Development of Computational Fluid Dynamics Based Artificial Neural Network Metamodels For Wastewater Disinfection

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 155

ii

Development of Computational Fluid Dynamics based Artificial Neural


Network Metamodels for Wastewater Disinfection

A Thesis
Submitted to the Faculty
of
Drexel University
by
Wangshu Wei
in partial fulfillment of the
requirements for the degree
of
Doctor of Philosophy
September 2018
iii

To My Family
iv

ACKNOWLEDGEMENTS
First of all, I want to thank my academic advisors, Professors Bakhtier Farouk, and

Charles Haas. I would not be successful without their kind support, guidance and

mentoring. With their consideration, I conquered all the challenges on the road. With

their help, I went through the most difficult time of my life during this period. With their

advice, I corrected the direction of the path of my life. They are not only my teachers in

research, but also my leaders in life.

Secondly, I am very very grateful to my parents. They sacrificed their life to realize my

dream. Their support was the strongest motivation for me to achieve what I believed.

During the past five years, we were ten thousand kilometers apart, we did not have much

time to be together. Their perseverance was the example that I am following all the time.

At last, I would say thankyou to my beloved wife, Yue Wang. It was her love, companion

and caring that made this thesis possible. Her love was my courage when facing

predicament, her companion was my consolation in this foreign place, her caring was my

lighthouse when I felt lost.

Thank PeroxyChem (Mr. Philip Block ), the department of Mechanical Engineering and

Mechanics and the L.D. Betz Chair Funds (via Professor Charles Haas) for financial

support.
v

Table of Contents
ACKNOWLEDGEMENTS .......................................................................................................... iv
LIST OF FIGURES...................................................................................................................... vii
LIST OF TABLES......................................................................................................................... xi
ABSTRACT .................................................................................................................................. xii
I. INTRODUCTION .................................................................................................................... 1
II. BACKGROUND AND LITERATURE REVIEW .............................................................. 5
2.1 CFD techniques in wastewater process ............................................................................. 5
2.2 Turbulent modeling in CFD ............................................................................................... 8
2.3 Applications of PAA in wastewater treatment................................................................ 13
2.4 ANN Based Metamodel ..................................................................................................... 16
III. COMPUTATIONAL FLUID DYNAMICS MODELING .............................................. 19
3.1 Introduction ....................................................................................................................... 19
3.2 Mathematical model .......................................................................................................... 19
3.3 Turbulent Modeling .......................................................................................................... 21
3.4 Grid Generation and Independence Study ..................................................................... 24
3.5 Validation of CFD models................................................................................................. 26
IV. MIXING OF TWO GASEOUS FLOWS AND MISCIBLE LIQUIDS IN A T-
JUNCTION ......................................................................................................................... 32
4.1 Introduction ....................................................................................................................... 32
4.2 Problem Statement ............................................................................................................ 34
4.3 Mathematical Formulation ............................................................................................... 38
4.4 Numerical Scheme, Boundary and Initial Conditions ................................................... 41
4.5 Results and Discussions..................................................................................................... 44
V. THERMAL MIXING OF TWO LIQUIDS IN A T-JUNCTION ..................................... 67
5.1 Introduction ....................................................................................................................... 67
5.2 Problem Description.......................................................................................................... 68
5.3 Methodology....................................................................................................................... 70
5.4 Results and Discussions..................................................................................................... 77
5.5 Conclusions ........................................................................................................................ 89
VI. CFD SIMULATIONS ON PILOT SCALE WASTEWATER TREATMENT
REACTOR WITH THE STUDY OF MIXING ENHANCEMENT....................................... 90
vi

6.1 Problem Description.......................................................................................................... 90


6.2 3-D CFD Model of the Pilot Scale Reactor ...................................................................... 91
6.3 Results and Discussions..................................................................................................... 94
6.4 Mixing Enhancement with Inlet Variations .................................................................... 95
6.5 Conclusions ...................................................................................................................... 103
VII. DEVELOPMENT OF ANN BASED METAMODEL ON WASTEWATER
TREATMENT PROCESSES WITH PAA ............................................................................. 104
7.1 Introduction ..................................................................................................................... 104
7.2 Methodology..................................................................................................................... 105
7.3 Development of 3-D CFD Model .................................................................................... 111
7.4 Results and discussions ................................................................................................... 114
7.5 Development of Metamodel based on ANN .................................................................. 117
7.6 VALIDATION OF ANN BASED METAMODEL....................................................... 121
7.7 Conclusions ...................................................................................................................... 128
VIII. SUMMARY AND SUGGESTIONS FOR FUTURE RESEARCH ........................... 131
8.1 Summary .......................................................................................................................... 131
8.2 Suggestion for Future Research ..................................................................................... 133
LIST OF REFERENCES.......................................................................................................... 135
VITA ........................................................................................................................................... 140
LIST OF PUBLICATIONS ...................................................................................................... 140
vii

LIST OF FIGURES
Figure 1 Techniques for metamodeling (Wang and Shan, 2007) .................................................. 18

Figure 2 The grid generation of a T-junction pipe (with 900,000 cells) ....................................... 25

Figure 3 Grid Independence Study of (a) T-junction mixing ........................................................ 26

Figure 4 Comparison of methane concentration from the simulations and the experiments at R =

11.0, d/D = 0.014, z/D = 3.0, Rem = 21,000, Rej = 4000, Sct = 0.9 ............................................. 27

Figure 5 Comparison of grayscale images (a) experiment (b) CFD simulation at 15:00 Sec ....... 29

Figure 6 T-junction simulation with inclined jet inlet (a) 3-degree forward; (b) 3-degree backward

....................................................................................................................................................... 31

Figure 7 Contour plot of tracer with a rough nozzle ..................................................................... 31

Figure 8 Schematic of 3-D T-junction geometry considered (not for scale) ................................. 35

Figure 9 Grid Generation of the T-junction with d/D = 0.1 .......................................................... 43

Figure 10 Normalized cm (maximum concentration) distribution of air-methane and water-PAA

solution cases along the axial length for d/D = 0.1 and 0.2 ........................................................... 44

Figure 11 Contour plot of normalized (a) methane, Sc = 0.75; (b) PAA, Sc = 700 concentration at

different cross-section plane. d/D = 0.2, Rem = 200, Rej = 40 ..................................................... 46

Figure 12 Plot of degree of laminar mixing along the downstream length for the air-methane and

the water-PAA solution cases, d/D = 0.1 and 0.2, Rem = 200, Rej = 40 ...................................... 48

Figure 13 Comparison of methane concentration from the simulations and the experiments

(Forney and Lee, 1982, Forney, Noureddine and Hanh, 1996) at R = 11.0, d/D = 0.014, z/D = 3.0,

Rem = 21,000, Rej = 4000, Sct = 0.9 ............................................................................................ 49

Figure 14 Normalized maximum concentration of methane at different axial locations (compared

with experimental data) ................................................................................................................. 50


viii

Figure 15 Contour plots of methane concentration at center vertical plane at different times d/D =

0.1, R = 3.73, Rem = 21,000, Rej = 4000, Sct = 0.9 ..................................................................... 52

Figure 16 Contour plot of methane concentration at the exit plane, time average from 3.600 sec to

3.620 sec; R = 2.45 and d/D = 0.2, Rem = 21,000, Rej = 6000, Sct = 0.9 .................................... 53

Figure 17 Comparison of liquid-liquid mixing and gas-gas mixing at R = 2.45, d/D = 0.2, Rem =

21000, Rej = 6000 Sct = 0.9 and R = 3.73, d/D = 0.1, Rem = 21000, Rej = 4000 Sct = 0.9 ........ 55

Figure 18 Contour plots of PAA solution concentration at center vertical plane at different times

(a) t =20.0 s (b) t =20.5 s (c) t = 21.0 s (d) t = 21.5 s and (e) t = 22.0 s; ....................................... 57

Figure 19 Contour plot of PAA concentration at the exit plane, with R = 3.73 and d/D = 0.1, time

average period: 49.80 sec to 50.00 sec, Rem = 21,000, Rej = 4000, Sct = 0.9 ............................. 58

Figure 20 Comparison of grayscale images of the side-view instantaneous concentration field for

R = 3.06, d/D = 0.1667, Rem = 20,850, Rej = 10,630. Screenshot of simulation is taken at 15.00

sec .................................................................................................................................................. 60

Figure 21 Instantaneous velocity field on the center plane of the tee mixer, R = 3.06, d/D =

0.1667, Rem = 20,850, Rej = 10,630. Screenshot of simulation is taken at 15.00 sec .................. 62

Figure 22 Time-averaged tracer concentration at vertical center plane, where R = 3.06,

Rem=20,850, Rej=10,630. Time averaging period from the present CFD simulation: 22.0 sec to

22.2 sec .......................................................................................................................................... 63

Figure 23 Contour plots of tracer (sodium fluorescein) concentration at center vertical plane at

different times (a) t =15.0 s (b) t =15.5 s (c) t = 16.0 s (d) t = 16.5 s and (e) t = 17.0 s; d/D =

0.1667, R = 3.06, Rem = 20,850, Rej = 10,740, Sct = 0.9 ............................................................ 65

Figure 24 Schematic of 3-D T-junction geometry considered (not for scale) ............................... 69

Figure 25 Grid Generation of the T-junction (a)prospective view; (b) Cross-sectional view of the

T-junction ...................................................................................................................................... 74

Figure 26 y+ at the top wall of the main pipe after the jet nozzle at y = 0.0381m. ....................... 76
ix

Figure 27 Visualized image of the interface between two flows in Z/A = 0 from (a) Experiment;

(b) CFD simulation result .............................................................................................................. 78

Figure 28 Mean velocity vector diagram in the symmetric plane of Z/A = 0. (The solid lines show

the time-averaged interface between the main and branch flows obtained by the flow

visualization.) (a) Experiment; (b) CFD simulation result ............................................................ 79

Figure 29 Profiles of the mean normalized temperature at (a) x = 0; (b) x = 1.0; (c) x = 3.0 ....... 81

Figure 30 Temperature contour at center plane at (a) t = 15.00 sec; (b) t = 15.30 sec; (c) t = 15.60

sec; (d) t = 15.90 sec; (e) t = 16.20 sec; at TR = 1.28, VR = 3.0. .................................................. 83

Figure 31 Time-averaged temperature contour plot at exit plane (15 pipe diameter downstream).

TR = 1.28, VR = 3.0. ..................................................................................................................... 84

Figure 32 Temperature contour at center plane at (a) t = 15.00 sec; (b) t = 15.30 sec; (c) t = 15.60

sec; (d) t = 15.90 sec; (e) t = 16.20 sec; at TR = 1.09, VR = 3.0. .................................................. 86

Figure 33 Time-averaged temperature contour plot at exit plane (15 pipe diameter downstream).

TR = 1.09, VR = 3.0. ..................................................................................................................... 86

Figure 34 Comparisons of the degree of mixing between TR = 1.28 and TR = 1.09, where (a) VR

= 3.0; (b) VR = 2.0 ........................................................................................................................ 88

Figure 35 The Pilot-scale Serpentine Reactor. (Photo courtesy of PeroxyChem)......................... 90

Figure 36 Zoomed-in Inlet pipe..................................................................................................... 91

Figure 37 (a) Geometric model of 3-D CFD simulations;............................................................. 92

Figure 38 Grid generation of serpentine contactor ........................................................................ 93

Figure 39 (a) mechanical diamgram of baffle plate. (b) CFD geometric model of baffle plate .... 93

Figure 40 Contour plots of PAA concentration at 5.0 sec simulating time at (a) center plane; (b)

exit ................................................................................................................................................. 94

Figure 41 Instanteous contour plot of PAA in the first channel of the serpentine contactor at 9.0

sec. ................................................................................................................................................. 95
x

Figure 42 Three jet inlet configurations. (a) Inserted straightly (b) inserted with 90-degree turn

downstream; (c) inserted with 90-degree turn upstream ............................................................... 96

Figure 43 Contour plots of the tracer concentration at center vertical plane (a) Inserted straightly

(b) inserted with 90-degree turn downstream; (c) inserted with 90-degree turn upstream............ 97

Figure 44 Contour plots of tracer concentration at cross-sectional planes (a) Inserted straightly (b)

inserted with 90-degree turn downstream; (c) inserted with 90-degree turn upstream ................. 98

Figure 45 Contour plots of tracer concentration at the outlet (a) inserted vertically; (b) inserted

with a 90-degree turn downstream; (c)inserted with a 90-degree turn upstream ........................ 100

Figure 46 The plot of the degree of mixing ................................................................................. 101

Figure 47 The plot of pressure head (a) inserted vertically; (b) inserted with a 90-degree turn

downstream; (c)inserted with a 90-degree turn upstream ........................................................... 102

Figure 48 (a): top view; (b) prospective view of a full-scale reactor. ......................................... 106

Figure 49 (a) Contour plot of PAA Residual concentration; (b) Contour plot of N survival ratio at

center horizontal plane ................................................................................................................ 115

Figure 50 Plot of CFD predicted N survival ratio at the exit....................................................... 116

Figure 51 Schematic of an ANN based metamodel structure...................................................... 117

Figure 52 Performance Plot of N survival ratio. ......................................................................... 120

Figure 53 Performance Plot of PAA residual concentration. ...................................................... 121

Figure 54 Validation plot of N survival ratio metamodel ........................................................... 123

Figure 55 validation plot of PAA residual concentration metamodel ......................................... 124

Figure 56 The configuration of the pilot scale reactor ................................................................ 125

Figure 57 Validation plot with experiments in the pilot scale reactor. A 1 log experimental error

is shown for comparison.............................................................................................................. 126


xi

LIST OF TABLES
Table 1 Listing of the main variables ............................................................. 37

Table 2 Listing of the main variables for the mixing problems ...................... 70

Table 3 List of variables ............................................................................... 109

Table 4 The range of the 11 variables ........................................................... 110

Table 5 Sample space generated by HSS ...................................................... 111

Table 6 The values of variables in the base model ....................................... 114

Table 7 Criteria for survival ratio ................................................................. 119

Table 8 Criteria for PAA residual concentration .......................................... 119

Table 9 Dataset for validation ....................................................................... 122

Table 10 Parameter table for the experiment after scale-up ......................... 125

Table 11 Parameter table with Sf = 3,000..................................................... 127

Table 12 Parameter table with Sf = 30,000................................................... 127

Table 13 Average bias and Mean squared error in the sensitivity study ...... 127
xii

ABSTRACT

Development of Computational Fluid Dynamics based Artificial Neural


Network Metamodels for Wastewater Disinfection
Wangshu Wei
Bakhtier Farouk, Advisor, Ph.D
Charles N Haas, Advisor, Ph.D

The mechanisms of the mixing of two streams are investigated numerically. The transport

processes of both mass and heat transfer in the mixing of either two gases or two miscible

liquids are studied with 3-D time-dependent computational fluid dynamics (CFD)

simulations. The flow fields are calculated for the T-junction (of two circular cross-

section pipes that meet orthogonally at a junction) which is one of the most common

mixing devices in the study of mass and heat transfer. For turbulent flow regimes, the

large eddy simulation (LES) technique is employed. Second-order differencing scheme is

applied in the simulation to minimize numerical diffusion. The results obtained by the

numerical simulations are verified with available experimental data in the literature for

air-methane mixing in a T-junction. In the wastewater disinfection process, Peracetic acid

(PAA) is an effective anti-microbial agent for disinfection and has a significant

applicability in wastewater treatment processes. Similar to other chemical disinfection

technologies, such as chlorination, and ozonation, the performance of PAA is determined

by hydraulics of a disinfection contactor, chemical kinetics of microbial inactivation, and

specific water characteristics. On-site physical modelling on pilot scale has been an

important and usual method to validate the kinetics and the performance for a particular

system. However, physical modelling is expensive and time consuming. CFD technique

offers an alternative to physical modelling. With increasing computational power on


xiii

personal computers (PCs), applying CFD simulation on a full-scale reactor is becoming

more and more achievable. With proven accuracy on predicting the performance of full

scale facilities, CFD is now playing a significant role in the wastewater treatment

industry. A transient CFD simulation of a full-scale contactor with reasonable accuracy

can be achieved within hours on PCs, which significantly lower the cost of undertaking

physical modelling experiments. Metamodels based on artificial neural network (ANN)

are computationally efficient mathematical approximations to a highly complex system

with multiple non-linear features. With a parametric study of CFD simulations providing

training dataset, the metamodel is able to be calibrated to predict the performance of a

variety contactors with different geometries, inlet conditions and chemical kinetics. In

this thesis, the results of microorganism inactivation and PAA decay in full scale

contactors using three-dimensional CFD models are presented. Then the results of the

simulations were applied to develop ANN-based metamodels. The performance of

metamodels were validated with both CFD simulations and pilot scale experiments.
1

I. INTRODUCTION

Mixing of fluid streams plays a key role in many areas in the industry and in nature. An

optimum condition of mixing is desired for enhancement of species mixing, heat and mass

transfer, combustion and other chemical reactions. For example, good mixing can promote

the efficiency of fuel combustion in an engine. In wastewater disinfection systems, good

mixing of a disinfectant with wastewater enhances the efficiency of treatment.

One of the most encountered mixing devices is a T-junction pipe. The T-junction containing

a transverse pipe and a 90º branch is widely used in mixing of two fluids (in gaseous or

liquid form). The geometry can be utilized in a chemical reactor, combustion chamber,

piping system, etc. Gas mixture flows appear in diverse applications. Examples include

pollutant dispersion, chemical processing, and combustor mixing and reaction. Previous

discussion was carried out on the mixing of two gases (air and methane) with no heat

transfer considered in a T-junction (Forney and Lee, 1982). However, it is also common

that the two streams are having different temperature. For example, cool air flow is mixing

with a hot air flow in an HVAC unit (Hirota, Mohri, Asano and Goto, 2010). Liquid-liquid

mixing problems with temperature variations are also common in industry and in nature. A

previous study of liquid-liquid mixing problem was the mixing of water with water-sodium

solution in a T-junction (Pan and Meng, 2001). T-junctions are also used for the mixing of

wastewater and disinfectant in wastewater treatment process, the mixing of hot water with

cooling water in a power plant (Naik-Nimbalkar, Patwardhan, Banerjee, Padmakumar and


2

Vaidyanathan, 2010), etc. Therefore, the hot water mixing with cooling fluid is considered

as a candidate in liquid-liquid mixing problem in a T-junction.

The mixing of two flows in a T-junction mixing chamber contains mass and heat transport

processes, thus the process is complex. Previous studies on T-Junction mixing showed the

mixing behaviors and flow structures between the main flow and the jet in detail (Wei,

Farouk and Haas, 2017). With increasing power of computing, CFD simulations are applied

in the present study to predict the flow properties and mixing characteristics of two fluids

(gas-gas or liquid-liquid) at a 90o T- junction of two pipes as well. As a balance on

computational cost and flow resolution, LES is becoming one of the most popular turbulent

modeling techniques in CFD. LES has been proven to be accurate on both mass transfer

and heat transfer processes.

In this study, CFD simulations of the mixing of two gaseous flows is carried out first for

validation purpose. In the gaseous mixing case, a cold air flow is introduced from the main

inlet, a stream of hot air is injected from the jet flow inlet. The main flow Reynolds number

is denoted by Rem, inlet jet Reynolds number is denoted by Rej. The mixing of two streams

is studied under turbulent flow conditions, where Rem ~ 15,000. After the fidelity of the

CFD model is validated, another series of simulations are focused on the mixing of two

water streams, i.e. cool water flow mixing with hot disinfectant solution (0.3% water-

peracetic acid mixture) to represent the scenario in a pilot scale wastewater treatment

reactor. In the liquid-liquid study, cold water flow is introduced from the main pipe inlet

and hot water is injected from the jet flow inlet. The mixing of the two liquid flows are

studied with a Reynold number of ~20,000.


3

Peracetic acid (PAA) is an efficient disinfectant for its wide spectrum of antimicrobial

ability. The significance of applying PAA as a disinfectant in wastewater treatment

industry has been emerging in recent years. The characteristics of PAA for microbial

wastewater disinfection such as ease of implementing treatment (without the need for

expensive capital investment), broad spectrum of activity even in the presence of

heterogeneous organic matter, absence of persistent toxic or mutagenic residuals or by-

products, no quenching requirement (i.e., no de-chlorination), short contact time, are all

desirable to the wastewater treatment industry, which makes PAA gain its popularity in

environmental engineering applications. CFD simulations can be used to predict performance

of chemical disinfection systems. However computational times remain sufficiently long to

prevent their use in optimal design. An artificial neural network (ANN) based metamodel is

considered as an alternative, A metamodel is a mathematical approximation to a multi-

dimensional system (Barton, 1994). Because of this nature, to perform metamodeling on

wastewater treatment process would reduce the computation cost dramatically (Hoque, Farouk

and Haas 2010; Wei, Farouk and Haas 2016). Therefore, in this study, ANN based metamodel to

approximate CFD results was assessed. The ANN metamodel was trained from a series of CFD

simulations of disinfection characteristics of PAA in a wastewater treatment contactor. The

design space was sampled using a quasi-random sampling technique. Quasi-random sampling is

one of the ‘space-filling’ designs (Wang and Shan, 2007). To generate space filling designs

for computer experiments or numerical simulations, sampling methods are applied to select

points in the design space t o create uniformly-spaced designs by minimizing discrepancy of

the sequence (Hoque, 2011). 40 different CFD cases on an 11-dimensional space are obtained

and used as input to the training process of metamodel development. Metamodels were

developed to predict disinfectant residual concentration and microbial inactivation rate on full

scale reactors. The performance of the ANN based metamodel is evaluated by comparison to
4

CFD simulation results and pilot-scale experimental measurements. As a mathematical

approximation method to a high dimensional nonlinear system, the ANN based metamodel shows

its ability to provide an efficient yet accurate solution to the wastewater disinfection process with

PAA.
5

II. BACKGROUND AND LITERATURE REVIEW


2.1 CFD techniques in wastewater process

After being applied for over twenty years, CFD has made significant contributions to the

design and optimization in the wastewater disinfection industry. CFD has been used, not

only in flow simulations, but also in reacting processes such as disinfection, and in the

optimal designs of reactors. Though it has demonstrated powerful abilities, computational

efficiency, and flexibility that can be used to resolve most kinds of problems, it is still

facing some challenges.

The first application of CFD in wastewater disinfection is to simulate flow. In most

research that applied CFD, the common stages of a complete disinfection process

simulation are: flow simulation, tracer transport simulation, reaction process simulation,

and inactivation simulation.

In order to study the hydraulic efficiency of the water treatment system, tracer transport

simulation is applied. In the simulation, tracer is always treated as an inert particle. Either

a Lagrangian approach of particle tracking or an Eulerian approach of solving a transport

equation for concentration has been applied. However, latter method is more widespread

since most CFD codes are based on Eulerian system. For reaction process simulation, the

main purposes are to predict inactivation, and also potentially formation of disinfectant

byproducts. With this purpose, a transport equation of all species has to be solved. In the

inactivation simulation, reactions between microorganisms and disinfectant are

introduced as what is done on the previous stage.


6

The first two-phase flow CFD code for ozone disinfection came into being at the end of

1990s (Cockx et al., 1999; Ta and Hague, 2004; Bolaños et al., 2008; Bartrand et al.,

2009; Talvy et al., 2011) Within last two decades, the development of CFD has been

conducted to solve more complicated problem, increase the accuracy, and save the

computation costs. Not only simulating flows in lab-scale contactors, but also applied to

full-scale reactors, CFD became the most reliable and most widely used method in this

field comparing to other modeling methods such as plug flow reactor (PFR) or

completely mixed flow reactor (CMFR) (Zhang et al., 2014). For parametric studies,

there are many parameters that would affect the disinfection process. The parameters

include: pH, temperature, flow rate, bubble size variation (for ozone disinfection), inlet

configurations, and so on. Optimization of contactor configuration aims to improve

disinfection efficiency economically. How to reduce short-circuiting, how to reduce dead

zones are the key to achieve the optimal efficiency. Increasing the number of baffles,

rearranging the reactors, adjusting the locations of inlet, outlet, and diffusers would make

contributions.

Though CFD techniques have been being developed rapidly in last two decades, there are

still several challenges that we are facing. These challenges include: 1) unsteady flow

structure effect, 2) multi-phase flow effect, 3) complexity of reaction system, 4)

uncertainty of inactivation kinetics, and 5) closure problem for chemical source terms.

These challenges may cause a lot of inaccuracies that could limit their applicability.

In order to overcome unsteady flow structure effect, the method has to be chosen

appropriately for each particular problem. For instance, in UV disinfection application,

Wols et al. (2010b) found that RANS wrongly predict local flow features around a UV
7

lamp which is a typical unsteady flow structure in a flow around a blunt body. LES was

employed and matched the experimentally measured velocity profile better than RANS.

Zhang et al. (2014b) investigated a baffled contactor and a column contactor which are

typically used for ozone and chlorine disinfection and showed that LES is a more reliable

strategy than RANS in simulating tracer transport in column contactors due to its ability

to better predict the spatial transition to turbulence characterizing the flow. Therefore, in

further study, higher resolution approaches such as LES, detached LES, or even DNS

should be involved if computational cost could be affordable. The multi-phase flow effect

doesn’t need to be considered in liquid-liquid flows, i.e. chlorine disinfection or peracetic

acid disinfection, as opposed to the ozone disinfection process. Most previous ozone

studies Cockx et al., 1999; Ta and Hague, 2004; Bolaños et al., 2008; Talvy et al., 2011).

neglected this effect because of two reasons: 1) unknown parameters, such as bubble size

distribution, mass transfer coefficients, models for closure of the two-phases, and so on;

2) simulation of single-phase flows are mathematically simpler with less computational

cost. However, this effect could not be neglect especially when studying in a column

contactor. The gas flows from bottom of the column upwards to the top because of

pressure gradient, while water goes against the direction that the ozone flows in. Thus, it

can impact the overall flow significantly. In future study, more simulations and

experiments of this effect need to be done so that we can understand the role of a gas

flow in disinfection process much better. The complexity of reaction system is caused by

the complex reaction mechanisms in the disinfection process. For example, once ozone

pumped into wastewater flow, many reactions are triggered. There is not only the ozone
8

decay reaction, a reaction between dissolved ozone and natural organic matters (NOM),

but also by-product formation reactions.

2.2 Turbulent modeling in CFD

For flow simulation, the biggest problem is how to describe a turbulent flow. Typical

Reynolds number in turbulent flow region for an internal flow is above 2,000, which is

most common scenario in the wastewater treatment process. Turbulent flow is very

complicated: it has a wide range of spatial and temporal scales from large spatial scales

that can be comparable to the size of flow, to the small scales even less than millimeters.

A similar situation arises for the temporal scales. Because of the random fluctuation in

turbulence, it is more difficult to model turbulent region. Therefore, a turbulence

modeling will help to reduce the difficulty of predicting the turbulent region. There are

three most popular turbulence models: Reynolds-averaged Navier–Stokes (RANS)

equations, Large-eddy Simulation (LES) and Direct Numerical Simulation (DNS).

DNS has the best accuracy of the three, because it solves Navier-Stokes equation directly

on the all spatial and temporal scales. However, DNS has a higher demand on fine grids

model, i.e. the grid for DNS should contain about Re9/4 points (Pope, 2000). For example,

if the Reynolds number for the mean flow is around 20,000, then it requires a meshing

with 4×109 cells, which make a great rise on computational time would take thousands of

hours computing on a PC. So far, we still cannot afford the computational cost today.

LES is an alternative to DNS, which has a good accuracy and an ability to predict the

spatial transition to turbulence characterizing the flow. RANS have a good computational

efficiency and accuracy on resolving mean flow.


9

In RANS methods, there are two major families, they are: k-ε and k-ω Family. k-ε

family has different models such as: Standard k-ε Model, RNG K-ε Model, Realizable k-

ε Model, Two-Layer k-ε Model (ESI-Group, 2017). According to CFD-ACE modules

2017 (ESI-Group, 2017), the k-ω Family contains such as k-ω SST, Wilcox etc. The

main idea is to describe the flow field with two introduced variables: one is for the

turbulent kinetic energy, k, and the other is either for the rate of dissipation, ε., or for the

specific rate of dissipation, ω.

There are two transported variables in k-ϵ model. The first variable determines the energy

in the turbulence and is called turbulent kinetic energy (k). The second variable is the

turbulent dissipation (ε) which describes how much the turbulent kinetic energy will

dissipate to the surrounding fluid. The idea of k-ϵ model is Reynolds decomposition,

which mathematically separates the time-averaged and fluctuating parts of the flow. The

flow velocity ui is to be decomposed as:

ui = ui + ui ' (0.1)

where ui is the averaged laminar velocity, ui’ is turbulent fluctuation.

Turbulent kinetic energy in per unit mass, k, is defined as:

1 2 2 2
k= (u + v + w ) (0.2)
2

Then the transport equations for k-ϵ models are:

∂( pk ) ∂( pu j k ) ∂  µt ∂k 
+ = ρ P − ρε + (µ + )  (0.3)
∂t ∂x j ∂x j  σ x ∂x j 

and
10

∂( pε ) ∂( pu jε ) ρ Pε ρε 2 ∂  µt ∂ε 
+ = Cε1 − Cε 2 + ( µ + )ε  (0.4)
∂t ∂x j k k ∂x j  σ ∂x j 

The production term P is defined as:

 ∂u ∂u 2 ∂um  2 ∂um
P =ν t  i + j − δ − k
 ∂x ∂x 3 ∂x ij  3 ∂x (0.5)
 j i m  m

The five constants used in this model are as follows:

C µ = 0.09,

C ε1 = 1.44,

Cε 2 = 1.92,

σ x = 1.0,

σ s = 1.3

Compared to RANS, LES solves a “filtered” set of Navier-Stokes equation instead

(Yang, 2014). Filtering of LES is a mathematical operation intended to remove a range of

small scales from the solution to the Navier-Stokes equations. Because the principal

difficulty in simulating turbulent flows comes from the wide range of length and time

scales, this operation makes turbulent flow simulation cheaper by reducing the range of

scales that must be resolved. The LES filter operation is low-pass, meaning it filters out

the scales associated with high frequencies. A LES filter can be applied to a spatial and

temporal field and the filter size is decided by maximum grid size and time step. Any

spatial fluctuation smaller than grid size will be filtered out, and similarly, any temporal
11

frequency higher than time step frequency will be filtered out as well. Therefore, filtered

governing equations become as following:

Filtered conservation of mass

∂ ui
=0 (0.6)
∂ xi

Filtered conservation of momentum

∂ ui ∂ 1 ∂p
+ (ui u j ) = − + ν∇ 2 ui + f i (0.7)
∂ t ∂x j ρ ∂ xi

Furthermore, the filtered advection term can be split up, following Leonard (Leonard,

1974),

u i u j = τ ijr + u i u j (0.8)

where τ ijr is the residual stress tensor. The filtered momentum equations become

(neglecting body force),

∂ ui ∂ 1 ∂p ∂ ∂ τ ijr
+ (u i u j ) = − + 2ν S ij − (0.9)
∂t ∂x j ρ ∂ xi ∂x j ∂x j

where S ij is the rate-of-strain tensor.

Applying the Smagorinsky model (Smagorinsky, 1963) for sub-grid scale turbulence,

1
τ ijr − τ iiδ ij = 2ν t S ij (0.10)
3

where ν t is turbulent eddy viscosity.

A relationship between ν t and the filter size Δ can be found as,


12

ν t ~ ∆2 2S ij S ij (0.11)

The Smagorinsky model gives,

ν t = cs2∆2 2S ij S ij (0.12)

where cs is called Smagorinsky constant, cs = 0.1 has been chosen in the present

simulations.

As shown clearly above, RANS is solving the averaged Navier-Stokes equation which

gives a much better efficiency but at the same time, it fails to predict the fine structure of

the flow. LES on the contrary, is able to capture the large scale structure in the flow

which results in a much better flow realization. Wols et al. (2010) showed that the

accuracy of RANS is weakened when dealing with unsteady flow structure. In this

research, RANS wrongly predicted local flow features around a UV lamp which was a

typical unsteady flow structure in a flow around a blunt body. In this case, LES was

employed and matched the experimentally measured velocity profile better than RANS.

In order to overcome unsteady flow structure effect, the method has to be chosen

appropriately for each particular problem. Zhang et al. (2014b) investigated a baffled

contactor and a column contactor which are typically used for ozone and chlorine

disinfection. Results showed that LES is a more reliable strategy than RANS in

simulating tracer transport in column contactors due to its ability to better predict the

spatial transition to turbulence characterizing the flow. Therefore, in further study, higher

resolution approaches such as LES, detached LES, or even DNS should be involved if

computational cost could be affordable.


13

2.3 Applications of PAA in wastewater treatment

Peracetic acid (PAA) is the peroxide of acetic acid (AA). The product is

commercially available as a quaternary equilibrium solution of acetic acid (CH3COOH),

hydrogen peroxide (HP, H2O2), PAA (CH3CO3H) and water, as shown in the following

equation (Alasri et al., 1992; Gehr et al., 2002; Block, 1991):

CH3CO2H + H2O2 ⇆ CH3CO3H + H2O

where CH3CO2H = acetic acid, CH3CO3H = peracetic acid; H2O2 = hydrogen peroxide.

Industrially, PAA is mass-produced from the reaction of acetic acid with hydrogen

peroxide in a sulfuric-acid environment, where sulfuric acid acts as a catalyst (Block,

1991). The PAA solution is always a mixture of PAA, AA and HP. Although hydrogen

peroxide is also considered as an effective disinfectant which can make additional

contribution to the disinfection power of the PAA mixture, PAA is a stronger

antimicrobial disinfectant compared to HP, and is able to rapidly react at low

concentrations against a wide spectrum of microorganisms (Baldry, 1983; Baldry and

French, 1989b; Fraser et al., 1984). It was found that HP required much larger doses than

PAA for the same level of disinfection (Wagner et al., 2002).

PAA in water is decomposed to AA, oxygen, and water, which is,

2CH3CO3H + H2O ⇆ 2CH3CO2H + H2O + O2

It has been shown that PAA produces no known toxic disinfection by-products (DBPs) in

water after disinfection (Baldry and Fraser, 1988; Monarca et al., 2001; Monarca et al.,

2002). This is one of the most important advantages of PAA over other typically used

chemical disinfectants in the treatment of wastewater or water such as gaseous chlorine,


14

sodium hypochlorite, chlorine dioxide, or ozone (Monarca et al., 2002). Once the

disinfectant is added to microbial containing wastewater, its consumption is driven by a

series of different reactions. There are three major reactions in which PAA is consumed

in water solution: instantaneous consumption, self-decay and the consumption during

reaction with microorganisms (Antonelli et al., 2013; Gehr et al., 2002; Yuan et al.,

1997).

Liberti and Notarnicola (1999) identified the PAA disinfectant activity as consequence of

the release of active oxygen, which likely oxidizes sensitive sulfhydryl and sulfur bonds

in proteins, enzymes, and other metabolites. According to the hypothesis advanced by

Pelczar and Chan (1986) the whole inactivation process is composed of four different

mechanisms:

1. Damage to the microorganisms’ cell wall;

2. Alteration of cytoplasmic membrane permeability;

3. Alteration of colloidal nature of protoplasm;

4. Inhibition of enzyme activity.

Gehr et al. (2002) concluded that for municipal primary effluents, PAA doses of 2 to 6

mg/L were needed to achieve 1000 colony forming units (CFU) per 100 ml of fecal

coliforms with a contact time of 60 min. For the secondary effluents, lower PAA doses of

0.6 to 4 mg/L were required to achieve 1000 CFU per 100 ml of fecal coliform. Lefevre

et al. (1992) recommended a dose in the range between 5 mg/L to 7 mg/L and 60 minutes

contact time in order to achieve effluent total coliform concentrations of less than 1000

CFU/100 mL and effluent fecal coliform concentrations of less than 100 CFU/100 mL.
15

PAA was also suggested to be applied with ultraviolet (UV) irradiation for wastewater

disinfection (Rajala-Mustonenet al., 1997). Because of the short residence time in UV

reactors, using PAA was an effective reinforcement to a short UV exposure. If

wastewater was in contact with PAA for a relatively longer time first, and then being

exposed to a short UV irradiation, a significant reduction of microorganisms could be

achieved.

Several factors that affecting PAA disinfection processes were identified. PAA can be

efficient in a wide range of temperatures. Nevertheless, in general, as in other chemical

reactions, the inactivation of microbial would be accelerated with an increased water

temperature during disinfection with PAA (Stampi et al., 2001). Another feature to be

considered is pH. With a lower pH value, PAA showed with greater activity (Colgan and

Gehr, 2001). An ideal environment for PAA against fecal coliforms was under neutral or

mildly acidic conditions (Baldry and French, 1989a).

Major disadvantages associated with PAA disinfection are the increases of organic

content in the effluent due to acetic acid and thus in the potential for microbial regrowth

(Lazarova et al. 1998). Another drawback to the use of PAA is its high cost, which is

partly due to limited production capacity worldwide. However, if the demand for PAA

increases, especially from the wastewater industry, mass production capacity might also

be increased in the future, thus lowering the cost. In such a case, in addition to having

environmental advantages, PAA may also become cost-competitive with chlorine. Thus,

PAA might replace chlorine-based disinfectants in wastewater treatment processes.


16

2.4 ANN Based Metamodel

A metamodel, first proposed by Blanning (1975) is a model of a complex numerical

simulation that serves as a surrogate for the more computationally intensive simulations.

Metamodeling allows for wider exploration of the input variables, improves the

understanding of the model to be generated and enables further studies for solution

optimization (Tabach, et al., 2007, Broad, et al., 2005). Areas that metamodels play a role

in are (Wang and Shan, 2007; Hoque, 2011):

1. Model approximation: Approximation of computationally intensive processes

across the entire design space;

2. Design space exploration: The design space is explored to enhance the engineers’

understanding of the design problem by working on a computationally inexpensive

metamodel;

3. Problem formulation: Metamodels can assist the formulation of an optimization

problem that is easier to solve or more accurate than otherwise;

4. Optimization support: Metamodeling can be applied and integrated to solve

various types of optimization problems that involve computation-intensive functions.

Meta modeling evolved from classical design of experiments theory (Wang and Shan,

2007). These methods include factorial or fractional factorial design, central

composite design, Box-Behnken design and Plackett-Burman design (Box, et al., 1978,

Montgomery, 2005). These methods tend to spread the sample points around boundaries

of the design space and leave a few at the center of the design space. For computer

experiments space filling design are recommended (Wang and Shan, 2007, Sacks, et al.,
17

1989). These are the orthogonal arrays, various Latin hypercube designs, Hammersley

sequences and uniform designs (Wang and Shan, 2007). Different types of metamodels

can be created such as regression models, artificial neural networks, radial basis

functions, and Kriging stochastic model (Wang and Shan, 2007). Hammersley sequences

and uniform designs belong to a more general group called low discrepancy sequences.

Hammersley sampling has been found to provide better uniformity than Latin

hypercube designs (Diwekar and Kalagnanam, 1997). Latin hypercube design is

uniform in one dimensional projection while the other methods tend to be more uniform

in the entire space. The number of sampling points or the appropriate sampling size is

dependent on the complexity of the function to be generated. In general, more sample

points mean more information of the function but at a higher expense.

Metamodels have been used in a wide variety of applications, including calibrating water

distribution systems, modeling of chemical reactors, modeling of aircraft operation

predicting groundwater contamination and combat simulation (Alam, et al., 2004).

Levandler and Payraudeau (2007) developed a metamodel for optimizing the design of

storm water detention basins. Galleli and Soncini – Sessa (2010) applied the

metamodeling design approach for the design of release policies in reservoir systems

serving irrigation districts by modeling the dynamics of the irrigation demand through

metamodels. The metamodeling approach was also applied to a water quality and

hydraulics – based optimization problem for water distribution systems with the objective

of developing a metamodel as a substitute for the simulation model, EPANET (Broad, et

al., 2005). The developed metamodel was approximately 700 times faster than the

simulation model and similar optimal solutions were obtained (Broad, et al., 2005).
18

ANNs have increasingly been applied instead of the conventional regression models

to develop metamodels. This is because ANNs can m o r e e f f i c i e n t l y represent

complex, nonlinear functions than other forms (such as polynomials). Tabach et al. (2007)

developed an ANN based metamodel to assess groundwater contamination involving the

release of toxic or hazardous materials.

Figure below summarizes the techniques for metamodeling (Wang and Shan, 2007).

Figure 1 Techniques for metamodeling (Wang and Shan, 2007)

The most common metamodeling approach is to apply the design of experiments to

identify an efficient set of computer runs or numerical experiments and then apply

regression analysis techniques to create a polynomial approximation to the computer code

(Wang and Shan, 2007, Simpson, et al., 2001). Furthermore, a better solution of producing

a metamodel of CFD results with ANNs has emerged with the Hoque’s study (Hoque,

2011). In the study, it was shown that the ANN based metamodeling performed better than

a polynomial fitting (Hoque, 2011).


19

III. COMPUTATIONAL FLUID DYNAMICS

MODELING

3.1 Introduction

The numerical simulations were conducted using the commercial code CFD-ACE+ (ESI

Group, 2017). It was necessary for the current situation to be able to track particles in a

transient flow. CFD-ACE+ had the required features. Computational fluid dynamics

(CFD) is a mathematical modeling procedure whereby the fluid parameters are calculated

by solving the Navier-Stokes equations. To obtain a solution of the equations

numerically, a discretization method was applied which approximates the differential

equations by a system of algebraic equations. The approximations were applied to small

domains in space (grid size) and/or time (size of time step). These algebraic equations

were solved through iterative techniques.

The CFD model applied to simulate the wastewater disinfection process with PAA

comprised of two main sections: transport equations and chemical reactions. The sections

below are discussing about the two sections respectively.

3.2 Mathematical model

The mathematical model in the simulations of wastewater treatment process is listed

below:

Continuity Equation

∂ui
=0 (1.1)
∂xi
20

Momentum Equation

∂ui ∂ 1 ∂p  ∂u ∂u 
+ (ui u j ) = − +ν  i + j  + Fi (1.2)
∂t ∂x j ρ ∂xi  ∂x ∂x 
 j i 

where ν is the kinematic viscosity, Fi is the body force term.

Species Equation

∂ci ∂c ∂  ∂ci 
+ uj i −  Di  = Si (1.3)
∂t ∂x j ∂x j  ∂x j 

where ci is the concentration of species i, Di is the mass diffusivity, Si is the source

term, which involved with chemical reactions.

Chemical Reactions in Wastewater Disinfection Process

To simplify the chemical kinetics, two reactions are mainly considered in the disinfection

process. They are: the reaction of PAA decay and the inactivation of microorganisms (N),

where N* represented inactivated microorganisms:

PAA → PAA* (1.4)

PAA+ N → PAA+ N* (1.5)

A recent batch study(Tizzoni, 2016) shows that PAA decays via immediate demand plus

first order decay in wastewater, and the inactivation of microorganisms with PAA follows

Hom’s power law (Hom, 1972). Therefore, the chemical kinetics may be written as,

d [PAA]
= −kd [PAA] (1.6)
dt

d[ N ]
= −mk *[PAA]n t m−1[ N ] (1.7)
dt
21

where kd is the decay rate constant of PAA, k* is the Hom’s law rate constant, m and

n are Hom’s law exponents.

Since an implicit form of equation (3.6) is needed to apply in CFD simulation, the Hom’s

power law can be transformed to an equivalent form only in state variables(Haas and

Joffe, 1994):

d[ N ]
= − mk
1
* m
n
[PAA] m [ − ln(
(
[ N ] 1− 1 m
)]
)
(1.8)
dt [ N ]0

3.3 Turbulent Modeling

Two turbulent modeling techniques are considered in the research. For full scale

wastewater treatment simulations, a k-ϵ model is applied. On the other hand, in order to

realize the flow field with higher fidelity, LES is chosen in studies of T-junction mixing.

k-ϵ model
The flow velocity ui is to be decomposed as:

ui = ui + ui ' (1.9)

'
where ui is the averaged laminar velocity, ui is turbulent fluctuation.

Turbulent kinetic energy in per unit mass, k, is defined as:

1 2 2 2
k= (u + v + w ) (1.10)
2

Then the transport equations for k-ϵ models are:

∂( pk ) ∂( pu j k ) ∂  µt ∂k 
+ = ρ P − ρε + (µ + )  (1.11)
∂t ∂x j ∂x j  σ x ∂x j 
22

and

∂( pε ) ∂( pu jε ) ρ Pε ρε 2 ∂  µt ∂ε 
+ = Cε1 − Cε 2 + ( µ + )ε  (1.12)
∂t ∂x j k k ∂x j  σ ∂x j 

The production term P is defined as:

 ∂u ∂u 2 ∂um  2 ∂um
P =ν t  i + j − δ − k
 ∂x ∂x 3 ∂x ij  3 ∂x (1.13)
 j i m  m

The five constants used in this model are as follows (ESI-Group, 2017):

C µ = 0.09,

C ε1 = 1.44,

Cε 2 = 1.92,

σ x = 1.0,

σ s = 1.3

LES - Filtered conservation of mass

∂ ui
=0 (1.14)
∂ xi

LES - Filtered conservation of momentum

∂ ui ∂ 1 ∂p
+ (ui u j ) = − + ν∇ 2 ui + f i (1.15)
∂ t ∂x j ρ ∂ xi
23

Furthermore, the filtered advection term can be split up, following Leonard (Leonard,

1974),

u i u j = τ ijr + u i u j (1.16)

where τ ijr is the residual stress tensor. The filtered momentum equations become

(neglecting body force),

∂ ui ∂ 1 ∂p ∂ ∂ τ ijr
+ (u i u j ) = − + 2ν S ij − (1.17)
∂t ∂x j ρ ∂ xi ∂x j ∂x j

where S ij is the rate-of-strain tensor.

Applying the Smagorinsky model (Smagorinsky, 1963) for sub-grid scale turbulence,

1
τ ijr − τ iiδ ij = 2ν t S ij (1.18)
3

where ν t is turbulent eddy viscosity.

A relationship between ν t and the filter size Δ can be found as,

ν t ~ ∆2 2S ij S ij (1.19)

The Smagorinsky model gives,

ν t = cs2∆2 2S ij S ij (1.20)

where cs is called Smagorinsky constant, cs = 0.1 has been chosen in the present

simulations.
24

3.4 Grid Generation and Independence Study

Mesh generation was carried out using CFD-Geom (ESI-Group, 2017), which is capable

to generate both structured and unstructured mesh. For a structured messing, codes use

quadrilaterals as elementary volume in 2-D whereas in 3-D hexahedra are generated. On

the other hand, unstructured grids require to generate triangles in 2D or tetrahedrals in a

3D domain as in Figure 2. Because the simplicity of the model geometries, most of the

grid generations are structured messing. The models in the study of mixing enhancement

with different inlet configurations (in Chapter 6) are combined with both techniques.

Grid independence is a key factor in numerical simulations. The term grid-independency

means that the numerical results of the simulation would not change with

increasing/decreasing the grid size which minimizing the error due to the numerical error.

For example, a model with coarsened mesh could have a larger truncation error.

Sometimes the error is too large to neglect. Therefore, to perform a grid independence

study is necessary.

In the T-junction mixing study, four CFD model were generated with 300,000; 600,000;

900,000 and 1,200,000 structured cells respectively. Figure 2 below showed the grid

generation of 900,000 cells.


25

(a)

(b)

Figure 2 The grid generation of a T-junction pipe (with 900,000 cells)

Following the experimental work of Forney et al. (Forney and Lee, 1982, Forney,

Noureddine and Hanh, 1996), simulations of mixing of two gases (air and methane) in a

T-junction were carried out for grid independence study by carefully matching the

parameters with their experiments. As an indicator of mixing thoroughness, the cross-


26

sectional maximum concentrations were collected for comparison. The results are shown

in Figure 3.

Figure 3 Grid Independence Study of (a) T-junction mixing

It is clear that the model with 300,000 performed most poorly among four simulations. It

over-predicted the mixing thoroughness. However, when increasing the grid size from

900,000 to 1,200,000, there was very little difference between the two simulations.

Therefore, the model with 900,000 structured cells was used in the T-junction

simulations.

3.5 Validation of CFD models

In order to test if CFD simulations are providing reasonable predictions, validation is a

necessary step. One way to validate the CFD simulations is to compare the CFD

predictions with experimental results. In this study, the T-junction CFD model is test with

experimental measurements (Forney and Lee, 1982; Pan and Meng, 2001).
27

Forney et al. discussed the optimum mixing conditions of two gaseous flow in a T-

junction. In the experiments (Forney and Lee, 1982, Forney, Noureddine and Hanh,

1996), the two flow streams are air and methane-air mixture. The main flow Reynolds

number Rem = 21,000, and the jet flow Reynolds number Rej varies from 4,000 to 8,000.

Figure 4 below shows the normalized concentration profile of methane (c/cm) along the

center vertical line at the axial location where z/D = 3 and R (jet-to-pipe inlet velocity

ratio) = 11.0.

Figure 4 Comparison of methane concentration from the simulations and the experiments

at R = 11.0, d/D = 0.014, z/D = 3.0, Rem = 21,000, Rej = 4000, Sct = 0.9
28

Pan and Meng (Pan and Meng, 2001) conducted two sets of experimental measurements

on T-junction mixing in liquid-liquid flows. The main flow is water and the jet flow is a

sodium fluorescein solution in water. The Reynolds number of the main flow is fixed at

20,850 in both cases; Reynolds numbers of the jet flow are 10,630 and 17,500 (Pan and

Meng, 2001). Simulations were carried out for the case with main flow Reynold number

equals to 20,850 and the jet flow Reynolds number of 10,630.

The Figures below are the comparison of grayscale images of the side-view instantaneous

concentration field for R = 3.06, d/D = 0.1667, Rem = 20,850, Rej = 10,630. Screenshot of

simulation is taken at 15.00 sec.

(a) Instantaneous image (Pan and Meng, 2001)


29

(b) Present simulations


Figure 5 Comparison of grayscale images (a) experiment (b) CFD simulation at 15:00

Sec

As shown above, there is noticeable difference between the two plots. CFD is not able

to predict the fine flow structure. Because of the nature of LES, small structures are filtered

out. Furthermore, a roll-up vortex on the left of the jet in the experiment plot (marked by

red frame) did not show up in CFD simulations. These differences might be caused by

imperfect geometry lineup, or the residual swirl in either the jet or the main flow, etc. The

numerical errors contribute to the difference as well, for example, LES filtered out small

structures, which is one of the reason that simulation predicts less diffusion. Besides, the

difficulty on matching experimental time and simulation time could be another factor as

well. Though the simulation conditions were carefully matched with the experimental

conditions (Pan and Meng, 2001), the predicted concentration field lacks some of the

details found in the experimental work. The LES filters out the sub-grid level fluctuations.

The alternate here is to carry out the direct numerical simulations (DNS) which can become

prohibitively expensive in time and resources. In addition, the simulations are for ideal
30

conditions, assuming no upstream or laboratory disturbances. Such effects are unavoidable

in most experiments. These could contribute to the mismatch of the finer details of the

experimental and computational results presented here.

To test out the assumptions above, a series of 2-D cases with imperfect geometries were

carried out. In the plots below, two cases were considered first with inclined jet inlet. This

could be a common scenario in laboratory. Bartend et al., found a perfect CFD model of a

vertical reactor was not providing a good prediction. In order to get a better correlations

with his experiment, 10 inclination on the vertical tank was made (Bartend, 2009).

In this case, we first looked at the two cases with 30 inclination forward and backward

(towards upstream and downstream).

(a)
31

(b)

Figure 6 T-junction simulation with inclined jet inlet (a) 3-degree forward; (b) 3-degree

backward

With only 30 difference, there is some significant difference between the two results.

Especially in the case with 3-degree forward inclination, the roll-up vortex showed up.

Furthermore, when installed the T-junction in the laboratory, the intersection of the two

pipes might not be smooth ideally. Therefore, to add a small roughness by blocking the

nozzle with 5% width is considered.

Figure 7 Contour plot of tracer with a rough nozzle

The simulation showed that the development of the roll-up vortex and providing more

flow structure. The observation from Figure 6 to 7 supported the previous assumption.
32

IV. MIXING OF TWO GASEOUS FLOWS AND

MISCIBLE LIQUIDS IN A T-JUNCTION

4.1 Introduction

Mixing of fluid streams plays a key role in many areas in the industry and in nature. An

optimum condition of mixing is desired for enhancement of species mixing, heat and mass

transfer, combustion and other chemical reactions. For example, good mixing can promote

the efficiency of fuel combustion in an engine. In wastewater disinfection systems, good

mixing of a disinfectant with wastewater enhances the efficiency of treatment. The T-

junction geometry is widely used in industry and research for mixing of two dissimilar

fluids (in gaseous or liquid form).

Gas mixing problems are common in engineering. Gas mixture flows appear in diverse

applications. Examples include pollutant dispersion, chemical processing, and combustor

mixing and reaction. The first systematic study of gas mixing with a T-junction was

conducted by Chilton and Genereaux (Chilton and Genereaux, 1930) who used smoke

visualization techniques to determine optimum mixing conditions. They concluded that

right angle configurations were as effective as any other geometry for good mixing. Chilton

and Genereaux also found that when the ratio of the velocity of the secondary flow to the

velocity of the main flow was 2-3 for their geometries, satisfactory mixing was obtained in

2-3 pipe diameters. Narayan (Narayan, 1971) used quantitative methods to measure the

degree of mixing of air-carbon dioxide feed streams in three pipeline mixers. Narayan

(Narayan, 1971), like Chilton and Genereaux (Chilton and Genereaux, 1930), found it was
33

possible to achieve quality mixing in a few diameters with perpendicular jet injection

devices but that parallel flow devices required up to 250 pipe diameters.

Following the work of Forney et al. (Forney and Lee, 1982, Forney, Noureddine and

Hanh, 1996) we consider the mixing of two gases (air and methane) in a T-junction.

Mixing two different liquids is important in the field of microfluidics. A micromixer is

one of the most basic devices in microfluidics. Due to the small size of the device,

pressure driven flows in simple channels (i.e. with smooth walls) are laminar and mostly

uniaxial, so that the confluent liquids tend to flow side by side. As the Schmidt number

(the ratio between kinematic viscosity ν and molecular diffusivity between two species

D1-2, ν/D1-2) is typically large for two miscible liquids – the mixing between two parallel

laminar liquid streams is weak. The subscripts 1 and 2 in D1-2 stand for the main flow and

the jet flow components. T-shaped micromixers have been investigated extensively in

recent years, as they are quite suitable to carry out fundamental studies to understand

mixing of liquids at the microscale (Nguyen and Wu, 2005).

Liquid-liquid mixing problems are also common in industry and in nature. Recently liquid-

liquid mixing problems with T-junctions have been intensely studied for problems in

microchannels (Raj, Mathur and Buwa, 2010). CFD techniques have been widely applied

in simulating wastewater disinfection process in the past decade since they can provide the

high accuracy on predicting reactor hydraulics, concentrations of bacteria and disinfectant

(Zhang, Tejada-Martínez and Zhang, 2014). T-junctions are also used for the mixing of

wastewater with disinfectants. Wastewater disinfection is a complex process where mixing

of the disinfectant and the wastewater is vital in process performance. Treatment techniques

vary in industrial, municipal and agricultural wastewater treatment process, due to the
34

different contaminations and competing reactions. There are many disinfection

technologies being applied. Each technique has its own particular usage. In all cases of

chemical disinfection, it is necessary to mix a strong solution of disinfectant with a larger

volume of fluid. This current work is motivated by the use of peracetic acid (PAA) as a

disinfectant.

4.2 Problem Statement

Computational fluid dynamic (CFD) simulations are carried out in the present study to

predict the flow properties and mixing characteristics of two fluids (gas-gas or liquid-

liquid) at a 90o T- junction of two pipes of dissimilar diameters. Following the work of

Forney et al. (Forney and Lee, 1982, Forney, Noureddine and Hanh, 1996), we first

consider the mixing of two gases (air and methane) in a T-junction. For the turbulent flow

field predictions the LES technique is used to predict the spatial-temporal evolution of the

mixing process. A schematic of the mixing problem considered in the T-junction is shown

below (Figure 8). The two flows are labeled as the main flow (in the larger pipe with

diameter D) and the jet flow (in the smaller pipe with diameter d).
35

Figure 8 Schematic of 3-D T-junction geometry considered (not for scale)

The main pipe with D = 11.43 cm is placed horizontally along the z-axis. The jet pipe

with smaller diameter is on the top and placed vertically along the x-axis. The diameter of

the inlet jet can be varied in order to control the jet-to-pipe diameter ratio i.e. d/D. The

distance from injecting point to the outlet is 15 times of main pipe diameter to ensure the

flow is fully developed when it exits.

In general, the larger pipe contains the main solvent and a solute is injected into the main

flow via the T-junction. Forney et al. (Forney and Lee, 1982) investigated gas-phase mixing

in a T-junction. In a succeeding paper (Forney, Noureddine and Hanh, 1996), they showed

that an optimized mixing condition can be determined by varying the geometry of the T-

junction, i.e. the jet-to-pipe diameter ratio d/D and inlet velocity ratio R = U j / U m ). Based

on the jet and the main flows, we define two Reynolds numbers viz. Rej ( U j d / ν and Rem

( U m D / ν , where U j
and U m are the mean inlet velocity components for the jet and main

flows respectively and ν is the kinematic viscosity of the mixing components (assumed to

be the same). The laminar Schmidt number ν /D1-2 is around 0.7 for gas-gas mixing (say
36

air/methane mixing) problems; however, the corresponding laminar Schmidt number is two

to three orders higher in liquid-liquid mixing problems indicating a low degree of mixing

between two liquids in laminar flows (Raj, Mathur and Buwa, 2010). In general, the

laminar mixing characteristics of two gases beyond the T-junction is stronger than that of

two liquids. For turbulent flows however, the Schmidt number is defined as ν t/Dt,1-2 where

ν t is the so-called eddy diffusivity and Dt,1-2 is the (sub-grid scale) turbulent mass

diffusivity (Tominagaa and Stathopoulosh, 2007).

In this study, for gas-gas mixing in a T-junction, we consider the mixing of air and methane

(0.3% methane) following the work of Forney and co-workers (Forney and Lee, 1982). A

series of three-dimensional CFD/LES simulations were carried out to simulate the air-

methane experiments. To simulate the liquid-liquid mixing, we consider two cases: (a)

mixing of water with a PAA-water solution and (b) mixing of water with a sodium

fluorescein solution in water (Pan and Meng, 2001). Wastewater contains microorganisms

(e.g., coliforms and E.coli) and the inactivation of microorganisms can be achieved by PAA

(Wei, Farouk and Haas, 2016). The case of mixing water with a sodium fluorescein solution

in water (Pan and Meng, 2001) is of particular interest as the experimental study provides

results that can be compared to the present simulations.

CFD simulations of mixing of two gases is considered first. In the gas-gas mixing case,

air flows in the main pipe, and 0.3% methane-air mixture is injected from the jet flow

inlet (see Table 1). The jet-to-pipe diameter ratio, d/D is chosen from following values:

0.014, 0.025, 0.1, and 0.2. The mixing of air and methane is studied under both laminar

and turbulent flow conditions. In laminar flows, we consider Rem ~ 200 and Rej ~ 40. In

turbulent cases we consider, Rem ~ 20,000 and Rej ~ 4,000.


37

Table 1 Listing of the main and the ranges of the parameters

Item Symbol Gas-gas mixing Liquid-liquid

mixing

Main pipe Rem 200 (laminar), 200 (laminar),


Reynolds 21,000 (turbulent) 21,000 and 20,850
number (turbulent)
Inlet jet Rej 40 (laminar), 40 (laminar),
Reynolds 4000 – 6000 (turbulent)
number
Jet-to-pipe d/D 0.014, 0.1, and 0.2 0.1 and 0.2;
diameter ratio 0.1667
Laminar Sc 0.75 700
Schmidt number
Turbulent Sct ScSGS(Smagorinsky, ScSGS(Smagorinsk
Schmidt number 1963) y, 1963)
Jet-to-pipe inlet R (= 2.0 (laminar) 2.0 (laminar)
velocity ratio U j / Um ) 2.45,3.73, and 11.0 2.45, 3.06, and
(turbulent) 3.73 (turbulent)

With the same geometry, two additional series of simulations were carried out on the

mixing of two miscible fluids, i.e. (a) water and peracetic acid solution (PAA) and (b)

water and sodium fluorescein solution in water. In the liquid-liquid study, water flows in

the main pipe and PAA solution (or sodium fluorescein solution in water) is injected from

the jet flow inlet. The mixing of the two fluids (gas-gas or liquid-liquid) are studied in

laminar and turbulent flow conditions. The Reynolds number ranges (for the laminar and

turbulent flows) in liquid-liquid case are kept in the same range as in the gas-gas mixing

simulations. For the water- sodium fluorescein solution in water (Pan and Meng, 2001),

the case with Rem = 20,850, and Rej = 10,630 was considered for comparison with the

present predictions. All parameters for the mixing problems considered are specified in
38

the Table 1. In the case of PAA, reactions were neglected due to the relatively short

residence times in the system being studied.

4.3 Mathematical Formulation

Laminar flows

The governing equations of mass, momentum and concentration for laminar mixing

problem in a T-junction are described below. It is assumed that the kinematic viscosities of

the two mixing species are close to each other.

Conservation of mass:

∂ui
=0 (2.1)
∂xi

Conservation of momentum

∂ui ∂ 1 ∂p
+ (ui u j ) = − +ν∇2ui (2.2)
∂t ∂x j ρ ∂xi

where ν is the kinematic viscosity.

Concentration conservation

∂ci ∂c ∂ 2 ci
+ u j i = D1− 2 (2.3)
∂t ∂x j ∂x j ∂x j

where the index ‘i’ can be either 1 or 2 for the normalized mass concentrations (c1 or

c2) of the two mixing components such that

c1 + c2 = 1.0 (2.4)
39

Turbulent flows

For simulating turbulent flows in the T-junction, we need to invoke a turbulence model.

There are three most popular turbulence models: the Reynolds-averaged Navier–Stokes

(RANS) equations, the Large-eddy Simulation (LES) and the Direct Numerical

Simulation (DNS). As a compromise between the accuracy of results and resource

requirements, we have used the LES model to simulate the turbulent mixing problems in

the T-junction.

Compared to DNS, LES solves a “filtered” set of Navier-Stokes equation instead (Yang,

2014). Filtering of LES is a mathematical operation intended to remove a range of small

scales from the solution to the Navier-Stokes equations. Because the principal difficulty

in simulating turbulent flows comes from the wide range of length and time scales, this

operation makes turbulent flow simulation cheaper by reducing the range of scales that

must be resolved. The LES filter operation is low-pass, meaning it filters out the scales

associated with high frequencies. A LES filter can be applied to a spatial and temporal

field and the filter size is decided by maximum grid size and time step. Any spatial

fluctuation smaller than grid size will be filtered out, and similarly, any temporal

frequency higher than time step frequency will be filtered out as well. Therefore, filtered

governing equations become as following:

Filtered conservation of mass

∂ ui
=0 (2.5)
∂ xi
40

Filtered conservation of momentum

∂ ui ∂ 1 ∂p
+ (ui u j ) = − + ν∇ 2 ui + f i (2.6)
∂ t ∂x j ρ ∂ xi

Furthermore, the filtered advection term can be split up, following Leonard(Leonard,

1974),

uiu j = τijr + ui u j (2.7)

where τ ijr is the residual stress tensor. The filtered momentum equations become

(neglecting body force),

∂ ui ∂ 1 ∂p ∂ ∂ τ ijr
+ (u i u j ) = − + 2ν S ij − (2.8)
∂t ∂x j ρ ∂ xi ∂x j ∂x j

where S ij is the rate-of-strain tensor.

Applying the Smagorinsky model (Smagorinsky, 1963) for sub-grid scale turbulence,

1
τ ijr − τ iiδ ij = 2ν t S ij (2.9)
3

where ν t is turbulent eddy viscosity.

A relationship between ν t and the filter size Δ can be found as,

ν t ~ ∆2 2S ij S ij (2.10)
41

The Smagorinsky model gives,

ν t = cs2∆2 2S ij S ij (2.11)

where cs is called Smagorinsky constant, cs = 0.1 has been chosen in the present

simulations.

For the filtered concentration conservation equation,

∂ ci ∂ ∂q j
+ (u j ci ) = D1− 2 ∇ 2 c i − (2.12)
∂t ∂x j ∂x j

where D1-2 is diffusion coefficient, and qj is the sub-grid scale term, representing a

mass flux term which is critical to study mixing characteristics (Calmet and Magnaudet,

1997) . Applying the Smagorinsky model as the sub-grid scale model to the filtered

conservation equation,

νt ∂c
qj = (2.13)
ScSGS ∂x j

where ScSGS is the sub-grid-scale Schmidt number (~ henceforth the turbulent

Schmidt number Sct). A value of Sct = 0.9 is used in the present simulations.

4.4 Numerical Scheme, Boundary and Initial Conditions

The time-dependent simulations were carried out using the CFD-ACE+ software (CFD-

ACE, 2016) with a time step equal to 10-3 sec was used for the gas mixing cases and a

time step equal to 10-2 sec was set for the liquid mixing cases. A second order blended

upwind scheme was applied for the convective-diffusive terms and Crank-Nicolson

scheme with a blending factor of 0.7 was used for the temporal terms in the governing
42

equations. The internal relaxations for velocities and passive scalar were 0.2. The

convergence criterion for each variable is the same (set to 0.001 for the present

simulations) which is a requirement for the CFD-ACE+ solver (Meganathanm, 2018).

The grid generation (see Figure 9 below for a sample grid structure used in the study) was

accomplished in CFD-GEOM (the grid generator for CFD-ACE+ solver)(CFD-ACE,

2016). To balance the computation cost and the resolution of an eddy structure, 900,000

structured cells were applied in this model.

(a)
43

(b)
Figure 9 Grid Generation of the T-junction with d/D = 0.1

(a) cross-sectional view passing through the jet flow pipe; (b) perspective view.

At inlets of the domain, normal velocities in the profile of 1/7th power law were specified

for turbulent simulations. For laminar mixing on the other hand, a flathead velocity

profile was assigned. No-slip conditions were assigned for the walls. Initially, the main

pipe and was filled up with air (gas-gas flows) or water (liquid-liquid flows), while the jet

inlet was specified as air-methane mixture or PAA solution or sodium fluorescein

solution in water (Pan and Meng, 2001). At t ≥ 0.0 s, jet tracers, air-methane mixture or

PAA solution or sodium fluorescein solution in water was injected from jet inlet and air

or water started flowing from main inlet as well. The simulations were carried out for a

period of three times the theoretical residence time to diminish the effect of initial

conditions and ensure the flows become quasi-steady.


44

4.5 Results and Discussions

Laminar mixing of gas-gas and liquid-liquid streams

For laminar cases, the simulations were obtained for the geometry shown in Figure 8.

Mixing of air-methane (0.3% methane by volume in the jet flow) and water-PAA solution

(0.3% by volume of the jet flow) were studied with low Reynolds numbers, with Rem =

200, and Rej = 40 with jet-to-pipe inlet velocity ratio R = 1.0, and d/D = 0.1 and 0.2. Plots

of normalized maximum concentration of the solute cm - methane (for the gas-gas flows)

and PAA solution (for the water-PAA solution) cases are shown in Figure 10 below.

Figure 10 Normalized cm (maximum concentration) distribution of air-methane and water-PAA

solution cases along the axial length for d/D = 0.1 and 0.2
45

From Figure 10, it is evident that the mixing is faster for the gas-gas case (Sc = 0.7)

compared to the liquid-liquid case (Sc = 700.0) in laminar flows. The present simulations

also provide the nature of the 3-D mixing process in the flow geometry.

For the case described in Figure 8, the spatial distribution of the mixing components for

the air-methane the water-PAA solutions are shown in Figures 11 (a) and (b) respectively.

While the mixing process is quite vigorous for the air-methane case as evidenced in Figure

11 (a), the mixing of PAA solution in water is significantly less intense as shown in Figure

11 (b).

(a)
46

(b)
Figure 11 Contour plot of normalized (a) methane, Sc = 0.75; (b) PAA, Sc = 700 concentration at

different cross-section plane. d/D = 0.2, Rem = 200, Rej = 40

For the laminar liquid-liquid mixing, the PAA solution contours at downstream planes

tend to form a horse-shoe type vortex structure that retains its identity further

downstream. A recent research on thermal mixing of two water streams in a T-junction

(Karthick Selvam, 2017) had observed similar vortex structure at the near field of the T-

junction.

To evaluate the difference in the behavior of the two mixing cases, i.e. air-methane and

water-PAA solution, a degree of non-mixing parameter φ is defined as:

std (ci )
φ= (14)
mean (ci )
47

where std (ci) is the standard deviation of the tracer concentration on a given cross-section

plane; mean (ci) is the averaged tracer concentration on the same plane. A higher value of

φ would thus indicate a lower level of mixing. Figure 12 below shows the variation of φ

as a function of the normalized axial length (z/D ranging from 0.0 to 15.0) for the gas-gas

and liquid mixing cases for d/D values of 0.1 and 0.2 respectively. The degree of mixing is

good for the gas-gas (Sc = 0.75) mixing cases for d/D = 0.1 and d/D = 0.2 cases. On the

other hand, the water-PAA mixing is poor under laminar flow conditions for d/D = 0.1 and

d/D = 0.2 cases. From comparisons above, it is clear that Sc value is affecting the mixing

of two streams. With about three orders of magnitude difference in Sc value, the flow with

smaller Sc value (air-methane) has a much more thorough mixing.


48

d/D=0.2
0.2

d/D=0.1
0.1

Figure 12 Plot of degree of laminar mixing along the downstream length for the air-methane and

the water-PAA solution cases, d/D = 0.1 and 0.2, Rem = 200, Rej = 40

Turbulent mixing of gas-gas and liquid-liquid streams

Turbulent mixing of two gases

Forney et al. discussed the optimum mixing conditions of two gaseous flow in a T-

junction. In the experiments (Forney and Lee, 1982, Forney, Noureddine and Hanh,

1996), the two flow streams are air and methane-air mixture. The main flow Rem =

21,000, and the jet flow Rej varies from 4,000 to 8,000. Figure 13 below shows the

normalized concentration profile of methane (c/cm) along the center vertical line at the
49

axial location where z/D = 3 and R (jet-to-pipe inlet velocity ratio) = 11.0. The results

were time-averaged with last 0.02s (20 time-steps) in the simulation.

Figure 13 Comparison of methane concentration from the simulations and the experiments

(Forney and Lee, 1982, Forney, Noureddine and Hanh, 1996) at R = 11.0, d/D = 0.014, z/D = 3.0,

Rem = 21,000, Rej = 4000, Sct = 0.9

and R = 20, d/D = 0.014, z/D = 3.0, Rem = 21,000, Rej = 7700, Sct = 0.9

In Figure 13, the predictions agree well with the measured values reported in (Forney and

Lee, 1982, Forney, Noureddine and Hanh, 1996). However, slight deviations are

noticeable. This may be caused by a possible systematic error on locating sampling points
50

during the experiment; as well as numerical errors viz. mesh not fine enough, under-

prediction of turbulence intensity and therefore of turbulent mixing etc.

Figure 14 below shows the normalized concentration profile of methane along the center

horizontal line where R and d/D are 3.73 and 0.1 respectively. The results were time-

averaged with last 0.02s in the simulation, which were the average of 20 time steps.

Figure 14 Normalized maximum concentration of methane at different axial locations (compared

with experimental data)

with R = 3.73, d/D = 0.1, Rem = 21,000, Rej = 6000, Sct = 0.9

Both the plots in Figure 14 show good agreement of CFD predictions with the

experimental data (Forney and Lee, 1982, Forney, Noureddine and Hanh, 1996), which

indicate the fidelity of the current CFD grid generation together with the applied
51

numerical algorithm in resolving the mixing characteristics of two turbulent streams in

the T-junction pipe.

A series of contour plots (at different times with a 0.01 s interval) displaying the mixing

behavior of air-methane mixing (with R = 3.73 and d/D = 0.1) as predicted by the LES

are shown in Figures 15 (a) – (e). The boxed zones in the figures show the progression

and decay of a mixing structure as it flows downstream.

(a)

(b)

(c)
52

(d)

(e)
Figure 15 Contour plots of methane concentration at center vertical plane at different times d/D =

0.1, R = 3.73, Rem = 21,000, Rej = 4000, Sct = 0.9

(a) t =3.005s (b) t = 3.015s (c) t = 3.025s (d) t = 3.035s and (e) t = 3.045s;

From the simulations, it appears that the jet is unable to fully penetrate the main pipe

diameter, however, the large eddy structures of the methane jet can be identified

downstream. The boxed zones in Figures 15 (a) – (e) shows the progression and decay of

a mixing structure as it flows downstream. It is also interesting to note that the upper part

of the main pipe has little or no methane concentration in the downstream region.
53

Figure 16 shows the time-averaged contour plot of methane concentration from 3.600 sec

to 3.620 sec at the exit plane of the geometry given by Forney et al. (Forney and Lee,

1982, Forney, Noureddine and Hanh, 1996) . It is interesting to note that the degree of

mixing of methane at the exit plane is rather non-uniform and shows wide variations

(perhaps caused by the presence of eddy structures). Though the methane is injected from

the top region of the main pipe, the resulting flow field produces a methane rich region at

the bottom region of the main pipe at the exit which is at 15-pipe-diameter downstream

from the jet pipe. Recall Fig 1, the exit plane is marked in red on the schematic.

Figure 16 Contour plot of methane concentration at the exit plane, time average from 3.600 sec to

3.620 sec; R = 2.45 and d/D = 0.2, Rem = 21,000, Rej = 6000, Sct = 0.9

Turbulent mixing of two liquid streams

For turbulent mixing of two liquid streams, the simulations were carried out for the

geometry shown in Figure 8. Turbulent mixing of water-PAA solution were studied with
54

Rem = 21,000, and Rej = 6,000 with jet-to-pipe inlet velocity ratio R = 2.0, and d/D = 0.2.

For the turbulent mixing of water-sodium fluorescein solution in water, we carried out the

simulations following the parameter values as specified in the paper by Pan and Wang

(Pan and Meng, 2001), in which cases Rem = 20,850, and Rej = 10,630 respectively.

Turbulent mixing of Water/PAA solution in water

Figure 17 below shows the cm (maximum concentration of the solute) for liquid-liquid

(water – PAA/water) mixing in a T-junction under similar geometric conditions as for

laminar cases described earlier. For the results shown in Figure 12, R = 2.45, and d/D = 0.2

with Rem = 21000, Rej = 6000 Sct = 0.9, and R = 3.73, d/D = 0.1 with Rem = 21000, Rej =

4000 Sct = 0.9. The corresponding results for turbulent mixing of the air-methane case is

also shown in the Figure 17 by dash lines. Since the effect of the laminar diffusivity plays

a minor role in the turbulent mixing problems, the mixing characteristics for turbulent

water-PAA solution and air-methane cases are almost identical.


55

Figure 17 Comparison of liquid-liquid mixing and gas-gas mixing at R = 2.45, d/D = 0.2, Rem =

21000, Rej = 6000 Sct = 0.9 and R = 3.73, d/D = 0.1, Rem = 21000, Rej = 4000 Sct = 0.9

Recalling the laminar flow mixing results (for air-methane and water-PAA solution) shown

in Figures 11, the turbulent flow mixing results (gas-gas mixing vs. liquid-liquid mixing)

appear to be identical; however, this may be misleading as the maximum concentration

value cm may not be the best indicator for mixing effectiveness.

Figures 18 below show a series of contour plots (at different times with a 0.5 s interval)

displaying the water-PAA mixing (with R = 3.73 and d/D = 0.1). The series of plots are

given at 0.5 s interval. For the mixing of water and PAA solution, the main flow Rem =

21,000, and the jet flow Rej ~ 4,000, d/D = 0.1, and R =3.73. In Figure 15 (shown

earlier), for gas-gas turbulent mixing, the series of plots are given at 0.01 s interval. Even
56

though the Rem and Rej for the gas-gas and the liquid-liquid cases are the same, the U m

for the liquid-liquid case is about an order of magnitude smaller than in the gas-gas case

as the kinematic viscosity of air is about an order of magnitude higher than that of water.

The growth and evolution of the large eddies are less pronounced in the liquid-liquid

cases as shown below.

(a)

(b)

(c)
57

(d)

(e)

Figure 18 Contour plots of PAA solution concentration at center vertical plane at different times

(a) t =20.0 s (b) t =20.5 s (c) t = 21.0 s (d) t = 21.5 s and (e) t = 22.0 s;

d/D = 0.1, R = 3.73, Rem = 21,000, Rej = 4000, Sct = 0.9

Because the residence time of water-PAA solution mixing ( U m = 0.11 m/s) is much

larger than residence time of air-methane mixing ( U m =2.0 m/s), the liquid-liquid case

was computed for a larger time to ensure the flow-field reached quasi-steady state. Each

simulation is conducted from t = 0.00 sec to three times of the theoretical residence time

to reach the quasi-steady state. For example, the theoretical residence time of a water-

PAA solution mixing case is about 16.6 sec, then numerical calculation is carried up to

50.0 sec. Even though the Rem and Rej for the gas-gas and the liquid-liquid cases are the
58

same, the U m for the liquid-liquid case is about an order of magnitude smaller than the

gas-gas case as the kinematic viscosity of air is about an order of magnitude higher than

that of water.

Figure 19 shows the time-averaged contour plots of PAA concentration from 49.80 sec to

50.00 sec at the exit plane of the geometry for the turbulent mixing of water-PAA

solution. It is interesting to find that the degree of mixing of PAA solution at the exit

plane is again good only for the bottom half of the pipe cross-section. Comparing with

gas mixing results shown in Figure 16, the liquid mixing shows more uniform and less

affected by the eddy structures.

Figure 19 Contour plot of PAA concentration at the exit plane, with R = 3.73 and d/D = 0.1, time

average period: 49.80 sec to 50.00 sec, Rem = 21,000, Rej = 4000, Sct = 0.9
59

Water-sodium fluorescein solution mixing

Pan and Meng (Pan and Meng, 2001) conducted two sets of experimental measurements

on T-junction mixing in liquid-liquid flows. The main flow is water and the jet flow is a

sodium fluorescein solution in water. The Reynolds number of the main flow is fixed at

20,850 in both cases; Reynolds numbers of the jet flow are 10,630 and 17,500 (Pan and

Meng, 2001). Simulations were carried out for the case with main flow Reynold number

equals to 20,850 and the jet flow Reynolds number of 10,630.

Figures 20 (a) and (b) show the grayscale image of the instantaneous concentration

field for R = 3.06. The image covers a cross section at the symmetry plane in the cross-jet

region. The jet can be found entering from the top and turns to the right with the pipe flow.

The simulation results in Fig 20 (b) showed lack of structure compared to the experimental

measurements. The inflow side (left of the jet) did not show any sign of roll-up vortex

either. These differences might be caused by imperfect geometry lineup, or the residual

swirl in either the jet or the main flow, etc. The numerical errors contribute to the difference

as well, for example, LES filtered out small structures, which is one of the reason that

simulation predicts less diffusion. Besides, the difficulty on matching experimental time

and simulation time could be another factor as well. Though the simulation conditions were

carefully matched with the experimental conditions (Pan and Meng, 2001), the predicted

concentration field lacks some of the details found in the experimental work. The LES

filters out the sub-grid level fluctuations. The alternate here is to carry out the direct

numerical simulations (DNS) which can become prohibitively expensive in time and

resources. In addition, the simulations are for ideal conditions, assuming no upstream or

laboratory disturbances. Such effects are unavoidable in most experiments. These could
60

contribute to the mismatch of the finer details of the experimental and computational results

presented here.

(a) Instantaneous image (Pan and Meng, 2001)

(b) Present simulations


Figure 20 Comparison of grayscale images of the side-view instantaneous concentration field for

R = 3.06, d/D = 0.1667, Rem = 20,850, Rej = 10,630. Screenshot of simulation is taken at 15.00

sec
61

Figure 21 (a) shows the instantaneous two-dimensional velocity field measurements (Pan

and Meng, 2001) on the center plane of the T-junction (side view, covering the pipe

diameter). The prediction from CFD simulation in Fig 14 (b) shows some fluctuation,

however (in part due to averaging) it does not provide as many fine structures as found in

the experimental measurement. The darker part on Fig 14 (b) shows high velocity region,

which has similar pattern of fluctuations.

(a) Instantaneous image (Pan and Meng, 2001)


62

(b) Present simulations

Figure 21 Instantaneous velocity field on the center plane of the tee mixer, R = 3.06, d/D =

0.1667, Rem = 20,850, Rej = 10,630. Screenshot of simulation is taken at 15.00 sec

Figure 22 (a) below shows the measured (Pan and Meng, 2001) time-averaged tracer

concentration at the vertical center plane, where R = 3.06, Rem=20,850, Rej=10,630. The

contours from CFD simulation shown in Figure 22 (b) predict a similar distribution of

tracer concentration as found in the measurements. The simulation, however, predicts less

diffusion of the tracer compared to the experimental measurement. Some probable causes

might again be slight misalignment in the experimental setup, or the residual swirl in either

the jet or the main flow, numerical errors introduced by the differencing (spatial and

temporal) schemes used, etc.


63

(a) Measured data (Pan and Meng, 2001)

(b) Present simulations

Figure 22 Time-averaged tracer concentration at vertical center plane, where R = 3.06,

Rem=20,850, Rej=10,630. Time averaging period from the present CFD simulation: 22.0 sec to

22.2 sec
64

Finally, Figures 23 below show a series of contour plots (at different times with a 0.05 s

interval) displaying the water-sodium fluorescein solution mixing (with R = 3.06 and d/D

= 0.1667). For the mixing of water-sodium fluorescein solution, the main flow Rem =

20,850, and the jet flow Rej = 10,630, d/D = 0.1667, and R = 3.06, d/D = 0.1667. Even

though the Rem for the water-PAA solution and the water-sodium fluorescein solution

cases are nearly the same (Figure 18 vs. Figure 23), the Rej for two cases are significantly

different.

(a)

(b)

(c)
65

(d)

(e)

Figure 23 Contour plots of tracer (sodium fluorescein) concentration at center vertical plane at

different times (a) t =15.0 s (b) t =15.5 s (c) t = 16.0 s (d) t = 16.5 s and (e) t = 17.0 s; d/D =

0.1667, R = 3.06, Rem = 20,850, Rej = 10,740, Sct = 0.9

Comparing the results shown in Figure 18 (a) – (e), (Rej = 4,000), with the above results

in Figures 23 (a) – (e), (Rej = 10,750), it is evident that the downstream eddy structures

are more vigorous in the later cases resulting in better mixing.

4.6 Conclusions

The transport processes that are involved in the mixing of either two gases or two miscible

liquids in a T-junction mixer are investigated. The turbulent flow fields are calculated for

the T-junction with the LES model employing CFD-ACE+. In the mathematical model the
66

transport of species is described with a species concentration variable that is a function of

both space and time. The results obtained by numerical simulations are compared with two

well-defined experiments (Forney and Lee, 1982, Pan and Meng, 2001) for gas-gas and

liquid-liquid mixing respectively. Good mixing is obtained in a T-junction if the branch

inlet flow is designed to penetrate to the opposite pipe wall in the mixer.

The comparison between laminar air-methane and water-PAA solution cases showed that

with similar Rem and Rej, the air-methane case (with a smaller Sc) has improved mixing

beyond the T-junction, compared to the water-PAA solution case (with a higher Sc). LES

simulations for turbulent mixing showed differences between the mixing behavior of gas-

gas and liquid-liquid streams for similar geometrical and flow conditions. However, the

mixing of two turbulent flow streams (indicated by the maximum concentration value of

the solute cm) downstream of a T-junction are not affected by the Sc values even when Sc

for the gas and liquid-liquid streams vary by about three orders of magnitude. Cross-

sectional contour plots show that even with such high Reynolds number (as found in this

study), there is still a relative large concentration gradient at the pipe exit (~15 pipe

diameters). Additional mixing devices like baffle plates are used in the industry to improve

mixing. For the liquid-liquid cases, the comparisons of flow structures between CFD

predictions and experimental measurements (Pan and Meng, 2001) showed qualitative

agreement only.
67

V. THERMAL MIXING OF TWO LIQUIDS IN A T-

JUNCTION

5.1 Introduction

One of the most encountered mixing devices is a T-junction pipe. The T-junction containing

a transverse pipe and a 90º branch is widely used in the mixing of two fluids (in gaseous

or liquid form). The geometry can be utilized as a chemical reactor, combustion chamber,

piping system, etc. Gas mixture flows appear in diverse applications. Examples include

pollutant dispersion, chemical processing, and combustor mixing and reaction. Previous

discussion was carried out on the mixing of two gases (air and methane) with no heat

transfer considered in a T-junction (Forney and Lee, 1982). However, it is also common

that the two streams are having different temperature. For example, cool air flow is mixing

with a hot air flow in an HVAC unit (Hirota, Mohri, Asano and Goto, 2010). Liquid-liquid

mixing problems with temperature variations are also common in industry and in nature. A

previous study of liquid-liquid mixing problem was the mixing of water with water-sodium

solution in a T-junction (Pan and Meng, 2001). T-junctions are also used for the mixing of

wastewater and disinfectant in wastewater treatment process, the mixing of hot water with

cooling water in a power plant (Naik-Nimbalkar, Patwardhan, Banerjee, Padmakumar and

Vaidyanathan, 2010), etc. Therefore, the hot water mixing with cooling fluid is considered

as a candidate in liquid-liquid mixing problem in a T-junction.

The mixing of two flows in a T-junction mixing chamber contains mass and heat transport

processes, thus the process is very complicated. It is necessitated to entirely reveal the

detailed characteristic and the mechanism of the mixing of hot and cold flows in the T-
68

junction. Previous studies on T-Junction mixing showed the mixing behaviors and flow

structures between the main flow and the jet in detail (Wei, Farouk and Haas, 2017). With

the development of computing power, computational fluid dynamics (CFD) simulations

are applied in the present study to predict the flow properties and mixing characteristics of

two fluids (gas-gas or liquid-liquid) at a 90o T- junction of two pipes as well. As a balance

on computational cost and flow resolution, large eddy simulation (LES) is becoming one

of the most popular turbulent modeling technique in CFD. LES has been proven to be

accurate on both mass transfer and heat transfer processes.

In this study, CFD simulations of the mixing of two gaseous flows is carried out first for

validation purpose. In the gaseous mixing case, a cold air flow is introduced from the

main inlet, a stream of hot air is injected from the jet flow inlet. Main flow Reynolds

number is denoted by Rem, inlet jet Reynolds number is denoted by Rej. The mixing of

two streams is studied under turbulent flow conditions, where Rem ~ 15,000. After the

fidelity of the CFD model is validated, another series of simulations are focused on the

mixing of two water streams, i.e. cool water flow mixing with hot disinfectant solution

(0.3% water-peracetic acid mixture) to represent the scenario in a pilot scale wastewater

treatment reactor. In the liquid-liquid study, cold water flow is introduced from the main

pipe inlet and hot water is injected from the jet flow inlet. The mixing of the two liquid

flows are studied with a Reynold number of ~20,000.

5.2 Problem Description

CFD simulations are carried out in the present study to predict the flow properties and

mixing characteristics of two fluids at a 90o T- junction of two pipes of dissimilar diameters.

Previous works in the T-junction mixing problem were either focused on velocity ratio
69

(Sroka and Forney, 1989) or temperature difference (Hirota, Mohri, Asano and Goto,

2010). In this study, both species and temperature are considered in a T-junction. For the

turbulent flow field predictions, the LES technique, which has been proven to be a better

alternative to RANS, is used to predict the spatial-temporal evolution of the mixing

process. A schematic of the mixing problem considered in the T-junction is shown below.

The two flows are labeled as the main flow (in the larger pipe with diameter D) and the jet

flow (in the smaller pipe with diameter d). The main pipe with D = 7.62 cm is placed

horizontally along the x-axis. The jet pipe with smaller diameter is on the top and placed

vertically along the y-axis.

Figure 24 Schematic of 3-D T-junction geometry considered (not for scale)

The diameter of the inlet jet is 1.27 cm making the diameter ratio d/D = 1/6. The distance

from injecting point to the outlet is about 15 times of main pipe diameter to ensure the flow

is fully developed when it exits. The main inlet and jet velocity is marked by Um and Uj

respectively. The ratio between the two velocities is denoted by VR, where VR = UM / UJ.
70

Similarly, the temperatures for the two inlets are Tm and Tj, the ratio of the two is TR, where

TR = Tm/ Tj. Table 1 below lists out the major variables in this study.

Table 2 Listing of the main variables for the mixing problems

Item Symbol Value

Main pipe Reynolds number Rem 15,000 for validation


~20,850
Inlet jet Reynolds number Rej 7,500 for validation
~10,630
Jet-to-pipe diameter ratio d/D 1/6

Jet-to-pipe inlet velocity VR 2;


ratio 3
Jet-to-pipe inlet temperature TR 1.09;
ratio 1.28

5.3 Methodology

Mathematical Model

For simulating turbulent flows in the T-junction, we need to invoke a turbulence model.

We have used the Large-eddy simulation (LES) model to simulate the turbulent mixing

problems in the T-junction. LES solves a “filtered” set of Navier-Stokes equation (Yang,

2014). Filtering of LES is a mathematical operation intended to remove a range of small

scales from the solution to the Navier-Stokes equations. Because the principal difficulty in

simulating turbulent flows comes from the wide range of length and time scales, this

operation makes turbulent flow simulation cheaper by reducing the range of scales that

must be resolved. The LES filter operation is low-pass, meaning it filters out the scales

associated with high frequencies. A LES filter can be applied to a spatial and temporal field

and the filter size is decided by maximum grid size and time step. Any spatial fluctuation

smaller than grid size will be filtered out, and similarly, any temporal frequency higher
71

than time step frequency will be filtered out as well. The operation of “filtering” applied to

the flow is a convolution with a spatial filter over entire domain, which is,

f ( x, t ) = ∫ f (ξ , t ) G ( x − ξ , ∆ ) d ξ (5.1)
D

where f is filtered value of a field variable f, ∆ is the filter width and G( x − ξ , ∆)dξ is

the filter function.

In the simulation,

 1 if x − ξ ≤ ∆
G( x − ξ , ∆) =  ∆ 2 (5.2)
 0 otherwise

With this operation, variables such as velocity u were divided into large scale which is

denoted by an overbar and sub grid scale (SGS) denoted by a prime, which is,

u = u +u' (5.3)

Therefore, the filtered governing equations for a Newtonian incompressible flow become

as following:

Filtered conservation of mass

∂ρ ∂
+ ( ρ ui ) = 0 (5.4)
∂t ∂xi

Filtered conservation of momentum

∂ui ∂ 1 ∂p ∂u ∂u j
+ (u i u j ) = − +ν ( i + ) + Si (5.5)
∂t ∂x j ρ ∂ xi ∂ x j ∂ xi
72

whereν is kinematic viscosity, Si is body force, here represents gravitational force.

Furthermore, the filtered advection term can be split up, following Leonard (Leonard,

1974),

uiu j = τijr + ui u j (5.6)

where τ ijr is residual stress tensor, then the filtered momentum equations become

(neglecting body force),

∂ ui ∂ 1 ∂p ∂ ∂τ ijr
+ (u i u j ) = − + 2ν S ij − (5.7)
∂t ∂x j ρ ∂xi ∂x j ∂x j

where S ij is the rate-of-strain tensor, and,

1 ∂ ui ∂ u j
S ij = ( + ) (5.8)
2 ∂ x j ∂ xi

Applying the Smagorinsky model (Smagorinsky, 1963) as sub-grid scale model,

1
τ ijr − τ iiδ ij = 2ν t S ij (5.9)
3

where ν t is turbulent eddy viscosity.

A relationship between ν t and the filter size Δ can be found as,

ν t ~ ∆2 2S ij S ij (5.10)

The Smagorinsky model gives,

ν t = cs2∆2 2S ij S ij (5.11)

where cs is called Smagorinsky constant, cs = 0.1 has been chosen in this simulation.
73

Filtered conservation of energy

∂ ( ρ h) ∂ ∂ ∂T
+ ( ρ hu j ) = ( k eff ) (5.12)
∂t ∂x j ∂x j ∂x j

where keff is effective coefficient that includes turbulent mixing contribution in addition to

molecular conduction. Hence,

µ t Cp
k eff = k + (5.13)
Prt

where k is thermal conductivity, Cp is specific heat, Prt is turbulent Prandtl number.

Recommended value of Prt is 0.9 according to CFD-ACE+(ESI-Group, 2017).

Meshing, Numerical Scheme and Boundary Conditions

The grid meshing is generated in the CFD-GEOM. To balance the computational cost and

the resolution of flow characteristics, a grid with 900,000 structured cells is addressed to

this model. In order to capture the small structure at near field, more condensed meshing

is assigned to the region (see Fig. 2(b)). Then the length scale of the grid is gradually

increased to the exit.


74

(a)

(b)

Figure 25 Grid Generation of the T-junction (a) prospective view; (b) Cross-sectional view of the

T-junction
75

When generating the meshing grid, as suggested by Zang (Zang, 1991), near-wall

configuration should have the first node being located between 0 < y+< 1, and more than

three nodes should be placed with 0 < y+ < 10 at the same time, where y+ stands for the

normalized wall coordinate derived by equation (14).

τw
y
ρ
y+ = (5.14)
υ

where τ w is the wall shear.

The plot in Figure 26 shows the y+ at the top wall of the main pipe along after the jet

nozzle, where y = 0.0381 m. It is noticeable that in this model, each first node near the

top wall is below 1.0, which follows the suggestion from Zang(Zang, 1991).
76

Figure 26 y+ at the top wall of the main pipe after the jet nozzle at y = 0.0381m.

The simulations are carried out in transient mode with a time step equal to 10-3 sec. A

second order blended upwind scheme is applied for the convective-diffusive terms and

Crank-Nicolson scheme with a blending factor of 0.7 is used for the temporal terms in the

governing equations. The internal relaxations for velocities and enthalpy are 0.2. The

convergence criterion was set to 0.001.

At inlets of the domain, normal uniform velocities are specified. No-slip conditions are

assigned for the walls. Initially, the main pipe is filled up with cold water, while the jet

inlet is specified with hot flow (mixture of water with tracer). At t = 0.0 s, hot flow is

introduced from jet inlet and cold flow is injected from main inlet as well. The simulation

is carried out for three times of the theoretical residence time to diminish the effect of
77

initial conditions and ensure the flows becomes quasi-steady(Wei, Farouk and Haas,

2016). Simulations of this study are operated in the CFD-ACE+.

5.4 Results and Discussions

Validation

Previous experimental work on a T-junction geometry was carried out by Hirota et al.,

(Hirota, Mohri, Asano and Goto, 2010). In their experiments, the temperature of cold and

hot air flow is 12ºC and 60ºC respectively. By matching the experimental parameters, a

CFD simulation is carried out first for validation.

(a)
78

(b)

Figure 27 Visualized image of the interface between two flows in Z/A = 0 from (a) Experiment;

(b) CFD simulation result

As shown in Figure 27, the interface of hot and cold fluid predicted by CFD simulation

matches the experimental measurement. Because of the nature of LES, small structures

were filtered out. This is the reason that CFD didn’t provide enough information on those

small fluctuations. Furthermore, mean velocity vector plot shows agreement between the

CFD simulation and experimental result as well. The interface line in CFD is about the

same location as in experiments. Besides, CFD simulation also predicts the recirculation

area around the same location.


79

(a)

(b)

Figure 28 Mean velocity vector diagram in the symmetric plane of Z/A = 0. (The solid lines show

the time-averaged interface between the main and branch flows obtained by the flow

visualization.) (a) Experiment; (b) CFD simulation result

Figures 29 below are comparing the time averaged temperature profile at various

locations on the center plane between CFD simulation predictions and experimental

measurements. The time averaged temperature profile plots show that the prediction from

CFD simulation is agreeing with experiments fairly well. These comparisons from

Figures 27 to 29 validate the CFD model involved in this study.


80

(a)

(b)
81

(c)

Figure 29 Profiles of the mean normalized temperature at (a) x = 0; (b) x = 1.0; (c) x = 3.0

CFD Results of Thermal Mixing

Characteristics of the Temperature Field

In Figures 30 (a) to (e) below, a series of contour plots displaying the characteristics of

the mixing of cold water with hot water-PAA solution (with TR = 1.28 and VR = 3.0) as

predicted by the LES are shown in Figures 30 (a) – (e).


82

(a)

(b)

(c)

(d)
83

(e)

Figure 30 Temperature contour at center plane at (a) t = 15.00 sec; (b) t = 15.30 sec; (c) t = 15.60

sec; (d) t = 15.90 sec; (e) t = 16.20 sec; at TR = 1.28, VR = 3.0.

From the simulations, it appears that the jet is unable to fully penetrate the main pipe

diameter at the near field. The large eddy structures of the hot liquid jet can be identified

downstream. The boxed zones in Figures 30 (a) – (e) shows the progression and decay of

a mixing structure as it flows downstream. It is noticeable that the upper part of the main

pipe has little heat transfer until 7-8 main pipe diameters downstream the nozzle. On the

other hand, the jet flow takes about 4 pipe diameters to reach the bottom.

When looking at the exit plane, which locates at about 15 main pipe-diameter

downstream after the jet nozzle, the time averaged temperature contour in Figure 31

shows rather non-uniform and shows wide variations of 6 degree Kelvin from lowest

286K to a highest of 292K.


84

Figure 31 Time-averaged temperature contour plot at exit plane (15 pipe diameter downstream).

TR = 1.28, VR = 3.0.

Another series of CFD simulation with TR = 1.09 and VR = 3.0 shows the comparable

results. By observation, the mixing is not through at the upper half of the pipe neither.

However, the jet flow reaches the bottom of the pipe at about 2 pipe diameter which is

considerably earlier than the case with TR = 1.28.


85

(a)

(b)

(c)

(d)
86

(e)

Figure 32 Temperature contour at center plane at (a) t = 15.00 sec; (b) t = 15.30 sec; (c) t = 15.60

sec; (d) t = 15.90 sec; (e) t = 16.20 sec; at TR = 1.09, VR = 3.0.

Figure 33 Time-averaged temperature contour plot at exit plane (15 pipe diameter downstream).

TR = 1.09, VR = 3.0.
87

The plot in Figure 32 shows the time-averaged temperature contour plot with TR = 1.09,

VR = 3.0, which has a similar pattern as in Figure 30. It is observed that the hot flow is

accumulating at the second half of the pipe at the exit, despite slightly smaller density.

Comparing with TR = 1.28, the mixing of TR = 1.09 shows more uniform and less

affected by the eddy structures (Figure 33).

Comparisons of the Concentration Field

In order to describe the thoroughness of the thermal mixing, a derived variable is

introduced: Φ, degree of mixing, which can be calculated by:

std (ci )
φ= (5.15)
Mean(ci )

where std(ci) represents the standard deviation of the concentration at certain cross-

sectional plane, Mean(ci) is the average concentration at the cross-sectional plane.

As shown in Figure 34 (a) and (b), with same velocity ratio, the simulation with higher

temperature difference has a slightly smaller degree of mixing, which indicates a better

mix happening downstream.


88

(a)

(b)

Figure 34 Comparisons of the degree of mixing between TR = 1.28 and TR = 1.09, where (a) VR

= 3.0; (b) VR = 2.0


89

5.5 Conclusions

In this study, a series of numerical simulation of turbulent mixing between a low

temperature main flow and a hot vertical jet in a T-junction were carried out. LES with

sub-grid scale modeling has shown to reveal the characteristics of the mixing process

successfully. CFD simulations show the effects due to density difference cannot be

overlooked in the study where temperature difference exists. The CFD predictions for

validation showed that the agreement with previously published experimental

measurement is good regarding both the temperature field and the velocity field.

Further simulations show that the shear layer formed in the near field after jet injection is

essential for the thermal mixing. With the increase of the temperature ratio between the

two flows, the gradient of the shear increases correspondingly, which promotes the

thermal mixing furthermore.


90

VI. CFD SIMULATIONS ON PILOT SCALE

WASTEWATER TREATMENT REACTOR

WITH MIXING ENHANCEMENT

6.1 Problem Description

In wastewater treatment industry, to study the characteristics of disinfectants are usually

relied on physical experiments. Studies on pilot-scale reactors reveal the chemical

kinetics and hydraulics of the wastewater treatment processes. With geometric scaling-

up, the performance of such disinfectant in a full-scale contactor can be predicted. The

pilot-scale serpentine reactor shown in Fig. 35 was built by PeroxyChem, LLC. This

serpentine reactor was designed to be transported among different wastewater treatment

facilities for studying the kinetics of disinfecting local wastewater with PAA.

Figure 35 The Pilot-scale Serpentine Reactor. (Photo courtesy of PeroxyChem)


91

On this pilot-scale reactor, the inlet pipe is a 90o T-junction, shown in Fig 36. The

transvers pipe with 5.08-centimeter-diameter pipe is filled with wastewater. A vertical

0.635-centimeter-diameter injector which injecting PAA into the center of the main pipe

at a very slow speed (~0.001m/s) located in the middle. Total length of the T-junction

pipe is 5 meters.

Figure 36 Zoomed-in Inlet pipe.

6.2 3-D CFD Model of the Pilot Scale Reactor

Based on the geometry provided, a 3-D CFD geometric model with 600,000 Cartesian

cells first was created as shown in Fig 37 (a) and (b). Square shaped pipe addressed in

this model to approximate circular pipe could reduce computational cost considerably.

The time-dependent simulations were carried out with a time step equal to 0.1 sec. A

second order blended upwind scheme was applied for the convective-diffusive terms and
92

Crank-Nicolson scheme with a blending factor of 0.7 was used for the temporal terms in

the governing equations. The internal relaxations for velocities and passive scalar are 0.2.

The convergence criterion was set to 0.001. Main pipe inlet had a mixture of water and

microbial with a velocity of 0.77 m/s. The injection pipe has 10 mg/L PAA with a

velocity of 0.003 m/s.

(a)

(b)

Figure 37 (a) Geometric model of 3-D CFD simulations;

(b) Zoomed-in view of near field of the injector. The serpentine contactor was a 12-inch

diameter PVC pipe with 3 180° turns. Four channels formed a 90 ft long contact pass to

retain the contact time. Each turn has two 90° bends and one 19-inch straight section, in

this way, it could be fitted on a trailer truck to relocate conveniently. Over all dimensions

of the device was 22×4×4 ft. Meshing for the serpentine is shown in Fig 38.
93

Figure 38 Grid generation of serpentine contactor

Because of the difference in diameter, the flow pass from inlet T-junction would be

expanded. A ¼-inch thick baffle plate placed 4 inches from the expansion. The baffle

plate contained 72 0.5-inch-diameter holes, shown in Fig 39(a). To simplify the model,

the circular holes were accommodated to 15 1-inch-long square shaped structure. The

porosity was kept the same, as in Fig 39 (b).

(a) (b)

Figure 39 (a) mechanical diamgram of baffle plate. (b) CFD geometric model of baffle plate
94

6.3 Results and Discussions

After simulating three residence times, the system became quasi-steady. Contour plots of

PAA concentration on center plane and exit cross-section showed the structure of the

mixing in pipeline. From the plots, it was obvious that within a 5-meter pipeline, PAA

was not mixing well with water. Further mixing enhancement method such as static

mixer was required for this process.

(a)

(b)

Figure 40 Contour plots of PAA concentration at 5.0 sec simulating time at (a) center plane; (b)

exit
95

Same numerical scheme was carried out in this simulation. Fig 40 showed the PAA

structure in the first channel at 9.0 sec simulating time.

Figure 41 Instanteous contour plot of PAA in the first channel of the serpentine contactor at 9.0

sec.

6.4 Mixing Enhancement with Inlet Variations

6.4.1 Problem Description

For mixing enhancement study with different inlet configurations, three types of jet inlet

are proposed, they are: (a) inserted straightly to the middle of the main pipe (b) inserted

with 90-degree turn downstream; (c) inserted with 90-degree turn upstream as in Figure

42 below:

(a)
96

(b)

(c)

Figure 42 Three jet inlet configurations. (a) Inserted straightly (b) inserted with 90-degree

turn downstream; (c) inserted with 90-degree turn upstream

Boundary conditions of the simulation are listed in the Table below.


Table 3 Listing of the main variables for the mixing enhancement study

Item Symbol Value

Main pipe Reynolds number Rem ~20,000


Inlet jet Reynolds number Rej ~10,000
Jet-to-pipe diameter ratio d/D 1/6
Jet-to-pipe inlet velocity R (= 3.0
ratio U j / Um )

6.4.2 Results and Discussions

Figures 43 and 44 below show contour plots of the tracer concentration at center vertical

plane and cross-sectional planes respectively.


97

(a)

(b)

(c)

Figure 43 Contour plots of the tracer concentration at center vertical plane (a) Inserted

straightly (b) inserted with 90-degree turn downstream; (c) inserted with 90-degree turn

upstream
98

(a)

(b)

(c)

Figure 44 Contour plots of tracer concentration at cross-sectional planes (a) Inserted

straightly (b) inserted with 90-degree turn downstream; (c) inserted with 90-degree turn

upstream
99

From both comparisons, the model with a 90-degree turn upstream jet inlet has the best

mixing performance among the three. The jet flow in the T-junction with a vertical jet

inlet is injected to the bottom. Carrying a large momentum with high velocity, it hits on

the bottom and reflected upward. In the case with 90-degree turn to downstream, on the

other hand, the system is lack of turbulence, which results from the interaction of main

flow and jet flow, so that the mixing performed most poorly. When examining the outlet

contour plots, the difference is even obvious, shown in Figure 45.

(a)

(b)
100

(c)

Figure 45 Contour plots of tracer concentration at the outlet (a) inserted vertically; (b)

inserted with a 90-degree turn downstream; (c) inserted with a 90-degree turn upstream

Following the study of T-junction mixing, as an indicator of mixing thoroughness, the

plot of degree of mixing is derived. Shown in Figure 46.


101

Figure 46 The plot of the degree of mixing

As shown in this plot, after 10-pipe diameter downstream, the degree of mixing in the

case with 90-degree turn upstream is much smaller than the other two simulations.

Pressure head is another important characteristics. It relates to operation cost directly. A

large pressure head requires a larger capacity pump, and a larger energy consumption.

Figures 47 below are the pressure head plots.


102

(a)

(b)

(c)

Figure 47 The plot of pressure head (a) inserted vertically; (b) inserted with a 90-degree

turn downstream; (c) inserted with a 90-degree turn upstream

The pressure head of each case is 914.4 Pa, 539.5 Pa and 1,829 Pa respectively. The

pressure head of the case with the best mixing results appears to be the largest. This is

because the energy is dissipated due to the intense turbulence and so that it requires the

pressure head.
103

6.5 Conclusions

In this study, we showed the simulation of a pilot scale wastewater treatment reactor.

However, the mixing of wastewater with disinfectant (PAA solution) did not well-mixed

condition even after 20-pipe-diameter in the main pipe. Therefore, a study of mixing

enhancement with different inlet configuration was carried out. It was shown that the one

with a 90-degree turn upstream provided the best performance of mixing. However, it

requires the largest pressure head at the same time.


104

VII. DEVELOPMENT OF ANN BASED

METAMODEL ON WASTEWATER

TREATMENT PROCESSES WITH PAA

7.1 Introduction

The design of chemical disinfection systems requires the balancing of multiple

constraints such as adequate inactivation of microorganisms and achievement of low

effluent residuals (in wastewater) to minimize ecological toxicity (Blatchley, Hunt,

Duggirala, Thompson, Zhao and Halaby, 1997) or (in drinking water) minimize the

formation of disinfection byproducts (Craun, Bull and Symons, 1994). This inactivation

must be achieved at a reasonable cost, and also must be capable of meeting the objective

(minimize toxicity) given the variability of water or wastewater quality and disinfection

kinetics.

While there is a growing literature to show that computational fluid dynamics (CFD) can

be used to describe inactivation kinetics and disinfectant chemistry in treatment

systems(Zhang, Tejada-Martínez and Zhang, 2014, Greene, Farouk and Haas, 2004),

these models frequently require hours of computer time to run for a particular

configuration. These speeds are inadequate to serve as an engine for the optimal

probabilistic design of disinfection processes.

In prior works, it has been shown in other applications that a metamodel could be used to

approximate CFD results over a range of physical and design variables(Hoque, Farouk

and Haas, 2011, Hoque, Farouk and Haas, 2011, Hoque, Farouk and Haas, 2010).
105

Metamodels are simple approximations to the results of a complex computational model

and have gained interest over the past two decades(Barton, 1994). The objective of this

work was to develop a proof of concept for the use of artificial neural network based

metamodels to approximate the CFD model predictions for microbial inactivation in a

wastewater disinfection contactor where peracetic acid was used as the disinfectant.

7.2 Methodology

The development of the present metamodel requires following five steps: (1) identify 11

input variables, the 11-D performance space and 2 outputs, i.e. PAA residual

concentration and the number of survival microorganisms; (2) determine the range of

each variable in interest; (3) generate the dataset by using a quasi-random sampling

technique to cover most scenarios in the design and operational space; (4) conduct

numerical experiments, i.e. 3-D CFD simulations configured with each set of parameters;

and finally (5) train the metamodel with numerical results and determine a configuration

of the metamodel to best serve the purpose.

Physical Model of Wastewater Treatment Reactor

As shown in Figures 48 (a) and (b), a full-scale wastewater treatment reactor is modeled

as an open rectangular tank with dimensions of length L, width W and depth D. An inlet

pipe of wastewater is placed horizontally near the bottom and an inlet tube for PAA is

located on top, dispensing disinfectant to the top surface. A weir at certain height locates

at the outlet wall to maintain a certain depth of the flow. Different number of baffles

forming a serpentine control the residence time (θ). Here b is to stand for the number of

baffles. The baffle slit width (bx) is described by the ratio with total length (L).
106

Moreover, there is a sharp edge weir designed near the entrance to form a mixing

chamber and therefore to enhance the mixing of wastewater with disinfectant.

(a)

(b)

Figure 48 (a): top view; (b) prospective view of a full-scale reactor.


107

Chemical Reactions in Wastewater Disinfection Process

To simplify the chemical kinetics, two reactions are mainly considered in the disinfection

process. They are: the reaction of PAA decay and the inactivation of microorganisms (N),

where N* represented inactivated microorganisms:

PAA → PAA* (4.1)

PAA + N → PAA + N* (4.2)

A recent batch study(Tizzoni, 2016) shows that PAA decays via immediate demand plus

first order decay in wastewater, and the inactivation of microorganisms with PAA

follows Hom’s power law(Hom, 1972). Therefore, the chemical kinetics may be written

as:

d [PAA]
= − k d [PAA] (4.3)
dt

d[ N ]
= − mk * [PAA]n t m −1[ N ] (4.4)
dt

where kd is the decay rate constant of PAA, k* is the Hom’s law rate constant, m and n

are Hom’s law exponents.

Since an implicit form of equation (4) is needed to apply in CFD simulation, the Hom’s

power law can be transformed to an equivalent form only in state variables(Haas and

Joffe, 1994):

d[N] 1 n [N] (1− 1m)


= −mk * m [ PAA] m [− ln( )] (4.5)
dt [N]0

Determination of the 11-D performance space


108

As demonstrated above, a model of the disinfection process in a full-scale reactor

contains both physical and chemical parameters, which divided into three types:

hydraulic characteristics (flow rate, residence time), reactor dimensions and PAA decay

and inactivation kinetics of microorganisms. Considering variables in each part, 12

dimensional variables are selected. As a function of the variables discussed in last

section, both PAA residual concentration and the number of microorganisms can be

written as:

[N ]
f (log( )) = f (Q,θ , L,W , D, b, bx, kd , k * , m, n,[PAA]0 ) (4.6)
[ N ]0

f ([PAA]) = f (Q , θ , L , W , D , b, bx , k d , k * , m , n,[PAA]0 ) (4.7)

Converting the variables of reactor geometry to dimensionless variables, and substituting

D by expression:

D=
( Qgθ )
( LgW ) (4.8)

The two target variables therefore can be described as,

[N]
f (log( )) = f (Q,θ , L ,W , b, bx, kd , k * , m, n,[PAA]0 ) (4.9)
[ N ]0 W (b + 1) D

f ([PAA]) = f (Q , θ , L ,W , b , bx , k d , k * , m , n ,[PAA]0 ) (4.10)


W (b + 1) D

All variables involved in this research are given in Table 4 below.


109

Table 4 List of variables

No. Variable Symbol Unit


1 Flow rate Q MGD (million-gallon per day)
2 Residence time θ min
3 Length L m
4 Width W m
5 Depth D m
6 Number of baffle b count
7 Baffle slit width bx Dimensionless (% of L)
8 PAA decay rate kd min-1
9 Hom law constant k* min-m (mg/L)-n
10 Hom law constant m dimensionless
11 Hom law constant n dimensionless
12 PAA inlet concentration [PAA]0 mg/L

A wastewater treatment contactor normally processes flows in the multi-million-gallon

per day (MGD) range. The range of the flow rate Q in this study is set from 10 to 100

MGD. The hydraulic residence time of a contactor varies between 15 and 75 mins. On

the other hand, the geometry of a contactor can vary from a square to a long narrow

rectangle (from top view). Therefore, the ratio of l/w is limited within the range of 1.0 to

12.0; while w/(b+1)/d is limited within between 0.8 to 2.0. The number of baffle varies

from 1 to 10. The baffle slit width varies from 5% to 15% of the length of the reactor.

Typically, the applied PAA dose has a range of 0.5-10 mg/L. The range of the

computational space for the present study is shown in Table 5.


110

Table 5 The range of the 11 variables

No. Variable min max


1 Q 10 100
2 θ 15 75
3 L/W 1.0 12.0
4 W/(b+1)/D 0.8 2.0
5 b 1 10
6 bx 5% 15%
7 [PAA] 0 0.5 10
8 kd 0.011 0.045
9 k* 1.177 3.401
10 m 0.235 0.701
11 n 0.328 0.820

For numerical experiments, space-filling designs play a significant role. The quality of

the space-filling affects the performance of the developed model. Four types of space-

filling designs have been used more often in the literature: orthogonal arrays, Latin

hypercube designs, Hammersley sequences, and uniform designs(Wang and Shan, 2007).

Hammersley sequence sampling (HSS) has been found to provide better uniformity than

other technique for a multi-dimensional performance space (Anders and Korn). It is

considered as a low-discrepancy sequence. The idea of discrepancy is applied to choose

the appropriate quasi-Monte Carlo sequence, of which the Hammersley sequence is an

example. 40 data points are generated for each variable by the HSS. Detail of the

generated dataset is shown in Table 6. This table is derived in MATLAB. Furthermore,

flow in the reactor always maintains a minimum scour velocity to ensure microorganisms

mixing well with disinfectant. The velocity can be calculated by

Q
V= (4.11)
W
Dg
b +1
111

Any dataset with scour velocity smaller than 0.3 m/s is discarded and replaced by a

resampled dataset.

Table 6 Sample space generated by HSS

Q θ l/w w/(b+1)/d b bx [PAA]o k k* m n


20 36 8.8 0.8 9 14% 5.5 0.040 2.059 0.594 0.407
74 61 9.3 1.7 6 9% 3.8 0.041 2.506 0.693 0.541
85 57 8.1 1.6 5 8% 9.7 0.027 1.871 0.609 0.465
30 26 10.4 0.8 10 9% 7.2 0.029 1.774 0.435 0.386
81 32 2.5 1.0 8 10% 5.6 0.021 1.611 0.351 0.450
64 19 8.6 1.8 9 8% 9.0 0.042 2.999 0.308 0.762
70 71 6.3 1.6 8 5% 4.8 0.015 1.728 0.601 0.610
81 67 5.2 1.5 7 14% 6.9 0.034 3.271 0.516 0.535
84 50 11.5 1.1 5 9% 2.5 0.012 2.908 0.639 0.721
39 48 10.9 1.0 5 9% 7.8 0.044 2.591 0.597 0.683
87 17 2.2 1.5 9 10% 8.1 0.040 2.182 0.424 0.605
20 71 11.5 1.3 8 8% 7.1 0.020 1.229 0.297 0.491
93 69 10.9 1.2 8 7% 3.9 0.013 3.090 0.254 0.453
48 67 10.3 1.2 8 7% 0.7 0.039 2.772 0.674 0.415
70 65 9.7 1.1 7 6% 9.8 0.032 2.454 0.632 0.377
25 63 9.2 1.1 7 6% 6.6 0.025 2.137 0.589 0.340
82 61 8.6 1.0 7 5% 3.4 0.018 1.819 0.547 0.791
37 59 8.0 1.0 7 14% 8.7 0.012 1.501 0.505 0.753
96 52 6.3 0.8 6 13% 7.7 0.030 3.038 0.377 0.640
52 17 7.4 1.1 10 13% 1.7 0.013 1.994 0.582 0.485
75 15 6.8 1.1 9 12% 7.0 0.038 1.676 0.539 0.447
35 52 11.4 1.7 6 6% 1.2 0.035 2.539 0.535 0.520
58 50 10.8 1.7 6 5% 9.3 0.029 2.221 0.493 0.482
94 46 9.7 1.6 5 14% 3.0 0.015 1.585 0.408 0.407
49 44 9.1 1.5 5 13% 8.2 0.041 1.268 0.366 0.369
72 42 8.5 1.5 5 13% 5.1 0.034 3.129 0.324 0.331
47 73 6.9 1.7 8 6% 7.9 0.021 2.046 0.643 0.648
92 15 7.5 1.7 8 6% 2.7 0.028 2.364 0.686 0.686
31 19 2.8 1.5 9 10% 2.8 0.014 2.500 0.466 0.642
86 71 5.7 1.0 9 11% 0.6 0.025 3.219 0.455 0.372
26 15 2.3 0.8 8 13% 9.1 0.023 2.552 0.528 0.805
94 21 4.1 1.0 9 5% 7.3 0.043 1.326 0.655 0.430
34 16 4.4 1.5 4 6% 7.7 0.015 2.435 0.421 0.506
21 24 4.7 1.0 9 6% 2.1 0.018 1.644 0.697 0.468
68 32 11.5 1.5 3 7% 4.1 0.028 2.039 0.416 0.750
64 32 1.8 1.6 9 5% 1.5 0.032 3.103 0.667 0.738
76 21 3.4 1.6 10 11% 6.0 0.021 2.817 0.508 0.680
14 26 11.7 0.8 7 8% 4.6 0.034 2.837 0.686 0.337
42 15 1.7 1.4 9 9% 5.0 0.033 1.864 0.381 0.567
83 38 7.4 1.3 4 11% 7.2 0.021 2.493 0.239 0.744

7.3 Development of 3-D CFD Model

Fluid flow modeling

Since water flow can be treated as incompressible flow, the continuity equation can be

written as:
112

∂u ∂v ∂w
+ + =0 (4.12)
∂x ∂y ∂z

Besides. the equation of conservation of momentum can be written as:

∂u ∂u ∂u ∂u 1 ∂p
+u +v +w =− + ν∇ 2u + f x
∂t ∂x ∂y ∂z ρ ∂x
∂v ∂v ∂v ∂v 1 ∂p
+u +v +w = − +ν∇ 2 v + f y (4.13)
∂t ∂x ∂y ∂z ρ ∂y
∂w ∂w ∂w ∂w 1 ∂p
+u +v +w =− + ν∇ 2 w + f z
∂t ∂x ∂y ∂z ρ ∂z

Turbulence Modeling

Typical Reynolds number in a CFD simulation in the study is around 30,000~100,000,

which is in turbulent flow region for an internal flow. Because of the random fluctuation

in turbulence, it is more difficult to model turbulent region. Therefore, a turbulence

modeling will help to reduce the difficulty of predicting the turbulent region. There are

three most popular turbulence models: Reynolds-averaged Navier–Stokes (RANS)

equations, Large-eddy Simulation (LES) and Direct Numerical Simulation (DNS). k-ϵ

model, which is one of RANS models was applied in this study.

There are two transported variables in k-ϵ model. The first variable determines the energy

in the turbulence and is called turbulent kinetic energy (k). The second variable is the

turbulent dissipation (ε) which describes how much the turbulent kinetic energy will

dissipate to the surrounding fluid. The idea of k-ϵ model is Reynolds decomposition,

which mathematically separates the time-averaged and fluctuating parts of the flow. The

flow velocity ui is to be decomposed as:

ui = ui + ui ' (4.14)
113

'
where ui is the averaged laminar velocity, ui is turbulent fluctuation.

Turbulent kinetic energy in per unit mass, k, was defined as:

1 2 2 2
k= (u + v + w ) (4.15)
2

Chemical species transport model

Disinfection is a process to achieve a significant reduction of the number of

microorganisms in wastewater. Hence, there are two major reactions in this study, which

are the decay of disinfectant and the inactivation of microorganisms.

Both of the two reactants in equation (1) and (2) can be predicted by solving chemical

species transport equation.

∂[PAA] ∂[PAA] ∂[PAA] ∂[PAA] ∂ 2 [PAA] ∂ 2 [PAA] ∂ 2 [PAA]


+u +v +w = DPAA−Water { + + } + SPAA
∂t ∂x ∂y ∂z ∂x 2 ∂y 2 ∂z 2

(4.16)

∂[ N ] ∂[ N ] ∂[ N ] ∂[ N ] ∂ 2 [N] ∂ 2 [ N ] ∂ 2 [ N ]
+u +v +w = DN −Water { 2 + + } + SN (4.17)
∂t ∂x ∂y ∂z ∂x ∂y 2 ∂z 2

where Γ is diffusivity, S is source/sink term.

Since microorganisms have a negligible diffusivity in a dilute flow, Equation (17)

becomes:

∂[MO] ∂[MO] ∂[MO] ∂[MO]


+u +v +w = SMO (4.18)
∂t ∂x ∂y ∂z
114

The chemically reactive time-dependent three-dimensional turbulent flow model was

solved using the CFD-ACE+.

7.4 Results and discussions

One of the advantages of CFD in post-processing is that it is very easy to visualize the

data. In the first run as the based model, the values of each variable were chosen as

shown in Table 7.

Table 7 The values of variables in the base model

Variable Value
Q 30 MGD
θ 30 min
L/W 2.75
W/(b+1)/D 0.667
b 9
bx 10%
[PAA] 0 10 mg/L
kd 0.028/min
k* 2.289 min-m (mg/L)-n
m 0.468
n 0.574

Figure 49 (a) shows a contour plot of PAA concentration on the middle horizontal plane.

The contour plot of MO survival ratio on the same plane is shown in Figure 49 (b).
115

(a)

(b)

Figure 49 (a) Contour plot of PAA Residual concentration; (b) Contour plot of N survival ratio at

center horizontal plane


116

As shown on Figure 49(a), about 60% of PAA has been consumed during the process

accompanied by a -1.6-logarithm microbial inactivation at the exit of the reactor which is

shown in Figure 49(b).

To ensure the flow in the time-dependent simulation has reached the quasi-steady state,

the survival ratio at the exit is tracked during the entire simulation. Plot of survival ratio

is shown as Figure 50.

Figure 50 Plot of CFD predicted N survival ratio at the exit

As shown on Figure 50, the plot of N survival ratio has a dramatic decrease in the second

residence time. And the curve flats out around 90 minutes, which is about three

theoretical residence time. Therefore, as a criterion, all time-dependent CFD simulations

in this research are carried out from zero to three theoretical residence time.
117

7.5 Development of Metamodel based on ANN

Inspired by neural system in animals, ANN is a structure that data flows between neurons

with different weights and biases (Hopfield, 1988). There are three layers in general,

which are input layer, hidden layer and output layer respectively. Each neuron in the

hidden layer represents a nonlinear transfer function (Basheer and Hajmeer, 2000). In this

study, hyperbolic tangent function is addressed. The number of the nodes in the hidden

layer is to be determined. The neural network is to be trained by optimizing the weight

and bias values to best fit the data. Figure 51 shows the structure of the ANN model.

Figure 51 Schematic of an ANN based metamodel structure

To evaluate the performance of an ANN based metamodel, the following diagnostic plots

were employed: metamodel prediction values (ANN fitted values) versus observed values

(numerical simulation results). A plot appears close to a 45º line indicating an accurate

model. The sum of squares of errors (SSE) is an important aspect to assess the

performance of the model as well. By definition, SSE can be expressed by:


118

SSE = ∑ (Y − Yo )
2
(4.19)

where Y is the predicted value and Yo is observed value.

Besides SSE, there are two more criteria for assessing the adequacy of fit of a model,

which are, Cross Validation SSE (CV) and Akaike Information Criterion (AIC) (Akaike,

1978). Cross-validation is conducted by parsing the data into two groups. Firstly, the

metamodel is trained or fitted to one of the groups. The fitted model then is used to

predict the target values of the other group of data. This test is repeated for all groups of

data. This tests the model structures propensity to be over-fitted. Cross-validation

evaluations were conducted to analyze the artificial neural networks. For AIC, this

criteria weights the overall error in the outputs to the number of parameters in the model

(Akaike, 1978). The weighting function for the AIC is:

AIC = n ln(sse / n) + 2K + (2K(K +1)) / (n − K −1) (4.20)

where n is the number of data points, sse is the sum squared error, and K is the

number of total parameters in the system.

Both criteria are used to test if the data was over-fitted. SSE decreases monotonically

with the increase of the number of nodes in the hidden layer. The cross-validation SSE

initially decreases, but then is expected to increase when over-fitting occurs. AIC, which

is more computationally efficient than CV, is a penalized sum of squares designed to

correct for additional parameters and is expected to behave in a similar fashion to CV.

Training ANN model with 11 features of interests as input, starting from 1 node in the

hidden layer, the survival ratio of survival ratio and PAA residual as two outputs
119

separately, table 8 and 9 record the change of each criterion for survival ratio and PAA

residual concentration respectively.

Table 8 Criteria for survival ratio

Nodes # SSE AIC SSE from CV


1 2.585 43.9751 6.2833
2 0.890 180.102 6.1343
3 0.309 -3199 3.6332
4 0.068 -300.52 5.4968

Table 9 Criteria for PAA residual concentration

Nodes # SSE AIC SSE from CV


1 7.292 1.8641 38.6509
2 3.601 51.506 34.0300
3 1.641 -254.54 19.4365
4 0.855 -1706.8 41.3964

Table 8 shows that the model with three nodes in the hidden layer provides the best

performance with respect to both AIC and CV. In Table 9 on the other hand, there is a

discrepancy between AIC and Cross Validation wherein there is no optimal AIC

achieved. This study accords more significance to the results of cross validation in the

interpretation. Therefore, the model with three nodes in the hidden layer is chosen.

Figures 52 and 53 show observed values versus metamodel prediction values of MO

survival ratio and PAA residual, respectively.


120

Figure 52 Performance Plot of N survival ratio.


121

Figure 53 Performance Plot of PAA residual concentration.

7.6 VALIDATION OF ANN BASED METAMODEL

Numerical Validation

In order to test and validate the metamodels for MO and PAA, an additional 15 CFD runs

have been conducted. These 15 CFD cases are derived by the same quasi-random

sampling technique to generate points distinct from the 40 in the training set. Table 10

shows the 11 parameters of the 15 runs.


122

Table 10 Dataset for validation

No. Q θ l/w w /(b+1)/d b bx [PAA]o k k* m n


1 23 87 9.8 2.0 9 9% 4.1 0.023 1.456 0.393 0.348
2 63 51 9.8 1.7 6 10% 7.0 0.015 2.824 0.274 0.578
3 65 96 10.4 1.8 6 11% 1.7 0.022 3.142 0.316 0.616
4 67 15 11.0 1.8 7 11% 4.9 0.028 1.281 0.358 0.654
5 69 60 11.6 1.9 7 12% 8.0 0.035 1.598 0.401 0.692
6 19 49 3.5 0.9 8 15% 4.2 0.037 3.187 0.612 0.392
7 26 66 5.2 1.1 9 6% 5.2 0.024 1.962 0.235 0.506
8 28 43 5.8 1.1 9 7% 8.4 0.031 2.279 0.278 0.543
9 30 88 6.4 1.2 10 8% 3.1 0.038 2.597 0.320 0.581
10 71 95 7.0 1.0 7 9% 9.2 0.036 2.415 0.243 0.360
11 15 58 8.2 1.1 7 11% 4.3 0.016 3.051 0.328 0.436
12 17 36 8.7 1.2 7 11% 7.5 0.023 3.369 0.370 0.474
13 19 81 9.3 1.3 8 12% 2.2 0.030 1.196 0.413 0.511
14 22 24 9.9 1.3 8 12% 5.3 0.037 1.514 0.455 0.549

Figure 54 and 55 below display ANN fitted values of MO survival ratio and PAA

residual concentration versus the values from CFD simulations respectively. SSE for each

target is 0.8605, 2.1143 respectively. Then the mean squared error (MSE) of two models

are 0.2395, and 0.375 respectively, which can be considered with a reasonable accuracy

for estimating the system.


123

Figure 54 Validation plot of N survival ratio metamodel


124

Figure 55 validation plot of PAA residual concentration metamodel

Validation with Experimental Results

In 2006, D. Santoro(Santoro, 2005) conducted a series of experiments with PAA on a

pilot scale reactor(Santoro, 2005). The configuration of the pilot scale reactor is shown

on Figure 56.
125

Figure 56 The configuration of the pilot scale reactor used in the experiment

This pilot scale reactor has geometric similarity to the configurations simulated in the

CFD study. However, it is substantially smaller. The pilot scale reactor treats 0.0025

MGD, far below the range of the CFD study design space. Before utilizing the

experimental data, a scale-up operation is required. A physical similarity approach was

used with a scaled-up flow of 25.2 MGD. The other parameters after scale-up are shown

in the table below. The process kinetics were also those used from Santoro’s study16.

Table 11 Parameter table for the experiment after scale-up

L/w w/(b+1)/d Q θ b bx

2.38 2.1 25.2 MGD 30 min 3 10%

∗ m n [PAA] 0

0.015 min-1 1.75 0.57 0.54 Varying


126

Then applying the parameter set to the metamodel, the validation plot is obtained by

comparing the experimental N survival ratio with the metamodel prediction.

Figure 57 Validation plot with experiments in the pilot scale reactor. A 1 log experimental error is

shown for comparison.

The error bars on Figure 57 showed the typical error in measuring the number of

microorganisms as ±1 logarithm. The predictions from metamodel are smaller than

experimental measurements. But with considering the error bars, there are still 10 cases

out of 15 cases overlapping the ideal fitting line. When computing Reynolds number

(Re), the pilot-scale flow is about 5,300; on the other hand, the scaled-up model is around
127

250,000. The Froude number (Fr) for each set up is 0.008 and 0.036 respectively. Such

difference in both Re and Fr may have contributed to the discrepancy that was observed

from the plot.

To determine the influence of scale factor in this comparison, a sensitivity study on scale-

up factor (Sf) is carried out. Besides the original case with Sf = 10,000, two more cases

are considered: one with Sf = 3,000 (which represents about the lower limit of size in the

ANN metamodel development); the other one with Sf = 30,000. The parameters after

scale-up are shown in table 12 (Sf = 3,000) and table 13 (Sf = 30,000).

Table 12 Parameter table with Sf = 3,000

L/w w/(b+1)/d Q θ b bx
2.38 2.1 7.56 MGD 30 min 3 10%
∗ m n [PAA] 0
-1
0.015 min 1.75 0.57 0.54 Varying

Table 13 Parameter table with Sf = 30,000

L/w w/(b+1)/d Q θ b bx
2.38 2.1 75.6 MGD 30 min 3 10%
∗ m n [PAA] 0
-1
0.015 min 1.75 0.57 0.54 Varying

After applied these parameter sets into metamodel prediction, the results of sensitivity

study are shown below in Table 14.

Table 14 Average bias and Mean squared error in the sensitivity study

Scale factor Average bias Mean Squared Error


(MSE)
3,000 -0.36330 0.61455
10,000 -0.78968 1.00477
30,000 -0.75931 1.05911
128

As shown on Table 14, the prediction with Sf = 3,000 has the smallest average bias and

the least mean squared error. With increasing Sf value, difference in both Re and Fr

(between experimental pilot and the ANN model) is getting larger, resulting in increasing

discrepancy. This finding supports this hypothesis.

7.7 Conclusions

In this study, a series of CFD simulation were conducted as numerical experiments in

order to provide a relatively reasonable size data set to apply in the ANN based

metamodels train process, which are for predicting N survival ratio and PAA residual

concentration. The metamodels showed their accuracy of predicting the results of

wastewater disinfection with PAA in full scale reactor.

Evaluated by multiple criterions such as SSE, cross validation and AIC, the metamodels

with three nodes in the hidden layer were shown to be the best models for predicting N

survival ratio and PAA residual concentration. 15 more CFD simulations conducted for

validation showed that the metamodels can provide a reasonable accuracy of prediction

with a lot less time consuming compared with full-scale numerical simulations.

The nature of an ANN based metamodel is a mathematical interpretation of a highly

dimensional system, therefore it can be considered as a good reference in the design of a

wastewater treatment reactor due to its effectiveness. Comparing the mathematical model

with a full scale transient CFD simulation, the computation time for a reasonable

approximation is shorten from days to seconds, which is providing a superior solution to

the industrial use. A wide adaptable metamodel is suitable for any similar wastewater

treatment reactor configuration. For example, at given geometry and chemistry situations,

by varying the configuration of the reactor such as baffle number, baffle slit width,
129

residence time and flow rate, ANN based metamodels will provide a reasonable

prediction of the outcome for every case.

Nomenclature

The following symbols are used in this Chapter:

b = number of baffle;

bx = baffle slit width in percentage of reactor length;

d = depth of wastewater in the reactor;

kd = decay rate constant of PAA;

k* = Hom’s law constant;

L = length of reactor;

m = Hom’s law constant;

n = Hom’s law constant;

N = number of microorganisms;

N* = number of inactivated microorganisms

[PAA] = local concentration of PAA;

[PAA]0 = inlet concentration of PAA;

[PAA]* = concentration of decayed PAA;

Q = Flow rate;

Re = Reynold’s number;
130

θ = Theoretical residence time;

Γ = molecular diffusivity.
131

VIII. SUMMARY AND SUGGESTIONS FOR

FUTURE RESEARCH

8.1 Summary

The transport processes that are involved in the mixing of either two gases or two miscible

liquids in a T-junction mixer are investigated. The turbulent flow fields are calculated for

the T-junction with the LES model employing CFD-ACE+. In the mathematical model the

transport of species is described with a species concentration variable that is a function of

both space and time. The results obtained by numerical simulations are compared with two

well-defined experiments (Forney and Lee, 1982, Pan and Meng, 2001) for gas-gas and

liquid-liquid mixing respectively. Good mixing is obtained in a T-junction if the branch

inlet flow is designed to penetrate to the opposite pipe wall in the mixer.

The comparison between laminar air-methane and water-PAA solution cases showed that

with similar Rem and Rej, the air-methane case (with a smaller Sc) has improved mixing

beyond the T-junction, compared to the water-PAA solution case (with a higher Sc). LES

simulations for turbulent mixing showed differences between the mixing behavior of gas-

gas and liquid-liquid streams for similar geometrical and flow conditions. However, the

mixing of two turbulent flow streams (indicated by the maximum concentration value of

the solute cm) downstream of a T-junction are not affected by the Sc values even when Sc

for the gas and liquid-liquid streams vary by about three orders of magnitude. Cross-

sectional contour plots show that even with such high Reynolds number (as found in this

study), there is still a relatively large concentration gradient at the pipe exit (~15 pipe

diameters). Additional mixing devices like baffle plates are used in the industry to
132

improve mixing. For the liquid-liquid cases, the comparisons of flow structures between

CFD predictions and experimental measurements (Pan and Meng, 2001) showed

qualitative agreement only. CFD simulations show the effects due to density difference

cannot be overlooked in the study where temperature difference exists. The CFD

predictions for validation showed that the agreement with previously published

experimental measurement is good regarding both the temperature field and the velocity

field. Further simulations show that the shear layer formed in the near field after jet

injection is essential for the thermal mixing. With the increase of the temperature ratio

between the two flows, the gradient of the shear increases correspondingly, which

promotes the thermal mixing furthermore.

A series of CFD simulation were conducted as numerical experiments in order to provide

a relatively reasonable size data set to apply in the ANN based metamodels training

process, which are for predicting N survival ratio and PAA residual concentration. The

metamodels showed their accuracy of predicting the results of wastewater disinfection

with PAA in full scale reactor.

Evaluated by multiple criterions such as SSE, cross validation and AIC, the metamodels

with three nodes in the hidden layer were shown to be the best models for predicting N

survival ratio and PAA residual concentration. 15 more CFD simulations conducted for

validation showed that the metamodels can provide a reasonable accuracy of prediction

with a lot less time consuming compared with full-scale numerical simulations.

The nature of an ANN based metamodel is a mathematical interpretation of a highly

dimensional system, therefore it can be considered as a good reference in the design of a

wastewater treatment reactor due to its effectiveness. Comparing the mathematical model
133

with a full scale transient CFD simulation, the computation time for a reasonable

approximation is shorten from days to seconds, which is providing a superior solution to

the industrial use. A wide adaptable metamodel is suitable for any similar wastewater

treatment reactor configuration. For example, at given geometry and chemistry situations,

by varying the configuration of the reactor such as baffle number, baffle slit width,

residence time and flow rate, ANN based metamodels will provide a reasonable

prediction of the outcome for every case.

8.2 Suggestion for Future Research

8.2.1 Regarding the CFD modeling

• The simulations could be run in cluster machines with parallel processing which

would reduce computation time significantly.

• In the current study of T-junction mixing LES was applied with the Smagorinsky

model coupled with 2nd order differentiating method, future simulations could use the

dynamic model instead and a higher order scheme such as 3rd order or Quadratic

Upstream Interpolation for Convective Kinematics (QUICK) scheme.

• When comparing the CFD simulation results with experimental works, it might

not be a good idea to compare the flow contour or velocity vector. Even a very small

change in geometry, which is common in laboratory, could result in a large difference. If

time-averaged quantities are available, validations should be carried out with mean

quantities.
134

• In the future study of T-junction mixing, instead of sticking onto a vertical T-

shaped pipes, varying the angle of jet pipe with main pipe is to be discovered. For

example, a Y-shaped pipeline.

8.2.2 Regarding the simulations of full scale reactor

• In current study, the simulations were carried out with k-ϵ model. In the future, a

more advanced turbulent model scheme should be applied.

• In the future study, other parameters such as temperature, pH, or gravity should be

considered in the CFD modeling as well.

8.2.3 Regarding the development of ANN based metamodel

• We show that the ANN based metamodeling can provide reasonably accurate

predictions with a small database. In future study, when facing the similar scenario, i.e.

limited data points of a high-dimensional system, to develop the ANN based metamodel

is a better approach than a polynomial regression.

• To determine the number of the hidden nodes, the criteria do not always coincide.

The weight of each criterion should be evaluated individually. For example, in this study,

the cross-validation values did not follow the same trend as AIC values. To fulfill the

purpose of this research, cross-validation was treated with more weight.


135

LIST OF REFERENCES
[1] Aieta E M, Berg J D, Roberts P V, and Cooper R C. Comparison of chlorine dioxide
and chlorine in wastewater disinfection. Journal of Water Pollution Control Federation,
pages 810-822, 1980.

[2] Akaike, H., 1978, "On the likelihood of a time series model", Statistician, 27, pp. 217-
235
[3] Alasri A, Roques C, Michel G, Cabassud C, Aptel P. Bactericidal properties of
peracetic acid and hydrogen peroxide, alone and in combination, and chlorine and
formaldehyde against bacterial water strains. Can J Microbiol 1992;38:635 – 42.
[4] Anders, U. and Korn, O., 1999, "Model Selection in Neural Networks", Neural
Networks, 12, pp. 309-323
[5]Antonelli, M., Turolla, A., Mezzanotte, V., & Nurizzo, C. (2013). Peracetic acid for
secondary effluent disinfection: a comprehensive performance assessment.
[6] Baldry MGC. The bactericidal, fungicidal, and sporicidal properties of hydrogen
peroxide and peracetic acid. J Appl Bacteriol 1983;54: 417 – 23.
[7] Baldry MGC, French MS. Activity of peracetic acid against sewage indicator
organisms. Water Sci Technol 1989a;21:1747 – 9.
[8] Baldry MGC, French MS. Disinfection of sewage effluent with peracetic acid. Water
Sci Technol 1989b;21:203 – 6.
[9] Barton, R. R. (1994). "Metamodeling: A state of the art review." Simulation
Conference Proceedings, 1994. WinterLake Buena Vista, FL, USA, 237-244.
[10] Bartrand, T.A., 2006. High Resolution Experimental Studies and Numerical
Analysis of Fine Bubble Ozone Disinfection Contactors (Doctoral dissertation).
Available from ProQuest Dissertations & Theses Database. (UMI No. 3239812).
[11] Bartrand, T.A., Farouk, B., Haas, N.C., 2009. Countercurrent gas/liquid flow and
mixing: implications for water disinfection. Int. J. Multiph . Flow. 35 (2), 171-184.
[12] Basheer, I. A., and Hajmeer, M. (2000). "Artificial neural networks: fundamentals,
computing, design, and application." Journal of Microbiological Methods, 43, 3-31.
[13] Blatchley, E., Hunt, B., Duggirala, R., Thompson, J., Zhao, J., and Halaby, T.
(1997). "Effects of Disinfectants on Wastewater Effluent Toxicity." Water Res., 31, 8.
[14] Block SS. Disinfection, sterilization, and preservation. 4th ed. Philadelphia:
Lea&Febiger Pubs; 1991.
[15] Calmet, I., and Magnaudet, J., 1997, "Large-eddy simulation of high-Schmidt
number mass transfer in a turbulent channel flow," Physics of Fluids, 9, p. 438.
136

[16] Chilton, T. H., and Genereaux, R. P., 1930, "The Mixing of Gases for Reaction,"
AIChE Transuctions, 25, p. 103.
[17] Colgan S, Gehr R. Disinfection. Water Environ Technol 2001;13:29 – 33.
[18] Craun, G. F., Bull, R. J., and Symons, J. M. (1994). "Balancing chemical and
microbial risks of drinking water disinfection, Part 1. Benefits and potential risks."
Journal of Water Supply: Research and Technology - AQUA, 43, 8.
[19] E. ToolBox, Water - Thermal Conductivity, in,
https://www.engineeringtoolbox.com/water-liquid-gas-thermal-conductivity-temperature-
pressure-d_2012.html, 2018.
[20] ESI-Group, 2016, "CFD-ACE, 2016, "ESI-US RD (2016) CFD-ACE+ modules
manual V2016. Huntsville, AL," ESI Group.
[21] ESI-Group, 2017, "CFD-ACE, 2017, "ESI-US RD (2017) CFD-ACE+ modules
manual V2017. Huntsville, AL," ESI Group.
[22] Fraser JAL, Godfree AF, Jones F. Use of peracetic acid in operational sewage sludge
disposal to pasture. Water Sci Technol 1984;17:451 – 66.
[23] Forney LJ, Lee HC. Optimum Dimensions for Pipeline Mixing at a T-Junction.
AiChE Journal. 1982;28(6):8.
[24] Forney, L. J., Noureddine, N., and Hanh, X. V., 1996, "Optimum jet mixing in a
tubular reactor," AIChE Journal, 42(11), pp. 3113 - 3122.
[25] Gehr R, Cochrane D, French M. Peracetic acid as a disinfectant for municipal
wastewaters: encouraging performance results from physicochemical as well as
biological effluents. Proc of the US water environment federation disinfection
conference; 2002.
[26] Greene, D. J., Farouk, B., and Haas, C. N. (2004). "CFD Design Approach for
Chlorine Disinfection Processes." J Am Water Works Assoc. (96), 13.
[27] Haas, C. N. and Joffe, J., 1994, "Disinfection under Dynamic Conditions:
Modification of Hom's Model for Decay", Environ. Sci. Technol., 28, pp. 1367-1369
[28] Hirota, M., Mohri, E., Asano, H., and Goto, H., 2010, "Experimental study on
turbulent mixing process in cross-flow type T-junction," International Journal of Heat
and Fluid Flow, 31, pp. 776-784.
[29] Hopfield, J. J. (1988). "Artificial neural networks." IEEE Circuits and Devices
Magazine, 4(5), 3 - 10.
[30] Hoque, S., Farouk, B., and Haas, C. N. (2011). "Development of metamodels for
predicting aerosol dispersion in ventilated spaces." Atmos Environ., 45, 12.
[31] Hoque, S., Farouk, B., and Haas, C. N. (2011). "Development of Artificial Neural
Network Based Metamodels for Inactivation of Anthrax Spores in Ventilated Spaces
Using Computational Fluid Dynamics." J Air Waste Manag Assoc., 61, 15.
137

[32] Hoque, S., Farouk, B., and Haas, C. N. (2010). "Multiple Linear Regression Model
Approach for Aerosol Dispersion in Ventilated Spaces Using Computational Fluid
Dynamics and Dimensional Analysis." Journal of Environmental Engineering, 136(6),
638-649.
[33] Lazarova V, Janex ML, Fiksdal L, Oberg C, Barcina I, Pommepuy M. Advanced
wastewater disinfection technologies: short and long term efficiency. Water Sci Technol
1998;38:109 – 17.
[34] Lefevre F, Audic JM, Ferrand F. Peracetic acid disinfection of secondary effluents
discharged off coastal seawater. Water Sci Technol 1992;25:155 – 64.
[35] Leonard A. On the Energy Cascade in Large-Eddy Simulations of Turbulent Fluid
Flows. Adv. Geophys. 1974;18:12
[36] Liberti L, Notarnicola M. Advanced treatment and disinfection for municipal
wastewater reuse in agriculture. Water Sci Technol 1999;40:235 – 45.
[37] Liberti L, Lopez A, Notarnicola M, Barnea N, Pedahzur R, Fattal B. Comparison of
advanced disinfecting methods for municipal wastewater reuse in agriculture. Water Sci
Technol 2000;42:215 – 20.
[38] Monarca S, Feretti D, Zerbini I, Zani C, Alberti A, Richardson SD, et al. Studies on
mutagenicity and disinfection by-products in river drinking water disinfected with
peracetic acid or sodium hypochlorite. Proc of IWA world conference, Berlin, Germany;
2001.
[39] Monarca S, Richardson SD, Feretti D, Grottolo M, Thruston AD, Zani C, et al.
Mutagenicity and disinfection by-products in surface drinking water disinfected with
peracetic acid. Environ Toxicol Chem 2002;21:309 – 18.
[40] Narayan, B. C., 1971, "Experimental Study of the Rates of Turbulent Mixing in Pipe
Flow," M.S. Thesis, University of Tulsa, Tulsa, OK
[41] Naik-Nimbalkar VS, Patwardhan AW, Banerjee I, Padmakumar G, Vaidyanathan G,
Thermal mixing in T-junctions, Chemical Engineering Science, 65(22) (2010) 5901-
5911.
[42] Nguyen, N. T., and Wu, Z. G., 2005, "Micromixers - a review," Journal of
Micromechanics and Microengineering, 15, pp. 1-16.
[43] Pan G, Meng H. An Experimental Study of Turbulent Mixing in a Tee Mixer Using
PIV and PLIF. AIChE Journal. 2001;47(12)
[44] Raj R, Mathur N, Buwa VV. Numerical Simulations of Liquid−Liquid Flows in
Microchannels. Ind. Eng. Chem. Res. 2010;49(21):10606-10614
[45] Rajala-Mustonen RL, Toivola PS, Heinonen-Tanski H. Effect of peracetic acid and
UV irridation on the inactivation of coliphages in wastewater. Water Sci Technol
1997;35:237 – 41.
138

[46] Santoro, D. (2005). "CFD Modeling of Wastewater Disinfection Process by PAA."


Ph.D, Technical University of Bari.
[47] Selvam, P. K., 2017, "Thermal mixing characteristics of flows in horizontal T-
junctions," Ph. D thesis, University of Stuttgart.
[48] Simpson, T.W., Peplinski, J.D., Koch, P.N., Allen, J.K., 2001. Metamodels for
computerbased engineering design:survey and recommendations. Engineering with
Computers 17, 129-150.
[49] Smagorinsky J. General Circulation Experiments with the Primitive Equations.
Monthly Weather Review. 1963;91(3):99-164.
[50] Sroka LM, Forney LJ. Fluid Mixing with a Pipeline Tee: Theory and Experiment.
AIChE Journal. 1989;35:9.
[51] Stampi S, De Luca G, Zanetti F. Evaluation of the efficiency of peracetic acid in the
disinfection of sewage effluents. J Appl Microbiol 2001;91:833 – 8.
[52] Tizzoni, M., 2016, "Wastewater disinfection with peracetic acid, Master's thesis",
School of Civil, Environment and Land Management Engineering, Politechnico Di
Milano, Milan, Italy,
[53] Tominagaa Y, Stathopoulosh T. Turbulent Schmidt numbers for CFD analysis with
various types of fiowfield. Atmospheric Environment 2007;41:8051-8099.
[54] US Environmental Protection Agency. Wastewater technology fact sheet: ozone
disinfection., 1999.

[55] Wagner M, Brumelis D, Gehr R. Disinfection of wastewater by hydrogen peroxide


or peracetic acid: Development of procedures for measurement of residual disinfectant
and application to a physiochemically treated municipal effluent. Water Environ Res
2002;74:33 – 50.
[56] Wang, G. G. and Shan, S., 2007, "Review of Metamodeling Techniques in Support
of Engineering Design Optimization", Trans. ASME
[57] Wei W, Farouk B, Haas CN. Optimized Design OF Wastewater Disinfection
Reactors based on an Artificial Neural Network metamodel. ASME 2016 International
Mechanical Engineering Congress and Exposition. Phoenix, AZ, 2016
[58] Wei W, Farouk B, Haas C N. Mixing of Gases and Miscible Liquids in a T-Junction.
ASME 2017 International Mechanical Engineering Congress and Exposition. Tampa, FL,
US, 2017
[59] Wols, B.A., Hofman, J.A.M.H., Uijttewaal, W.S.J., Rietveld, L.C., Stelling, G.S.,
van Dijk, J.C., 2008a. A particle tracking technique to estimate disinfection efficacy in
drinking water treatment plants. In: 6th International Conference on CFD in the Oil and
Gas, Metallurgical and Process Industries, Trondheim.
139

[60] Wols, B.A., Uijttewaal, W.S., Rietvelda, L.C., Stellinga, G.S., van Dijka, J.C.,
Hofman, J.A.M.H., 2008b. Residence time distributions in ozone contactors. Ozone Sci.
Eng. J. Int. Ozone Assoc. 30 (1), 49-57.
[61] Wols, B.A., Hofman, J.A.M.H., Uijttewaal, W.S.J., Rietveld, L.C., van Dijk, J.C.,
2010a. Evaluation of different disinfection calculation methods using CFD. Environ.
Model. Softw. 25, 573e582.
[62] Wols, B.A., Shao, L., Uijttewaal, W.S.J., Hofmana, J.A.M.H., Rietveld, L.C., van
Dijk, J.C., 2010b. Evaluation of experimental techniques to validate numerical
computations of the hydraulics inside a UV bench-scale reactor. Chem. Eng. Sci. 65 (15),
4491-4502.
[63] Wols, B.A., Uijttewaal, W.S.J., Hofmana, J.A.M.H., Rietveld, L.C., van Dijk, J.C.,
2010c. The weaknesses of a k-epsilon model compared to a large-eddy simulation for the
prediction of UV dose distributions and disinfection. Chem. Eng. J. 162 (2), 528-536.
[64] Wols, B.A., Hofman, J.A.M.H., Beerendonk, E.F., Uijttewaal, W.S.J., van Dijk, J.C.,
2011. A systematic approach for the design of UV reactors using computational fluid
dynamics. AIChE J. 57 (1), 193-207.
[65] Wols, B.A., Hofman-Caris, C.H.M., Harmsen, D.J.H., Beerendonk, E.F., van Dijk,
Chan, J.C., Blatchley, E.R., 2012. Comparison of CFD, biodosimetry and Lagrangian
actinometry to assess UV reactor performance. Ozone Sci. Eng. 34(2), 81-91.
[66] Xu P, Janex M L, Savoye P, Cockx A, and Lazarova V. Wastewater disinfection by
ozone: main parameters for process design. Water Research, 36(4):1043-1055, 2002.
[67] Yang Z. Large-eddy simulation: Past, present and the future. Chinese Journal of
Aeronautics. 2014;28(1):14.
[68] Yuan Z, Ni Y, van Heiningen ARP. Kinetics of peracetic acid decomposition: Part I.
Spontaneous decomposition at typical pulp bleaching conditions. Can J Chem Eng
1997;75:37 – 41.
[69] Zang, T. A., 1991, "Numerical simulation of the dynamics of turbulent boundary
layers: perspectives of a transition simulator," Philos. Trans. Roy. Soc. Lond. A,
336(1647), pp. 95-102.
[70] Zhang J, Tejada-Martínez AE, Zhang Q. Developments in computational fluid
dynamics-based modeling for disinfection technologies over the last two decades: A
review. Environmental Modelling & Software 2014;58:71-85.
[71] Zhiyin, Y., 2015, "Large-eddy simulation: Past, present and the future," Chinese
Journal of Aeronautics, 28(1), pp. 11-24.
140

VITA
Wangshu Wei was born in the city of Shenyang, China. He started his academic career in

Huazhong University of Science and Technology in Wuhan, China, from where he

received his bachelor’s degree in mechanical engineering in 2012. Then he continued his

study towards a doctoral degree in Drexel University since the Fall of 2013. The research

area he focused on was the 3-dimensional numerical flow realization with mass and heat

transfer especially in wastewater treatment process. Aiming at identifying the parameters

that affecting the mixing behavior of fluids, his research was to provide a solution to

enhance the mixing process. Furthermore, he proposed a computational efficient

mathematical model of wastewater disinfection process with Peracetic Acid, which is

providing a superior solution to the industrial use.

LIST OF PUBLICATIONS

a). Wei W, Farouk B, Haas C N. Optimized Design OF Wastewater Disinfection Reactors

based on an Artificial Neural Network metamodel. ASME 2016 International Mechanical

Engineering Congress and Exposition. Phoenix, AZ, US, 2016

b). Wei W, Farouk B, Haas C N. Mixing of Gases and Miscible Liquids in a T-Junction.

ASME 2017 International Mechanical Engineering Congress and Exposition. Tampa, FL,

US, 2017

c). Wei W, Farouk B, Haas C N. Gas mixing vs. liquid mixing in a T-junction: a

comparative study. Submitted to Journal of AIChE


141

d). Wei W, Farouk B, Haas C N. Development of a CFD Based Artificial Neural Network

Metamodel in Wastewater Disinfection Process with Peracetic Acid. Submitted to

Journal of Environmental Engineering

e) Wei W, Farouk B, Haas C N. Thermal Mixing of two liquids in a T-junction. Submitted

to International Journal of Mass and Heat Transfer


142

You might also like