Download as pdf or txt
Download as pdf or txt
You are on page 1of 192

University of Surrey

Faculty of Engineering and Physical Sciences

Department of Chemical and Process Engineering

Advanced Oxidation Processes for the


Treatment of Real Slaughterhouse Wastewater
After a Biological Treatment

Pello Alfonso-Muniozguren

Principal supervisor: Dr. Judy Lee

Second supervisor: Dr. Devendra Saroj

Third supervisor: Dr. Madeleine Bussemaker

April 2020
II
STATEMENT OF ORIGINALITY

This thesis and the work to which it refers are the results of my own efforts.

Any ideas, data, images or text resulting from the work of others (whether

published or unpublished) are fully identified as such within the work and

attributed to their originator in the text, bibliography or in footnotes. This thesis

has not been submitted in whole or in part for any other academic degree or

professional qualification. I agree that the University has the right to submit my

work to the plagiarism detection service TurnitinUK for originality checks.

Whether or not drafts have been so-assessed, the University reserves the right to

require an electronic version of the final document (as submitted) for assessment

as above.

Signature: Pello Alfonso-Muniozguren

Date: 03/04/2020

3
SUMMARY

Water shortage is increasing worldwide and becoming a social and

environmental problem. To address the increasing need of water human

population is currently facing, the use of advanced oxidation processes (AOPs)

after a biological treatment seem to be a right approach for the treatment of

industrial wastewaters. AOPs such as ozone (O3), ultrasound (US), ultraviolet light

(UV) or electrochemical oxidation produce in situ powerful oxidants (i.e. OH

radicals) that can easily degrade recalcitrant organic matter and inactivate

microorganisms. Among different industries, current industrialised livestock

agriculture has one of the highest consumptions of water and produces up to ten

times more polluted wastewaters in comparison to domestic sewage. Therefore,

the present study looks into the use of single and combined AOPs for the treatment

of real slaughterhouse wastewaters.

O3 proved to be a powerful oxidant, reducing significantly chemical oxygen

demand (COD) and biological oxygen demand (BOD), as well as lowering colour

and total suspended solids (TSS) to minimum values. O3 also acted as a strong

disinfectant, inactivating most of the microorganisms present in the wastewater.

After an activated sludge process, a complete inactivation of total coliforms (TC)

was obtained with 17 min of ozonation at an O3 inlet concentration of 71 mg O3/Lgas.

For total viable counts (TVC), a drastic reduction was observed after 30 min of

ozonation (5 log inactivation).

The addition of US only increased ozonation performance slightly (44%

COD and 74% BOD reduction), principally due to the high inlet O3 concentration

4
supplied to the system. US alone showed a modest effect in organic matter

removal, reducing 18% COD, 50% BOD and 25% TSS with 300 kHz at an applied

power of 40 W. A minor removal was achieved for TC (< 1 log), while an

insignificant reduction in TVC was measured.

The application of UVC light (254 nm) and hydrogen peroxide (H2O2)

individually did not reduce any dissolved organic carbon (DOC), while the

combination of UVC and H2O2 led to a large synergy. The coupled system reached

26% DOC abatement with an optimised H2O2 concentration after 3 h of treatment,

principally due to the attack of OH radicals formed through the cleavage of H2O2

molecules by UVC. DOC removal was further increased to 41% when O3 was used

as a pre-treatment. This was attributed to the reduction in colour, turbidity and

TSS by O3 pre-treatment, which lowered UV inner filtering effects and hence,

increased the performance of the subsequent UVC/H2O2 process. The application

of a pre-ozonation step also improved organics removal by converting original

organic compounds into easily oxidisable compounds. O3, UVC and H2O2 were

later combined in a single process, but no synergy was observed with a DOC

reduction of 43% and the COD was reduced down to 61 mg O2/L. However, less

O3 (mg O3/Lwastewater) was used with the combined O3/UVC/H2O2 system compared

to that supplied during O3 pre-treatment (O3+UVC/H2O2), increasing O3

performance.

Electrochemical oxidation alone reduced DOC values by 15%, while the use

of an electrochemical cell along with UVC and H2O2 led to a DOC removal of ~81%.

The significant production of OH radicals and electrochemically generated

oxidants during the combined process would be responsible for such a high

removal. A similar removal percentage was achieved when O3 was used as a pre-

treatment prior to EO and UVC combined. Colour values below 25 mg Pt-Co/L

were achieved, typical acceptable colour limit values for treated wastewaters
leaving wastewater treatment plants. O3 pre-treatment also reduced the

production of hazardous compounds such as ClO4- in subsequent treatments.

The individual application of AOPs did not show high efficiency in organic

matter removal, while the combination of different AOPs reached direct discharge

limits set by the European Union and showed potential to be used for water

treatment intended for agricultural irrigation. Particularly, the coupling of UVC

and H2O2 showed a high synergy, with and without O3 pre-treatment, as well as

the combination of EO, UVC and H2O2, where the highest DOC removal was

observed.

6
ACKNOWLEDGEMENTS

Nire gurasoei, Amaia eta Gotzon, bihotzez. Beti ondoan izateagatik. Nire

nahi eta kapritxoak zuen beharren aurretik jartzeagatik. Zuen

eskuzabaltasunagatik. Zuen maitasunagatik. Irakatsitako guztiagatik. Zuengadik

jaso dudan guztiagatik. Zaretenagatik. Naizenagatik eta izango naizenagatik.

Bihotz betez, mila esker.

I would also like to send a special thank to my supervisor, Dr Judy Lee, for

providing the trust and freedom I needed during the last three years. It’s been a

very intense but highly interesting learning experience, in different countries and

with different people. All this would have not been possible without her support.

I don’t want to miss the opportunity to thank Dr Devendra Saroj, Dr

Madeleine Bussemaker and Dr Claudio Avignon-Rossa for their support and

advice. Always ready to give a helping hand. Mohd Bohari and Elodie Richard are

also acknowledged for their help during lab work at CPE.

I would like to thank Professor Lombraña, Cristian and Leire, for welcoming

me to their research lab at the University of the Basque Country. I am also grateful

of having met interesting people, now good friends, during my stay in the LSRE

LCM at the Universidade do Porto, specially Salva, Jose, Jennyfer, Bruna, Joao and

Daniela.
Last but not least, I would like to thank to people from the CPE department.

A special thank and a big hug go to Vasilis, Yiannis, Silvia, Anna, Oliver and

Bethany. Everything is easier when surrounded by good people.

8
PREFACE

Scientific publications

The following articles have been published and/or submitted to international

Q1 journals with the experimental results obtained during the completion of this

thesis.

Chapter 4: P. Alfonso-Muniozguren, J. Lee, M. Bussemaker, R. Chadeesingh,

C. Jones, D. Oakley, D. Saroj, A combined activated sludge-filtration-ozonation

process for abattoir wastewater treatment, Journal of Water Process Engineering,

25 (2018) 157-163.

Chapter 5: P. Alfonso-Muniozguren, M.H. Bohari, A. Sicilia, C. Avignone-

Rossa, M. Bussemake, D. Saroj, J. Lee, Tertiary treatment of real abattoir

wastewater using combined acoustic cavitation and ozonation, Ultrasonics

Sonochemistry, 64 (2020) 104986.

Chapter 7: P. Alfonso-Muniozguren; S. Cotillas; R.A Boaventura; F.C. Moreira;

J. Lee; V.J.P. Vilar, Single and combined electrochemical oxidation driven processes

for the treatment of slaughterhouse wastewater, Journal of Cleaner Production

(accepted).
Attended conferences and summer schools

Oral and poster communications presented in international conferences and

summer schools are listed below:

16th Meeting of the European Society of Sonochemistry (April 2018, Besançon,

France). Poster communication.

20th International Conference on Water Pollution and Treatment (May 2018,

London, UK). Oral communication.

13th International Chemical and Biological Engineering Conference (October

2018, Aveiro, Portugal). Poster communication.

3rd European Summer School on Environemntal Applications of Advanced

Oxidation Processes (June 2019, Alcoy, Spain). Oral communication.

6th conference on Environmental Applications of Advanced Oxidation

Processes (June 2019, Portoroz, Slovenia). Oral communication.

4th Iberoamerican Conference on Advanced Oxidation Technologies

(November 2019, Natal, Brazil). Oral communication.

10
TABLE OF CONTENTS

ADVANCED OXIDATION PROCESSES FOR THE TREATMENT OF REAL

SLAUGHTERHOUSE WASTEWATER AFTER A BIOLOGICAL TREATMENT... I

STATEMENT OF ORIGINALITY .....................................................................................3

SUMMARY ............................................................................................................................4

ACKNOWLEDGEMENTS ..................................................................................................7

PREFACE ................................................................................................................................9

LIST OF FIGURES ..............................................................................................................15

LIST OF TABLES ................................................................................................................19

ABBREVIATIONS..............................................................................................................20

SYMBOLS ............................................................................................................................22

CHAPTER 1 INTRODUCTION .......................................................................................24

BACKGROUND .............................................................................................................24

CHAPTER 2 LITERATURE REVIEW .............................................................................27

INTRODUCTION ...........................................................................................................27

SLAUGHTERHOUSE WASTEWATER .............................................................................27

ACTIVATED SLUDGE ...................................................................................................28

2.3.1 Slaughterhouse wastewater treatment 30

ADVANCED OXIDATION PROCESSES ..........................................................................33

OZONATION ................................................................................................................34

2.5.1 Slaughethouse wastewater treatment 36

ULTRASOUND .............................................................................................................37

2.6.1 Synthetic waters 41


2.6.2 Real wastewaters 43

2.6.3 Disinfection 44

PHOTOLYSIS................................................................................................................ 46

ELECTROCHEMICAL OXIDATION ............................................................................... 48

COMBINATION OF ADVANCED OXIDATION PROCESSES ........................................... 50

2.9.1 Ozone and ultrasound 50

2.9.2 UV and hydrogen peroxide 54

2.9.3 Ozone and UV/hydrogen peroxide 55

2.9.4 Electro-advanced oxidation processes 58

OUTLINE OF THE THESIS AND OBJECTIVES .............................................................. 59

CHAPTER 3 MATERIALS AND METHODS .............................................................. 62

INTRODUCTION .......................................................................................................... 62

CHEMICALS ................................................................................................................ 62

ANALYTICAL METHODS ............................................................................................. 63

3.3.1 Colour, turbidity and pH 63

3.3.2 KI dosimtery 64

3.3.3 Inorganics 64

3.3.4 Free oxidants 64

3.3.5 Standard and Hach methods: organics, inorganics and microorganisms 65

REAL SLAUGHTERHOUSE WASTEWATER................................................................... 70

A COMBINED ACTIVATED SLUDGE-FILTRATION-OZONATION PROCESS ................. 71

3.5.1 Experimental setup 71

COMBINED ACOUSTIC CAVITATION AND OZONATION ........................................... 73

3.6.1 Experimental setup 73

OZONE AS A PRE-TREATMENT AND COMBINED TO UVC/H2O2 .............................. 75

3.7.1 Experimental setup 75

3.7.2 Experimental procedure 77

SINGLE AND COMBINED ELECTROCHEMICAL OXIDATION DRIVEN PROCESSES ...... 79

3.8.1 Experimental setup 79

3.8.2 Experimental procedure 80

12
CHAPTER 4 A COMBINED ACTIVATED SLUDGE-FILTRATION-OZONATION

PROCESS FOR SLAUGHTERHOUSE WASTEWATER TREATMENT .................82

INTRODUCTION ...........................................................................................................82

RESULTS AND DISCUSSION .........................................................................................83

4.2.1 Slaughterhouse wastewater characterisation 83

4.2.2 Activated sludge process 83

4.2.3 A combined activated sludge-filtration-ozonation process 85

CONCLUSIONS.............................................................................................................95

CHAPTER 5 TERTIARY TREATMENT OF REAL SLAUGHTERHOUSE

WASTEWATER USING COMBINED ACOUSTIC CAVITATION AND

OZONATION ......................................................................................................................96

INTRODUCTION ...........................................................................................................96

RESULTS AND DISCUSSION .........................................................................................96

5.2.1 Slaughterhouse wastewater characterisation 96

5.2.2 OH radical yield from ozonation and sonication 97

5.2.3 Effect of ozonation on water quality indicators 98

5.2.4 Effect of sonication on COD and BOD 99

5.2.5 Impact of sonication on TSS 101

5.2.6 Effect of sonication on TC and TVC 102

5.2.7 Effect of combined ozonation-sonication on COD and BOD 104

5.2.8 Effect of combined ozonation-sonication on TSS 107

5.2.9 Effect of combined ozonation-sonication on TC and TVC 108

CONCLUSIONS...........................................................................................................110

CHAPTER 6 OZONE AS A PRE-TREATMENT AND COMBINED TO UVC/H2O2 AS

TERTIARY TREATMENT FOR REAL SLAUGHTERHOUSE WASTEWATER .112

INTRODUCTION .........................................................................................................112

RESULTS AND DISCUSSION .......................................................................................113

6.2.1 Slaughterhouse wastewater characterisation 113

6.2.2 Ozonation pre-treatment 114

6.2.3 Ozonation pre-treatment followed by combined UVC and H2O2 117


6.2.4 Combined ozone, UVC and H2O2 system 123

CONCLUSIONS .......................................................................................................... 128

CHAPTER 7 SINGLE AND COMBINED ELECTROCHEMICAL DRIVEN

PROCESSES FOR THE TREATMENT OF SLAUGHTERHOUSE WASTEWATER129

INTRODUCTION ........................................................................................................ 129

RESULTS AND DISCUSSION....................................................................................... 130

7.2.1 Slaughterhouse wastewater characterisation 130

7.2.2 Effects of EAOPs on DOC and produced oxidants 131

7.2.3 Effects of EAOPs on COD, TSS and colour 137

7.2.4 Effects of EAOPs in combination with a pre-ozonation step on DOC and produced

oxidants 139

7.2.5 Effects of EAOPs in combination with a pre-ozonation step on nitrogen compounds


144
CONCLUSIONS .......................................................................................................... 147

CHAPTER 8 CONCLUSIONS ....................................................................................... 149

INTRODUCTION ........................................................................................................ 149

PERFOMANCE OF THE APPLIED PROCESSES ............................................................ 150

OPTIMISATION.......................................................................................................... 152

FUTURE PROSPECTS .................................................................................................. 153

APPENDIX A .................................................................................................................... 155

APPENDIX B..................................................................................................................... 160

APPENDIX C .................................................................................................................... 164

REFERENCES ................................................................................................................... 167

14
LIST OF FIGURES

Figure 2.1. Typical activated sludge process............................................................................ 28

Figure 2.2. Possible processes of a bubble in an acoustic field. Adapted from [72]. .......... 38

Figure 2.3. Sonochemical bubble-size distribution for 213, 355, 647, 875, 1056, and 1136 kHz

[76].......................................................................................................................................... 39

Figure 2.4. Mechanisms of: (a) direc electrolysis and (b) indirect/mediated electrolysis. . 49

Figure 2.5. Possible mechanisms of electrochemical and photo-electrochemical treatment

of secondary wastewater [48]. ............................................................................................ 58

Figure 3.1. A schematic of the experimental setup. ................................................................ 71

Figure 3.2. A schematic diagram of the experimental setup: activated sludge-ozonation

system (ultrasonic plate transducer off), activated sludge-sonication system

(ozonation off) and activated sludge-ozonation-sonication system (ozonation and

ultrasonic plate transducer on). ......................................................................................... 73

Figure 3.3. Schematic of the experimental setup. (a) O3 pre-treatment; (b) UVC/H2O2; (c)

Combined O3/UVC/H2O2. ................................................................................................... 75

Figure 3.4. Scheme of the experimental setup: (a) ozonation system and (b) electrochemical

system. ................................................................................................................................... 80

Figure 4.1. MLSS and MLVSS progress as a function of time in the ASP, as well as

MLVSS/MLSS ratio. 24 h HRT and 13 days SRT. ............................................................ 85

Figure 4.2. Average COD (a) and BOD (b) values for every step of the AFO process at 24 h

and 12 h ASP HRT with 17 min ozonation time. The average COD and BOD values

for the raw wastewater are 1804 ± 204 mg O2/L and 651 ± 89 mg O2/L, respectively.

................................................................................................................................................ 87

Figure 4.3. TSS content for every step of the AFO process at 24 h and 12 h ASP HRT with

17 min ozonation time. The average TSS value for the raw wastewater is

250 ± 90 mg/L. ....................................................................................................................... 88


Figure 4.4. TC progress as a function of ozonation exposure after ASP (empty symbols)

and ASP-filtration (solid symbols). () 12 h HRT and () 24 h HRT. 71 ± 17 mg O3/L

injected. Dotted line indicative of maximum acceptable levels of faecal coliforms for

unrestricted irrigation [191, 192]. ....................................................................................... 92

Figure 4.5. TVC progress as a function of ozonation exposure after ASP (empty symbols)

and ASP-filtration (solid symbols). () 12 h HRT and () 24 h HRT. 71 ± 17 mg O3/L

injected. Dotted line indicative of maximum acceptable levels of bacterial

contamination in drinking water [193]. ............................................................................ 94

Figure 5.1. I3- concentration as a function of time for ozonation (71 mg O3/L), sonication

(44, 300 and 1000 kHz at 40 W) and ozonation-sonication treatments (71 mg O3/L and

44, 300, 1000 kHz at 40 W). The insert shows the zoomed in plot of I3- concentration as

a function of time for sonication only systems. Sample volume 400 mL and 0.1 M KI.

................................................................................................................................................ 98

Figure 5.2. COD and BOD values as a function of sonication time for three different

frequencies (44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed as

standard deviation. ............................................................................................................ 101

Figure 5.3. TSS values as a function of sonication time for three different frequencies (44,

300 and 1000 kHz) at 40 W applied power. Error bars expressed as standard deviation.

.............................................................................................................................................. 102

Figure 5.4. TC and TVC log survival values as a function of sonication time for three

different frequencies (44, 300 and 1000 kHz) at 40 W applied power. Error bars

expressed as standard deviation. ..................................................................................... 103

Figure 5.5. COD and BOD values as a function of O3 alone (71 mg /L) and O3/US

(71 mg O3/L and 44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed

as standard deviation. ....................................................................................................... 105

Figure 5.6. TSS values as a function of O3 alone (71 mg /L) and O3/US (71 mg O3/L and 44,

300 and 1000 kHz) at 40 W applied power. Error bars expressed as standard deviation.

.............................................................................................................................................. 108

Figure 5.7. TC and TVC log survival values as a function of O3 alone (71 mg /L) and O3/US

(71 mg O3/L and 44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed

as standard deviation. ....................................................................................................... 109

16
Figure 6.1. (a) DOC (empty symbols) and COD (filled symbols) as a function of time. O3

inlet concentration: (∆, ▲) 45 mg O3/Lgas; (□, ■) 70 mg O3/Lgas; (◊, ♦) 100 mg O3/Lgas.

Injected gas at a flowrate of 0.3 Lgas/min. (b) Turbidity, TSS and colour as a function of

O3 pre-treatment time. 100 mg O3/Lgas injected at a flowrate of 0.3 Lgas/min. Error bars

expressed as standard deviation...................................................................................... 116

Figure 6.2. (a) Relative DOC concentration as a function of time. (b) H2O2 concentration as

a function of time. (■) 10 min O3 + UVC; (●) 10 min O3 + H2O2 300 mg H2O2/L constant

concentration; (◊) UVC/H2O2 300 mg H2O2/L constant concentration; (∆) 10 min

O3 + UVC/H2O2 300 mg H2O2/L constant concentration (H2O2 added during the 3 h

process to maintain the concentration constant); (□) 10 min O3 + UVC/H2O2 490 mg

H2O2/L initial addition....................................................................................................... 119

Figure 6.3. (a) Relative DOC and (b) colour as a function of time. (∆) 8.1 mg O3/Lgas injected;

(○) 8.1 O3/Lgas injected combined with UVC; (×) 8.1 mg O3/Lgas injected combined with

H2O2; (□) 8.1 mg O3/Lgas injected combined with UVC and H2O2. Initial injection of

490 mg H2O2/Leffluent. ........................................................................................................... 124

Figure 6.4. (a) Relative DOC concentration as a function of time. (b) Relative consumed O3

(injected O3 minus exhaust O3) as a function of time (Consumed O3/Consumed O3 time

zero). (◊) 5.6 mg O3/Lgas injected for combined O3,UVC and H2O2; (□) 8.1 mg O3/Lgas

injected for combined O3,UVC and H2O2; (∆) 12.1 mg O3/Lgas injected for combined

O3,UVC and H2O2; (○) 10 min O3 (100 mg O3/Lgas injected) followed by combined UVC

and H2O2. 490 mg H2O2/L single initial addition for all experiments. ........................ 127

Figure 7.1. (a) Normalised DOC decay and (b) concentration of free oxidants as a function

of Q for the treatment of slaughterhouse wastewater by the following EAOPs: () EO,

() EO/H2O2, () EO/UVC and () EO/UVC/H2O2. Operating conditions: constant

current density of 100 mA/cm2, solution temperature of 25 °C, initial solution pH of

7.5, initial volume of 1.4 L, initial H2O2 addition of 850 mg/L and 11 W UVC lamp.

.............................................................................................................................................. 132

Figure 7.2. Concentration of ClO3- and ClO4- at the reaction end for the treatment of the

slaughterhouse wastewater by the following EAOPs: () EO, () EO/UVC, ()

EO/H2O2 and ( ) EO/UVC/H2O2. Operating conditions: constant current density of


100 mA/cm2, solution temperature of 25 °C, initial solution pH of 7.5, initial volume

of 1.4 L, initial H2O2 addition of 850 mg/L and 11 W UVC lamp. ............................... 136

Figure 7.3. Colour decay as a function of Q for the treatment of slaughterhouse wastewater

by the following EAOPs: () EO, () EO/H2O2, () EO/UVC and () EO/UVC/H2O2.

Operating conditions: constant current density of 100 mA/cm2, solution temperature

of 25 °C, initial solution pH of 7.5, initial volume of 1.4 L, initial H2O2 addition of 850

mg/L and 11 W UVC lamp. .............................................................................................. 138

Figure 7.4. Bio-treated slaughterhouse wastewater before (a) and after O3 pre-treatment.

10 min O3 pre-treatment at 100 mg O3/Lgas and a flowrate of 0.3 Lgas/min. ................ 140

Figure 7.5. Comparison between EAOPs for the treatment of the slaughterhouse

wastewater with and without pre-ozonation in terms of (a) normalised DOC decay as

a function of Q after O3 pre-treatment and (b) concentration of ClO4- at the reaction

end. Processes: (, ) EO, (, ) 10’O3+EO, (, ) EO/UVC, (, ), 10’O3+EO/UVC,

(, ) EO/UVC/H2O2 and (, ) 10’O3+EO/UVC/H2O2. Operating conditions:

constant current density of 100 mA/cm2, solution temperature of 25 °C, initial solution

pH of 7.5, initial volume of 1.4 L, initial H2O2 addition of 850 mg/L and 11 W UVC

lamp. .................................................................................................................................... 142

Figure 7.6. Concentration of free oxidants as a function of Q for the treatment of the

slaughterhouse wastewater with pre-ozonation by the following EAOPs: ()

10’O3+EO, () 10’O3+EO/UVC and () 10’O3+EO/UVC/H2O2. Operating conditions:

constant current density of 100 mA/cm2, solution temperature of 25 °C, initial solution

pH of 7.5, initial volume of 1.4 L, initial H2O2 addition of 850 mg/L and 11 W UVC

lamp. .................................................................................................................................... 144

Figure 7.7. (a) NO3- and (b) NH4+ concentration as a function of Q for the treatment of the

slaughterhouse wastewater by the following EAOPs: () EO, () 10’O3+EO, ()

EO/UVC, () 10’O3+EO/UVC, () EO/UVC/H2O2 and () 10’O3+EO/UVC/H2O2.

Operating conditions: constant current density of 100 mA/cm2, solution temperature

of 25 °C, initial solution pH of 7.5, initial volume of 1.4 L, initial H2O2 addition of 850

mg/L and 11 W UVC lamp. .............................................................................................. 145

Figure 8.1. Schematic of the processes used for the treatment of slaughterhouse

wastewater. ......................................................................................................................... 150

18
LIST OF TABLES

Table 2.1. Physico-chemical characteristics of slaughterhouse wastewater prior to a

biological process. Modified after Arvanitoyannis and Ladas [14]. ............................. 28

Table 2.2. Loading and operation parameters for activated sludge processes of a municipal

wastewater treatment plant. Adapted from [3]. .............................................................. 32

Table 4.1. Characteristics of raw wastewater collected from the slaughterhouse. The

averages were calculated from 8 samples collected over a period of 2 months. ........ 83

Table 4.2. Concentration of COD, BOD, TSS and P before and after 4-7 µm filter paper

filtration during the non-steady state phase of the ASP. ............................................... 89

Table 5.1. Characteristics of treated wastewater. Sample variability from eight different

samples shown as standard deviation. ............................................................................. 97

Table 5.2. COD, BOD, TSS, TC and TVC values before and after the combined ozonation-

sonication process, as well as direct discharge limits (COD, BOD and TSS) and

drinking water standards (TC and TVC)........................................................................ 110

Table 6.1. Physico-chemical characteristics of the pig slaughterhouse wastewater after

secondary biological treatment. Average values from minimum ten samples. ....... 114

Table 6.2. Percentage reduction of DOC by applying combined UVC and H 2O2 after 3

hours. Three different ozonation pre-treatment times (5, 10 and 15 min) at an inlet O3

concentration of 100 mg/Lgas at 0.3 Lgas/min. A constant concentration of H2O2 at three

H2O2 dosages: 150, 300 and 450 mg/L. Standard deviation measured for at least two

independent experiments. ................................................................................................ 121

Table 7.1. Physico-chemical characteristics of the collected slaughterhouse wastewater.

.............................................................................................................................................. 130

Table 7.2. kDOC values for all the applied processes along with the corresponding interval

of adjustment, residual variance (S2R) and coefficient of determination (R2). ........... 134

19
ABBREVIATIONS

AFO Activated sludge-filtration-ozonation

AOPs Advanced oxidation processes

ASP Activated sludge process

BDD Boron-doped diamond

BOD Biological oxygen demand

CFU Colony forming units

COD Chemcial oxygen demand

DIC Dissolved inorganic carbon

DO Dissolved oxygen

DOC Dissolved organic carbon

EAOPs Electrochemical advanced oxidation processes

E. coli Escherichia coli

EO Electrochemical oxidation

EU European Union

HC Hydrodynamic cavitation

HRT Hydraulic retention time

KI Potassium iodide

MLSS Mixed liquor suspended solids

MLVSS Mixed liquor volatile suspended solids

MPN Most probable number

PTFE Polytetrafluoroethylene

SM Standard method

SRT Solid retention time

20
SS Suspended solids

TC Total coliforms

TDC Total dissolved carbon

TKN Total Kjeldahl nitrogen

TN Total nitrogen

TOC Total organic carbon

TP Total phosphorus

TSS Total suspended solids

TVC Total viable counts

UK United Kingdom

US Ultrasound

UV Ultraviolet light

UVC Ultraviolet light type C

WWTPs Wastewater treatment plants


SYMBOLS

A Absorbance in the spectrophotometer in absorbance

units

B1 Dissolved oxygen of blank before BOD5 test in mg O2/L

B2 Dissolved oxygen of blank after BOD5 test in mg O2/L

c Concentration of I3- in molarity (M)

CO3,I-g Constant inlet ozone concentration in mg O3/Lgas

CO3,O-g Outlet/off-gas ozone concentration in mg O3/Lgas

cp Specific heat capacity in J/g K

hv UV light

D1 Dissolved oxygen of diluted sample before BOD5 test in

mg O2/L

D2 Dissolved oxygen of diluted sample after BOD5 test in

mg O2/L

f Ratio between seeded volume and total volume for

BOD5 test

I Applied intensity to the electrochemical cell in ampere

kDOC Degradation kinetic of dissolved organic carbon in min-1

M Mass of liquid in kg

ODT Transferred ozone dose in mg O3/Leffluent

Q Specific charge of electrochemical cell in

amperes hours/Leffluent

QASP Wastewater flowrate in the activated sludge process in

L/day

22
Qg Applied gas flowrate in L/min

Qw Excess sludge in the activated sludge process in L/day

R Ratio between wastewater treated volume and total

volume for BOD5 test

R2 Coefficient of determination in mg2/L2

S2R Residual variance

t Treatment time of the electrochemical cell in hours


𝑣 Treated sample volume in L

V Average reactor volume in the activated sludge process

in L

VL Volume of water in the reactor in L

W Width of the cuvette in m

X Mixed liquor suspended solids in the activated sludge

process in mg/L

Xr Produced biomass in the activated sludge process in g/L


ɛ Absorptivity of I3- in L/mol cm

λ Wavelength in nm
Chapter 1

Introduction

Background

Water pollution is becoming a worldwide concern due to new and tighter

environmental regulations. The exponentially growing human population is

also leading to an increase in water shortage and this, at the same time,

becoming an important issue not only for the environment, but also for the

normal functioning of our society. Therefore, the increasing need for fresh

water needs to be addressed, targeting principally the main wastewater

producer industries. It is estimated that the equivalent to 15% of global water

consumption could potentially come from reused wastewaters [1].

In order to meet certain water discharge or reuse regulations, wastewater

must be treated using a combination of physical, chemical and biological

processes. The treatment usually combines primary, secondary and tertiary

treatments. Secondary treatments usually consist of biological processes that

can be aerobic or anaerobic. Compared to anaerobic systems, aerobic

bioreactors are usually more expensive and more difficult to operate due to the

need for aeration systems and sludge disposal. Aerobic system, however, are

more efficient at removing organic matter [2-4]. The last step, the so-called

24
tertiary treatment which may include microbial disinfection, is used as a polishing

step for organic matter and nutrient removal. Chemicals such as chlorine, chlorine

dioxide, chloramines and hydrogen peroxide (H2O2) are currently used for tertiary

treatment, as well as advanced oxidation processes (AOPs) [5, 6].

The type and combination of processes used are governed by the wastewater

quality and regulatory limits [7]. Within the European Union (EU), standards for

direct discharge from urban wastewater treatment plants (WWTPs) are subjected

to 91/271/EEC Council Directive and dictates the maximum acceptable pollutant

levels as follows: 5 day biological oxygen demand (BOD) 25 mg O2/L, chemical

oxygen demand (COD) 125 mg O2/L, total suspended solids (TSS) 35 mg/L, total

phosphorus (TP) 1-2 mg/L and total nitrogen (TN) 10-15 mg/L [8]. There are no

regulations at the EU level on water reuse for agriculture irrigation, although steps

are being taken to implement a common water reuse legislation [9].

Agricultural sector accounts for 92% of the total freshwater footprint of

humanity and one third of that consumption is directly related to animal products.

The global animal production requires 2422 Gm3 of water per year, leading to a

grey water (pollution of surface or groundwater) production of 159 Gm 3 [10, 11].

The amount of wastewater produced in a slaughterhouse depends on the type and

amount of animals processed, consuming as an average 1.5–10 m3water/tmeat for pigs,

2.5–40 m3water/tmeat for cattle, and 6–30 m3water/tmeat for poultry [12]. With the

production of animal products increasing yearly [13], so does the consumption of

water. This then leads to the increase in the generation of wastewater which can

vary considerably in terms of organic content and microbial population [14-17].

This type of wastewater is characterised by high organic and suspended solids

loading, as well as strong colour and high turbidity. This makes slaughterhouse

wastewater a potential source of harmful organic matter for groundwater

pollution. At the same time, the high nutrient load this type of wastewater carries
can lead to eutrophication of natural water bodies (lack of oxygen due to

uncontrolled growth of algae), preventing the growth of other aquatic plants and

animals [18]. Slaugterhouse wastewaters can also hold non-biodegradable and

recalcitrant substances [19], making biological treatment methods alone

insufficient. Taking this into account, slaughterhouse wastewater has been

catalogued by the United States Environmental Protection Agency as one of the

most harmful wastewaters for the environment [20, 21], rendering its treatment

and disposal an economic and public health necessity [22]. It is therefore necessary

to develop clean and efficient technologies that allow a proper treatment of

slaughterhouse wastewaters before direct discharge or reuse, since this kind of

wastewaters are not completely treated by conventional physical-chemical

processes such as biological systems. In this context, AOPs can be considered a

good option after a biological treatment due to the production of powerful

oxidants such as hydroxyl radicals (•OH) that are able to degrade the most

recalcitrant organic pollutants [23].

This thesis aims to bridge the literature gap by assessing on the potential of

AOPs and electro advanced oxidation processes (EAOPs) as tertiary treatment

methods for the treatment of slaughterhouse wastewater. Considering the nature

of this type of wastewaters, the single and combined AOPs such as ozone (O3),

ultrasound (US), ultraviolet light (UV) and electrochemical oxidation (EO) could

serve as additional treatment methods to current treatment technologies, analysing

possible synergies between the systems. The ultimate aim is to treat the wastewater

to a level where it can be discharged directly into water bodies, as well as assess

the potential to reach water reuse applications. The proposed process

combinations also give a direction towards advancing AOPs application in

slaughterhouse industry for effluent treatment.

26
Chapter 2

Literature review

Introduction

This chapter provides the fundamental aspects of the different wastewater

treatment technologies (activated sludge, O3, US, UV, H2O2 and EO) employed in

this thesis. This is followed by a literature review on the individual and combined

use of the abovementioned processes for the treatment of slaughterhouse

wastewater (organic matter and nutrient removal, as well as microbial

disinfection), showing also the performance obtained by many published studies

when dealing with other real wastewaters and synthetic samples. The most

relevant results are highlighted, and possible research gaps underlined.

Slaughterhouse wastewater

Considerable variations occur in the waste/organic content and microbial

population of slaughterhouse wastewater and it mainly depends on

slaughterhouse cleaning procedures, type and amount of chemicals used and

treatment processes, as well as production systems [24]. Present in the literature

(Table 2.1) are studies confirming the variability of wastewater quality from animal

origin.
Table 2.1. Physico-chemical characteristics of slaughterhouse wastewater prior to a

biological process. Modified after Arvanitoyannis and Ladas [14].

COD BOD Total


Location pH TSS (mg/L) E. coli Study
(mg O2/L) (mg O2/L) Coliforms
Canada - 7685 - 1742 - - [25]
Spain - 40300 24900 3450 - - [26, 27]
Brazil - 2000-6200 1300-2300 850-6300 - - [28]
Austria - 5800 2200-9800 2400-94700 - - [29]
Spain - 3969 1730 2580 - - [30]
Spain - 3980-7125 2035-4200 285-2660 - - [31]

Mexico 7.7 4306 2733 1900 2x107 9x104 [15]


(MPN/100mL) (MPN/100mL)
Canada 4.9-8.1 1250-15900 610-4635 300-2800 - - [21]

Canada 6.1-6.4 1900-2200 2000-2400 1300-1700 (4.6-7.0)×103 290-340 [17]


(CFU/mL) (CFU/mL)
Colombia - 9039 5242 2972 - - [32]

Activated sludge

The activated sludge process (ASP) is a biological treatment method that was

first developed in the United Kingdom (UK) in 1914 [2]. It is an aerobic process

(Figure 2.1) composed of an aeration tank (aeration basin), working as a bioreactor,

a settling tank (clarifier), where the activated sludge solids (biomass) and

wastewater are separated, and a return activated sludge system, essential to keep

microbial population active.

Figure 2.1. Typical activated sludge process.

28
In the aerobic bio-reactor wastewater is mixed with a slurry of microorganisms.

This mixture receives the name of mixed liquor and is composed of bacteria, fungi,

worms and other microorganisms, as well as suspended particles and impurities.

The suspended particles in the mixture are termed mixed liquor suspended solids

(MLSS), while mixed liquor volatile suspended solids (MLVSS) account for the

microbial suspension in the aeration tank [2, 3]. MLVSS/MLSS ratio, therefore,

expresses microorganisms concentration in the aeration tank with respect to the

total suspended solid content and is used to evaluate sludge activity. A low value

(0.2-0.3) would be indicative of a high particle concentration or a low microbial

community, while 0.75 is a common MLVSS/MLSS ratio used in WWTPs [33]. In

the bioreactor, different microorganisms degrade organic matter mainly into

carbon dioxide and water, although other compounds are also formed; e.g. nitrate

and sulphate. The diffused and/or mechanically injected air (or pure oxygen) keeps

the solution mixed and provides oxygen for the bacteria to oxidise the organic

matter and grow [3, 34]. In order to keep the population of microorganisms active,

a minimum dissolved oxygen (DO) concentration of 2 mg/L is used in the activated

sludge processes [34, 35].

The hydraulic retention time (HRT) sets the time that the mixed liquor spends

in the aeration tank before passing to the clarifier. Typical values for the HRT can

go from one hour and up to 30-40 hours, depending on the activated sludge setup,

the quality of the influent and the quality of the effluent to be met. Once in the

clarifier, suspended flocs and particles settle and the clarified effluent is either

discharged or further treated via tertiary treatment. Dependant on the produced

sludge per day and the required microbial population, a certain amount of sludge

is pumped back to the aeration tank (returned activated sludge) to maintain the

required microbial population constant. The excess sludge collected in the clarifier

is disposed of (waste activated sludge). The amount of time that a specific bacteria

floc remains in the activated sludge system is defined by the solid retention time
(SRT), also called sludge age or mean cell residence time. A minimum time (usually

at least one day) is required for microorganism to achieve bioflocculation. Short

SRT may cause settling issues in the clarifier while long SRT values can lead to the

production of a large amount of sludge, requiring larger reactor volumes [3, 34,

36].

Both the HRT and SRT are defined by Eqs. (1-2):

V
HRT = (1)
Q𝐴𝑆𝑃

V×X
SRT = Q (2)
w ×Xr

Where V is average reactor volume, QASP is the wastewater flowrate, and X and

𝑋𝑟 are mixed liquor suspended solids and produced biomass (concentration of

particulate material in the wastage stream), respectively, while 𝑄𝑤 is the excess

sludge production that needs to be removed per day. HRT and SRT are the main

two parameters used to control process efficiency. Other factors affecting the

biological activity and therefore treatment performance are type and concentration

of microorganisms, DO concentration, food (biomass) and temperature of the

aeration tank, as well as BOD, COD and TSS content [3, 34, 36].

2.3.1 Slaughterhouse wastewater treatment

Slaughterhouse wastewater treatment with ASP has been studied in the

literature for more than 40 years proving the efficiency of the method in reducing

COD, BOD and TSS. With the aim of increasing the efficiency of an ASP, different

approaches have been taken regarding SRT and feeding pattern [37], DO

concentration [35], filamentous bacteria control [38, 39], as well as the evaluation

30
of an activated sludge treatment plant itself [40] and a full-scale activated sludge

system [32].

The variation in the operating parameters can be large depending principally

on influent wastewater quality and type of ASP. This leads to different organic

matter and nutrient removals, as well as specific sludge properties and thus, a

significant variation in ASP performance. For a SRT between 5 and 25 days,

reported values of COD removal range between 93% and 97% [37, 41],

recommending a SRT between 5 and 20 days for a complete mixture during the

process [42]. A 2 day SRT led to a poor BOD reduction of 46% in a full scale ASP

process [32], observing that effluent and sludge properties would be negatively

affected at SRT below 3 days [37]. At low SRT values, a microbial population

cannot be maintained because new cells are removed faster than they are produced

[37], reducing the treatment efficiency of the system. A SRT of up to 6 months with

a HRT of 10 days has also been reported in the literature for the treatment of real

slaughterhouse wastewaters and obtained 96% COD and above 99% BOD removal

[43]. The long contact time provided would allow micoorganisms to almost

completly degrade organic matter present in the effluent, although this is not a

common approach in industry as it leads to long treatment times that are usually

not economically favorable.

For urban WWTPs, general values for an ASP and improved activated sludge

processes are shown in Table 2.2 [3]. It is imporant to remark the need for a higher

HRT when dealing with slaughterhouse wastewater due to the presence of less

readily biodegradable molecules [42]. Thus, compared to values presented in Table

2.2 (5-14 h), longer HRT of 29 h [32] and 36 h [37] have been reported for

slaughterhouse wastewater treatment using a conventional ASP.


Table 2.2. Loading and operation parameters for activated sludge processes of a

municipal wastewater treatment plant. Adapted from [3].

BOD loading HRT SRT MLSS BOD removal


Process
(mg/L day) (h) (days) (mg/L) efficiency (%)
Conventional ASP 500-1500 5-14 3-10 2000-3000 90-95
Extended aeration ASP 250-300 20-30 20-30 2000-6000 90-95
High rate ASP 1600-16000 2.5-3.5 0.5-10 5000-8000 60-70
Pure O2 ASP 1600-4000 1-10 3-10 3000-8000 90-95

Another important parameter in the ASP is the feeding pattern. With a 10 day

SRT, a difference in COD removal efficiency was highlighted between continuous

(94%) and intermittent (98%) feeding, obtaining a less variable sludge properties

with the latter [37]. Additionally, a more constant and smaller number of

filamentous microorganisms (filamentous microorganisms have slow settling

properties due to the formation of poorly compacted flocs [44]) was observed with

intermittent feeding, increasing sludge settling properties and thus, improving the

operational performance of the system. The food to microorganism ratio and

treatment temperature can also affect the performance of an ASP [45].

Regarding ASP efficiency in terms of BOD, COD, TSS, TN and total Kjeldahl

nitrogen (TKN) removal for slaughterhouse wastewater, data has been widely

published throughout the years. Lovett et al. [37] reported a maximum COD and

TKN reduction of 96% and 97% at a 20 days SRT, respectively, leading to effluent

concentrations of 54 mg O2/L for COD and 3 mg/L for TKN. BOD values at 5 and

10 days SRT were 45 and 33 mg O2/L, respectively, showing a high efficiency for

organic matter and nutrient removal.

Improved ASP processes have also been used for the treatment of

slaughterhouse wastewater. In order to achieve a better sludge settleability and an

increase in COD and BOD reduction, a two-phase biological treatment (activated

sludge/contact aeration process) was employed with an average removal efficiency

32
for COD, BOD and TSS of 96%, 97% and 95%, respectively [46]. The average value

in the effluent was 39.8 mg O2/L for COD and 18.8 mg O2/L for BOD, notably lower

than those obtained from the activated sludge system alone

(COD = 150 - 200 mg O2/L). TSS concentration was also reduced from 80-100 mg/L

to 22 mg/L. The combined system increased MLSS concentration and lowered the

F:M ratio, providing an increase in the microbial community and thus, removing

more organic matter from the system. Similar removal percentages for COD, BOD

and TSS have also been reported elsewhere [32, 35, 40-43, 45, 46]. This way, ASP

has proved to be effective at reducing organic matter indicators, among other

parameters, even for high organic load slaughterhouse wastewaters.

Advanced oxidation processes

AOPs are based on the in-situ production of reactive species, such as •OH,

which are able to degrade the most recalcitrant organic pollutants and increase

biodegradability [23]. After fluorine (3.03 V), the •OH (2.80 V) is the strongest

known oxidant and capable of completely mineralising most organic matter [47].

Different methods fall within the name of AOPs, generating in situ •OH and

further reacting to produce other reactive species such as H2O2 and superoxide

(O2- ). AOPs can be divided into ozone (O3), ultrasonic, photo(cata)lytic and Fenton

processes, while the combination of multiple AOPs is commonly used for

synergistic effects [5, 47]. Electro-advanced oxidation processes is another subset

of AOPs where electrical energy input is coupled with a supporting electrolyte to

form strong oxidising species that can remove recalcitrant compounds.

Specifically, electrochemical oxidation (EO) with non-active anodes, such as the

boron-doped diamond (BDD) anode, favours the production of large amounts of

•OH [48, 49]. EO performance can be improved by coupling it with different AOPs
since these combined technologies can produce synergistic effects for the removal

of organic matter [50-52].

Ozonation

O3 is a triatomic allotrope of oxygen, formed naturally in the atmosphere from

the combination of bimolecular and atomic oxygen and protects the surface of the

planet from UV-B and UV-C radiation [53]. O3 is an unstable gas and therefore, it

has to be produced in situ when used for medical or water treatment purposes,

among others. O3 is highly toxic in gas form (a few tenths of ppm can cause to

expose individuals headache, coughing, eye and skin irritation, etc.), while no data

on health hazards has been reported when dissolved in liquids. It is highly reactive

and able to oxidise compounds directly and indirectly [54]. During ozonation in

liquid media, oxidation can occur through direct reaction involving molecular O3

(pH < 7 low O3 decomposition rate, pH < 4 direct O3 oxidation pathway) and via

an indirect pathway through •OH formed during O3 decomposition (pH > 7 high

O3 decomposition rate, pH > 10 radical pathway dominated degradation). The

radical chain mechanisms follows three different steps, the so called initiation

reaction, chain propagation, and termination. The first one deals with the

decomposition of O3 that is usually accelerated by initiators such as H2O2, OH-

(Eqs. 3-4), Fe2+ or UV radiation [53, 54].

O3 + OH− → HO2• + O2-• (3)

O3 + OH− → HO2- + O2 (4)

The chain reaction starts when a radical is formed after O3 decomposition (e.g.

superoxide ion radical, O2-•, key for free radical propagation). Due to an unpaired

34
electron, these highly reactive species react rapidly with other molecules forming

other radicals and thus, starting a chain reaction (Eqs. 5-12) [53].

HO2• → O2-• + H+ (5)

O2-• + H+ → HO2• (6)

O3 + O2-• → O3-• + O2 (7)

O3-• + H+ → HO3• (8)

HO3• → O3-• + H+ (9)

HO3• → HO• + O2 (10)

O3 + HO• → HO4• (11)

HO4• → HO2• + O2 (12)

When two radicals neutralise each other by unpairing their missing electron,

the chain reaction terminates (Eq. 13) [53].

HO4• + HO4•→ H2O2• + 2 O3 (13)

During this process, there are many compounds that can act as promoters (such

as methanol or humic acids that propagate the radical chain through the formation

of O2-•) and some others, such as carbonate and bicarbonate, that act as •OH

scavengers [53, 54].

O3 is unstable in water and selectively attacks organic and inorganic

compounds. •OH, on the contrary, react non-selectively with many dissolved

compounds (organic and inorganic contaminants) and the water matrix [53-59]. By

oxidation of the specific cell wall components and subsequent DNA damage by O3

and •OH, O3 kills bacteria and disinfects water [59, 60]. O3 can also increase the

biodegradability of organic pollutants converting recalcitrant organic matter into


more readily biodegradable one [61-63]. This can be achieved using O3 alone or in

combination with other AOPs such as ultrasound (US), UV and EO [64, 65].

2.5.1 Slaughethouse wastewater treatment

To the best of the authors’ knowledge, there are only few reports in the

literature on slaughterhouse wastewater treatment with O3 [17, 66-69]. For the

treatment of real red-meat processing wastewater, Wu and Doan [17] used a

screening system to remove particles larger than 1 mm as the only pre-

treatment before ozonation, reporting a 99% inactivation of total coliforms

(103 CFU/mL initial), aerobic bacteria (104 CFU/mL initial) and E. coli

(105 CFU/mL) after 8 min of ozonation with an applied O3 dose of

23.09 mg/min L (0.63 Lgas/min). They also reported a reduction in COD by only

10.7% (1900-2200 mg O2/L initial). O3 did not manage to degrade organic matter

and disinfect water efficiently, not improving light transmissibility nor TSS

reduction. The poor results obtained are probably due to the high initial COD

and BOD content, having not used any biological treatment prior to ozonation.

Millamena [66] relied on coagulation and filtration processes as a pre-treatment

method reporting a COD reduction of 25% (334 mg O2/L after pre-treatment)

after applying O3 at 0.11 g O3/h for 60 min to the pre-treated samples. Compared

to a previous study, where a similar injected O3 dosage was applied (116 mg O3

[17] vs 108 mg O3 [66]), a higher COD removal was obtained when initial COD

was lower, leading also to a lower cocentration in the final COD value. The

highest reduction in COD was reported by Proesmans et al. [68], where they

combined a biological-ozonation system for slaughterhouse wastewater

treatment, achieving a 66% COD reduction (30 mg O2/L final COD) with the

ozonation step. This indicates the importance of a secondary treatment to lower

the COD in order to achieve a higher ozonation efficiency. O3 also appears to

work efficiently prior to biological treatments, increasing, among other things,


36
the biodegradability of the effluent [66, 68]. Hodúr et al. [69] investigated the

effects of an ozonation pre-treatment in a subsequent ultrafiltration process for

the treatment of a meat industry wastewater and found 10 min O3 pre-treatment

reduced fouling and increased water flux through the membrane. COD was

also reduced roughly from 500 mg O2/L in the feed to below 150 mg O2/L in the

permeate after 5 min O3 pre-treatment. Due to a large O3 demand of raw or

primary effluents, O3 has been mainly used as a secondary or tertiary treatment

method [70].

Ultrasound

US is an acoustic (mechanical) wave whose frequency is above the upper

audible limit of an average person, usually 20 kHz. Figure 2.2 describes the

cavitation events that occur when a liquid is irradiated with US. Pre-existing

bubble nuclei act as a source for cavitation. When the US pressure is above the

threshold for cavitation, these bubble nuclei grow and coalescence, and once they

reach a resonance size, the bubbles undergo violent inertial collapse [71-73].

During the inertial collapse, the bubble core can reach temperatures of 10,000 K

and pressures of up to 1000 atm [71]. Due to the high temperatures reached upon

bubble collapse, water vapour inside the bubble dissociates to form reactive radical

species such as •OH (sonochemistry), as well as the emission of light

(sonoluminescence) [71, 72]. Along with chemical effects, mechanical effects can

also occur via formation of localised microjets (with velocities of up to 120 m/s),

formed when bubbles collapse asymmetrically near a surface. This can generate

extreme shear forces that can contribute to water treatment by tearing apart

microorganisms and disinfect water [73].


Figure 2.2. Possible processes of a bubble in an acoustic field. Adapted from [72].

For a given frequency, increasing acoustic power increases the number of active

bubbles (bubbles capable of producing sonochemistry and sonoluminescence

upon collapse) and size of individual bubbles. At higher acoustic powers bubbles

are subjected to greater negative and positive pressures during expansion

(rarefaction) and compression cycles leading to the formation of bigger bubbles

(greater net mass transfer into the bubble). At its maxium size is when a bubble

carries its maxiumum potential energy, converting it during bubble collapse into

heat, emission of light and sound, as well as the formation of radicals [72, 74].

Increasing applied frequency increases the number of cavitation bubbles produced

although higher powers are needed to form active bubbles. At higher frequencies

(shorter acoustic cycles), however, smaller bubbles are formed and undergo lower

compression [74]. That is, increasing frequency the reduction in the bubbles

expansion phase leads to less vopour entering the bubble and thus, resulting in the

formation of smaller bubbles [75]. The effect of frequency in bubble radius and

population is shown in Figure 2.3 [76]. Sonochemical effects peak at medium-high

frequencies (200-500 kHz) [77-80] due to a balance between bubble collapse

38
intensity (larger at lower frequencies) and the number of active bubbles (more at

higher frequencies) [71, 72].

Figure 2.3. Sonochemical bubble-size distribution for 213, 355, 647, 875, 1056, and

1136 kHz [76].

Two types of piezoelectric ceramic transducers are usually used for water

treatment via acoustic cavitation: (i) horn type, operated usually at low frequencies

(20 to 100 kHz); and (ii) plate transducers, with a wider frequency range (20 kHz

to 3 MHz) and a larger emitting surface area. The former provides localised high

energy density coming from the tip of the probe, while the latter, plate transducers,

produce soundwaves that can propagate through the entire bulk of the solution

[75, 81].

Sonochemical effects can be quantified by measuring the concentration of free

radicals (mainly •OH) via potassium iodide (KI) dosimetry [82-85], where I- ions

in a KI aqueous solution are oxidised by those radicals following Eqs. (14-15) to


form I3- ions. Accounting for the concentration of I3- (colorimetric method) gives an

indication of the production rate of free radicals in the solution, and thus, the

chemical effects produced during the process. Additionally, H2O2 can be formed

by a combination of •OH (Eq. 16). The absorbance of the I3- ion can be converted

into concentration following the Beer Lambert law [86], as shown in Appendix A.

2•OH + 2I−→ 2OH− + I2 (14)

I2 + I−→ I3− (15)

•OH + •OH → H2O2 (16)

In such a way, it has been reported [85, 87, 88] that the concentration of I3- and

H2O2 has a linear relationship with the concentration of •OH and is the method

commonly used to quantify the formation of •OH in a solution. The highest I 3-

formation is found to be dependent on the applied frequency, peaking at around

300 kHz [89].

On the other hand, the electric power supplied to the ultrasonic emitter (e.g.

horn or plate transducer) may differ from the power provided to the liquid sample

inside the reactor, as some of it may be wasted as heat (energy conversion of US is

transducer-dependent). In order to quantify the real power delivered to the system,

researchers account for the so called calorimetric power (energy disipated in the

solution) [88]. To account for that, electric power is supplied to the US emitting

device, mesuring a temperature change in the solution (dT/dt), while considering

its liquid mass (M) and specific heat capacity (cp, usully used that of water)

following Eq. (17).

dT
Powercal (W) = × cp × M (17)
dt

40
The effectiveness and therefore, the use of US in water and wastewater

treatment varies depending on the frequency and type of sonicator, nature of the

water to be treated, as well as the target of the treatment method. Low frequencies

(< 100 kHz) are mainly applied for water disinfection (using principally “horn”

type reactors due to higher physical effects produced ), while medium (⁓300 kHz)

and high frequencies (⁓800 kHz) are used for the removal of different compounds

(higher •OH production) [90].

2.6.1 Synthetic waters

Phenol is a contaminant easily found in the environment and chemical

industry wastewaters [79, 80]. Numerous reports [77-80] in the literature have

shown that there exists an optimal US frequency for an effective degradation

of phenol. Pétrier and Francony [77] reported that the highest phenol

decomposition rate is achieved with 200 kHz compared to frequencies of 20,

500 and 800 kHz. The highest formation of H2O2 is also observed with 200 kHz,

indicating a linkage between H2O2 formation and phenol degradation. Others

have also reported a similar optimum frequency between 200 and 500 kHz for

phenol degradation [77-80].

According to Lifka et al. [91], the correlation between H2O2 formation and

degradation is true for low-volatile compounds such as phenol and

4 - chlorophenol. Low-volatile compounds would not evaporate into the

bubble and would be oxidised by •OH in the liquid layer adjacent to the

bubble. Therefore, for low-volatile compounds a higher degradation is

expected when the •OH production rates are high (i.e. frequencies of

200 - 500 kHz).


As observed with phenol (low volatile and hydrophobic), the

decomposition of 1,4-dioxane (low volatile and hydrophilic) is also directly

linked to the formation of H2O2, showing the highest decomposition rate at

358 kHz [92].

On the contrary, for a high-volatile compound such as carbon tetrachloride

(CCl4), the decomposition rate decreased with decreasing frequency [77], with

the highest decomposition rate recorded at 800 kHz that was the highest

frequency studied. According to the authors, CCl4 degradation occurs within

the cavitation bubble (mainly through pyrolytic reactions) and therefore,

increasing frequency reduces bubble size providing a higher surface area to

volume ratio. This enhances the evaporation of the volatile compound into the

bubble leading to a higher CCl4 decomposition. Similarly, Pétrier et al. [77] and

Lifka et al. [91] reported that the degradation rate of high-volatile compounds

would increase with increasing frequency, up to around 800 kHz. Above that

frequency, the degradation speed would decrease [91].

At the same time, carbon removal increases with increasing ultrasonic

irradiation time and power. This is true independent of the type of organic

compound/wastewater sonicated (phenol [77, 93], chlorophenol [90],

diclofenac [94, 95], landfill leachate [96], methyl orange [97], azobenzene [97]).

However, saturation effects (no further increase in degradation rate above a

certain input power) have been observed when exceeding a certain power

density [91, 98].

42
2.6.2 Real wastewaters

Although no published reports were found on the use of US alone for real

slaughterhouse wastewater treatment, US has been used in the direct treatment of

different real wastewaters such as landfill leachate [96], palm oil mill effluent [99],

hospital wastewater [100] and municipal wastewater [101, 102]. US has also been

used in combination with secondary and tertiary processes to treat wastewaters

[102, 103].

A study by Abdurahman et al. [104] reported a 96.5% COD reduction

(8000 - 25400 mg O2/L initial COD) using US (25 kHz) assisted-membrane

anaerobic system for the treatment of real slaughterhouse wastewater. In addition,

the destructive effect of US have shown to lower TSS in municipal wastewaters

[105]. Wang et al. [96] used a 20 kHz ultrasonic horn at an applied power of

50 - 150 W for the treatment of 1 L landfill leachate (4770 mg O2/L, 350 mg O2/L and

680 mg/L for initial COD, BOD5 and NH3-N, respectively, at pH 8.1). Increasing

applied power increased both COD and ammonia nitrogen removal, reaching 14%

and above 80% for COD and ammonia, respectively, after 180 min treatment at

150 W. The authors highlighted that COD was primarily removed by •OH

(thermal rections such as pyrolysis had small contribution), while NH3-N removal

happend principally by pyrolysis. Manickam et al. [99] obtained a slight reduction

in COD of palm oil mill effluent using a 37 kHz ultrasonic bath alone, increasing

significantly COD removal by the addition of H2O2. On the other hand,

Naffrechoux et al. [101] obtained low COD removal of a municipal wastewater

with a combined US (500 kHz and 100 W) and UV reactor, slightly lowering COD

from 285 to 280 mg O2/L after 90 min and to 133 mg O2/L after 240 min (non-filtered

samples). It is important to underline that a lower COD reduction was achieved

with filtered (0.45 µm) wastewater after 240 min treatment (from 120 to 90 mg

O2/L). According to the authors, “suspended solids probably play an important part in
the cavitation enhancement and then in radical production”. The research group led by

Torres-Palma showed the effectiveness of ultrasonic treatment in the removal of

contaminants of emerging concern from hospital [100] and municipal effluents

[102].

2.6.3 Disinfection

Disinfection efficiency also varies with applied ultrasonic frequency. Most

of the studies found in the literature analyse the effect of low frequency US

(< 100 kHz) on microbial inactivation.

Hua and Thompson [106] analysed US disinfection efficiency of E. coli

using 5 different frequencies: 20 kHz using a horn type sonicator and 205, 358,

618 and 1071 kHz using a plate transducer. The best performance was achieved

with a frequency of 205 kHz with 4 log reduction in E. coli. Increasing the

frequency to 1071 kHz resulted in a decrease in the disinfection efficiency.

Thus, for E. coli inactivation with increasing ultrasonic frequency seems to

follow a similar trend to that observed for H2O2 and free radical formation,

attributing principally E.coli removal to the formation and attack of free

radicals such as •OH [106]. This is further supported by the addition of •OH

scavengers which decreased significantly E. coli and S. Mutans removal rate,

linking microbial disinfection to chemical effects [107]. The chemical attack,

however, would be enhanced by the physical disruption of the microbial cells

when cavitation bubbles collapse [106-108]. On the contrary, Joyce et al. [109]

obtained a higher inactivation performance of E. coli with 40 kHz (80.9 % dead

cells by flow cytometry) compared to 580 kHz (5.3 % dead cells by flow

cytometry) after 15 min irradiation at a similar acoustic intensity. It was

explained that at low frequencies, the cavitation collapse intensity is higher

caused by the formation of larger bubbles. At higher frequencies, smaller

44
bubbles are formed due to a shorter acoustic cycle and thus, leading to a lower

collapse intensity. That is, larger mechanical effects are produced at low

frequencies [109]. Thermal effects have also been used to explain the observed

microbial disinfection at low frequencies [110]. It is unclear the reason behind

the discrepencies observed from varying frequencies but it is important to

highlight that the difference in the employed sonicator type makes the

comparison between frequencies difficult. What is common in all the

aforementioned studies [106-108] is that increasing irradiation time increased

E. coli inactivation.

The disinfection efficiency of a horn type sonicator was investigated at two

frequencies, 24 and 80 kHz, reporting almost no difference on E.coli

inactivation between the two frequencies under study [111]. The disinfection

efficiency and rate constants increased with increasing applied ultrasonic

power, which is in agreement with other studies [107, 110, 112]. The influence

of high frequency irradiation (1 and 3 MHz) has also been reported, measuring

no significant difference in E. coli and Saccharomyces cerevisiae removal

percentages [113].

The disinfection efficiency is strongly influenced by solution temperature

[112, 114]. US irradiation can lead to increase in the solution temperature and

result in increased microbial inactivation (above 50°C for E. coli O157:H7) [112].

This would therefore make it difficult to determine if the effect is due to

temperature or cavitation induced chemical and physical effects.

The effect of solids in bacterial disinfection performance has also been

reported in the literature for US treatment. Suspended solids can act as

nucleation sites, increasing the number of cavitation events. This leads to a

higher cavitational activity and therefore, increasing disinfection efficiency [73,


101]. However, several reports have varied the TSS between 17.5 to 590 mg/L

and found sonication to have insignificant effect on the disinfection efficiency

[112, 115, 116]. Thus, it is currently unclear from the literature the exact role

suspended solids play in microbial disinfection.

The type and initial amount of microorganisms affects US disinfection

efficiency [117, 118]. The studies reported a decrease in cell inactivation rate

when increasing initial concentration of bacterial sample for the same

irradiation time (using synthetic samples with no suspended solids) [117, 118].

Both studies used a low frequency horn type sonicator (27.5 kHz [117] and

30 kHz [118]) to perform the experiments.

On the other hand, different bacterial reduction percentages were obtained

depending on the type of microorganism treated [115, 116, 118-120]. Neis and

Blume [119] obtained an inactivation of 2.9 log units for E. coli (gram-negative)

compared to a much lower reduction of 0.9 log units for Streptococci (gram-

positive) at 400 W/L and 60 min irradiation (20 kHz horn). The differences in

bacterial reduction would be due to cell wall structure of the treated bacteria:

gram-positive bacteria have a thicker cell wall than gram-negative bacteria and

therefore, gram-positive bacteria can withstand more the physical and

chemical effects produced by cavitation bubbles [115, 116, 118-120].

Photolysis

Ultraviolet light (UV) is an electromagnetic wave (irradiation) with a

wavelength range between 400 nm (shorter than visible light) and 10 nm (greater

than x-rays). UV light can be used for pollutant degradation and microbial

disinfection through direct and indirect photolysis. In direct photolysis, the

46
irradiated UV light must be absorbed by the target contaminant (photo-sensitive

compound), leading it to an excited state. The excited molecule may react with

other compounds and eventually be degraded. When O3 or H2O2 are present in

water, indirect photolysis may occur. •OH are produced by UV irradiation (no

need for the target compound to absorb the incident UV light) through the

decomposition of O3 or H2O2 (two •OH formed with the cleavage of a single H2O2

molecule via UV photolysis, as shown by Eq. 18), leading to compound

degradation or microbial disinfection via •OH attack [121, 122].

H2 O2 + hv → 2 ● OH (18)

In direct photolysis application for water treatment, C type UV light (UVC,

wavelength range between 200-280 nm) is of interest, where both the target

pollutants and the water constituents absorb the radiation. When an external

oxidant such as H2O2 is present in water, low-pressure mercury lamps are usually

used (peak emission at 254 nm). The maximum absorbance of H2O2 is at around

220 nm and due to its low-absorption or molar coefficient (measures how strongly

a chemical species absorbs light at a particular wavelength), a high concentration

of H2O2 is required to generate •OH in sufficient quantities. At high

concentrations, however, H2O2 acts as •OH scavenger and reduces the efficiency

of the process [122].

Real wastewater effluents from biological processes are rarely treated by UV

alone, as treatment efficiency is significantly reduced for coloured waters with a

considerable concentration on TSS. In this case, light transmissibility and hence,

UV efficiency would be hindered due to inner filtering effects [123, 124]. For

microbial disinfection, photolytic processes are well established [125-127], where

waters with light colour and a low amount of suspended solids are usually
encountered. UV light is also used when targeting specific pollutants (e.g.

pharmaceuticals) in synthetic [128, 129] and real waters [130, 131].

Electrochemical oxidation

Electrochemical oxidation consists on the use of two electrodes (electrochemical

cell), operating as anode (positive charge) and cathode (negative charge), and

connected to a power source. When energy is supplied and providing a supporting

electrolyte (such as water or wastewater), strong oxidising species are formed that

degrade organic matter and remove compounds. The main operating parameters

in these processes are the electrode material and the current density. In wastewater

treatment, one of the most commonly used electrodes are those with diamond

coatings (non-active anodes) since they favour the production of large amounts of

free •OH from water oxidation over the anode surface (Eq. 19) [48, 132]. Non-active

anodes are electrodes that do not participate in anodic reactions. They only serve

as an inert substrate, where the principal reaction is water discharge. •OH

produced by water discharge on non-active anodes are associated with the

oxidation of organic compounds in aqueous medium [49]. These radicals present

a high oxidation potential, favouring the complete mineralisation of the organic

matter present in the effluents [47]. For this reason, this technology has been tested

in the treatment of different industrial and urban wastewaters polluted with

different organic pollutants such as pharmaceuticals, pesticides, hormones, dyes,

etc. [133-135].

H2O → •OH + H+ + e- (19)

Furthermore, the use of diamond anodes favours the production of powerful

oxidants from the oxidation of ions naturally contained in wastewater. These

48
species can also contribute to the degradation of organic matter present in the

effluents, increasing the process efficiency. Therefore, different mechanisms can

simultaneously occur during electrolysis with diamond anodes for the removal of

organic matter (Figure 2.4): (i) direct electrolysis over anode surface and (ii)

mediated electrolysis by hydroxyl radicals and/or electrogenerated inorganic

oxidants [48].
(a) (b)
Interface Interface
Anode (+) Anode (+) H 2O
R R
R* R
•OH
e- e-
RO RO
RO* RO

(+) Electrolyte (+) Electrolyte

Figure 2.4. Mechanisms of: (a) direc electrolysis and (b) indirect/mediated electrolysis.

Powerful oxidants such as hypochlorite (ClO-), hypoclorous acid (HClO) or

peroxodisulfate (S2O82-) can be electrochemically generated following Eqs. (20-23)

when chloride (Cl-) or sulfate (SO42-) ions are present in wastewater [133].

2Cl− → Cl2 + 2e− (20)

Cl2 + H2O → HClO + Cl− + H+ (21)

HClO ↔ClO− + H+ (22)

2SO42- → S2O82- + 2e− (23)

According to Sánchez et al. [136], the efficiency obtained by in-situ

electrogenerated oxidants would be increased due to some sort of activation (e.g.

by UV irradiation) of the oxidants produced. Similarly, some oxidants can also be

formed by the reaction of ions with •OH when employing BDD anodes [48, 133],

as shown by Eqs. (24-25).


Cl− + •OH → ClO− + H+ + e− (24)

2SO42- + •OH → S2O82- + •OH- + e− (25)

Combination of advanced oxidation processes

2.9.1 Ozone and ultrasound

It has been reported that combining US with O3 leads to an important increase

in total organic carbon (TOC) removal efficiency, as well as decreasing the time

needed to achieve a specific removal percentage [80]. According to the authors, the

synergistic effect of the combined system (US/O3) comes from the rapid generation

of free radical species after O3 decompostion and the attack of •OH on relatively

refractory reaction by-products. That is, a higher TOC removal is achieved with

US/O3 due to an increase in mass transfer of O3 into the solution and an increase in

the decomposition process (higher production of •OH), as well as the generation

of radicals by US [137].

The use of O3 and US, either combined or in subsequent processes, has been

reported for real wastewater treatment [138, 139] and disinfection [140-143],

although the use of a combined system for the treatment of real slaughterhouse

wastewater is still missing in the literature.

A combination of hydrodynamic cavitation (HC) and O3 (9.41 g O3/h at a

flowrate of 15 L/min) was employed by Boczkaj et al. [138] for the treatment of

wastewater from bitumen production (initial COD of 8000-12000 mg O2/L). The

temperature of the solution was maintained constant at 40±2˚C. After 1 h of

treatment, 30% reduction in COD was measured with the combined system (8%

for HC alone and O3 alone). Increasing the treatment time to 2 h, 35% COD removal

50
was achieved by US/O3, while roughly 10% removal was obtained with HC and O3

alone, showing a significant improvement with the combined system. For BOD,

similar values were measured with a 45% reduction after 3 h of US/O3 vs 13% and

19% with HC and O3 alone, respectively. The authors stated that •OH attack was

the main degradation mechanisms of the organic pollutants. The combination of

the two processes increased the performance of the system due to the formation of

radicals formed through O3 decomposition under HC.

Ibáñez et al. [139] investigated the treatment of a real wastewater from a

sewage at pilot-plant scale. A combination of US and O3 was used in a 2.3 m3 steel

pipe reactor, targeting principally emerging contaminants. COD was not reduced

during the treatment, probably due to a short contact time with O3 and a low COD

value that ranged between 30-40 mg O2/L. The authors mentioned that O3 alone

proved to be efficient in the removal of most of the emerging contaminants. The

addition of US did not improve the efficiency, being its use “practically

unnecessary”.

The coupled system has also been used for real wastewater disinfection.

Burleson et al. [140] used a 40 kHz transducer (150 W) coupled to an O3 generator

(injection flowrate 152 cm3/h) to treat 1.7 L of synthetic and real (20 mg O2/L BOD

and 45 mg O2/L COD) effluents coming from a wastewater treatment plant. For the

sythetic wastewater, no difference was found in microbial inactivation using O3

alone and US/O3 as O3 alone was enough to inactivate bacteria. While for real

wastewater, a higher efficiency was achived with US/O3 compared to O3 alone (up

to 4 log difference). The observed differences between synthetic and real

wastewater suggest that the complexity of the real wastewater plays a role in the

final performance of the combined system and highlights the importance to work

with real effluents.


Dahi [141] employed redistilled water (synthetic), a secondary effluent from a

biological sewage treatment plant and a 5 times diluted secondary effluent and

treated with a combined US (20 kHz horn at 160 W) and O3 (1-4 mg O3/L dose)

system. It was reported that US enhanced the action of O3 (above all for the

treatment of secondary effluent) by increasing mass transfer rate and O3

decomposition (increase in •OH production), leading to a reduction in O3

sterilisation dose. US also enhanced the effect of O3 in chemical oxidation, being

•OH attack the principle mechanism for bacterial inactivation.

Another study [142] utilised plate transducers at frequencies of 20, 28 and

40 kHz (0-99 W/L) combined with O3 (0.24 mg O3/L) for the treatment of a

secondary effluent from a wastewater treatment plant. The combined system was

significantly more effective than using O3 alone (2 log difference in bacterial

inactivation). Combined with O3, similar microbial inactivation values were

obtained with the three frequencies. An increase in bacterial log reduction was

measured with increasing applied US power up to 55 W/L. Further inceasing

applied power did not increase the efficiency of the system. The authors

argumented that at the highest US energy density, a large volume of cavitation

bubbles was obtained and this resulted in an insufficient collapse of bubbles. This

effect led to a dissipation of energy with no apparent consequence in bacterial

reduction. From the reported literature [140-142] it can be highlighted the

enhanced action of O3 obtained by US application for the treatment of real

wastewaters. In the combined US/O3, O3 concentration in the effluent was

increased by the ultrasound induced enhancement in the interphase transport of

O3. Besides increasing O3 transfer into the solution, US could have broken up

particulate organic matter and bacteria clusters, exposing them to the oxidising

power of O3 [140]. The degradation of a more complex water (real wastewater)

52
would be improved with the coupled system by the increase in O3 dissolution and

a subsequent increase in •OH production via O3 decomposition, as real

wastewaters tend to carry more suspended solids and refractory organic

compounds that makes more difficult its degradation [144]. US also reduced O3

demand of the secondary effluent. Sonication alone, on the contrary, did not

inactivate bacteria neither in synthetic nor in real wastewaters.

Jyoti and Pandit [143] studied the effect of sonication (22 kHz horn at 240 W

and 100 mL samples for 15 min and 20.4 kHz bath at 120 W and 2 L samples for

15 min) and HC after ozonation of filtered bore well water. A stock solution of O3

with concentration of 50 mg O3/L was prepared to achieve 0.5, 1, 2, 3 and 4 mg O3/L

in 1 L samples prior to sonication. Variation in bacterial population did not have a

significant effect on the experimental results. The authors reported that the higher

the O3 concentration in water, the lower the time to reach a certain inactivation

level and the higher the disinfection percentage. They also measured a faster O3

decomposition in water in the presence of US (after 15 min, 55% O3 decomposition

without US and 75 and 80% with US horn and bath, respectively). Synergistic

effects were however only observed up to 2 mg O3/L. The study also highlighted

that an increase in bacterial cell wall permeability was obtained with cavitation

activity, increasing disinfection efficiency. When O3 was added to water samples

before subjecting it to sonication, the overall disinfection rate obtained was higher

than that obtained when US was applied (horn or the bath). This could be

explained as follows:

• Disruption of microbial flocs due to US [141, 145].

• Increase in microbial cell wall permeability due to sonication [141] that

enhances O3 attack.
• Faster penetration of O3 into the microorganisms due to an ultrasonic

acceleration of diffusion [146].

• Disinfection due to free radical attack after O3 decomposition increased by

the application of US [141].

2.9.2 UV and hydrogen peroxide

The combination of UV and H2O2 leads to the production of •OH with a yield

of two radicals formed per photon absorbed by 254 nm (Eq. 18). Despite having a

low molar extinction coefficient, the quantum yield (defined as the number of

defined events per photon absorbed by the system) of H2O2 is high, leading to a

high production of •OH when sufficient H2O2 is present in solution [122, 147].

Literature has shown the importance of H2O2 concentration in a combined

UVC/H2O2 process for TOC removal in slaughterhouse wastewater [22, 148]. The

optimum H2O2 concentration depends on water quality parameters such as initial

TOC [22, 148] and concentration of solids [149]. Using synthetic slaughterhouse

wastewater, Barrera et al. [150] reported a significant reduction in TOC with UVC

alone (48%) compared to the 51% TOC reduction when 1000 mg H2O2/L was added

to the UVC photolytic process and 14% TOC removal with 1000 mg H2O2/L alone.

They attributed the main mechanism of TOC removal to direct photolysis and that

the slight reduction in the TOC for the combined UVC/H2O2 was due to the

decrease in the H2O2 concentration during the process, resulting in a lower •OH

production. Therefore, the research article highlighted the importance of keeping

a constant concentration of H2O2 throughout the entire UVC/H2O2 process to

increase system’s performance. At the same time, they showed by CFD simulations

the relevance of a good mixing in order to obtain a good reaction rate for TOC

degradation. The authors also mentioned the importance of further studies on the

54
treatment of secondary effluent of slaughterhouse wastewater to examine the

applicability of AOPs (specifically UVC/H2O2) in real situations.

For real slaughterhouse wastewater, Bustillo-Lecompte et al. [151] reported a

total 91% TOC reduction during the anaerobic baffled reactor, activated sludge and

UV/H2O2 combined process, with a 75% reduction in TOC during the UV/H2O2

process alone. Similarly, Luiz et al. [152] treated real slaughterhouse wastewater

after coagulation and observed a higher COD reduction of 48% with the combined

UVC/H2O2 system compared to 31% COD reduction with UVC alone, showing

possible differences in the efficiency of the UVC/H2O2 process between real [152]

and synthetic [150] slaughterhouse wastewaters.

2.9.3 Ozone and UV/hydrogen peroxide

Although the combined UVC/H2O2 has shown to be effective for treating real

slaughterhouse wastewater, the efficiency of photolytic processes would be

hindered by the high turbid and colour of the wastewater. This is due to the poor

UVC light transmissibility and inner filtering effects [123, 124]. O3 is an effective

colour removing agent [153], and if used as pre-treatment or coupled to UVC/H2O2,

could greatly improve the treatment efficiency. However, O 3 is an expensive

oxidant [154] and optimisation of ozonation pre-treatment time is important. To

the best of the authors’ knowledge, there are no reports on the literature using O3

as a pre-treatment and combined to a UVC/H2O2 process for the treatment of real

slaughterhouse wastewater.

As it happens when combining UV and H2O2, the combination of O3 with H2O2

also leads to the formation of •OH, producing two •OH from two O3 molecules

(Eq. 26) [54].


2O3 + H2O2 → 2•OH + 3O2 (26)

When O3 and UV radiation are coupled, the oxidation mechanism is initiated

by the photolysis of O3. This then leads to the production of H2O2, as shown by

Eq. 27 [53, 54].

O3 + H2O + hv → H2O2 + O2 (27)

When H2O2 is added to a O3/UV processes, O3 decomposition is accelerated

and •OH generation rate increases [155].

Taking the above into account, different oxidation mechanisms should be

considered when combining O3, UVC and H2O2: direct photolysis of the target

pollutant (considering it absorbs light at the applied wavelength), direct attack of

molecular O3, as well as the production of •OH and H2O2 through ozonation, and

the production of •OH through the combination of UV and H2O2, among others

[53, 54].

To the best of the authors’ knowledge, no complete studies were found on the

literature where O3, UVC and H2O2 were coupled for the treatment of

slaughterhouse wastewater. There are, however, several research articles

combining the three technologies used for real wastewater treatment such as

potato chips manufacturing plant effluent [156], hospital wastewater [157] and

winery wastewater [158].

Arslan et al. [156] employed two UVC lamps of 2.2 W each (285 nm) combined

with a varying O3 injection of 0-20 mg/L and a H2O2 concentration of 0-1500 mg/L

in batch mode (2 L). Potato chips manufacturing effluent samples (504 mg O2/L

56
COD, 100 mg/L dissolved organic carbon (DOC), 145 mg/L TSS and 1691 Pt-Co at

pH 7.7) were collected after an anaerobic reactor and treated with the combined

setup for 90 min. In the absence of H2O2, high concentrations of O3 were necessary

to obtain organic matter removal. When O3 and H2O2 concetrations ranged

between 8-17 mg/L and 400 - 1100 mg/L, respectively, 80% TOC removal was

achieved. Best pH was found between 5.5 and 8.2 (above 80% TOC reduction) and

at low O3 concentrations, H2O2 impact was found to be very weak. Under acidic

condition, no TOC removal could be seen, indicating •OH oxidation as the most

probable degradation mechanism. Arslan et al. [157] also used a similar combined

O3/UVC/H2O2 for the pre-treatment of hospital wastewater. In this case, the

interaction between O3 concentration and H2O2 dosage was found to be the most

important factor affecting the process performance.

Lucas et al. [158] utilised a pilot scale reactor (9 L) for the treatment of winery

wastewater (4600 mg O2/L COD and 1200 mg/L TOC at pH 4) using different

combination of O3, UV and H2O2. After 3 h of treatment, COD degradation by UV

alone was insignificant, O3 alone obtained 12% reduction, O3/UV combined

resulted in 21% COD removal, and O3/UVC/H2O2 (COD/H2O2 = 4) led to 35% COD

reduction. According to the authors, “the improvement happens due to the photolysis

of ozone, the enhanced mass transfer of O3 and the generation of OH radicals”. Increasing

pH to 10, higher efficiencies were achieved by O3/UV and O3/UVC/H2O2 processes,

probably due to a higher production of •OH. However, the natural pH of 4 of the

wastewater was mantained for further experiments in order to minimise expensive

pH adjustments in an industrial process. Increasing H2O2 concentration

(COD/H2O2 = 2) provided sufficient H2O2 in solution to optimise the O3/UVC/H2O2

process and obtained a TOC removal of 88%. Further increasing H2O2

concentration (COD/H2O2 = 1) led to •OH scavenging effects, reducing the final

performance of the system. In this case, the removal of TOC followed the trend of

COD removal.
2.9.4 Electro-advanced oxidation processes

The performance of an electrochemical cell can be improved by coupling it with

other advanced oxidation processes such as O3, photolysis or UVC/H2O2. These

combined technologies can produce synergistic effects for the removal of organic

matter [50, 51, 159] and decrease the energy and operational costs. For instance, UV

light can be used as an external source for the generation of free radicals (Figure

2.5) through the photoactivation of electrochemically generated species (Eqs. 28-

30) [48, 160].

ClO- + hv→ •O- + •Cl (28)

O- + H2O + hv → OH- + •OH


• (29)

S2O82- + hv →2(SO4−)• (30)

Figure 2.5. Possible mechanisms of electrochemical and photo-electrochemical

treatment of secondary wastewater [48].

Literature showed the potential electrocoagulation offers as a secondary

treatment for slaughterhouse wastewater [161, 162], having found just few articles

58
on the use of electrochemical oxidation and EAOPs as post-treatment method for

the treatment of slaughterhouse wastewater. Awang et al. [163] obtained 85% COD

removal (220 mg O2/L initial COD) with 55 min of electro-oxidation treatment at

30 mA/cm2 using aluminium electrodes. Davarnejad and Nasiri [164] achieved 92%

COD reduction after 55 min of electro-Fenton treatment using iron electrodes. In

turn, Vidal et al. [165] reported up to 88% COD removal by a solar photoelectro-

Fenton process after an anaerobic digestion of synthetic slaughterhouse

wastewater.

Outline of the thesis and objectives

It is clear from the literature that several research articles have been published

on the use of an ASP for the treatment of real and synthetic slaughterhouse

wastewasters, showing high efficiency and realiability in organic matter and

nutrient removal. However, the effluent is still above the direct discharge limits set

by the EU. The use of single or combined AOPs such as O3, US, UVC/H2O2 and EO

has potential to reach these discharge limits and meet water reuse standards if

applied after a biological treatment, but complete and detailed studies on these

applications for organic matter removal and microbial disinfection of real

slaughterhouse wastewater are still lacking. These includes effect of treatment

times, injection dosages (O3 and H2O2) and methods of injection, applied

ultrasound powers, reactor type, influence of wastewater matrix (organic load, TSS

content, colour and turbidity, wastewater origin), organic matter degradation

mechanisms, etc.

In order to reach direct discharge limits set by the EU and evaluate the

efficiency and realiability of the employed treatment methods, different water

quality indicators such as DOC, COD, BOD, TN, TP, TSS, colour, turbidity, as well
as electrochemically generated ions and free oxidants need to be monitored. The

extent of microbial disinfection needs to be tracked with the measurement of total

coliforms (TC) and total viable counts (TVC) to assess the potential in microbial

inactivation. These parameters will be measured under different applications of

single and combined AOPs by investigating the effect of injected chemical

concentration (O3 and H2O2), treatment time and applied ultrasound frequency,

power and reactor type, with the objective of reducing water parameters down to

minimums and assess possible synergistic effects. The thesis is therefore organised

as follows:

Chapter 3 provides the materials employed and the methods followed during

the experimental work carried out. Employed reactors and equipment are defined

and treatment procedures explained in terms of reactor volume, residenence time,

concentration of injected chemicals, applied powers, etc. Characteristics of the

chemicals are also mentioned, as well as applied analytical techniques.

Chapter 4 investigates the use of O3 as a tertiary treatment after a combined

activated sludge-filtration process. Two HRT (12 and 24 hours) and a 13-day SRT

were used to evaluate the performance of the ASP, followed by filtration. The

purpose of the filtration system was to separate any solids/sludge coming from the

ASP, as well as to show the possible extent of the application of a separation

process after the ASP. O3 was applied at 71 mg O3/Lgas for different periods of time

to investigate the organic matter removal and the disinfection efficiency, assessed

by measuring TC and TVC.

Chapter 5 examines the application of acoustic cavitation with and without O3.

The effect of sonication frequency was evaluated by using three different

frequencies of 44, 300 and 1000 kHz and an applied power of 40 W. The injected O3

dose was fixed at 71 mg O3/Lgas and the treatment time varied from 1 to 60 min. A

60
20 kHz ultrasonic horn was also utilised at two different applied powers and

compared to higher frequencies.

The effect of combining O3 and UVC/H2O2 is explored in Chapter 6. Three

different ozonation pre-treatment times were utilised (5, 10 and 15 min) with the

aim of adjusting best pre-treatment time for a given O3 inlet concentration

(100 mg O3/Lgas). An UVC/H2O2 system was further applied to the O3 pre-oxidised

effluent, evaluating the performance of different H2O2 concentrations at 150, 300

and 450 mg H2O2/L during the 3 h of treatment. In order to assess any possible

synergy between the applied process, O3, UVC and H2O2 were later combined in a

single configuration, reducing injected O3 concentration to 5.6, 8.1 and 12.1 mg/Lgas

(3 h injection).

Chapter 7 includes the results on the use of EO and EO related processes, either

alone or in combination with pre-ozonation. Besides EO, the following processes

were applied: EO with the addition of H2O2, EO assisted by UVC light and EO in

the presence of UVC light and H2O2. A BDD anode (electrochemical cell) was

employed, as well as an annular channel photoreactor for enhanced illumination

and mass transfer. An O3 pre-treatment (100 mg O3/Lgas) was also applied prior to

the abovementioned EAOPs, monitoring free oxidants and inorganic anions and

cations formed during EO combined processes.

Finally, Chapter 8 presents the main conclusions of the thesis, where a

comparison of the employed AOPs is given with a highlight in the most important

parameters and process combinations studied. Possible optimisation points of the

employed setups are also considered, as well as suggestions and proposals for

future research.
Chapter 3

Materials and Methods

Introduction

This chapter provides materials, methods and the different setups employed

during the experiments carried out. The sections are divided into the different

wastewater treatment configurations employed, highlighting treatment

technologies and analytical techniques.

Chemicals

KI (> 99%, Fisher Chemical) was used as received. COD, ammonia, nitrate and

total dissolved nitrogen measurement kits were purchased from Hach. For

phosphorus measurements, Molybdovanadate (Hach product number 2076049)

was used as a reagent. H2O2 was used as an external oxidant (HYPE-30P-1K0 from

Labbox, 30% purity v/v) and sodium sulphite (Na2SO3) added in a Na2SO3-to-H2O2

molar ratio of 1:1 [166] after sample collection to quench the H2O2 and cease the

oxidation process before DOC analysis. Ammonium metavanadate (Merck) was

used as a colorimetric reagent for H2O2 determination. Sulfuric acid (H2SO4), starch,

sodium thiosulfate (Na2S2O3), hydrochloric acid (HCl) and sodium hydroxide

62
(NaOH) were supplied by Merck, Probalab, Alfa Aesar and VWR Chemicals,

respectively.

For microbial platecount, nutrient agar (peptone 5 g/L, yeast extract 3 g/L, NaCl

5 g/L, agar 15 g/L, 1 L reagent-grade water at pH 6.8±0.2) was used for TVC growth

and lauryl tryptose broth (tryptose 20 g/L, lactose 5 g/L, dipotassium hydrogen

phosphate 2.75 g/L, potassium dihydrogen phosphate 2.75 g/L, NaCl 5 g/L, sodium

lauryl sulphate 0.1 g/L, 1 L reagent-grade water at pH 6.8±0.2) for TC. Pure and dry

oxygen used to generate O3 was supplied by Linde (HiQ® Oxygen 4.5,

Purity ≥ 99.995%). Ultrapure water was obtained from a Millipore® Direct-Q

system (18.2 MΩ cm resistivity at 25 °C).

Analytical methods

3.3.1 Colour, turbidity and pH

Colour in the Pt-Co scale was determined by measuring the absorbance at

400 nm in a VWR UV-6300PC double beam spectrophotometer and computed

using a calibration curve. Before determination, samples were filtered through

0.45 µm Nylon filters from Whatman. This spectrophotometer was employed in all

ultraviolet-visible measurements. Turbidity (SM 2130 B) was evaluated using a

Nephelometer to measure the intensity of light scattered of a specific sampled and

compared to a standard reference. pH, temperature and conductivity were

measured by HI 9829 Multiparameter portable meter from Hanna Instruments.


3.3.2 KI dosimtery

For •OH production analysis, 400 mL of 0.1 M KI were sonicated and/or

ozonated following a KI dosimetry method explained in Section 2.6 and elsewhere

[82, 84]. The concentration of H2O2 was determined by the colorimetric

metavanadate method explained by Nogueira et al. [167], where 1 mL of

ammonium metavanadate (NH4VO3) was added (solution: 7 gr of NH4VO3 and

20 mL of 98% H2SO4 in a 1 L flask) to a 10 mL sample and absorbance measured at

450 nm (colour change into a red-orange).

3.3.3 Inorganics

The concentration of inorganic anions and cations was measured according to

the procedure given by Moreira et al. [168]. Inorganic anions were quantified by

ion chromatography by injecting 10 µL sample in a Dionex ICS-2100 LC equipped

with an IonPac®AS11-HC 250 mm × 4 mm column at 30°C and an anionself-

regenerating suppressor (ASRS®300, 4 mm) under isocratic elution of 30 mM

NaOH at a flow rate of 1.5 mL min-1. Inorganic cations were measured by ion

chromatography, injecting 25 µL sample in a Dionex DX-120 LC equipped with an

IonPac®CS12A 250 mm × 4 mm column at ambient temperature and a cation self-

regenerating (CSRS®Ultra II, 4 mm) suppressor under isocratic elution of 20 mM

methanesulfonic acid at aflow rate of 1.0 mL min-1. Before determination, samples

were filtered through 0.45 µm Nylon filters from Whatman.

3.3.4 Free oxidants

Free oxidants reported in Chapter 7 were determined iodometrically taking

into account the procedure reported by Kolthoff and Carr [169]. The procedure to

follow was: (i) 10 mL sample were taken from the recirculation vessel and mixed

64
with 4 mL sulphuric acid (H2SO4, 20% v/v), (ii) 0.5 g KI were added to the mixture,

(iii) 0.5 mL starch (20% v/v) were dropped into the mixture, darkening the colour

of the mixture, and (iv) 0.1 M sodium thiosulfate (Na2S2O3) were added drop by

drop to the mixture until a change in colour into transparent-white was observed.

The volume of sodium thiosulfate was annotated to compute the amount of free

oxidants [170].

3.3.5 Standard and Hach methods: organics, inorganics and microorganisms

The concentration of organic matter was measured as 5-day BOD (standard

method (SM) 5210 B) and COD (SM 5220 D). Shimadzu TOC-VCSN analyser was

used to determine total dissolved carbon (TDC) and dissolved inorganic carbon

(DIC) values. DOC was then calculated by subtracting DIC from TDC. Before

determination of TDC and DIC, samples were filtered through 0.45 µm Nylon

filters from Whatman. Phosphorus (Hach method 8114, adapted from SM 4500 P

E), total dissolved nitrogen (Hach method 10072), ammonia (Hach method 10031)

and nitrate (Hach method 8039) were measured by Hack colorimetric kits. TSS (SM

2540 D), MLSS (SM 2540 D) and MLVSS (SM 2540 E) contents were also

determined. Additionally, TC (SM 9222 B) and TVC (SM 9215 C) were analysed

before and after ozonation, sonication and ozonation-sonication experiments to

evaluate the disinfection efficiency of the process [171]. Standard and Hach

methods adopted are further explained below.

3.3.5.1 BOD (SM 5210 B)

The method consists of measuring principally the molecular oxygen used

during a 5 day incubation period for the biochemical degradation of organic

matter. For slaughterhouse wastewater, sample dilution was necessary to

provide a proper balance between the oxygen demand and supply. The method
follows the next procedure: (i) dilution of the initial wastewater sample

depending on initial BOD and DO concentration (a blank with no wastewater

is also needed), (ii) fill up an airtight bottle of specific volume (250 mL) until

overflows and incubate it for 5 days in the dark, at a constant temperature of

20 ± 1°C and at pH between 6.5 and 7.5, (iii) measure DO concentration before

and after incubation period and compute BOD in mg O2/L following Eq. (31):

(𝐷1 −𝐷2 )−(𝐵1 −𝐵2 )×𝑓


𝐵𝑂𝐷 = (31)
𝑅

Where D1, D2 are DO of diluted samples before and after the test; B1, B2 are

DO of control samples (blank) before and after the test; f is the ratio (Vd/V)

between seeded volume (Vd) and total volume (V); R is the ratio (Vww/V)

between the wastewater treatment volume (Vww) and the total volume (V).

3.3.5.2 COD (SM 5220 D)

The method consists of measuring the amount of dichromate ion (Cr 2O72-)

reduced to chromic ion (Cr3+) by a change in absorbance at 420 nm (colorimetric

method). Cr2O72 oxidises COD material in the sample during a 2 h digestion at

150°C. Organic (predominant) and inorganic compounds are accounted after a

digestion period (including suspended solids) and the amount of oxidant

consumed is given in its oxygen equivalence (mg O2/L). Hach COD Kits (COD

range 0-1500 mg O2/L) were used for COD mesurements, adding 2 mL samples

to each of the kit vials. Final COD values were calculated from a calibration

curved obtained with COD standards of 20, 400, 1000 and 2000 mg O2/L.

66
3.3.5.3 Phosphorus (Hach Method 8114 adapted from SM 4500 P E)

Molybdovanadate (Hach product number 2076049) was used as a

colorimetric reagent, adding 0.5 mL to a 10 mL sample. Molybdate reacts with

orthophosphate to produce a mixed phosphate/molybdate complex and in the

presence of vanadium, a change into an intense yellow colour is observed

(molybdovanadophosphoric acid). The absorbance was measured at 430 nm

and phosphorus (P) concentration given by the Hach spectrophotometer

method 8114.

3.3.5.4 TSS and MLSS (SM 2540 D)

A weighed standard fiber-glass filter (4-7 µm) is used to filter a sample

(100 mL) and the solids retained on the filter are dried in an oven at 103 - 105°C

for at least 2 h. The TSS and MLSS (mg/L) are represented by the increase in

weight of the filter.

3.3.5.5 MLVSS (SM 2540 E)

The residue from MLSS method is ignited in a furnance for at least 20 min

at 550°C to evaporate all volatile solids. The weight lost after ignition accounts

for the volatile solids in the residue.

3.3.5.6 Total dissolved nitrogen (Hach method 10072)

A TN persulphate reagent powder pillow was added to a TN hydroxide

digestion reagents vials (alkaline persulfate digestion converts all forms of

nitrogen to nitrate) and a 0.5 mL of sample added (a blank was also prepared

with deionised water). The vials were digested for 30 min at 105°C. Once at

room temperature, TN reagent A powder pillow was added and after 3 min, a
second TN reagent B powder pillow was poured (sodium metabisulfite is

added to eliminate halogen oxide interferences). After 2 min, 2 mL of the

digested sample was pipetted to a TN reagent C vial (nitrate reacts with

chromotropic acid under strongly acidic conditions to form a yellow complex)

and the absorbance mesaured at 410 nm (colorimetric method). Total dissolved

nitrogen concentration was obtained from Hach spectrophotometer method

10072.

3.3.5.7 Ammonia (Hach method 10031)

0.1 mL of sample was added to a reagent test tube (blank was also prepared

with deionised water). Ammonia salicylate reagent powder pillow was later

added (monochloramine is formed by the combination of ammonia compounds

and chlorine, and reacts with salicylate to form 5-aminosalicylate), as well as

the content of ammonia cyanurate regent powder pillow. After a proper

mixing, vials were left for a 20 min reaction time (5‑aminosalicylate is oxidised

in the presence of a sodium nitroprusside catalyst) and the absorbance of the

distiguished green-coloured solution measured at 655 nm. Concentration of

ammonia was given by the Hach spectrophotometer method 10031.

3.3.5.8 Nitrate (Hach method 8039)

A nitrateVer 5 reagent powder pillow was added to a 10 mL sample (blank

was also prepared with deionised water). After 1 min of reaction time, the cell

was shaken vigorously and a 5 min reaction time was allowed. Finally, the

absorbance was measured at 500 nm and the concentration given by the Hach

spectrophotometer method 8039. In this case, nitrate (NO3-) is reduced to nitrite

(NO2-) by cadmium metals and nitrite ion subsequently reacts with sulfanilic

68
acid to form an intermediate diazonium salt. The salt then couples with gentisic

acid to form an amber colored solution.

3.3.5.9 Total coliforms (SM 9222 B)

TC refers to a group of gram-negative, rod-shape bacteria, including

thermotolerant coliforms and bacteria of faecal origin mostly found in the

environment. They can be damaged by lack of nutrients, as well as current

water treatment methods (chlorine disinfection, ozonation, UV light, etc.).

Therefore, TC are considered useful indicators of bacterial and viral waterborn

pathogens, as it is impractical to analyse water samples for all known microbial

pathogens that may contaminate water. TC are used to control water quality

and the integrity of the distribution systems in waters intended for human

consumption. The absence of TC reduces the likelihood that fecal pathogens are

present [172, 173].

A 0.45 µm filter paper was used to filter 100 mL diluted wastewater samples

(up to 104 times dilutions applied to make microbial colony plate count

possible). At the same time, 2 mL of prepared medium (lauryl tryptose broth)

were poured in a sterilised absorbent pad and the pad placed in a 50 × 9 mm

sterilised petri dish. The filter paper was then placed on top of the pad and

incubated overnight in the dark at 35°C. After the incubation period, bacteria

producing red colonies (considered member of the coliform group) were

counted and colony forming units per mL sample (CFU/mL) computed.

3.3.5.10 Total viable counts (SM 9215 C)

The heterotrophic group of bacteria comprises a wide range of bacteria that

use organic carbon source to grow, including all microorganisms capable of


growing in solid agar medium. Heterotrophic bacteria colony counts (also

known as total plate count or total viable count) provide an indication of the

overall load of aerobic and facultative anaerobic bateria in water samples [174].

Wastewater samples were diluted up to 106 times and 0.1 mL of diluted

samples were spread using a sterilised glass rod in a sterilised 100 × 15 mm petri

dish where agar had been previously added and solidified. Petri dishes were

covered and incubated for 5 days at room temperature (25 ± 2 °C) in a sterilised

fume cupboard. Colonies were counted after 5 days and results given in

CFU/mL.

Real slaughterhouse wastewater

Wastewater samples employed in Chapters 4 and 5 were taken directly from a

cattle slaughterhouse located in the county of Surrey, UK. The wastewater

contained not only animal residues (blood, fat, viscera, manure, among others), but

also onsite sewage, and traces of floor cleaning products. The wastewater collected

was partially treated on site by a grit removal system, followed by coagulation-

flocculation, where ferric chloride solution was used as a coagulant and Polygold

CE662 as a flocculation agent and processed further by dissolved air flotation. This

onsite pre-treated effluent will be referred to as “raw wastewater” during results

section in Chapter 4. To account for wastewater variability, the slaughterhouse

effluent was sampled at least once per week over a two-month period and stored

at 4°C prior to use. After an onsite pre-treatment, the wastewater was treated using

a lab scale ASP. This onsite pre-treated and biologically treated effluent will be

referred to as “treated wastewater” during results section in Chapter 5.

70
The wastewater used in Chapters 6 and 7 was collected from a pig

slaughterhouse near Porto, Portugal. Before collection, the wastewater was treated

on site by the following treatment train: (i) grit removal, (ii) degreasing, (iii)

biological process in an activated sludge biological reactor, and (iv) settling.

A combined activated sludge-filtration-ozonation process

3.5.1 Experimental setup

The activated sludge-filtration-ozonation (AFO) system used is shown in

Figure 3.1. O3 was used after filtered and non-filtered effluent from the ASP.

Figure 3.1. A schematic of the experimental setup.

3.5.1.1 Activated sludge reactor

Slaughterhouse wastewater samples were fed at a rate of 1 L/day into an

activated sludge reactor (6 L glass reactor) in a semi-batch mode and with a SRT
of 13 days. The aeration (5 L/min) in the ASP was ceased for 30 min in order to

allow the sludge to settle before removing the bio-treated effluent (from the top

of the reactor) and sludge (settled sludge).

3.5.1.2 Filtration

Once the ASP reached steady state, the effluent was filtered through a filter

paper of pore size ranging between 4 and 7 µm (Whatman cellulose filters,

grade 595). The purpose of the filtration system was to separate any

solids/sludge coming from the ASP, as well as to show the possible extent of

the application of a separation process as a polishing step after the ASP.

3.5.1.3 Ozonation

400 mL of the filtrate was exposed to a fixed inlet concentration of

71 ± 17 mg O3/Lgas (produced by an Okamizu Food Detoxifier V.2 at a rate of

2.3 L air/min), which was injected into the filtrate via an air stone diffuser

placed at the bottom of a conical flask. The exhaust O3 leaving the reaction

vessel (O3 residence time < 1 second) was measured with Aeroqual S-200 ozone

meter and trapped in two subsequent 1 L bottles of 0.1 M KI solution.

Ozonation was carried out at room temperature (22°C ± 1) and varying

exposure time from 1 to 60 min. To avoid airborne contamination, ozonation

experiments and subsequent sample analyses were run within a fume cabinet.

The HRT of 24 h was later halved (12 h) to account for COD, BOD, TSS, P, TC

and TVC variation.

72
Combined acoustic cavitation and ozonation

3.6.1 Experimental setup

The setup used to establish different treatment combinations such as

activated sludge-ozonation, activated sludge-sonication and activated sludge-

ozonation-sonication systems, is shown in Figure 3.2.

Figure 3.2. A schematic diagram of the experimental setup: activated sludge-ozonation

system (ultrasonic plate transducer off), activated sludge-sonication system (ozonation

off) and activated sludge-ozonation-sonication system (ozonation and ultrasonic plate

transducer on).

3.6.1.1 Activated sludge reactor

The same procedure to that explained in section 3.5.1.1 was employed here,

with a feeding rate of 2 L/day, a HRT of 24 h and a SRT of 13 days. 400 mL of the

effluent was used during the experimental study.


3.6.1.2 Ozonation

For the ozonation experiments, the treated wastewater was exposed to a

fixed O3 inlet concentration of 71 ± 17 mg/Lgas (produced by an Okamizu Food

Detoxifier V.2 at a rate of 2.3 L air/min), which was injected into the sample via an

air stone diffuser placed at the bottom of a jacketed cylindrical glass vessel (15 cm

height and 6.7 cm inner diameter). O3 was injected from 1 to 60 min (O3 residence

time < 1 second). The exhaust O3 leaving the reaction vessel was measured with

Aeroqual S-200 ozone meter and trapped in two subsequent 1 L bottles of 0.1 M KI

solution. The KI solution bottles were used to prevent human exposure to O3, as

well as to avoid O3 to be released into the atmosphere. Ozonation experiments and

subsequent sample analyses were run within a fume cabinet.

3.6.1.3 Sonication

The same jacketed cylindrical glass vessel and sample volume used for

ozonation was used to run the sonication and the combined ozonation-sonication

experiments. Three US transducers (Honda Electronics Co. LTD) with resonance

frequencies of 44, 300 and 1000 kHz were used in this study. The selected

transducer was fixed at the bottom of the glass vessel (Figure 3.2) and driven at its

resonance frequency by a power amplifier (T&C Power Conversion AG1006)

coupled with an impedance matching unit. The same applied power of 40 W was

used for all three frequencies and the calorimetric powers measured at this applied

power were 17.1 W for 44 kHz, 34.4 W for 300 kHz, and 34.2 W for 1000 kHz.

During the combined ozonation-sonication experiments, O3 was injected at

the same concentration and rate as mentioned above. The temperature of the

solution was kept at 16 ± 3°C for all the experiments with a temperature control

74
system (Julabo FL300) set at 10°C and varying treatment time from 1 to 60 min. To

avoid airborne contamination, all the experiments and subsequent sample analyses

were run within a fume cabinet.

Ozone as a pre-treatment and combined to UVC/H2O2

3.7.1 Experimental setup

The schematic of the different systems (O3 pre-treatment, UVC/H2O2 and

combined O3/UVC/H2O2) used in the present study is shown in Figure 3.3. All the

units were connected by polytetrafluoroethylene (PTFE) tubing.

Figure 3.3. Schematic of the experimental setup. (a) O3 pre-treatment; (b) UVC/H2O2; (c)

Combined O3/UVC/H2O2.
3.7.1.1 Ozonation pre-treatment

The O3 pre-treatment was carried out in a separate reactor (Figure 3.3a),

with the ozonation system comprised of a BMT 802N ozone generator able to

produce up to 4 g O3/h, BMT DH3b de-humidifier and BMT heated catalyst ozone

destroyer. The O3 was generated in-situ from pure and dry oxygen and injected

through an air stone diffuser placed at the bottom of a cylindrical glass reactor

(ozonation column, internal diameter of 7.3 cm, maximum fluid column height of

37 cm and a volume capacity of around 1.5 L). Injected and exhaust O3 was

analysed by a BMT 964 ozone analyser and the gas flowrate measured by a gas

mass flow meter from Alicat Scientific. All the system was contained within a fume

cupboard to avoid O3 exposure.

3.7.1.2 Combined UVC and H2O2 system

For the UVC/H2O2 process (Figure 3.3b), a FluHelik photoreactor was

employed in a horizontal configuration (previously described by Moreira et al.

[175]) characterised by tangential inlet/outlet pipes that promote the generation of

a helical motion of fluid, and a concentric quartz tube filled by a low pressure 11 W

UVC lamp (HNS 11W G5 Osram Puritec) and powered by a power supply by

Paralab. The volumetric flowrate (75 L/h) and UVC power supply used in the

present study have been based on previously reported optimum conditions [175,

176]. The recirculation vessel for the effluent is a 1.5 L cylindrical glass vessel that

is placed on top of a magnetic stirrer (LBX S20 series). The recirculation vessel was

connected to a thermostatic bath (Julabo F12-EH) using a water jacket and the

temperature of the solution kept constant at 25 ± 1°C. Slaughterhouse wastewater

was pumped from the recirculation vessel into the FluHelik photoreactor by a gear

pump (Ismatec, model MVP-Z).

76
3.7.1.3 Combined ozone, UVC and H2O2 system

For the O3, UVC and H2O2 combined process (Figure 3.3c), the UVC/H2O2

setup (Figure 3.3b without the recirculation vessel) was coupled to the ozonation

pre-treatment system (Figure 3.3a). The ozonation column was used as the

recirculation vessel and the FluHelik photoreactor placed in a vertical

configuration to increase O3 residence time with the inlet located in the lowest part

of the glass cylinder and the outlet on its top.

3.7.2 Experimental procedure

3.7.2.1 Ozonation pre-treatment

A volume of 1.1 L of slaughterhouse wastewater was first added to the

ozonation column. O3 was then injected at a flowrate of 0.3 Lgas/min and an inlet O3

concentration of 45, 70 and 100 mg/Lgas for different pre-treatment times of 5, 10

and 15 min. During the experiments, injection O3 flowrate and inlet O3

concentration were kept constant and both inlet and outlet O3 concentrations

monitored, as well as temperature and pH.

3.7.2.2 Combined UVC and H2O2 system

The 1.1 L of O3 pre-treated effluent was transferred to the recirculation

vessel of the UVC/H2O2 setup (Figure 3.3b) for further treatment. The effluent was

pumped from the bottom of the recirculation vessel to the FluHelik photoreactor.

After passing through the photoreactor, the treated wastewater was pumped back

to the recirculation vessel. A certain volume of H2O2 was initially added to give an

initial oxidant concentration of 150, 300 or 450 mg H2O2/L and continuously stirred

at 400 rpm to provide a proper mixing. During the 3 h of treatment (UVC/H2O2

process), pH, temperature and H2O2 concentration were continuously monitored.


Samples were withdrawn from the recirculation vessel every 15 min. H2O2

concentration was maintained constant throughout the duration of the experiment

by replacing the consumed oxidant. A single initial addition of 490 mg H2O2/L was

also applied.

3.7.2.3 Combined ozone, UVC and H2O2 system

In this setup, the treatment volume had to be increased from 1.1 L to 1.6 L

in order to maintain a water column height in the ozonation column similar to the

previous O3 pre-treatment tests. Slaughterhouse wastewater together with

490 mg H2O2/L was poured into the ozonation column and recirculated at a fixed

flowrate of 75 L/h. To have a homogeneous mixing along the whole setup,

wastewater was recirculated from the ozonation column to the FluHelik and back

for 5 min before O3 was injected into the ozonation column at a flowrate of

0.3 Lgas/min, same as the O3 pre-treatment.

Compared to ozonation pre-treatment experiments, a much longer

ozonation time of 3 h and larger effluent volume of 1.6 L were used. Therefore, to

have the same amount of O3 injected per litre of wastewater as 10 min of 100 mg/Lgas

O3 pre-treatment (272 mg O3/Lwastewater), O3 injection concentration was reduced

from 100 mg/Lgas to 8.1 mg/Lgas. For that, an O3 dilution system was constructed

following supplier’s recommendation. This approach was taken in order to make

the best use of O3, as synergistic effects between O3 and UVC/H2O2 have been

reported [154, 156]. The impact of different O3 dosages was later measured by

reducing and increasing the inlet concentration to 5.6 mg/Lgas (189 mg O3/Lwastewater)

and 12.1 mg/Lgas (408 mg O3/Lwastewater). pH, temperature and H2O2 concentration, as

well as inlet and outlet O3 gas concentration were monitored continuously.

Samples were taken every 15 min after passing the FluHelik reactor.

78
Single and combined electrochemical oxidation driven processes

3.8.1 Experimental setup

Figure 3.4 shows the scheme of the two experimental units used in the

current study. The ozonation system (Figure 3.4a) was the one reported in section

3.7.1.1. All the system was maintained within a fume cupboard to avoid O3

exposure.

The electrochemical system (Figure 3.4b), used for EO, EO/UVC, EO/H2O2

and EO/UVC/H2O2 experiments, was mainly composed of: (i) a FluHelik

photoreactor, (ii) a continuous flow electrochemical filter-press cell, MicroFlowCell

from ElectroCell (Tarm, Denmark), detailed described elsewhere [177], equipped

with a 10 cm2 BDD anode and a 10 cm2 platinum (Pt) cathode (inter-electrode gap

of ~3.7 mm), and coupled to a MLINK DPS3005 power supply (0-5 A, 0 - 30 V) to

provide constant current density (galvanostatic mode), (iii) a 1.5 L recirculation

cylindrical glass vessel thermostatically controlled and magnetically stirred at

400 rpm to provide solution homogenisation, and (iv) a gear pump (Ismatec, model

BVP-Z) to flow the solution throughout the system at 18 L/h. The BDD electrode

comprised a conductive niobium sheet with 2 mm thickness coated with a BDD

thin film of around 5 μm thickness. The system units were connected by PTFE

tubing.
Figure 3.4. Scheme of the experimental setup: (a) ozonation system and (b) electrochemical

system.

3.8.2 Experimental procedure

The ozonation pre-treatment was carried out by filling the bubble column

with 1.4 L of the collected slaughterhouse wastewater and injecting a constant inlet

O3 concentration of 100 mg O3/Lgas at a gas flow rate of 0.3 L/min for 10 min. Injected

and exhausted O3 were continuously monitored, as well as temperature and pH.

To carry out EAOPs, a volume of 1.4 L of slaughterhouse wastewater (as

collected or pre-treated by ozonation) was placed into the recirculation vessel and

the gear pump was switched on to pump the solution through the system. Then,

the thermostatic bath was switched on and set at a temperature that allowed

solution to reach 25 ± 1°C. In EO/H2O2 and EO/H2O2/UVC processes, a content of

850 mg H2O2/L was initially added and after proper homogenisation (10 min) an

initial control sample was collected. For EO and EO/H2O2 processes, the reaction

was started by switching on the power supply at a constant current density of

80
100 mA/cm2. For EO/UVC and EO/H2O2/UVC processes, the UVC lamp was also

switched on. Samples were taken at different time intervals. During reaction, the

temperature of the thermostatic bath was regulated to keep the inner solution at

25 ± 1°C. pH was not adjusted during reaction.


Chapter 4

A combined activated sludge-filtration-

ozonation process for slaughterhouse

wastewater treatment

Introduction

The importance of a biological process (ASP) and the relevance of a filtration

system has been evaluated in this Chapter, as well as the performance of O 3 as a

post-treatmet method. This study has been carried out in order to reach direct

discharge limits and assess the potential of the combined process to reach water

reuse purposes. For that, along with organic matter indicators, microbial

disinfection has been monitored by the measurement of TC and TVC before and

after ozonation.

82
Results and discussion

4.2.1 Slaughterhouse wastewater characterisation

The average values of physico-chemical characteristics of the raw wastewater

collected from the slaughterhouse are shown in Table 4.1. The averages were

calculated based on 8 samples collected over a 2-month period. The variation in

the COD values were relatively small and varied between 1680 and 2047 mg O2/L.

BOD values were approximately three times lower compared to COD, and varied

between 466 and 786 mg O2/L. TSS values had a slightly higher fluctuation with

values ranging from 110 to 412 mg/L.

Table 4.1. Characteristics of raw wastewater collected from the slaughterhouse. The

averages were calculated from 8 samples collected over a period of 2 months.

Parameter (units) Concentration

COD (mg O2/L) 1804 ± 204

BOD (mg O2/L) 651 ± 89

TSS (mg/L) 250 ± 90

P (mg/L) 115 ± 25

pH 5.3 ± 0.1

4.2.2 Activated sludge process

Shown in Figure 4.1 are the values for MLSS in the ASP (24 h HRT) presented

as a function of time, which increase and reach a steady state at approximately

3300 mg/L MLSS in about 40 days, when a balance between new formed cells and

dead cells was obtained. Oxygen was not a limiting substrate for the growth of
heterotrophic and autotrophic microorganisms [36] as the system was close to

oxygen saturation values [178] with an average DO value of 8.1 mg O2/L.

Comparable MLSS values were obtained in a similar study by Chen and Lo [46]

setting the best operational MLSS and SRT values at 3200 mg/L and 9.94 days,

respectively. Moreover, the MLSS and SRT values (13 days present study) herein

obtained also fall within the range of operational parameters for ASP

recommended by the United States Environmental Protection Agency [3].

Also shown in Figure 4.1 is the MLVSS, whose values also exhibit the expected

similar increase and plateau pattern as observed for the MLSS. It is worthy to note

the low values for the MLVSS and MLVSS/MLSS ratio, with the latter ranging

between 0.26 and 0.32 (24 h HRT). Such low values indicate high content of

inorganic matter in the raw wastewater and is attributed to a poor slaughterhouse

wastewater pre-treatment, i.e., poor coagulation and flocculation. Others [32, 37]

obtained MLVSS/MLSS values between 0.65 and 0.85, indicating a better

separation of inorganics prior to the ASP.

84
4500 0.35
MLVSS MLSS MLVSS/MLSS
4000
0.30
3500
0.25
MLSS and MLVSS [mg/L]

3000

MLVSS/MLSS
2500 0.20

2000 0.15

1500
0.10
1000
0.05
500

0 0.00
0 10 20 30 40 50 60 70 80 90
Time [days]

Figure 4.1. MLSS and MLVSS progress as a function of time in the ASP, as well as

MLVSS/MLSS ratio. 24 h HRT and 13 days SRT.

4.2.3 A combined activated sludge-filtration-ozonation process

4.2.3.1 COD and BOD reduction

Once the steady state was established for the ASP, with and without a filtration

step, the mixed liquor from the ASP was ozonated for 17 min. The COD and BOD

of the liquid at each treatment step are plotted in Figure 4.2. It can be seen that the

filtration step had a negligible effect during the AFO process (24 h HRT ASP) which

reduced the COD and BOD levels down to 128 ± 41 mg O2/L and 12 ± 1 mg O2/L,

respectively. This corresponds to an average reduction of 92.9% (COD) and 98.1%

(BOD) relative to the raw wastewater. In this AFO process, the 17 min of ozonation

worked as a polishing step to slightly reduce the COD level while not affecting the

final BOD value. In this particular case, carbon was reduced almost entirely during
the biological process, resulting in a reduction of 90.1% (COD) and 97.4% (BOD).

The filtration step had no significant effect on COD and BOD. In addition, both the

COD and BOD were measured at longer ozonation times (30 and 60 min), but the

values were similar to those measured at 17 min. O3 has affinity towards

double/triple bonds and aromatic rings, degrading organic matter into single bond

compounds. After these bonds are broken down into simple compounds, the

oxidation process of O3 slows down and becomes less efficient [55, 179]. Therefore,

it is concluded that the COD and BOD values have reached a minimum at 17 min

and further ozonation up to 60 min had no effect on the COD and BOD.

The observed COD reduction by ozone in this study is low compared to that

reported by Proesmans et al. [68], where ozonation of slaughterhouse wastewater

reduced the COD from 89 mg O2/L (after biological treatment) to 30 mg O2/L (after

ozonation). Proesmans et al. [68] have used an inlet O3 concentration of up to

200 mg O3/Lgas (influent flowrate of 1.2 L/min). This O3 dose is significantly higher

than the one used in the present study and thus, higher injected O3 could result in

a higher COD removal [61]. The reactor the authors employed enhanced

O3·diffusion into the wastewater, potentially increasing ozonation performance.

Research has also shown the potential of O3 in reducing COD and BOD by applying

it to slaughterhouse wastewater without a previous biological treatment [17, 66].

The reduction of the ASP HRT from 24 h to 12 h had a minimal effect on organic

load removal where similar COD values were observed for both HRTs (Figure 4.2).

For the BOD values, decreasing the HRT increased the BOD from 17 mg O 2/L to

32 mg O2/L. This is attributed to the bacteria having a shorter residence time in the

ASP reactor for BOD reduction (such as hard BOD or less readily biodegradable

organic matter)[180, 181]. Although minimal, this led to a decrease in the final

reduction in the AFO process for BOD when the HRT was reduced from 24 h (98%)

to 12 h (97%).

86
(a)
250
12h HRT ASP 24h HRT ASP

200
COD [mg/L]

150

100

50

0
Activated Sludge (AS) AS+Filtration AS+Filtration+Ozone AS+Ozone

(b)
40
12h HRT ASP 24h HRT ASP
35

30
BOD [mg/L]

25

20

15

10

0
Activated Sludge (AS) AS+Filtration AS+Filtration+Ozone AS+Ozone

Figure 4.2. Average COD (a) and BOD (b) values for every step of the AFO process at

24 h and 12 h ASP HRT with 17 min ozonation time. The average COD and BOD values

for the raw wastewater are 1804 ± 204 mg O2/L and 651 ± 89 mg O2/L, respectively.
4.2.3.2 TSS reduction

Besides removing organic matter, the AFO system also efficiently reduced TSS.

Figure 4.3 depicts a final reduction of 98.8% after applying 17 min ozonation to the

filtrate using a HRT of 24 h during the ASP. Such a reduction is translated into a

TSS average value of 3 ± 2 mg/L, where similar values have been reported suitable

for irrigation [1]. 90.5% of the reduction was measured after the ASP, decreasing

TSS value from 250 mg/L (raw effluent) to 24 mg/L (mixed liquor).

70
12h HRT ASP 24h HRT ASP

60

50
TSS [mg/L]

40

30

20

10

0
Activated Sludge (AS) AS+Filtration AS+Filtration+Ozone AS+Ozone

Figure 4.3. TSS content for every step of the AFO process at 24 h and 12 h ASP HRT

with 17 min ozonation time. The average TSS value for the raw wastewater is

250 ± 90 mg/L.

As observed with carbon, filtration by 4-7 µm pore size filter paper did not

significantly reduce TSS and no further reduction to the TSS was measured when

the O3 exposure time was increased from 17 min to 60 min. However, from the

experimental results obtained during the preliminary phase (non-steady state for

the ASP), a notable reduction in all the measured parameters (COD, BOD, TSS and

88
P) was obtained after filtration compared to those before the filtration step (Table

4.2). Therefore, it could be underlined that filtration could work as a backup

process, along with ozonation, if the ASP fails (e.g. inappropriate settling

performance in the clarifier increasing solid concentration in the effluent).

Table 4.2. Concentration of COD, BOD, TSS and P before and after 4-7 µm filter paper

filtration during the non-steady state phase of the ASP.

COD (mg O2/L) BOD (mg O2/L) TSS (mg/L) P (mg/L)


Before filtration 270 ± 15 31 ± 10 30 ± 3 17 ± 5
After filtration 230 ± 10 15±1 16 ± 5 10 ± 5

Analysing samples from the 12 h HRT in the ASP, TSS reduction after the

combined system falls to 90% in comparison to that of 24 h HRT (98.8%), with a

final value of 22 mg/L after the AFO. This value is notably higher than the value of

3 mg/L for the 24 h HRT. A similar explanation used to address the different BOD

results could be applied to the TSS. That is, in the 12 h HRT ASP, bacteria had less

time to degrade the solids in the raw influent leading to a higher value in the ASP

effluent (24 mg/L and 60 mg/L after 24 h and 12 h HRT ASP, respectively).

Therefore, decreasing HRT could have adversely affected the solubilisation of

colloidal and particulate BOD, resulting in the increase in the final TSS

concentration [182].

4.2.3.3 P, pH and total dissolved nitrogen

P analyses show that the ASP reduced P values down to 19.5 ± 7 mg/L (83%),

further lowering P values to 13.2 ± 9 mg/L (88.5%) with filtration. The small

reduction during the filtration step would come from the removal of P attached to

suspended solids. After applying O3 for 30 min, P values in the order of

1.9 ± 1 mg/L (98.4% reduction) were measured. The reduction in P during

ozonation was probably related to the removal of P containing suspended solids


by flotation, as P does not directly react with O3 [183]. As observed with TSS,

increasing ozonation time up to 60 min did not further decrease P concentration.

The change in ASP HRT did not make a significant difference either, showing a

high P uptake by aerobic microorganisms [184].

In regards to pH, raw wastewater samples had a pH of around 5.3 while the

pH increased to 8.0 after the ASP. This increase in pH during the ASP could be

attributed to the formation of free ammonia during the ASP and its subsequent

reaction with CO2, produced during the aerobic process. This results in ammonium

bicarbonate, increasing alkalinity and generating a buffering capacity in the system

[15]. Additionally, ozonation caused a negligible increase in the pH to a value of

8.3, agreeing with already published data [1, 185]. During ozonation in liquid

media, oxidation can occur through direct molecular O3 (direct O3 oxidation

pathway when pH < 4) and via an indirect pathway through •OH formed during

O3 decomposition (radical pathway dominated degradation when pH > 10) [53, 54].

That is, at pH between roughly 4 and 8.5, both •OH and molecular O3 would take

part in the oxidation process of compounds when O3 is applied. Therefore, the

slight increase in pH to a value of 8.3 did not affect O3 mechanisms for

slaughterhouse wastewater degradation.

The total dissolved nitrogen content was also measured and was reduced by

30% (from 224 ± 95 mg/L for raw wastewater) after the ASP to a value of

160 ± 27 mg/L. Ammonia (NH4-N) reduction also occurred mainly in the

bioreactor, reducing NH4-N by 22% to a value of 116 ± 12 mg/L. In order to remove

nitrogen in a biological reactor, nitrification (oxidation of NH 4-N into nitrite (NO2- )

and nitrate (NO3-)) and denitrification (reduction of NO3- to nitrogen gas) processes

are required. Nitrification principally occurs in aerobic environments, while anoxic

conditions (absence of free or dissolved oxygen) are needed for denitrification. A

long SRT is necessary for nitrification due to a slow growth rate of nitrifiers

90
(principally Nitrosomonas and Nitrobacter), being the conversion of NH4-N to

NO2- and NO3- inhibited by certain heavy metals and polymers (such as those used

as coagulant and flocculant). This could explain the low NH4-N removal obtained.

At the same time, a low TN removal could be expected from the proposed ASP due

to a high DO concentration [186]. The removed portion of the TN could have come

from the denitrification produced in specific anoxic zones e.g. inside suspended

flocs or due to an uneven distribution of DO/improper mixing of the anaerobic

reactor. Finally, nitrate was reduced by 85% (initial value of 13.6 mg/L) during the

biological step (denitrification), achieving a value of 2 ± 1 mg/L after the ASP.

During filtration and ozonation, no change in the total dissolved nitrogen and

NH4-N was measured. Ammonia only rects with O3 when not protonated and

shows low rate constants at pH around 8, showing low oxidation rates when

produced secondary oxidants (e.g. •OH) are used in competing reactions [183].

The ASP was not design for nitrogen removal and therefore, the efficiency obtained

in the AFO was lower than those found in the literature for nitrogen reduction [22,

37, 45, 187].

4.2.3.4 Microbial counts

To assess disinfection efficiency, samples were analysed for TC and TVC. Prior

to the ozonation step, the average values of TC and TVC in the filtrate were found

to be 1.4×104 CFU/100 mL and 4.6×109 CFU/100 mL, respectively. These are typical

values reported in the literature for slaughterhouse wastewater [15, 43, 188].

4.2.3.4.1 Total coliforms

The TC as a function of ozonation time shown in Figure 4.4 reveal that

within the first two minutes of ozonation the number of TC remained constant.

This behaviour has been observed in the literture and is attributed to the
consumption of dissolved O3 by the presence of an initial high amount of organic

matter in the wastwater [189].

A complete inactivation of TC (4 log reduction) was achieved after dosing

the filtrate samples with O3 for 10 minutes (Figure 4.4). O3, however, has shown

potential to achieve [7, 190] up to 6 log of bacterial inactivation (Faecal Coliforms,

TC and E. coli mainly) for industrial and municipal wastewater.

Neither filtration nor ASP HRT reduction significantly affected the final

disinfection efficiency (a complete inactivation of TC was achieved with and

without filtration for both 12 h and 24 h ASP HRT after applying O3 for 10 min).

4.5

4
TC log survival [CFU/ 100 mL]

3.5

2.5

1.5

0.5

0
0 2 4 6 8 10 12 14 16 18
Ozonation time [min]

Figure 4.4. TC progress as a function of ozonation exposure after ASP (empty

symbols) and ASP-filtration (solid symbols). () 12 h HRT and () 24 h HRT.

71 ± 17 mg O3/L injected. Dotted line indicative of maximum acceptable levels of faecal

coliforms for unrestricted irrigation [191, 192].

92
The World Health Organisation recommends less than 1000 geometric mean

number of faecal coliforms per 100 mL for unrestricted irrigation for edible crops,

sports fields, and public parks [191]. When it comes to drinking water standards,

the maximum acceptable concentration of TC is zero per 100 mL, noting that for

water put into bottles or containers the limit is set at zero per 250 mL [193]. Both

limits were met after an O3 contact time of 10 minutes in the AFO system.

4.2.3.4.2 Total viable counts

O3 could not completely remove TVC, even after increasing the O3 exposure

time up to 1 hour, as seen in Figure 4.5. However, 5 log of inactivation was obtained

when subjecting the filtrate to 30 min of ozonation. The average TVC after

ozonation was 6×103 CFU/100 mL. Increasing O3 contact time to 1 hour did not

further inactivate TVC.

10

8
TVC log survival [CFU/100 mL]

0
0 5 10 15 20 25 30 35 40 45 50 55 60 65
Ozonation time [min]
Figure 4.5. TVC progress as a function of ozonation exposure after ASP (empty

symbols) and ASP-filtration (solid symbols). () 12 h HRT and () 24 h HRT.

71 ± 17 mg O3/L injected. Dotted line indicative of maximum acceptable levels of bacterial

contamination in drinking water [193].

The lack of TVC removal by further ozonation is attributed to the few TVC

that remained alive after 30 min of O3 exposure being resistant to O3. These TVC

could have developed defence mechanisms against O3 and be resistant to its

oxidation [194, 195], producing O3 resistant coatings [196] or the formation of O3

resistant pigments and biofilms [197]. Microbes could also have remained within

the suspended solids left in wastewater where the solids acted as a barrier against

O3 oxidation [194, 198]. The setup of the ozonation chamber may have also played

a role in the mixing and transfer efficiency of O3 in wastewater [199], resulting in a

lower TVC removal. It is important to mention that after 30 min contact time with

O3, all the TVC values in the wastewater were below the drinking water limit, set

by the European Commission at 100 CFU/mL [193].

Paper filtration had negligible effect on TC and TVC (Fig. 4.4 and Fig. 4.5),

however, this is specific to the type of filtration used in this study. For example,

Liberti et al. [200] had reported a 1 log difference on bacterial inactivation using O3

between clarified and clarified-filtered feeds when using a multilayer pressure

filter (deep bed sand filter) filled with high purity silica sand and gravel. The better

solid and microbial removal efficiency of the deep bed sand filter compared to the

filter paper used in the present study could be explained by the biofilm formation

which can improve microbial adsorption. Additionally, Liberti et al. [200] used a

dose between 7 and 15 mg O3/Lwastewater, while in this study a much higher O3 dose

was applied (71 mg O3/Lgas at 2.3 Lgas/min). Such a high O3 dose could have led to

the inactivation of almost all the microorganisms (< 100 CFU/ml), regardless the

amount of TSS before ozonation. Thus, comparing the amount of TSS between the

94
24 h ASP HRT (24 mg/L) and 12 h ASP HRT (60 mg/L), no difference in the final

TVC was measured after ozonating both feeds. At the same time, Venosa et al. [199]

concluded that filtration substantially improves coliform (total and faecal)

inactivation if secondary effluent is of poor quality, while not being necessary

effluent filtration before ozonation if total COD values are already low enough.

Conclusions

In the combined AFO system, filtration had no significant effect on organic

matter and TSS removal, while ozonation worked as a polishing step on the

reduction of COD, BOD and TSS. Filtration, however, showed the potential to

reduce the aforementioned parameters for the non-steady state ASP effluent and

therefore, it could work as a backup process, along with O3, if the ASP fails. At the

same time, the ASP HRT could be further reduced with a minimal effect on

substrate (BOD) removal efficiency, increasing the overall efficiency of the process.

The experimental results obtained with the AFO system proved the process is

effective in reducing organic load, as well as TSS and P for direct discharge [8]. TC

and TVC for drinking water standards were also met [193]. Based on these results

and considering the possible improvements for the ASP (e.g. SRT and HRT

adjustment, nitrogen reduction, etc.) and the ozonation chamber (e.g. increase O 3

transfer efficiency), it is reasonable to say that the combined AFO system has the

potential to reach water reuse levels.


Chapter 5

Tertiary treatment of real slaughterhouse

wastewater using combined acoustic

cavitation and ozonation

Introduction

No reports where found in the literature for the application of ultrasound on

the treatment of real slaughterhouse wastewater. Therefore, this chapter

investigates into the invidual and combination of O3 and ultrasound to treat real

slaughterhouse wastewater to reach direct discharge limits, with the possibility of

using the coupled system for direct water reuse applications. For that, a variation

in the applied frequency, ultrasonic reactor type and treatment time has been

evaluated.

Results and discussion

5.2.1 Slaughterhouse wastewater characterisation

Physico-chemical characteristics of the treated wastewater are shown in Table

5.1. After the lab scale ASP, slaughterhouse wastewater quality was not suitable

for direct discharge, exceeding COD (125 mg O2/L), BOD (25 mg O2/L) and TSS

96
(35 mg/L) values set by the EU [8]. Therefore, further treatment was required to

meet reported limits prior to discharge.

Table 5.1. Characteristics of treated wastewater. Sample variability from eight

different samples shown as standard deviation.

Parameters (units) Treated wastewater

COD (mg O2/L) 242 ± 48

BOD (mg O2/L) 53 ± 21

TSS (mg/L) 70 ± 23

TC (CFU/mL) 4.1×103 ± 3.8×103

TVC (CFU/mL) 4.1×107 ± 2.2×107

pH 7.7 ± 0.2

5.2.2 OH radical yield from ozonation and sonication

The rate of •OH production is important for understanding the effect of

sonication, ozonation and combined ozonation-sonication treatment has on the

degradation of organic matter. The common method of evaluating the amount of

•OH yield is via KI dosimetry, where the concentration of I3- can be assumed to be

proportional to the concentration of •OH (the production of •OH is believed to be

the dominant reaction, although other oxidising species could be measured) [201].

Figure 5.1 shows that sonication at 300 kHz gave the highest •OH yield compared

to 1000 kHz and 44 kHz. This optimum frequency for sonochemical yield has been

previously reported [77-80] and attributed to the balance between increase in

population of cavitation bubbles and decrease in cavitation collapse intensity with

increasing frequency [71, 72]. However, the •OH yield under sonication alone is

considerably lower compared to ozonation alone and for the combined ozonation-

sonication, similar •OH production rate as ozonation only were obtained. It is


interesting to note that a linear •OH production is obtained for the sonication and

ozonation systems, but for combined ozonation-sonication, the production rate

slightly reduces as the treatment time increases (time > 30 min).

900
44 kHz + O3 300 kHz + O3
800
1000 kHz + O3 O3 alone
700 44 kHz 300 kHz
1000 kHz
600
I3- [µM]

500 100

80
400
60
300 40

200 20

0
100 0 10 20 30

0
0 10 20 30 40 50 60 70
Treatment time [min]

Figure 5.1. I3- concentration as a function of time for ozonation (71 mg O3/L), sonication

(44, 300 and 1000 kHz at 40 W) and ozonation-sonication treatments (71 mg O3/L and 44,

300, 1000 kHz at 40 W). The insert shows the zoomed in plot of I 3- concentration as a

function of time for sonication only systems. Sample volume 400 mL and 0.1 M KI.

5.2.3 Effect of ozonation on water quality indicators

For O3 alone experiments, a substantial reduction in organic matter was

observed after 60 min of treatment. COD and BOD were reduced down to 133 ± 20

and 17 ± 4 mg O2/L, respectively (33% COD and 74% BOD removal), whereas TSS

decreased from 60 ± 11 mg/L to 20 ± 5 mg/L. At the same time, a complete

inactivation of TC was obtained after 30 min ozonation, while 537 ± 151 CFU/mL

of TVC were measured after 60 min ozonation (4 log reduction). These results agree

98
with those reported in Chapter 4, although the slightly higher final values obtained

herein can be attributed to the higher organic load and TSS content in the

slaughterhouse wastewater used in the experiments presented in this Chapter. The

time to obtain a complete inactivation of TC would also be affected by a higher

initial organic matter content, as explain in Section 4.2.3.4.1 Although minimal, the

difference in the type of reactor used between the two studies (Chapter 4 and

Chapter 5) could have also played a role in the O3 transfer efficieny into the liquid

media.

A slight increase in pH of 0.2 (initial pH 7.7) was measured after 60 min of

ozonation with no apparent change in degradation mechanisms (•OH and

molecular O3 attack). This is consistent with reported findings [53], showing that

an increase in pH from 6.7 to 7.9 (wastewater from a wastewater treatment plant)

does not affect significantly •OH generation. This strongly suggests that both •OH

and molecular O3 would play a role in organic matter removal.

5.2.4 Effect of sonication on COD and BOD

Shown in Figure 5.2 are COD and BOD values during 60 min sonication time

for the three frequencies (44, 300 and 1000 kHz) under study. Only the experiments

at 300 kHz showed a notable decrease in organic matter concentration with an 18%

reduction in COD to 228 ± 13 mg O2/L in 60 min and a 50% reduction in BOD to

13 ± 1 mg O2/L after 10 min. Further treatment up to 60 min showed no change in

the BOD value. The biodegradable carbon (measured by BOD) was probably easier

to oxidise within the first minutes of sonication and further increasing sonication

up to 60 min did not reduce BOD further, as the BOD value was already low. The

difference in the final COD and BOD values for the three frequencies could be

explained by the •OH production rate. It is well known that •OH production rate

peaks at around 300 kHz [75], producing more •OH compared to 44 and 1000 kHz
for a given applied power. As observed in Figure 5.1, after 60 min of sonication

and 40 W applied power, the concentration of •OH with 300 kHz was three times

higher than that of 1000 kHz and six times higher than with 44 kHz. The increase

in •OH concentration induces more chemical effects and therefore, leads to a better

degradation efficiency [202]. Similarly, Pétrier and Francony [77] reported that

higher phenol decomposition rate is achieved with 200 kHz compared to

frequencies of 20, 500 and 800 kHz. The highest formation of H2O2 (comparable to

•OH production) is also observed with 200 kHz, indicating a link between •OH

formation and phenol degradation. Others have also reported similar optimum

frequencies between 200-550 kHz for phenol degradation [77-80]. Berlan et al. [79]

obtained a complete removal of phenol with 541 kHz (30 W calorimetric power)

after 100 min (100 mg/L initial phenol concentration), while Lesko et al. [80]

achieved a 10% TOC reduction after 400 min sonication at 358 kHz and 100 W

applied power. The latter was substantially improved in combination with O3. It is

important to highlight that the aforementioned experiments were carried out with

synthetic waters (simple water mixtures), whereas the present study showed

results for complex water matrixes (real wastewaters).

100
350 120

300
100

250
80

BOD [mg O2/L]


COD [mg O2/L]

200
60
150

40
100

50 20

0 0
0 10 20 30 40 50 60 70
Treatment time [min]
44kHz COD 300kHz COD 1000kHz COD 44kHz BOD 300kHz BOD 1000kHz BOD

Figure 5.2. COD and BOD values as a function of sonication time for three different

frequencies (44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed as

standard deviation.

5.2.5 Impact of sonication on TSS

It can be seen in Figure 5.3 that TSS values decreased progressively as

sonication time increased. After 60 min of treatment, an average TSS removal of

25% was achieved, independently of the applied frequency. Although the lowest

TSS value (45 ± 9 mg/L) was obtained with 300 kHz, it had a lower initial TSS value.

The lack of frequency effect on the TSS suggests that although mechanical effects

are higher at lower frequencies (44 kHz), cavitation intensities at higher

frequencies (300 and 1000 kHz) are strong enough to fragment suspended solids

into soluble particles. That is, even though stronger cavitation is produced with
low frequencies, the higher number of cavitation events produced with high

frequencies [72] could lead to similar TSS reduction.

90

80

70

60
TSS [mg/L]

50

40

30

20

10

0
0 10 20 30 40 50 60 70
Treatment time [min]
44kHz TSS 300kHz TSS 1000kHz TSS

Figure 5.3. TSS values as a function of sonication time for three different frequencies

(44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed as standard deviation.

5.2.6 Effect of sonication on TC and TVC

The disinfection performance of US plate transducers is shown in Figure 5.4.

After 60 min of sonication at the three frequencies studied, less than 1 log reduction

in TC and insignificant removal in TVC was measured. The high amount of

suspended solids and the complex matrix of the treated wastewater would be

responsible for the poor disinfection performance of US. Literature, however, has

shown the potential ultrasonic plate transducers have in microbial inactivation. A

5 log reduction was obtained by Hua and Thompson [106] using a 205 kHz multi-

102
frequency reactor and a sample volume of 300 mL. These experiments, nonetheless,

were carried out at an approximate total power of 128 W (3 times the power used

in the present study) and with ultrapure water inoculated with E. coli.

Consequently, the treated effluent had no organic matter or any other inorganic

compound that could reduce the disinfection performance of ultrasound.

Additionally, the tests were performed in oxygenated solutions where the type and

amount of the dissolved gas can influence the formation of free radicals and thus,

lead to a better performance.

7
TC and TVC log survival [CFU/mL]

0
0 10 20 30 40 50 60 70
Treatment time [min]
44kHz TVC 300kHz TVC 1000kHz TVC 44kHz TC 300kHz TC 1000kHz TC

Figure 5.4. TC and TVC log survival values as a function of sonication time for three

different frequencies (44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed

as standard deviation.

A 20 kHz ultrasonic horn was also employed to treat two different sample

volumes (50 and 100 mL) of slaughterhouse wastewater for 30 seconds at two
different applied powers (210 W for 30% amplitude and 350 W for 50% amplitude).

The results are shown in Apendix A. No significant difference was observed in

COD, BOD and TSS, or in TC and TVC before and after the treatment. The short

treatment times applied would not be enough to degrade organic matter and TSS,

as well as disinfect microorganisms through mechanical and chemical effects

produced. Increasing treatment time (30 min) at a high applied power (1120 W/L),

up to 4 log E. coli inactivation can be obtained with a 20 kHz horn [112]. Although

low frequency ultrasonic horns have shown the ability to inactivate

microorganisms [112], the complex water matrix slaughterhouse wastewater

presents and the short treatment times employed during the present study would

make ultrasonic horns not effective for microbial disinfection. Incresing treatment

time at high applied powers could make the implementation of ultrasonic horns in

treatment plants effective [112] but economically difficult.

5.2.7 Effect of combined ozonation-sonication on COD and BOD

The combined ozonation-sonication experiments (Figure 5.5) performed

slightly better for COD removal (44% reduction after 60 min treatment with a final

value of 114 ± 15 mg O2/L) compared to the results achieved with O3 alone. It is

important to highlight that frequency variation does not play a role in COD

removal for the combined system. The slight increase in performance of the

combined system could be explained by the additional benefit of US at increasing

O3 mass transfer into the aqueous solution (i.e. increasing molecular O3

concentration in the water matrix). The improvement would not be caused by

direct •OH attack, as the results obtained show no difference in •OH production

rates between O3 alone and ozonation-sonication systems (Figure 5.1). As with O3

alone experiments, an insignificant increase in pH of 0.2 was observed. The slight

difference in COD and BOD removal between O3 alone experiments in Figure 4.2

104
(Chapter 4) and Figure 5.5 could be related, as stated earlier, to the initial organic

matter load (higher in this Chapter), as well as in the different reactors employed.

250

200
COD and BOD [mg O2/L]

150

100

50

0
0 10 20 30 40 50 60 70
Treatment time [min]
44kHz+O3 COD 300kHz+O3 COD 1000kHz+O3 COD O3 COD

44kHz+O3 BOD 300kHz+O3 BOD 1000kHz+O3 BOD O3 BOD

Figure 5.5. COD and BOD values as a function of O3 alone (71 mg /L) and O3/US

(71 mg O3/L and 44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed as

standard deviation.

The small difference in efficiency between the combined system and O3 alone

could be related to the high O3 dose used. Considering the short residence time of

the ozonated bubbles within the reactor (1 second approximately), it is unlikely

that US would cause a significant increase in the performance of the system. In

addition, the degassing of the system by US [203] could have also reduced, to some

extent, the dissolved O3 in the water matrix thus reducing the degradation

efficiency in the combined configuration. However, since O3 was continuously

injected during the experiment, the degassing effect of US in the combined system
would be minor. Boczkaj et al. [138] used a similar O3 dose (9.41 g/h vs 9.8 g/h of

our system, equivalent to injecting 71 mg /Lair at 2.3 Lair/min) coupled with HC (2

mm diameter throat Venturi tube at 6-10 bar inlet pressure and a flowrate of

470 - 590 L/h) to treat 5 L of real wastewater from bitumen production for 6 h. In

that case, the initial COD concentration ranged between 8000 and 12000 mg O 2/L

(compared to 195 mg O2/L here). Consequently, this combination of O3 and HC led

to a substantial improvement in performance after 60 min of treatment (30% COD

removal with the combined system compared to approximately 8% reduction for

O3 and HC alone). The reported improved performance under the combined

system could be attributed to the much higher COD which would result in a faster

consumption of O3 and •OH. Therefore, the application of US would increase O3

mass transfer into the aqueous phase, replacing the consumed O3 and leading to a

higher COD removal percentage. On the contrary, Ibañez et al. [139] reported no

additional COD reduction when combining O3 (7 - 12 mg O3/L) and US for urban

wastewater treatment. In that case, no significant difference was found on the

reduction of different contaminants between O3 alone and O3-US combined. The

reported COD concentrations (26-50 mg O2/L) were substantially lower than those

given by Boczkaj et al. [138] and the present study.

When it comes to the use of O3/US systems to treat wastewaters, it appears that

the role US plays in performance efficiency is inversely related to the O3 dose used.

It is also important to take into account the initial concentration of organic carbon

(the initial COD value seems to be crucial in the performance improvement of the

combined system compared to individual systems) and sonication power. Thus,

Weavers et al. [204] and Barbier and Petrier [205] obtained a significant increase in

carbon removal adding US to low O3 injection dosages (0.01 g/h and 7-8 mg/L,

respectively) compared to O3 alone. Whereas Tezcanli-Guyer and Ince [206] and

Destaillats et al. [207] showed only a slightly better mineralisation performance

106
with ozonation-sonication compared to ozonation alone when using a

substantially higher O3 dose of 40 mg/L and 15 mg/L, respectively.

BOD reduced progressively as the treatment time increased with no significant

difference between the two systems (O3 alone and ozonation-sonication). Final

BOD values of 14 ± 5 mg O2/L were measured after 60 min of treatment for both

systems, leading to a reduction in BOD of 74% independent of the frequency

(Figure 5.5). Similar values were obtained after applying 300 kHz alone (13 ± 1 mg

O2/L) with a final BOD reduction of 50%.

5.2.8 Effect of combined ozonation-sonication on TSS

As observed with BOD, TSS values did not show any difference between the

combined system (independent of the frequency used) compared to O3 alone,

achieving 18 ± 6 mg/L after 60 min (Figure 5.6). This value represents a 78%

reduction, substantially higher than that obtained with US alone (25%). The

organic fraction of suspended solids is composed of natural organic compounds

which are prone to hydrolysis and oxidation due to ozonation. O3 reacts with the

organic fraction of suspended solids to convert it into dissolved organic matter,

and further mineralisation occurs at higher O3 dose. Additionally, O3 can also

remove suspended solids by flotation.


80

70

60

50
TSS [mg/L]

40

30

20

10

0
0 10 20 30 40 50 60 70
Treatment time [min]

44kHz+O3 TSS 300kHz+O3 TSS 1000kHz+O3 TSS O3

Figure 5.6. TSS values as a function of O3 alone (71 mg /L) and O3/US (71 mg O3/L and

44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed as standard deviation.

5.2.9 Effect of combined ozonation-sonication on TC and TVC

As opposed to the disinfection efficiency of US alone, both the combined

system with 44, 300 and 1000 kHz, as well as O3 alone were able to achieve a

complete inactivation of TC after 30 min (Figure 5.7). The high O3 dose applied

would be responsible for the observed TC inactivation, as stated in Chapter 4. In

this case and as opposed to results shown in Chapter 4 (Figure 4.4), 10 min

ozonation where not enough to achieve a complete inactivation of TC. This is

probably due to a higher initial organic load and TSS concentration of the samples

used in Chapter 5. Regarding TVC removal, Figure 5.7 shows a slight improvement

for the combined system compared to O3 alone.

108
8

7
TC and TVC Log survival [CFU/mL]
6

0
0 10 20 30 40 50 60 70
Treatment time [min]

44kHz+O3 TVC 300kHz+O3 TVC 1000kHz+O3 TVC O3 TVC

44kHz+O3 TC 300kHz+O3 TC 1000kHz+O3 TC O3 TC

Figure 5.7. TC and TVC log survival values as a function of O3 alone (71 mg /L) and

O3/US (71 mg O3/L and 44, 300 and 1000 kHz) at 40 W applied power. Error bars expressed

as standard deviation.

It is important to emphasise that even though the ozonation-sonication process

did not show any synergism, it was the only treatment method that reduced the

COD, BOD and TSS levels down to direct discharge limits. Additionally, drinking

water standards were met (99 ± 23 CFU/mL) with ozonation-sonication

experiments, while O3 alone (537 ± 151 CFU/mL) was insufficient to reach the

minimum value of 100 CFU/mL (Table 5.2) set by the Council Directive 98/83/EC

on the quality of water intended for human consumption [193].


Table 5.2. COD, BOD, TSS, TC and TVC values before and after the combined

ozonation-sonication process, as well as direct discharge limits (COD, BOD and TSS) and

drinking water standards (TC and TVC).

COD BOD TSS TC TVC

(mg O2/L) (mg O2/L) (mg/L) (CFU/mL) (CFU/mL)

2.0×10 7±
Initial value 198 ± 34 65 ± 3 60 ± 11 27.1 ± 4.1
1.1×107

Final value O3 133 ± 20 17 ± 4 20 ± 5 0 537 ± 151

Reduction O3 33% 74% 67% 100% 5 log

Final value
114 ± 15 14 ± 5 18 ± 6 0 99 ± 23
US +O3

Reduction
44% 78% 70 % 100% 5 log
US +O3

Direct discharge
125 25 35 - -
limits [8]

Drinking water
- - - 0 100
standards [193]

Conclusions

US alone was not efficient in reducing organic carbon and inactivating

microorganisms when treating real wastewater such as that coming from a

slaughterhouse. 300 kHz was the only frequency showing a reduction in organic

matter, obtaining 228 mg O2/L COD and 13 mg O2/L BOD, while no microbial

inactivation was measured after 60 min of treatment. Combining US with O3, on

the contrary, led to a significant increase in COD and BOD removal (down to 114

and 14 mg O2/L, respectively), as well as achieving complete inactivation of TC and

a 5 log reduction in TVC. Furthermore, a substantial performance improvement

was seen in TSS removal using ozonation-sonication. The reduction percentage in

110
TSS was increased from 25% (sonication alone) to 78% (ozonation-sonication),

obtaining 18 mg/L with the coupled system. O3 was, however, the main treatment

process leading organic matter removal and microbial disinfection. Nevertheless,

the combined system was the only treatment method in our study (compared to

sonication and ozonation alone) able to reach direct discharge limits for COD, BOD

and TSS, as well as meeting drinking water standards for microbial disinfection.
Chapter 6

Ozone as a pre-treatment and combined to

UVC/H2O2 as tertiary treatment for real

slaughterhouse wastewater

Introduction

After an on-site biological treatment, the treated effluent was further processed

in the lab using UVC and H2O2. Considering the importance of H2O2 concentration

and the addition mode in the performance of the process, the system was subjected

to different H2O2 concentrations and injection systems. The relevance of light

filtering effects were also addressed by subjecting the combined process to pre-

ozonated and non pre-ozonated samples, varying ozonation pre-treatment time

and dose. In order to evaluate any possible synergy between the applied processes,

O3, UVC and H2O2 were later combined in a single configuration. For that, organic

macroparameters were monitored to evaluate the potential of the system and

compared to values reported in direct water discharge regulations.

It is important to highlight that the oxygen consumed during COD analysis is

not specific of organic matter oxidation, as the oxygen used by some inorganic

112
chemicals can also be accounted [171]. Therefore, in order to have a better

understanding of organic matter degradation in view of a complete degradation to

CO2 (mineralisation), DOC was measured, along with COD, during experiments

carried out in Chapters 6 and 7. DOC was also monitored in order to have an

indication of organic matter concentration, as high concentrations of H 2O2 may

interfere in COD analysis leading to erroneous final values [208, 209]. Additionally,

the adoption of direct measurements for organic matter indication have been

suggested as water quality indices [210].

Results and discussion

6.2.1 Slaughterhouse wastewater characterisation

The physico-chemical characteristics of the wastewater collected from the pig

slaughterhouse are summarised in Table 6.1. From the table one can highlight the

following main characteristics of the collected slaughterhouse wastewater, similar

to the cattle slaughterhouse wastewater after the lab scale ASP employed in

Chapters 4 and 5: (i) organics content exceeding the emission limit value imposed

by the European legislation [8] for discharge from urban WWTPs in terms of COD

(186 versus 125 mg O2/L), (ii) solids content slightly above the European emission

limits [8] in regards to the TSS content (40 versus 35 mg/L), and (iii) visible dark

brown colour, corresponding to 254 mg Pt-Co/L, which is above the typical

acceptable colour limit values for treated wastewaters leaving WWTPs, i.e.

25 - 75 mg Pt-Co/L [211, 212], depending on the nature of the receiving water body

(river, sea, lake, etc.).


Table 6.1. Physico-chemical characteristics of the pig slaughterhouse wastewater after

secondary biological treatment. Average values from minimum ten samples.

Parameter (units) Collected slaughterhouse wastewater


DOC (mg/L) 38±8
COD (mg O2/L) 186±21
TSS (mg/L) 40±13
BOD (mg O2/L) 47±4
pH 7.8±0.1
Turbidity (NTU) 15±1
Colour (Pt/Co) 254±33
Conductivity (µS/cm) 5420±120
Cl- (mg/L) 974±43
NO3- (mg/L) 126±32
NO2- (mg/L) 49±35
PO43- (mg/L) 96±15
SO42- (mg/L) 50±2
NH4+ (mg/L) 152±57

6.2.2 Ozonation pre-treatment

The effect of different injected O3 dosages on the DOC and COD is shown in

Figure 6.1a. No reduction in DOC occurred for the three injected O3 dosages (45,

70 and 100 mg/Lgas) for 30 min, while a reduction in COD with increasing O3

dosages was observed. O3 reacts fast on unsaturated bonds (double/triple bonds

and aromatic rings) to break large molecules into smaller molecules. Once these

bonds are broken the oxidation process reactions slow down and ozonation

becomes less efficient. This indicates organic matter degradation (reduction in

COD) up to a point, where refractory by-products for O3 attack were probably

formed (such as carboxylic acids). [55, 179, 213]. Therefore, O3 did not manage to

degrade further and mineralise these refractory by-products and hence, DOC

concentration remained constant [185, 214, 215]. During O3 pre-treatment, some

experiments (Figure 6.1a) showed a slight increase in DOC concentration after

ozonation. The rise in DOC indicates the dissolution of particulate organic matter

114
during the oxidation step (hence the reduction in TSS), increasing the carbon

content of water [216].

Comparing COD values obtained after 30 min ozonation from Chapters 4 and

5 (12248 mg O3/Lwastewater), a slightly higher COD removal was obtained in Chapter

6 with 100 mg O3/Lgas (818 mg O3/Lwastewater). Even though a significantly lower

injected O3 dose was used per litre slaughterhouse wastewater in this Chapter, the

higher O3 mass transfer efficiency probably obtained from the diffuser and the

ozonation column employed (higher O3 residence time) would have led to a better

O3 performance. This also highlights the excess O3 dosage employed in Chapters 4

and 5.

100 mg O3/Lgas was used to investigate the O3 pre-treatment time on TSS, colour

and turbidity values, which were observed to decrease until a plateau was reached

at approximately 10 min (Figure 6.1b). The reduction in TSS and turbidity can be

attributed to the reaction between O3 and the organic fraction of the suspended

solids (susceptible to hydrolysis and oxidation through O 3), converting it into

dissolved organic matter [217]. Colour reduction presumably occurred when O3

attacked and degraded chromophoric molecules [218] present in slaughterhouse

wastewater.
(a)
200
DOC [mg/L] and COD [mg O2/L] 180
160
140
120
100
80
60
40
20
0
0 5 10 15 20 25 30 35
Time [min]

(b)
60 350
TSS
Turbidity 300
TSS [mg/L] and turbidity [NTU]

50 Colour
250
40

Colour [Pt/Co]
200
30
150
20
100
10 50

0 0
Slaughterhouse 5' O3
5'O3 10' O3
10'O3 15' O3
15'O3
wastewater

Figure 6.1. (a) DOC (empty symbols) and COD (filled symbols) as a function of time.

O3 inlet concentration: (∆, ▲) 45 mg O3/Lgas; (□, ■) 70 mg O3/Lgas; (◊, ♦) 100 mg O3/Lgas.

Injected gas at a flowrate of 0.3 Lgas/min. (b) Turbidity, TSS and colour as a function of O3

pre-treatment time. 100 mg O3/Lgas injected at a flowrate of 0.3 Lgas/min. Error bars

expressed as standard deviation.

116
6.2.3 Ozonation pre-treatment followed by combined UVC and H2O2

The coupling of UVC and H2O2 (Figure 6.2a) showed a large synergistic

decrease in DOC of 41% compared to no DOC reduction from the individual

application of UVC and H2O2 (after 10 min of 100 mg O3/Lgas pre-treatment). UVC

and H2O2 individual experiments without O3 pre-treatment did not reduce DOC

either (Figure B 2 in Appendix B). During the H2O2 individual experiment (without

UVC), Figure 6.2b shows that the H2O2 was not consumed while a reduction in

H2O2 concentration was observed during the UVC/H2O2 experiment

(300 mg H2O2/L constant concentration coupled with UVC). In order to keep a

constant concentration of H2O2 at 300 mg/L for the combined UVC/H2O2 process,

H2O2 was added four times during the 3 h of treatment. The reduction in H2O2

suggests •OH attack as the main degradation mechanism, probably due to the

production of •OH through the cleavage of H2O2 molecules by UVC light [23].

These results differ from those reported by Barrera et al. [150] where they

mentioned direct photolysis to be the main mechanisms for TOC removal. The

difference could be attributed to the use of real (present study) and synthetic [150]

slaughterhouse wastewaters, highlighting the importance of using real industrial

samples. Comparing the two studies, DOC degradation mechanisms could have

completely changed when moving from synthetic to real wastewaters. Compared

to synthetic waters, real samples contain TSS and high turbidity that would lead to

light filtering effects [123, 124]. Therefore, DOC removal by direct photolysis

would be significantly reduced when treating real industrial effluents. The

difference in organic matter composition and other dissolved compounds present

in the wastewaters could also be considerable, with more complex compounds

found on real slaughterhouse wastewater [217, 219]. Thus, the high oxidation

potential •OH poses would mineralise these complex molecules, while direct

photolysis could be enough to reduce DOC significantly in synthetic


slaughterhouse wastewaters. Luiz et al. [14] reported a significant increase in COD

removal (from 31 to 48%) when adding H2O2 to a UV254 process (H2O2/COD = 2) for

real slaughterhouse wastewater treatment. The authors also mentioned that

H2O2/UV removed aromatic compounds 5.2 times faster than the application of UV

alone, stating that the effectiveness of the H2O2/UV process is highly dependent on

water quality parameters.

118
(a)
1.2

1.0

0.8
DOC/DOC0

0.6

0.4

0.2

0.0
0 20 40 60 80 100 120 140 160 180
Time [min]

(b)
600

500

400
H2O2 [mg/L]

300

200

100

0
0 20 40 60 80 100 120 140 160 180
Time [min]
Figure 6.2. (a) Relative DOC concentration as a function of time. (b) H2O2 concentration

as a function of time. (■) 10 min O3 + UVC; (●) 10 min O3 + H2O2 300 mg H2O2/L constant

concentration; (◊) UVC/H2O2 300 mg H2O2/L constant concentration; (∆) 10 min

O3 + UVC/H2O2 300 mg H2O2/L constant concentration (H2O2 added during the 3 h process

to maintain the concentration constant); (□) 10 min O3 + UVC/H2O2 490 mg H2O2/L initial

addition.
In order to evaluate the role O3 pre-treatment time could play in DOC removal

of the UVC/H2O2 process, non-ozonated samples, as well as 5, 10 and 15 min

ozonated samples (100 mg O3/Lgas at 0.3 L gas/min) were subsequently treated by

UVC/H2O2 for 3 h (constant concentration of H2O2). Results showed (Table 6.2) that

increasing O3 pre-treatment time, DOC removal of the UVC/H2O2 process

increased, obtaining best results with 10 min O3 pre-treatment. It is important to

highlight the difference between 10 min ozonated (41% DOC removal) and non-

ozonated (26% DOC removal) UVC/H2O2 post-treatment (300 mg H2O2/L),

showing a significant improvement of the process’ performance (Figure 6.2a). This

is attributed to the minimum colour and turbidity values (Figure 6.1b) reached

with 10 min O3 pre-treatment time, which would lead to an increase in light

transmissibility and the subsequent increase in •OH production through the

cleavage of more H2O2 molecules by UVC. At the same time, solids present in the

slaughterhouse wastewater would act as •OH scavengers [149] and therefore, by

reducing the TSS with 10 min O3 pre-treatment (Figure 6.1b) would lower this

scavenging effect. The degradation of organic matter achieved during O 3 pre-

treatment (conversion of original organic compounds into easily oxidisable

compounds) would also lead to an increase in DOC removal during post-

treatment. This would explain the difference in DOC removal between UVC/H2O2

processes with and without O3 pre-treatment. The O3 pre-treatment also reduced

consumed H2O2 per mg DOC removed of the subsequent UVC/H2O2 process,

lowering it from 52.4 mg H2O2 consumed/mg DOC removed for non pre-ozonated

UVC/H2O2 experiments to 28.5 mg H2O2 consumed/mg DOC for UVC/H2O2 trials

with pre-ozonation.

120
Table 6.2. Percentage reduction of DOC by applying combined UVC and H2O2 after 3

hours. Three different ozonation pre-treatment times (5, 10 and 15 min) at an inlet O3

concentration of 100 mg/Lgas at 0.3 Lgas/min. A constant concentration of H2O2 at three H2O2

dosages: 150, 300 and 450 mg/L. Standard deviation measured for at least two independent

experiments.

5'O3 + 10'O3 + 15'O3 +


UVC/H2O2
UVC/H2O2 UVC/H2O2 UVC/H2O2
150 mg H2O2/L - - 30±3 34±0.3
300 mg H2O2/L 26±0.3 32±0.9 41±3 35±0.8
450 mg H2O2/L 21±0.3 31±3 29±0.6 41±0.4

The most efficient H2O2 concentrations was later explored for the process under

study, keeping a constant concentration of H2O2 at three different values (150, 300

and 450 mg/L) for the whole duration of the experiment. Table 6.2 shows there

exists an optimum concentration of H2O2 with different ozonation pre-treatment

times. With 10 min O3 pre-treatment, an optimum H2O2 concentration was found

at 300 mg H2O2/L (41% DOC removal), whereas with 15 min, the maximum DOC

degradation was obtained with a H2O2 concentration of 450 mg/L (40.5% DOC

removal). However, a longer O3 pre-treatment time and a higher H2O2

concentration were required for the latter to reach a similar DOC reduction. With

10 min O3 pre-treatment time, increasing H2O2 concentration up to 450 mg/L would

provide an excess of the oxidant and contribute to radical scavenging effects, while

reducing H2O2 concentration to 150 mg/L would lead to a lower production of

•OH [220, 221]. That is, 150 mg H2O2/L would not be enough to meet H2O2 demand
during the UVC/H2O2 process, probably due to a low molar extinction coefficient

of H2O2 [23]. It is important to highlight that after 10 min ozonation, colour values

obtained prior to UVC experiments at 450 mg H2O2/L were higher (80 mg Pt-Co/L)

than those measured before 300 and 150 mg H2O2/L experiments (25 mg Pt-Co/L).

Therefore, the higher initial colour values, along with H2O2 scavenging effects
would be responsible for the lower DOC removal obtained (29%) with UVC/H 2O2

at 450 mg H2O2/L.

With the best configuration (10 min O3 pre-treatment followed by UVC/H2O2 at

a constant 300 mg H2O2/L), 490 mg H2O2/L were consumed during the 3 h

experiment. Therefore, 490 mg H2O2/L were injected at the beginning of the

UVC/H2O2 process (single H2O2 addition) after 10 min O3 pre-treatment. Similar

DOC degradation rates and reduction percentages were obtained in both

experiments (41% ± 3 constant H2O2 concentration vs 42% ± 0.9 single H2O2 initial

addition. Figure 6.2a). This was translated into a DOC reduction from 36.1 mg/L to

20.5 mg/L during the UVC/H2O2 process (490 mg/L H2O2 single addition). Residual

H2O2 was, however, reduced to more than half with the latter (from 273 mg/L to

102 mg/L, respectively. Figure 6.2b). This suggests that no major difference in DOC

removal would be observed for the system under study when H2O2 concentration

falls in the range between 150 and 450 mg/L. As DOC degradation progresses

during the UVC/H2O2 process (reduction in organic carbon as a function of

treatment time), fewer •OH would be needed to degrade the remaining DOC. That

is, a higher amount of H2O2 would make no difference in the reduction of DOC.

Lowering H2O2 to values below 150 mg/L, H2O2 consumption rate was significantly

lowered (Figure B 1a in Appendix B). This would subsequently reduce

considerably the production of •OH and hence, preventing further DOC removal

(Figure B 1b). This result is in disagreement with a previous study [150], where the

importance of keeping a constant concentration of H2O2 throughout the entire

treatment process was highlighted for a synthetic slaughterhouse wastewater.

Barrera et al. [150] mentioned that the decay of H2O2 during the 2.5 h of treatment

would lead to a low production of •OH, making the role of radicals on TOC

removal minimal. The present study, on the contrary, showed a significant

improvement in DOC reduction when UVC and H2O2 were simultaneously

122
applied with a single initial addition of 490 mg H2O2/L, attributing DOC removal

principally to •OH attack.

6.2.4 Combined ozone, UVC and H2O2 system

In order to increase process efficiency, O3, UVC and H2O2 (490 mg/L single

initial addition) were simultaneously applied. Figure 6.3a shows that the only

combination able to reduce DOC considerably was O3/UVC/H2O2 using an injected

O3 concentration of 8.1 mg/Lgas. In this case, O3 acted as a colour removing agent

(Figure 6.3b). Colour, however, did not play a major role in the reduction of DOC,

as similar colour removal rates and final values were obtained with O 3, O3/UVC

and O3/UVC/H2O2. Once again, the interaction between UVC and H2O2 seemed to

be the main cause for DOC removal. Although O3/UV and O3/H2O2 showed low

DOC removal, literature has shown these systems can degrade organic matter

significantly by reducing COD of a real livestock wastewater, attributing COD

removal principally to •OH attack [209]. Therefore, O3/UVC and O3/H2O2 can still

degrade organic matter present in slaughterhouse wastewater without obtaining a

significant mineralisation.
(a)
1.0
0.9
0.8
0.7
DOC/DOC0
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0 20 40 60 80 100 120 140 160 180
Time [min]

(b)
500
450
400
350
Colour [Pt/Co]

300
250
200
150
100
50
0
0 20 40 60 80 100 120 140 160 180
Time [min]

Figure 6.3. (a) Relative DOC and (b) colour as a function of time. (∆) 8.1 mg O3/Lgas

injected; (○) 8.1 O3/Lgas injected combined with UVC; (×) 8.1 mg O3/Lgas injected combined

with H2O2; (□) 8.1 mg O3/Lgas injected combined with UVC and H2O2. Initial injection of

490 mg H2O2/Leffluent.

124
With the aim of adjusting O3 injected dose, higher (12.1 mg O3/Lgas) and lower

(5.6 mg O3/Lgas) concentrations were employed combined with UVC and H2O2. As

shown by Figure 6.4a, no significant difference was found between the three O3

injected dosages in terms of DOC reduction. That is, combining O3/UVC/H2O2 with

an injected O3 concentration of 5.6 mg/Lgas 43% ± 2 DOC removal was obtained

(DOC removal from 35.5 to 20.4 mg/L), similar to that achieved with 8.1 mg O3/Lgas

(42% ± 4) and 12.1 mg O3/Lgas (39.2%). An increase in O3 dosage can lead to an

increase in •OH scavenging through the production of hydroperoxyl radical

(•HO2) and oxygen (O2) [154]. That is, •OH scavenging by an increase in O3 dose

could be a sink for •OH production [222]. Consumed H2O2 per removed mg DOC

was also measured similar for the three O3 inlet concentrations under study

(29.5 ± 4 mg H2O2 consumed/mg DOC removed). Comparing DOC removals with

results of 10 min O3 followed by UVC/H2O2 with single initial addition of

490 mg H2O2/L, no clear deviation in the DOC reduction rate was observed either.

This suggests that 5.6 mg O3/Lgas would be enough for the maximum removal of

DOC using the combined O3/UVC/H2O2 system. COD was also reduced to

61 mg O2/L and colour down to 20 mg Pt-Co/L, achieving colour values below

typical acceptable limits for treated wastewaters leaving wastewater treatment

plants (WWTPs), i.e. 25-75 mg Pt-Co/L [211, 212]. At the same time, no synergy was

found with the addition of O3 (combined O3/UVC/H2O2 vs 10 min O3 pre-treatment

and UVC/H2O2 post-treatment) because O3 would attack specific organic

compounds compared to UVC and H2O2. That is, molecular O3 would attack

selectively specific compounds (either as pre-treatment or combined with

UVC/H2O2) and would not increase significantly the production of •OH when

combined with UVC/H2O2. The pH remained below 8.5 during the combined

process (initial pH of 7.8) without significantly increasing O 3 decomposition into

•OH [223], while the combination of O3 and H2O2 produce •OH at a slower rate

compared to UVC/H2O2 [154]. The majority of •OH production and DOC removal
would therefore come from the combination of UVC/H2O2, as stated earlier. At the

same time, Figure 6.4b shows that a better use of the injected O3 was made during

the O3/UVC/H2O2 process at the lowest injected O3 inlet concentration (5.6 mg/Lgas),

reducing O3 consumption from 273 mg O3/Lwastewater (8.1 mg O3/Lgas injected) to

189 mg O3/Lwastewater and thus, making the process more efficient in terms of O 3

usage. Following the trend of O3 consumption (injected minus exhaust O3), one

could expect a similar behaviour down to an injected O3 concentration of

approximately 3 to 4 mg/Lgas. This would be translated into an increase in O3

consumption percentage and thus, a more efficient use of O 3 supplied to the

system. Colour and SUVA254 variation also showed a similar trend during the four

scenarios under study (Figure B 3 in Appendix B).

126
(a)
1.2

DOC/DOC0 0.8

0.6

0.4

0.2

0
0 20 40 60 80 100 120 140 160 180
Time [min]

(b)
1.0
Consumed O3/Consumed O3 initial

0.8

0.6

0.4

0.2

0.0
0 20 40 60 80 100 120 140 160 180
Time [min]

Figure 6.4. (a) Relative DOC concentration as a function of time. (b) Relative consumed

O3 (injected O3 minus exhaust O3) as a function of time (Consumed O3/Consumed O3 time

zero). (◊) 5.6 mg O3/Lgas injected for combined O3,UVC and H2O2; (□) 8.1 mg O3/Lgas injected

for combined O3,UVC and H2O2; (∆) 12.1 mg O3/Lgas injected for combined O3,UVC and

H2O2; (○) 10 min O3 (100 mg O3/Lgas injected) followed by combined UVC and H2O2.

490 mg H2O2/L single initial addition for all experiments.


Conclusions

O3 pre-treatment followed by a UVC/H2O2 process proved efficient in reducing

DOC values for real slaughterhouse wastewater. This way, with 10 min O 3 pre-

treatment and 3 h UVC/H2O2 at a constant concentration of 300 mg H2O2/L, DOC

reduction greater than 40% was obtained. The removal of colour, turbidity and TSS

during O3 pre-treatment led to an improved performance of the UVC/H2O2 process

due to an increase in UVC light transmissibility. A similar DOC removal

percentage was achieved with a single initial addition of 490 mg H2O2/L, showing

no need to keep H2O2 concentration constant. Additionally, residual H2O2

concentration was reduced to half. O3 was also combined simultaneously with

UVC/H2O2 during 3 h, obtaining again a DOC removal above 40%. In this case,

injected O3 concentration was reduced first to 8.1 and later to 5.6 mg O 3/Lgas,

showing in both cases around 42% DOC removal. This was translated into a final

DOC value of 20.5 mg/L after 3 h of treatment. The main synergy in all the

configurations (O3 with UVC/H2O2, either as a pre-treatment or combined) came

from the combination of UVC with H2O2 and is attributed to •OH attack, showing

no significant difference for O3 when used as a pre-treatment or combined. A better

use of O3, however, was made when 5.6 mg O3/Lgas were injected in combination

with UVC/H2O2, making overall the process more efficient. Direct discharge limits

were met for COD (61 mg O2/L) and TSS (9 mg/L), achieving colour values

(20 mg Pt- Co/L) below typical acceptable colour limit values for treated

wastewaters leaving WWTPs. Therefore, high process efficiency and high

treatment quality is reported in the present study for real slaughterhouse

wastewater. This way, this process combination gives a direction towards

advancing AOPs application in slaughterhouse treatment industry, with

applications for direct discharge and great potential for water reuse purposes.

128
Chapter 7

Single and combined electrochemical

driven processes for the treatment of

slaughterhouse wastewater

Introduction

The bio-treated effluent was further processed in the lab by an electrochemical

cell individually and combined with UVC and/or H2O2. The performance of the

different combinations and possible synergies were assessed measuring DOC

abatement capacity, colour removal and monitoring produced hazardous by-

porducts. An O3 pre-treatment was also applied prior to EAOPs with the aim of

increasing the performance of the different processes by producing more easily

oxidisable compounds and increasing light transmissibility.


Results and discussion

7.2.1 Slaughterhouse wastewater characterisation

Similar to the wastewater reported in Chapter 6, organic matter, TSS and colour

values shown in Table 7.1 exceed the emission limit values imposed by the

European legislation for direct discharge from urban WWTPs [8].

Table 7.1. Physico-chemical characteristics of the collected slaughterhouse

wastewater.

Collected Ozonated
Parameter (units) slaughterhouse slaughterhouse
wastewater wastewater
pH 7.5 7.9
DOC (mg/L) 31 31
COD (mg O2/L) 241 177
TSS (mg/L) 39 21
Turbidity (NTU) 16 n.d.
Colour (visual aspect) Dark brown Transparent
Colour (mg Pt-Co/L) 230 37
Conductivity (µS/cm) 5451 5424
Cl (mg/L)
- 968 943
NO (mg/L)
3- 200 58
NO2- (mg/L) 79 < 0.01
NH4+ (mg/L) 96 199
PO (mg/L)
43- 82 47
SO (mg/L)
42- 50 53
Na (mg/L)
+ 767 780
K (mg/L)
+ 84 83
Ca (mg/L)
2+ 28 28
Mg (mg/L)
2+ 11 11

130
7.2.2 Effects of EAOPs on DOC and produced oxidants

Figure 7.1a shows DOC removal as a function of the specific charge (Q) during

the treatment of slaughterhouse wastewater by EO, EO/H2O2, EO/UVC and

EO/UVC/H2O2 processes at a current density of 100 mA/cm2. The EO process was

able to remove DOC, although poorly. This is attributed to the low extent of

organics oxidation by direct electron transfer to the anode surface, and the

production of low amounts of the following agents and/or a small participation of

these agents on the organics oxidation: (i) physisorbed •OH at the BDD anode

surface, generated by water oxidation via Eq. (19), and (ii) active chlorine species,

such as hypochlorous acid (HClO), hypochlorite (ClO-) and chlorine (Cl2),

generated from the oxidation of chloride (Cl-) ions (Cl- content of the effluent of

968 mg/L) at the anode via Eqs. (20-21) [224]. At the pH registered during reactions

(pH 7 to 8), the dominant active chlorine species is HClO, and ClO- is also present.

HClO can be electrochemically converted into chlorate ion (ClO3-) via Eq. (23).

ClO3- and Cl- can be oxidised to perchlorate (ClO4-) via Eqs. (24, 32-35). ClO3- and

ClO4- are very poor oxidants and hazardous for human health [225, 226].
(a)
1.2

1.0

0.8
DOC/DOC0

0.6

0.4

0.2

0.0
0 2 4 6 8
Q [Ah/L]
(b)
0.5

0.4
Oxidants [mM]

0.3

0.2

0.1

0.0
0 2 4 6 8
Q [Ah/L]
Figure 7.1. (a) Normalised DOC decay and (b) concentration of free oxidants as a

function of Q for the treatment of slaughterhouse wastewater by the following EAOPs: ()

EO, () EO/H2O2, () EO/UVC and () EO/UVC/H2O2. Operating conditions: constant

current density of 100 mA/cm2, solution temperature of 25 °C, initial solution pH of 7.5,

initial volume of 1.4 L, initial H2O2 addition of 850 mg/L and 11 W UVC lamp.

132
6HClO + 3H2 O → 2ClO− − + −
3 + 4Cl + 12H + 1.5O2 + 6e (32)

ClO− + •OH→ ClO− +


2 + H + e

(33)

ClO− − + −
2 + •OH → ClO3 + H + e (34)

ClO− − + −
3 + •OH → ClO4 + H + e (35)

There was a perceptible increase in DOC to values above the initial one (values

at time zero) during the first instants of EO process, which can be attributed to the

dissolution of some suspended organic matter by •OH and free oxidants formed

during the treatment process. Organic solids dissolution also occurred during the

application of the other EAOPs, although with a less noticeable DOC increase.

The generation of low amounts of HClO and ClO- was confirmed by the low

amount of oxidants determined by iodometric titration during the EO process

(Figure 7.1b). Note that the iodometric titration method was able to detect oxidants

such as HClO, ClO-, Cl2 , chlorine dioxide (ClO2), persulphate and H2O2 and unable

to determine ClO3- and ClO4- since the pH of the method was not sufficiently acidic

[227, 228].

Table 7.2 displays the pseudo-first order kinetic constants for the DOC removal

for all the experiments included in this study, determined according to Moreira et

al. [229] and explained further in Appendix C.


Table 7.2. kDOC values for all the applied processes along with the corresponding

interval of adjustment, residual variance (S2R) and coefficient of determination (R2).

DOC

Time interval kDOC S2R


Process R2
(min) (10-3 min-1) (mg2/L2)

H2O2 Null DOC removal

UVC Null DOC removal

No fitting of a pseudo-first-order kinetic model to


UVC/H2O2
experimental data

No fitting of a pseudo-first-order kinetic model to


EO
experimental data

EO/H2O2 0, 30-180-480 0.78±0.06 3.3 0.959

EO/UVC 0, 60-480 1.27±0.08 3.2 0.979

EO/UVC/H2O2 0-480 3.5±0.2 19 0.972

10’O3+EO 0-480 1.13±0.04 1.9 0.988

10’O3+EO/UVC 0-480 2.3±0.2 13 0.970

10’O3+EO/UVC/H2O2 0-480 3.5±0.1 3.0 0.994

The addition of H2O2 to EO slightly improved the DOC removal (Figure 7.1a

and Table 7.2). This can be ascribed to the oxidising potential of H2O2 [230]. H2O2

may have degraded some by-products formed during the EO process and/or may

have reacted with electrogenerated ClO3-, favouring its decomposition to ClO2 (Eq.

36), a powerful disinfectant which can contribute to the mineralisation process

[231].

2H+ + 2ClO3- + H2O2 → 2ClO2 + 2H2O + O2 (36)

Figure 7.1b confirms the presence of higher amounts of oxidants in the

EO/H2O2 process compared to the EO process. Note that H2O2 was unable to

oxidise the original organics of the slaughterhouse wastewater since a null DOC

134
decay was achieved when adding H2O2 alone to the slaughterhouse wastewater

(Figure B 2 in Appendix B).

The supply of UVC light during EO was more beneficial than the addition of

H2O2 in regards to the mineralisation extent (Figure 7.1a and Table 7.2). This can

be attributed to the homolysis of the generated HClO according to Eq. (37), with

the consequent production of extra •OH and chlorine radicals (Cl●) [232].

Furthermore, UVC light may have been able to degrade some organic by-products.

Note that a blank experiment revealed that UVC light alone was unable to degrade

the organics available in the slaughterhouse effluent matrix (Figure B 2 in

Appendix B).

HClO + hv →•OH+ Cl● (37)

The amount of total oxidants in the EO/UVC and EO processes was quite

similar (Figure 7.1b). HClO is an intermediate compound and so the occurrence of

HClO homolysis may have not affected expressively the total amount of oxidants

in the EO/UVC process.

The coupling of EO, UVC radiation and H2O2 oxidant led to the highest organic

matter removal (Figure 7.1a and Table 7.2). This can be ascribed to the homolysis

of H2O2 under UVC irradiation via Eq. (18) [133], resulting in the generation of high

amounts of •OH.

The total amount of oxidants was quite similar for EO/UVC/H2O2 and EO/H2O2

processes (Figure 7.1b) and can be attributed mainly to H2O2 and ClO2 oxidants.

The H2O2 was more rapidly degraded in the EO/UVC/H2O2 process (Figure C 1 in

Appendix C) due to the homolysis of H2O2 under UVC light.


Figure 7.2 shows the contents of ClO3- and ClO4- determined by ion

chromatography. ClO4- was present in high amount, much higher than ClO3-, since

ClO4- is a final product of Cl- oxidation and ClO3- follows a typical trend of

intermediate compound during the electrochemical processes.

12

10
Concentration [mM]

0
ClO3-
ClO3- ClO
ClO4--
4

Figure 7.2. Concentration of ClO3- and ClO4- at the reaction end for the treatment of the

slaughterhouse wastewater by the following EAOPs: () EO, () EO/UVC, () EO/H2O2

and ( ) EO/UVC/H2O2. Operating conditions: constant current density of 100 mA/cm2,

solution temperature of 25 °C, initial solution pH of 7.5, initial volume of 1.4 L, initial H2O2

addition of 850 mg/L and 11 W UVC lamp.

The lower amount of ClO4- in the EO/H2O2 process compared to that of the EO

can be attributed to the reaction of H2O2 with HClO, which may have prevented a

portion of HClO to be electrochemically oxidised to ClO 3- via Eq. (32), thereby

avoiding ClO4- production via Eq. (35). During the EO/UVC process, the occurrence

of HClO homolysis may have led to the accumulation of lower amounts of ClO 4-

compared to the EO process since less HClO was available to be converted into

136
ClO3- and ClO4-. Figure 7.2 also reveals a lower accumulation of ClO4- in the

EO/UVC/H2O2 process compared to those observed in the other EAOPs, likely

mainly due to the reaction of active chlorine species with H2O2. These results show

that the coupling of EO, UVC radiation and H2O2 oxidant not only increased the

process efficiency in terms of organics degradation but also decreased the

concentration of undesirable oxidation by-products.

7.2.3 Effects of EAOPs on COD, TSS and colour

After the application of a specific charge of 7.3 Ah/L (480 min of reaction), COD

values were reduced from 241 mg O2/L to < 125 mg O2/L, i.e. the emission limit

value imposed in the EU [8]. For the EO/UVC and EO/UVC/H2O2, COD values as

low as 40 and < 10 mg O2/Lwere found, respectively. As for DOC, COD

degradation was principally due to •OH and free oxidants attack. The amount of

suspended solids also decreased during EAOPs. TSS contents below 15 mg/L were

achieved after applying a specific charge of 7.3 Ah/L for all tested EAOPs.

Regarding colour (Figure 7.3), an abrupt colour decay during the first instants

of reaction for all the AOP processes was observed. This decay corresponded to

⁓60 mg Pt-Co/L for EO and EO/UVC processes and to ⁓150 mg Pt-Co/L for

EO/H2O2 and EO/UVC/H2O2 processes, after 10 min of reaction (0.12 Ah/L). These

results indicate the presence of readily oxidisable coloured organic and/or

inorganic compounds in the effluent matrix, part of them susceptible to oxidation

by H2O2 (either directly or by the action of ClO2 produced from ClO3- in the

presence of H2O2). The coloured organic compounds were converted into non-

coloured organic by-products as pointed by the absence of remarkable DOC

decays simultaneously with these dramatic colour removals. For longer reaction

times, processes could be arranged in the following sequence according to their


ability to remove colour: EO < EO/H2O2 < EO/UVC < EO/UVC/H2O2. This

behaviour was similar to that found in terms of mineralisation. To reach a colour

of 25 mg Pt-Co/L, it took ⁓400 min (6 Ah/L) for the EO/H2O2 process, ⁓260 min

(3.7 Ah/L) for the EO/UVC process and ⁓120 min (1.6 Ah/L) for the EO/UVC/H2O2

process.

1.2

1.0

0.8
Colour/Colour0

0.6

0.4

0.2

0.0
0 2 4 6 8
Q [Ah/L]
Figure 7.3. Colour decay as a function of Q for the treatment of slaughterhouse

wastewater by the following EAOPs: () EO, () EO/H2O2, () EO/UVC and ()

EO/UVC/H2O2. Operating conditions: constant current density of 100 mA/cm2, solution

temperature of 25 °C, initial solution pH of 7.5, initial volume of 1.4 L, initial H2O2 addition

of 850 mg/L and 11 W UVC lamp.

138
7.2.4 Effects of EAOPs in combination with a pre-ozonation step on DOC and

produced oxidants

As a second approach, an ozonation pre-treatment was carried out before the

application of EAOPs. The main aim of the pre-treatment was to reduce colour and

TSS, and assess the influence of these changes on the efficiency of EAOPs. Results

indicated maximum colour (Figure 7.4) and TSS removals were achieved with an

ozonation time of 10 min using an inlet O3 dose of 100 mg O3/Lgas at a flow rate of

0.3 L/min (Figure 6.1b). Therefore, an ozonation pre-treatment under the same

conditions was applied to the slaughterhouse wastewater before the application of

EAOPs. Under these conditions, a transferred O3 dose (ODT) of ⁓120 mg O3/Leffluent

was estimated after the 10 min reaction via Eq. (38). At 10 min reaction, the off-gas

concentration was ⁓68 mg O3/Lgas.

Qg t
ODT = VL
∫0 (CO3 ,I‑g ‑ CO3 ,O‑g ) dt (38)

With CO3,I-g being the constant inlet O3 concentration (100 mg O3/Lgas) and CO3,O- g

being the outlet/off-gas concentration (mg O3/Lgas), Qg being the applied gas flow

rate (L/min) and VL being the volume of slaughterhouse wastewater in the reactor

(L).
Figure 7.4. Bio-treated slaughterhouse wastewater before (a) and after O3 pre-

treatment. 10 min O3 pre-treatment at 100 mg O3/Lgas and a flowrate of 0.3 Lgas/min.

Characteristics of the slaughterhouse wastewater before and after ozonation are

displayed in Table 7.1. Colour was reduced from 230 to 37 mg Pt-Co/L (dark brown

to transparent colour), indicating a high ability of O3 to convert coloured organic

and/or inorganic compounds into non-coloured ones. TSS were reduced from 39

to 21 mg/L, suggesting the dissolution of organics and/or inorganics. Both colour

and TSS were in agreement with legislated/permissible limits for final wastewater

discharge into water bodies. DOC content remained constant, whereas COD,

which takes into account dissolved and suspended organic matter, decreased from

241 to 177 mg O2/L, remaining above the European discharge limit into

waterbodies (Directive no. 91/271/CEE [8]), i.e. 125 mg O2/L. This indicates the

ability of O3 not only to dissolve suspended organics but also to degrade them.

Figure 7.5a compares DOC decay profiles as a function of specific charge

achieved for EO, EO/UVC and EO/UVC/H2O2 processes using the collected

slaughterhouse wastewater with and without pre-ozonation. The ozonation pre-

treatment enhanced the DOC removal in EO and EO/UVC processes. This

improvement was also confirmed by the kDOC values shown in Table 7.2, where for

the EO/UVC process the kDOC was 1.8 times higher when using the pre-ozonated

140
wastewater. These results can be mainly attributed to the presence of readily

oxidisable organic compounds in the ozonated effluent. The improvement of light

transmissibility in the ozonated effluent may have not contributed to a large

enhancement of HClO homolysis via Eq. (28) in the EO/UVC process since the

superiority of the EO/UVC over the EO process in regards to DOC removal was

even slightly higher for the non-ozonated wastewater (35% versus 27% after

480 min of reaction). For the EO/UVC/H2O2 process, the DOC decay was similar

using both wastewaters, and furthermore, the superiority of the EO/UVC/H 2O2

process over the EO/UVC one was much less pronounced for the pre-ozonated

wastewater. That suggests a higher susceptibility of organic compounds in the pre-

ozonated effluent for oxidation, with oxidants produced in the EO/UVC process

being almost enough to maximise the DOC decay.


(a)
1.2

1.0

0.8
DOC/DOC0

0.6

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8
Q [Ah/L]

(b)
1200

1000
Concentration [mg/L]

800

600

400

200

0
ClO4- (without
ClO3- O3) - (with O )
ClO4ClO4- 3

Figure 7.5. Comparison between EAOPs for the treatment of the slaughterhouse

wastewater with and without pre-ozonation in terms of (a) normalised DOC decay as a

function of Q after O3 pre-treatment and (b) concentration of ClO4- at the reaction end.

Processes: (, ) EO, (, ) 10’O3+EO, (, ) EO/UVC, (, ), 10’O3+EO/UVC, (, )

EO/UVC/H2O2 and (, ) 10’O3+EO/UVC/H2O2. Operating conditions: constant current

density of 100 mA/cm2, solution temperature of 25 °C, initial solution pH of 7.5, initial

volume of 1.4 L, initial H2O2 addition of 850 mg/L and 11 W UVC lamp.

142
Figure 7.5b reveals the accumulation of much lower contents of ClO 4- for

EAOPs carried out with the ozonated wastewater. The differences observed in

chlorine speciation could be related to the interaction between O3 and chlorine

species. Hypochlorite can react with O3 favouring the production of chloride or

chlorine dioxide (Eq. 39-40) [233, 234]. Hence, it is possible to avoid the potential

promotion of hypochlorite to chlorate and perchlorate during the electrochemical

treatment due to a competitive reaction between the electro-oxidation of

hypochlorite and its reaction with O3.

O3+ClO- → 2O2 + Cl- (39)

O3 + ClO- → O2 + ClO2- (40)

This means that O3 used as a pre-treatment for slaughterhouse wastewater can

limit the formation of undesirable by-products during the electrochemical

processes. The amount of total oxidants along the various EAOPs for the pre-

ozonated wastewater (Figure 7.6) was quite similar to that of the wastewater

without pre-ozonation (Figure 7.1b), indicating no influence of the wastewater

matrix on these oxidants generation.


0.5

Oxidants [mM] 0.4

0.3

0.2

0.1

0.0
0 2 4 6 8
Q [Ah/L]
Figure 7.6. Concentration of free oxidants as a function of Q for the treatment of the

slaughterhouse wastewater with pre-ozonation by the following EAOPs: () 10’O3+EO,

() 10’O3+EO/UVC and () 10’O3+EO/UVC/H2O2. Operating conditions: constant current

density of 100 mA/cm2, solution temperature of 25 °C, initial solution pH of 7.5, initial

volume of 1.4 L, initial H2O2 addition of 850 mg/L and 11 W UVC lamp.

7.2.5 Effects of EAOPs in combination with a pre-ozonation step on nitrogen

compounds

Nitrate (NO3-) and ammonium (NH4+) ions were monitored during EAOPs

carried out with the slaughterhouse wastewater with and without pre-ozonation

(Figure 7.7). During O3 pre-treatment, as well as within the first minutes of

electrolysis with no ozonation pre-treatment, the concentration of NO3- rapidly

decreased (Figure 7.7a) and was probably reduced to NH4+ (Figure 7.7b). During

electrochemical oxidation, this behaviour can be explained by the electrochemical

reduction of NO3- to NH4+ over the cathode surface (Eqs.41-42). Afterwards, a linear

increase was registered in the concentration of NO3-, which can be related to the

144
release of nitrogen to the wastewater from the oxidation of the organic matter

(Eqs. 43-45) [235].

(a)
5

4
O3 pre-treatment

3
NO3- [mM]

0
-1 0 1 2 3 4 5 6 7 8
Q [Ah/L]

(b)
14

12

10
NH4+ [mM]

4
O3 pre-treatment
2

0
-1 0 1 2 3 4 5 6 7 8
Q [Ah/L]

Figure 7.7. (a) NO3- and (b) NH4+ concentration as a function of Q for the treatment of

the slaughterhouse wastewater by the following EAOPs: () EO, () 10’O3+EO, ()

EO/UVC, () 10’O3+EO/UVC, () EO/UVC/H2O2 and () 10’O3+EO/UVC/H2O2.

Operating conditions: constant current density of 100 mA/cm2, solution temperature of

25 °C, initial solution pH of 7.5, initial volume of 1.4 L, initial H2O2 addition of 850 mg/L

and 11 W UVC lamp.


NO− −
3 + 6H2 O + 8e ⇄ NH3 + 9OH

(41)

NH3 + H2 O ⇄ NH+4 + OH− (42)

N2 + 2O2 + 2e− → 2NO−


2 (43)

NO− + −
2 + 2H ⇄ 2NO + NO3 + H2 O (44)
1
NO−
2 + O2 → NO−
3 (45)
2

On the other hand, if no pre-treatment was applied, NH4+ concentration showed

an initial increase, which was previously attributed to the electroreduction of NO3- ,

followed by a decrease. To explain this decrease, it is important to highlight the

influence of chlorine species on nitrogen speciation. Specifically, the

electrogenerated ClO- can react with NH4+, favouring the production of

chloramines (Eqs. 46-48) [236, 237]. These species present an oxidant capacity that

can also contribute to the mineralisation of the organic matter present in effluents.

NH+4 + ClO− → NH2 Cl + H2 O (46)

NH2 Cl + ClO− → NHCl2 + OH− (47)

NHCl2 + ClO− → NCl3 + OH− (48)

It is worth mentioning that the total amount of nitrogen was higher than the

European discharge limit into waterbodies, i.e. 10 mg/L (Directive no. 91/271/CEE

[8]), for the slaughterhouse wastewater with and without pre-ozonation (Table 7.1).

NH4+, NO2- and NO3- species contributed to around 160 mg/L of nitrogen for both

wastewaters and additional nitrogen could come from organic matter and

undissolved compounds. This problem can be solved by enhancing the biological

stage.

146
Conclusions

The applied EAOPs were able to remove recalcitrant organics, suspended

solids and colour from the pre-treated slaughterhouse wastewater. EAOPs could

be arranged in the following sequence in regards to their efficiency for

mineralisation: EO < EO/H2O2 < EO/UVC < EO/UVC/H2O2. After 480 min of

reaction (7.3 Ah/L of specific charge), COD values below the emission limit

imposed in the European Union, i.e. 125 mg O2/L, were found for all EAOPs.

Suspended solids dissolution occurred during EAOPs. A TSS value below 35 mg/L,

i.e. the European TSS emission limit, was also achieved for all the EAOPs. A

portion of the coloured compounds was easily removed at the first reaction

instants for all the EAOPs, but some coloured matter was more persistent. For long

reaction times, EAOPs could be ordered in the same sequence in terms of their

ability for colour removal as for mineralisation. To reach a colour of

25 mg Pt- Co/L, it took more than 480 min for the EO process, ⁓400 min for the

EO/H2O2 process, ⁓260 min for the EO/UVC process and ⁓120 min for the

EO/UVC/H2O2 process.

The addition of an ozonation stage prior to EAOPs converted the original

organic compounds into easily oxidisable compounds, which enhanced the ability

of all EAOPs for organic compounds removal. As a result, the superiority of the

EO/UVC/H2O2 process over the EO/UVC one was not as evident for the pre-

ozonated slaughterhouse wastewater as for the non-ozonated wastewater. Beyond

that, ozonation itself was able to reduce suspended solids and colour to below the

legislated/permissible limits for treated wastewaters leaving WWTPs. Regarding

the production of hazardous compounds such as ClO4-, the pre-ozonation step was

beneficial.
Finally, future research needs to be carried out focusing on the design of robust

systems that allow to combine electrochemical advanced oxidation processes to

avoid the formation of undesirable by-products in slaughterhouse wastewater. The

enhancement of the biological treatment stage should be carried out in order to

provide effluent nitrification/denitrification, thereby allowing to comply with the

total nitrogen discharge limit.

148
Chapter 8

Conclusions

Introduction

The aims and objectives proposed in the present study have been successfully

addressed by implementing different single and combined lab scale AOPs for the

treatment of slaughterhouse wastewater (Figure 8.1). During the experiments

carried out, the efficiency of the processes under study has been evaluated by

measuring different organic matter indicators (DOC, COD and BOD), inorganics

(P, TN, anions, cations, etc.), extent of disinfection (TC and TVC), applied powers

and amount of injected oxidising agents (O3 and H2O2), meeting direct water

dischage limits and showing a great potential to be used in water reuse

applications. This way, a thorough study on slaughterhouse wastewater treatment

by AOPs has been incorporated to the literature, opening the door to the

application of AOPs in the meat industry for effluent treatment.


Figure 8.1. Schematic of the processes used for the treatment of slaughterhouse

wastewater.

Perfomance of the applied processes

The experimental results obtained in the AFO system (Chapter 4) highlighted

the importance of using a biological process (ASP in particular) prior to applying

AOPs (O3 in this case) to wastewaters with a high organic load. When an ASP is

efficiently designed for organic matter and nutrient removal, O3 would work as a

polishing step without the need of applying a filtration system prior to O 3

application. O3, however, showed a great potential for microbial inactivation and

met drinking water standards for TC and TVC within 30 min. Ozonation results

also highligthed there exists a maximum O3 injetion time (30 min) where no further

TVC inactivation would be achieved exceeding that limit. Thus, the AFO process

reduced COD, BOD and TSS, by 93%, 98% and 99%, respectively, with a complete

inactivation of TC and 5 log reduction in TVC.

150
US, on the contrary, was not efficient in reducing organic carbon and microbial

indicators, obtaining no removal with 44 and 1000 kHz and 18% reduction in COD

with 300 kHz after 1 h of treatment. The combination of US and O3, however,

reduced significantly the parameters under study. The performance of US in the

combined system seems to be inversely related to the applied O3 dose. In this case,

a high concentration of O3 was injected and thus, the addition of US would not

increase substantially O3 mass tranfer to the liquid bulk and •OH production, as

highligthed in section 5.2. Nonetheless, the coupled system was the only treatment

method, compared to US and O3 alone, able to reach direct discharge limits (COD,

BOD and TSS) and drinking water standards (TC and TVC).

O3 was also employed as a pre-treatment method prior to UVC/H2O2, proving

effective in the removal of colour, turbidity and TSS and hence, increasing

considerably the performance of the subsequent treatment. For slaugherhouse

wastewater, the individual application of UVC and H2O2 showed no DOC removal,

while the combination of the two exhibited a large synergy. It is also important to

underline the optimisation of H2O2 concentration and injection system for an

efficient process. The simultaneous application of O3, UVC and H2O2 did not

increase DOC reduction, obtaining similar DOC removal percentages (⁓ 40%)

when O3 was used as a pre-treatment and combined. As with other coupled AOPs,

the O3 combined UVC/H2O2 process, either as a pre-treatment or administered

simultaneously, met direct discharge limits for COD (61 mg O 2/L) and TSS

(9 mg/L), achieving also colour values (20 mg Pt-Co/L) below typical acceptable

colour limit values for treated wastewaters leaving WWTPs .

The application of EAOPs has also been presented in this study, being able to

remove recalcitrant organics, suspended solids and colour after a biological

treatment. EO alone, however, did not show high DOC removal percentages,
highlighting the importance of combining different AOPs. Thus, EO/UVC system

led to an increase in the production of •OH and chlorine radicals, notably

increasing the efficiency of the process. The highest DOC removal (above 80%) was

obtained by the EO/UVC/H2O2 process, increasing the production of •OH by the

cleavage of H2O2 molecules under UVC irradiation. Colour values of 25 mg Pt-Co/L

were achieved with the two combined systems.

As it happened with the study on UVC/H2O2, a 10 min pre-ozonation was also

applied preciding EAOPs. In this case, the pre-treatment increased the

performance of the EO and EO/UVC principally due to the formation of readily

oxidisable organic compounds. The efficiency of the EO/UVC/H2O2 process,

however, was not improved, although the accumulation of ClO4- was reduced with

pre-ozonated samples for all the EAOPs under study. O3 pre-treatment could limit

the formation of undesirable by-products during electrochemical processes, as

stated in section 7.2.4.

Optimisation

It is always important to know what kind of wastewater one is dealing with in

order to choose best treatment method (type and combination of AOPs); i.e.

wastewaters with chloride (Cl-) or sulfate (SO42-) ions to enhance the performance

of an electrochemical process, light colour and low turbid waters when applying

UV, type of US reactor for organic matter removal or microbial inactivation, etc.

One important parameter to be optimised, either during the processes presented

in this study or in any other, is the O3 dose (inlet concentration and injection

flowrate). It has been shown that O3 injection dose and time play an important role

in the optimisation of the process, either alone or in combination with other AOPs.

Besides improving the efficiency of the process in terms of organic matter removal

152
and microbial inactivation (depends on the effluent quality to be met), optimising

the use of O3 would significantly reduce treatment costs when applied in real

treatment plants. The transfer efficiency of O3 into the bulk of the solution, an

important parameter affecting the efficiency of the process, could be also increased

by improving reactor design or injection system. When O3 is used with other AOPs,

injection place could also improve treatment performance, although Chapter 6

shows no significant difference in DOC removal between O3 pre-treatment and

combined.

Similarly, the concentration of H2O2 (related to the water quality we are dealing

with) and the injection system (continuous, intermittent or single initial addition)

would be another two parameters to take into account to increase treatment plant

efficiency and reduce costs. When combined with UVC, low H2O2 concentrations

would not produce •OH in sufficient quantity, while an excess of the oxidant

would act as radical scavenger.

Future prospects

Several papers have been published on the use of different AOPs, targeting

most of them specific pollutants in synthetic solutions. Future experimental studies

with real samples are encouraged in the present work, having already shown

differences in removal percentages and possibly degradation behaviours between

synthetic and real waters. The importance of designing AOP systems with

upscaling possibilities (looking for a practical use of new treatment systems) needs

to be highlighted, as well as research articles on real wastewaters at pilot or real

scale. With that, accurate and reliable assessment of the economical viability of

AOPs could be addressed and hence, move forward in the application of AOPs in

real industries. This way, new and more stringent regulations (i.e. a common EU
regulation on water to be used for irrigation, lowering of current parameter limits,

new limits for trihalomethane formation and emerging contaminants, etc.) coming

sooner or later into force could be met in a more efficient way and hence, tackle to

some extent water supply issues worldwide population is currently facing.

154
Appendix A

A 20 kHz ultrasonic horn (ultrasonic processor composed of a ½ inch

titanium alloy ultrasonic horn attached to a CL-334 converter and connected to a

20 kHz generator, Fisher Scientific, with 700 W max power output) was employed

to treat two different sample volumes (50 and 100 mL) of slaughterhouse

wastewater (Table 5.1) for 30 seconds at two different applied powers (210 W for

30% amplitude and 350 W for 50% amplitude). No significant difference was

measured in COD, BOD and TSS (Figure A 1 and Figure A 2), as well as in TC

(Figure A 3 and Figure A 4) and TVC (Figure A 5 and Figure A 6) before and after

the treatment.
1000
COD BOD TSS
900

800

700
Concentration [mg/L]

600

500

400

300

200

100

0
Activated sludge 30 sec sonication (50 mL) 30 sec sonication (100 mL)

Figure A 1. COD, BOD and TSS values as a function of sonication time at 20 kHz and

210 W applied power (30% amplitude). Error bars expressed as standard deviation.

1000
COD BOD TSS
900

800
Concentration [mg/L]

700

600

500

400

300

200

100

0
No sonication 30 sec (50 mL sample) 30 sec (100 mL sample)

Figure A 2. COD, BOD and TSS values as a function of sonication time at 20 kHz and

350 W applied power (50% amplitude). Error bars expressed as standard deviation.

156
3E+03
50 mL sample 100 mL sample

2E+03
Total coliforms [CFU/mL]

2E+03

1E+03

5E+02

0E+00
Activated sludge 5 sec sonication 10 sec sonication 30 sec sonication

Figure A 3. TC survival values as a function of sonication time at 20 kHz and 210 W

applied power (30% amplitude). Error bars expressed as standard deviation.

3E+03
50 mL sample 100 mL sample

3E+03
Total coliforms [CFU/mL]

2E+03

2E+03

1E+03

5E+02

0E+00
Activated sludge 5 sec sonication 10 sec sonication 30 sec sonication

Figure A 4. TC survival values as a function of sonication time at 20 kHz and 350 W

applied power (30% amplitude). Error bars expressed as standard deviation.


1E+08
50 mL sample 100 mL sample

1E+08

1E+08
Total viable counts [CFU/mL]

8E+07

6E+07

4E+07

2E+07

0E+00
Activated sludge 5 sec sonication 30 sec sonication

Figure A 5. TVC survival values as a function of sonication time at 20 kHz and 210 W

applied power (30% amplitude). Error bars expressed as standard deviation.

1E+08
50 mL sample 100 mL sample

1E+08

1E+08
Total viable counts [CFU/mL]

8E+07

6E+07

4E+07

2E+07

0E+00
Activated sludge 5 sec sonication 30 sec sonication

Figure A 6. TVC survival values as a function of sonication time at 20 kHz and 350 W

applied power (30% amplitude). Error bars expressed as standard deviation.

158
The absorbance of the I3- ion during KI dosimetry can be converted into

concentration following the Beer Lambert law [86], as expressed by Eq. (49).

𝐴 =ɛ×𝑐×𝑊 (49)

Where A is the absorbance value obtained from the spectrophotometer

(355 nm), ɛ accounts for I3- absorptivity (26303 dm3/mol cm [88]), c is the

concentration of I3-and W is the width of the cuvette used in the spectrophotometer.


Appendix B

The wastewater was collected from a pig slaughterhouse and biologically

treated prior to O3, UVC and H2O2 application. DOC, colour and SUVA254 reduction

are shown below as a function of time for the processes under study. SUVA 254 is

the specific ultraviolet absorbance at 254 nm where the aromatic nature of the

solution is normalised over the total organic load [238].

160
(a)
600

500

400
H2O2 [mg/L]

300

200

100

0
0 20 40 60 80 100 120 140 160 180 200 220 240
Time [min]

(b)
1.2

1.0

0.8
DOC/DOC0

0.6

0.4

0.2

0.0
0 20 40 60 80 100 120 140 160 180 200 220 240
Time [min]

Figure B 1. (a) H2O2 concentration as a function of time. (b) Relative DOC concentration

as a function of time. (□) 10 min O3 + UVC/H2O2 490 mg H2O2/L initial addition.


1.2

1.0

0.8
DOC/DOC0

0.6

0.4

0.2

0.0
0 20 40 60 80 100 120 140 160 180
Time [min]

Figure B 2. Relative DOC concentration as a function of time. No ozonation pre-

treatment. (■) UVC alone; (●) H2O2 alone, 300 mg H2O2/L.

162
(a)
300

250
Colour [Pt/Co]

200

150

100

50

0
0 20 40 60 80 100 120 140 160 180 200
Time [min]

(b)
3.0

2.5

2.0
SUVA254

1.5

1.0

0.5

0.0
0 20 40 60 80 100 120 140 160 180 200
Time [min]
Figure B 3. (a) Colour and (b) SUVA254 as a function of time. (◊) 5.6 mg O3/Lgas injected

for combined O3, UVC and H2O2; (□) 8.1 mg O3/Lgas injected for combined O3, UVC and

H2O2; (∆) 12.1 mg O3/Lgas injected for combined O3, UVC and H2O2; (○) 10 min O3

(100 mg O3/Lgas injected) followed by combined UVC and H2O2. 490 mg H2O2/L single

initial addition for all experiments.


Appendix C

Electrochemical oxidation was employed individually and combined to UVC

and H2O2 for the treatment of real slaughterhouse wastewater. The specific electric

charge (Q) applied to an electrochemical cell is a measure of the efficiency of an

electro-oxidation processes [239] and its mathematical equation reads as follows

(Eq. 50):

𝐼×𝑡
𝑄= (50)
𝑣

Where I is applied intensity in ampere, t is treatment time in hours, v is effluent

volume in litres and Q is measured in Ah/L.

164
A pseudo-first-order kinetic model was fitted to the DOC data as a simple

mathematical model to quantitatively compare the processes efficiency (Table 7.2).

The kinetic model was adjusted by a nonlinear regression method using Fig.P

software for Windows from Biosoft. The pseudo-first-order kinetic constants for

DOC removal (kDOC), in min−1, were calculated via Eq. (51):

[DOC]t = [DOC]0 × e-kDOC × t (51)

Where [DOC]t is the DOC content after time t and [DOC]0 is the DOC content just

before the reaction begins (time zero).

The fitting was performed by minimising the sum of the squared deviations

between experimental and predicted values. The fitting was assessed by

calculating the relative standard deviations, the coefficient of determination (R2)

and the residual variance (S2R).


Different H2O2 consumption rates are shown in Figure C 1 for EO/ H2O2 and

EO/UVC/H2O2 processes during the treatment of slaughterhouse wastewater.

900
800
700
600
H2O2 [mg/L]

500
400
300
200
100
0
0 100 200 300 400 500
Time [min]
Figure C 1. Concentration of H2O2 as a function of time for the treatment of

slaughterhouse wastewater at an specific charge of 7.3 Ah/L by (◊) EO/H2O2 and (Δ)

EO/H2O2/UVC (11W).

166
REFERENCES

[1] S.B. Martínez, J. Pérez-Parra, R. Suay, Use of Ozone in Wastewater Treatment

to Produce Water Suitable for Irrigation, Water Resources Management, 25 (2011)

2109-2124.

[2] D. H.F. Liu, B.G. Liptak, Wastewater treatment, Boca Raton: Lewis Publishers,

London, 2000.

[3] Environmental Protection Agency, Wastewater treatment manuals. Primary,

secondary and tertiary treatment. Ardcavan, Wexford, Ireland, 1997.

[4] N.F. Gray, Water Technology: An Introduction for Environmental Scientists

and Engineers, in: Arnold (Ed.), London, 1999, pp. 548.

[5] Y. Deng, R. Zhao, Advanced Oxidation Processes (AOPs) in Wastewater

Treatment, Current Pollution Reports, 1 (2015) 167-176.

[6] M. Mohajerani, M. Mehrvar, F. Ein-Mozaffari, An overview of the integration

of advanced oxidation technologies and other processes for water and wastewater

treatment, International Journal of Engineering, 3 (2009) 120-146.

[7] M.L. Janex, P. Xu, P. Savoye, J.M. Laine, V. Lazarova, Ozonation as a wastewater

disinfection process to meet reuse regulations, Proceedings of the ozone world

congress, International ozone association, 1999, pp. 81-92.

[8] L. Council Directive 91/271/EEC concerning urban waste-water treatment, 30

European Commission, 1991, pp. 40-52.

[9] I. Amec Foster Wheeler Environment & Infrastructure UK Ltd, ACTeon,

IMDEA and NTUA, EU-level instruments on water reuse. European Comission,

Publications Office of the European Union, Luxembourg, 2016.


[10] M.M. Mekonnen, A.Y. Hoekstra, The green, blue and grey water footprint of

crops and derived crop products, UNESCO-IHE, Delft, The Netherlands, Value of

Water, Research Report Series, (2010).

[11] P.W. Gerbens-Leenes, M.M. Mekonnen, A.Y. Hoekstra, The water footprint of

poultry, pork and beef: A comparative study in different countries and production

systems, Water Resources and Industry, 1-2 (2013) 25-36.

[12] K. Valta, T. Kosanovic, D. Malamis, K. Moustakas, M. Loizidou, Overview of

water usage and wastewater management in the food and beverage industry,

Desalination and Water Treatment, 53 (2015) 3335-3347.

[13] M.M. Mekonnen, A.Y. Hoekstra, A global assessment of the water fooprint of

farm animal products, Ecosystems, 15 (2012) 401-415.

[14] I.S. Arvanitoyannis, D. Ladas, Meat waste treatment methods and potential

uses, International Journal of Food Science & Technology, 43 (2008) 543-559.

[15] E. Padilla-Gasca, A. Lopez-Lopez, J. Gallardo-Valdez, Evaluation of Stability

Factors in the Anaerobic Treatment of Slaughterhouse Wastewater, Journal of

Bioremediation & Biodegradation, 02 (2011) 114-119.

[16] C.F. Bustillo-Lecompte, M. Mehrvar, Treatment of an actual slaughterhouse

wastewater by integration of biological and advanced oxidation processes:

Modeling, optimization, and cost-effectiveness analysis, Journal of Environmental

Management, 182 (2016) 651-666.

[17] J. Wu, H. Doan, Disinfection of recycled red-meat-processing wastewater by

ozone, Journal of Chemical Technology & Biotechnology, 80 (2005) 828-833.

[18] WWAP (United Nations World Water Assessment Programme). 2017. The

United Nations World Water Development Report 2017. Wastewater: The

Untapped Resource. Paris, UNESCO.

[19] C.F. Bustillo-Lecompte, M. Mehrvar, Slaughterhouse wastewater

characteristics, treatment, and management in the meat processing industry: A

review on trends and advances, J Environ Manage, 161 (2015) 287-302.

168
[20] USEPA, Effluent Limitations Guidelines and New Source Performance

Standards for the Meat and Poultry Products Point Source Category,

Environmental Protection Agency (EPA): Federal Register, 2004.

[21] C.F. Bustillo-Lecompte, M. Mehrvar, E. Quinones-Bolanos, Cost-effectiveness

analysis of TOC removal from slaughterhouse wastewater using combined

anaerobic-aerobic and UV/H2O2 processes, J Environ Manage, 134 (2014) 145-152.

[22] C.F. Bustillo-Lecompte, M. Mehrvar, E. Quinones-Bolanos, Combined

anaerobic-aerobic and UV/H2O2 processes for the treatment of synthetic

slaughterhouse wastewater, Environmental Science and Health. Part A,

Toxic/hazardous substances & environmetal engineering, 48 (2013) 1122-1135.

[23] W.H. Glaze, J.-W. Kang, D.H. Chapin, The Chemistry of Water Treatment

Processes Involving Ozone, Hydrogen Peroxide and Ultraviolet Radiation, Ozone:

Science & Engineering, 9 (1987) 335-352.

[24] R.d. Pozo, D.O. Ta, H. Dulkadiro?lu, D. Orhon, V. Diez, Biodegradability of

slaughterhouse wastewater with high blood content under anaerobic and aerobic

conditions, Journal of Chemical Technology & Biotechnology, 78 (2003) 384-391.

[25] D.P. Cassidy, E. Belia, Nitrogen and phosphorus removal from an abattoir

wastewater in a SBR with aerobic granular sludge, Water Research, 39 (2005) 4817-

4823.

[26] R. Borja, C. Banks, A. Martin, Influence of the organic volumetric loading rate

on soluble chemical oxygen demand removal in a down-flow fixed-bed reactor

treating abattoir wastewater, Journal of Chemical Technology and Biotechnology,

64 (1995a) 361-366.

[27] R. Borja, C.J. Banks, Z. Wang, Effect of organic loading rate on anaerobic

treatment of slaughterhouse wastewater in a fluidised-bed reactor, Bioresource

Technology, 52 (1995b) 157-162.

[28] C.E.T. Caixeta, M.C. Cammarota, A.M.F. Xavier, Slaughterhouse wastewater

treatment: evaluation of a new three-phase separation system in a UASB reactor,

Bioresource Technology, 81 (2002) 61-69.


[29] W. Fuchs, H. Binder, G. Mavrias, R. Braun, Anaerobic treatment of wastewater

with high organic content using a stirred tank reactor coupled with a membrane

filtration unit, Water Research, 37 (2003) 902-908.

[30] D. Obaja, S. Macé, J. Costa, C. Sans, J. Mata-Alvarez, Nitrification,

denitrification and biological phosphorus removal in piggery wastewater using a

sequencing batch reactor, Bioresource Technology, 87 (2003) 103-111.

[31] M.I. Aguilar, J. Sáez, M. Lloréns, A. Soler, J.F. Ortuño, V. Meseguer, A. Fuentes,

Improvement of coagulation–flocculation process using anionic polyacrylamide as

coagulant aid, Chemosphere, 58 (2005) 47-56.

[32] S.L. Pabón, J.H. Suárez Gélvez, Starting-up operating a full-scale activated

sludge system for slaughterhouse wastewater, Revista Ingeniería e Investigación,

29 (2009) 53-58.

[33] J. Fan, F. Ji, X. Xu, Y. Wang, D. Yan, X. Xu, Q. Chen, J. Xiong, Q. He, Prediction

of the effect of fine grit on the MLVSS/MLSS ratio of activated sludge, Bioresour

Technol, 190 (2015) 51-56.

[34] David H.F. Liu, B.G. Liptak, Wastewater treatment, CRC Press1999.

[35] S.M. Travers, D.A. Lovett, Activated sludge treatment of abattoir wastewater-

II: Influence of dissolved oxygen concentration, Water research, 18 (1984) 435-439.

[36] M. Henze, M.C.M. van Loosdrecht, G.A. Ekama, D. Brdjanovic, Biological

Wastewater Treatment: Principles, Modelling and Design, 2008.

[37] D.A. Lovett, S.M. Travers, K.R. Davey, Activated sludge treatment of abattoir

wastewater-I: Influence of sludge age and feeding pattern, Water Research, 18

(1984) 429-434.

[38] N.Z. Al-Mutairi, Aerobic selectors in slaughterhouse activated sludge systems:

a preliminary investigation, Bioresour Technol, 100 (2009) 50-58.

[39] V. Ferreira, C. Martins, M.O. Pereira, A. Nicolau, Use of an aerobic selector to

overcome filamentous bulking in an activated sludge wastewater treatment plant,

Environ Technol, 35 (2014) 1525-1531.

170
[40] N.Z. Al-Mutairi, F.A. Al-Sharifi, S.B. Al-Shammari, Evaluation study of a

slaughterhouse wastewater treatment plant including contact-assisted activated

sludge and DAF, Desalination, 225 (2008) 167-175.

[41] T.H. Hsiao, J.S. Huang, Y.I. Huang, Process kinetics of an activated-sludge

reactor system treating poultry slaughterhouse wastewater, Environ Technol, 33

(2012) 829-835.

[42] M.R. Johns, Developments in wastewater treatment in the meat processing

industry: a review, Bioresource technology, 54 (1995) 203-216.

[43] M.M. Um, O. Barraud, M. Kerouredan, M. Gaschet, T. Stalder, E. Oswald, C.

Dagot, M.C. Ploy, H. Brugere, D. Bibbal, Comparison of the incidence of

pathogenic and antibiotic-resistant Escherichia coli strains in adult cattle and veal

calf slaughterhouse effluents highlighted different risks for public health, Water

Research, 88 (2016) 30-38.

[44] M. Sezgin, D. Jenkins, D.S. Parker, A Unified Theory of Filamentous Activated

Sludge Bulking, Journal (Water Pollution Control Federation), 50 (1978) 362-381.

[45] J.F. Heddle, Activated sludge treatment of slaughterhouse wastes with protein

recovery, Water research, 13 (1979) 581-584.

[46] C.K. Chen, S.L. Lo, Treatment of slaughterhouse wastewater using an

activated sludge/contact aeration process, Water Science & Technology, 47 (2003)

285-292.

[47] S.C. Ameta, P.B. Punjabi, A. Kumar, R. Ameta, Advanced Oxidation Processes:

Basics and Applications, in: D.G. Rao, R. Senthilkumar, J.A. Byrne, S. Feroz (Eds.)

Wastewater Treatment: Advanced Processes and Technologies, IWA publishing

2012, pp. 46.

[48] S. Cotillas, M.J. Martín de Vidales, J. Llanos, C. Sáez, P. Cañizares, M. Rodrigo,

Electrolytic and electro-irradiated processes with diamond anodes for the

oxidation of persistent pollutants and disinfection of urban treated wastewater,

Journal of Hazardous Materials, 319 (2016) 93-101.


[49] B. Marselli, J. Garcia-Gomez, P.-A. Michaud, M.A. Rodrigo, C. Comminellis,

Electrogeneration of hydoxyl radicals on boron-doped diamond electrodes,

Journal of Electrochemical Society, 150 (2003) D79-D83.

[50] D. Montanaro, R. Lavecchia, E. Petrucci, A.t. Zuorro, UV-assisted

electrochemical degradation of coumarin on boron-doped diamond electrodes,

Chemical Engineering Journal, 323 (2017) 512-519.

[51] C. Qiu, S. Yuan, X. Li, H. Wang, B. Bakheet, S. Komarneni , Y. Wang,

Investigation of the synergistic effects for p-nitrophenol mineralization by a

combined process of ozonation and electrolysis using a boron-doped diamond

anode, Journal of Hazardous Materials, 280 (2014) 644-653.

[52] C.A. Martínez-Huitle, M.A. Rodrigo, I. Sirés, O. Scialdone, Single and Coupled

Electrochemical Processes and Reactors for the Abatement of Organic Water

Pollutants: A Critical review, Chemical Reviews, 115 (2015) 13362-13407.

[53] F.J. Beltran, Ozone Reaction Kinetics for Water and Wastewater Systems,

Lewis Publ CRC Press, Boca raton, FL 2004.

[54] C. Gottschalk, J.A. Libra, A. Sau, Ozonation of water and waste water: A

practical guide to understanding ozone and its applications, Wiley2009.

[55] U. von Gunten, Ozonation of drinking water: Part I. Oxidation kinetics and

product formation, Water Research, 37 (2003) 1443-1467.

[56] R. Gehr, M. Wagner, P. Veerasubramanian, P. Payment, Disinfection efficiency

of peracetic acid, UV and ozone after enhanced primary treatment of municipal

wastewater, Water Research, 37 (2003) 4573-4586.

[57] X. Jin, S. Peldszus, P.M. Huck, Reaction kinetics of selected micropollutants in

ozonation and advanced oxidation processes, Water Research, 46 (2012) 6519-6530.

[58] S.J. Masten, S.H.R. Davies, The use of ozonation to degrade organic

contaminants in wastewaters, Environmental science & technology, 28 (1994) 180-

185.

[59] N.F. Gray, Ozone disinfection, Microbiology of waterborne diseases,

Elsevier/Academic Press2014, pp. 599-615.

172
[60] U. von Gunten, Ozonation of drinking water: Part II. Disinfection and by-

product formation in presence of bromide, iodide or chlorine, Water Research, 37

(2003) 1469-1487.

[61] S. Tripathi, V. Pathak, D.M. Tripathi, B.D. Tripathi, Application of ozone based

treatments of secondary effluents, Bioresource Technology, 102 (2011) 2481-2486.

[62] B. Langlais, B. Legube, H. Beuffe, M. Dore, Study of the nature of the by-

products formed and the risks of toxicity when disinfecting a secondary effluent

with ozone, Water Science & Technology, 25 (1992) 135-143.

[63] A. Ried, J. Mielcke, A. Wieland, The Potential Use of Ozone in Municipal

Wastewater, Ozone: Science & Engineering, 31 (2009) 415-421.

[64] J. Gomes, R. Costa, R.M. Quinta-Ferreira, R.C. Martins, Application of

ozonation for pharmaceuticals and personal care products removal from water, Sci

Total Environ, 586 (2017) 265-283.

[65] I. Oller, S. Malato, J.A. Sanchez-Perez, Combination of Advanced Oxidation

Processes and biological treatments for wastewater decontamination--a review, Sci

Total Environ, 409 (2011) 4141-4166.

[66] O.M. Millamena, Ozone treatment of slaughterhouse and laboratory

wastewaters, Aquacultural Engineering, 11 (1991) 23-31.

[67] A. Roux, Renovation of wastewater for direct re-use in an abattoir,

Department of Chemical and Environmental Engineering, MSc thesis, University

of Pretoria, 1996.

[68] P. Proesmans, R. De Vil, R. Gerards, L. Vriends, Advanced treatment of

industrial wastewaters: combination of biological treatment and ozonation,

Mededelingen-faculteit landbouwkundige en toegepaste biologische

wetenschappen, 62 (1997) 1729-1736.

[69] C. Hodúr, M. Ábel, Z. Kiss, G. Szabó, Z. László, EFFECTS OF OZONATION

ON THE ULTRAFILTRATION OF MEAT INDUSTRY WASTEWATER,

Environmental Engineering & Management Journal (EEMJ), 17 (2018) 267-272.


[70] P. Paraskeva, N.J.D. Graham, Ozonation of municipal wastewater effluents,

Water Environment Research 74 (2002) 569-581.

[71] Y.G. Adewuyi, Sonochemistry: Environmental science and engineering

applications, Industrial & engineering chemistry research, 40 (2001) 4681-4715.

[72] J. Lee, Importance of Sonication and Solution Conditions on the Acoustic

Cavitation Activity, in: M. Ashokkumar (Ed.) Handbook of Ultrasonics and

Sonochemistry, Spring Sicence + Business Media, Singapore 2016, pp. 137-175.

[73] P.R. Gogate, Application of cavitational reactors for water disinfection: current

status and path forward, J Environ Manage, 85 (2007) 801-815.

[74] P. Kanthale, M. Ashokkumar, F. Grieser, Sonoluminescence, sonochemistry

(H2O2 yield) and bubble dynamics: Frequency and power effects, Ultrasonics

Sonochemistry, 15 (2008) 143-150.

[75] R.J. Wood, J. Lee, M.J. Bussemaker, A parametric review of sonochemistry:

Control and augmentation of sonochemical activity in aqueous solutions, Ultrason

Sonochem, 38 (2017) 351-370.

[76] A. Brotchie, F. Grieser, M. Ashokkumar, Effect of Power and Frequency on

Bubble-Size Distributions in Acoustic Cavitation, Physical Review Letters, 102

(2009) 084302.

[77] C. Petrier, A. Francony, Ultrasonic waste-water treatment: incidence of

ultrasonic frequency on the rate of phenol and carbon tetrachloride degradation,

Ultrasonics Sonochemistry, 4 (1997) 295-300.

[78] C. Pétrier, M.-F. Lamy, A. Francony, A. Benahcene, B. David, Sonochemical

degradation of phenol in dilute aqueous solutions: Comparison of the reaction

rates at 20 and 487 kHz, Journal of Physical Chemistry, 98 (1994) 10514-10520.

[79] J. Berlan, F. Trabelsi, H. Delmas, A.M. Wilhelm, J.F. Petrignani, Oxidative

degradation of phenol in aqueous media using ultrasound, Ultrasonics

Sonochemistry, 1 (1993) 97-102.

[80] T. Lesko, A.J. Colussi, M.R. Hoffmann, Sonochemical decomposition of

phenol: evidence for synergistic effect of ozone and ultrasound for the elimination

174
of total organic carbon from water, Environmental Science & Technology, 40 (2006)

6818-6823.

[81] L.H. Thompson, L.K. Doraiswamy, Sonochemistry: Science and Engineering,

Ind. Eng. Chem. Res., 38 (1999) 1215-1249.

[82] E.J. Hart, A. Henglein, Free radical and free atom reactions in the sonolysis of

aqueous Iodide and formate solutions, Journal of Physical Chemistry, 89 (1985)

4342-4347.

[83] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, Photocatalytic production of

H2O2 and organic peroxides in aqueous suspensions of TiO2, ZnO, and desert

sand, Environmental Science & Technology, 22 (1988) 798-806.

[84] A. Weissler, H.W. Cooper, S. Snyder, Chemical effect of ultrasonic waves:

oxidation of potassium iodide solution by carbon tetrachloride, Journal of the

American Chemical Society, 72 (1950) 1769-1775.

[85] K.R. Morison, C.A. Hutchinson, Limitations of the Weissler reaction as a model

reaction for measuring the efficiency of hydrodynamic cavitation, Ultrason

Sonochem, 16 (2009) 176-183.

[86] J.M. Parnis, K.B. Oldham, Beyond the Beer–Lambert law: The dependence of

absorbance on time in photochemistry, Journal of Photochemistry and

Photobiology A: Chemistry, 267 (2013) 6-10.

[87] A. Henglein, Sonochemistry: historical developments and modern aspects,

Ultrasonics, 25 (1987) 6-16.

[88] S. Koda, T. Kimura, T. Kondo, H. Mitome, A standard method to calibrate

sonochemical efficiency of an individual reaction system, Ultrasonics

Sonochemistry, 10 (2003) 149-156.

[89] M. Ashokkumar, D. Sunartio, S. Kentish, R. Mawson, L. Simons, K. Vilkhu, C.

Versteeg, Modification of food ingredients by ultrasound to improve functionality:

A preliminary study on a model system, Innovative Food Science & Emerging

Technologies, 9 (2008) 155-160.


[90] L. Paniwnyk, O. Larpparisudthi, T.J. Mason, Degradation of water pollutants

using ultrasound, 20th international congress on acoustic, ICA 2010Sydney,

Australia, 2010.

[91] J. Lifka, B. Ondruschka, J. Hofmann, The use of ultrasound for the degradation

of pollutants in water: aquasonolysis - A review, Engineering in Life Sciences, 3

(2003) 253-262.

[92] M.A. Beckett, I. Hua, Impact of ultrasonic frequency on aqueous

sonoluminescence and sonochemistry, Journal of Physical Chemistry, A 105 (2001)

3796-3802.

[93] S. Okouchi, O. Nojima, T. Arai, Cavitation-induced degradation of phenol by

ultrasound, Water Science & Technology, 26 (1992) 2053-2056.

[94] V. Naddeo, V. Belgiorno, D. Kassinos, D. Mantzavinos, S. Meric, Ultrasonic

degradation, mineralization and detoxification of diclofenac in water: optimization

of operating parameters, Ultrasonics Sonochemistry, 17 (2010) 179-185.

[95] V. Naddeo, V. Belgiorno, D. Ricco, D. Kassinos, Degradation of diclofenac

during sonolysis, ozonation and their simultaneous application, Ultrasonics

Sonochemistry, 16 (2009) 790-794.

[96] S. Wang, X. Wu, Y. Wang, Q. Li, M. Tao, Removal of organic matter and

ammonia nitrogen from landfill leachate by ultrasound, Ultrasonics

Sonochemistry, 15 (2008) 933-937.

[97] J.M. Joseph, H. Destaillants, H.-M. JHung, M.R. Hoffmann, The sonochemical

degradation of azobenzene and related azo dyes: rate enhancements via Fenton's

reactions, Journal of Physical Chemistry, A 104 (2000).

[98] R.A. Shrestha, T.D. Pham, M. Sillanpaa, Effect of ultrasound on removal of

persistent organic pollutants (POPs) from different types of soils, Journal of

Hazardous Matererials, 170 (2009) 871-875.

[99] S. Manickam, N.b. Zainal Abidin, S. Parthasarathy, I. Alzorqi, E.H. Ng, T.J.

Tiong, R.L. Gomes, A. Ali, Role of H2O2 in the fluctuating patterns of COD

(chemical oxygen demand) during the treatment of palm oil mill effluent (POME)

176
using pilot scale triple frequency ultrasound cavitation reactor, Ultrasonics

Sonochemistry, 21 (2014) 1519-1526.

[100] E.A. Serna-Galvis, J. Silva-Agredo, A.M. Botero-Coy, A. Moncayo-Lasso, F.

Hernández, R.A. Torres-Palma, Effective elimination of fifteen relevant

pharmaceuticals in hospital wastewater from Colombia by combination of a

biological system with a sonochemical process, Science of The Total Environment,

670 (2019) 623-632.

[101] E. Naffrechoux, S. Chanoux, P.J. Suptil, Sonochemical and photochemical

oxidation of organic matter, Ultrasonics Sonochemistry, 7 (2000) 255-259.

[102] E.A. Serna-Galvis, A.M. Botero-Coy, D. Martínez-Pachón, A. Moncayo-Lasso,

M. Ibáñez, F. Hernández, R.A. Torres-Palma, Degradation of seventeen

contaminants of emerging concern in municipal wastewater effluents by

sonochemical advanced oxidation processes, Water Research, 154 (2019) 349-360.

[103] S. Parthasarathy, R.R. Mohammed, C.M. Fong, R.L. Gomes, S. Manickam, A

novel hybrid approach of activated carbon and ultrasound cavitation for the

intensification of palm oil mill effluent (POME) polishing, Journal of Cleaner

Production, 112 (2016) 1218-1226.

[104] N.H. Abdurahman, Y.M. Rosli, N.H. Azhari, The potential of ultrasonic

membrane anaerobic system(UMAS) in treating slaughterhouse wastewater,

Journal of Engineering and Applied Sciences, 11 (2016) 2653-2659.

[105] J.H. Gibson, H. Hon, R. Farnood, I.G. Droppo, P. Seto, Effects of ultrasound

on suspended particles in municipal wastewater, Water Res, 43 (2009) 2251-2259.

[106] I. Hua, J.E. Thompson, Inactivation of E. coli by sonication at discrete

ultrasonic frequencies, Water Research, 34 (2000) 3888-3893.

[107] S. Koda, M. Miyamoto, M. Toma, T. Matsuoka, M. Maebayashi, Inactivation

of Escherichia coli and Streptococcus mutans by ultrasound at 500kHz, Ultrason

Sonochem, 16 (2009) 655-659.


[108] S.S. Phull, A.P. Newman, J.P. Lorimer, B. Pollet, J.T. Mason, The development

and evaulation of ultrasound in the biocidal treatment of water, Ultrasonics

Sonochemistry, 4 (1997) 157-164.

[109] E. Joyce, A. Al-Hashimi, T.J. Mason, Assessing the effect of different

ultrasonic frequencies on bacterial viability using flow cytometry, J Appl

Microbiol, 110 (2011) 862-870.

[110] B.A. Madge, J.N. Jensen, Disinfection of wastewater using 20-kHz ultrasound

unit, Water Environment Research, 74 (2002) 159-169.

[111] A. Antoniadis, I. Poulios, E. Nikolakaki, D. Mantzavinos, Sonochemical

disinfection of municipal wastewater, J Hazard Mater, 146 (2007) 492-495.

[112] V.M. Gómez-López, M.I. Gil, A. Allende, J. Blancke, L. Schouteten, M.V.

Selma, Disinfection Capacity of High-Power Ultrasound Against E. coli O157:H7

in Process Water of the Fresh-Cut Industry, Food and Bioprocess Technology, 7

(2014) 3390-3397.

[113] M.S. Limaye, W.T. Coakley, Clarification of small volume microbial

suspensions in an ultrasonic standing wave, Journal of Applied Microbiology, 84

(1998) 1035-1042.

[114] S.Z. Salleh-Mack, J.S. Roberts, Ultrasound pasteurization: the effects of

temperature, soluble solids, organic acids and pH on the inactivation of

Escherichia coli ATCC 25922, Ultrason Sonochem, 14 (2007) 323-329.

[115] S. Drakopoulou, S. Terzakis, M.S. Fountoulakis, D. Mantzavinos, T. Manios,

Ultrasound-induced inactivation of gram-negative and gram-positive bacteria in

secondary treated municipal wastewater, Ultrason Sonochem, 16 (2009) 629-634.

[116] P. Foladori, B. Laura, A. Gianni, Z. Giuliano, Effects of sonication on bacteria

viability in wastewater treatment plants evaluated by flow cytometry--fecal

indicators, wastewater and activated sludge, Water Res, 41 (2007) 235-243.

[117] I. Tsukamoto, B. Yim, C.E. Stavarache, M. Furuta, K. Hashiba, Y. Maeda,

Inactivation of Saccharomyces cerevisiae by ultrasonic irradiation, Ultrasonics

Sonochemistry, 11 (2004) 61-65.

178
[118] N.C. Sesal, O. Kekeç, Inactivation of Escherichia coli and Staphylococcus

aureus by ultrasound, J Ultrasound Med, 33 (2014) 1663-1668.

[119] U. Neis, T. Blume, Ultrasonic disinfection of wastewater effluents for high-

quality reuse, IWA Regional symposium on water recycling in Mediterranean

regionIraklio, Greece, 2002.

[120] T. Blume, U. Neis, Improved wastewater disinfection by ultrasonic pre-

treatment, Ultrasonics Sonochemistry, 11 (2004) 333-336.

[121] O. Legrini, E. Oliveiros, A.M. Braun, Photochemical processes for water

treatment, Chemical Reviews, 93 (1993) 671-698.

[122] S. Parsons, Advanced Oxidation Processes for Water and Wastewater

Treatment, IWA Publishing, London, 2004.

[123] A.M. Nienow, J.C. Bezares-Cruz, I.C. Poyer, I. Hua, C.T. Jafvert, Hydrogen

peroxide-assisted UV photodegradation of Lindane, Chemosphere, 72 (2008) 1700-

1705.

[124] B. Wu, M. Yang, R. Yin, S. Zhang, Applicability of light sources and the inner

filter effect in UV/acetylacetone and UV/H2O2 processes, Journal of Hazardous

Materials, 335 (2017) 100-107.

[125] E. Friedler, Y. Gilboa, Performance of UV disinfection and the microbial

quality of greywater effluent along a reuse system for toilet flushing, Sci Total

Environ, 408 (2010) 2109-2117.

[126] C.P. Gerba, D.M. Gramos, N. Nwachuku, Comparative inactivation of

enteroviruses and adenovirus 2 by UV light, Appl Environ Microbiol, 68 (2002)

5167-5169.

[127] W.A. Hijnen, E.F. Beerendonk, G.J. Medema, Inactivation credit of UV

radiation for viruses, bacteria and protozoan (oo)cysts in water: a review, Water

Res, 40 (2006) 3-22.

[128] S. Jiao, S. Zheng, D. Yin, L. Wang, L. Chen, Aqueous photolysis of tetracycline

and toxicity of photolytic products to luminescent bacteria, Chemosphere, 73

(2008) 377-382.
[129] M. Trapido, I. Epold, J. Bolobajev, N. Dulova, Emerging micropollutants in

water/wastewater: growing demand on removal technologies, Environ Sci Pollut

Res Int, 21 (2014) 12217-12222.

[130] N. De la Cruz, L. Esquius, D. Grandjean, A. Magnet, A. Tungler, L.F. de

Alencastro, C. Pulgarin, Degradation of emergent contaminants by UV, UV/H2O2

and neutral photo-Fenton at pilot scale in a domestic wastewater treatment plant,

Water Res, 47 (2013) 5836-5845.

[131] N. De la Cruz, J. Gimenez, S. Esplugas, D. Grandjean, L.F. de Alencastro, C.

Pulgarin, Degradation of 32 emergent contaminants by UV and neutral photo-

fenton in domestic wastewater effluent previously treated by activated sludge,

Water Res, 46 (2012) 1947-1957.

[132] B. Marselli, J. Garcia-Gomez, P.A. Michaud, M.A. Rodrigo, C. Comninellis,

Electrogeneration of hydroxyl radicals on boron-doped diamond electrodes,

Journal of the Electrochemical Society, 150 (2003) D79-D83.

[133] F.C. Moreira, R.A.R. Boaventura, E. Brillas, V.J.P. Vilar, Electrochemical

advanced oxidation processes: A review on their application to synthetic and real

wastewaters, Applied Catalysis B: Environmental, 202 (2017) 217-261.

[134] E. Brillas, C.A. Martínez-Huitle, Decontamination of wastewaters containing

synthetic organic dyes by electrochemical methods. An updated review, Applied

Catalysis B: Environmental, 166-167 (2015) 603-643.

[135] M.A. Rodrigo, N. Oturan, M.A. Oturan, Electrochemically assisted

remediation of pesticides in soils and water: a review, Chemical reviews, 114 (2014)

8720-8745.

[136] A. Sánchez, J. Llanos, C. Sáez, P. Cañizares, M.A. Rodrigo, On the

applications of peroxodiphosphate produced by BDD-electrolyses, Chemical

Engineering Journal, 233 (2013) 8-13.

[137] S. Song, M. Xia, Z. He, H. Ying, B. Lu, J. Chen, Degradation of p-nitrotoluene

in aqueous solution by ozonation combined with sonolysis, J Hazard Mater, 144

(2007) 532-537.

180
[138] G. Boczkaj, M. Gagol, M. Klein, A. Przyjazny, Effective method of treatment

of effluents from production of bitumens under basic pH conditions using

hydrodynamic cavitation aided by external oxidants, Ultrason Sonochem, 40 (2018)

969-979.

[139] M. Ibanez, E. Gracia-Lor, L. Bijlsma, E. Morales, L. Pastor, F. Hernandez,

Removal of emerging contaminants in sewage water subjected to advanced

oxidation with ozone, J Hazard Mater, 260 (2013) 389-398.

[140] G.R. Burleson, T.M. Murray, M. Pollard, Inactivation of viruses and bacteria

by ozone, with and without sonication, Applied microbiology, 29 (1974) 340-344.

[141] E. Dahi, Physicochemical aspects of disinfection of water by means of

ultrasound and ozone, Water research, 10 (1975) 677-684.

[142] Y. Zhao, Z.F. Li, Y. Zhang, Impacts of Ultrasound and Ozone Disinfection of

WWTPs Secondary Effluent, Advanced Materials Research, 610-613 (2012) 1735-

1738.

[143] K.K. Jyoti, A.B. Pandit, Ozone and cavitation for water disinfection,

Biochemical Engineering Journal, 18 (2004) 9-19.

[144] M. Ak, O. Gunduz, Comparison of Organic Matter Removal from Synthetic

and Real Wastewater in a Laboratory-Scale Soil Aquifer Treatment System, Water,

Air, & Soil Pollution, 224 (2013) 1467.

[145] D. Paul, L.W. Canter, Evaluation of photochemical oxidation technology for

remediation of ground water contaminated with organics, Journal of

Environmental Science and Health . Part A: Environmental Science and

Engineering and Toxicology, 25 (1990) 953-985.

[146] R.M.G. Boucher, M.A. Pisano, G. Tortora, E. Sawicki, Synergistic Effects in

Sonochemical Sterilization, Applied Microbiology, 15 (1967) 1257-1261.

[147] M.I. Stefan, Advanced Oxidation Processes for Water Treatment:

Fundamentals and Applications, IWA Publishing, London, 2018.


[148] W. Cao, M. Mehrvar, Slaughterhouse wastewater treatment by combined

anaerobic baffled reactor and UV/H2O2 processes, Chemical Engineering Research

and Design, 89 (2011) 1136-1143.

[149] P. Puspita, F.A. Roddick, N.A. Porter, Decolourisation of secondary effluent

by UV-mediated processes, Chemical Engineering Journal, 171 (2011) 464-473.

[150] M. Barrera, M. Mehrvar, K.A. Gilbride, L.H. McCarthy, A.E. Laursen, V.

Bostan, R. Pushchak, Photolytic treatment of organic constituents and bacterial

pathogens in secondary effluents of synthetic slaughterhouse wastewater,

Chemical Engineering Research and Design, 90 (2012) 1335-1350.

[151] C.F. Bustillo-Lecompte, M. Mehrvar, Treatment of an actual slaughterhouse

wastewater by integration of biological and advanced oxidation processes:

Modeling, optimization, and cost-effectiveness analysis, J Environ Manage, 182

(2016) 651-666.

[152] D.B. Luiz, A.K. Genena , H.J. José, R.F.P.M. Moreira, H.F. Schröder, Tertiary

treatment of slaughterhouse effluent: degradation kinetics applying UV radiation

or H2O2/UV, Water Science & Technology, 60 (2009) 1869-1874.

[153] V. Preethi, K.S. Parama Kalyani, K. Iyappan, C. Srinivasakannan, N.

Balasubramaniam , N. Vedaraman, Ozonation of tannery effluent for removal of

cod and color, Journal of Hazardous Materials, 166 (2009) 150-154.

[154] M.S. Lucas, J.A. Peres, G. Li Puma, Treatment of winery wastewater by

ozone-based advanced oxidation processes (O3, O3/UV and O3/UV/H2O2) in a pilot-

scale bubble column reactor and process economics, Separation and Purification

Technology, 72 (2010) 235-241.

[155] I. Sirés, E. Brillas, M.A. Oturan, M.A. Rodrigo, M. Panizza, Electrochemical

advanced oxidation processes: today and tomorrow. A review, Environmental

Science and Pollution Research, 21 (2014) 8336-8367.

[156] A. Arslan, E. Topkaya, B. Ozbay, I. Özbay, S. Veli, Application of O3/UV/H2O2

oxidation and process optimization for treatment of potato chips manufacturing

wastewater, Water and Environment Journal, 31 (2016) 64-71.

182
[157] A. Arslan, S. Veli, D. Bingöl, Use of response surface methodology for

pretreatment of hospital wastewater by O3/UV and O3/UV/H2O2 processes,

Separation and Purification Technology, 132 (2014) 561-567.

[158] M.S. Lucas, J.A. Peres, G. Li Puma, Treatment of winery wastewater by

ozone-based advanced oxidation processes (O3, O3/UV and O3/UV/H2O2) in a

pilot-scale bubble column reactor and process economics, Separation and

Purification Technology, 72 (2010) 235-241.

[159] C.A. Martínez-Huitle, M.A. Rodrigo, I. Sirés, O. Scialdone, Single and

Coupled Electrochemical Processes and Reactors for the Abatement of Organic

Water Pollutants: A Critical Review, Chemical Reviews, 115 (2015) 13362-13407.

[160] S. Cotillas, J. Llanos, O.G. Miranda, G.C. Díaz-Trujillo, P. Canizares, M.A.

Rodrigoa, Coupling UV irradiation and electrocoagulation for reclamation of

urban wastewater, Electrochimica Acta, 140 (2014) 396-403.

[161] K. Eryuruk, U.T. Un, U.B. Ogutveren, Electrochemical treatment of

wastewater from poultry slaughtering and processing by using iron electrodes,

Journal of Cleaner Production, 172 (2018) 1089-1095.

[162] M. Asselin, P. Drogui, H. Benmoussa, J.F. Blais, Effectiveness of

electrocoagulation process in removing organic compounds from slaughterhouse

wastewater using monopolar and bipolar electrolytic cells, Chemosphere, 72 (2008)

1727-1733.

[163] Z.B. Awang, M.J.K. Bashir, S.R.M. Kutty, M.H. Isa, Post-Treatment of

Slaughterhouse Wastewater using Electrochemical Oxidation, Research Journal of

Chemistry and Environment, 15 (2011) 229-237.

[164] R. Davarnejad, S. Nasiri, Slaughterhouse wastewater treatment using an

advanced oxidation process: Optimization study, Environ. Pollut., 223 (2017) 1-10.

[165] J. Vidal, A. Carvajalb, C. Huiliñir, R. Salazara, Slaughterhouse wastewater

treatment by a combined anaerobic digestion/solar photoelectro-Fenton process

performed in semicontinuous operation, Chemical Engineering Journal, 378 (2019)

122097.
[166] W. Liu, S.A. Andrews, M.I. Stefan, J.R. Bolton, Optimal methods for

quenching H2O2 residuals prior to UFC testing, Water Research, 37 (2003) 3697-

3703.

[167] R.F.P. Nogueira, M.C. Oliveira, W.C. Paterlini, Simple and fast

spectrophotometric determination of H2O2 in photo-Fenton reactions using

metavanadate, Talanta, 66 (2005) 86-91.

[168] F.C. Moreira, J. Soler, M.F. Alpendurada, R.A.R. Boaventura, E. Brillas, V.J.P.

Vilar, Tertiary treatment of a municipal wastewater toward pharmaceuticals

removal by chemical and electrochemical advanced oxidation processes, Water

Res, 105 (2016) 251-263.

[169] I.M. Kolthoff, E.M. Carr, Volumetric Determination of Persulfate in the

Presence of Organic Substances Analytical Chemistry, 25 (1953).

[170] P. Cañizares, C. Sáez, A. Sánchez-Carretero, M.A. Rodrigo, Synthesis of novel

oxidants by electrochemical technology, Journal of Applied Electrochemistry, 39

(2009) 2143-2149.

[171] APHA, American Public Health Association. Standard Methods for the

Examination of Water and Wastewater, 20 ed., Washington, DC, 1998.

[172] J. Bartram, R. Ballance, O. World Health, P. United Nations Environment,

Water quality monitoring : a practical guide to the design and implementation of

freshwater quality studies and monitoring programs / edited by Jamie Bartram and

Richard Ballance, London : E & FN Spon, 1996.

[173] U.S.E.P. Agency, EPA Protocol for the Review of Existing National Primary

Drinking Water Regulations, in: D.C. Office of Water Regulations and Standards:

Washington (Ed.), 2003, pp. 1-60.

[174] S. Verhille, Understanding microbial indicators for drinking water

assessment: interpretation of test results and public health significance, National

collaborating centre for environmental health, (2013) 1-12.

[175] F.C. Moreira, E. Bocos, A.G.F. Faria, J.B.L. Pereira, C.P. Fonte, R.J. Santos,

J.C.B. Lopes, M.M. Dias, M.A. Sanroman, M. Pazos, R.A.R. Boaventura, V.J.P. Vilar,

184
Selecting the best piping arrangement for scaling-up an annular channel reactor:

An experimental and computational fluid dynamics study, Sci Total Environ, 667

(2019) 821-832.

[176] J.C. Espíndola, R.O. Cristóvão, S.R.F. Araújo, T. Neuparth, M.M. Santos, R.

Montes, J.B. Quintana, R. Rodil, R.A.R. Boaventura, V.J.P. Vilar, An innovative

photoreactor, FluHelik, to promote UVC/H2O2 photochemical reactions: Tertiary

treatment of an urban wastewater, 2019, pp. 197-207.

[177] F.C. Moreira, S. Garcia-Segura, R.A.R. Boaventura, E. Brillas, V.J.P. Vilar,

Degradation of the antibiotic trimethoprim by electrochemical advanced oxidation

processes using a carbon-PTFE air-diffusion cathode and a boron-doped diamond

or platinum anode, Applied Catalysis B: Environmental, 160-161 (2014) 492-505.

[178] G. Olsson, J.F. Andrews, The dissolved oxygen profile—A valuable tool for

control of the activated sludge process, Water Research, 12 (1978) 985-1004.

[179] D.P. Saroj, A. Kumar, P. Bose, V. Tare, Y. Dhopavkar, Mineralization of some

natural refractory organic compounds by biodegradation and ozonation, Water

Research, 39 (2005) 1921-1933.

[180] H. Abbas, H. Seif, A. Moursy, Effect of hydraulic retention time on the

activated sludge system, Sixth International Water Technology

ConferenceAlexandria, Egypt. 2001, pp. 277-284.

[181] Y. Zhang, X. Wang, M. Hu, P. Li, Effect of hydraulic retention time (HRT) on

the biodegradation of trichloroethylene wastewater and anaerobic bacterial

community in the UASB reactor, Applied microbiology biotechnology, 99 (2015)

1977-1987.

[182] M. Gerardi, Settleability problems and loss of solids in the activated sludge

process. New Jersey, USA: John Wiley and Sons, 2002. 179 p., 2002.

[183] J. Hoigné, H. Bader, W.R. Haag, J. Staehelin, Rate constants of reactions of

ozone with organic and inorganic compounds in water—III. Inorganic compounds

and radicals, Water Research, 19 (1985) 993-1004.


[184] J.P. Kerrn-Jespersen, M. Henze, Biological phosphorus uptake under anoxic

and aerobic conditions, Water Research, 27 (1993) 617-624.

[185] C. Nebel, D. Gottschling, R.L. Hutchinson, T.J. McBride, D.M. Taylor, J.L.

Pavoni, M.E. Tittlebaum, H.E. Spencer, M. Fleischman, Ozone disinfection of

industrial municipal secondary effluents, Water pollution control federation, 45

(1974) 2493-2507.

[186] S. Jeyanayagam, True confessions of the biological nutrient removal process,

Florida Water Resources Journal, (2005) 37-46.

[187] P. Fongsatitkul, D.G. Wareham, P. Elefsiniotis, P. Charoensuk, Treatment of

a slaughterhouse wastewater: effect of internal recycle rate on chemical oxygen

demand, total Kjeldahl nitrogen and total phosphorus removal, Environmental

Technology, 32 (2011) 1755-1759.

[188] P.F. Wu, G.S. Mittal, Characterization of provincial inspected slaughterhouse

wastewater in Ontario, Canada, Canadian biosystems engineering, 53 (2001) 9-18.

[189] Arayan L., Alisha W. B. Reyes, Huynh T. Hop, Huy T. Xuan, Eun J. Baek, Han

S. Yang, Hong H. Chang, Suk Kim, Antimicrobial effect of different concentrations

of ozonate water in the sanitation of water experimentally inoculated with

Escherichia coli, Preventive Veterinary Medicine, 41 (2017) 84-87.

[190] G.R. Finch, D.W. Smith, Ozone dose-response of Escherichia coli in activated

sludge effluent, Water Research, 23 (1989) 1017-1025.

[191] D. Mara, S. Cairncross, Guidelines for the safe use of wastewater and excreta

in agriculture and aquaculture, World Health Organization, Geneva, 1989.

[192] L. Liberti, M. Notarnicola, Advanced treatment and disinfection for

municipal wastewater reuse in agriculture, Water Science and Technology, 40

(1999) 235-245.

[193] O. Council Directive 98/83/EC on the quality of water intended for human

consumption, European Union, 1998, pp. 42.

186
[194] S. Patil, P. Bourke, J.M. Frias, B.K. Tiwari, P.J. Cullen, Inactivation of

Escherichia coli in orange juice using ozone, Innovative Food Science & Emerging

Technologies, 10 (2009) 551-557.

[195] N. Czekalski, S. Imminger, E. Salhi, M. Veljkovic, K. Kleffel, D. Drissner, F.

Hammes, H. Burgmann, U. von Gunten, Inactivation of Antibiotic Resistant

Bacteria and Resistance Genes by Ozone: From Laboratory Experiments to Full-

Scale Wastewater Treatment, Environmental Science & Technology, 50 (2016)

11862-11871.

[196] S.B. Young, P. Setlow, Mechanisms of Bacillus subtilis spore resistance to and

killing by aqueous ozone, Applied microbiology, 96 (2004) 1133-1142.

[197] S. Hess, C. Gallert, Sensitivity of antibiotic resistant and antibiotic susceptible

Escherichia coli, Enterococcus and Staphylococcus strains against ozone, Water

Health, 13 (2015) 1020-1028.

[198] J.P. Dietrich, F.J. Loge, T.R. Ginn, H. Basagaoglu, Inactivation of particle-

associated microorganisms in wastewater disinfection: modeling of ozone and

chlorine reactive diffusive transport in polydispersed suspensions, Water

Research, 41 (2007) 2189-2201.

[199] A.D. Venosa, M.C. Meckes, E.J. Opatken, Disinfection of filtered and

unfiltered secondary effluent in two ozone contactors, Environment International,

4 (1981) 299-311.

[200] L. Liberti, M. Notarnicola, A. Lopez, Advanced Treatment For Municipal

Wastewater Reuse In Agriculture. III - Ozone Disinfection, Ozone: Science &

Engineering, 22 (2000) 151-166.

[201] R. Pflieger, S. Nikitenko, C. Cariros, R. Mettin, Characterization of Cavitation

Bubbles and Sonoluminescence, Springer2019.

[202] R.A. Torres-Palma, E.A. Serna-Galvis, Sonolysis, Advanced Oxidation

Processes for Waste Water Treatment2018, pp. 177-213.


[203] T. Tuziuti, S. Hatanaka, K. Yasui, T. Kozuka, H. Mitome, Effect of ambient-

pressure reduction on multibubble sonochemiluminescence, J Chem Phys, 116

(2002) 6221-6227.

[204] L.K. Weavers, F.H. Ling, M.R. Hoffmann, Aromatic compound degradation

in water using a combination of sonolysis and ozonolysis, Environmental science

& technology, 32 (1998) 2727-2733.

[205] P.F. Barbier, C. Petrier, Study at 20 kHz and 500 kHz of the ultrasound-ozone

advanced oxidation system: 4-Nitrophenol degradation, journal of advanced

oxidationtechnologies, 1 (1996).

[206] G. Tezcanli-Guyer, N.H. Ince, Individual and combined effects of ultrasound,

ozone and UV irradiation: a case study with textile dyes, Ultrasonics, 42 (2004) 603-

609.

[207] H. Destaillats, A.J. Colussi, J.M. Joseph, M.R. Hoffmann, Synergistic effects of

sonolysis combined with ozonolysis for the oxidation of azobenzene and methyl

orange, Journal of Physical Chemistry, 104 (2000) 8930-8935.

[208] I. Talinli, G.K. Anderson, Interference of hydrogen peroxide on the standard

cod test, Water Research, 26 (1992) 107-110.

[209] E. Lee, H. Lee, Y.K. Kim, K. Sohn, K. Lee, Hydrogen peroxide interference in

chemical oxygen demand during ozone based advanced oxidation of anaerobically

digested livestock wastewater, International Journal of Environmental Science &

Technology, 8 (2011) 381-388.

[210] Lee Seong-Tae, Lee Young-Han, Hong Kwang-Pyo, Lee Sang-Dae, Kim Min-

Kyeong, Park Jong-Hwan, Seo Dong-Cheol, Comparison of BOD, COD, TOC and

DOC as the Indicator of Organic Matter Pollution of Agricultural Surface Water in

Gyeongnam Province, Korean Journal of Soil Science and Fertilizer 46 (2013) 327–

332.

[211] USEPA, Quality Criteria for Water 1986, United States Environmental

Protection Agency, Office of Water Regulations and Standards Washington, DC

20460, 440/5-86-001, 1986.

188
[212] Y. Anjaneyulu, N. Sreedhara Chary, D. Samuel Suman Raj, Decolourization

of Industrial Effluents – Available Methods and Emerging Technologies – A

Review, Reviews in Environmental Science and Bio/Technology, 4 (2005) 245-273.

[213] R. Lee, M.L. Coote, Mechanistic insights into ozone-initiated oxidative

degradation of saturated hydrocarbons and polymers, Physical Chemistry

Chemical Physics, 18 (2016) 24663-24671.

[214] S. Cortez, P. Teixeira, R. Oliveira, M. Mota, Ozonation as polishing treatment

of mature landfill leachate, Journal of Hazardous Materials, 182 (2010) 730-734.

[215] P. Paraskeva, N.J.D. Graham, Treatment of a secondary municipal effluent by

ozone, UV and microfiltration: microbial reduction and effect on effluent quality,

Desalination, 186 (2005) 47-56.

[216] P. Puspita, F.A. Roddick, N.A. Porter, Decolourisation of secondary effluent

by UV-mediated processes, Chemical Engineering Journal, 171 (2011) 464-473.

[217] P. Alfonso-Muniozguren, M. Hazzwan Bohari, A. Sicilia, C. Avignone-Rossa,

M. Bussemaker, D. Saroj, J. Lee, Tertiary treatment of real abattoir wastewater

using combined acoustic cavitation and ozonation, Ultrasonics Sonochemistry, 64

(2020) 104986.

[218] B. Riaño, M. Coca, M.C. García-González, Evaluation of Fenton method and

ozone-based processes for colour and organic matter removal from biologically

pre-treated swine manure, Chemosphere, 117 (2014) 193-199.

[219] S. Casani, M. Rouhany, S. Knøchel, A discussion paper on challenges and

limitations to water reuse and hygiene in the food industry, Water Research, 39

(2005) 1134-1146.

[220] F.C. Moreira, E. Bocos, A.G.F. Faria, J.B.L. Pereira, C.P. Fonte, R.J. Santos,

J.C.B. Lopes, M.M. Dias, M.A. Sanromán, M. Pazos, R.A.R. Boaventura, V.J.P. Vilar,

Selecting the best piping arrangement for scaling-up an annular channel reactor:

An experimental and computational fluid dynamic study, Science of the Total

Environment, 667 (2019) 821-832.


[221] J.C. Espindola, R.O. Cristovao, S.R.F. Araujo, T. Neuparth, M.M. Santos, R.

Montes, J.B. Quintana, R. Rodil, R.A.R. Boaventura, V.J.P. Vilar, An innovative

photoreactor, FluHelik, to promote UVC/H2O2 photochemical reactions: Tertiary

treatment of an urban wastewater, Science of the Total Environment, 667 (2019)

197-207.

[222] E.J. Rosenfeldt, K.G. Linden, S. Canonica, U. von Gunten, Comparison of the

efficiency of OH radical formation during ozonation and the advanced oxidation

processes O3/H2O2 and UV/H2O2, Water Research, 40 (2006) 3695-3704.

[223] F.J. Beltran, J.F. Garcia-Araya, J. Frades, P. Alvarez, O. Gimeno, Effects of

single and combined ozonation with hydrogen peroxide or UV radiation on the

chemical degradation and biodegradability of debittering table olive industrial

wastewaters, Water Research, 33 (1999) 723.

[224] M. Panizza, G. Cerisola, Direct and mediated anodic oxidation of organic

pollutants, Chemical Reviews, 109 (2009) 6541-6569.

[225] A. Sánchez-Carretero, C. Sáez, P. Cañizares, R. M.A., Electrochemical

production of perchlorates using conductive diamond electrolyses, Chemical

Engineering Journal, 166 (2011) 710-714.

[226] G.M. Brown, B. Gu, The Chemistry of Perchlorate in the Environment

(Chapter 2), in: B. Gu, J.D. Coates (Eds.) Perchlorate: Environmental Occurrence,

Interactions and Treatment, Springer, United States, 2006.

[227] D.V. Girenko, A.A. Gyrenko, N.V. Nikolenko, Potentiometric determination

of chlorate impurities in hypochlorite solutions, Int. J. Anal. Chem., 2019 (2019) 7.

[228] N.N. Greenwood, A. Earnshaw, Chemistry of the Elements (2nd Edition),

Butterworth-Heinemann, United Kingdom, 1997.

[229] C.F. Moreira, R.A.R. Boaventura, E. Brillas, V.J.P. Vilar, Degradation of

trimethoprim antibiotic by UVA photoelectro-Fenton process mediated by Fe(III)-

carboxylate complexes, Applied Catalysis B: Environmental, 162 (2015) 34-44.

190
[230] P. Cañizares, C. Sáez, J. Lobato, R. Paz, M.A. Rodrigo, Effect of the Operating

Conditions on the Oxidation Mechanisms in Conductive-Diamond Electrolyses,

Journal of The Electrochemical Society, 154 (2007) E37.

[231] S. Cotillas, J. Llanos, M.A. Rodrigo, P. Cañizares, Use of carbon felt cathodes

for the electrochemical reclamation of urban treated wastewaters, Appl. Catal., B.,

162 (2015) 252-259.

[232] Y. Feng, D.W. Smith, J.R. Bolton, Photolysis of aqueous free chlorine species

(HOCl- and OCl-) with 254 nm ultraviolet light, Journal of Environmental

Engineering and Science, 6 (2007) 277-284.

[233] A.V. Levanov, O.Y. Isaikina, R.B. Gasanova, A.S. Uzhel, V.V. Lunin, Kinetics

of chlorate formation during ozonation of aqueous chloride solutions,

Chemosphere, 229 (2019) 68-76.

[234] M.S. Siddiqui, Chlorine-ozone interactions: formation of chlorate, Water

Research, 30 (1996) 2160-2170.

[235] S. Cotillas, E. Lacasa, C. Sáez, P. Cañizares, M.A. Rodrigo, Disinfection of

urine by conductive-diamond electrochemical oxidation, Applied Catalysis B:

Environment, 229 (2018) 63-70.

[236] S. Cotillas, E. Lacasa, C. Sáez, P. Cañizares, M.A. Rodrigo, Disinfection of

urine by conductive-diamond electrochemical oxidation, Appl. Catal., B., 229

(2018) 63-70.

[237] G. Perez, J. Saiz, R. Ibanez, A.M. Urtiaga, I. Ortiz, Assessment of the formation

of inorganic oxidation by-products during the electrocatalytic treatment of

ammonium from landfill leachates, Water Res, 46 (2012) 2579-2590.

[238] Y. Liu, Q. Wang, S. Zhang, J. Lu, S. Yue, NOM reactivity with chlorine in low

SUVA water, Journal of Water Supply: Research and Technology-Aqua, 60 (2011)

231-239.

[239] Á. Anglada, A. Urtiaga, I. Ortiz, Contributions of electrochemical oxidation

to waste-water treatment: fundamentals and review of applications, Journal of

Chemical Technology & Biotechnology, 84 (2009) 1747-1755.


192

You might also like