.
Gillot, S., Heduit, A., 2008. Prediction of alpha factor values for
fine pore aeration systems. Water Sci. Technol. 57 (8),
1265e1269.
Grijspeerdt, K., Verstraete, W., 1997. Image analysis to estimate
the settleability and concentration of activated sludge. Water
Res. 31 (5), 1126e1134.
Gujer, W., Henze, M., Mino, T., van Loosdrecht, M., 1999. Activated
sludge model No. 3. Water Sci. Technol. 39 (1), 183e193.
Henze, M., Gujer, W., Mino, T., van Loosdrecht, M., 2000. Activated
Sludge Models ASM1, ASM2, ASM2d and ASM3. Scientific and
Technical Report 9, IWA Task Group on Mathematical
Modelling for Design and Operation of Biological Wastewater
Treatment. IWA Publishing, London.
Hong, K.H., Chang, D., Hur, J.M., Han, S.B., 2003. Novel phased
isolation ditch system for enhanced nutrient removal and its
optimal operating strategy. J. Environ. Sci. Health Part AToxic/Hazard. Subst. Environ. Eng. 38 (10), 2179e2189.
Insel, G., Artan, N., Orhon, D., 2005. Effect of aeration on nutrient
removal performance of oxidation ditch systems. Environ.
Eng. Sci. 22 (6), 802e815.
Jin, B., Wilen, B., Lant, P., 2004. Impacts of morphological, physical
and chemical properties of sludge flocs on dewaterability of
activated sludge. Chem. Eng. J. 98 (1e2), 115e126.
Krishna, C., Van Loosdrecht, M.C.M., 1999. Substrate flux into
storage and growth in relation to activated sludge modeling.
Water Res. 33 (14), 3149e3161.
Kulkarni, A.A., 2007. Mass transfer in bubble column reactors:
effect of bubble size distribution. Industrial Eng. Chem. Res. 46
(7), 2205e2211.
Kumar, B., Rao, A.R., 2009. Oxygen transfer and energy dissipation
rate in surface aerator. Bioresour. Technol. 100 (11), 2886e2888.
Le Moullec, Y., Potier, O., Gentric, C., Leclerc, J.P., 2011. Activated
sludge pilot plant: comparison between experimental and
predicted concentration profiles using three different
modelling approaches. Water Res. 45 (10), 3085e3097.
Littleton, H.X., Daigger, G.T., Strom, P.F., 2007. Application of
computational fluid dynamics to closed-loop bioreactors: II.
Simulation of biological phosphorus removal using
computational fluid dynamics. Water Environ. Res. 79 (6),
613e624.
Liu, Y., Liu, Y., Wang, Z., Yang, S., Tay, J., 2005. Influence of
substrate surface loading on the kinetic behaviour of aerobic
granules. Appl. Microbiol. Biotechnol. 67 (4), 484e488.
Mantziaras, I.D., Stamou, A., Katsiri, A., 2011. Effect of operational
cycletime length on nitrogen removal in an alternating
oxidation ditch system. Bioprocess Biosyst. Eng. 34 (5), 597e606.
Mena, P., Ferreira, A., Teixeira, J.A., Rocha, F., 2011. Effect of some
solid properties on gas-liquid mass transfer in a bubble
column. Chem. Eng. Process. 50 (2), 181e188.

214

w a t e r r e s e a r c h 5 3 ( 2 0 1 4 ) 2 0 0 e2 1 4

Orhon, D., Ates, E., Sozen, S., Cokgor, E.U., 1997. Characterization
and COD fractionation of domestic wastewaters. Environ.
Pollut. 95 (2), 191e204.
Pipes, W.O., 1969. Types of activated sludge which separate
poorly. J. Water Pollut. Control Fed. 41 (5), 714e724.
Schmid, M., Thill, A., Purkhold, U., Walcher, M., Bottero, J.Y.,
Ginestet, P., Nielsen, P.H., Wuertz, S., Wagner, M., 2003.
Characterization of activated sludge flocs by confocal laser
scanning microscopy and image analysis. Water Res. 37 (9),
2043e2052.
Squires, G.L., 2001. Practical Physics. Cambridge University Press,
Cambridge, UK.
Stamou, A.I., 1994. Modelling oxidation ditches using the IAWPRC
activated sludge model with hydrodynamic effects. Water Sci.
Technol. 30 (2), 185e192.
Stamou, A.I., 1997. Modelling of oxidation ditches using an
open channel flow 1-D advection-dispersion equation and
ASM1 process description. Water Sci. Technol. 36 (5),
269e276.
Stamou, A.I., Katsiri, A., Mantziaras, I., Boshnakov, K.,
Koumanova, B., Stoyanov, S., 1999. Modelling of an
alternating oxidation ditch system. Water Sci. Technol. 39
(4), 169e176.

Stenstrom, M.K., Gilbert, R.G., 1981. Effects of alpha-factor, betafactor and theta-factor upon the design, specification and
operation of aeration systems. Water Res. 15 (6), 643e654.
Su, K.Z., Yu, H.Q., 2006. A generalized model for aerobic granulebased sequencing batch reactor. 1. Model development.
Environ. Sci. Technol. 40 (15), 4703e4708.
Wang, J., Qiu, Z., Chen, Z., Li, J., Zhang, Y., Wang, X., Zhang, B.,
2012. Comparison and analysis of membrane fouling between
flocculent sludge membrane bioreactor and granular sludge
membrane bioreactor. PLoS One 7 (7).
Yang, Y., Yang, J., Zuo, J., Li, Y., He, S., Yang, X., Zhang, K., 2011.
Study on two operating conditions of a full-scale oxidation
ditch for optimization of energy consumption and effluent
quality by using CFD model. Water Res. 45 (11), 3439e3452.
Zhang, D., Guo, L., Xu, D., Chen, Y., 2010. Simulation of
component distributions in a full-scale carrousel oxidation
ditch: a model coupling sludge-wastewater two-phase
turbulent hydrodynamics with bioreaction kinetics. Environ.
Eng. Sci. 27 (2), 159e168.
Zhou, X., Guo, X., Han, Y., 2012. Enhancing nitrogen removal in an
Orbal oxidation ditch by optimization of oxygen supply:
practice in a full-scale municipal wastewater treatment plant.
Bioprocess Biosyst. Eng. 35 (7), 1097e1105.

Potrebbero piacerti anche