Accepted Manuscript Metabolic modelling of full-scale enhanced biological phosphorus removal sludge Ana B. Lanham, Adrian Oehmen, Aaron M. Saunders, Gilda Carvalho, Per H. Nielsen, Maria A.M. Reis PII:

S0043-1354(14)00603-4

DOI:

10.1016/j.watres.2014.08.036

Reference:

WR 10843

To appear in:

Water Research

Received Date: 20 May 2014 Revised Date:

1 August 2014

Accepted Date: 23 August 2014

Please cite this article as: Lanham, A.B., Oehmen, A., Saunders, A.M., Carvalho, G., Nielsen, P.H., Reis, M.A.M., Metabolic modelling of full-scale enhanced biological phosphorus removal sludge, Water Research (2014), doi: 10.1016/j.watres.2014.08.036. 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.

ACCEPTED MANUSCRIPT 1

Metabolic modelling of full-scale enhanced biological

2

phosphorus removal sludge

3

Ana B. Lanhama, Adrian Oehmena,*, Aaron M. Saundersc, Gilda Carvalhoa,b, Per H.

4

Nielsenc, Maria A. M. Reisa

5

a

6

[email protected]; [email protected]; [email protected];

7

[email protected]

8

b Instituto de Biologia Experimental e Tecnológica, Apt.12, 2781-901 Oeiras, Portugal

9

c Center for Microbial Communities, Department of Biotechnology, Chemistry and

SC

RI PT

REQUIMTE/CQFB, Chemistry Department FCT-UNL, 2829-516 Caparica, Portugal,

Environmental Engineering, Aalborg University, Denmark, [email protected];

11

[email protected]

12

* Corresponding author – Tel: +351212948571 /Fax: +351212948550

M AN U

10

13 14 15

Abstract

16

This study investigates, for the first time, the application of metabolic models

17

incorporating

18

accumulating organisms (GAOs) towards describing the biochemical transformations of

19

full-scale enhanced biological phosphorus removal (EBPR) activated sludge from

20

wastewater treatment plants (WWTPs). For this purpose, it was required to modify

21

previous metabolic models applied to lab-scale systems by incorporating the anaerobic

22

utilisation of the TCA cycle and the aerobic maintenance processes based on sequential

23

utilisation of polyhydroxyalkanoates, followed by glycogen and polyphosphate. The

24

abundance of the PAO and GAO populations quantified by fluorescence in situ

25

hybridisation served as the initial conditions of each biomass fraction, whereby the

26

models were able to describe accurately the experimental data. The kinetic rates were

27

found to change among the four different WWTPs studied or even in the same plant

28

during different seasons, either suggesting the presence of additional PAO or GAO

29

organisms, or varying microbial activities for the same organisms. Nevertheless, these

30

variations in kinetic rates were largely found to be proportional to the difference in

31

acetate uptake rate, suggesting a viable means of calibrating the metabolic model. The

32

application of the metabolic model to full-scale sludge also revealed that different

accumulating

organisms

(PAOs)

and

glycogen

AC C

EP

TE D

polyphosphate

ACCEPTED MANUSCRIPT Accumulibacter clades likely possess different acetate uptake mechanisms, as a

34

correlation was observed between the energetic requirement for acetate transport across

35

the cell membrane with the diversity of Accumulibacter present. Using the model as a

36

predictive tool, it was shown that lower acetate concentrations in the feed as well as

37

longer aerobic retention times favour the dominance of the TCA metabolism over

38

glycolysis, which could explain why the anaerobic TCA pathway seems to be more

39

relevant in full-scale WWTPs than in lab-scale systems.

40

RI PT

33

Keywords : metabolic modelling; polyphosphate accumulating organisms (PAOs);

42

glycogen accumulating organisms (GAOs); anaerobic TCA cycle; glycolysis;

43

maintenance processes

SC

41

44

M AN U

45

1

Introduction

47

The biological removal of phosphate (also known as enhanced biological phosphorus

48

removal or EBPR) has been incorporated into several wastewater treatment plant

49

(WWTP) configurations and provides a more economical and sustainable alternative to

50

chemical precipitation methods of P removal (Mino et al., 1998; Oehmen et al., 2007).

TE D

46

51

For more than 20 years, an effort has been made to develop and apply activated sludge

53

models (ASM) to describe and predict the activated sludge processes, which are a useful

54

tool for plant design and optimisation (Henze et al., 2000). While ASM models use a

55

grey-box approach and focus on macroscopic processes, a different modelling approach,

56

relying on metabolic and biochemical pathways, describes the energy, redox and mass

57

balances of the cell processes (Smolders et al., 1994a). When comparing the two

58

strategies, ASM models require a plant-tailored calibration procedure that can affect a

59

higher number of variables, whereas metabolic models have been reported to require a

60

simpler calibration procedure, since all of the equations for the microbial processes are

61

inter-dependant (Seviour et al., 2010). Both approaches have been combined in the

62

Technical University of Delft model (TUDP) and successfully applied for full-scale

63

WWTPs, describing anaerobic, anoxic and aerobic processes of polyphosphate

64

accumulating organisms (PAOs) (Van Veldhuizen et al., 1999; Brdjanovic et al., 2000;

65

Meijer et al., 2001).

AC C

EP

52

ACCEPTED MANUSCRIPT 66 EBPR is carried out by PAOs using three different internal storage compounds that

68

generate energy and/or reducing power, i.e., polyphosphate (poly-P), glycogen and

69

polyhydroxyalkanoate (PHA). Anaerobically, volatile fatty acids (VFAs) like acetate

70

and propionate are converted into PHA by consuming poly-P and glycogen, while PHA

71

is then consumed for P uptake and glycogen production under anoxic and aerobic

72

conditions. Additionally, PAOs have to withstand competition from glycogen

73

accumulating organisms (GAOs), for which external parameters, such as temperature,

74

pH, COD:P ratio and carbon source, play a significant role (Oehmen et al., 2007).

75

Hence, the initial metabolic models developed for PAOs (Smolders et al., 1994a;

76

Smolders et al., 1995; Kuba et al., 1996; Murnleitner et al., 1997), were expanded to

77

include the metabolic pathways of two main GAOs, i.e., Competibacter and

78

Defluviicoccus vanus-related organisms, as well as the effects of temperature, carbon

79

source and pH on their metabolism (Lopez-Vazquez et al., 2009). Moreover, the

80

denitrification capacities of Accumulibacter, the main PAO known, and Competibacter-

81

and Defluviicoccus-GAOs were further incorporated in the model by Oehmen et al.

82

(2010b).

M AN U

SC

RI PT

67

TE D

83

However, these new additions have only been validated in lab-scale systems containing

85

enrichments of PAOs and GAOs and have not previously been tested on full-scale

86

sludge. While simplified metabolic model calibration strategies have been proposed

87

based on lab-scale results (Oehmen et al., 2010b), it is necessary to test these theories

88

using full-scale sludge in order to evaluate their applicability to more complex

89

situations. It is noteworthy to mention that when modelling the performance of full-

90

scale systems, there is an added complexity, since not only could there be unknown

91

PAOs and/or GAOs whose contribution to the phosphorus removal process is still

92

unknown, but known PAOs such as Tetrasphaera could be active, whose metabolism

93

related to EBPR is still largely unclear. Adding to this complexity is the fact that

94

wastewater influents contain a wider diversity of organic carbon sources that are subject

95

to much variability, and PAOs and GAOs make up a much smaller fraction of the total

96

microbial community in full-scale sludge as compared to lab-scale systems.

AC C

EP

84

97 98

Although PAOs have been typically modelled as using the glycolysis pathway as their

99

sole source of anaerobic reducing power generation, it has been suggested that the role

ACCEPTED MANUSCRIPT of the anaerobic TCA cycle in real WWTPs might be greater than expected as compared

101

to lab-scale results (Pijuan et al., 2008; Lanham et al., 2013). In fact, Zhou et al. (2009)

102

have shown that the TCA might have a particularly prominent role when PAOs face

103

conditions of glycogen shortage. Since WWTPs deal with variable influent

104

compositions and often with limited carbon substrate availability, this might be the

105

reason for a greater reliance on the TCA cycle in WWTPs as opposed to lab-scale

106

reactors (Lanham et al., 2013). Therefore, in order to improve the applicability of

107

metabolic models, particularly with respect to full-scale situations, the relevance of

108

incorporating the TCA cycle activity into the model should be assessed.

RI PT

100

SC

109

Furthermore, in previous metabolic models the aerobic maintenance processes predict

111

cell decay at low PHA levels, which is not consistent with experimental findings.

112

Experiments on the endogenous metabolism of PAOs (Lopez et al., 2006; Lu et al.,

113

2007; Vargas et al., 2013) observed that the aerobic maintenance processes were

114

dependant on glycogen and polyphosphate degradation following PHA depletion, with

115

minimal cell decay. This is a particularly relevant factor to include when applying the

116

model to full-scale systems, where the level of polymers stored by the sludge is much

117

lower as compared to lab-scale systems.

TE D

118

M AN U

110

In this study, a simplified version of the metabolic models previously developed by

120

Lopez-Vazquez et al. (2009) and Oehmen et al. (2010b) was adapted in order to

121

incorporate the anaerobic TCA utilisation of PAOs, in addition to the previously

122

implemented glycolytic pathway. The resulting model was tested by describing the

123

anaerobic/aerobic chemical transformations observed in activated sludge batch tests fed

124

with acetate as carbon source. The tests were carried out with sludge from four different

125

EBPR WWTPs with differing microbial compositions (including different fractions of

126

Accumulibacter, Competibacter and Defluviicoccus) and metabolisms, as shown in

127

Lanham et al. (2013). Special attention was paid to the required calibration procedure

128

necessary in order to describe the activity of each biomass, and where possible,

129

simplified calibration procedures that could be applicable to the modelling industry

130

were evaluated. In addition, theoretical simulation studies were conducted between

131

PAOs using solely the TCA cycle anaerobically (PAO_TCA) and PAOs using

132

glycolysis (PAO_Glyc) in order to better understand the conditions which may lead to

133

the use of one metabolic pathway over the other. Thus, this study is also relevant to

AC C

EP

119

ACCEPTED MANUSCRIPT 134

improve our knowledge about factors that influence the microbial metabolism in EBPR

135

systems, which is necessary in order to better understand and optimise the performance

136

of the process.

137 138

2

Materials and Methods

139

2.1

140

Four WWTPs were modelled in order to describe the anaerobic and aerobic chemical

141

transformations observed in activated sludge tested in lab-scale batch tests, fed with

142

acetate, at neutral pH (7) and at 20°C. The WWTPs studied had either an A2/O

143

configuration (Portuguese WWTPs: PT_1 and PT_2) or a Biodenitro configuration

144

coupled to a return sludge side-stream hydrolysis process (RSS) (Danish WWTPs:

145

DK_1 and DK_2). Experiments were carried out in winter and summer for the

146

Portuguese WWTPs and only in winter for the Danish WWTPs (Lanham et al., 2013).

147

Two of the WWTPs (PT_1 and DK_2) only had Accumulibacter-PAOs (approximately

148

4% as determined by quantitative fluorescence in situ hybridisation (qFISH)), whereas

149

PT_2 had significant amounts of Defluviicoccus- and Competibacter-GAOs (4-8%).

150

DK_1 presented one time point with almost 1% of Competibacter. Also, Type I and

151

Type II Accumulibacter were quantified using specific oligonucleotide probes, Acc-I-

152

444 and Acc-II-444 targeting type-I and type-II Accumulibacter-PAOs (Flowers et al.,

153

2009). The biovolume of these Accumulibacter types was determined as described in

154

Lanham et al., 2013. A complete account of the WWTPs characterisation, the batch test

155

results and the microbial population quantification can be seen in Lanham et al. (2013).

158

RI PT

SC

M AN U

TE D

EP

AC C

156 157

Experimental results

2.2

Model description

159

The model developed in this study was based on a simplified version of previous

160

metabolic models defined by Lopez-Vazquez et al. (2009) and Oehmen et al. (2010b)

161

and was compiled using AQUASIM software (v. 2.1, Reichert (1994)).

162

The present model focused only on acetate as the sole external carbon source, converted

163

into polyhydroxyalkanoate (PHA). It does not address pH nor temperature dependencies

164

(since these were controlled at 7 and 20oC in all batch tests) and aims at describing the

ACCEPTED MANUSCRIPT 165

anaerobic and aerobic transformations of Accumulibacter-PAOs (abbreviated to ACC)

166

and Competibacter and Defluviicoccus-GAOs (abbreviated to GB and DEF).

167 The model describes the acetate ( SHAc ) and the phosphate ( SPO4 ) concentrations

169

observed in the medium, the concentration of PHA ( X PHA ), glycogen ( X GLY ) and

170

polyphosphate ( X PP ) inside the bacterial cells, as well as the concentration of PAOs and

171

GAOs ( X ACC , X GB or X DEF ). The initial values for these parameters in the model were

172

based on the experimental results. The initial concentrations of the PAO and GAO

173

biomass fractions were based on the active biomass concentrations (as given by the

174

volatile suspended solids (VSS) minus the organic storage polymers, PHA and

175

glycogen, cf., Smolders et al. (1994a), converted to CmM via the biomass formula of

176

CH1.84O0.5N0.19 (Zeng et al., 2003a)), and multiplied by the fraction of Accumulibacter

177

(XACC), Competibacter (XGB)and Defluviicoccus (XDEF) detected by qFISH (Oehmen et

178

al., 2010b). In systems containing PAOs and GAOs simultaneously, their initial

179

fraction of glycogen and PHA was estimated based on the specific anaerobic yields for

180

each compound (cf., Table 1) in each type of organism as exemplified in Eq. 1 for the

181

GB initial PHA fraction of Competibacter ( X PHA ,ini ).

182 183

X

GB PHA ,ini

= X PHA,ini ×

GB fGB × YPHA , HAc

GB ACC DEF f GB × YPHA , HAc + f ACC × YPHA, HAc + f DEF × YPHA, HAc

(1)

EP

184

TE D

M AN U

SC

RI PT

168

where f GB , f ACC and f DEF are the fraction of each of these organisms as exemplified in

186

Eq. 2 for Competibacter.

187 188

AC C

185

f GB =

X GB

X GB + X ACC + X DEF

(2)

189 190

The anaerobic stoichiometry of PAOs was based on acetate uptake coupled with

191

polyphosphate hydrolysis and phosphate release, glycogen degradation and PHA

192

production. Anaerobic maintenance processes were modelled as polyphosphate

193

hydrolysis and subsequent phosphate release, which was followed by glycogen

ACCEPTED MANUSCRIPT 194

degradation, if low polyphosphate levels were attained. A complete account of the

195

anaerobic reactions and kinetics in the model is given in Appendix A-I.

196 The anaerobic yields were defined based on the utilisation of the anaerobic TCA cycle

198

or the glycolysis pathway, as determined experimentally by the anaerobic glycogen per

199

acetate yield obtained in the activated sludge batch tests (Lanham et al., 2013). The

200

overall reactions for the TCA cycle or the glycolysis stoichiometry were based on the

201

Comeau-Wentzel (Comeau et al., 1986; Wentzel et al., 1986) and Mino (Mino et al.,

202

1987) models, respectively, as presented by Smolders et al. (1994) and shown in Eqs. 3

203

and 4, respectively. Greater explanation of the incorporation of the TCA cycle

204

stoichiometry is detailed in Section 3.1.1.

205

SC

RI PT

197

The overall acetate uptake reaction where the TCA cycle is incorporated is shown

207

below (C-mol basis):

M AN U

206

208 209

1 Acetate + (0.5 + α ACC ) HPO3 + ( − 0.5 + α ACC ) H 2O → 3 0.89 PHA + 0.11CO2 + (0.5 + α ACC ) H 3 PO4

TE D

210

(3)

211

The overall acetate uptake reaction where glycolysis is incorporated is shown below

212

(conversion of acetate, glycogen and polyphosphate into PHB, CO2 and phosphate):

213

215 216 217

 5  1.33PHA + 0.17CO2 + (0.25 + α ACC ) H 3 PO4 +  − (0.25 + α ACC )  H 2O  12 

(4)

AC C

214

EP

Acetate + 0.5Glycogen + (0.25 + α ACC ) HPO3 →

where αACC is the energy of transport of one C-mol of acetate across the cell membrane.

218

GAOs were modelled in the same way as PAOs, except with a different stoichiometry

219

(see Table 1) and excluding the processes dependant on polyphosphate or phosphate.

220

The overall anaerobic acetate uptake of GAOs is shown in Eq. 5 (Filipe et al., 2001a;

221

Zeng et al., 2002):

222

5 1     Acetate+ (1 + 2α GAO )Glycogen→ 1.75 + α GAO PHA +  0.25 + α GAO CO2 (5) 3 3    

ACCEPTED MANUSCRIPT 223 The aerobic stoichiometry and kinetics of the PAO model are based on the method

225

proposed by Murnleitner et al. (1997), where PHA is degraded to contribute to

226

phosphate uptake, polyphosphate formation and glycogen replenishment. Growth is

227

determined in the model from the difference between the degraded PHA and the

228

glycogen and polyphosphate produced. The aerobic stoichiometry depends on the ATP

229

production yield per NADH oxidised with oxygen through oxidative phosphorylation,

230

which is known as the P/O ratio or δ. The same processes are applied to the GAO

231

models with the exception of the phosphate uptake and the polyphosphate formation

232

processes. As explained in more detail in Section 3.1.3, the aerobic maintenance

233

processes rely sequentially on PHA, glycogen and polyphosphate as energy sources.

234

The kinetic processes and parameters utilised in this model, summarised in Appendix

235

A-H, were mostly consistent with those presented in Lopez-Vazquez et al. (2009) and

236

Oehmen et al. (2010b), except for the differences specified in Section 3.1.

M AN U

SC

RI PT

224

237 2.3

Model calibration and validation strategy

239

The PAO model was first calibrated using the averaged results from two replicate batch

240

tests of plant PT_1, in winter time, at standard conditions (acetate as the carbon source,

241

pH=7 and 20ºC). The calibration procedure was performed on the key kinetic

242

max parameters, namely the maximum anaerobic acetate uptake rate ( qHAc ), the aerobic

243

polyphosphate formation rate ( q PP ), the aerobic PHA consumption rate ( q PHA ) and the

244

aerobic glycogen production rate ( qGLY ).The calibrated values were determined based

245

on a deviation between the experimental values and the modelled results of less than

246

10%, calculated by the normalised root mean squared deviation (NRMSD) (cf. 2.4).

247

The resulting calibrated model was then used to describe the experimental results

248

obtained for the other experiments at different time-points or in different WWTPs. In

249

order to achieve this, the strategy detailed in Oehmen et al. (2010b) was used, where

250

max qHAc was first recalibrated, if needed, followed by adjusting all other aerobic kinetic

251

rates proportionally. During this validation procedure, the initial concentrations of

252

acetate, phosphate, PHA, glycogen and polyphosphate were set to the initial

253

experimental values.

AC C

EP

TE D

238

ACCEPTED MANUSCRIPT 254

The Competibacter model and the Defluviicoccus model were first calibrated in one

255

batch test of plant DK_1 and two averaged replicate batch tests of plant PT_2 (summer),

256

respectively. The same procedure as for PAOs was applied to adjust the models to the

257

experimental results of plant PT_2 (winter).

258 2.4

Error and sensitivity analysis

260

The deviation for n data points of the values predicted by the model ( ximod el at time

261

point i ) and the experimental values ( xiexp at time point i ) was determined by

262

calculating the normalised root mean squared deviation (NRMSD) as stated in Eq. 6: n

∑ (x

exp i

i =1

NRMSD =

− ximod el ) 2

n exp − xmin

M AN U

263

SC

RI PT

259

exp max

x

264

exp exp where xmax and xmin are the maximal and minimal experimental values measured for

265

that parameter.

266

(6)

Sensitivity analyses were performed to assess the impact of the aerobic PHA

268

degradation rate and of the initial polyphosphate concentration by varying the value of

269

these parameters by ± 50% and then determining the effect on growth, levels of storage

270

compounds (PHA, glycogen and polyphosphate) and on phosphorus removal in long-

271

term simulations of 40 d.

272

274 275

2.5

Simulation studies

AC C

273

EP

TE D

267

276

Long-term simulations were performed by defining a sequencing batch reactor (SBR) in

277

AQUASIM as a mixed reactor compartment (maximum volume = 8L), with a sludge

278

retention time (SRT) of 10 d, a hydraulic retention time of approximately 0.6 d, and a 7-

279

h cycle. The simulations were executed for 40 d to ensure that repeatable profiles were

280

achieved (4 x SRT), similar to what is described in Oehmen et al. (2010b).

281

ACCEPTED MANUSCRIPT The simulations were performed either for sensitivity analysis purposes (see 2.4) or for

283

microbial competition estimates. In the latter, simulations were performed to determine

284

the theoretical competition between the two anaerobic metabolisms that are at the focus

285

of this work. Two theoretical distinct PAO organisms were defined, one using only the

286

anaerobic TCA metabolism and the other only using the glycolysis metabolism. While it

287

is recognised that PAOs can potentially perform both processes simultaneously, this

288

approach allowed examination of the factors that influence the competition between

289

different metabolic pathways. Due to hypotheses suggested by Lanham et al. (2013), the

290

competition between the two metabolisms was assessed at different acetate

291

concentrations in the feed, from 0.25 to 5 C-mM, and at different aerobic phase

292

durations, from 3 to 12 h. Simulations that varied the acetate concentration were

293

performed at a constant aerobic phase duration of 4 h and simulations that varied the

294

duration of the aerobic phase were performed at a constant acetate concentration in the

295

feed of 2 C-mM.

296

3

298

3.1

299

Results and Discussion Model development

TE D

297

M AN U

SC

RI PT

282

The goal of this study was primarily to adapt metabolic models developed for lab-scale

301

systems to describe the processes occurring in full-scale sludge. Although lab-scale

302

systems are essential to achieve a deeper understanding about key metabolic

303

transformations, the higher complexity of full-scale WWTPs might require adjustments

304

to the models developed upon lab-scale data, regarding the diversity of influent

305

characteristics, configurations and operating conditions existing in WWTPs. In

306

particular for EBPR systems, the diversity of PAOs and GAOs and of their metabolic

307

activities in real plants might be much higher than that obtained in lab-scale reactors.

308

Moreover, the enrichments obtained in the latter can contain an abundance of these

309

specific populations that are amplified as compared to full-scale systems, thus diluting

310

possible competition and synergetic interactions of flanking populations. In this study,

311

some parameters and concepts had to be adjusted from the models developed by Lopez-

312

Vazquez et al. (2009) and Oehmen et al. (2010b), as detailed below.

313

AC C

EP

300

ACCEPTED MANUSCRIPT 314

3.1.1

Incorporating the TCA cycle stoichiometry

315 Both Pijuan et al. (2008) and Lanham et al. (2013), two EBPR studies concerning full-

317

scale activated sludge, have observed the partial utilisation of the anaerobic TCA cycle

318

to different extents, as shown by the stoichiometric yields, which, as reviewed by

319

Oehmen et al. (2010a), can indicate that different biochemical pathways are being

320

employed. Smolders et al. (1994b) proposed two metabolic models for the anaerobic

321

metabolism of PAOs, one using the TCA cycle and the other employing glycolysis. The

322

anaerobic TCA cycle utilisation is associated with a higher phosphorus release to

323

ACC acetate uptake yield ( YPO 4, HAc ) and a lower glycogen consumption and PHA production

324

ACC ACC per acetate uptake ( YGLY , HAc and YPHA, HAc , respectively) than what is observed when using

325

the glycolytic pathway (Smolders et al., 1994b, cf., Table 1). Lanham et al. (2013)

326

observed different degrees of TCA cycle utilisation in different plants, and even in the

327

same plant at different periods in time, which were proportional to the glycogen

328

consumption to acetate uptake yield. Furthermore, this study verified correlations of the

329

glycogen consumption yield coefficients with the PHA production and P release yields

330

that supported the glycolysis and TCA models. Therefore, the incorporation of the TCA

331

ACC ACC cycle metabolism was performed by adjusting the YPO 4, HAc and the YPHA, HAc yields such

332

ACC that they are dependent on the YGLY , HAc yields observed in the batch tests.

SC

M AN U

TE D

333

RI PT

316

Another important aspect is that, within the cases where the anaerobic TCA cycle was

335

ACC relevant, a discrepancy was observed between some of the YPO values obtained 4 , HAc

336

experimentally and those predicted by the TCA model. The reason for this could likely

337

ACC stem from the fact that YPO is dependent on the energetic requirement for acetate 4 , HAc

338

uptake across the cell membrane (αACC), which may vary depending on whether PAOs

339

employ the TCA cycle or glycolysis. Thus, the αACC value that Smolders et al. (1994)

340

found using the glycolysis pathway (via a lab-scale culture) is not necessarily

341

transferable to the TCA cycle model. This energetic requirement (αACC), is known to be

342

dependent on pH (Smolders et al., 1994a; Filipe et al., 2001b) but also on the VFA

343

uptake mechanism of the cell (Oehmen et al., 2010a). Since pH was controlled at 7 in

344

all experiments, we hypothesised that the reason for this occasional discrepancy was

345

due to the αACC parameter changing as a function of the PAO population, with different

AC C

EP

334

ACCEPTED MANUSCRIPT Accumulibacter sub-groups possessing different VFA uptake mechanisms. Supporting

347

this hypothesis is the fact that a correlation was observed in this study between the αACC

348

parameter and the abundance of Accumulibacter Type I and Type II vs. total

349

Accumulibacter (Table 2). In cases where total Accumulibacter correlates closely

350

(between 75 and 100%) with the sum of Type I plus Type II, a higher αACC, and

351

ACC consequently a higher YPO was observed. When Type I plus Type II accounted for 4 , HAc

352

between 25 and 75% of the total Accumulibacter population, the αACC agrees well with

353

that predicted by the glycolysis model. Interestingly, Zhou et al. (2009) also found a

354

ACC value than the one predicted by Smolders et al. (1994) when the TCA higher YPO 4 , HAc

355

cycle was active, corroborating the results of this study. Therefore, two new terms were

356

ACC introduced into the model to describe the value of YPO 4, HAc in the case of the glycolysis

357

and TCA cycle pathways (Eqs. 7 and 8). The first term, defined in equation 7, accounts

358

ACC for a higher YPO 4, HAc value when the TCA cycle is active instead of glycolysis, due to the

359

necessity to increase the hydrolysis of polyphosphate to accommodate for the ATP

360

production that is no longer being generated by glycolysis. This is dependent on the

361

ACC observed glycogen per acetate yield ( YGLY , HAc ) and the model predicted yields via

362

glycolysis and the TCA cycle shown in Table 1. The second term impacts the αACC

363

parameter, as shown in equation 8. This term is expressed as a function of f ACCI , ACCII ,

364

which describes the composition of the Accumulibacter population (i.e. whether the

365

total Accumulibacter population assessed through the PAOMIX probes was mainly

366

comprised of Type I plus Type II as assessed through the FISH probes, or of other

367

clades). Furthermore, it is important to note that this increase in ATP requirement for

368

acetate transport was not observed when the glycolysis pathway was active, but only

369

when the TCA cycle was used and the total Accumulibacter population was well

370

described by the sum of Type I and Type II, as reflected in equation 8 through

371

ACC dependence on the YGLY , HAc term. This suggests that different Accumulibacter sub-

372

populations possess different acetate uptake mechanisms. Additionally, when

373

comparing the VFA uptake energetic requirements of PAOs and GAOs (

374

Table 3), it can be observed that α appears to be dependent on the level of glycolysis

375

activity of the cell, in addition to the VFA uptake mechanism. For instance, PAOs and

376

GAOs are capable of using additional VFA uptake mechanisms besides the direct use of

AC C

EP

TE D

M AN U

SC

RI PT

346

ACCEPTED MANUSCRIPT ATP (see Oehmen et al. (2010a) for a review). These mechanisms become more

378

prevalent when higher quantities of glycogen are used by the cell, and thus, they are

379

more significant in GAOs than PAOs. Furthermore, Defluviicoccus GAOs have been

380

found to have two simultaneous extra proton motive force (pmf) generation capacities

381

(i.e. through fumarate reductase and methylmalonyl-CoA decarboxylase), while

382

Competibacter GAOs only have one (through fumarate reductase) (Saunders et al.

383

2007; Burow et al. 2008). This correlates very well with the ATP requirement for

384

transport found in these organisms through metabolic models (Filipe et al. 2001;

385

Oehmen et al. 2006).

RI PT

377

387

ACC YPO 4, HAc =

SC

386 ACC YGLY 1 , HAc ACC _ Gly ACC _ TCA ACC Y + (YGLY − ) × (1 − ) + α HAc , HAc GLY , HAc ACC _ Gly 4 YGLY , HAc

389

α

ACC HAc

M AN U

388

(7)

ACC YGLY 1 = −1.1 + 0.19 × pH + (1 − ACC,_HAc ) × f ACCI , ACCII 4 YGLY , HAcGly

390

(8)

where f ACCI , ACCII is 1 for a higher coverage of total Accumulibacter by Types I and II

392

(75-100% of total Accumulibacter) and 0 for a lower coverage (25-75% of total

393

Accumulibacter). As previously reported by Smolders et al. (1994a), the range of α is

394

constrained between 0 and 0.5 mol ATP / C-mol acetate uptake. The values of the yields

395

for the two Accumulibacter metabolisms (ACC_TCA and ACC_Gly) are stated in Table

396

1.

EP

TE D

391

397

ACC ACC YPHA , HAc was defined using a similar strategy as for YPO4 , HAc (Eq. 9), where more PHA is

399

produced via the glycolysis pathway per mole of acetate uptake as compared to the TCA

400

cycle.

401

AC C

398

ACC PHA, HAc

Y

=Y

ACC _ TCA PHA , HAc

+ (Y

ACC _ Gly PHA, HAc

−Y

ACC _ TCA PHA , HAc



ACC YGLY , HAc ACC _ Gly YGLY , HAc

(9)

402 403

3.1.2

Anaerobic maintenance processes

404

ANA The anaerobic maintenance coefficient ( mACC ) was determined in a blank test without

405

the addition of acetate, and its value was determined by calculating the rate of ATP

ACCEPTED MANUSCRIPT production from the rate of P release observed during this test. The batch tests

407

conducted in WWTPs that carried out a chemical phosphate precipitation polishing step,

408

namely DK_1 and DK_2, presented a relatively high anaerobic P release, which could

409

be partially due to anaerobic dissolution of iron-phosphate precipitates caused by iron-

410

reducing bacteria (Nielsen, 1996). Since it was not possible to distinguish between the P

411

release attributed to anaerobic maintenance and chemical dissolution in the tests

412

performed, it was opted to model this additional P-release as an increased anaerobic

413

maintenance coefficient, instead of adding a new chemical P release process (Table 4).

414

ANA Therefore, the anaerobic maintenance coefficient ( mACC ) should be considered as a

415

lumped parameter for the Danish WWTPs.

SC

RI PT

406

416 3.1.3

418

In previous models, the aerobic maintenance was modelled as a function of cell decay,

419

which is indirectly related to PHA degradation (Murnleitner et al., 1997). For this

420

reason, at low PHA concentrations, the model predicted cell decay, which contrasts with

421

studies where the endogenous mechanisms of PAOs and GAOs were characterised

422

(Lopez et al., 2006; Lu et al., 2007; Vargas et al. 2013). Based on the results of

423

Brdjanovic et al. (1998), Lopez et al. (2006), Lu et al. (2007) and Vargas et al. (2013),

424

the aerobic maintenance process was expanded to include a sequential dependence on

425

PHA, glycogen and then polyphosphate. These studies have shown that PAOs tend to

426

rely on energy generation from their storage polymers prior to cell decay, an effect that

427

is of high importance particularly in the substrate-limited conditions that normally exist

428

in full-scale WWTPs, unlike lab-scale systems. The kinetic equations of the three

429

sequential aerobic maintenance processes are defined in Eqs. 10 to 12:

430

AER ACC mACC , PHA = mPHA × X ACC ×

TE D

EP

AC C

431 432

Aerobic maintenance processes

M AN U

417

m

AER ACC ,GLY

=m

ACC GLY

ACC X PHA ACC ACC X PHA + K PHA

(10)

ACC ACC   X GLY X PHA × X ACC × 1 − ACC × ACC ACC  ACC  X PHA + K PHA  X GLY + KGLY

(11)

433 434

ACC  X PHA AER ACC mACC = m × X × 1 −  , PP PP ACC ACC ACC  X PHA + K PHA

ACC   X GLY × 1 −   ACC ACC   X GLY + KGLY

ACC  X PP ×  ACC ACC  X PP + K PP

(12)

ACCEPTED MANUSCRIPT 435 436

AER AER AER where, mPHA , mGLY and mPP are the aerobic maintenance coefficients based on PHA,

437

AER glycogen and polyphosphate respectively. mPHA was defined according to Lopez-

438

AER AER Vazquez et al. (2009) (Eq.15), while mGLY and mPP are defined in this study. From

439

Eqs. 13 and 14, the degradation of one C-mol of glycogen or of one P-mol of

440

polyphosphate produces

441

proportional to how much glycogen or polyphosphate would be needed to produce the

442

AER ATP requirements for aerobic maintenance, defined as the constant variable mATP

443

(Smolders et al., 1994a), hence obtaining the expressions presented in Eqs. 16 and 17.

444 Glycogen degradation to produce ATP (C-mol basis):

M AN U

445 446 447

1 2 1 1 1+ δ Glycogen + O2 → PHA + CO2 + H 2O + ATP 4 3 3 3 2

448

450 451

HPO3 + H2 0 → ATP + H2 PO4

452 AER mPHA =

AER 12 × mATP 6 + 27δ

454

456 457

AER 2 × mATP 1+ δ

AC C

455

EP

453

(13)

Polyphosphate degradation to produce ATP (Smolders et al. 1994b):

TE D

449

SC

RI PT

1+ δ or 1 ATP-mol, respectively. These coefficients are 2

AER mGLY =

AER AER mPP = mATP

(14)

(15)

(16)

(17)

458 459

The stoichiometric matrix for these maintenance coefficients is given in Table 5.

460 461

It was assumed that the aerobic maintenance processes would be similar in GAOs, as in

462

PAOs, i.e. with the sequential utilisation of PHA and glycogen for aerobic maintenance,

ACCEPTED MANUSCRIPT 463

excluding of course the process dependent on polyphosphate. This is consistent with the

464

results found by Vargas et al. (2013), where GAOs aerobic maintenance processes

465

relied on glycogen degradation following PHA depletion.

466 467

3.2

Model calibration and validation in the different WWTPs

RI PT

468

The PAO model was first calibrated using results from PT_1 during the winter (Figure

470

1). The strategy used in this study was to input the abundance of Accumulibacter, as

471

determined by quantitative FISH, multiplied by the active biomass concentration

472

(Oehmen et al., 2010b). This allowed to account for the microbial community covered

473

by the model, which in the case of a WWTP sludge is only a minor fraction of the total

474

community (in this study, usually approximately 5% of the total population, and never

475

surpassing 12%). This step alone yielded a fairly correct description of the chemical

476

transformations, with some adjustment required with respect to the kinetic parameters.

M AN U

SC

469

477

Similarly to lab-scale studies, the calibration procedure involved the adjustment of the

479

max , the q PHA , the q PP and the qGLY parameters (Table 6). In general, the rates were qHAc

480

slightly lower than the values obtained by Lopez-Vazquez et al. (2009), except for

481

glycogen, which had a slightly faster rate than found in lab-scale systems (Table 6).

482

After the calibration procedure, the model fitted very well with the experimental results,

483

with error values below 5% for PHA, glycogen and P, as determined by the NRMSD

484

(Table 8). The strong fitting of the model to the data further supports the anaerobic TCA

485

cycle activity in some sludges (e.g. Figure 1a). It should be noted that both the

486

glycolysis and TCA models are based on a steady-state assumption for intermediates

487

including NADH, ATP and acetyl-CoA. Some studies have found that the levels of

488

some intermediates in cells (e.g. NADH) can fluctuate over time and space in WWTPs

489

(Wos and Pollard, 2009). Nevertheless, if NADH were to be consumed anaerobically in

490

PAOs, without generation through the TCA cycle (or glycolysis), then acetate would be

491

completely converted to PHA at a ratio of 1 C-mol PHA/C-mol HAc (since some

492

acetyl-CoA must be oxidised to CO2 to generate reducing equivalents via the TCA

493

cycle), leading to an underestimation of anaerobic PHA production by the model. Such

494

a discrepancy was not observed in this study, which validates the hypothesis that the

495

TCA cycle and/or glycolysis were responsible for anaerobic NADH generation.

AC C

EP

TE D

478

ACCEPTED MANUSCRIPT 496 Other minor modifications from the lab-scale models are shown in Table 4, notably of

498

max which was the need to increase the maximal glycogen fraction ( fGLY ), which was found

499

to exceed the previously estimated threshold (0.27 C-mol/C-mol as used in Lopez-

500

Vazquez et al. (2009)) when measured in the activated sludge. This parameter was

501

introduced in the model to prevent the prediction of an unrealistically high level of

502

glycogen accumulation (Meijer et al., 2002). However, considering the low quantity of

503

PAOs and GAOs in full-scale sludge, the measured glycogen value might be more

504

highly influenced by the presence of other hydrolysable sugar-polymers resulting, for

505

example, from exopolymeric substances, from the presence of other EBPR bacteria (e.g.

506

Tetrasphaera or unidentified populations) or other populations present in the sludge. It

507

is noticeable that despite the fact that glycogen levels are approximately 1 C-mmol/L in

508

sludge from PT_1, the microorganisms hardly consume glycogen anaerobically (Figure

509

1), which could suggest a depletion of their internal storage reserves and that the

510

remaining glycogen quantified resulted from other glucose sources.

M AN U

SC

RI PT

497

511

The calibration of the GAO models necessitated the inclusion of PAOs as well, since in

513

the set of WWTPs modelled, there was no situation where GAOs, either Competibacter

514

or Defluviicoccus, were present without PAOs. The DK_1 winter_3 experiment, with a

515

population of Accumulibacter and some Competibacter, was used for calibrating the

516

Competibacter model, whereas PT_2 summer, with a population containing

517

Accumulibacter, Competibacter and Defluviicoccus was used to calibrate the

518

Defluviicoccus model. Despite this challenge, the phosphate profile was effectively

519

described by introducing, as initial concentrations, the GAO and Accumulibacter

520

abundance in the proportions determined by qFISH, confirming once more the success

521

of this methodology (Figure 1 and Table 8). The stoichiometry of glycogen and PHA

522

was also generally well described. It should be noted that the calibrated anaerobic and

523

aerobic kinetic parameters of Competibacter were consistently higher than those of

524

Defluviicoccus (Table 7), which is consistent with the results of Lopez-Vazquez et al.

525

(2009) and Oehmen et al. (2010b). Indeed, these studies showed that it was necessary to

526

model each group of GAOs independently, primarily due to their different carbon

527

source preferences (Competibacter were able to take up acetate at a much faster rate

528

than propionate, while Defluviicoccus preferred propionate uptake).

AC C

EP

TE D

512

ACCEPTED MANUSCRIPT 529 The next step was to apply the calibrated models to other experiments. Once more, the

531

strategy of inputting the PAO and GAO abundance as determined by FISH proved very

532

effective in determining the overall cycling of the target parameters (Figure 2 and Table

533

8). However, in some situations the model required some adjustments in order to

534

correctly describe the experimental data. When applying the model to the Danish plants,

535

the anaerobic maintenance coefficient had to be adjusted in order to also account for a

536

higher phosphorus release due to the presence of chemical precipitants (see 3.1.2 and

537

Table 4). Additionally, the acetate consumption rate had to be adjusted for most

538

experiments. The aerobic kinetic parameters were also adjusted but in the same

539

proportion as for the acetate uptake rate as shown in Table 6 and Table 7. This

540

corroborates the findings from lab-scale studies presented in (Oehmen et al., 2010b).

541

Interestingly, different kinetics were observed in Portuguese plants when comparing

542

summertime (faster) and wintertime (slower) experiments. In Denmark, despite the fact

543

that the experiments were all performed in winter-like conditions, DK_1 revealed the

544

same kinetics as the summertime Portuguese experiments, whereas DK_2 agreed with

545

the Portuguese wintertime kinetics (Table 6 and Table 7), suggesting this effect might

546

not entirely be related to seasonal effects, but could reflect either different levels of

547

activity of PAO and GAO cells, or the presence of other PAOs and GAOs besides

548

Accumulibacter, Competibacter and Defluviicoccus.

SC

M AN U

TE D

549

RI PT

530

One exception to the proportionality between parameters was the aerobic PHA

551

degradation rate, which remained constant in most experiments (except in DK_2, Table

552

6) and therefore was independent of the overall activity factor discussed above. Since

553

PHA is not an exclusive polymer of PAOs and GAOs, it could be produced or

554

consumed by other organisms present in the sludge, thus influencing the rates observed.

555

Nevertheless, Meijer et al. (2002) already discussed that the model was more sensitive

556

to stoichiometry than to kinetics, in steady state conditions. In fact, when varying the

557

aerobic PHA degradation rate by ± 50%, a difference of less than 10% was observed for

558

the PAO biomass concentration and

Metabolic modelling of full-scale enhanced biological phosphorus removal sludge.

This study investigates, for the first time, the application of metabolic models incorporating polyphosphate accumulating organisms (PAOs) and glycoge...
1MB Sizes 0 Downloads 5 Views