CFD-aided modelling of activated sludge systems – A critical review

CFD-aided modelling of activated sludge systems – A critical review

Accepted Manuscript CFD-Aided Modelling of Activated Sludge Systems – A Critical Review Anna M. Karpinska, John Bridgeman PII: S0043-1354(15)30336-5 ...

3MB Sizes 0 Downloads 65 Views

Accepted Manuscript CFD-Aided Modelling of Activated Sludge Systems – A Critical Review Anna M. Karpinska, John Bridgeman PII:

S0043-1354(15)30336-5

DOI:

10.1016/j.watres.2015.11.008

Reference:

WR 11635

To appear in:

Water Research

Received Date: 25 June 2015 Revised Date:

1 November 2015

Accepted Date: 2 November 2015

Please cite this article as: Karpinska, A.M., Bridgeman, J., CFD-Aided Modelling of Activated Sludge Systems – A Critical Review, Water Research (2015), doi: 10.1016/j.watres.2015.11.008. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 1

Title: CFD-Aided Modelling of Activated Sludge Systems –

2

A Critical Review.

3 4

Authors: Anna M. Karpinskaa, John Bridgemana

5

a

6

Edgbaston, Birmingham, B15 2TT, United Kingdom

7

Email

8

[email protected]

9

Corresponding

RI PT

[email protected];

author:

Anna

([email protected])

SC

addresses:

M.

Karpinska

M AN U

10

– School of Civil Engineering, University of Birmingham,

11 Abstract

13

Nowadays, one of the major challenges in the wastewater

14

sector is the successful design and reliable operation of

15

treatment

16

efficiencies to comply with effluent quality criteria, while

17

keeping the investment and operating cost as low as possible.

18

Although conceptual design and process control of activated

TE D

12

which

guarantee

high

treatment

AC C

EP

processes,

19

sludge plants are key to ensuring these goals, they are still

20

based

21

experience, dominated often by rule of thumb. This review

22

paper discusses the rationale behind the use of Computational

23

Fluid Dynamics (CFD) to model aeration, facilitating

24

enhancement of treatment efficiency and reduction of energy

25

input. Several single- and multiphase approaches commonly

on

general

empirical

1

guidelines

and

operators’

ACCEPTED MANUSCRIPT used in CFD studies of aeration tank operation, are

27

comprehensively described, whilst the shortcomings of the

28

modelling assumptions imposed to evaluate mixing and mass

29

transfer in AS tanks are identified and discussed. Examples and

30

methods of coupling of CFD data with biokinetics, accounting

31

for the actual flow field and its impact on the oxygen mass

32

transfer and yield of the biological processes occurring in the

33

aeration tanks, are also critically discussed. Finally, modelling

34

issues, which remain unaddressed, (e.g. coupling of the AS

35

tank with secondary clarifier and the use of population balance

36

models to simulate bubbly flow or flocculation of the activated

37

sludge), are also identified and discussed.

M AN U

SC

RI PT

26

38 Keywords

40

Activated Sludge; aeration; Computational Fluid Dynamics;

41

hydrodynamics; multiphase models; turbulence

Nomenclature

AC C

43

EP

42

TE D

39

2

ACCEPTED MANUSCRIPT mass fraction of phase  [-]



sum of interfacial forces shared by the phases [N]



sum of interfacial forces between the continuous and dispersed phases [N] diffusion flux of ′ species [kg m-2 s-1]



turbulent kinetic energy, [m2 s-2]

M AN U

SC



RI PT



volumetric mass transfer coefficient [h-1]



mass transfer from phase  to  [kg s-1]



mass transfer from phase  to  [kg s-1]

TE D



pressure [Pa]



averaged pressure field [Pa]

AC C

EP





net rate of production of ′ species due to chemical



Reynolds number [-]

reaction [kg m-3 s-1]

3

ACCEPTED MANUSCRIPT 

source term representing rate of creation of ′ species from dispersed phase and any user-defined sources [kg

time [s]



particle relaxation time [s]



velocity [m s-1]

 ,  , 

drift velocity for phase  [m s-1]

M AN U

,

SC



RI PT

m-3 s-1]

velocity components [m s-1]

averaged velocity term [m s-1]

′

fluctuating velocity term [m s-1]



velocity of mixture [m s-1]

AC C

EP

TE D

̅



velocity of the particle [m s-1]



velocity of phase  [m s-1]



velocity of phase  relative to velocity of phase  [m s-1]

4

ACCEPTED MANUSCRIPT velocity of phase  [m s-1]



distance [m]

 ,

direction) [m] mass fraction of the ’-th species [-]

SC

!



spatial coordinates (displacement in the streamwise

Greek letters

"∗ $

damping function coefficient [-]

turbulent kinetic energy dissipation rate [m2 s-3] dynamic viscosity of the fluid [Pa s]

EP

%

phasic volume fraction [-]

TE D

"

M AN U

44 45

RI PT

,

dynamic viscosity of the mixture [Pa s]

AC C

% % %& '

'

dynamic viscosity of the phase  [Pa s] turbulent viscosity [Pa s]

fluid density [kg m-3] density of the mixture [kg m-3]

5

ACCEPTED MANUSCRIPT '

density of the particle [kg m-3] density of the phase  [kg m-3]

(

specific turbulence dissipation (frequency)

'

density of the phase  [kg m-3]

[s-1]

)* 

′ +

,



refers to drift velocity index or counter species index

index or counter

refers to turbulence kinetic energy index or counter

refers to mixture

EP

-

M AN U

Indices

TE D

47

SC

46

refers to particle



refers to phase 



refers to turbulence

refers to phase 

AC C



RI PT

'

48 49 50

Abbreviations

51

ADV – Acoustic Doppler Anemomentry

52

AS – Activated Sludge 6

54

ASM1 – Activated Sludge Model No. 1

55

ASP – Activated Sludge Plant

56

BOD – Biochemical Oxygen Demand

57

CFD – Computational Fluid Dynamics

58

COD – Chemical Oxygen Demand

59

CPU – Central Processing Unit

60

CSTR – Continuous Stirred Tank Reactor

61

DO – Dissolved Oxygen

62

LDV – Laser Doppler Velocimetry

63

LES – Large |Eddy Simulation

64

MDV– Mono-directional Velocimetry

65

MLSS – Mixed Liquor Suspended Solids

66

MRF – Multiple Reference Frames

67

PBM – Particle Balance Model

68

PDA – Particle Dynamic Analysis

69

PFR – Plug Flow Reactor

70

PID – Proportional-Integral-Derivative

SC

ASM – Activated Sludge Model

AC C

EP

TE D

M AN U

53

RI PT

ACCEPTED MANUSCRIPT

71

QMOM – Quadrature Method of Moments

72

RAM – Random Access Memory

73

RANS – Reynolds Averaged Navier- Stokes Simulation

74

RSM – Reynolds Stress Model

75

RTD – Residence Time Distribution

76

SCADA – Supervisory Control and Data Acquisition

77

SGS – Sub-grid Scale

7

ACCEPTED MANUSCRIPT SMM – Standard Method of Moments

79

SRT – Solids Retention Time

80

TSS – Total Suspended Solids

81

URANS – Unsteady Reynolds Averaged Navier- Stokes

82

Simulation

83

WWTP – Wastewater Treatment Plant

84

RI PT

78

1. Introduction

86

To comply with global water policy focussed on responsible

87

management of water resources and protection of public health,

88

wastewater collected from municipalities and communities

89

must be treated to achieve levels imposed by discharge permits

90

and maximum daily loads, allowing its subsequent return to

91

receiving water bodies, or to the land or even to be reused. In

92

the last century, the application of scientific knowledge and

93

engineering practice led to significant developments in the

94

wastewater

95

treatment based on aerobic biological methods, specifically the

96

activated sludge (AS) process (Ardern and Lockett 1914),

97

which is now a well-documented standard for many wastewater

98

treatment utilities. The objectives of secondary biological

99

wastewater treatment in the AS process have also expanded

100

from an early emphasis on high levels of Biological Oxygen

101

Demand (BOD) and Total Suspended Solids (TSS) removal to

102

cover enhanced nutrients (N and P) removal, as the process

TE D

M AN U

SC

85

particularly in

biological

secondary

AC C

EP

sector,

8

ACCEPTED MANUSCRIPT itself has flexibility and numerous modifications can be tailored

104

to meet specific requirements.

105

Undoubtedly, one of the foremost challenges in the wastewater

106

sector is the successful design and reliable operation of

107

treatment plant, which guarantee high treatment efficiencies in

108

order to meet the effluent quality standards defined by

109

regulators, while keeping the investment and operating cost as

110

low as possible (Brouckaert and Buckley 1999, Do-Quang et al.

111

1999). One of the characteristic features of ASP is continuous

112

operation of the aeration process and sludge and nitrates

113

recycling, and thus the process performance relies on a steady

114

energy supply for operation of air blowers and sludge and

115

mixed liquor recirculation pumps. Aeration accounts for the

116

largest fraction of a total wastewater treatment plant’s (WWTP)

117

energy bill, ranging from 45 to 75% (Reardon 1995, Rieger et

118

al. 2006) and in extreme cases of stringent effluent nitrogen

119

criterion requiring enhanced nitrification, even up to 85%.

120

Thus, aeration has a significant effect on the operation and

AC C

EP

TE D

M AN U

SC

RI PT

103

121

maintenance budget of water utilities (WEF 2009). As a

122

consequence of the global emphasis on the water-energy-food-

123

climate change nexus there is an urgent need to reduce energy

124

usage

125

conservation measures, engineering practices and management

126

programs. According to guidelines (EPA 2013, WEF 2009)

127

opportunities for improving energy efficiency in wastewater

at

WWTPs

by

imposing

9

cost-effective

energy

ACCEPTED MANUSCRIPT treatment utilities can be obtained by optimizing aeration

129

processes or equipment upgrades, which focus on replacing

130

items such as blowers with more efficient models; replacing the

131

whole aeration system with less energy intensive systems (e.g.

132

replacement of the surface aeration system by porous diffusers

133

in full floor coverage configuration); and operational

134

modifications, involving reduction of the energy requirements

135

to perform specific functions by modification of the aeration

136

control systems, e.g. on-off operation allowing formation of the

137

anoxic conditions for denitrification. The last option facilitates

138

greater savings than equipment upgrades, and may not require

139

capital investment. Nevertheless, current best available aeration

140

technology (i.e. membrane diffusers supplied by atmospheric

141

air) are characterized by relatively low Standard Oxygen

142

Transfer Efficiencies of around 40 up to 60% (EPA 1989,

143

Mueller et al. 2002, Taricska et al. 2009), which reduce with

144

time as a result of fouling and scaling. Consequently, a wide

145

range of multidisciplinary approaches contribute to current

AC C

EP

TE D

M AN U

SC

RI PT

128

146

research aiming to improve the development, troubleshooting

147

and management of aeration systems.

148

2. Current Trends in Engineering Practice

149

2.1. Aeration Control and Design Assumptions

150

When investigating performance and energy expenditure of AS

151

plants it is clear that dissolved oxygen (DO) concentration is a

152

key process variable, which controls both, nutrient removal (in 10

ACCEPTED MANUSCRIPT the case of biological nutrient removal plants) and thus effluent

154

quality, and the operating cost of the utility. While operating

155

DO profiles and nitrogen patterns in the AS system are usually

156

obtained from biokinetics modelling with Activated Sludge

157

Model No. 1- ASM1 (Henze et al. 2000) employed by

158

supervisory control and data acquisition (SCADA) systems, the

159

robust modelling of the optimization strategies requires

160

implementation of complex computational algorithms and off-

161

line optimization techniques, which allow for coupling of the

162

biological process with pre-defined control variables such as

163

minimal DO concentration in the aeration tank or effluent

164

nitrogen criterion (Åmand and Carlsson 2012, Chachuat et al.

165

2005, Cristea et al. 2011, De Araújo et al. 2011, Fernández et

166

al. 2011, Fikar et al. 2005, Holenda et al. 2007, Holenda et al.

167

2008).

168

Although hydraulic design of wastewater treatment systems is a

169

crucial step to assure reliable and energy-optimised operation

170

of the process, it is usually labelled as a “low-tech” task, which

AC C

EP

TE D

M AN U

SC

RI PT

153

171

is based on empirical guidelines without sound theoretical basis

172

(Bosma and Reitsma 2007, Pereira et al. 2012, Stamou 2008).

173

Thus, in the majority of designs, flow behaviour in the unit

174

process tanks is predicted via the ideal reactor model while the

175

actual reactor hydrodynamics are not taken into account

176

(Stamou 2008). The classic example for Activated Sludge

177

Plants (ASPs) is the assumption of a completely mixed flow 11

ACCEPTED MANUSCRIPT regime in aerobic, anoxic and anaerobic tanks. Other

179

commonly practised rules of thumb are (Samstag and Wicklein

180

2012, Tchobanoglous et al. 2003) the assumption that in AS

181

basins equipped with diffused aeration systems, the air

182

requirement to ensure good mixing will vary from 1.2 – 1.8 m3

183

h-1 / m3 of tank volume; and typical power requirements for

184

maintaining a completely mixed flow regime with mechanical

185

aerators varies from 13 to 26 W per m3 of tank volume. None

186

of these assumptions consider the impact of tank hydraulics

187

(cross-section, depth or presence of baffles), energy input, or

188

any variable affecting mixing, such as local density gradients

189

due to solids transport. Furthermore, none of these guidelines

190

defines clearly hydraulic features and performance of

191

“completely mixed” wastewater treatment systems. The one

192

commonly used criterion for “good mixing” in AS process

193

control is that variations of solids concentration across the

194

complete mixed tank profile should be less than 10% (Samstag

195

and Wicklein 2014) However, the proper design of such “well

AC C

EP

TE D

M AN U

SC

RI PT

178

196

mixed” AS systems requires a sophisticated analysis of flow

197

behaviour accounting for mixing patterns in the tank,

198

distribution of the oxygen, solids and determination of local

199

densities.

200

3. Dynamic Behaviour of AS Tanks

201

Dynamic modelling of wastewater treatment plants has been

202

shown to be a powerful tool providing detailed insight into the 12

ACCEPTED MANUSCRIPT 203

unit process and system behaviour, useful for optimization

204

studies

205

evaluation, operational optimization, or controller design) and

206

model-based process control (Langergraber et al. 2004).

207

Nonetheless, the majority of the ASP design procedures are still

208

limited to the empirical principles and static models, such as

209

the ‘classic’ ATV-A-131 guideline (ATV-DVWK 2000), while

210

control strategies are based on a simple or cascade

211

proportional-integral-derivative (PID) controllers for DO and

212

ammonia. While not often used for design and control of ASP

213

due to its complexity, the systemic approach based on well-

214

established ASM models focuses mainly on the reactions of

215

biochemical conversion within the ideal reactors: usually one or

216

a cascade of Continuous Stirred Tank Reactors (CSTRs) or

217

Plug Flow Reactor (PFR) (Abusam and Keesman 1999, Le

218

Moullec et al. 2011, Makinia and Wells 2000, Pereira et al.

219

2009). Such approaches enable only quantitative prediction of

220

the oxygen consumption and nutrients removal and qualitative

conceptual

process

design,

performance

AC C

EP

TE D

M AN U

SC

RI PT

(e.g.

221

assessment of biomass growth and decay. Nevertheless, as the

222

overall biochemical conversion reactions occurring in AS are of

223

orders greater than zero, the wastewater treatment efficiency in

224

such non-ideal systems will depend on the bioreactor’s

225

hydrodynamics (Le Moullec et al. 2008), spatial distribution of

226

oxygen, and temperature. Therefore, a more accurate forecast

227

of the actual local scale phenomena occurring within the tank,

13

ACCEPTED MANUSCRIPT (such as DO patterns and solids profiles) is of crucial

229

importance to enhance system design, process efficiency and

230

energy requirements, to predict failures of existing aeration

231

systems and finally, to design upgrades and enhance control

232

strategy, such as fine-tuning blower operation.

233

The physics of typical AS systems is complex, not only due to

234

presence of the multiphase (gas-liquid-solid) flow, comprising

235

mixed liquor and air/oxygen, but also due to the different

236

length scales between the sludge flocs, bubbles and tank

237

geometry; and furthermore, different velocities of the phases

238

imparted by mixers and aerators yielding turbulent Reynolds

239

numbers (Karpinska Portela 2013, Pereira et al. 2012). The

240

fluid flow in such bioreactors is governed by the vessel

241

geometry, physical properties of its contents (phase, density,

242

viscosity) and operating condition variables (flow rates and

243

concentrations). At the same time, fluid flow governs the local

244

concentration of components, interphase contact and mass

245

transfer, reaction conversions and performance (Nopens and

AC C

EP

TE D

M AN U

SC

RI PT

228

246

Wicks 2012).

247

In engineering practice, whilst assessment of the flow regime,

248

and thus overall mixing phenomena in AS bioreactors, can be

249

achieved via experimental assessment of local flow velocities

250

through a tracer technique, the dimensions of full scale units

251

generally render this unfeasible (Pereira et al. 2012, Stamou

14

ACCEPTED MANUSCRIPT 2008), making simulation an attractive alternative. Several

253

factors contribute to the increasing popularity of modelling in

254

engineering practice, as it is a cost and time efficient solution,

255

which allows evaluation of process performance (whether a

256

new unit or modification will operate properly); prediction of

257

consequences

258

quantification of bottlenecks in liquid or solid handling lines in

259

the AS system. Thus, prediction of Residence Time

260

Distributions (RTDs) of settled sewage is a fundamental tool to

261

help understand and analyse a flow system providing realistic

262

information on tank hydrodynamics (Danckwerts 1953,

263

Levenspiel 1999, Nauman 2007). Furthermore, the RTD yields

264

information about the macromixing within the reactor, allowing

265

recognition of mixing behaviours that are not plug flow or

266

complete mixing regimes, and that are usually described by

267

tank-in-series models or even more complex arrangements of

268

the unit reactors (Pereira et al. 2012). Consequently, successful

269

biological

implementation;

isolation

and

EP

TE D

M AN U

SC

before

RI PT

252

treatment

modelling

combining

AC C

wastewater

270

hydrodynamics, mass transfer and biochemical reactions

271

kinetics remains one of the major goals in wastewater

272

engineering (Le Moullec et al. 2010a, b, Morchain et al. 2014,

273

Pereira et al. 2012).

274

4. Computational Fluid Dynamics

275

Since the 1970s, increasing computational power has been

276

accompanied by a rapid development in the software intended 15

ACCEPTED MANUSCRIPT 277

for solution of fluid flow problems. Nowadays, a wide range of

278

software suites intended for the solution of complex fluid flow

279

problems is commercially available. Initially, CFD was almost

280

exclusively

281

industries allowing simulation of the processes occurring in

282

combustion chambers of rocket engines; physico-chemical

283

processes in the flow around rocket airframe and supersonic

284

aircrafts. Subsequently CFD found applications with chemical

285

engineers, mainly for design of reaction vessels, and in the last

286

two decades, application of CFD has been extended to the civil

287

and environmental engineering sectors (Kochevsky 2004).

288

Recent developments in multiphase flow research have seen a

289

steady growth in the application of CFD modelling in

290

wastewater treatment, with a focus on the design of pumping

291

stations, headworks, screens, grit chambers, flow splitters, AS

292

tanks, clarifiers and digesters. Undoubtedly, one of the great

293

opportunities of CFD modelling of ASPs is the analysis of the

294

multiphase flow behaviour and prediction of the impact of a

with

aerospace

and

mechanical

AC C

EP

TE D

M AN U

SC

RI PT

associated

295

wide range of operating parameters on the local scale

296

phenomena, such as flow field coupled with interfacial mass

297

transfer and chemical reaction. In addition to that, CFD has

298

gained popularity over traditional wastewater treatment

299

modelling approaches, as it is a high-precision technique

300

allowing evaluation of the engineering systems, which are

301

expensive, difficult or even dangerous to reproduce in

16

ACCEPTED MANUSCRIPT laboratory-scale, pilot-scale or field conditions. Therefore,

303

CFD-aided modelling can be used as a robust tool for the

304

design of a new facility or the optimization or retrofitting of

305

existing ASPs, leading to enhanced performance and energy-

306

optimised operation of the utility, facilitating time, economic

307

cost and manpower savings (De Gussem et al. 2014, Do-Quang

308

et al. 1999, Essemiani et al. 2004, Guimet et al. 2004, Laurent

309

et al. 2014).

310

4.1. Hydrodynamics – RANS and URANS

311

Various options exist for the numerical simulation of the

312

turbulent flow with CFD codes.

313

In most engineering practice, time-averaged properties of the

314

flow are able to provide the required information. Steady

315

Reynolds Averaged Navier-Stokes (RANS) simulations and

316

unsteady RANS (URANS) focus on the representation of the

317

effects of turbulence on the mean flow properties by solving

318

transport equations for the averaged flow quantities with whole

319

range of the turbulent scales being modelled. Thus this

AC C

EP

TE D

M AN U

SC

RI PT

302

320

modelling approach greatly reduces required computational

321

effort and resources, and is widely adopted for practical

322

engineering applications, including the wastewater sector

323

(Karpinska Portela 2013).

324

In RANS and URANS the flow patterns within the AS tank are

325

obtained from the solution of nonlinear partial differential

326

equations, expressing balances of mass and momentum. 17

ACCEPTED MANUSCRIPT 327

Therefore, the flow is governed by the following mass

328

conservation equation: .' =0 . 

RI PT

329

(1)

and momentum conservation equation, which for RANS is:

330

(2)

M AN U

. + <−'  = . 

SC

. .̅ . . . 2 . 1'  2 = − + 5% 6 + − 9 :; .  .  .  .  .  3  . 

and for time dependent, transient flow (URANS) is:

TE D

.̅ . + 1'  2 . . 

=−

.̅ . . . 2 . + 5% 6 + − 9 :; .  .  .  .  3  .  (3)

. <−'  = . 

EP AC C

+

331

332 333 334

where ̅ is the averaged pressure field, ' and % are the fluid

density and viscosity, respectively,  is the time,

,



and



are the spatial coordinates,  ,  and  are the velocity components and 9 is the Kronecker delta.

18

ACCEPTED MANUSCRIPT 335

Here, the turbulent mean velocity field is described by

336

Reynolds decomposition of the velocity using time averaged

337

term, ̅ , and a fluctuating term,  :  = ̅ +

RI PT

(4)

RANS and URANS equations are linearized and solved. The

339

flow structures originating from the momentum transfer by the

340

fluctuating velocity field, which are smaller than the numerical

341

grid discretization and represented by the term

342

namely Reynolds stresses, are unclosed, and thus they must be

343

modelled.

344

Turbulence models

345

The

346

commercial CFD software suites, which are used for Reynolds

347

stresses closure are: standard, renormalized group (RNG) and

M AN U models

included

in

popular

TE D

turbulence

realizable − $ models; standard and shear stress transport

(SST) − ( models; and Reynolds stress model (RSM)

AC C

349

principal

−'  ,

EP

348

SC

338

350

(Bridgeman et al. 2009).

351

The two-equation models (i.e. − $ and − (), solve

352

additional transport equations for two turbulence quantities:

353 354 355 356

velocity scale- turbulent kinetic energy, , and turbulence length scale- either its dissipation rate, $, or the specific

frequency, ( (Pope 2000). From among the two-equation

models, the standard − $ (> − $) has found the broadest 19

ACCEPTED MANUSCRIPT range of applicability in both, academia and industrial sectors,

358

due to its robustness, relatively low computational requirements

359

and satisfactory accuracy. The characteristic feature of all two-

360

equation models is that modelling of the Reynolds stresses

361

employs the Boussinesq hypothesis relating these stresses to the

362

mean deformation rates and thus mean velocity gradients, and

363

assuming locally isotropic turbulence (Rodi 1993):

RI PT

357

SC

. . 2 . −'  = %& 6 + : − ?' + %& @9 .  .  3 .  

365

where %& is the turbulent viscosity computed as a function of

M AN U

364

(5)

and $:

TE D

%& = 'AB

C $

(6)

where AB is a model constant (variable function in a different

367

turbulence model).

368

A comprehensive description of the > − $ and the relatively

369

recently developed improved models, namely Renormalized

AC C

EP

366

370

Group (RNG) − $ and realizable − $, can be found in the

371

literature (Launder and Spalding 1974, Orszag et al. 1996, Shih

372

et al. 1995).

373

The second widely used two-equation model, introduced by

374

Wilcox, is the − ( model, also based on the isotropic eddy

20

ACCEPTED MANUSCRIPT

376 377

viscosity hypothesis described by Equation (5), but where the

turbulent viscosity is computed as a function of and (.

Knowing, that ( = $ ⁄ , Equation (6) takes the following form: %& = '" ∗

(

(7)

RI PT

375

where " ∗ is a coefficient related to the use of functions

379

damping the turbulent viscosity causing a low-Reynolds-

380

number correction.

381

Contrary to the − $ model, the standard − ( model uses

382

enhanced wall treatment to solve low-Re-number flows in the

383

viscous layer in near-wall region. Its improved variant, SST

M AN U

SC

378

− ( model, considered to be the most accurate from two-

385

equation eddy viscosity models, provides modified turbulent

386

viscosity formulation which account for the transport of the

387

principal turbulent shear stress.

TE D

384

A wider discussion on the

standard and SST − ( models can be found in Wilcox

389

(Wilcox 1998, Wilcox and Traci 1976) and Menter (Menter

390

1993, 1994).

391

The Reynolds Stress Model (RSM) (Launder et al. 1975,

392

Versteeg and Malalasekera 1995) is the most elaborate and

393

complex turbulence model, referred as the Second Order

394

Closure. In RSM the isotropic eddy viscosity hypothesis is

395

discarded and the RANS equations are closed by solving

396

transport equations for the Reynolds stresses, together with an

AC C

EP

388

21

ACCEPTED MANUSCRIPT equation for the dissipation rate, yielding seven additional

398

transport equations to be solved in a 3D scheme. A detailed

399

description of the RSM turbulent closure development can be

400

found in the literature (Launder et al. 1975), whilst a

401

summarised description of the turbulence models commonly

402

used for RANS closure has been summarized in Table 1

403

(Bridgeman et al. 2009).

AC C

EP

TE D

M AN U

SC

RI PT

397

22

ACCEPTED MANUSCRIPT

Table 1. Comparison of turbulence models used in ASP modelling. Adapted from Bridgeman et al. (2009) with permission from Taylor &

405

Francis Ltd (www.tandfonline.com/) Standard E − F (GE − F)

Comments

Renormalized Group (RNG) E−F

• Based on the statistical methods, not observed

Advantages

• Semi-empirical modelling of and $. • Valid only for fully developed turbulent flow cores (molecular diffusion ignored).

• Simplest and complete turbulence model. • Excellent performance for many flows. • Well established in academia and industry. • Robust, economic in terms of computational effort and satisfactory accuracy in diverse turbulent flow issues.

• • •

TE D

• Improved performance for swirling

and high-strained flows compared to > − $.

EP



fluid behaviour. Mathematics is highly abstruse. Texts only quote model equations which result from it. Effects of small-scale turbulence represented by means of random forcing function in NavierStokes equations. Procedure systematically removes small scales of motion by expressing their effects in terms of larger scale motions and a modified viscosity. Similar in form to > − $, but modified $ equation to describe high-strain flows better. Differential equation solved for %& (changes AB from 0.09 to 0.0845 at high Re).

AC C



M AN U

SC

Model

RI PT

404

406

23

Disadvantages • Poor performance in some scenarios (strong streamline curvature, vortices, rotating flows, flow separation, adverse pressure gradients). • Assumes locally isotropic turbulence. • Poor prediction of the lateral expansion in 3D wall jets.

• Less stable than > − $.

ACCEPTED MANUSCRIPT

Table 1. (Continued). Advantages • Suited for planar and rounded jets, swirling and separating flows and wall-bounded flows with strong adverse pressure gradients.

Standard E − H

• Recent development. In highly strained flows, the normal Reynolds stresses become negative (unrealizable condition), so %& uses variable AB . • AB is function of local strain rate and fluid rotation. • Different source and sink terms in transport equations for eddy dissipation. • Good for spreading rate of round jets.

• Specific dissipation rate is ( = $/ . • > − $ solves for dissipation of turbulent kinetic energy, − ( solves for rate at which dissipation occurs. • Resolves near wall region without wall functions so can be applied through boundary layer.

408

SC

M AN U

• As − ( except from gradual change from

− ( to inner region of boundary layer to high Re version of > − $ in outer part. • Modified %& formulation to account for transport effects of principal turbulent shear stresses.

• Not recommended to use with multiple reference frames.

• Valid throughout to boundary layer, subject to fine grid resolution. • Accounts for the stream-wise pressure gradients. • Applicable for detached, separated flows and fully turbulent flows.

• Pressure induced separation is typically predicted to be excessive and early.

• The most accurate from twoequation eddy viscosity models. • Suitable for adverse pressure gradients and pressure-induced flow separation. • Accounts for the transport of the principal shear stresses.

• Less suitable for free shear flows.

TE D

Shear Stress Transport (SST) E − H

Disadvantages

RI PT

Realizable E − F

Comments

EP

Model

AC C

407

409 410 24

ACCEPTED MANUSCRIPT

Table 1. (Continued). Comments

Advantages

Reynolds Stress Model (RSM)

• The most general and complex of all models solving transport of the Reynolds stresses- so called seven-equation model • Isotropic eddy viscosity hypothesis is discarded.

• Accurate calculation of the mean flow properties and all Reynolds stresses. • Accounts for the streamline curvature, rotation and rapid changes in strain rate yielding superior results to two-equation models for complex flows, e.g. with stagnation points.

412

SC

AC C

EP

TE D

413 414

Disadvantages

RI PT

Model

M AN U

411

25

• Computationally expensive. • Not always more accurate than two-equation models. • Harder to obtain converged result. • Reliability of RSM predictions are still limited by the closure assumptions employed to model pressure-strain and dissipationrate terms.

ACCEPTED MANUSCRIPT 4.1.1. Alternative approaches - LES and DNS

416

Large Eddy Simulation (LES) is an alternative approach in

417

hydrodynamic modelling where large, three-dimensional

418

unsteady scale motions affected by the flow geometry, i.e. large

419

eddies, are directly and explicitly solved in time-dependent

420

simulation using space-filtered Navier-Stokes equations. LES is

421

one of the most expensive simulation options and requires

422

a refined grid to accurately resolve eddies in the boundary

423

layer. A filtering operation, analogous to the Reynolds

424

decomposition in RANS, is based on the decomposition of the

SC

M AN U

425

RI PT

415

velocity into the resolved (filtered) component ̅ J , K and the residual, so called subgrid-scale (SGS) component   J , K

427

(Pope 2000). The accuracy of the LES model is the result of

428

modelling only the SGS motions- the smallest eddies, which

429

tend to have more universal properties (Karpinska Portela

430

2013). The model commonly used to model small eddies is the

431

Smagorinsky SGS model (Smagorinsky 1963).

432

LES has been most successful for high-end applications where

AC C

EP

TE D

426

433

the RANS models fail to meet the required goals, e.g.

434

modelling of combustion and mixing. Although it provides

435

improved accuracy, wide application of LES approach to solve

436

flow related issues is still limited due to the large mesh

437

requirements and high computational costs.

438

Direct Numerical Simulation (DNS) is a high-fidelity tool

439

offering the explicit solution of the whole range of turbulent 26

ACCEPTED MANUSCRIPT time and length scales, from the Kolmogorov scales to large

441

motion scales transporting most of the kinetic energy within the

442

domain (Orszag 1970). As a result, the computational cost of

443

DNS is extremely high even at low Reynolds numbers,

444

preventing this approach from being used as a general-purpose

445

design tool, and making it impractical for most industrial flow

446

conditions, especially large multiphase AS systems. Hence the

447

CFD modelling of the AS tanks requires more computationally

448

efficient and thus simplified methods.

449

4.2. Multiphase Modelling

450

A robust understanding of the physics and biochemical

451

processes in the gas-liquid-solid environment of the AS tank

452

relies on accurate assessment of the transport phenomena and

453

character of interactions between the phases. Numerical

454

simulation of the multiphase flow in AS system is usually

455

enabled by Euler-Euler or Euler-Lagrange approaches. Table 2

456

outlines the relevant characteristics of the multiphase models

457

used in modelling of the AS systems.

AC C

EP

TE D

M AN U

SC

RI PT

440

458

In Eulerian (Eulerian-Eulerian or two-fluid) model, the phases

459

are treated mathematically as separate and interpenetrating

460

continua, hence the introduction of the phasic volume fractions-

461

" . The sum of volume fractions is equal to unity

27

ACCEPTED MANUSCRIPT M

L " = 1

(8)

NO

Each phase is governed by a set of continuity and momentum

463

conservation equation, which have similar structure for all

464

phases.

Thus,

Q-phase

considering

system,

RI PT

462

the

mass

conservation equation for phase  is (Joshi 2001, Ratkovich

466

2010): M

SC

465

M AN U

. 1" ' 2 + ∇1" ' S 2 − L1 −  2 = 0 .   NO

467 468

(9)

where ' is density and the term " ' is the effective density of the phase , S denotes its velocity,  and  are mass

transfer mechanisms from phase  to  and from  to ,

470

respectively.

471

The generalized momentum conservation equation for phase 

472

can be written in the simplified form as (Azzopardi et al. 2011,

AC C

EP

TE D

469

473

Ratkovich 2010): . 1" ' S 2 + ∇1" ' S S 2 .   

= −" ∆ + ∇% 1∇" S + ∇" SU 2 + ' VS + S

28

(10)

ACCEPTED MANUSCRIPT

474

where  is pressure shared by all phases, % denotes shear viscosity of phase , S represents the sum of interfacial forces

476

between the continuous and dispersed phases.

477

Considering gas-liquid system, the fluid flow around the

478

bubbles is characterized by the occurrence of relative motion

479

between the phases yielding local pressure and shear stress

480

gradients. As a result, relative motion of the bubbles will be

481

affected by a drag force which is predominant in conditions of

482

the uniform flow. In case of non-uniform bubble motion, the

SC

RI PT

475

concept of interfacial forces S needs to account for drag and

484

various non-drag forces, such as virtual mass force, lateral lift

485

force, wall lubrication force, turbulent dispersion force, Basset

486

force and momentum transfer associated with mass transfer.

487

Accordingly, the closure of Equation (11) requires correct

488

assessment of the interfacial forces, typically using analytical

489

models, empirical correlations and coefficients (e.g. drag

490

coefficient), described comprehensively in Clift et al. (1978),

491

Joshi (2001), Azzopardi et al. (2011) and Yang and Mao

492

(2014). Many of these relations are case-specific and based on

493

limited data what yields difficulties in exploring complex

494

multiphase systems (Azzopardi et al. 2011, Manninen et al.

495

1996) and influencing the outcomes of the CFD simulations.

496

The Eulerian model is commonly used in multiphase systems,

497

where momentum exchange between the phases is significant,

AC C

EP

TE D

M AN U

483

29

ACCEPTED MANUSCRIPT e.g. gas-liquid flow in aeration tanks or solid-liquid flow in AS

499

systems, especially in the case of zones with high solids

500

concentrations, where solid-solid interactions are relevant;

501

transitional zones with steep solids concentrations gradients

502

(from high to low concentration), where momentum of the solid

503

phase is relevant to model its dissipation (Samstag et al. 2015).

504

The Volume of Fluid (VOF) model is a single-fluid approach

505

based on a surface tracking technique that solves the

506

momentum equations for the continuous phase while the

507

dispersed phase follows the closure conditions of the volume

508

fraction for the incompressible flow (Wang et al. 2013). While

509

the phases are immiscible, the fields for all variables and

510

properties are shared and represented by volume-average

511

values. Therefore, the continuity equation has the following

512

form (Ratkovich 2010, Vedantam et al. 2006):

TE D

M AN U

SC

RI PT

498

EP

." + ∇1" S 2 = 0 .

(11)

where the computation of the primary-phase volume fraction is

514

based on the constraint defined by Equation (9).

515

The single momentum equation, which is solved throughout the

516

domain yielding velocity field shared by the phases, is

517

expressed as:

AC C

513

30

ACCEPTED MANUSCRIPT . J'SK + ∇J'SSK = −∆ + ∇W%J∇S + ∇S U KX + 'VS + S (12) .

518

The properties ' and % appearing in the transport equations are

519

determined by the presence of the component phases in each

control volume. Considering two-phase ( and ) system if the

RI PT

520

volume fraction of phase  is being tracked, the density in each

522

cell is given by (Ratkovich 2010, Vedantam et al. 2006):

523

M AN U

' = " ' + 11 − " 2'

SC

521

(13)

The relationship described by Equation (13) is based on the

524

fact, that for an Q-phase system, the volume-fraction-averaged

525

density is

(14)

TE D

' = L " '

All other properties (e.g volume-fraction-averaged viscosity, %)

527

are computed in the same manner.

528

VOF is designed for modelling multiphase systems where the

AC C

EP

526

529

position of the interface between the immiscible fluids is of

530

interest. This approach has found application in modelling of

531

stratified and free-surface flows, filling, sloshing, and motion

532

of large bubbles (slug flow). Examples of VOF application to

533

model Membrane Bioreactors (MBRs) can be found in the

534

literature (Andersson et al. 2011, Ratkovich et al. 2009, Wang

535

et al. 2013). 31

ACCEPTED MANUSCRIPT The Eulerian-algebraic (slip mixture or algebraic slip) model is

537

a simplified multiphase model used for two or more phases,

538

treated as interpenetrating continua. This single-fluid approach

539

can be used to model phases moving at different velocities- by

540

using concept of slip (or drift) velocities.

541

Here, the continuity and momentum equations are solved for

542

the mixture and algebraic equations are used to solve relative

543

velocities to describe the dispersed phases. Thus, the continuity

544

equation for the mixture is (Manninen et al. 1996):

M AN U

SC

RI PT

536

.' + ∇J' S K = 0 .

545

where the mixture density, ' is defined as: M

TE D

' = L " ' NO

AC C 547

(16)

and the mass-averaged mixture velocity S is:

EP

546

(15)

M

1 S = L " ' S ' NO

(17)

The mass fraction of phase  is defined as:  =

" ' '

(18)

548

The momentum equation for the mixture is obtained from the

549

following formula (Ratkovich 2010):

32

ACCEPTED MANUSCRIPT . J' S K + ∇J' S S K .  

U KX = −∆ + ∇W% J∇S + ∇S + ' VS + S

NO

550

where % is mixture viscosity equal to: M

%  = L " % 

M AN U

NO

SC

+ ∇ YL " ' S, S, Z

(19)

RI PT

M

(20)

551

and S, is drift velocity for the secondary phase  expressed

552

as follows:

(21)

TE D

S, = S − S

The slip velocity of the secondary phase (K relative to the

554

velocity of the primary phase (K is computed as follows:

AC C

EP

553

555

S = S − S

(22)

The relation between drift and slip velocities can be written as: M

S, = S − L  S NO

(23)

556

The basic assumption of the algebraic slip mixture model is that

557

a local equilibrium between the phases should be reached over

558

a short spatial length scale. For the phases moving with the 33

ACCEPTED MANUSCRIPT same velocity, the mixture approach can be used to model

560

homogeneous multiphase flow (Manninen et al. 1996, Wicklein

561

and Samstag 2009). However, this approach is usually not

562

recommended in cases when the interface laws are not well

563

known (Ratkovich 2010). This model has been successfully

564

used to simulate sediment-induced density currents, and thus

565

solid-liquid mixing in AS tanks, and turbulent transport of

566

suspended solids in secondary clarifiers (Samstag et al. 2015,

567

Wicklein and Samstag 2009, Yeoh and Tu 2009), but also a

568

gas-liquid turbulent flow in closed loop bioreactors (Xu et al.

569

2010, Yang et al. 2011).

570

Another method to deal with multiphase flow in AS systems is

571

via the modelling of turbulent transport of the secondary phase

572

within a continuous primary phase. The transported phase

573

(species) is treated as an active or passive scalar (density,

574

viscosity, temperature, dissolved oxygen or solids), hence the

575

effects of its gradients across the domain can be either coupled

576

to the momentum equation as an equation of state or treated as

AC C

EP

TE D

M AN U

SC

RI PT

559

577

a passive property transported by the fluid flow.

578

Therefore prediction of local mass fraction of each species, ! ,

579

is modelled by solving conservation equation describing

580

convection and diffusion of the ′-th species (Cartland Glover et

581

al. 2000):

34

ACCEPTED MANUSCRIPT . J'! K + ∇J'S! K = −∇S +  +  .

582 583

(24)

where S is the ′-th diffusive mass flux of species;  is net

rate of production of ′-th specie due to chemical reaction;  is

the source term which injects ′-th specie into the domain by

585

addition from the dispersed phase plus any user-defined

586

sources.

587

The active/passive scalar approach is used to model solid-liquid

588

interactions in the systems with lower solids concentrations

589

(Combest et al. 2011, De Clercq 2003, Samstag et al. 2015,

590

Wicklein and Samstag 2009) and to model oxygen mass

591

transfer in aerated tanks (Fayolle et al. 2007).

592

In the Lagrangian model, the governing phase (fluid) is treated

593

as a continuum by solving time-averaged Navier-Stokes

594

equations (Eulerian reference frame), while the behaviour of

595

secondary phase (e.g. solids) is predicted by tracking a large

596

number of particles using random-walk Lagrangian trajectory

AC C

EP

TE D

M AN U

SC

RI PT

584

597

calculations for dispersed phase through the flow field of the

598

continuous phase. Here, the particles can exchange momentum,

599

mass, and energy with the fluid phase. The particles’

600

trajectories are predicted by integrating the force balance on the

601

single particle and recording the particle position. This force

602

balance, which equates the particle inertia with the forces

603

acting on it is defined as (Bridgeman et al. 2009): 35

ACCEPTED MANUSCRIPT .S S − S VSJ' − 'K = + + S .  '

(25)

604

where S is particle velocity and S is fluid phase velocity, the

605

term

606

particle relaxation time, ' and ' are particle and fluid

_`

represents drag force per unit particle mass,  is

RI PT

\S][ \S^ [

densities, respectively; and S relates to additional forces that

608

are relevant under special circumstances; for example virtual

609

mass and pressure gradient forces, or forces on particles that

610

arise due to rotation of the reference frame.

611

Additionally, particle-particle interactions, such as collisions

612

and momentum exchange can be enabled through the

613

introduction of drag coefficients (Bridgeman et al. 2009,

614

Karpinska Portela 2013). This approach is suitable to simulate

615

gas-liquid flow in aeration tanks, however should not be used

616

when the volume fraction of the secondary (particulate) phase

617

exceeds 10-12% (Sokolichin et al. 1997), as e.g. solid-liquid

618

flow in complete mixed AS systems.

619

Apart from multiphase modelling, particle tracking within the

620

Lagrangian reference frame is a useful tool to simulate RTDs in

621

process tanks (Danckwerts 1953, Levenspiel 1999), and so to

622

assess the macromixing, established by convective flow

623

patterns (Karpinska Portela 2013, Le Moullec et al. 2008).

624

However, the main limitation here is the number of particles

625

being tracked within the simulated system, and in the case of a

AC C

EP

TE D

M AN U

SC

607

36

ACCEPTED MANUSCRIPT 626

full-scale AS tank, tracking too many particles will result in

627

extensive computational

628

workability of different approaches in multiphase modelling,

629

the Lagrangian approach is the most expensive, involving long

630

computational times and requiring a large number of CPUs

631

(Central Processing Units), so limiting its popularity in the

632

simulation of wastewater treatment tanks.

when

comparing

SC

633

Thus,

RI PT

times.

AC C

EP

TE D

M AN U

634

37

ACCEPTED MANUSCRIPT Table 2. Summary of approaches to multiphase modelling. Concept

Modelling

Applicability

Issues

Eulerian

Each phase modeled as a separate fluid.

A set of averaged, volume fractionweighted NavierStokes (NS) equations per phase. Momentum transfer terms and constitutive equations to be modelled.

Theoretically, every type of flow; depending on the additional terms’ modelization.

Additional terms’ modelization is determinant, but their modelling is difficult.

VOF

Each phase modeled as a separate fluid. The interface between the phases is tracked.

Interface tracked via a continuity equation and the domains of the single phases are defined. A set of phase-specific NS equations with momentum exchange terms is solved for each domain.

Flows where the interphase surface is clearly defined (e.g. a single, large bubble inside a liquid).

Inapplicable if the interphase surface is too complex (e.g. bubbly flow where bubble dimensions are smaller than single cell size).

Mixture

Both phases treated as a whole.

Single set of NS equations.

Homogeneous fluids or nonhomogeneous fluids that are treated as homogeneous.

Inapplicable in every case in which there is clear distinction between the phases.

Flows where at least one phase is clearly dispersed into a principal phase. Particles smaller than mesh cell size.

Computational expense not a priori computable, and proportional to the particles number. May be prohibitive for high particle numbers.

SC

M AN U

TE D

Liquid phase treated with Eulerian approach. Every particle (bubble) tracked along trajectory.

AC C

EulerianLagrangian

RI PT

Model

Effective mixture density and viscosity to be modelled.

EP

635

A set of averaged NS equations for the liquid phase. A set of Newton’s 2nd Law (N2L) equations applied to all particles. Momentum transfer terms in both NS and N2L to be modeled.

636

38

ACCEPTED MANUSCRIPT 637

5. Application of CFD in ASP

638

Biological wastewater treatment systems require enhanced

639

transfer of oxygen into the AS tanks to maintain aerobic

640

processes

641

Enhanced oxygen mass transfer in most common ASP

642

configurations i.e. channel or closed-loop bubbly flow AS

643

reactors aerated by diffusers, can be achieved by maximisation

644

of the surface area of the interface between the dispersed phase

645

(air/oxygen bubbles) and the continuous phase (mixed liquor).

646

Nonetheless, in most wastewater utilities, designs of AS tank

647

and aeration system, and furthermore process control are based

648

on static models and simple PID controllers, and so a major

649

concern

650

hydrodynamics and its impact on DO profiles within the basin.

651

Therefore,

652

efficiencies, hydraulic and process design, and reliable

653

operation of AS systems rely on an improved understanding of

654

the bioreactors’ behaviour, with emphasis on micro- and

in

biodegrading-nitrifying

biomass.

the

accurate

TE D

remains

M AN U

SC

RI PT

occurring

of

biochemical

of

tank

conversion

AC C

EP

improvement

determination

655

macro-scale mixing and, in particular, on the analysis of the

656

interactions between the mixed liquor circulation imparted by

657

mechanical agitation and the fluid and sludge flocs motions

658

induced by the air bubbles. The key advantage of CFD as a

659

virtual modelling technique is powerful visualization capability

660

allowing detailed characterization of the local-scale phenomena

661

in varying operating conditions. Hence, CFD can be used as a 39

ACCEPTED MANUSCRIPT robust tool for the design of either new efficient and energy

663

optimised unit processes at WWTPs or for “tune for benefit”

664

optimization of performance and even retrofitting of existing

665

AS systems.

666

The complete CFD simulation of a biologically active gas-

667

liquid-solid AS system is challenging, due to the complex

668

hydrodynamics and biochemical reactions of conversions

669

involved, resulting in massive computational resources required

670

in terms of RAM and CPU usage and with long computational

671

times involved. Furthermore, increasing complexity of the

672

model involved, mesh resolution and solution accuracy may

673

lead distinctly higher costs of the CFD analysis, yet lower than

674

a capital investment in a new layout. To avoid overly long and

675

complex CFD runs (Karpinska Portela 2013, Le Moullec et al.

676

2010b), common engineering practice has been to model the

677

AS processes separately (e.g. aeration system performance or

678

flow field within the basin) and afterwards to couple the results

679

(Pereira et al. 2012). In addition to that, depending on the

AC C

EP

TE D

M AN U

SC

RI PT

662

680

purpose of the CFD analysis, data collection for model

681

calibration and validation using advanced measurement

682

techniques, may be absolutely required and yet time and

683

resource consuming (Nopens et al. 2012).

684

It is worth mentioning, that despite the current successful

685

spread and usage of CFD modelling, potential risk of its misuse

40

ACCEPTED MANUSCRIPT due to poor model choice, wrong setup assumptions or

687

interpretation of the results, has been also recognized (Nopens

688

et al. 2012).

689

5.1. Development of New Aeration Devices

690

CFD simulations have been used as a high-tech design tool in

691

the development of new aeration devices and optimisation of

692

their spatial arrangement in wastewater treatment tanks.

693

A computationally inexpensive modelling procedure, based on

694

3D steady-state and single-phase flow simulations with the

SC

RI PT

686

> − $ turbulence model was used by Morchain et al. (2000) to

696

study the impact of the spatial distribution of the cross-flow

697

hydro-ejectors on the recirculation and the oxygen mass

698

transfer in a large tank for wastewater treatment. Although the

699

cross-flow hydro-ejectors generate two-phase flow, it was

700

shown experimentally that the momentum transfer is not

701

significantly affected by the presence of bubbles and thus the

702

velocity field could be obtained using a single-phase model,

703

while the oxygen transfer in the tank was enabled by the

AC C

EP

TE D

M AN U

695

704

introduction of transport species.

705

The same single-flow modelling scheme was applied to modify

706

geometry and operating scenarios of a curved blade mechanical

707

aerator intended for use in oxidation ditches (Bhuyar et al.

708

2009). The results, which were in good agreement with the

709

experimental data, allowed optimization of blade design and

710

the aerators’ submergence, and reductions to rotational speed 41

ACCEPTED MANUSCRIPT 711

range, yielding enhanced aeration efficiencies when compared

712

with conventional mechanical devices.

713

Despite higher computational costs, transient gas-liquid simulations involving URANS with the > − $ turbulence

715

model and one of the Euler-Euler multiphase models have

716

found applications in the development of new aeration

717

techniques, as they provide direct and more accurate analysis of

718

the multiphase reactor systems. Xu et al. (2010) simulated an

719

oxidation ditch equipped with cylindrical airlift aerators. Here

720

the CFD simulations of the fluid flow in an airlift oxidation

721

ditch served to verify the feasibility of the preliminary design

722

and to assess its applicability for municipal wastewater

723

treatment. The results from the simulations (which were

724

validated in a bench-scale ditch) emphasized the suitability of

725

the

726

configurations.

727

Karpinska Portela (2013) used CFD to focus on overcoming

728

oxygen transfer efficiency limitations of membrane diffusers

729

via the introduction of an independent external aeration unit,

730

designed as a continuous flow component included in the

731

mixed liquor recirculation loop. Similar to the previous work,

732

CFD

733

configurations provided an excellent design tool for selection of

734

the most efficient geometry, characterized by the highest value

TE D

M AN U

SC

RI PT

714

aeration

system

for

deep

tank

ditch

AC C

EP

proposed

simulations

of

aeration

42

within

several

device

ACCEPTED MANUSCRIPT of SOTE. The aeration process parameters determined from

736

pilot-scale aeration tests matched the predictions obtained from

737

the CFD simulations.

738

A more detailed description of the CFD approaches used in the

739

development of aeration devices is shown in Table 3.

RI PT

735

740

AC C

EP

TE D

M AN U

SC

741

43

ACCEPTED MANUSCRIPT Table 3. CFD studies of new aeration devices. Author

Code

Device

Scale

Dim.

Turbulence Extra model model

Bhuyar et al. (2009)

Fluent

Curved blade mechanical aerator

Lab

3D

RANS + > − $

Xu et al. (2010)

Fluent

Airlift aerator

Lab

3D

URANS + > − $

Karpinska Portela (2013)

Fluent

Pressurized Lab Aeration Chamber

3D

URANS + > − $

743 744 745 746

750 751 752 753 754 755

EP

749

AC C

748

TE D

747

44

Validation measurements

RI PT

Transport species + chemical reaction Mixture

SC

M AN U

742

Mixture

Particle Dynamic Analysis (PDA)

measurements

ACCEPTED MANUSCRIPT 756 757

5.2. Evaluation of the performance of the existing aeration systems Robust and energy efficient operation of complex AS reactor

759

systems relies on an understanding of the multiphase nature of

760

the AS tank and on the assessment of the impact of the aeration

761

system on the mixing, biological treatment efficiency and

762

associated energy expenditure, while taking into account the

763

number, type and spatial distribution of the aeration and mixing

764

devices. This section considers CFD approaches used to

765

determine the dynamic behaviour of aeration tanks, and for

766

which descriptions are summarized in Table 4.

767

5.2.1. Gas-liquid Models

768

When considering fluid flow within an aeration tank, the use of

769

gas-liquid

770

straightforward prediction of the multiphase velocity field

771

induced by the aerators and mixers, bubble sizes and local gas

772

hold-up, i.e. process parameters that have significant impact on

773

the oxygen mass transfer in bioreactors, but either require time-

774

and resource-expensive experimental techniques or are not

775

taken into account in the analysis of the experimental results

776

(Fayolle et al. 2007, Gillot et al. 2005, Gillot and Héduit 2000,

777

Le Moullec et al. 2008, Vermande et al. 2007).

778

Hence the focus of early CFD works was on a better

779

understanding of mass transfer phenomena in full scale closed-

TE D

M AN U

SC

RI PT

758

models

enables

relatively

fast

and

AC C

EP

CFD

45

ACCEPTED MANUSCRIPT loop aeration tanks equipped with membrane diffusers and

781

agitated by means of horizontal slow speed impellers (Cockx et

782

al. 2001, Do-Quang et al. 1999). 2D two-phase flow in the

783

aeration tank was simulated using an Eulerian model for

784

imposed, constant gas bubble size. To simplify numerical

785

simulations, an equivalent uniform liquid velocity field was

786

imposed in the section of the impellers. The oxygen mass

787

transfer was obtained via introduction of the global volumetric

RI PT

780

mass transfer coefficient, , determined experimentally, as a

789

transport source term. Here, the correct estimation of the

790

absolute values of oxygen transfer requires assessment of the

791

most uncertain aeration process parameter, i.e. the alpha-factor,

792

quantifying the impact of the contaminants and process

793

conditions on the oxygen transfer rates- in this work- the

794

impact of the imposed constant bubble diameter on the gas-

795

liquid interface surface area (Nopens et al. 2015, Rosso et al.

796

2011, Rosso and Stenstrom 2006).

797

The

EP

TE D

M AN U

SC

788

from

the

simulations

showed

that

the

AC C

results

798

hydrodynamics of these tanks was controlled by mutual

799

competition between vertical gas plume and horizontal fluid

800

flow currents. Thus, in the absence of the horizontal flow

801

motion, the interactions between gas plumes released by

802

diffusers generate vertical, massive liquid loop circulations, i.e.

803

spiral flow, yielding low gas hold-up. Contrary to that, under

804

conditions of horizontal liquid motion imparted by e.g. rotating

46

ACCEPTED MANUSCRIPT impellers,

the

increase

in

wastewater

velocity causes

806

neutralization of the spiral flow patterns, inclination of the

807

vertical gas plume, and dispersion of the bubbles and, as a

808

consequence, longer contact times between the phases yielding

809

better gas retention in the tank. As a result, the increase of the

810

global mass transfer coefficient in the closed-loop tanks was

811

found to be associated exclusively with fluid circulation.

812

The impact of the fluid velocity on oxygen mass transfer was

813

later confirmed in experimental studies (Abusam et al. 2002). It

814

was reported that even small changes in the axial velocity may

815

have a dramatic effect on the oxygen profiles, and thus on the

816

ammonia removal in oxidation ditches.

817

closed-loop AS systems, horizontal fluid velocity should be

818

treated as an important process variable to control total nitrogen

819

removal efficiency.

820

A clear shortcoming of the discussed work (Cockx et al. 2001,

821

Do-Quang et al. 1999) is that the 2D modelling is insufficient

822

to represent mutual interactions between the neighbouring gas

M AN U

SC

RI PT

805

AC C

EP

TE D

Therefore in these

823

plumes causing formation of the more complex, three-

824

dimensional flow structures. Moreover, the impact of the flow

825

circulation on the size of the bubbles released by membrane

826

diffusers was neglected; yet this plays a key role in the oxygen

827

mass transfer within the aeration tank. Nevertheless, despite

828

these weaknesses and the lack of experimental validation,

829

conclusions from this work provided valuable research

47

ACCEPTED MANUSCRIPT background for many successive numerical and experimental

831

studies.

832

Subsequent studies (Fayolle et al. 2007) addressed development

833

of a more complete numerical tool, to predict oxygen mass

834

transfer in a number of pilot- and full-scale oxidation ditches of

835

various configurations and where aeration is dissociated from

836

mixing by the introduction of membrane diffusers and slow

837

speed mixers. The aeration was simulated using Eulerian model

838

for water and air. Mixing was modelled using fixed value

839

method by imposing flow characteristics induced by agitation

840

on the grid zones corresponding to mixers location. This

841

method gives good prediction of the average experimental axial

842

velocity. Similar to earlier CFD works (Cockx et al. 2001, Do-

843

Quang et al. 1999), Fayolle et al. (2007) showed that oxygen

844

mass transfer in closed-loop AS basins aerated with diffusers is

845

linked to the number, type, spatial distribution and performance

846

of the agitators (Fayolle et al. 2007, Fayolle et al. 2010, Fayolle

847

et al. 2006, Vermande et al. 2007),

The impact of the

AC C

EP

TE D

M AN U

SC

RI PT

830

848

ascending bubbles released by the diffusers on the velocity

849

profile along the oxidation ditch can be seen in Figure 1. It was

850

also showed that implementation of more complex modelling

851

approach accounting for the impact of the bubble size on the

852

axial fluid velocities and on the global gas hold-up profile

853

within the tank, was able to reproduce precisely the values of

854

with high accuracy (≤ 5%). Nonetheless, it was found, that 48

ACCEPTED MANUSCRIPT the predicted accuracy depends on the assumed value of inlet

856

bubble size, emphasizing the necessity either to measure in situ

857

bubble diameter or apply an appropriate numerical model

858

estimating bubble sizes at the diffuser level. A shortcoming of

859

Fayolle et al. (2007) is that although the proposed CFD

860

protocol appears to be suitable for the prediction and

861

optimization of oxygenation capacities in a full scale AS tank,

862

it is based on clean water-air simulations. However, for robust

863

analysis of agitation system performance, the impact of the

864

velocity on the local solids content and density gradients should

865

be considered.

866

Yang et al. (2011) focused on predicting the flow pattern and

867

oxygen mass transfer in a multichannel oxidation ditch aerated

868

with horizontal rotors and agitated with submerged mixers (a

869

Carrousel). Experimental field data showed that under existing

870

operational conditions, DO concentration in the anoxic zone

871

exceeded 1.0 mg/L, inhibiting the denitrification process. Thus,

872

the aim of the CFD study was to determine an operational

AC C

EP

TE D

M AN U

SC

RI PT

855

873

regime for the surface aerators, which would lead to formation

874

of stronger DO gradients in the anoxic zone, yielding a

875

reduction in energy expenditure for aeration while maintaining

876

high treatment efficiencies. The modelling procedure involved

877

a two-phase mixture approach, enhanced with additional

878

transport equations related to the transfer of oxygen introduced

879

by rotors and biochemical reaction of oxygen consumption for 49

ACCEPTED MANUSCRIPT BOD removal. To reduce computational expenses related to the

881

modelling of a number of rotating devices, the large surface

882

aerators were simulated using moving wall model, while

883

actuator disk (fan) approach was used to simulate submerged

884

mixers. Although an energy efficient operating scheme for the

885

ditches was developed, this study was based on average flow

886

and constant influent quality parameters, thus the validity of the

887

optimal operating conditions deduced from the steady-state is

888

limited.

889

More recent CFD studies on the aeration of conventional AS

890

tanks (Gresch et al. 2011) provided further and more detailed

891

analysis of the flow field induced by porous diffusers. The CFD

892

approach proposed in this work consisted of the gas-liquid

893

Eulerian multiphase model, where the physical properties of the

894

continuous phase were approximated to those of activated

895

sludge. Moreover, instead of solving additional transport

896

equations for oxygen transfer, hydrodynamic simulations were

897

enhanced with a biokinetic model for nitrification. The results

AC C

EP

TE D

M AN U

SC

RI PT

880

898

from CFD studies validated experimentally in a full-scale plug

899

flow AS tank complemented conclusions from earlier work

900

(Cockx et al. 2001, Do-Quang et al. 1999) and provided a

901

comprehensive description of the spiral flow generated by

902

changing diffuser layout, its impact on the velocity field, air

903

hold-up (Figure 2), and ammonia degradation. It was also

904

shown that, contrary to the closed-loop basins, in plug flow AS 50

ACCEPTED MANUSCRIPT lanes without flow boosters, the flow field determines the

906

aeration efficiency and the intensity of the longitudinal mixing,

907

which are significantly reduced by either occurrence of non-

908

aerated zones at the sidewalls or the rolling motion of the fluid

909

generated by changes in diffusers layout.

910

5.2.2. Density-coupled Models

911

In recent years, multiphase modelling of AS tanks based on

912

gas-liquid neutral density simulations has become common

913

practice for evaluation of both aeration and mixing, mainly

914

because it enables faster setup of the lab-scale validation,

915

usually involving use of tap water and air only. However, CFD

916

studies on a sequencing batch reactor equipped with

917

conventional jet aeration system (Samstag et al. 2012) showed

918

that use of neutral density may lead to over-prediction of the

919

degree of mixing. In this work, a density-coupled CFD model

920

incorporating solids settling and transport was calibrated to

921

field data and used to evaluate capacity of the jet aeration

922

system in keeping the solids suspended and to determine power

AC C

EP

TE D

M AN U

SC

RI PT

905

923

consumption for pumping considering two operating modes:

924

mixing with and without air. The authors found that in order to

925

generate complete information about activated sludge mixing,

926

complementary

927

modelling are required to predict local density gradients due to

928

the impact of the flow regime on solids transport. While this

929

work underlines the importance of the analysis of the aeration

CFD

studies

51

based

on

density-couple

ACCEPTED MANUSCRIPT systems with respect to the activation sludge mixing, the

931

approach is rarely used by the wastewater modelling

932

community for AS tanks, and only limited works can be found

933

in the literature (Jensen et al. 2006, Laursen 2007, Samstag et

934

al. , Samstag et al. 1992, Samstag et al. 2012, Wicklein and

935

Samstag 2009).

936

Recently Xie et al. (2014) considered prediction of the mixing

937

and suspended solids distribution in a full-scale Carrousel ditch

938

equipped with surface aerators and submerged impellers. The

939

simulation procedure involved a solid-liquid mixture model,

940

where the sludge settling was coupled through the slip velocity.

941

Similar to the earlier work by Yang et al. (2011), surface rotors

942

and submerged mixers were simulated using moving wall and

943

fan models. The resulting mixing patterns and solids profiles

944

throughout the ditch, shown in Figure 3, facilitated

945

identification of the stagnation regions affected by the sludge

946

settling, and resulted in optimization of the operation scenario.

947

Previously, Fan et al. (2010) simulated a single-channel

948

oxidation ditch aerated by means of inversed umbrella surface

949

aerators using a solid-liquid two-fluid Eulerian model. The

950

operating aerators were simulated using Multiple Reference

951

Frames (MRF) approach. This work also focused on a detailed

952

description of the mixing regime induced by the operating

953

aeration system, aiming to select the optimal range of aerator

AC C

EP

TE D

M AN U

SC

RI PT

930

52

ACCEPTED MANUSCRIPT speed ensuring the most uniform distribution of the solids

955

within the ditch, and thus prevents solids settling.

956

Thus, the literature shows that the use of density-coupling is

957

well founded to assess sludge mixing in agitated AS systems,

958

however its applicability to study bubbly bioreactors is still

959

uncertain.

960

5.2.3. Residence Time Distribution of AS tanks

961

Although rarely used to simulate the aeration process itself,

962

CFD modelling based on the Lagrangian multiphase approach

963

with particle tracking can be used as a tool for evaluation of the

964

aeration system performance through the assessment of the

965

macromixing, and thus the overall transport phenomena in a

966

bioreactor. The literature offers numerous examples of studies

967

on aeration and/or agitation systems responsible for different

968

flow patterns within the tank, and thus having an impact on the

969

mixing at micro- and macro-scale in AS systems. Considering

970

different advection paths of the fluid elements within an aerated

971

continuous flow channel or closed-loop bioreactors for

AC C

EP

TE D

M AN U

SC

RI PT

954

972

wastewater

treatment,

RTD

973

simulations can be used successfully to predict such adverse

974

phenomena as segregation of the flow resulting in short-

975

circuiting (channelling), recycling of the flow, or formation of

976

dead zones. RTD data can also be used to assess the mixing

977

time, which can be of crucial importance for oxygen transfer

978

and impact on chemical and biochemical reaction yield. Data 53

data

obtained

from

CFD

ACCEPTED MANUSCRIPT can also be used for troubleshooting of the reactor and

980

improved design of the future vessels (Brannock 2003,

981

Karpinska et al. 2015, Karpinska Portela 2013, Kjellstrand

982

2006).

983

Glover et al. (2006) considered oxidation ditches aerated by

984

bottom diffusers and agitated by slow-speed rotors and

985

determined reactor model structure by implementation of a

986

CFD-generated RTD curve in Fluent into WEST® software

987

environment based on the systemic approach. The fluid flow in

988

the ditch was obtained from the Eulerian gas-liquid model, and

M AN U

SC

RI PT

979

the turbulence simulated with > − $ model. It was found that

990

the RTDs of the ditch behaviour can be approximated by 90

991

CSTRs (plug flow) for non-aerated conditions and by 20

992

CSTRs (perfectly mixed) for aerated conditions (Glover et al.

993

2006). The work showed that closed-loop reactors with

994

different internal recycling rates, such as oxidation ditches, can

995

be modelled by one or series of CSTRs, the number of which

996

can be determined from CFD-RTD studies. In this way, a

AC C

EP

TE D

989

997

suitable reactor model structure for ASM studies can be

998

predicted.

999

Le Moullec et al. (2008) considered the assessment of RTDs in

1000

a cross-flow channel reactor aerated with porous tube, in which

1001

the hydrodynamics was modelled using a gas-liquid Eulerian

1002

model and the turbulence models used were > − $ and RSM. 54

ACCEPTED MANUSCRIPT It was shown that in the channel-type reactors for wastewater

1004

treatment, turbulence has a dominant role in axial dispersion

1005

(90%), while dispersion due to convection is negligible.

1006

Furthermore it was shown that only the RTDs obtained from

1007

the RSM model are in good agreement with the experimental

1008

data treated by curve fitting for a plug flow with axial

RI PT

1003

dispersion model, while the > − $ model underestimated the

1010

value of dispersion coefficient by around 50%, as seen in

1011

Figure 4. This particular case is the effect of the default

1012

constants in Lagrangian stochastic particle motion model,

1013

which are determined by different closure assumptions. Thus,

M AN U

SC

1009

both models, > − $ and RSM, will use different AB which are

1015

directly linked to the modelling of the turbulent dispersion,

1016

what emphasizes the need for careful attribution of values to

1017

parameters, through the proper model calibration.

1018

Macromixing assessment is a useful tool to describe actual

1019

reactor performance. CFD studies assessing the impact of

1020

turbulence model selection on the RTDs of oxidation ditches

AC C

EP

TE D

1014

1021

aerated with slot injectors (Karpinska Portela 2013, Pereira et

1022

al. 2012), considered the fluid flow simulated with RANS and

1023

URANS with the > − $ model and also LES with

1024

Smagorinsky’s SGS model. This work shows the limitations of

1025

some approaches in computation of the RTD, as the turbulence

1026

model involved to simulate hydrodynamics has a large impact

1027

on prediction of the tracer’s trajectories. The RTD simulations 55

ACCEPTED MANUSCRIPT based on the average flow field, RANS and URANS, led to

1029

overestimation of channelling effects. The flow dynamics

1030

underlies mixing at all scales, both macro- and micro-, and thus

1031

this should always be accounted for. Furthermore, is has been

1032

shown that CFD data for AS modelling must align with all

1033

dynamic components of the flow, thus LES simulations, which

1034

demand more computational resource, should be used for some

1035

specific purposes.

1036

5.2.4. Full model: CFD-ASM

1037

Few workers have focused on the development of a complete

1038

three-phase CFD model to simulate actual physical-chemical-

1039

biological processes within different AS system configurations

1040

(Glover et al. 2006, Le Moullec et al. 2010a, b, Le Moullec et

1041

al. 2011, Lei and Ni 2014). Here, the integration of

1042

hydrodynamics with biokinetics is accomplished by embedding

1043

an ASM model (e.g. ASM1) in a CFD model through

1044

introduction of the additional species of transport with source

1045

terms comprising bioreaction rates. The CFD-ASM simulations

AC C

EP

TE D

M AN U

SC

RI PT

1028

1046

permit the quantification of interactions and transport

1047

phenomena between the water-gas, water-sludge and gas-

1048

sludge phases occurring within the AS reactor and accounting

1049

for the flow field, oxygen mass transfer, growth, decay and

1050

metabolic activity of the AS biomass, and sludge settling.

1051

Therefore it is possible to predict simultaneously the system

1052

hydrodynamics and its impact on the reactions occurring in 56

ACCEPTED MANUSCRIPT nitrifying-denitrifying-biodegrading biomass and represented

1054

by the concentration profiles, in order to optimize aeration,

1055

mixing or system layout leading to efficient process operation

1056

with reduced exploitation costs.

1057

The first attempt at CFD-ASM1 coupling reported in the

1058

literature (Glover et al. 2006), based on the Eulerian gas-liquid

1059

approach, facilitated prediction of poor performance of the

1060

diffused aeration system in a full scale oxidation ditch system,

1061

giving low nitrifying capacity of the sludge. These studies

1062

demonstrated the robustness of the CFD-ASM1 approach;

1063

however it was emphasized, that the validity of this modelling

1064

procedure requires further work to study different AS systems

1065

and aeration/agitation scenarios.

1066

Le Moullec et al. (2010a) provided a robust discussion on the

1067

modelling issues concerning simulations of the hydrodynamics

1068

and biokinetics in a channel AS tank aerated by porous tube

1069

using Eulerian two-fluid model coupled to the ASM1. The

1070

authors discussed the trade-off between a model which can be

1071

implemented and run in the realistic time, and the extent of

1072

simplifications required to facilitate this. The authors assumed

1073

that the channel reactor was in a pseudo steady-state with

1074

constant biomass content; air bubbles with constant diameter;

1075

and the activated sludge flocs being perfectly soluble in the

1076

liquid phase. These simplifications gave rise to a series of

AC C

EP

TE D

M AN U

SC

RI PT

1053

57

ACCEPTED MANUSCRIPT 1077

errors in model output; specifically, incorrect estimation of the

1078

local

1079

overestimation of the ammonia values (wrong estimation of the

1080

autotrophic fraction of biomass in the system).

1081

The conclusions from this work emphasized the importance of

1082

the bubble size distribution for the oxygen mass transfer and for

1083

the molar diffusivity of oxygen in mixed liquor. Furthermore, a

1084

new concept for the approximation of AS floc properties

1085

(highly hydrated solid phase) to a liquid phase, was proposed.

1086

The authors also highlighted that, as the hydrodynamics-

1087

biokinetics coupling requires high number of CPUs and long

1088

computational times, a compromise between mesh resolution

1089

and solution accuracy has to be found.

1090

In more recent work on three-phase CFD-ASM1 simulations of

1091

a Carrousel oxidation ditch aerated and agitated by means of

1092

mechanical aerators and mixers, Lei and Ni (2014) treated the

1093

AS flocs as a pseudo-solid phase. Here, the solid-liquid-gas

1094

flow field was obtained with a three-phase mixture model. The



underestimation

of

the

nitrate

and

AC C

EP

TE D

M AN U

SC

RI PT

value;

1095

results from the simulations, (i.e. velocity, DO, organics and

1096

nutrients profiles) are in good agreement with the validation

1097

data. Crucially, this modelling approach allowed the prediction

1098

of physical kinematics of sludge settling. Concentration maps

1099

of the solids (near the bottom), DO, COD, ammonia and

1100

nitrates (in the mid-depth) in the horizontal cross-section

1101

through the oxidation ditch, are shown in Figure 5. 58

ACCEPTED MANUSCRIPT It can be concluded, that the CFD-ASM1 simulations provide

1103

more reliable results than those obtained from oversimplified

1104

ASM models, which fail to represent the flow dynamics within

1105

a system, particularly regarding DO profiles along an AS basin,

1106

which have an impact on the biological nutrient removal

1107

(Pereira et al. 2009). Thus, the CFD-ASM data can be

1108

considered more suitable for the design and scale-up of

1109

bioreactors. However, besides high computational costs,

1110

another emerging drawback of CFD-ASM modelling approach

1111

is its limited feasibility due to stringent convergence criteria

1112

and equilibrium solution (Nopens and Wicks 2012). As a result,

1113

the procedures of coupling hydrodynamics data obtained from

1114

CFD simulations with the ASM simulations have been also

1115

intensively studied (Glover et al. 2006, Karpinska Portela 2013,

1116

Le Moullec et al. 2010b, Pereira et al. 2012). Here the RTD

1117

curves, actual HRT values, corrected recycle ratio, local

1118

velocities and other hydrodynamics characteristics obtained

1119

from the CFD simulations of an analysed AS system can be

AC C

EP

TE D

M AN U

SC

RI PT

1102

1120

used to generate a suitable reactor model, i.e. in terms of

1121

number of CSTRs in series, recirculation rate and the flow

1122

pattern between each of the reactors, where the ASM can be

1123

implemented.

1124

Nowadays, an alternative modelling approach using a

1125

compartmental model, based on the description of the AS

1126

reactor system using a network of interconnected sections, i.e. 59

ACCEPTED MANUSCRIPT compartments, is emerging. The bi-directional flow rates

1128

between the compartments are computed from the flow field

1129

obtained in the CFD simulations and accounting for the local

1130

values of velocities and mixing due to turbulence. When

1131

comparing with CFD models, compartmental models are

1132

inexpensive in terms of RAM and CPU usage. Nevertheless, as

1133

they are derived from steady-state CFD simulations, the results

1134

must be experimentally validated. In addition, it is still

1135

necessary to develop a more detailed biokinetic model to apply

1136

this approach in modelling of the full-scale industrial

1137

bioreactors, including ASPs (Le Moullec et al. 2010b, Le

1138

Moullec et al. 2011, Nopens and Wicks 2012, Pereira et al.

1139

2012).

M AN U

SC

RI PT

1127

AC C

EP

TE D

1140

60

ACCEPTED MANUSCRIPT

Table 4. CFD modelling of aeration in Activated Sludge tanks. Reference

Code

Scale

Dim. Mesh size

Model

Multiphase model

Extra model

Validation

Numerical modelling of an oxidation ditch aerated with diffusers and agitated by impellers.

Cockx et al. (2001), Do-Quang et al. (1999)

Astrid

Full

2D

n/a

RANS + > − $

Gas-liquid Eulerian

Transport eq. for oxygen

-

CFD studies of the oxidation ditches- coupling of hydrodynamic with biokinetics modelling.

Glover et al. (2006)

Fluent Lab pilot, full

3D

n/a

RANS + > − $

Gas-liquid Eulerian

DO, COD, NH4, NO3 measurements

29k452k

URANS + > − $

Gas-liquid Eulerian

Lagrangian, transport eq. for oxygen + complete biokinetics (ASM1) Transport eq. for oxygen (each phase)

Prediction of the oxygen transfer in oxidation ditches.

Fayolle et al. (2007)

Fluent Pilot, full

3D

RTD of an aerated channel bioreactor.

Le Moullec et al. (2008)

Fluent Lab

3D

350k

Gas-liquid Eulerian

Lagrangian

LDV, tracer experiments

3D

50k

RANS + > − $/RSM

Simulation of the hydrodynamics and reactions in activated sludge channel reactor.

Le Moullec et al. (2010a), b), Le Moullec et al. (2011)

Fluent Lab

Gas-liquid Eulerian

Transport eq. for oxygen + biokinetics (ASM1)

LDV, bubble size measurements

SC

M AN U

TE D

EP

RI PT

Aim

AC C

1141

61

RANS + > − $

MDV, DO measuremets

ACCEPTED MANUSCRIPT

Reference

Code

Impact of the surface aeration on the solids distribution in the oxidation ditch.

Fan et al. (2010)

Fluent Lab

3D

70,3k

RANS + > − $

Effect of aeration patterns on the flow field in the conventional AS tank.

Gresch et al. (2011)

CFX

Full

3D

400k

URANS + SST

−(

Modification of the operation conditions in the oxidation ditch to enhance energy efficiency.

Yang et al. (2011)

Fluent Full

3D

Evaluation of jet aeration and mixing in sequence batch reactors.

Samstag et al. (2012)

Fluent Full

3D

RTD of an oxidation ditch aerated with hydrojets.

Karpinska Portela (2013); Karpinska et al. (2015)

Fluent Pilot

3D

Multiphase model

Extra model

Validation

Solid-liquid Eulerian

-

PDA

Gas-liquid Eulerian

Biokineticsconsumption of NH4

ADV, reactive tracer experiments

Gas-liquid mixture

Transport eq. for oxygen + biokineticsconsumption of BOD

MDV, DO measurements

URANS + > − $

Gas-liquid mixture

Density-coupled model- solids settling

MLSS measurements

RANS + > − $, URANS + > − $, LES + Smagorinsky SGS

-

Lagrangian

-

1,66M RANS + > − $

TE D

EP

Model

SC

Scale Dim. Mesh size

RI PT

Aim

M AN U

Table 4. Continuation.

600k

AC C

1142

62

ACCEPTED MANUSCRIPT

1143

Aim

Reference

Code

Model

Studies on flow field and sludge settling in Carrousel oxidation ditch.

Xie et al. (2014)

Fluent Full

3D

1,53M URANS + > − $

Fluent Pilot

3D

Multiphase model

Extra model

Validation

Solid-liquid mixture

Slip velocity- sludge settling

MDV, MLSS measurements

Gas-liquidsolid mixture

Transport eq. for oxygen + biokinetics- (ASM1)

ADV, DO, COD, MLSS, NH4, NO3 measurements

SC

Scale Dim. Mesh size

RI PT

Table 4. Continuation.

M AN U

1144

A complete model to predict Lei and Ni (2014) hydrodynamics, oxygen transfer and biokinetic reactions in an oxidation ditch.

162k

RANS + > − $

ADV - Acoustic Doppler Velocimetry, MDV - Mono-directional Velocimetry, PDA - Particle Dynamic Analysis, LDV - Laser Doppler

1146

Velocimetry.

AC C

EP

TE D

1145

63

ACCEPTED MANUSCRIPT 6. Unaddressed Issues in CFD Modelling of AS Tanks

1148

Despite the fact that work has taken place on the CFD

1149

modelling of AS systems for over 15 years now, there are a

1150

number of issues which remain unaddressed and which, if

1151

successfully overcome, would enhance model fidelity and

1152

robustness further.

1153

6.1. Secondary Settler

1154

In engineering practice, the behaviour of the AS tanks is

1155

inseparably linked to the performance of the secondary settler,

1156

as the efficient removal of BOD and nutrients in AS process

1157

depends on the Solid Retention Times (SRTs), concentration of

1158

Mixed Liquor Suspended Solids (MLSS), and the composition

1159

of biomass. The rationale behind the dynamic modelling of a

1160

coupled aeration tank- clarifier system is the prediction of

1161

interconnected flow and mass transport in unsteady flow

1162

loading conditions, and thus evaluation of the impact of

1163

dynamic changes in return sludge flow rates on the

1164

concentration patterns (Patziger et al. 2012).

AC C

EP

TE D

M AN U

SC

RI PT

1147

1165

Secondary settling is referred to in the literature as “the most

1166

sensitive and complicated process in activated sludge plants”

1167

(Ji et al. 1996). In fact, the complete CFD modelling of a

1168

clarifier is not a feasible task, since many physico-biochemical

1169

phenomena

1170

hydrodynamics, turbulence, flocculation, sludge rheology,

1171

settling characteristics, heat exchange and temperature, and

must

be

considered

64

simultaneously,

i.e.

ACCEPTED MANUSCRIPT finally, biokinetics (Plósz et al. 2012). The most critical issue in

1173

the modelling of clarifiers is the inherent unpredictability of

1174

activated sludge settleability and missing data regarding the

1175

particulate fraction. Additionally, even the most advanced

1176

models for clarifiers are still based on empirical equations

1177

describing sludge settling. The variability of settling behaviour

1178

which is not predicted by models affects the actual SRTs of AS

1179

system, yielding a potential source of error in wastewater

1180

treatment plant models (Plósz et al. 2011, Plósz et al. 2012).

1181

Consequently, there is always a risk that the simultaneous

1182

modelling of clarifiers may affect the results obtained for the

1183

AS tank. Nonetheless, a CFD simulation of a simplified 2D AS

1184

tank- clarifier system has been reported in the literature

1185

(Patziger et al. 2012), where the focus was on the solids

1186

transport between both units in conditions of variable inflow

1187

and wet weather, but neglecting biokinetics and oxygen mass

1188

transfer. The necessity for further model enhancement and

1189

validation was also emphasized.

AC C

EP

TE D

M AN U

SC

RI PT

1172

1190

When considering the complexity of CFD models predicting

1191

clarifier behaviour simultaneously with a complete solid-liquid-

1192

gas model of the AS tank, computing time, availability of CPU

1193

resources and predictive capability must be recognised as

1194

confounding issues. As a result, common modelling practice is

1195

to avoid the interference between the two unit processes, and

1196

assume constant separation efficiency for the clarifier

65

ACCEPTED MANUSCRIPT performance (Le Moullec et al. 2010a, b, Le Moullec et al.

1198

2011) predicted in agreement with the guideline (Copp 2001).

1199

Moreover, in engineering practice, the design of clarifiers is

1200

based on assumption of certain sludge properties, hence the

1201

CFD models of settlers only have been frequently used for

1202

optimization of design and retrofitting purposes (De Clercq

1203

2003, Stamou et al. 2009). However, it will be worthwhile to

1204

pursue CFD analysis of integrated aeration tank- secondary

1205

settler systems, when models capable of predicting the

1206

character of activated sludge flocculation are regularly

1207

incorporated into sludge settling models.

1208

6.2. Population Balance Model

1209

The AS system can be described as an ensemble of populations

1210

of individual entities (bubbles, flocs, biomass cells), having

1211

specific properties (size, density, viscosity, enzymatic activity).

1212

In such a system, two kinds of behaviour may be recognized:

1213

interactions of the individual entities with the environment (e.g.

1214

interfacial oxygen transfer, shear induced break-up), and

AC C

EP

TE D

M AN U

SC

RI PT

1197

1215

mutual interactions between the individual entities (e.g.

1216

coalescence, aggregation). The character of these interactions is

1217

a function of one or more properties of the entities, which vary

1218

within the population. Thus, it is more correct to refer to this

1219

variation as “distributed properties”, as they can be represented

1220

by a distribution instead of a scalar (Nopens et al. 2015).

1221

Nevertheless, in order to reduce overall complexity of the 66

ACCEPTED MANUSCRIPT model framework associated with aeration tanks, a common

1223

procedure is to assume non-distributed scalar properties

1224

implying that all individuals behave in exactly the same way

1225

(Nopens et al. 2015, Sobremisana et al. 2011). The classic

1226

example is the assumption of fixed bubble diameter (Fayolle et

1227

al. 2007, Glover et al. 2006, Gresch et al. 2011, Le Moullec et

1228

al. 2010a) and uniform floc size (Fan et al. 2010).

1229

When considering an AS tank, recent experimental studies have

1230

demonstrated the influence of floc size on the flocculation

1231

behaviour and thus settleability of the activated sludge (Nopens

1232

et al. 2015), as well as on the biokinetic reaction environment

1233

within the AS floc (Sobremisana et al. 2011). Therefore

1234

estimation of floc size distribution may provide further insight

1235

into modelling of the activated sludge mixing (Samstag et al.

1236

2012) and assessment of aeration tank settling, an important

1237

process parameter allowing the prediction of hydraulic capacity

1238

and treatment efficiency during high hydraulic loads associated

1239

with wet weather flows (Nielsen et al. 2000, Sharma et al.

AC C

EP

TE D

M AN U

SC

RI PT

1222

1240

2013).

1241

Depending on the bubble flow regime in an AS tank, the

1242

intensity

1243

deformation promotes a large diversity in the shapes and sizes

1244

of the bubbles (Karpinska Portela 2013, Shaikh and Al-Dahhan

1245

2007, Takács 2005). At the same time, the impact of the bubble

1246

of

collisions,

agglomerations,

breakups

and

size on , superficial gas velocity, and thus gas hold-up and 67

ACCEPTED MANUSCRIPT oxygen mass transfer in aeration tanks has been also recognized

1248

(Fayolle et al. 2006, Vermande et al. 2007). However, in some

1249

cases involving modelling of aeration in conventional AS

1250

tanks, the assumption of non-distributed scalar properties is still

1251

predominant, while in others, i.e. modelling of clarifiers or

1252

bubble columns, it may lead to predictions that diverge

1253

significantly from the real systems (Sobremisana et al. 2011).

1254

Consequently,

1255

bubble/particle size distributions require the use of population

1256

balance models (PBM) to describe variations in populations of

1257

entities. An assessment of several solution methods for PBM,

1258

namely the discrete class size method (Hounslow et al. 1988),

1259

the standard method of moments- SMM (Randolf and Larson

1260

1971) and the quadrature method of moments- QMOM

1261

(Marchisio et al. 2003), can be found in the literature

1262

(Bridgeman et al. 2009). It should be highlighted, that although

1263

development of PBMs is at advanced stage and the coupling of

1264

QMOM with hydrodynamics requires only a small number of

models

which

incorporate

AC C

EP

TE D

M AN U

SC

multiphase

RI PT

1247

1265

computationally inexpensive scalar equations to be tracked, its

1266

application in modelling of aeration systems with suspended

1267

solids is still scarce and limited to very few examples

1268

(Bridgeman et al. 2009, Nopens et al. 2015, Sobremisana et al.

1269

2011), associated almost exclusively with secondary settlers

1270

(Griborio and McCorquodale 2006, Nopens et al. 2005).

1271

However, PBM-CFD coupling has been successfully exploited

68

ACCEPTED MANUSCRIPT 1272

by the chemical engineering community to study bubble

1273

columns, airlift and stirred bioreactors (Dhanasekharan et al.

1274

2005, Morchain et al. 2014, Wang 2011).

1275 7. Conclusions

1277

In the last few years, as a result of increasing availability and

1278

accessibility of commercial and open-source software suites,

1279

the use of CFD has evolved into a robust and precise technique

1280

for design, optimization and control of the AS systems. The

1281

following key conclusions can be put forward from this review:

SC



M AN U

1282

RI PT

1276

The complete CFD simulation of the complex multiphase flow in AS tanks remains a challenge, due

1284

to the high CPU and RAM requirements and limited

1285

feasibility resulting from the imposed convergence

1286

criteria. Although there is still no unequivocal protocol

1287

on CFD methodology, the most computationally

turbulence model has been adopted as the standard for

AC C

1289

efficient scenario, RANS/URANS closed by > − $

EP

1288

TE D

1283

1290

1291

the modelling of AS tanks;



Different CFD models serve different applications.

1292

Examples from the literature have demonstrated the

1293

potential and robustness of single flow simulations in

1294

design and optimization of the AS systems equipped

1295

with mechanical and jet aeration systems;

69

ACCEPTED MANUSCRIPT 1296



The Euler-Euler approach has been extensively

1297

explored for the prediction and optimization of

1298

oxygenation capacities in AS tanks equipped with

1299

diffused aeration systems; •

Lagrangian approach with particle tracking has been

RI PT

1300

used to determine RTDs of the tanks- a tool for

1302

evaluation of the aeration and mixing performance and

1303

troubleshooting of the reactor design;

1304



SC

1301

The neutral density modelling of AS tanks became common practice used for design and evaluation of

1306

aeration and mixing systems, despite leading to over-

1307

prediction of the degree of mixing. Hence the necessity

1308

to include density-coupled modelling of aeration tanks

1309

to predict accurately local density gradients due to the

1310

impact of the flow regime on solids transport; A small number of works concerns a complete threephase CFD model coupled with ASM1 aiming to quantify the transport and mutual interactions between

AC C

1313

TE D

1312



EP

1311

M AN U

1305

1314

water-gas-sludge phases within the AS reactor. Despite

1315

providing more reliable results than those obtained

1316

from ASM, this approach relies on the trade-off

1317

between a model which can be implemented and run in

1318

a realistic timeframe, and the extent of simplifications

1319

and the solution accuracy, which may lead to a series of

1320

output errors;

70

ACCEPTED MANUSCRIPT 1321



Potential possibilities of coupling of the CFD data with

1322

ASM based codes have been explored in order to

1323

generate a suitable tank-in-series model, where the

1324

biokinetic model can be implemented; •

There are several areas in modelling practice, which

RI PT

1325

remain unaddressed and require further study, e.g.

1327

modeling of coupled aeration tank- clarifier system to

1328

predict interconnected flow and mass transport in

1329

unsteady flow conditions and CFD-PBM coupling to

1330

assess the impact of the AS floc/air bubble size on

1331

mixing, settling, gas hold-up and oxygen mass transfer

1332

in the aeration tank.

M AN U

SC

1326

TE D

1333 Acknowledgements

1335

The research work of Dr. Anna M. Karpinska Portela was

1336

funded by the College of Engineering and Physical Sciences,

1337

University of Birmingham, UK.

EP

1334

AC C

1338 1339

References

1340 1341 1342 1343

Abusam, A. and Keesman, K.J. (1999) Effect of number of CSTR’s on the modelling of oxidation ditches; steady state and dynamic analysis . Proceedings of 13th Forum for Applied Biotechnology, pp. 91-94, Sept. 22-23, Gent, Belgium.

1344 1345 1346 1347

Abusam, A., Keesman, K.J., Spanjers, H., Van Straten, G. and Meinema, K. (2002) Effect of oxidation ditch horizontal velocity on the nitrogen removal process. European Water Management Online European Water Association (EWA).

71

ACCEPTED MANUSCRIPT Åmand, L. and Carlsson, B. (2012) Optimal aeration control in a nitrifying activated sludge process. Water Research 46(7), 21012110.

1351 1352 1353 1354

Andersson, B., Andersson, R., Håkansson, L., Mortensen, M., Sudiyo, R. and Van Wachem, B. (2011) Computational fluid dynamics for engineers, Cambridge: Cambridge University Press, Cambridge.

1355 1356 1357

Ardern, E. and Lockett, W.T. (1914) Experiments on the oxidation of the sewage without the aid of filters. Journal of the Society of Chemical Industry 33(10), 523-539.

1358 1359 1360

ATV-DVWK (2000) Standard ATV-DVWK- A 131E Dimensioning of Single-Stage Activated Sludge Plants, German Association for Water, Wastewater, Waste, Hennef, Germany.

1361 1362 1363

Azzopardi, B.J., Mudde, R.F., Lo, S., Morvan, H., Yan, Y. and Zhao, D. (2011) Hydrodynamics of gas-liquid reactors: normal operation and upset conditions, John Wiley & Sons, Ltd, Chichester, UK.

1364 1365 1366 1367 1368

Bhuyar, L.B., Thakre, S.B. and Ingole, N.W. (2009) Design characteristics of curved blade aerator w.r.t. aeration efficiency and overall oxygen transfer coefficient and comparison with CFD modeling. International Journal of Engineering, Science and Technology 1(1), 1-15.

1369 1370 1371

Bosma, A.J. and Reitsma, B. (2007) Hydraulic design of activated sludge tanks with CFD. Proceedings of the COMSOL Users Conference, October 23-24, Grenoble, France.

1372 1373 1374

Brannock, M. (2003) Computational Fluid Dynamics tools for the design of mixed anoxic wastewater treatmnent vessels. PhD Thesis, University of Queensland, Brisbane, Australia.

1375 1376 1377 1378

Bridgeman, J., Jefferson, B. and Parsons, S.A. (2009) Computational Fluid Dynamics modelling of flocculation in water treatment: A review. Engineering Applications of Computational Fluid Mechanics 3(2), 220-241.

AC C

EP

TE D

M AN U

SC

RI PT

1348 1349 1350

1379 1380 1381 1382

Brouckaert, C.J. and Buckley, C.A. (1999) The Use of Computational Fluid Dynamics for improving the design and operation of water and wastewater treatment plants Water Science & Technology 40(4-5), 81-89.

1383 1384 1385

Cartland Glover, G.M., Generalis, S.C. and Thomas, N.H. (2000) CFD and bubble column reactors: simulation and experiment. Chemical papers 54(6A), 361-369.

1386 1387 1388 1389

Chachuat, B., Roche, N. and Latifi, M.A. (2005) Long-term optimal aeration strategies for small-size alternating activated sludge treatment plants. Chemical Engineering and Processing: Process Intensification 44(5), 591-604.

72

ACCEPTED MANUSCRIPT Clift, R., Grace, J.R. and Weber, M.E. (1978) Bubbles, drops, and particles, Academic Press Inc., London, UK.

1392 1393 1394 1395

Cockx, A., Do-Quang, Z., Audic, J.M., Liné, A. and Roustan, M. (2001) Global and local mass transfer coefficients in waste water treatment process by computational fluid dynamics. Chemical Engineering and Processing 40(2), 187-194.

1396 1397 1398 1399

Combest, D.P., Ramachandran, P.A. and Dudukovic, M.P. (2011) On the gradient diffusion hypothesis and passive scalar transport in turbulent flows. Industrial & Engineering Chemistry Research 50(15), 8817-8823.

1400 1401 1402

Copp, J.B. (2001) The COST Simulation Benchmark: Description and simulator manual. Office for Official Publications of the European Community, Luxembourg.

1403 1404 1405

Cristea, S., de Prada, C., Sarabia, D. and Gutiérrez, G. (2011) Aeration control of a wastewater treatment plant using hybrid NMPC. Computers & Chemical Engineering 35(4), 638-650.

1406 1407

Danckwerts, P.V. (1953) Continuous flow systems. Chemical Engineering Science 2(1), 1-13.

1408 1409 1410 1411

De Araújo, A.C.B., Gallani, S., Mulas, M. and Olsson, G. (2011) Systematic approach to the design of operation and control policies in activated sludge systems. Industrial and Engineering Chemistry Research 50(14), 8542-8557.

1412 1413 1414 1415

De Clercq, B. (2003) Computational Fluid Dynamics of settling tanks:Development of experiments and rheological, settling, and scraper submodels. PhD Thesis, University of Ghent, Ghent, Belgium.

1416 1417 1418 1419

De Gussem, K., Fenu, A., Wambecq, T. and Weemaes, M. (2014) Energy saving on wastewater treatment plants through improved online control: case study wastewater treatment plant AntwerpSouth. Water Science & Technology 65(5), 1074-1079.

1420 1421 1422 1423

Dhanasekharan, K.M., Sanyal, J., Jain, A. and Haidari, A. (2005) A generalized approach to model oxygen transfer in bioreactors using population balances and computational fluid dynamics. Chemical Engineering Science 60(1), 213-218.

1424 1425 1426 1427

Do-Quang, Z., Cockx, A., Liné, A. and Roustan, M. (1999) Computational fluid dynamics applied to water and wastewater treatment facility modeling. Environmental Engineering and Policy 1(3), 137-147.

1428 1429 1430

EPA (1989) Design Manual: Fine pore aeration systems, United States Environmental Protection Agency, Research and Development - Risk Reduction Engineering Laboratory, Cincinnati, OH

1431 1432

EPA (2013) Energy efficiency in water and wastewater facilities - A guide to developing and implementing greenhouse gas reduction

AC C

EP

TE D

M AN U

SC

RI PT

1390 1391

73

ACCEPTED MANUSCRIPT 1433 1434

programs United Washington, DC.

1435 1436 1437 1438

Essemiani, K., Vermande, S., Marsal, S., Phan, L. and Meinhold, J. (2004) Optimization of WWTP units using CFD - a tool grown for real scale application. van Loosdrecht, M. and Clement, J. (eds), pp. 183-192, IWA Publishing.

1439 1440 1441

Fan, L., Xu, N., Wang, Z. and Shi, H. (2010) PDA experiments and CFD simulation of a lab-scale oxidation ditch with surface aerators. Chemical Engineering Research and Design 88(1), 23-33.

1442 1443 1444

Fayolle, Y., Cockx, A., Gillot, S., Roustan, M. and Héduit, A. (2007) Oxygen transfer prediction in aeration tanks using CFD. Chemical Engineering Science 62(24), 7163-7171.

1445 1446 1447 1448

Fayolle, Y., Gillot, S., Cockx, A., Bensimhon, L., Roustan, M. and Heduit, A. (2010) In situ characterization of local hydrodynamic parameters in closed-loop aeration tanks. Chemical Engineering Journal 158(2), 207-212.

1449 1450 1451 1452 1453

Fayolle, Y., Gillot, S., Cockx, A., Roustan, M. and Heduit, A. (2006) In situ local parameter measurements for CFD modelling to optimize aeration. Proceedings of WEFTEC 06 - 79th Annual WEF Technical Exhibition and Conference, pp. 3314-3325, Oct. 21-25, Dallas, TX, USA.

1454 1455 1456 1457

Fernández, F.J., Castro, M.C., Rodrigo, M.A. and Cañizares, P. (2011) Reduction of aeration costs by tuning a multi-set point on/off controller: A case study. Control Engineering Practice 19(10), 12311237.

1458 1459 1460

Fikar, M., Chachuat, B. and Latifi, M.A. (2005) Optimal operation of alternating activated sludge processes. Control Engineering Practice 13(7), 853-861.

1461 1462 1463 1464

Gillot, S., Capela-Marsal, S., Roustan, M. and Héduit, A. (2005) Predicting oxygen transfer of fine bubble diffused aeration systems Model issued from dimensional analysis. Water Research 39(7), 1379-1387.

1465 1466 1467

Gillot, S. and Héduit, A. (2000) Effect of air flow rate on oxygen transfer in an oxidation ditch equipped with fine bubble diffusers and slow speed mixers. Water Research 34(5), 1756-1762.

1468 1469 1470 1471

Glover, G.C., Printemps, C., Essemiani, K. and Meinhold, J. (2006) Modelling of wastewater treatment plants - how far shall we go with sophisticated modelling tools? Water Science & Technology 53(3), 79-89.

1472 1473 1474

Gresch, M., Armbruster, M., Braun, D. and Gujer, W. (2011) Effects of aeration patterns on the flow field in wastewater aeration tanks. Water Research 45(2), 810-818.

Environmental

Protection

Agency,

AC C

EP

TE D

M AN U

SC

RI PT

States

74

ACCEPTED MANUSCRIPT Griborio, A. and McCorquodale, J.A. (2006) Optimum design of your center well: use of a CFD model to understand the balance between flocculation and improved hydrodynamics. Proceedings of WEFTEC 06 - 79th Annual WEF Technical Exhibition and Conference, pp. 5735-5746, Oct. 21-25, Dallas, TX, USA.

1480 1481 1482 1483

Guimet, V., Savoye, P., Audic, J.-M. and Do-Quang, Z. (2004) Advanced CFD tool for Wastewater: today complex Modelling & tomorrow easy-to-use Interface. Van Loosdrecht, M. and Clement, J. (eds), pp. 165-172, IWA Publishing.

1484 1485 1486

Henze, M., Gujer, W., Mino, T. and van Loosdrecht, M.C.M. (2000) Activated Sludge Models ASM1, ASM2, ASM2d and ASM3. IWA Scientific and Technical Report No. 9, London, UK.

1487 1488 1489 1490

Holenda, B., Domokos, E., Rédey, Á. and Fazakas, J. (2007) Aeration optimization of a wastewater treatment plant using genetic algorithm. Optimal Control Applications and Methods 28(3), 191208.

1491 1492 1493 1494

Holenda, B., Domokos, E., Rédey, Á. and Fazakas, J. (2008) Dissolved oxygen control of the activated sludge wastewater treatment process using model predictive control. Computers & Chemical Engineering 32(6), 1270-1278.

1495 1496 1497

Hounslow, M.J., Ryall, R.L. and Marshall, V.R. (1988) A discretized population balance for nucleation, growth, and aggregation. AIChE Journal 34(11), 1821-1832.

1498 1499 1500 1501

Jensen, M.D., Ingildsen, P., Rasmussen, M.R. and Laursen, J. (2006) Computational Fluid Dynamics modelling of hydraulics and sedimentation in process reactors during aeration tank settling. Water Science & Technology 53(12), 257-264.

1502 1503 1504

Ji, Z., McCorquodale, J.A., Zhou, S. and Vitasovic, Z. (1996) A dynamic solids inventory model for activated sludge systems. Water Environment Research 68(3), 329-337.

1505 1506 1507

Joshi, J.B. (2001) Computational flow modelling and design of bubble column reactors. Chemical Engineering Science 56(21–22), 5893-5933.

1508 1509 1510 1511

Karpinska, A.M., Dias, M.M., Boaventura, R., A. and Santos, R.J. (2015) Modeling of the hydrodynamics and energy expenditure of oxidation ditch aerated with hydrojets using CFD codes. Water Quality Research Journal of Canada 50(1), 83-94.

1512 1513 1514

Karpinska Portela, A.M. (2013) New design tools for Activated Sludge process. PhD Thesis, FEUP, University of Porto, Porto, Portugal.

1515 1516 1517 1518

Kjellstrand, R. (2006) Hydraulic Behaviour in an Activated Sludge tank. From tracer test through hydraulic modelling to full-scale implementation. Lic. of Eng. Thesis, Chalmers University of Technology, Göteborg, Sweden

AC C

EP

TE D

M AN U

SC

RI PT

1475 1476 1477 1478 1479

75

ACCEPTED MANUSCRIPT Kochevsky, A.N. (2004) Possibilities of simulation of fluid flows using the modern CFD software tools. arXiv preprint physics/0409104.

1522 1523 1524 1525

Langergraber, G., Rieger, L., Winkler, S., Alex, J., Wiese, J., Owerdieck, C., Ahnert, M., Simon, J. and Maurer, M. (2004) A guideline for simulation studies of wastewater treatment plants. Water Science & Technology 50(7), 131-138.

1526 1527 1528

Launder, B.E., Reece, G.J. and Rodi, W. (1975) Progress in the development of a Reynolds-stress turbulent closure. Journal of Fluid Mechanics 68, 537-566.

1529 1530 1531

Launder, B.E. and Spalding, D.B. (1974) The numerical computation of turbulent flows. Computer Methods in Applied Mechanics and Engineering 3(2), 269-289.

1532 1533 1534 1535 1536

Laurent, J., Samstag, R.W., Ducoste, J., Griborio, A., Nopens, I., Batstone, D.J., Wicks, J., Saunders, S. and Potier, O. (2014) A protocol for the use of computational fluid dynamics as a supportive tool for wastewater treatment plant modelling. Water Science & Technology 70(10), 1757-1584.

1537 1538 1539 1540

Laursen, J. (2007) WWTPprocess tank modelling - The integration of advanced hydrodynamic and microbiological models. PhD Thesis, Department of Civil Engineering, Aalborg University, Aalborg, Denmark.

1541 1542 1543 1544

Le Moullec, Y., Gentric, C., Potier, O. and Leclerc, J.P. (2010a) CFD simulation of the hydrodynamics and reactions in an activated sludge channel reactor of wastewater treatment. Chemical Engineering Science 65(1), 492-498.

1545 1546 1547 1548

Le Moullec, Y., Gentric, C., Potier, O. and Leclerc, J.P. (2010b) Comparison of systemic, compartmental and CFD modelling approaches: Application to the simulation of a biological reactor of wastewater treatment. Chemical Engineering Science 65(1), 343-350.

1549 1550 1551 1552

Le Moullec, Y., Potier, O., Gentric, C. and Leclerc, J.P. (2011) Activated sludge pilot plant: Comparison between experimental and predicted concentration profiles using three different modelling approaches. Water Research 45(10), 3085-3097.

1553 1554 1555 1556

Le Moullec, Y., Potier, O., Gentric, C. and Pierre Leclerc, J. (2008) Flow field and residence time distribution simulation of a cross-flow gas-liquid wastewater treatment reactor using CFD. Chemical Engineering Science 63(9), 2436-2449.

1557 1558 1559 1560

Lei, L. and Ni, J. (2014) Three-dimensional three-phase model for simulation of hydrodynamics, oxygen mass transfer, carbon oxidation, nitrification and denitrification in an oxidation ditch. Water Research 53(0), 200-214.

1561 1562

Levenspiel, O. (1999) Chemical Reaction Engineering, Wiley & Sons, New York, NY.

AC C

EP

TE D

M AN U

SC

RI PT

1519 1520 1521

76

ACCEPTED MANUSCRIPT Makinia, J. and Wells, S.A. (2000) A general model of the activated sludge reactor with dispersive flow - I. Model development and parameter estimation. Water Research 34(16), 3987-3996.

1566 1567

Manninen, M., Taivassalo, V. and Kallio, S. (1996) On the mixture model for the multiphase flow. VTT Publications 288, 1-67.

1568 1569 1570

Marchisio, D.L., Vigil, R.D. and Fox, R.O. (2003) Quadrature method of moments for aggregation–breakage processes. Journal of Colloid and Interface Science 258(2), 322-334.

1571 1572 1573 1574

Menter, F.R. (1993) Zonal two equation k-w turbulence models for aerodynamic flows. Proceedings of 23rd Fluid Dynamics, Plasmadynamics and Lasers Conference, July 6-9, Orlando, FL, USA.

1575 1576

Menter, F.R. (1994) Two-equation eddy-viscosity turbulence models for engineering applications. AIAA Journal 32(8), 1598-1605.

1577 1578 1579 1580

Morchain, J., Gabelle, J.-C. and Cockx, A. (2014) A coupled population balance model and CFD approach for the simulation of mixing issues in lab-scale and industrial bioreactors. AIChE Journal 60(1), 27-40.

1581 1582 1583

Morchain, J., Maranges, C. and Fonade, C. (2000) CFD modelling of a two-phase jet aerator under influence of a crossflow. Water Research 34(13), 3460-3472.

1584 1585

Mueller, J.A., Boyle, W.C. and Pöpel, H.J. (2002) Aeration: principles and practice, CRC Press, Boca Raton, FL, USA.

1586 1587

Nauman, E.B. (2007) Residence Time Distributions, pp. 535-574, John Wiley & Sons, Inc., New Jersey, NJ, USA.

1588 1589 1590

Nielsen, M.K., Bechmann, H. and M., H. (2000) Modelling and test of aeration tank settling (ATS). Water Science & Technology 41(9), 179-184.

1591 1592 1593 1594 1595

Nopens, I., Batstone, D.J., Griborio, A., Samstag, R.W., Wicklein, E. and Wicks, J. (2012) Computational Fluid Dynamics (CFD): What is Good CFD-Modeling Practice and what can be the added value of CFD models to WWTP modeling?, pp. 7400-7405, Sept. 29-Oct. 3, New Orleans, LA, USA.

1596 1597 1598

Nopens, I., Koegst, T., Mahieu, K. and Vanrolleghem, P.A. (2005) PBM and activated sludge flocculation: From experimental data to calibrated model. AIChE Journal 51(5), 1548-1557.

1599 1600 1601 1602

Nopens, I., Torfs, E., Ducoste, J., Vanrolleghem, P.A. and Gernaey, K.V. (2015) Population balance models: a useful complementary modelling framework for future WWTP modelling. Water Science & Technology 71(2), 159-167.

1603 1604

Nopens, I. and Wicks, J. (2012) Integration of CFD and Biokinetics. CFD Workshop- Presentation. 3rd IWA/WEF Wastewater Treatment

AC C

EP

TE D

M AN U

SC

RI PT

1563 1564 1565

77

ACCEPTED MANUSCRIPT 1605 1606

Modelling Canada.

1607 1608

Orszag, S.A. (1970) Analytical theories of turbulence. Journal of Fluid Mechanics 41(2), 363-386.

1609 1610 1611 1612

Orszag, S.A., Staroselsky, I., Flannery, W.S. and Zhang, Y. (eds) (1996) Simulation and modeling of turbulent flows. Chapter 4: Introduction to Renormalization Group Modeling of Turbulence, Oxford University Press Inc., New York, NY, USA.

1613 1614 1615 1616

Patziger, M., Kainz, H., Hunze, M. and Józsa, J. (2012) Influence of secondary settling tank performance on suspended solids mass balance in activated sludge systems. Water Research 46(7), 24152424.

1617 1618 1619 1620

Pereira, J.P., Karpinska, A., Gomes, P.J., Martins, A.A., Dias, M.M., Lopes, J.C.B. and Santos, R.J. (2012) Activated Sludge Models coupled to CFD simulations. Dias, R., Lima, R., Martins, A.A. and Mata, T.M. (eds), pp. 153-173, Bentham Science Publishers Ltd.

1621 1622 1623

Pereira, J.P., Karpinska, A.M., Dias, M.M. and Santos, R.J. (2009) The dissolved oxygen profile in oxidation ditches: a numerical study on its effect on nutrient removal, pp. 97-104, Krakow, Poland.

1624 1625 1626 1627 1628

Plósz, B.G., De Clercq, J., Nopens, I., Benedetti, L. and Vanrolleghem, P.A. (2011) Shall we upgrade one-dimensional secondary settler models used in WWTP simulators? – An assessment of model structure uncertainty and its propagation. Water Science & Technology 63(8), 1726–1738.

1629 1630 1631 1632 1633 1634

Plósz, B.G., Nopens, I., Rieger, L., Griborio, A., De Clercq, J., Vanrolleghem, P.A., Daigger, G.T., Takács, I., Wicks, J. and Ekama, G.A. (2012) A critical review of clarifier modelling: State-of-the-art and engineering practices. Proceedings of WWTmod2012 - 3rd IWA/WEF Wastewater Treatment Modelling Seminar, pp. 27-30, Feb. 26-28, Mont-Sainte-Anne, Québec, Canada.

1635

Pope, S.B. (2000) Turbulent Flows, New York, NY, USA.

Mont-Sainte-Anne,

QC,

AC C

EP

TE D

M AN U

SC

RI PT

Seminar-WWTmod2012,

1636 1637

Randolf, A.D. and Larson, M.A. (1971) Theory of particulate processes, Academic Press, New York, NY, USA.

1638 1639 1640 1641 1642

Ratkovich, N. (2010) Understanding hydrodynamics in Membrane Bioreactor systems for wastewater treatment: Two-phase empirical and numerical modelling and experimental validation. PhD thesis, Faculty of Bioscience Engineering, Ghent University, Ghent, Belgium.

1643 1644 1645 1646 1647

Ratkovich, N., Hunze, M., Futselaar, H. and Nopens, I. (2009) Use of CFD to model and optimize the hydrodynamics of an airlift MBR side-stream module. Proceedings of WEFTEC 09 - 82th Annual WEF Technical Exhibition and Conference, pp. 2812-2818, Oct. 1014, Orlando, FL, USA

78

ACCEPTED MANUSCRIPT Reardon, D.J. (1995) Turning down the power. Civil Engineering 65(8), 54-56.

1650 1651 1652

Rieger, L., Alex, J., Gujer, W. and Siegrist, H. (2006) Modelling of aeration systems at wastewater treatment plants. Water Science & Technology 53(4-5), 439-447.

1653 1654

Rodi, W. (1993) Turbulence models and their application in hydraulics., IAHR Monograph, Delft, The Netherlands.

1655 1656 1657 1658

Rosso, D., Lothman, S.E., Jeung, M.K., Pitt, P., Gellner, W.J., Stone, A.L. and Howard, D. (2011) Oxygen transfer and uptake, nutrient removal, and energy footprint of parallel full-scale IFAS and activated sludge processes. Water Research 45(18), 5987-5996.

1659 1660

Rosso, D. and Stenstrom, M.K. (2006) Surfactant effects on α-factors in aeration systems. Water Research 40(7), 1397-1404.

1661 1662 1663 1664

Samstag, R.W., Ducoste, J.M., Griborio, A., Nopens, I., Batstone, D.J., Wicks, J.D., Saunders, S., Wicklein, E., Kenny, G. and Laurent, J. (2015) CFD for wastewater treatment: A State of the Art review. Submitted manuscript.

1665 1666 1667

Samstag, R.W., McCorqudale, J.A. and Zhou, S.P. (1992) Prospects for Transport Modelling in Process Tanks. Water Science & Technology 26(5-6), 1401-1410.

1668 1669 1670 1671

Samstag, R.W. and Wicklein, E. (2012) Application of CFD to Activated Sludge mixing analysis. CFD Workshop- Presentation. 3rd IWA/WEF Wastewater Treatment Modelling SeminarWWTmod2012, February 26-28 Mont-Sainte-Anne, QC, Canada.

1672 1673 1674 1675

Samstag, R.W. and Wicklein, E. (2014) A Protocol for optimization of Activated Sludge mixing. Proceedings of WEFTEC - 87th Annual WEF Technical Exhibition and Conference, Sept. 27-Oct. 1, New Orleans, LA, USA.

1676 1677 1678 1679 1680

Samstag, R.W., Wicklein, E., Reardon, R.D., Leetch, R.J., Parks, R.M. and Groff, C.D. (2012) Field and CFD analysis of jet aeration and mixing. Proceedings of WEFTEC 2012 - 85th Annual WEF Technical Exhibition and Conference, pp. 4113-4139, Sept. 29-Oct. 3, New Orleans, LA, USA.

1681 1682 1683

Shaikh, A. and Al-Dahhan, M.H. (2007) A Review on flow regime transition in bubble columns. International Journal of Chemical Reactor Engineering 5(1), 1-68.

1684 1685 1686 1687 1688

Sharma, A.K., Guildal, T., Thomsen, H.A., Mikkelsen, P.S. and Jacobsen, B.N. (2013) Aeration tank settling and real time control as a tool to improve the hydraulic capacity and treatment efficiency during wet weather: results from 7 years' full-scale operational data. Water Science & Technology 67(10), 2169-2176.

1689 1690

Shih, T.-H., Liou, W.W., Shabbir, A., Yang, Z. and Zhu, J. (1995) A New K-epsilon eddy viscosity model for high Reynolds number

AC C

EP

TE D

M AN U

SC

RI PT

1648 1649

79

ACCEPTED MANUSCRIPT turbulent flows: Model development and validation. Computers & Fluids 24(3), 227-238.

1693 1694

Smagorinsky, J. (1963) General circulation experiments with the primitive equations. Monthly Weather Review 93(3), 99-164.

1695 1696 1697 1698 1699 1700

Sobremisana, A.P., Ducoste, J.J. and de los Reyes III, F.L. (2011) Combining CFD, floc dynamics and biological reaction kinetics to model carbon and nitrogen removal in an activated sludge system. Proceedings of WEFTEC 2011 - 84th Annual WEF Technical Exhibition and Conference, pp. 3272-3282, Oct. 15-19, Los Angeles, CA, USA.

1701 1702 1703 1704

Sokolichin, A., Eigenberger, G., Lapin, A. and Lübert, A. (1997) Dynamic numerical simulation of gas-liquid two-phase flows Euler/Euler versus Euler/Lagrange. Chemical Engineering Science 52(4), 611-626.

1705 1706 1707

Stamou, A.I. (2008) Improving the hydraulic efficiency of water process tanks using CFD models. Chemical Engineering and Processing: Process Intensification 47(8), 1179-1189.

1708 1709 1710

Stamou, A.I., Theodoridis, G. and Xanthopoulos, K. (2009) Design of secondary settling tanks using a CFD model. Journal of Environmental Engineering 135(7), 551-561.

1711 1712

Takács, G. (2005) Gas Lift: Manual, PennWell Books, Tulsa, OK, USA.

1713 1714 1715 1716

Taricska, J.R., Huang, J.Y.C., Chen, P., Hung, Y.-T. and Zou, S.-W. (2009) Biological treatment processes. Wang, L.K., Pereira, N.C., Hung, Y.-T. and Shammas, N.K. (eds), The Humana Press, Totowa, NJ.

1717 1718 1719

Tchobanoglous, G., Burton, F.L., Stensel, H.D. and Eddy, M. (eds) (2003) Wastewater engineering: treatment and reuse, McGraw-Hill New York, NY.

1720 1721 1722

Vedantam, S., Joshi, J.B. and Koganti, S.B. (2006) Threedimensional CFD simulation of stratified two-fluid Taylor-Couette flow. The Canadian Journal of Chemical Engineering 84(3), 279-288.

1723 1724 1725 1726

Vermande, S., Simpson, K., Essemiani, K., Fonade, C. and Meinhold, J. (2007) Impact of agitation and aeration on hydraulics and oxygen transfer in an aeration ditch: Local and global measurements. Chemical Engineering Science 62(9), 2545-2555.

1727 1728 1729

Versteeg, H.K. and Malalasekera, W. (1995) An introduction to Computational Fluid Dynamics. The Finite Volume Method, Pearsons Prentice Hall, Harlow, Essex, UK.

1730 1731 1732

Wang, T. (2011) Simulation of bubble column reactors using CFD coupled with a population balance model. Frontiers of Chemical Science and Engineering 5(2), 162-172.

AC C

EP

TE D

M AN U

SC

RI PT

1691 1692

80

ACCEPTED MANUSCRIPT Wang, Y., Waite, T.D. and Leslie, G.L. (2013) Handbook of membrane reactors. Basile, A. (ed), pp. 532-568, Woodhead Publishing, Cambridge, UK.

1736 1737 1738

WEF (2009) Energy conservation in water and wastewater treatment facilities, Water Environment Federation Manual of Practice No. 32; McGraw-Hill, Inc., New York, NY.

1739 1740 1741 1742

Wicklein, E. and Samstag, R.W. (2009) Comparing commercial and transport CFD models for secondary sedimentation. Proceedings of WEFTEC 09 - 82th Annual WEF Technical Exhibition and Conference, pp. 6066-6081, Oct. 10-14, Orlando, FL, USA

1743 1744 1745

Wilcox, D.C. (1998) Re-assessment of the scale-determining equation for advanced turbulence models. AIAA Journal 26(11), 1299-1310.

1746 1747 1748

Wilcox, D.C. and Traci, R.M. (1976) A complete model of turbulence. Proceedings of 9th Fluid and PlasmaDynamics Conference, July 14-16, San Griego, CA, USA.

1749 1750 1751 1752

Xie, H., Yang, J., Hu, Y., Zhang, H., Yang, Y., Zhang, K., Zhu, X., Li, Y. and Yang, C. (2014) Simulation of flow field and sludge settling in a full-scale oxidation ditch by using a two-phase flow CFD model. Chemical Engineering Science 109(0), 296-305.

1753 1754 1755 1756

Xu, N., Fan, L., Pang, H. and Shi, H. (2010) Feasibility study and CFD-aided design for a new type oxidation ditch based on airlift circulation. The Canadian Journal of Chemical Engineering 88(5), 728-741.

1757 1758

Yang, C. and Mao, Z.-S. (2014) Numerical simulation of multiphase reactors with continuous liquid phase, Academic Press, Oxford, UK.

1759 1760 1761 1762

Yang, Y., Yang, J., Zuo, J., Li, Y., He, S., Yang, X. and Zhang, K. (2011) Study on two operating conditions of a full-scale oxidation ditch for optimization of energy consumption and effluent quality by using CFD model. Water Research 45(11), 3439-3452.

1763 1764

Yeoh, G.H. and Tu, J. (2009) Computational techniques for multiphase flows, Elsevier.

AC C

EP

TE D

M AN U

SC

RI PT

1733 1734 1735

1765 1766

81

ACCEPTED MANUSCRIPT Figure captions Figure 1. Streamlines coloured by liquid velocity magnitude ܷ௅ in tank (a) without aeration (ܷ௅ = 0.35 ms−1) (b) without aeration (ܷ௅ = 0.27 ms−1) and (c) with aeration (ܷ௅ = 0.23 ms−1). Reprinted from Fayolle et al. (2007). Copyright (2007) with permission from Elsevier.

RI PT

Figure 2. Snapshot of air volume fractions at two cross sections with different diffuser patterns. Reprinted from Gresch et al. (2011). Copyright (2011) with permission from Elsevier.

Figure 3. Contours of volume fraction of solid phase at different height: (A) top of the tank, (B) middle of the tank, and (C) bottom of the tank. Reprinted from Xie et al. (2014). Copyright (2014) with permission from Elsevier.

M AN U

SC

Figure 4. Comparison between experimental and simulated RTD obtained with the RSM and the k – ε turbulence models and the particle tracking method for a liquid flowrate of 3.6 L min-1 and a gas flowrate of 15 L min-1. Reprinted from Le Moullec et al. (2008). Copyright (2008) with permission from Elsevier.

AC C

EP

TE D

Figure 5. The contour plots of: (a) MLSS, (b) DO, (c) COD, (d) ammonia and (e) nitrate concentration distribution, where “+” indicates measured data values at the sampling points in a Carrousel oxidation ditch. Reprinted from Lei and Ni (2014). Copyright (2014) with permission from Elsevier.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Highlights

EP

TE D

M AN U

SC

RI PT

CFD analysis is a robust tool for evaluation and optimization of the AS systems. Turbulence models and multiphase approaches used to study AS tanks are presented. Pitfalls of modelling assumptions and simplifications are identified. Methods and examples of coupling of the CFD data with biokinetics are discussed. Unaddressed challenges in modelling of the AS systems are critically discussed.

AC C

• • • • •