Ferrate as a New Treatment Chemical for Removal of ...

171
Montclair State University Montclair State University Montclair State University Digital Montclair State University Digital Commons Commons Theses, Dissertations and Culminating Projects 1-2017 Ferrate as a New Treatment Chemical for Removal of Effluent Ferrate as a New Treatment Chemical for Removal of Effluent Organic Matter (EfOM) and Emerging Micro-Pollutants in Treated Organic Matter (EfOM) and Emerging Micro-Pollutants in Treated Municipal Wastewater for Water Reuse Municipal Wastewater for Water Reuse Nanzhu Li Montclair State University Follow this and additional works at: https://digitalcommons.montclair.edu/etd Part of the Environmental Sciences Commons Recommended Citation Recommended Citation Li, Nanzhu, "Ferrate as a New Treatment Chemical for Removal of Effluent Organic Matter (EfOM) and Emerging Micro-Pollutants in Treated Municipal Wastewater for Water Reuse" (2017). Theses, Dissertations and Culminating Projects. 37. https://digitalcommons.montclair.edu/etd/37 This Dissertation is brought to you for free and open access by Montclair State University Digital Commons. It has been accepted for inclusion in Theses, Dissertations and Culminating Projects by an authorized administrator of Montclair State University Digital Commons. For more information, please contact [email protected].

Transcript of Ferrate as a New Treatment Chemical for Removal of ...

Page 1: Ferrate as a New Treatment Chemical for Removal of ...

Montclair State University Montclair State University

Montclair State University Digital Montclair State University Digital

Commons Commons

Theses, Dissertations and Culminating Projects

1-2017

Ferrate as a New Treatment Chemical for Removal of Effluent Ferrate as a New Treatment Chemical for Removal of Effluent

Organic Matter (EfOM) and Emerging Micro-Pollutants in Treated Organic Matter (EfOM) and Emerging Micro-Pollutants in Treated

Municipal Wastewater for Water Reuse Municipal Wastewater for Water Reuse

Nanzhu Li Montclair State University

Follow this and additional works at: https://digitalcommons.montclair.edu/etd

Part of the Environmental Sciences Commons

Recommended Citation Recommended Citation Li, Nanzhu, "Ferrate as a New Treatment Chemical for Removal of Effluent Organic Matter (EfOM) and Emerging Micro-Pollutants in Treated Municipal Wastewater for Water Reuse" (2017). Theses, Dissertations and Culminating Projects. 37. https://digitalcommons.montclair.edu/etd/37

This Dissertation is brought to you for free and open access by Montclair State University Digital Commons. It has been accepted for inclusion in Theses, Dissertations and Culminating Projects by an authorized administrator of Montclair State University Digital Commons. For more information, please contact [email protected].

Page 2: Ferrate as a New Treatment Chemical for Removal of ...

FERRATE AS A NEW TREATMENT CHEMICAL FOR REMOVAL OF EFFLUENT

ORGANIC MATTER (EFOM) AND EMERGING MICRO-POLLUTANTS IN

TREATED MUNICIPAL WASTEWATER FOR WATER REUSE

A DISSERTATION

Submitted to the Faculty of

Montclair State University in partial fulfillment

of the requirement

for the degree of Doctor of Philosophy

by

NANZHU LI

Montclair State University

Upper Montclair, NJ

2017

Dissertation Chair: Dr. Yang Deng

Page 3: Ferrate as a New Treatment Chemical for Removal of ...

Copyright © 2017 by Nanzhu Li. All rights reserved.

Page 4: Ferrate as a New Treatment Chemical for Removal of ...
Page 5: Ferrate as a New Treatment Chemical for Removal of ...

iv

Abstract

Rapidly expanding population, escalating water consumption, and dwindling

water resources make water shortage a crisis on a global level, particularly in arid and

semi-arid areas or where water sources are highly contaminated. Water reuse,

accomplished by a variety of water reclamation technologies, is a strategically sound

approach to this crisis. In practice, secondary effluent that is the treated wastewater

discharged from municipal wastewater treatment plants (WWTPs) is a preferred

reclaimed water source due to its high quality and stable quantity. However, various

pollutants, effluent organic matters (EfOM) and emerging contaminants in particular, are

still present in secondary effluent. They should be effectively removed before the

reclaimed water is safely used.

In this dissertation study, ferrate (Fe(VI)) is proposed as a new treatment agent for

treatment of secondary effluent, with an emphasis of removing EfOM and emerging

contaminants. Fe(VI) is an environmentally friendly treatment agent with multiple

treatment mechanisms including oxidation, coagulation, adsorption, precipitation and

disinfection. Bench-scale studies were performed to investigate ferrate(VI) treatment of

secondary effluent for the purpose of water reclamation. Special attention was paid to

EfOM and a model emerging contaminant, i.e. mefenamic acid (MEF), which represent

two challenging traditional and emerging contaminants in secondary effluent,

respectively. Initial efforts were made to preliminarily evaluate the treatment

performance of ferrate(VI) for the removal of different secondary effluent contaminants

under different operating conditions. Thereafter, ferrate(VI) reactions with EfOM and

Page 6: Ferrate as a New Treatment Chemical for Removal of ...

v

MEF were mechanistically investigated. All the aforementioned experiments were carried

out in a batch mode. Afterwards, the treatment performance of ferrate(VI) treatment of

secondary effluent was studied in a continuous-flow reactor, which is more commonly

selected in engineering practices. Besides removal of the common wastewater

contaminant, focus would be specially on coagulative behaviors of ferrate(VI) resultant

particles, which could not be studied in a batch mode due to a different hydraulic flow

state. Moreover, preliminary cost analysis was made to compare ferrate(VI) treatment

and ozonation (a common chemical oxidation option in water reuse) for water

reclamation. Finally, implications of ferrate(VI) for environmental management were

discussed and future research suggestion was identified.

Page 7: Ferrate as a New Treatment Chemical for Removal of ...

vi

Acknowledgements

I would like to express my sincere respect and gratitude to my advisor, Dr. Yang

Deng. He is always available for my questions on research and life during my PhD study.

He showed me how to be a good scientist and critical thinking.

Special thanks also go to my committee members Dr. Dibyendu Sarkar (Stevens

Institute of Technology), Dr. Huan Feng (Montclair State University), and Dr. Yueqiang

Liu (Weston Solutions, Inc.) for their guidance and contributions on my research.

I would like to thank the funding and financial support from New Jersey Water

Resources Research Institute (NJWRRI) and Montclair State University Graduate

Assistantship. The work cannot be completed without the facility resource from the

Department of Earth and Environmental Studies and Department of Chemistry and

Biochemistry.

I am grateful to Lei Zheng, Dr. Chanil Jung, Dr. Dongyu Lyu, Dr. Yali Song and

Dr. Xin Huang for their assist on my laboratory work. I want to thank Dr. Laying Wu, Dr.

Xiaona Li and Dr. Kevin Olsen for their guidance on instrumental analysis. A special

appreciation goes to Dr. Jinshan Gao from Chemistry Department for discussion and

collaboration in the research project. I would like to thank all EAES colleagues, faculties

and staffs who shared the adventure in the past five years.

Last but not least, I want to thank my parents, husband, family members and

friends who have been support me in the past years for this long journey.

Page 8: Ferrate as a New Treatment Chemical for Removal of ...

vii

Table of Contents

CHAPTER 1 BACKGROUND AND SIGNIFICANCE .................................................... 1

1.1 Problem Statement and Research Needs ................................................................... 1

1.2 EfOM and emerging contaminants in water reuse .................................................... 5

1.3 Ferrate chemistry and its application in water and wastewater treatment ................. 6

1.4 Nature and Scope of Research ................................................................................ 11

1.5 References ............................................................................................................... 13

CHAPTER 2 RATIONALE, OBJECTIVES, AND HYPOTHESIS ................................ 23

2.1 Rationales ................................................................................................................ 23

2.2 Objectives and hypothesis ....................................................................................... 25

2.3 References ............................................................................................................... 27

CHAPTER 3 PRELIMINARY STUDIES ON FERRATE(VI) TREATMENT OF

SECONDARY EFFLUENT ............................................................................................. 28

3.1 Introduction ............................................................................................................. 28

3.2 Materials and Methods ............................................................................................ 30

3.3 Results and Discussion ............................................................................................ 33

3.4 Conclusions ............................................................................................................. 48

3.5 References ............................................................................................................... 50

CHAPTER 4 FERRATE(VI) REACTIONS WITH EFFLUENT ORGANIC MATTER 53

4.1 Introduction ............................................................................................................. 53

4.2 Materials and Methods ............................................................................................ 55

4.3 Results and Discussion ............................................................................................ 57

4.4 Conclusions ............................................................................................................. 77

4.5 References ............................................................................................................... 79

CHAPTER 5 FERRATE(VI) OXIDATION OF MEFENAMIC ACID (MEF) IN

WATER: KINETICS AND REACTION MECHANISMS ............................................. 81

5.1 Introduction ............................................................................................................. 81

Page 9: Ferrate as a New Treatment Chemical for Removal of ...

viii

5.2 Materials and methods ............................................................................................ 84

5.3 Results and Discussion ............................................................................................ 87

5.4 Conclusions ........................................................................................................... 106

5.5 References ............................................................................................................. 107

CHAPTER 6 PERFORMANCE OF FERRATE(VI) FOR TREATMENT OF

SECONDARY EFFLUENT IN A CONTINUOUS-FLOW REACTOR ....................... 110

6.1 Introduction ........................................................................................................... 110

6.2 Materials and methods .......................................................................................... 113

6.3 Results and Discussion .......................................................................................... 119

6.4 Conclusions ........................................................................................................... 134

6.5 References ............................................................................................................. 135

CHAPTER 7 ENVIRONMENTAL IMPLICATIONS .................................................. 137

7.1 Preliminary cost analysis ....................................................................................... 137

7.2 Implications in Environmental Management ........................................................ 141

7.3 Limitations of Ferrate(VI) Technology ................................................................. 147

7.4 Overall Conclusions .............................................................................................. 148

7.5 Future Research ..................................................................................................... 151

7.6 References ............................................................................................................. 154

Page 10: Ferrate as a New Treatment Chemical for Removal of ...

ix

LIST OF TABLES

Table 3-1 Conversion of DOD and Fe(VI) mass concentrations

Table 3-2 Basic parameters of secondary effluent

Table 5-1 Structures and physiochemical properties of mefenamic acid

Table 5-2 MEF removal efficiencies at different Fe(VI) doses in PBS and SE matrixes

Table 5-3 Comparison of different oxidative processes for the degradation of MEF

Table 5-4 Rate constant k and R2 in different kinetics models for Fe(VI) decomposition

due to the reactions with MEF and Ops

Table 5-5 Fragments (m/z) and chemical formula of OPs identified by multiple-stage

tandem MS

Table 6-1 Basic water quality parameters of secondary effluent

Table 7-1 Cost comparison between Fe(VI) treatment and ozonation + FeCl3 for

treatment of secondary effluent

Page 11: Ferrate as a New Treatment Chemical for Removal of ...

x

LIST OF FIGURES

Fig. 1-1 Chemical structure of ferrate(VI) anion

Fig. 2-1 Conceptual scheme of ferrate(VI) application for water reuse

Fig. 3-1 Preliminary tests for ferrate(VI) treatment of secondary effluent: (a) jar tests; and

(b) secondary effluent after ferrate(VI) treatment and an overnight settling ( ferrate(VI)-

induced sludge settled at the bottom).

Fig. 3-2 COD removal under different DOD and initial pH (COD0 = 33.0 mg/L for pH

8.0 and 34.4 mg/L for pH 5.0)

Fig. 3-3 UV254 removal under different DOD and initial pH (initial UV254 = 0.115 cm-1

for pH 8.0 and 0.130 cm-1 for pH 5.0)

Fig. 3-4 TP removal under different DOD and initial pH (TP0 = 4.84 mg/L for pH 8.0 and

5.19 mg/L for pH 5.0)

Fig. 3-5 Log removal of bacteria indicators during ferrate(VI) treatment of secondary

effluent: (a) DOD = 0.1; and (b) DOD = 3.0 (pH = 8.0, initial total coliform = 2.46 × 105

MPN/100 mL, E.Coli. = 3.29 × 104 MPN/100 mL)

Fig. 3-6 Removal of E3 and OTC during ferrate(VI) treatment of secondary effluent: (a)

pH 8.0; and (b) pH 5.0 (initial E3 = 3.83 mg/L; and initial OTC = 2.00 mg/L)

Fig. 4-1 Residual ferrate(VI) at different reaction times (pH = 7.5): (a) ferrate(VI) decay

at different initial ferrate(VI) doses in secondary effluent; and (b) comparison of

ferrate(VI) decay in secondary effluent and DI water (5.00 mg/L Fe(VI))

Page 12: Ferrate as a New Treatment Chemical for Removal of ...

xi

Fig. 4-2 Experimental data vs. model data for ferrate(VI) decomposition kinetics (pH =

7.5): (a) 1st order reaction; and (b) 2nd order reaction

Fig. 4-3 2nd order reaction rate constant (k) at different ferrate(VI) doses

Fig. 4-4 Turbidity and EfOM at different chemical doses during Fe(VI) or Fe(III)

treatment of the secondary effluent at the initial pH of 7.5: (a) turbidity; and (b) removal

efficiencies of COD and UV254 (initial pH = 7.5; initial COD = 31.7-33.1 mg/L; initial UV254

= 0.135 – 0.142 cm-1)

Fig. 4-5 MW fractions of Fe(VI) and Fe(III)-treated secondary effluent in terms of

UV254: (a) Fe(VI) treatment; and (b) Fe(III) treatment ( initial pH = 7.5)

Fig. 4-6 UV absorbance of the untreated, Fe(VI)-treated and Fe(III)-treated secondary

effluents ( initial pH = 7.5; chemical dose = 15.0 mg/L)

Fig. 4-7(a) Fluorescence excitation-emission matrix (EEM) images of untreated

secondary effluent (X-axis is the excitation wavelength and Y-axis is the emission

wavelength: regions I and II, proteins at Ex/Em of < 250 nm/< 380 nm; region III, fulvic

acid-like materials at Ex/Em of < 250 nm/> 380 nm; region IV, soluble microbial

byproduct-like materials at Ex/Em of 250-280 nm/< 380 nm; and region V, humic acid-

like organics at Ex/Em of > 250 nm/> 380 nm.

Fig. 4-7(b) Fluorescence excitation-emission matrix (EEM) images of untreated and

Fe(VI) treated secondary effluent (initial pH = 7.5)

Page 13: Ferrate as a New Treatment Chemical for Removal of ...

xii

Fig. 4-7(c) Fluorescence excitation-emission matrix (EEM) images of untreated and

Fe(III) treated secondary effluent (initial pH = 7.5)

Fig. 5-1 MEF removal rate under different initial ferrate(VI) dose in buffer and secondary

effluent

Fig. 5-2 Kinetics of overall ferrate(VI) decomposition and ferrate(VI) decomposition due

to self-decay and reactions with MEF and its OPs: (a) [Fe(VI)]:[MEF] = 3:1; (b)

[Fe(VI)]:[MEF] = 4:1; (c) [Fe(VI)]:[MEF] = 5:1; and (d) [Fe(VI)]:[MEF] = 6:1. (pH

=7.5, initial MEF = 0.018 mM, i.e. 4.3 mg/L, and [Fe(VI)] = 0.054, 0.072, 0.090, and

0.108 mM, corresponding to 3.0, 4.0, 5.0 and 6.0 mg/L, respectively).

Fig. 5-3 k vs. [Fe(VI)]:[MEF] (pH =7.5, initial MEF = 0.018 mM)The rate constant k2’

has a negative correlation again the initial Fe (VI):MEF ratio

Fig. 5-4 Structure identification and functional group determination of MEF OPs after

ferrate(VI) oxidation

Fig. 5-5 Multiple-stage tandem mass spectrometry studies of m/z 138 ion in the positive

mode. EIC: extracted ion chromatography of m/z 138; MS1: m/z 138 is the parent ion of

interest; MS2: CID of m/z 138 ion; MS3: CID of m/z 120 ion.

Fig. 5-6 Multiple-stage tandem mass spectrometry studies of m/z 256 ion in the positive

mode. EIC: extracted ion chromatography of m/z 256; MS1: m/z 256 is the parent ion of

interest; MS2: CID of m/z 256 ion; MS3: CID of m/z 238 ion; MS4: CID of m/z 223 ion.

Page 14: Ferrate as a New Treatment Chemical for Removal of ...

xiii

Fig. 5-7 Multiple-stage tandem mass spectrometry studies of m/z 138 ion in the positive

mode. EIC: extracted ion chromatography of m/z 138; MS1: m/z 138 is the parent ion of

interest; MS2: CID of m/z 138 ion; MS3: CID of m/z 120 ion.

Fig. 5-8 Multiple-stage tandem mass spectrometry studies of m/z 256 ion in the positive

mode. EIC: extracted ion chromatography of m/z 256; MS1: m/z 256 is the parent ion of

interest; MS2: CID of m/z 256 ion; MS3: CID of m/z 238 ion; MS4: CID of m/z 223 ion.

Fig. 5-9 Proposed pathway of MEF degradation with Fe(VI)

Fig. 6-1 Scheme of the continuous flow reactor design (CSTR (inlet zone) + PFR

(reaction zone + sludge zone + outlet zone)

Fig. 6-2 The secondary effluent reservoir (left) and Fe(VI) stock solution container

(right)

Fig. 6-3 Flow patterns in the continuous flow reactor using a food-grade dye

Fig. 6-4 Residua turbidity at different contact times with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

Fig. 6-5 TEM images of suspended Fe(VI)-induced particles in the reaction and settling

zone during Fe(VI) treatment of secondary effluent (initial pH = 7.23, Fe(VI) = 5.00

mg/L, and contact time = 120 min)

Page 15: Ferrate as a New Treatment Chemical for Removal of ...

xiv

Fig. 6-6 Zeta potentials of suspended particles in the reaction and settling zone during

Fe(VI) treatment of secondary effluent (initial pH = 7.23; relative standard deviations are

below 5%, not shown in the figure)

Fig. 6-7 Numbers of suspended particles in secondary effluent during Fe(VI) treatment of

secondary effluent (initial pH = 7.23; 5.00 mg/L Fe(VI); relative standard deviations are

below 5%, not shown in the figure)

Fig. 6-8 Size distributions of suspended particles in secondary effluent during Fe(VI)

treatment of secondary effluent (initial pH = 7.23; 5.00 mg/L Fe(VI))

Fig. 6-9 Total Fe in the reaction and settling zone with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

Fig. 6-10 Particulate Fe in the reaction and settling zone with different Fe(VI) doses

during ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

Fig. 6-11 Total P in the reaction and settling zone with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

Fig. 6-12 Dissolved P in the reaction and settling zone with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

Fig. 6-13 COD removal with different Fe(VI) doses during ferrate(VI) treatment of

secondary effluent (initial pH = 7.23, COD0 = 21.0 mg/L, and reaction time = 2 hr)

Page 16: Ferrate as a New Treatment Chemical for Removal of ...

1

CHAPTER 1 BACKGROUND AND SIGNIFICANCE

1.1 Problem Statement and Research Needs

Current municipal water management involving wholesale transfers of water

among geographic areas often leads to a reliance on water conveyance and disposal with

a low efficiency of water and energy use (Englehardt et al. 2015). Meanwhile, rapidly

expanding population, escalating water consumption, and dwindling water resources have

severely aggravated the water shortage problem on a global scale, particularly in arid and

water-stressed countries and regions. The both situations are making water reuse to meet

current and future water demands (USEPA 2012, Watkinson et al. 2007). Presently,

United States Environmental Protection Agency (USEPA) and sixteen state and territorial

environmental agencies (e.g. New Jersey Department of Environmental Protection,

NJDEP) have issued their guidance to highly encourage water reuse through

implementation of a variety of water reclamation technologies (USEPA, 2012). Among

various reclaimed water sources, secondary effluent from municipal wastewater treatment

plants (WWTPs), i.e. biologically treated municipal wastewater, represents a stable non-

seasonal one. It is preferred as a reclaimed water source due to 1) it generally meets 87 of

the 93 numerical primary and secondary U.S. drinking water standards without further

treatment (Englehardt et al. 2001); and 2) the U.S. population generates ~ 121 million m3

sewage every day, of which 1/3 can be reused but only 7-8% is practically reclaimed

(NRC 2012, Intelligence 2010, Miller 2011), leaving a tremendous potential for

expanding the use of reclaimed water in the future((GWI) 2010).

Page 17: Ferrate as a New Treatment Chemical for Removal of ...

2

Secondary wastewater treatment primarily includes a primary settling process, a

subsequent aerobic biodegradation unit (e.g. activated sludge process or trickling

filtration), a secondary settling process, and disinfection. The treatment train has proven

to be technically and economically effective for the removal of particulate matters

indicated as total suspended solids (TSS), dissolved biodegradable organic matter

indicated as 5-day biochemical oxygen demand (BOD5), and pathogens (Tchobanoglous

et al. 2003). Challenges, however, have remained in the implementation of successful

water reuse projects using secondary effluent as a reclaimed water source. Secondary

effluent from WWTPs is a complex wastewater matrix with traditional (e.g., effluent

organic matter(EfOM), phosphorus (P), nitrogen (N), and pathogens) (Park et al. 2005,

Shon et al. 2004, Amy and Drewes 2007, Jarusutthirak and Amy 2001, Her et al. 2003,

Jarusutthirak and Amy 2007, Malhotra et al. 1964, Wang et al. 2005, Barth et al. 1968,

Wang et al. 2007, Kang et al. 2003, Jarvie et al. 2006, Harwood et al. 2005, Shuval et al.

1973, Watts et al. 1995, Tanaka et al. 1998, Liberti et al. 2003, Stevik et al. 1999, Rice

1974, Shon et al. 2006) and emerging contaminants (e.g., pharmaceutical and personal

care products (PPCPs)) (Watkinson et al. 2007, Shon et al. 2006, Bertanza et al. 2013,

Matamoros and Salvadó 2013, Matamoros et al. 2008, Terzić et al. 2008, Wintgens et al.

2008, Bolong et al. 2009, Huang and Sedlak 2001, Kolodziej et al. 2003, Snyder et al.

2003, Sedlak et al. 2000, Park et al. 2009, Nakada et al. 2004, Salste et al. 2007). EfOM

and pharmaceuticals represent examples in the two categories of wastewater-derived

contaminants, respectively. Without sufficient treatment, certain wastewater-derived

pollutants of ecological and health concerns can largely remain in reclaimed water,

Page 18: Ferrate as a New Treatment Chemical for Removal of ...

3

entering into potable water supplies through indirect or direct reuse or non-potable urban,

industrial, and agricultural scenarios, or fouling downstream reverse osmosis (RO)

systems (the most common water reclamation practice in the United States) in water

reclamation. For example, EfOM significantly contributes to organic fouling of RO

membranes in advanced wastewater treatment plants (Lee et al. 2006, Ang and Elimelech

2007).

A portfolio of treatment options, including engineered and managed natural

treatment processes, have been developed and applied to mitigate chemical and microbial

contaminants for water reclamation. However, existing treatment options are limited in

two aspects. Firstly, they often suffer from various technical or economic restrictions. RO

filtration requires intensive energy and generates a large volume of potentially toxic

concentrates, which need to be appropriately disposed of (NRC 2012). Furthermore,

membrane fouling caused by certain wastewater matrix constituents can reduce permeate

flux, increase energy consumption, and lead to frequent chemical cleaning and even

membrane replacement. Activated carbon adsorption is restricted by a high cost and

complex regeneration. Advanced oxidation processes (AOPs) are energy-intensive, in

which hydroxyl radicals unselectively react with target pollutants and co-existing

chemicals, thereby largely wasting chemical oxidants (Dickenson et al. 2009). Ozone has

a low solubility that limits its treatment efficiency, in addition to high cost, safety

concerns, and the production of disinfection byproducts (DBPs) such as bromate.

Secondly, the presence of multiple contaminants in secondary effluent requires a complex

treatment train composed of several treatment options. Each treatment in the system

Page 19: Ferrate as a New Treatment Chemical for Removal of ...

4

focuses on specific contaminants, thereby making the design and operation more complex

and costly. Therefore, there is an urgent demand to develop new, effective, low-cost, and

sustainable water reuse technologies.

Page 20: Ferrate as a New Treatment Chemical for Removal of ...

5

1.2 EfOM and emerging contaminants in water reuse

EfOM in secondary effluent is a complicate organic mixture, particulate and

dissolved, with a broad molecular weight (MW) range between 103 and 106 Da. The

complex mixture is primarily comprised of extracellular polymeric substances (EPS),

soluble microbial products (SMP), and natural organic matter (NOM) derived from

drinking water sources (Laspidou and Rittmann 2002). SMP is originally from the

biomass utilized in the bio-treatment units within WWTPs. NOM in secondary effluent

derives from the NOM in natural water, which is not removed from drinking water and

eventually enters into sewer systems. Although NOM is part of EfOM, EfOM has its

unique physical and chemical properties such as low specific ultraviolet absorbance

(SUVA), rich hydrophilic organic matter, high fluorescence index (FI) values, as well as

abundant polysaccharide and protein-like compounds.

The presence of EfOM in secondary effluent has profound impacts on water

reclamation: 1) many countries and states’ guidelines limit the maximum organic content

in reclaimed water (e.g. BOD5 ≤ 10 mg/L in urban unrestricted reclaimed water in New

Jersey) 2) EfOM as a principal sink of ferrate(VI) exerts a major fraction of chemical

dose; 3) EfOM significantly contributes to the RO membrane fouling in water

reclamation (RO is a widely used water reclamation technology) (Ang and Elimelech

2007, Tang et al. 2011); 4) EfOM serves as DBP precursors; 5) EfOM can bind metals in

wastewater to increase their solubility and bioavailability; and 6) it is a microbial

substrate to enhance biomass re-growth in water distribution networks (Shon et al.

2006).

Page 21: Ferrate as a New Treatment Chemical for Removal of ...

6

On the other hand, emerging contaminants are synthetic or naturally occurring

chemicals or microorganisms that are not commonly monitored in the environment but

have the potential to enter into the environment and cause known or suspected adverse

ecological and/or human health effects (Terzić et al. 2008, Wintgens et al. 2008, Bolong

et al. 2009, Daughton 2004). Examples of emerging contaminants recently found in

untreated and treated wastewater include PPCPs, endocrine disrupting compounds

(EDCs), brominated flame retardants, perfluorinated compounds (PFCs), and new DBPs

(e.g. NDMA). It should be noted that traditional WWTPs are not designed specifically for

the removal of these unregulated micro-pollutants. Consequently, many emerging

contaminants largely flow through WWTPs and enter into effluents. They are persistent,

unregulated, and of great concern in the environment even at traceable levels (Terzić et

al. 2008). The unwanted emerging contaminants should be substantially removed for

water reuse in order to obtain safe reclaimed water. Although activated carbon

adsorption and RO or nanomembrane filtration have been reported to effectively remove

certain emerging contaminants, the former one only transfers the undesirable

contaminants from one phase to another, and the latter one still keep these contaminants

in membrane concentrates. In contrast, chemical oxidation potentially provides an

ultimate solution through chemical degradation of them into less harmful organic

compounds and even nontoxic water and carbon dioxide (Englehardt et al. 2015).

1.3 Ferrate chemistry and its application in water and wastewater treatment

Ferrate(VI) , i.e. the oxyanion FeO42− containing iron in + 6 oxidation state , is

recognized as an environmentally friendly water treatment agent due to its multiple

Page 22: Ferrate as a New Treatment Chemical for Removal of ...

7

treatment mechanisms, the formation of non-toxic final products, and little production of

undesirable DBPs (Sharma 2002, Jiang 2007, Lee et al. 2004). It has a tetrahedral

structure and the four Fe-O bonds are equivalent with covalent character (Fig. 1-1)

(Hoppe et al. 1982). Although the Fe(VI) studies for water treatment began in the 1970s

(Waite and Gilbert 1978, Waite 1979, Gilbert et al. 1976), ferrate(VI) has recently

captured a renewed interest in environmental applications. Once added to water, Fe(VI) is

reduced to intermediate high valence iron species – more reactive Fe(V) and Fe(IV) - and

eventually to stable Fe(III) and even Fe(II) (Jiang 2007, Sharma et al. 2008, Jiang and

Lloyd 2002). Under typical water and wastewater treatment conditions, Fe(III)

immediately precipitates from water to initiate a unique in-situ coagulation (Graham et al.

2010). Besides chemical oxidation and coagulation, many other treatment mechanisms

concurrently proceed due to the production of Fe(III), including disinfection, adsorption

and precipitation. Therefore, in a single dose, Fe(VI) can replace multiple treatment units

and function as oxidant, coagulant, adsorbent, and disinfectant (Waite 2012a).

Particularly, ferrate(VI) is a unique oxidant with a reduction potential up to +2.2

V (Cyr and Bielski 1991, Lee et al. 2005a, Wood 1958). Its reactivity and stability are

acutely pH dependent, because Fe(VI) exists in four individual protonated forms

(H3FeO4+, H2FeO4, HFeO4

-, and FeO42-) with pK1 = 1.6 (Rush et al. 1996), pK2 = 3.5

(Rush et al. 1996, Carr et al. 1985), and pK3 = 7.3(Rush et al. 1996). Fe(VI) is unstable at

an acidic-neutral condition but relatively stable at an alkaline environment. For example,

the half-life of Fe(VI) is around 3 min at pH 7.1, but 2 hours at pH 9.2 (Li et al. 2005).

Another property of Fe(VI)-driven oxidation is selectivity. Similar to other selective

Page 23: Ferrate as a New Treatment Chemical for Removal of ...

8

oxidants, ferrate(VI) preferentially reacts with electron-rich organic moieties (ERMs),

such as activated aromatic compounds (e.g., phenol, aniline, and polycyclic aromatics),

organosulfur compounds, and deprotonated amines (Lee and von Gunten 2010, Yang et

al. 2012). Selective nature of Fe(VI) oxidation, to some degree, restricts its application,

but Fe(VI) is more efficient than nonselective OH· for transforming ERM-containing

pollutants, because less Fe(VI) is wasted by non-pollutants such as water and water

matrix constituents (Lee and von Gunten 2010).

Previous studies have demonstrated the effectiveness of ferrate(VI) in removing a

broad range of water pollutants, such as sewage organic substances (Jiang 2007, Jiang

and Lloyd 2002, Yngard et al. 2008, Alsheyab et al. 2010, Jiang et al. 2006a, Jiang et al.

2009, Jiang et al. 2012), phosphate (Lee et al. 2009), toxic inorganic substances (e.g.

arsenic and cyanides) (Costarramone et al. 2004, Fan et al. 2002, Lee et al. 2003, Lim and

Kim 2010, Prucek et al. 2013, Sharma 2010), algae (Ma and Liu 2002), bacteria and virus

(Jessen et al. 2008, Jiang et al. 2007, 2006b, Hu et al. 2012, Schink and Waite 1980), as

well as emerging pollutants (e.g. pharmaceuticals and microcystins) (Anquandah et al.

2013, Anquandah et al. 2011). Bench-, pilot-, and even full-scale experiments have been

performed to test ferrate(VI) for treatment of drinking water (Jiang et al. 2006b, Lim and

Kim 2009, Jiang et al. 2001, Qu et al. 2003), sewage (Waite 1979, Alsheyab et al. 2010,

Jiang et al. 2006a, Jiang et al. 2012, Jiang et al. 2007, 2006b, Bandala et al. 2009,

Cekerevac et al. 2010, Gombos et al. 2013), industrial wastewaters (Ciabatti et al. 2010),

ballast water (Jessen et al. 2008, Gillis et al. 2005), landfill leachate(Batarseh et al. 2007,

Gravesen 2013), biosolids (Zhang et al. 2012), and wastewater reuse. However, the

Page 24: Ferrate as a New Treatment Chemical for Removal of ...

9

knowledge in ferrate(VI) chemistry remains largely underdeveloped with little

understanding on mechanisms and little information for optimizing treatment sized

systems. Therefore, more efforts are highly needed to advance ferrate chemistry for

specific applications.

Page 25: Ferrate as a New Treatment Chemical for Removal of ...

10

Fig. 1-1 Chemical structure of ferrate(VI) anion

Page 26: Ferrate as a New Treatment Chemical for Removal of ...

11

1.4 Nature and Scope of Research

Simultaneous removal of multiple contaminants present in secondary effluent at a

reasonable cost is a great challenge in the water reuse industry. In particular, EfOM and

emerging contaminants are of great concern due to their chemical and biochemical

persistence. This dissertation focuses on revolutionizing water reuse through

development of ferrate(VI)-based technologies with at least four major benefits: (1)

multiple mechanisms allow the Fe(VI) treatment to simultaneously and effectively

addresses different contaminants; (2) a single treatment achieves different treatment

goals, thus simplifying design and operation, saving costs, and reducing the physical

footprint requirement; (3) ferrate(VI) is a safe oxidant without the production of DBPs,

advantageous over other oxidants (e.g. O3 and chlorine dioxide); and (4) recent advances

in ferrate(VI) production significantly reduce its manufacture costs (< $2/lb) and on-site

ferrate(VI) generators have become commercially available in the water treatment market

(Waite 2012b), making ferrate(VI) a potentially affordable water reclamation technology.

In this dissertation research, bench-scale studies were performed to investigate

ferrate(VI) treatment of secondary effluent for the purpose of water reclamation. Special

attention was paid to EfOM and a model emerging contaminant, i.e. mefenamic acid

(MEF), which represent two challenging traditional and emerging contaminants in

secondary effluent, respectively. Initial efforts were made to preliminarily evaluate the

treatment performance of ferrate(VI) for the removal of different secondary effluent

contaminants under different operating conditions. Thereafter, ferrate(VI) reactions with

EfOM and MEF were mechanistically investigated. All the aforementioned experiments

Page 27: Ferrate as a New Treatment Chemical for Removal of ...

12

were carried out in a batch mode. Afterwards, the treatment performance of ferrate(VI)

treatment of secondary effluent was studied in a continuous-flow reactor, which is more

commonly applied than a batch reactor in engineering practices. Besides removal of the

common wastewater contaminant, focus would be specially on coagulative behaviors of

ferrate(VI) resultant particles, which could not be studied in a batch mode due to a

different hydraulic flow state. Moreover, preliminary cost analysis was made to compare

ferrate(VI) treatment and ozonation (a common chemical oxidation option in water reuse)

for water reclamation. Finally, implications of ferrate(VI) for environmental management

were discussed and future research suggestion was identified.

Page 28: Ferrate as a New Treatment Chemical for Removal of ...

13

1.5 References

Alsheyab, M., Jiang, J.Q. and Stanford, C. (2010) Engineering Aspects of

Electrochemical Generation of Ferrate: A Step Towards Its Full Scale Application for

Water and Wastewater Treatment. Water Air and Soil Pollution 210(1-4), 203-210.

Amy, G. and Drewes, J. (2007) Soil aquifer treatment (SAT) as a natural and sustainable

wastewater reclamation/reuse technology: fate of wastewater effluent organic matter

(EfOM) and trace organic compounds. Environmental Monitoring and Assessment

129(1-3), 19-26.

Ang, W.S. and Elimelech, M. (2007) Protein (BSA) fouling of reverse osmosis

membranes: implications for wastewater reclamation. Journal of Membrane Science

296(1), 83-92.

Anquandah, G.A.K., Sharma, V.K., Knight, D.A., Batchu, S.R. and Gardinali, P.R.

(2011) Oxidation of Trimethoprim by Ferrate(VI): Kinetics, Products, and Antibacterial

Activity. Environmental Science & Technology 45(24), 10575-10581.

Anquandah, G.A.K., Sharma, V.K., Panditi, V.R., Gardinali, P.R., Kim, H. and Oturan,

M.A. (2013) Ferrate(VI) oxidation of propranolol: Kinetics and products. Chemosphere

91(1), 105-109.

Bandala, E.R., Miranda, J., Beltran, M., Vaca, M., Lopez, R. and Torres, L.G. (2009)

Wastewater disinfection and organic matter removal using ferrate (VI) oxidation. Journal

of Water and Health 7(3), 507-513.

Barth, E.F., Brenner, R.C. and Lewis, R.F. (1968) Chemical-biological control of

nitrogen and phosphorus in wastewater effluent. Journal (Water Pollution Control

Federation), 2040-2054.

Batarseh, E.S., Reinhart, D.R. and Daly, L. (2007) Liquid sodium ferrate and Fenton’s

reagent for treatment of mature landfill leachate. Journal of Environmental Engineering

133(11), 1042-1050.

Bertanza, G., Papa, M., Pedrazzani, R., Repice, C., Mazzoleni, G., Steimberg, N., Feretti,

D., Ceretti, E. and Zerbini, I. (2013) EDCs, estrogenicity and genotoxicity reduction in a

mixed (domestic+ textile) secondary effluent by means of ozonation: A full-scale

experience. Science of the Total Environment 458, 160-168.

Page 29: Ferrate as a New Treatment Chemical for Removal of ...

14

Bolong, N., Ismail, A., Salim, M.R. and Matsuura, T. (2009) A review of the effects of

emerging contaminants in wastewater and options for their removal. Desalination 239(1),

229-246.

Carr, J.D., Kelter, P.B., Tabatabai, A., Splichal, D., Erickson, J. and McLaughlin, C.

(1985) Properties of ferrate (VI) in aqueous solution: an alternate oxidant in wastewater

treatment, pp. 1285-1298.

Cekerevac, M.I., Nikolic-Bujanovic, L.N., Mirkovic, M.B. and Popovic, N.H. (2010)

APPLICATION OF ELECTROCHEMICALLY SYNTHESIZED FERRATE(VI) IN

THE PURIFICATION OF WASTEWATER FROM COAL SEPARATION PLANT.

Hemijska Industrija 64(5), 423-430.

Ciabatti, I., Tognotti, F. and Lombardi, L. (2010) Treatment and reuse of dyeing effluents

by potassium ferrate. Desalination 250(1), 222-228.

Costarramone, N., Kneip, A. and Castetbon, A. (2004) Ferrate(VI) oxidation of cyanide

in water. Environmental Technology 25(8), 945-955.

Cyr, J.E. and Bielski, B.H. (1991) The reduction of ferrate (VI) to ferrate (V) by

ascorbate. Free Radical Biology and Medicine 11(2), 157-160.

Daughton, C.G. (2004) Non-regulated water contaminants: emerging research.

Environmental Impact Assessment Review 24(7), 711-732.

Dickenson, E.R., Drewes, J.r.E., Sedlak, D.L., Wert, E.C. and Snyder, S.A. (2009)

Applying surrogates and indicators to assess removal efficiency of trace organic

chemicals during chemical oxidation of wastewaters. Environmental Science &

Technology 43(16), 6242-6247.

Englehardt, J.D., Amy, V.P., Bloetscher, F., Chin, D.A., Fleming, L.E., Gokgoz, S.,

Rose, J.B., Solo-Gabriele, H. and Tchobanoglous, G. (2001) Comparative Assessment of

Human and Ecological Impacts from Municipal Wastewater Disposal Methods in

Southeast Florida. Prepared for Florida Water Environment Association Utility Council.

Englehardt, J., Wu, T., Bloetscher, F., Deng, Y., Du Pisani, P., Eilert, S., Elmir, S., Guo,

T., Jacangelo, J. and LeChevallier, M. (2015) NET-ZERO WATER MANAGEMENT:

ACHIEVING ENERGY-POSITIVE MUNICIPAL WATER SUPPLY. Environmental

Science: Water Research & Technology.

Page 30: Ferrate as a New Treatment Chemical for Removal of ...

15

Fan, M.H., Brown, R.C. and Huang, C.P. (2002) Preliminary studies of the oxidation of

arsenic(III) by potassium ferrate. International Journal of Environment and Pollution

18(1), 91-96.

Gilbert, M., Waite, T.D. and Hare, C. (1976) An Investigation of the Applicability of

Ferrate Ion for Disinfection. Journal of American Water Works Association 68(9), 495-

497.

Gillis, N., Walters, L., Yeargain, M. and Reinhart, D. (2005) Effectiveness of Ferrate

(FeO4-2) as a ballast water disinfectant. Integrative and Comparative Biology 45(6),

1137-1137.

Gombos, E., Barkacs, K., Felfoldi, T., Vertes, C., Mako, M., Palko, G. and Zaray, G.

(2013) Removal of organic matters in wastewater treatment by ferrate (VI)-technology.

Microchemical Journal 107, 115-120.

Graham, N.J.D., Khoi, T.T. and Jiang, J.Q. (2010) Oxidation and coagulation of humic

substances by potassium ferrate. Water Science and Technology 62(4), 929-936.

Gravesen, C. (2013) Fe(VI) as a new chemical agent for treatment of landfill leachate.

Master Thesis, Montclair State University.

G.W.I. (2010) Media Analytics Ltd. Oxford, UK.

Harwood, V.J., Levine, A.D., Scott, T.M., Chivukula, V., Lukasik, J., Farrah, S.R. and

Rose, J.B. (2005) Validity of the indicator organism paradigm for pathogen reduction in

reclaimed water and public health protection. Applied and Environmental Microbiology

71(6), 3163-3170.

Her, N., Amy, G., McKnight, D., Sohn, J. and Yoon, Y. (2003) Characterization of DOM

as a function of MW by fluorescence EEM and HPLC-SEC using UVA, DOC, and

fluorescence detection. Water Research 37(17), 4295-4303.

Hoppe, M.L., Schlemper, E.O. and Murmann, R.K. (1982) Structure of dipotassium

ferrate(VI). Acta Crystallographica Section B: Structural Crystallography and Crystal

Chemistry 38(8), 2237-2239.

Huang, C.H. and Sedlak, D.L. (2001) Analysis of estrogenic hormones in municipal

wastewater effluent and surface water using enzyme‐linked immunosorbent assay and gas

chromatography/tandem mass spectrometry. Environmental Toxicology and Chemistry

20(1), 133-139.

Page 31: Ferrate as a New Treatment Chemical for Removal of ...

16

Hu, L.H., Page, M.A., Sigstam, T., Kohn, T., Marinas, B.J. and Strathmann, T.J. (2012)

Inactivation of Bacteriophage MS2 with Potassium Ferrate(VI). Environmental Science

& Technology 46(21), 12079-12087.

Intelligence, G.W. (2010) Municipal Water Reuse Markets 2010. Media Analytics Ltd.

Oxford, UK.

Jarusutthirak, C. and Amy, G. (2001) Membrane filtration of wastewater effluents for

reuse: effluent organic matter rejection and fouling. Water Science & Technology 43(10),

225-232.

Jarusutthirak, C. and Amy, G. (2007) Understanding soluble microbial products (SMP) as

a component of effluent organic matter (EfOM). Water Research 41(12), 2787-2793.

Jarvie, H.P., Neal, C. and Withers, P.J. (2006) Sewage-effluent phosphorus: a greater risk

to river eutrophication than agricultural phosphorus? Science of the Total Environment

360(1), 246-253.

Jessen, A., Randall, A., Reinhart, D. and Daly, L. (2008) Effectiveness and Kinetics of

Ferrate as a Disinfectant for Ballast Water. Water Environment Research 80(6), 561-569.

Jiang, J.Q., Lloyd, B. and Grigore, L. (2001) Preparation and evaluation of potassium

ferrate as an oxidant and coagulant for potable water treatment. Environmental

Engineering Science 18(5), 323-328.

Jiang, J.Q. and Lloyd, B. (2002) Progress in the development and use of ferrate(VI) salt

as an oxidant and coagulant for water and wastewater treatment. Water Research 36(6),

1397-1408.

Jiang, J.Q., Panagouopoulos, A., Bauer, M. and Pearce, P. (2006a) The application of

potassium ferrate for sewage treatment. Journal of Environmental Management 79(2),

215-220.

Jiang, J.Q., Wang, S. and Panagoulopoulos, A. (2006b) The exploration of potassium

ferrate(VI) as a disinfectant/coagulant in water and wastewater treatment. Chemosphere

63(2), 212-219.

Jiang, J.Q., Wang, S. and Panagoulopoulos, A. (2007) The role of potassium ferrate(VI)

in the inactivation of Escherichia coli and in the reduction of COD for water remediation.

Desalination 210(1-3), 266-273.

Page 32: Ferrate as a New Treatment Chemical for Removal of ...

17

Jiang, J.Q. (2007) Research progress in the use of ferrate(VI) for the environmental

remediation. Journal of Hazardous Materials 146(3), 617-623.

Jiang, J.Q., Stanford, C. and Alsheyab, M. (2009) The online generation and application

of ferrate(VI) for sewage treatment-A pilot scale trial. Separation and Purification

Technology 68(2), 227-231.

Jiang, J.Q., Stanford, C. and Mollazeinal, A. (2012) THE APPLICATION OF FERRATE

FOR SEWAGE TREATMENT: PILOT- TO FULL-SCALE TRIALS. Global Nest

Journal 14(1), 93-99.

Kang, S.K., Choo, K.H. and Lim, K.H. (2003) Use of iron oxide particles as adsorbents

to enhance phosphorus removal from secondary wastewater effluent. Separation science

and technology 38(15), 3853-3874.

Kolodziej, E.P., Gray, J.L. and Sedlak, D.L. (2003) Quantification of steroid hormones

with pheromonal properties in municipal wastewater effluent. Environmental Toxicology

and Chemistry 22(11), 2622-2629.

Laspidou, C.S. and Rittmann, B.E. (2002) A unified theory for extracellular polymeric

substances, soluble microbial products, and active and inert biomass. Water Research

36(11), 2711-2720.

Lee, S., Ang, W.S. and Elimelech, M. (2006) Fouling of reverse osmosis membranes by

hydrophilic organic matter: implications for water reuse. Desalination 187(1), 313-321.

Lee, Y., Um, I.H. and Yoon, J. (2003) Arsenic(III) oxidation by iron(VI) (ferrate) and

subsequent removal of arsenic(V) by iron(III) coagulation. Environmental Science &

Technology 37(24), 5750-5756.

Lee, Y., Cho, M., Kim, J.Y. and Yoon, J. (2004) Chemistry of ferrate (Fe(VI)) in aqueous

solution and its applications as a green chemical. Journal of Industrial and Engineering

Chemistry 10(1), 161-171.

Lee, Y., Yoon, J. and Von Gunten, U. (2005) Kinetics of the oxidation of phenols and

phenolic endocrine disruptors during water treatment with ferrate (Fe(VI)).

Environmental Science & Technology 39(22), 8978-8984.

Lee, Y., Zimmermann, S.G., Kieu, A.T. and von Gunten, U. (2009) Ferrate (Fe (VI))

application for municipal wastewater treatment: a novel process for simultaneous

micropollutant oxidation and phosphate removal. Environmental Science & Technology

43(10), 3831-3838.

Page 33: Ferrate as a New Treatment Chemical for Removal of ...

18

Lee, Y. and von Gunten, U. (2010) Oxidative transformation of micropollutants during

municipal wastewater treatment: Comparison of kinetic aspects of selective (chlorine,

chlorine dioxide, ferrate(VI), and ozone) and non-selective oxidants (hydroxyl radical).

Water Research 44(2), 555-566.

Liberti, L., Notarnicola, M. and Petruzzelli, D. (2003) Advanced treatment for municipal

wastewater reuse in agriculture. UV disinfection: parasite removal and by-product

formation. Desalination 152(1), 315-324.

Lim, M. and Kim, M.J. (2009) Removal of Natural Organic Matter from River Water

Using Potassium Ferrate(VI). Water Air and Soil Pollution 200(1-4), 181-189.

Lim, M. and Kim, M.J. (2010) Effectiveness of Potassium Ferrate (K2FeO4) for

Simultaneous Removal of Heavy Metals and Natural Organic Matters from River Water.

Water Air and Soil Pollution 211(1-4), 313-322.

Li, C., Li, X.Z. and Graham, N. (2005) A study of the preparation and reactivity of

potassium ferrate. Chemosphere 61(4), 537-543.

Malhotra, S., Lee, G.F. and Rohlich, G. (1964) Nutrient removal from secondary effluent

by alum flocculation and lime precipitation. International Journal of Air and Water

Pollution 8, 487-500.

Matamoros, V., García, J. and Bayona, J.M. (2008) Organic micropollutant removal in a

full-scale surface flow constructed wetland fed with secondary effluent. Water Research

42(3), 653-660.

Matamoros, V. and Salvadó, V. (2013) Evaluation of a coagulation/flocculation-lamellar

clarifier and filtration-UV-chlorination reactor for removing emerging contaminants at

full-scale wastewater treatment plants in Spain. Journal of environmental management

117, 96-102.

Ma, J. and Liu, W. (2002) Effectiveness and mechanism of potassium ferrate(VI)

preoxidation for algae removal by coagulation. Water Research 36(4), 871-878.

Miller, W.G. (2011) Water Reuse in the U.S.A.: Current Challenges and New

Initiatives.”. 8th IWA International Conference on Water Reclamation and Reuse,

Barcelona, Spain.

Nakada, N., Nyunoya, H., Nakamura, M., Hara, A., Iguchi, T. and Takada, H. (2004)

Identification of estrogenic compounds in wastewater effluent. Environmental

Toxicology and Chemistry 23(12), 2807-2815.

Page 34: Ferrate as a New Treatment Chemical for Removal of ...

19

NRC (2012) Water Reuse: Potential for Expanding the Nation's Water Supply Through

Reuse of Municipal Wastewater, The National Academies Press: Washington D.C.

Park, K.-Y., Ahn, K.-H., Kim, K.-p., Kweon, J.H. and Maeng, S.-K. (2005)

Characteristics and treatability of persistent colors in biologically treated wastewater

effluents. Environmental engineering science 22(5), 557-566.

Park, N., Vanderford, B.J., Snyder, S.A., Sarp, S., Kim, S.D. and Cho, J. (2009) Effective

controls of micropollutants included in wastewater effluent using constructed wetlands

under anoxic condition. Ecological Engineering 35(3), 418-423.

Prucek, R., Tucek, J., Kolarik, J., Filip, J., Marusak, Z., Sharma, V.K. and Zboril, R.

(2013) Ferrate(VI)-Induced Arsenite and Arsenate Removal by In Situ Structural

Incorporation into Magnetic Iron(III) Oxide Nanoparticles. Environmental Science &

Technology 47(7), 3283-3292.

Qu, J.H., Liu, H.J., Liu, S.X. and Lei, P.J. (2003) Reduction of fulvic acid in drinking

water by ferrate. Journal of Environmental Engineering-Asce 129(1), 17-24.

Rice, R.C. (1974) Soil clogging during infiltration of secondary effluent. Journal (Water

Pollution Control Federation), 708-716.

Rush, J.D., Zhao, Z. and Bielski, B.H. (1996) Reaction of ferrate (VI)/ferrate (V) with

hydrogen peroxide and superoxide anion--a stopped-flow and premix pulse radiolysis

study. Free radical research 24(3), 187-198.

Salste, L., Leskinen, P., Virta, M. and Kronberg, L. (2007) Determination of estrogens

and estrogenic activity in wastewater effluent by chemical analysis and the

bioluminescent yeast assay. Science of the Total Environment 378(3), 343-351.

Schink, T. and Waite, T.D. (1980) Inactivation of f2 virus with ferrate (VI). Water

Research 14(12), 1705-1717.

Sedlak, D.L., Gray, J.L. and Pinkston, K.E. (2000) Peer Reviewed: understanding

microcontaminants in recycled water. Environmental Science & Technology 34(23),

508A-515A.

Sharma, V.K. (2002) Potassium ferrate(VI): an environmentally friendly oxidant.

Advances in Environmental Research 6(2), 143-156.

Page 35: Ferrate as a New Treatment Chemical for Removal of ...

20

Sharma, V.K., Yngard, R.A., Cabelli, D.E. and Baum, J.C. (2008) Ferrate(VI) and

ferrate(V) oxidation of cyanide, thiocyanate, and copper(I) cyanide. Radiation Physics

and Chemistry 77(6), 761-767.

Sharma, V.K. (2010) Oxidation of Inorganic Compounds by Ferrate(VI) and Ferrate(V):

One-Electron and Two-Electron Transfer Steps. Environmental Science & Technology

44(13), 5148-5152.

Shon, H., Vigneswaran, S., Kim, I.S., Cho, J. and Ngo, H. (2004) The effect of

pretreatment to ultrafiltration of biologically treated sewage effluent: a detailed effluent

organic matter (EfOM) characterization. Water Research 38(7), 1933-1939.

Shon, H., Vigneswaran, S. and Snyder, S. (2006) Effluent organic matter (EfOM) in

wastewater: constituents, effects, and treatment. Critical reviews in environmental

science and technology 36(4), 327-374.

Shuval, H.I., Cohen, J. and Kolodney, R. (1973) Regrowth of coliforms and fecal

coliforms in chlorinated wastewater effluent. Water Research 7(4), 537-546.

Snyder, S.A., Westerhoff, P., Yoon, Y. and Sedlak, D.L. (2003) Pharmaceuticals,

personal care products, and endocrine disruptors in water: implications for the water

industry. Environmental engineering science 20(5), 449-469.

Stevik, T.K., Ausland, G., Jenssen, P.D. and Siegrist, R.L. (1999) Removal of E. coli.

during intermittent filtration of wastewater effluent as affected by dosing rate and media

type. Water Research 33(9), 2088-2098.

Tanaka, H., Asano, T., Schroeder, E.D. and Tchobanoglous, G. (1998) Estimating the

safety of wastewater reclamation and reuse using enteric virus monitoring data. Water

Environment Research 70(1), 39-51.

Tang, C.Y., Chong, T. and Fane, A.G. (2011) Colloidal interactions and fouling of NF

and RO membranes: a review. Advances in colloid and interface science 164(1), 126-

143.

Terzić, S., Senta, I., Ahel, M., Gros, M., Petrović, M., Barcelo, D., Müller, J., Knepper,

T., Martí, I. and Ventura, F. (2008) Occurrence and fate of emerging wastewater

contaminants in Western Balkan Region. Science of the Total Environment 399(1), 66-

77.

Tchobanoglous, G., Franklin, L. and Stensel, H.D. (2003) Wastewater engineering:

treatment and reuse. Metcalf & Eddy, Inc., New York.

Page 36: Ferrate as a New Treatment Chemical for Removal of ...

21

USEPA (2012) Guidelines for Water Reuse.

Waite, T.D. and Gilbert, M. (1978) Oxidative destruction of phenol and other organic

water residuals by iron(VI) ferrate. Journal Name: J. - Water Pollut. Control Fed.;

(United States); Journal Volume: 50:3, Medium: X; Size: Pages: 543-551.

Waite, T.D. (1979) Feasibility of Wastewater Treatment with Ferrate. Journal of the

Environmental Engineering Division 105(6), 1023-1034.

Waite, T.D. (2012a) On-Site Production of Ferrate for Water and Wastewater

Purification. American Laboratory 44(10), 26-28.

Waite, T.D. (2012b) Chemistry-On-Site Production of Ferrate for Water and Wastewater

Purification. American Laboratory 44(10), 26.

Watkinson, A., Murby, E. and Costanzo, S. (2007) Removal of antibiotics in

conventional and advanced wastewater treatment: Implications for environmental

discharge and wastewater recycling. Water Research 41(18), 4164-4176.

Wang, Y., Han, T., Xu, Z., Bao, G. and Zhu, T. (2005) Optimization of phosphorus

removal from secondary effluent using simplex method in Tianjin, China. Journal of

hazardous materials 121(1), 183-186.

Wang, L.-S., Hu, H.-Y. and Wang, C. (2007) Effect of ammonia nitrogen and dissolved

organic matter fractions on the genotoxicity of wastewater effluent during chlorine

disinfection. Environmental Science & Technology 41(1), 160-165.

Watts, R.J., Kong, S., Orr, M.P., Miller, G.C. and Henry, B.E. (1995) Photocatalytic

inactivation of coliform bacteria and viruses in secondary wastewater effluent. Water

Research 29(1), 95-100.

Wintgens, T., Salehi, F., Hochstrat, R. and Melin, T. (2008) Emerging contaminants and

treatment options in water recycling for indirect potable use. Water Science and

Technology 57(1), 99-108.

Wood, R.H. (1958) The heat, free energy and entropy of the ferrate (VI) ion. Journal of

the American Chemical Society 80(9), 2038-2041.

Yang, B., Ying, G.G., Zhao, J.L., Liu, S., Zhou, L.J. and Chen, F. (2012) Removal of

selected endocrine disrupting chemicals (EDCs) and pharmaceuticals and personal care

products (PPCPs) during ferrate(VI) treatment of secondary wastewater effluents. Water

Research 46(7), 2194-2204.

Page 37: Ferrate as a New Treatment Chemical for Removal of ...

22

Yngard, R.A., Sharma, V.K., Filip, J. and Zboril, R. (2008) Ferrate(VI) oxidation of

weak-acid dissociable cyanides. Environmental Science & Technology 42(8), 3005-3010.

Zhang, X.H., Lei, H.Y., Chen, K., Liu, Z., Wu, H. and Liang, H.Y. (2012) Effect of

potassium ferrate (K2FeO4) on sludge dewaterability under different pH conditions.

Chemical Engineering Journal 210, 467-474.

Page 38: Ferrate as a New Treatment Chemical for Removal of ...

23

CHAPTER 2 RATIONALE, OBJECTIVES, AND HYPOTHESIS

2.1 Rationales

Ferrate(VI) has a potential to serve as a new treatment agent to develop a new-

generation water reuse technology due to at least three reasons. Firstly, multiple treatment

mechanisms allow for the removal of different contaminants in secondary effluent (Fig.

2-1). Specifically, chemical oxidation can decompose EfOM and certain emerging

contaminants; disinfection can inactivate pathogenic microorganisms; coagulation can

remove particulate and colloidal particles, thus reducing TSS; precipitation can transform

unwanted dissolved inorganic species (e.g. phosphate) into particulates, which is

subsequently removed in a downstream solid-liquid separation ; and adsorption due to the

produced iron particles can adsorb certain dissolved or particulate pollutants (e.g. heavy

metals). Secondly, ferrate(VI) is a safe oxidant compared with other oxidants such as

chlorine and ozone in terms of the DBP production. It is well known that chlorine reacts

with NOM and EfOM to produce different DBP species (e.g. two EPA regulated DBPs -

trihalomethanes and haloacetic acids), while ozone can oxidize bromide to bromate

(another EPA regulated DBP) (Sedlak and von Gunten 2011). The contaminants are

produced after treatment and can reach end users through water distribution systems.

Thirdly, the final product of ferrate(VI) treatment is iron sludge. Similar to water

treatment residual (WTR) from drinking water treatment plants, the ferrate(VI)-based

iron sludge has a potential to be used for land application, as long as it passes certain

chemical leaching tests. If this is the case, this will greatly save landfill space and waste

disposal costs.

Page 39: Ferrate as a New Treatment Chemical for Removal of ...

24

Fig. 2-1 Conceptual scheme of ferrate(VI) application for water reuse

Secondary Effluent

• EfOM• P• Pathogens• TSS• Emerging

contaminants• ……

Chemical OxidationDisinfectionCoagulationPrecipitationAdsorption

Ferrate(VI)

Reverse osmosis

Non-potable reuse

Direct or indirect potable reuse

Concentrate

Page 40: Ferrate as a New Treatment Chemical for Removal of ...

25

2.2 Objectives and hypothesis

The long-term goal of this study is to develop new, technically viable, and cost

effective water reuse technologies to address both traditional and emerging contaminants

in secondary effluent. The overall objective of this dissertation research is to provide a

scientific basis for ferrate(VI) reactions with contaminants in biologically treated

secondary effluent, with a purpose of the development of ferrate(VI)-based water

reclamation technologies for water reuse. Particularly, the reaction mechanisms of Fe(VI)

with EfOM and a model emerging pollutant, i.e. MEF, will be investigated. The central

hypothesis is that ferrate(VI) is capable of effectively and simultaneously removing

different traditional and emerging contaminants in secondary effluent for water reuse

through multiple treatment mechanisms, thereby providing a new, technically reliable and

low-cost water reclamation technology. To achieve the overall objective, four specific

objectives were pursued, including:

1. To evaluate the technical feasibility of ferrate(VI) as a new treatment agent for

treatment of secondary effluent.

2. To understand the interactions between EfOM and Fe(VI) as well as

characterize chemical oxidation products of EfOM;

3. To investigate reaction kinetics and mechanisms of ferrate(VI) oxidation of

mefenamic acid (a model emerging contaminant) in water;

4. To evaluate the performance of ferrate(VI)-based secondary effluent treatment

in a continuous flow reactor.

Page 41: Ferrate as a New Treatment Chemical for Removal of ...

26

In this dissertation research, four tasks were designed to achieve the four aforementioned

specific objectives. The first three tasks were performed in a batch reactor, while the

fourth task was carried out in a continuous-flow reactor. The first task preliminarily

examined the treatment feasibility of ferrate(VI) for different contaminants in a

secondary effluent as well as evaluate the effects of solution pH and chemical dose on the

treatment results. The second task targeted at ferrate(VI) degradation of EfOM (a

traditional contaminant). The degradation products of EfOM after ferrate(VI) treatment

was particularly characterized to understand the reaction mechanisms of ferrate(VI)

oxidation of EfOM. In the third task, ferrate(VI) interactions with MEF (a model

emerging contaminant) were investigated. Ferrate(VI) decomposition kinetics were

studied. And the oxidation byproducts of MEF were identified to reveal the reaction

pathways. In the fourth task, the treatment performance of ferrate(VI) in a continuous-

flow reactor (a much more commonly applied reactor type in engineering practices) was

evaluated. In particular, the coagulation driven by ferrate(VI)-induced particles was

investigated, because the iron floc-driven coagulation and ensuing settling could not be

readily studied in a batch reactor in which the solution was under a complete mixing

state. Finally, preliminary cost analysis was made to compare ferrate(VI)-based

technologies with ozonation that is a common chemical oxidation practice in water reuse.

Based on these major findings in the aforementioned four tasks, the implication of this

dissertation research in environmental management was discussed, and future research

needs were identified.

Page 42: Ferrate as a New Treatment Chemical for Removal of ...

27

2.3 References

Sedlak, D.L. and von Gunten, U. (2011) The chlorine dilemma. Science 331(6013), 42-

43.

Page 43: Ferrate as a New Treatment Chemical for Removal of ...

28

CHAPTER 3 PRELIMINARY STUDIES ON FERRATE(VI)

TREATMENT OF SECONDARY EFFLUENT

3.1 Introduction

Ferrate(VI) is an emerging, safe, and multi-treatment agent for water and

wastewater treatment (Sharma 2002, Jiang 2007, Ghernaout and Naceur 2011). It is being

primarily used as an environmentally friendly oxidant due to its powerful oxidative

capacity and little production of DBPs (Ghernaout and Naceur 2011, Alsheyab et al.

2009). Besides chemical oxidation and disinfection, ferrate(VI) can initiate coagulation,

precipitation, and adsorption due to the in-situ production of Fe(III) in water. The

multiple treatment mechanisms allow ferrate(VI) to simultaneously remove different

contaminants from water and wastewater. With this unique property, ferrate(VI) is an

attractive option for water reuse using secondary effluent because a variety of

contaminants are present in biologically treated municipal wastewater. However, the

information regarding ferrate(VI) removal of secondary effluent contaminants is very

limited.

Although several hundreds of publications are available to report ferrate(VI) for

different environmental applications (Waite 2012b), the most of them focus only on a

single pollutant or a specific group of compounds. Comprehensive evaluations of

ferrate(VI) for various contaminants in a specific water matrix are very few. However,

such an evaluation is important to water reclamation with ferrate(VI), because this can

examine whether the removals of different contaminants of concern from secondary

Page 44: Ferrate as a New Treatment Chemical for Removal of ...

29

effluent can occur under the operating conditions related to wastewater treatment (e.g.

ferrate(VI) dose and solution pH). Another limitation in previous studies is the usage of

phosphate buffered solution. Phosphate was widely used in ferrate(VI) studies because it,

besides pH control, can complex the Fe(III) from Fe(VI) reduction and prevent the

formation of iron precipitates and sludge. It facilitates the studies on ferrate(VI)

oxidation, because ferrate(VI) can be directly monitored using spectrophotometry at 510

nm in the absence of particulate matters and the effect of iron sludge adsorption is not

considered. However, such a high phosphate concentration (typically a few to a few tens

of mM) is not present in a real secondary effluent. Application of phosphate in laboratory

tests may lead to the deviation of experimental data from data obtained in engineering

practices. For example, it has been noticed that ferrate(VI) decomposition rates can be

accelerated with an increasing phosphate concentration (Jiang et al. 2015). Another

drawback for the phosphate buffer solution is that ferrate(VI) precipitation of phosphate

and other dissolved species cannot be studied, because few iron precipitates are produced

due to the complexing effect of phosphate.

The objective of this chapter is to preliminarily evaluate ferrate(VI) removal of

different contaminants in secondary effluent. Phosphate or other buffer chemicals were

not used, in case that these externally added chemicals influenced ferrate(VI)

decomposition and treatment. Two factors were tested in this study, including pH and

ferrate(VI) doses.

Page 45: Ferrate as a New Treatment Chemical for Removal of ...

30

3.2 Materials and Methods

Chemicals and Reagents. Secondary effluent samples were collected from the

secondary clarifier of the Joint Meeting of Essex & Union Counties in Elizabeth, New

Jersey, which is a municipal WWTP using activated sludge process. Once collected, the

samples were transported to Montclair State University’s water treatment laboratory and

stored in a refrigerator at 4˚C until use. Potassium ferrate (K2FeO4, 96% purity) and all

other chemicals (reagent grade) were purchased from Sigma-Aldrich (St. Louis, MO, US)

or Fisher-Scientific (Pittsburgh, PA, US).

Fe(VI) treatment tests. Ferrate(VI) treatment tests were performed in 600 mL

glass beakers with 200 mL secondary effluent on a four paddle programmable jar tester

(Phipps & Bird - 7790-950) (Fig. 3-1). Estriol (E3) and oxytetracycline (OTC) were

spiked into the secondary effluent as model micro-pollutants. The treatment was initiated

through the addition of an appropriate weight of K2FeO4 to the secondary effluent. Initial

pH was adjusted to a designated level using 1 N NaOH or H2SO4 solution. In the first

oxidation phase, the solution was rapidly mixed at 150 rpm until all ferrate(VI) depleted.

Thereafter, the solution was gently stirred at 30 rpm to achieve the following coagulation.

In the coagulation phase, the iron flocs were formed. After 2-hr sedimentation, the

supernatant of treated effluent sample was collected. . Solution pH was not controlled

during the treatment.

In this study, dimensionless oxidant dose (DOD) was used to quantify ferrate(VI)

dose. DOD is defined as chemical equivalent ratio of the added oxidant to the initial COD

(COD0) as Eq.(3-1) (Jung et al. (in press)).

Page 46: Ferrate as a New Treatment Chemical for Removal of ...

31

𝐷𝑂𝐷 =𝐶ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑒𝑞𝑢𝑖𝑣𝑎𝑙𝑒𝑛𝑡 𝑜𝑓 𝑜𝑥𝑖𝑑𝑎𝑛𝑡

𝐶ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑒𝑞𝑢𝑖𝑣𝑎𝑙𝑒𝑛𝑡 𝑜𝑓 𝐶𝑂𝐷0 (3-1)

Equivalents of Fe(VI) and COD0 can be computed as follows.

𝐸𝑞𝑢𝑖𝑣𝑎𝑙𝑒𝑛𝑡 𝑜𝑓 𝐹𝑒(𝑉𝐼)𝑎𝑑𝑑𝑒𝑑 =𝐹𝑒(𝑉𝐼)

18.6 (3-2)

𝐸𝑞𝑢𝑖𝑣𝑎𝑙𝑒𝑛𝑡 𝑜𝑓 𝐶𝑂𝐷0

=𝐶𝑂𝐷0

8 (3-3)

Here, 18.6 and 8.00 are the equivalent weights (mg/meq) of Fe(VI) and COD,

respectively. Fe(VI) is ferrate(VI) mass concentration (mg/L as Fe). Considering that

COD0 is different from one secondary effluent sample to another, the use of DOD would

facilitate comparison of the treatment performances of an oxidation treatment technology

for different secondary effluent samples. Theoretically, the added oxidant just eliminates

all the COD at DOD = 1.00 as long as the four conditions are met: 1) the oxidant can

completely oxidize target pollutants; 2) electron transfer is the only chemical oxidation

mechanism; 3) co-existing chemical species do not compete with target pollutants for the

oxidant; and 4) the oxidant does not self-decay. The conversion of DOD and Fe(VI) mass

concentrations in this study is summarized in Table 3-1.

Table 3-1 Conversion of DOD and Fe(VI) mass concentrations

Fe(VI) (mg/L as Fe) DOD

0.00 0.00

7.7 0.10

19.2 0.25

38.4 0.50

76.8 1.00

153.6 2.00

Page 47: Ferrate as a New Treatment Chemical for Removal of ...

32

Analytical methods. Solution pH was measured with a pH meter (Thermo

Scientific Orion 5-Star Plus). All the contaminants excluding bacteria indicators were

analyzed prior to the filtration with 0.45µm syringe membrane filters (Millipore, nylon,

17 mm diameter). COD was measured colorimetrically following digestion (0.4–40 mg/L

range, HACH). UV254 absorbance was measured using a UV-Vis spectrophotometer

(HACH, DR 5000). Total phosphorus concentration was measured using PhosVer® 3

Phosphate Reagent Powder Pillows (HACH). E3 and OTC were measured using high

performance liquid chromatography (HPLC) (Li 2006). Total coliform and E.Coli. were

measured using IDEXX Colilert-18 test kits that are an EPA approved method for

simultaneous detection of total coliform and E.Coli. in water and wastewater. All the

experiments were run in triplicates.

(a)

Page 48: Ferrate as a New Treatment Chemical for Removal of ...

33

(b)

Fig. 3-1 Preliminary tests for ferrate(VI) treatment of secondary effluent: (a) jar tests;

and (b) secondary effluent after ferrate(VI) treatment and an overnight settling (

ferrate(VI)-induced sludge settled at the bottom).

3.3 Results and Discussion

3.3.1 Quality of secondary effluent

Quality of secondary effluent is shown in Table 3-2. The sample was an effluent

from an activated sludge process, which is the most common secondary treatment

practice in the United States (Tchobanoglous et al. 2003). As seen, the solution pH was at

a slightly alkaline condition. The sample contained EfOM with a COD range within 33.0-

34.4 mg/L. Total phosphorus (TP) varied between 4.84 and 5.19 mg/L as P. Two bacteria

Page 49: Ferrate as a New Treatment Chemical for Removal of ...

34

indicators were measured, including total coliform (2.46 × 105 MPN/100 mL) and E.Coli.

(3.29 ×104 MPN/100 mL).

Table 3-2. Basic parameters of secondary effluent

3.3.2 Treatment results

The removal of EfOM indicated as COD and UV254 absorbance is presented in

Fig. 3-2 and 3-3, respectively. As seen in Fig. 3-2, COD removal exhibited a biphasic

pattern. In the first phase, COD removal dramatically increased to 38% at pH 8.0 and

57% at pH 5.0, respectively, as the DOD increased to 0.5 and 1.0. This finding suggests

that ferrate(VI) rapidly reacted with EfOM through chemical oxidation and degraded

oxygen demanding materials. However, as the DOD further went up to 2.0, the increase

in COD removal was almost marginal, regardless of pH, suggesting that remaining

organic matters were recalcitrant to ferrate(VI) oxidation. As shown in Fig. 3-3, UV254

removals went up to 42% and 38% at pH 8.0 and 5.0, respectively, with the increasing

DOD from 0.0 to 0.5. At the two pH levels, the profiles of UV254 removal with DOD

were not obviously different at the low DOD range. As DOD further increased to 1.0, the

Parameters Values

pH 7.31-7.61

COD (mg/L) 33.0-34.4

UV254 (cm-1) 0.115-0.130

TP (mg/L as P) 4.84-5.19

NO3-N (mg/L) 4.1

NH3-N (mg/L) 20.2

Total coliform (MPN/100 mL) 2.46 × 105

E. coli (MPN/100 mL) 3.29 ×104

Page 50: Ferrate as a New Treatment Chemical for Removal of ...

35

UV254 removals reached to 49% and 60% at pH 8.0 and 5.0, respectively. However, the

increase of UV254 removal did not augmented as DOD further increased.

Different from unselective free radical oxidation in advanced oxidation processes,

ferrate(VI)-driven oxidation is selective. Ferrate(VI) is more reactive with electron-rich

moieties (ERMs) (Lee and von Gunten 2010, Yang et al. 2012), which are the functional

groups that release electron density to neighboring atoms from themselves via resonance

or inductive effects. However, ferrate(VI) poorly reacts with electron withdrawing

moieties (EWMs) that have an opposite effect, drawing electron density from

neighboring atoms to themselves. In this study, the similarity in the removal profiles of

COD and UV254 (UV254 indicates the abundance of aromatic structures and double bonds)

with DOD reflects the preferential reactivity of ferrate(VI). For example, at pH 5.0, as

DOD increased from 0.5 to 1.0, COD removal was increased from 40% to 58%, while the

UV254 removal went up from 38% to 60%. The slightly increased removal in COD or

UV254 at a high DOD range indicates that ERMs available to Fe(VI) oxidation became

less as the oxidation proceeded and residual organic molecules became more recalcitrant.

Moreover, the higher EfOM removal observed at an acidic condition suggest that HFeO4-,

the dominant ferrate(VI) species below pH 7.3, was more reactive with EfOM than

FeO42- that is a prevailing ferrate(VI) species above pH 7.3.

Page 51: Ferrate as a New Treatment Chemical for Removal of ...

36

Fig. 3-2 COD removal under different DOD and initial pH (COD0 = 33.0 mg/L for pH

8.0 and 34.4 mg/L for pH 5.0)

Fig. 3-3 UV254 removal under different DOD and initial pH (initial UV254 = 0.115 cm-1

for pH 8.0 and 0.130 cm-1 for pH 5.0)

0%

10%

20%

30%

40%

50%

60%

70%

0.0 0.5 1.0 1.5 2.0 2.5

CO

D r

emoval

(%)

DOD

pH 8.0

pH 5.0

0%

10%

20%

30%

40%

50%

60%

70%

0.0 0.5 1.0 1.5 2.0 2.5

UV

25

4re

moval

(%)

DOD

pH 8.0

pH 5.0

Page 52: Ferrate as a New Treatment Chemical for Removal of ...

37

The removal efficiencies of TP at different chemical doses and pH are shown in

Fig. 3-4. As DOD increased from 0.0 to 1.0, the TP removals were substantially

increased to 98% and 99% at pH 8.0 and 5.0, respectively. Within the chemical dose

range, TP removal was greater at pH 8.0 than at pH 5.0 at an specific dose. For example,

at DOD = 0.5, the TP removal at pH 8.0 was 79%, which was higher than 55% at pH 5.0.

Generally speaking, phosphate can be removed from water through the formation of Fe-P

precipitates and the adsorption to iron hydroxides (Tchobanoglous et al. 2003, Crittenden

et al. 2012). Iron has the lowest solubility around pH 8.0 (Stumm 1992). The observation

on the pH effect on TP removal is likely due to higher iron solubility at pH 5.0 that

disfavors the formation of Fe-P precipitates and/or iron hydroxides. Of note, the TP

removal was roughly linearly correlated with ferrate(VI) dose at a low DOD range of 0.0-

0.5 (i.e. 0.00-38.4 mg/L Fe(VI)): TP removed = 0.23 Fe(VI) , R2 = 0.76, for pH 8.0; TP

removed = 0.20 Fe(VI) , R2 = 0.98, for pH 5.0. Here, TP removed is the mass

concentration of removed TP (mg/L as P); and Fe(VI) is the mass concentration of

ferrate(VI) (mg/L as Fe).

Typically, raw wastewater contains 5.0 – 20.0 mg/L total P (TP), particulate or

dissolved, of which a majority is reactive P (orthophosphate), and the rest is nonreactive

P (acid hydrolysable and organic P) not readily removed by conventional secondary

treatment (Tchobanoglous et al., 2003). Unfortunately, the conventional secondary

treatment only removes 1.0-2.0 mg/L TP. Consequently, a majority of phosphorous

remains in secondary effluent. Although P may be beneficial for some water reuse

purpose (e.g. agricultural reuse), it is problematic in many reuse scenarios. Firstly, excess

Page 53: Ferrate as a New Treatment Chemical for Removal of ...

38

P can cause severe eutrophication when reclaimed water is for environmental reuse and

the receiving water bodies are environmentally sensitive (e.g. wetland restoration). For

example, the State of Florida requires TP in the reclaimed water for environmental reuse

below 1.0 mg/L (US EPA, 2012). Secondly, when reverse osmosis filtration (the most

common water reclamation technology) is applied, P can cause inorganic membrane

fouling due to the formation of calcium phosphate as well as bio-fouling because P as a

nutrient can support microbial growth. The most common P removal practices employ

enhanced biological phosphorus removal (EBPR) and/or chemical precipitation.

However, these methods only remove reactive P, but are ineffective for nonreactive P.

Here, ferrate(VI) treatment provides a promising alterative for the traditional options.

Although ferrate(VI) application typically targets at other pollutants in water reclamation,

the P removal provides a secondary benefit for ferrate(VI) treatment.

When Fe(VI) was reduced to Fe(III), the in-situ formed Fe(III) had a potential to

remove phosphate from water. Two mechanisms might be involved. The first one is the

direct precipitation between Fe(III) and phosphate to produce solid ferric phosphate. This

mechanism has been long recognized as the major pathway to remove orthophosphate

from wastewater with ferric salt addition, in which Fe(III) is dosed in a single step

(Crittenden et al., 2012). And the other mechanism is the adsorption of phosphate to the

produced iron hydroxide, which is gradually produced with the Fe(VI) reduction.

Adsorption of phosphate to iron oxides is ascribed to surface complexation theory

(Stumm 1992). The relevant reactions include the complexation between phosphate and

surface iron ions as well as the protonation reactions of adsorbed phosphate.

Page 54: Ferrate as a New Treatment Chemical for Removal of ...

39

Consequently, three surface phosphate complexes are likely formed, including ≡Fe-O-

PO3H2, ≡Fe-O-PO3H-, and ≡Fe-O-PO3

2-. Thermodynamic and spectrophonic

experimental observations support that these are inner-sphere surface complexation

reactions (Persson et al. 1996).

Page 55: Ferrate as a New Treatment Chemical for Removal of ...

40

Fig. 3-4 TP removal under different DOD and initial pH (TP0 = 4.84 mg/L for pH 8.0 and

5.19 mg/L for pH 5.0)

0%

20%

40%

60%

80%

100%

120%

0.0 0.5 1.0 1.5 2.0 2.5

TP

rem

oval

(%)

DOD

pH 8.0

pH 5.0

Page 56: Ferrate as a New Treatment Chemical for Removal of ...

41

The log removal of bacterial indicators during ferrate(VI) treatment of secondary

effluent at pH 8.0 is shown in Fig. 3-5(a) and (b). Fe(VI) dose was selected at its low and

high levels (DOD = 0.1 and 3.0, i.e. 7.7 and 153.6 mg/L, respectively). The initial total

coliform and E.Coli. concentrations were 2.46 × 105 and 3.29 × 104 MPN/100 mL,

respectively. Here the log removal is defined as below (Tchobanoglous et al. 2003,

Crittenden et al. 2012).

𝐿𝑜𝑔 𝑟𝑒𝑚𝑜𝑣𝑎𝑙 = − log(𝑟𝑒𝑚𝑎𝑖𝑛𝑖𝑛𝑔 𝑏𝑎𝑐𝑡𝑒𝑟𝑖𝑎 𝑐𝑜𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛

𝑖𝑛𝑡𝑖𝑎𝑙 𝑏𝑎𝑐𝑡𝑒𝑟𝑖𝑎 𝑐𝑜𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛) = − log(1 −

𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑎𝑡𝑒𝑑 𝑏𝑎𝑐𝑡𝑒𝑟𝑖𝑎 𝑐𝑜𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛

𝑖𝑛𝑡𝑖𝑎𝑙 𝑏𝑎𝑐𝑡𝑒𝑟𝑖𝑎 𝑐𝑜𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛) (3-4)

Log removal has been widely used to evaluate the treatment performance of a

disinfection process. As shown in Fig. 3-5(a), the log removal of total coliform increased

to 0.96 within the first minute. Thereafter, the log removal slightly increased to 1.20 (i.e.

93.7% inactivation) at 10 min. Meanwhile, the log removal of E.Coli. rapidly reached

1.61 within the first minute and then gradually went up to 2.52 (i.e. 99.7% inactivation)

at 10 min. As DOD was increased to 3.0, the log removals of total coliform and E.Coli.

were promptly increased to 2.37 (i.e. 99.6% inactivation) and 2.34 (i.e. 99.5 inactivation),

respectively. Afterwards, there was not an obvious variation for the removal of total

coliform or E.Coli. Higher inactivation efficiencies were achieved at higher chemical

doses, because a higher ferrate(VI) dose could achieve a greater oxidant exposure, which

is commonly used as a major operating parameter to control a disinfection process.

The concentration of E. coli was reduced by ferrate(VI) to 100 MPN/100 mL at

10-min disinfection at DOD=0.1 (7.7 mg/L as Fe). This result meets with the non-food

crop and processed food crops agricultural reuse criteria, restricted impoundment criteria,

and industrial reuse criteria in Texas (Type II) and Virginia (Level 2). Previous studies

Page 57: Ferrate as a New Treatment Chemical for Removal of ...

42

showthat ferrate(VI) is capable of inactivating indicator organisms (Cho et al., 2006;

Gombos et al., 2012; Jiang et al., 2006; Jiang et al., 2007), viruses (Schink and Waite,

1980), and bacteriophages (Hu et al., 2012; Kazama et al., 1994). These studies have

demonstrated that ferrate disrupts cell membranes and that it may cause some genome

damage in bacteriophage (Hu et al., 2012). Therefore, ferrate(VI)-induced cell damage is

a plausible reason to inactivate E. Coli. The damage may be caused by ferrate(VI)

oxidation of certain substances on cell membrane or DNA, such as sulfhydryl and sulfur

bonds in peptides.

Page 58: Ferrate as a New Treatment Chemical for Removal of ...

43

(a)

(b)

Fig. 3-5 Log removal of bacteria indicators during ferrate(VI) treatment of secondary

effluent: (a) DOD = 0.1; and (b) DOD = 3.0 (pH = 8.0, initial total coliform = 2.46 × 105

MPN/100 mL, E.Coli. = 3.29 × 104 MPN/100 mL)

0.00

0.50

1.00

1.50

2.00

2.50

3.00

0.0 2.0 4.0 6.0 8.0 10.0

Log r

emoval

Time (min)

Total coliform

E. Coli.

0.00

0.50

1.00

1.50

2.00

2.50

3.00

0.0 1.0 2.0 3.0 4.0 5.0

Log r

emoval

Time (min)

Total coliform

E. Coli.

Page 59: Ferrate as a New Treatment Chemical for Removal of ...

44

Ferrate(VI) removals of two selected emerging contaminants, i.e. E3 and OTC, at

pH 8.0 and 5.0 are shown in Fig. 3-6 (initial E3 = 3.83 mg/L; and initial OTC = 2.00

mg/L). DOD varied between 0.0 and 1.0. At DOD = 0.1 (i.e. 7.7 mg/L Fe(VI)), the

removal efficiencies of E3 and OTC reached 94% and 97% at pH 8.0 and 98% and 99%

at pH 5.0, respectively. For either contaminant, the slightly higher removal efficiency at

pH 5.0 was due to the higher reactivity of ferrate(VI) at an acidic condition than at an

alkaline environment. As DOD was increased to 0.25 or greater, the removal reached

over 99%, regardless of solution pH. These findings demonstrated the high reactivity of

ferrate(VI) toward E3 and OTC. In water, two ferrate(VI) species, i.e. HFeO4- and FeO4

2-,

have a potential to react with E3 or OTC in its dissociated and undissociated states. For

example, it is found that HFeO4- is more reactive for E3 in the both dissociated and

undissociated states, while dissociated E3 is more reactive in the reactions with Fe(VI)

than undissolved E3 (Li and Gao 2009).

The primary goal of the set of experiments was to examine the treatment feasibility

of ferrate(VI) for removal of certain micro-pollutants in secondary effluent, but not to

fundamentally investigate the interactions between ferrate(VI) and the target micro-

pollutants. These findings in Fig.3-6 were due to two reasons: 1) the both compounds

were highly reactive toward ferrate(VI) oxidation; and 2) the ferrate(VI) doses were

sufficiently high to effectively remove the compounds. The lowest ferrate(VI) dose was

DOD = 0.1 (i.e. 138.5 µM Fe(VI)), while E3 was 3.83 mg/ (i.e. 13.3 µM) and OTC was

2.00 mg/L (i.e. 4.35 µM). Hence, [Fe(VI)]:[E3] = 10.4:1 and [Fe(VI)]:[E3] = 31.9:1. Of

note, the lowest ferrate(VI) dose (138.5 µM, i.e. 7.70 mg/LFe(VI)) fell within a typical

Page 60: Ferrate as a New Treatment Chemical for Removal of ...

45

ferrate(VI) dose range for wastewater treatment. Therefore, the promising findings

showed that ferrate(VI) appeared to be effective for elimination of these unwanted micro-

pollutants as long as it is capable of degrading them. The study regarding the reactivity of

ferrate(VI) toward different functional groups was out of scope in this task. But a recent

study (Yang et al., 2012) provides a deep insight into what endocrine disrupting

compounds and PPCPs ferrate(VI) can chemically oxidize in a secondary effluent matrix.

A further study on ferrate(VI) degradation of a model micro-pollutant (MEF) would be

investigated in Chapter 5.

Oxidation results of traceable micro-pollutants obtained from bench scale tests

using deionized water are likely discounted in a real wastewater matrix because other

water matrix constituents, which have much higher concentrations than these micro-

pollutants, may scavenge the oxidant to “protect” the traceable micro-pollutants of

concern. Such a water matrix constituent in secondary effluent is EfOM. EfOM is present

in a secondary effluent with a typical COD at a few tens of mg/L. However, the

preliminary tests showed that ferrate(VI), even at a low dose, was sufficient to effectively

remove E3 or OTC, regardless of solution pH. Therefore, ferrate(VI) oxidation appeared

to be a potentially powerful process to address wastewater-derived emerging

contaminants in secondary effluent.

Page 61: Ferrate as a New Treatment Chemical for Removal of ...

46

(a)

(b)

Fig. 3-6 Removal of E3 and OTC during ferrate(VI) treatment of secondary effluent: (a)

pH 8.0; and (b) pH 5.0 (initial E3 = 3.83 mg/L; and initial OTC = 2.00 mg/L)

0%

20%

40%

60%

80%

100%

120%

0.0 0.2 0.4 0.6 0.8 1.0

Rem

ov

al

DOD

E3

OTC

0%

20%

40%

60%

80%

100%

120%

0.0 0.2 0.4 0.6 0.8 1.0

Rem

ov

al

DOD

E3

OTC

Page 62: Ferrate as a New Treatment Chemical for Removal of ...

47

To better understand the treatment potential of ferrate(VI) technology, results

from the preliminary experiments are compared with the guidelines for water reuse. For

EfOM, BOD5, rather than COD, has been used in different guidelines. For example,

BOD5 ≤, 30 mg/L for restricted urban reuse, environmental reuse, or industrial reuse (US

EPA, 2012). In this study, the lowest ferrate dose of 7.00 mg/L (DOD = 0.10) reduced

COD to 25.3 mg/L at pH 8.0 and 24.3 mg/L at pH 5.0. Because BOD5 is part of COD,

which indicates the 5-day biochemical degradable fraction of chemical oxygen demand,

the lowest ferrate(VI) dose sufficiently meets with the reuse guideline levels. TP is also

required for water reuse in some U.S. states. For example, TP ≤ 1 mg/L for

environmental reuse in Florida, North Carolina, and Washington (US EPA, 2012).

Preliminary results in this study show that a ferrate(VI) dose of 19.2 mg/L (DOD = 0.25)

is capable of removing TP below 1.00 mg/L from 4.84 mg/L, showing that ferrate(VI) is

an effective treatment for phosphorus in water reclamation. On the other hand, emerging

micro-pollutants are not regulated in water reuse. However, the almost complete removal

of E3 and OTC by ferrate(VI) suggests that ferrate(VI) technology is a reliable barrier for

these traceable contaminants in water.

Page 63: Ferrate as a New Treatment Chemical for Removal of ...

48

3.4 Conclusions

Preliminary tests were performed to evaluate the treatability of ferrate(VI) for

different contaminants in secondary effluent. The major findings in this study are

summarized as below.

1) Ferrate(VI) can chemically oxidize EfOM, thereby reducing COD and UV254. The

substantial reduction in UV254 suggests that ferrate(VI) favorably reacts with UV-

absorbing moieties in EfOM molecules, which are associated with aromatic

structures and double bonds. The treatment characteristic is likely related to the

selectivity of ferrate(VI) oxidation, because Fe(VI) more tends to react with

ERMs.

2) Ferrate(VI) also effectively removes phosphorus in secondary effluent through

transformation of dissolved P into particulate P.

3) Ferrate(VI) is an effective disinfecting agent. In this dissertation, it achieved a

1.20 log inactivation for total coliform and a 2.52 log removal for E. Coli. within

10 min at a dose of 7.7 mg/L Fe(VI).

4) Ferrate(VI) removal of certain micro-pollutants is rapid and sufficient. E3 and

OTC in this study was removed by 98% at a low ferrate(VI) dose of 7.7 mg/L.

5) Ferrate(VI)-driven oxidation was pH dependent. Generally, higher removal of

COD, UV254, bacteria indictors , E3 and OTC was achieved at an acidic pH

condition than at an alkaline environment, because HFeO4- (the dominant

ferrate(VI) species at pH < 7.30) was more reactive than FeO42- (the dominant

ferrate(VI) species at pH > 7.30).

Page 64: Ferrate as a New Treatment Chemical for Removal of ...

49

These findings suggest that ferrate(VI) is a promising agent to directly treat

secondary effluent for water reuse. It can concurrently and efficiently remove different

pollutants, including EfOM, P, bacteria indicators, and emerging micro-pollutants, in a

single dose. Solution pH and ferrate(VI) dose are two important operating parameters

affecting the treatment results.

Page 65: Ferrate as a New Treatment Chemical for Removal of ...

50

3.5 References

Alsheyab, M., Jiang, J.Q. and Stanford, C. (2009) On-line production of ferrate with an

electrochemical method and its potential application for wastewater treatment - A review.

Journal of Environmental Management 90 (3), 1350-1356.

Cho, M.; Lee, Y.; Choi, W.; Chung, H.; Yoon, J., (2006) Study on Fe(VI) species as a

disinfectant: Quantitative evaluation and modeling for inactivating Escherichia coli.

Water Research, 40 (19), 3580-3586.

Crittenden, J.C., Trussell, R.R., Hand, D.W., Howe, K.J. and Tchobanoglous, G. (2012)

MWH's Water Treatment: Principles and Design, John Wiley & Sons.

Dodd, M. C., (2012) Potential impacts of disinfection processes on elimination and

deactivation of antibiotic resistance genes during water and wastewater treatment. Journal

of Environmental Monitoring , 14 (7), 1754-1771.

Ghernaout, D. and Naceur, M.W. (2011) Ferrate(VI): In situ generation and water

treatment - A review. Desalination and Water Treatment 30 (1-3), 319-332.

Gombos, E., Felföldi, T., Barkács, K., Vértes, C., Vajna, B. and Záray, G., (2012)

Ferrate treatment for inactivation of bacterial community in municipal secondary

effluent. Bioresour. Technol. 107, 116-121.

Hu, L., Page, M. A., Sigstam, T., Kohn, T., Mariñas, B. J. and Strathmann, T. J., (2012)

Inactivation of Bacteriophage MS2 with Potassium Ferrate(VI). Environ. Sci. Technol.

46, (21), 12079-12087.

Jiang, J.Q., Panagoulopoulos, A. and Bauer, M.; Pearce, P., (2006) The application of

potassium ferrate for sewage treatment. J. Environ. Manag. 79, (2), 215-220.

Jiang, J.Q., Wang, S. and Panagoulopoulos, A., (2007) The role of potassium ferrate(VI)

in the inactivation of Escherichia coli and in the reduction of COD for water remediation.

Desalination, 210, (1–3), 266-273.

Jiang, J.Q. (2007) Research progress in the use of ferrate(VI) for the environmental

remediation. Journal of Hazardous Materials 146(3), 617-623.

Jiang, Y., Goodwill, J.E., Tobiason, J.E. and Reckhow, D.A. (2015) Effect of different

solutes, natural organic matter, and particulate Fe (III) on ferrate (VI) decomposition in

aqueous solutions. Environmental Science & Technology 49(5), 2841-2848.

Page 66: Ferrate as a New Treatment Chemical for Removal of ...

51

Jung, C., Deng, Y., Zhao, R. and Torrens, K. ((in press)) Chemical oxidation for

mitigation of UV-quenching substances (UVQS) from municipal landfill leachate: Fenton

process versus ozonation. Water Research.

Lee, Y. and von Gunten, U. (2010) Oxidative transformation of micropollutants during

municipal wastewater treatment: Comparison of kinetic aspects of selective (chlorine,

chlorine dioxide, ferrate(VI), and ozone) and non-selective oxidants (hydroxyl radical).

Water Research 44(2), 555-566.

Li, C. (2006) Mechanisms and performance of potassium ferrate in endocrine disrupting

chemicals, The Hong Kong Polytechnic University.

Li, C. and Gao, N. (2009) Modeling the aqueous reaction kinetics of estriol with ferrate.

Frontiers of Chemical Engineering in China 3(1), 39-45.

Kazama, F. (1994) Inactivation of coliphage Qβ by potassium ferrate. FEMS Microbiol.

Lett, 118, (3), 345-349.

Persson, P., Nilsson, N. and Sjöberg, S. (1996) Structure and bonding of orthophosphate

ions at the iron oxide–aqueous interface. Journal of Colloid and Interface Science 177

(1), 263-275.

Ramseier, M. K., von Gunten, U., Freihofer, P. and Hammes, F., (2011) Kinetics of

membrane damage to high (HNA) and low (LNA) nucleic acid bacterial clusters in

drinking water by ozone, chlorine, chlorine dioxide, monochloramine, ferrate(VI), and

permanganate. Water Res., 45, (3), 1490-1500.

Schink, T. and Waite, T. D., (1980) Inactivation of f2 virus with ferrate (VI). Water Res.,

14 (12), 1705-1717.

Sharma, V.K. (2002) Potassium ferrate(VI): an environmentally friendly oxidant.

Advances in Environmental Research 6(2), 143-156.

Stumm, W. (1992) Chemistry of the solid-water interface, Wiley, New York.

Tchobanoglous, G., Franklin, L. and Stensel, H.D. (2003) Wastewater engineering:

treatment and reuse. Metcalf & Eddy, Inc., New York.

US EPA (2012) Guidelines for Water Reuse

Waite, T.D. (2012) Chemistry-On-Site Production of Ferrate for Water and Wastewater

Purification. American Laboratory 44(10), 26.

Page 67: Ferrate as a New Treatment Chemical for Removal of ...

52

Yang, B., Ying, G.G., Zhao, J.L., Liu, S., Zhou, L.J. and Chen, F. (2012) Removal of

selected endocrine disrupting chemicals (EDCs) and pharmaceuticals and personal care

products (PPCPs) during ferrate(VI) treatment of secondary wastewater effluents. Water

Research 46(7), 2194-2204.

Zhou, S., Shao, Y., Gao, N., Zhu, S., Li, L., Deng, J. and Zhu, M., (2014) Removal of

Microcystis aeruginosa by potassium ferrate (VI): Impacts on cells integrity, intracellular

organic matter release and disinfection by-products formation. Chem. Eng. J., 251, 304-

309.

Page 68: Ferrate as a New Treatment Chemical for Removal of ...

53

CHAPTER 4 FERRATE(VI) REACTIONS WITH EFFLUENT

ORGANIC MATTER

4.1 Introduction

EfOM is a mixture of complex organic components (Liu and Fang 2002). Its

composition and negative impacts on water reclamation have been discussed in Chapter

1. As a major water matrix constituent, it inevitably reacts with ferrate(VI) that is added

into secondary effluent. The interactions between ferrate(VI) and EfOM include

ferrates(VI) consumption due to Fe(VI) reactions with EfOM as well as the EfOM

degradation. It is technically difficult to investigate the effect of EfOM alone on

ferrate(VI) decomposition because EfOM cannot be extracted from secondary effluent

and EfOM with a high purity grade, unlike NOM, is not commercially available. The

ferrate(VI) decomposition in secondary effluent is an overall effect of ferrate(VI) self-

decay and reactions with certain reducing agents, of which EfOM was a major

component. Studies on ferrate(VI) treatment of secondary effluent were reported

previously, but the information on ferrate decomposition kinetics in a secondary effluent

matrix is very limited (Lee and von Gunten 2010). Lee and von Gunten (2010) compared

the decomposition of four oxidants (40-45 µM) in a secondary effluent at pH 8.0. Results

showed that all the oxidants followed a 2nd order reaction pattern. Among the four

oxidants, ferrate(VI) decomposition was the slowest. The oxidation exposure of the four

oxidants followed the order: ferrate(VI) (4.5 × 10-2 M ) ≈ chlorine dioxide > chlorine (3.6

× 10-2 M-1 s-1) > ozone (6.0 × 10-4 M-1 s-1). Ferrate(VI) stability is beneficial because this

Page 69: Ferrate as a New Treatment Chemical for Removal of ...

54

may maintain a greater oxidant exposure for better oxidation of target pollutants.

Moreover, ferrate(VI) decomposition was very distinct from the decay of the other three

oxidants, which exhibited a biphasic behavior – a rapid decay at the onset followed by a

slow decomposition. In contrast, ferrate(VI) concentration decreased smoothly over the

reaction time.

A variety of treatment methods for reduction of EfOM in secondary effluent have

been intensively studied, including chemical coagulation, activated carbon adsorption and

advanced oxidation processes (Shon et al. 2006, Michael-Kordatou et al. 2015).

Coagulation can largely remove particulate matters in secondary effluent, thereby being

an effective method for the alleviation of particulate organic matter. However, the

treatment capability of coagulation or dissolved EfOM is very limited. Therefore, it is

often used as a pre-treatment prior to other treatments (e.g. activated carbon adsorption or

RO filtration). Activated carbon adsorption has proven to remove significant quantities

of dissolved EfOM from secondary effluent (Michael-Kordatou et al. 2015, Gur-Reznik

et al. 2008). For example, Gur-Reznik et al. (2008) removed 80-90% dissolved EfOM

from secondary effluent using granular activated carbon for mitigation of the downstream

RO fouling. Activated carbon favorably removes hydrophobic and biodegradable EfOM

and low-medium MW EfOM molecules with low SUVA (Shon et al. 2004). However,

activated carbon is particularly costly. Moreover, economical regeneration methods of

spent carbon need to be considered in practices to save operating costs. AOPs are another

option for removal of EfOM. Various AOPs are available for the application purpose,

such as the Fenton treatment, ozone-based AOPs (e.g. O3/H2O2 and O3/UV), UV-based

Page 70: Ferrate as a New Treatment Chemical for Removal of ...

55

AOPs (e.g. TiO2/UV), and emerging sulfate radical –based AOPs (e.g. heated activated

persulfate). Hydroxyl or sulfate radicals are among the strongest oxidizing agents with a

reduction potential over 2.80 V. They can rapidly degrade EfOM molecules and even

mineralize certain EfOM into water and carbon dioxide. A DOC removal of 20 - 90% has

been reported for free radical oxidation of EfOM during AOPs (Ito et al. 1998, Shon et al.

2003). However, AOPs are typically energy-intensive. And the free radicals are less

efficiently utilized because they unselectively react with many non-target constituents in

water to cause a chemical waste.

The objective of this study is to elucidate the interactions of ferrate(VI) and

EfOM in a secondary effluent matrix. Ferrate(VI) decomposition kinetics in a secondary

effluent was firstly studied. Afterwards, EfOM degradation products after ferrate(VI)

treatment were characterized to understand the reactivity of EfOM toward ferrate(VI)

oxidation.

4.2 Materials and Methods

Chemicals and Reagents. Potassium ferrate (K2FeO4, 96% purity) and all other

chemicals (reagent grade) were purchased from Sigma-Aldrich (St. Louis, MO, US) or

Fisher-Scientific (Pittsburgh, PA, US). Secondary effluent was collected from a

secondary clarifier prior to disinfection at a local municipal wastewater treatment plant

using activated sludge treatment (Elizabeth, NJ). Once collected, the sample was

delivered to Montclair State University’s water treatment laboratory and stored at 4oC in

a refrigerator until use. A concentrated ferrate(VI) stock solution (200 mg/L Fe(VI)) was

prepared by dissolving an appropriate mass of K2FeO4 in deionized water. The pH of this

Page 71: Ferrate as a New Treatment Chemical for Removal of ...

56

stock solution was around 9.0 at which Fe(VI) was relatively stable. The Fe(VI)

concentration was confirmed with the ABTS method(Lee et al. 2005b). A concentrated

ferric stock solution (1,000 mg/L Fe(III)) was prepared by dissolving an appropriate

amount of ferric chloride salt in deionized water. The Fe(VI) and Fe(III) stock solutions

were freshly prepared prior to use.

Fe(VI) and Fe(III) treatment tests. Ferrate(VI) treatment tests were performed in

600 mL glass beakers with 200 mL secondary effluent on a six-paddle programmable jar

tester (Phipps & Bird - 7790-950). The treatment was initiated through the addition of an

aliquot of K2FeO4 from the stock solution to the secondary effluent. Within the first 3

minutes, the solution was rapidly mixed at 150 rpm to completely disperse the added

ferrate(VI). The solution was gently stirred at 30 rpm in the following 60 minutes during

which Fe(VI) was completely decomposed. A low mixing speed during the slow mixing

also prevented the produced iron flocs from destruction. For the kinetics tests, 2 mL of

sample was collected at designated sampling times and then filtered through 0.45 µm

membrane for measurement of residual ferrate(VI). Thereafter, a 60-min settling cycle

started to allow for sedimentation of the large iron flocs. Finally, 100 mL supernatant was

collected for analysis. Control tests were performed with Fe(III) under the identical

conditions to understand the behavior of Fe(III)-based coagulation alone in the removal

of EfOM. The initial solution pH was 7.5. During the treatment, pH was not controlled

but monitored over time. The final pH was not beyond 8.10.

Analytical methods. Solution pH was measured with a pH meter (Thermo

Scientific Orion 5-Star Plus). Turbidity was determined with a portable turbidity meter

Page 72: Ferrate as a New Treatment Chemical for Removal of ...

57

(HACH, 2100Q). EfOM in the supernatant samples was analyzed prior to the filtration

with 0.45µm syringe membrane filters (Millipore, nylon, 17 mm diameter). COD was

measured colorimetrically using a set of COD test kits (0.4–40 mg/L ultralow range,

HACH). UV254 absorbance was measured using a UV-Vis spectrophotometer (HACH,

DR 5000). EfOM was sequentially fractionated using a stirred cell (Millipore, Model

8200) in terms of their molecular weight (MW) with 10 and 1 kDa UF membranes. Two

litters of Milli-Q water (18.2 MΩ/cm) passed through the UF membrane filters before

use. Fluorescence excitation-emission matrix (EEM) analyses were conducted with a LS-

55 fluorescence spectrometer (Perkin-Elmer, Norwalk, Connecticut, USA) under the

excitation wavelength of λex = 200–460 nm at 5 nm increments across an emission range

of λem = 240–590 nm at 2 nm intervals. Excitation and emission slit widths were set to 2.5

nm with a photomultiplier tube voltage of 800 V. All the experiments were run in

triplicates. Their relative standard deviations were below 5% (not shown in figures).

4.3 Results and Discussion

4.3.1 Ferrate(VI) decomposition kinetics in secondary effluent

Ferrate(VI) decomposition kinetics data at different ferrate(VI) doses is show in

Fig. 4-1(a). The ferrate(VI) decomposition exhibited a biphasic reaction pattern, having

a dramatic decrease at the initial phase followed by a subsequent slow decay. The

reaction times for Fe(VI) decrease by 90% were 400, 450, 600, and 650 s for 5.00, 10.00,

15.00, and 20.00 mg/L Fe(VI) (i.e., 0.09, 0.18, 0.27 and 0.36 mM), respectively. Two

kinetics models, i.e. 1st order and 2nd order reaction equations, were fitted with

experimental data, as shown in Fig. 4-2(a) and (b). At any specific ferrate(VI) dose,

Page 73: Ferrate as a New Treatment Chemical for Removal of ...

58

Fe(VI) decomposition well followed a 2nd order reaction pattern (R2 > 0.98) rather than a

1st order reaction pattern. Very clearly, as shown in Fig. 4-2(a), the experimental data did

not exhibit a linear relationship between ln(C/C0) and time for the 1st order fitting. The

observed 2nd order rate constants (k) at 0.09, 0.18, 0.27 and 0.36 mM Fe(VI) (i.e. 5.00,

10.00, 15.00 and 25.00 mg/L Fe(VI)) were 0.2239, 0.1030, 0.0475, and 0.0314 mM-1 s-1,

respectively.

Generally speaking, ferrate(VI) decomposition is due to two pathways: (1) 2nd order

reaction for homogenous self-decomposition; and (2) 1st order reaction for ferrate(VI)

reactions with different water matrix constituents (Lee et al. 2004). Comparison of

ferrate(VI) decay in secondary effluent and DI water is shown in Fig. 4-1(b) (Fe(VI) =

5.00 mg/.L). Although ferrate(VI) might react with different reducing agents in the

secondary effluent, the major ferrate(VI) sink, excluding self-decomposition, was EfOM

because its concentration was much greater than others (e.g. nitrite). In Fig. 4-1(b),

ferrate(VI) decay in DI water showed Fe(VI) consumption due to self-decomposition

alone (the diamond symbols), while the ferrate(VI) decomposition in secondary effluent

represented Fe(VI) consumption due to both self-decomposition and the reactions with

reducing agent (the majority was EfOM) (the square symbols). Therefore, the difference

between two curves was primarily caused by Fe(VI) reaction with EfOM. As shown, the

EfOM accounted for a significant fraction of Fe(VI) decomposition, particularly in the

initial phase. For example, at a ferrate(VI) dose of 5.00 mg/L, the Fe(VI) consumptions

due to self-decomposition and reactions with EfOM were 0.70 and 2.45 mg/L,

respectively, within the 1st min. This finding suggests that EfOM rapidly reacted with

Page 74: Ferrate as a New Treatment Chemical for Removal of ...

59

ferrate(VI) at the onset of treatment. Thereafter, the Fe(VI) consumption due to EfOM

was not obviously increased with time, suggesting that ferrate(VI) decomposition was not

greatly influenced by EfOM in the following phase. The data of k against ferrate(VI)

dose is presented in Fig. 4-3. As seen, k considerably dropped from 0.2239 to 0.0314

mM-1 s-1 with the increasing ferrate(VI) dose from 0.09 to 0.36 mM. The correlation

between the 2nd order rate constant and ferrate(VI)_dose was not linear. As the

ferrate(VI) dose went up, the extent of the decrease in the rate constant was reduced. The

kinetics parameter including reaction order, rate constant, and the relationship between k

and ferrate(VI) dose are important to determine Fe(VI) lifetime and exposure in a

secondary effluent matrix.

Page 75: Ferrate as a New Treatment Chemical for Removal of ...

60

(a)

(b)

Fig. 4-1 Residual ferrate(VI) at different reaction times (pH = 7.5): (a) ferrate(VI)

decay at different initial ferrate(VI) doses in secondary effluent; and (b) comparison of

ferrate(VI) decay in secondary effluent and DI water (5.00 mg/L Fe(VI))

0.00

5.00

10.00

15.00

20.00

0 500 1000 1500 2000

Res

idu

al

Fe(

VI)

(m

g/L

)

Time (s)

5.00 mg/l Fe(VI)

10.00 mg/l Fe(VI)

15.00 mg/l Fe(VI)

20.00 mg/l Fe(VI)

0.00

1.00

2.00

3.00

4.00

5.00

6.00

0 200 400 600 800 1000

Res

idu

al

Fe(

VI)

(m

g/L

)

Time (s)

secondary effluent

DI water

Page 76: Ferrate as a New Treatment Chemical for Removal of ...

61

(a)

(b)

Fig. 4-2 Experimental data vs. model data for ferrate(VI) decomposition kinetics (pH

= 7.5): (a) 1st order reaction; and (b) 2nd order reaction

-3.500

-3.000

-2.500

-2.000

-1.500

-1.000

-0.500

0.000

0 500 1000 1500 2000

ln(C

/C0)

Time (s)

0.09 mM

0.18 mM

0.27 mM

0.36 mM

1/C-1/C0 = 0.2239t

R² = 0.99

1/C-1/C0 = 0.103t

R² = 0.99

1/C-1/C0 = 0.0475t

R² = 0.98

1/C-1/C0 = 0.0314t

R² = 0.990.000

20.000

40.000

60.000

80.000

100.000

120.000

140.000

0 500 1000 1500 2000

1/C

-1/C

0

Time (s)

0.09 mM 0.18 mM0.27 mM 0.36 mM

Page 77: Ferrate as a New Treatment Chemical for Removal of ...

62

Fig. 4-3 2nd order reaction rate constant (k) at different ferrate(VI) doses

Fe(VI) (mM)

0.1 0.2 0.3 0.4

k (

mM

-1 s

-1)

0.00

0.05

0.10

0.15

0.20

0.25

Page 78: Ferrate as a New Treatment Chemical for Removal of ...

63

4.3.2 Ferrate(VI) reaction with effluent organic matter (EfOM) in secondary

effluent

Turbidity and EfOM of Fe(VI) or Fe(III)-treated secondary effluents at different

chemical doses are presented in Fig. 4-4 (a) and (b), respectively. As seen in Fig. 4-4 (a),

the secondary effluent turbidity was initially decreased from 6.53 to 2.58 NTU as the

Fe(VI) dose increased from 0.00 to 1.00 mg/L, and then gradually increased to 13.1 NTU

when the Fe(VI) dose further increased to 15.00 mg/L. In contrast, the turbidity in

Fe(III)-treated secondary effluent consistently dropped from 6.53 to 0.87 NTU with the

increasing Fe(III) dose from 0.00 to 15.00 mg/L. For any specific Fe(VI) dose, the Fe(III)

treatment achieved lower effluent turbidity than Fe(VI) treatment. The finding suggests

that coagulation and flocculation with Fe(III) produced flocs with the better settleability

than Fe(VI)-induced coagulation and flocculation. The mechanisms behind the

observation would be explored in Chapter 6.

The removal and transformation of EfOM with Fe(VI) or Fe(III) are presented in

Fig.4-4 (b). COD removal significantly went up from 0 to 12% as Fe(VI) dose increased

from 0.00 to 1.00 mg/L, and then slightly augmented to 15% when Fe(VI) further

increased to 15.0 mg/L. In contrast, for the Fe(III) treatment, the COD removal slightly

increased from 9% to 18% as the Fe(III) dose from 1.0 to 15.0 mg/L. For the COD

removal, Fe(VI) performed better than Fe(III) only at a low chemical dose (1.00 mg/L),

but provided an inferior treatment at a higher chemical dose (1.0-15.0 mg/L), thereby

indicating that the capability of Fe(VI) for COD removal was limited.

Page 79: Ferrate as a New Treatment Chemical for Removal of ...

64

However, the different patterns of the UV254 reduction were noticed between the

Fe(VI) and Fe(III) treatment. UV254 is a measurement of the amount of ultraviolet light

absorbed by aqueous constituents (e.g. natural organic matter or phenolic compounds),

especially aromatic organic compounds, in water. As the ferrate(VI) dose was increased

from 0.0 to 15.0 mg/L, the effluent UV254 removal was dramatically increased to 49%

and 23% in the Fe(VI) and Fe(III)-treated secondary effluents, respectively.

The different behaviors of Fe(VI) and Fe(III) in removal of COD and UV254 were

due to their different treatment mechanisms. In a typical Fe(VI) treatment system, Fe(VI)

decomposes via self-decomposition and reactions with water constituents (e.g. EfOM)

(Lee et al. 2004). Fe(VI) oxidation has proven to be very selective, so that ferrate(VI)

preferentially attacks organic compounds with ERMs (Lee and von Gunten 2010, Yang et

al. 2012). Accompanied with Fe(VI) reduction, Fe(III) is produced to initiate an in-situ

coagulation (Graham et al. 2010). Both chemical oxidation and coagulation are capable

of potentially removing EfOM. Graham et al. (2010) attempted to quantify the roles of

Fe(VI) oxidation and coagulation in the removal of humic acid (HA) through comparison

of the HA removals with Fe(VI) in the absence and presence of phosphate, which could

complex Fe(III) and prevent the formation of iron particles. However, the quantitative

information might not be accurate because phosphate can significantly alter the ferrate

stability in water, thereby changing the ferrate exposure. It is technically difficult to

separate the extents of organics removal due to oxidation and coagulation during the

Fe(VI) treatment (Graham et al. 2010).

Page 80: Ferrate as a New Treatment Chemical for Removal of ...

65

On the other hand, Fe(III) removed EfOM only through coagulative mechanism.

It should be noted that the coagulation performances with Fe(VI) and Fe(III) in secondary

effluent have been demonstrated to be greatly different (Jiang et al. 2015, Goodwill et al.

2015). Fe(III)-induced flocs tended to rapidly settle, while a majority of Fe(VI) resultant

particles remained suspended probably because of the formation of more stable iron

particles. In Fig.4-4(b), the stoichiometric relationship between COD removal and the

added Fe(III) in the Fe(III) treatment could be described as follows.

COD removal % = 0.016 Fe(III) (R2 = 0.96) (4-1)

The finding suggests that the Fe(III) coagulation efficiency was almost linearly increased

with Fe(III) dose. On the other hand, for Fe(VI) treatment, significant COD removal

occurring at 1.00 mg/L Fe(VI) was principally due to Fe(VI) oxidation, because

coagulation with iron was almost ineffective for organics removal at such a low dose

(Tchobanoglous et al. 2003). Generally, COD coagulation efficiency with Fe(III) is

theoretically increased with an increasing iron dose. However, the increase in COD

removal over 1.00 - 15.00 mg/L Fe(VI) was almost marginal, indicating that Fe(VI)-

induced coagulation was poor for removal of the secondary effluent COD. The UV254

reduction by Fe(III) was caused by direct Fe(III) coagulation through adsorption of UV-

absorbing compounds, while a better removal with Fe(VI) was achieved by the both

oxidative and coagulative mechanisms. Although the both effects were not separated in

this study, UV254 removal due to Fe(VI) oxidation was likely dominant because the

removal efficiency of Fe(VI)-induced coagulation for EfOM was very limited as

discussed above.

Page 81: Ferrate as a New Treatment Chemical for Removal of ...

66

(a)

(b)

Fig. 4-4 Turbidity and EfOM at different chemical doses during Fe(VI) or Fe(III)

treatment of the secondary effluent at the initial pH of 7.5: (a) turbidity; and (b) removal

efficiencies of COD and UV254 (initial pH = 7.5; initial COD = 31.7-33.1 mg/L; initial

UV254 = 0.135 – 0.142 cm-1)

0.00

2.00

4.00

6.00

8.00

10.00

12.00

14.00

0.0 5.0 10.0 15.0

Tu

rbid

ity

(N

TU

)

Fe (mg/L)

Fe(VI)

Fe(III)

0%

10%

20%

30%

40%

50%

60%

0.0 5.0 10.0 15.0

Rem

oval

effi

cien

cies

of

CO

D a

nd

UV

25

4

Fe (mg/L)

COD (Fe(VI))

COD (Fe(III))

UV254 (Fe(VI))

UV254 (Fe(III))

Page 82: Ferrate as a New Treatment Chemical for Removal of ...

67

Molecular weight (MW) fractions of EfOM in Fe(VI) and Fe(III)-treated

secondary effluents in terms of UV254 are shown in Fig. 4-5 (a) and (b), respectively.

EfOM compounds in untreated and treated samples were isolated into three groups, i.e.

high (> 10 k Da), medium (10-1 k Da), and low (< 1k Da) MW fractions. For the

untreated sample, the UV254 levels of high, medium and low MW molecules were 0.026,

0.026 and 0.082 cm-1, respectively. These findings suggested that low MW molecules

prevailed among the EfOM compounds, which accounted for 61% of the overall UV254.

As seen in Fig. 4-5 (a), UV254 in both high and medium MW groups gradually dropped

with the increasing Fe(VI) dose, suggesting that Fe(VI) treatment favorably removed

high and medium MW molecules rather than the low MW group. A significant UV254

reduction in the two groups was noticed at a low Fe(VI) dose at which coagulation

efficiency was insignificant owing to a too low Fe(III) dose (the Fe(VI) reduction

product). Therefore, chemical oxidation in the Fe(VI) treatment played an essential role

in the removal of UV254 due to the two groups of compounds. In contrast, UV254 removal

for the low MW EfOM molecules exhibited a different pattern with Fe(VI) dose. It

slightly increased from the original 0.082 to 0.088 cm-1 at 1.00 mg/L Fe(VI), and then

substantially dropped to 0.052 cm-1 as Fe(VI) increased to 15.00 mg/L. The UV254

increase observed at the low Fe(VI) dose was likely because part of high and medium

MW molecules were chemically transformed into low MW molecules after Fe(VI)

oxidation. The ensuing decrease in UV254 with increasing Fe(VI) dose indicated the

reactivity of ferrate(VI) towards these low MW EfOM molecules was high.

Consequently, the fraction of low MW molecules after the 15.00 mg/L Fe(VI) treatment

Page 83: Ferrate as a New Treatment Chemical for Removal of ...

68

was increased to 73%. Results of the control tests with Fe(III) are presented in Fig. 4-5

(b). Generally, the high and medium MW molecules were slightly removed with an

increasing Fe(III) dose. For example, UV254 due to high MW fraction was decreased from

0.026 to 0.014 cm-1 as the increasing Fe(III) was increased from 0.0 to 15.0 mg/L.

However, the removal of UV254 from the low MW group was very slight (7% removal)

over the tested Fe(III) range. Therefore, Fe(III)-driven coagulation was almost ineffective

for alleviating UV254 from low MW EfOM molecules.

Page 84: Ferrate as a New Treatment Chemical for Removal of ...

69

(a)

(b)

Fig. 4-5 MW fractions of Fe(VI) and Fe(III)-treated secondary effluent in terms of

UV254: (a) Fe(VI) treatment; and (b) Fe(III) treatment ( initial pH = 7.5)

0.000

0.010

0.020

0.030

0.040

0.050

0.060

0.070

0.080

0.090

0.100

0.0 1.0 5.0 10.0 15.0

UV

25

4(c

m-1

)

Fe(VI) (mg/L)

>10 K 1~10k <1k

0.000

0.010

0.020

0.030

0.040

0.050

0.060

0.070

0.080

0.090

0.0 1.0 5.0 10.0 15.0

UV

25

4(c

m-1

)

Fe(III) (mg/L)

>10 K 1~10k <1k

Page 85: Ferrate as a New Treatment Chemical for Removal of ...

70

UV absorbance spectra (200-400 nm) of the untreated, Fe(VI)-treated and Fe(III)-

treated secondary effluents are presented in Fig. 4-6. The three spectra were characterized

with a shoulder at 200-230 nm and a gradual decrease (tailing) over 230-400 nm. The

first band referred to Bz band was centered at 203 nm primarily due to the vibrational

perturbations in the π-electron system, while the band centered at 253 nm (ET band) was

a distinctive feature of the electronic spectra of aromatic compounds (Korshin et al.

1997). The shoulder and tailing spectra were obvious in the untreated secondary effluent,

but became less pronounced after Fe(VI) or Fe(III) treatment. Therefore, either treatment

could somewhat remove the assoicated chromospheres. However, the two treatments had

different removal efficiencies over different wavelength ranges. Fe(III) and Fe(VI)

appeared to preferentially alleviate the absorption in the Bz and ET bands, respectively,

suggesting that Fe(VI) or Fe(III) selectively removed polar functional groups such as

hydroxyl, carbonyl, carboxyl and ester groups. UV253/UV203 is an indicator of the

presence of activated aromatic rings in EfOM. The original UV253/UV203 was 0.163 in

untreated secondary effluent, but dropped to 0.103 and 0.139 after Fe(VI) and Fe(III)

treatment, respectively, again validating that the both treatments favorably targeted at the

destruction of aromatic rings. The lower UV253/UV203 achieved by Fe(VI) was likely

because these aromatic structures were better removed through Fe(VI)-driven oxidation

than Fe(III) coagulation.

The aforementioned findings indicate that ferrate(VI) tends to lower the aromatic

degrees primarily via chemical oxidation. The reactions between EfOM and Fe(VI) can

lead to additional consumption of ferrate(VI) in advanced wastewater treatment practices

Page 86: Ferrate as a New Treatment Chemical for Removal of ...

71

to reduce the degree of ferrate(VI) exposure. Therefore, the impact of EfOM on Fe(VI)

exposure needs to be evaluated to accurately determine the ferrate(VI) dose for target

pollutants during applications. Meanwhile, although ferrate(VI) does not preferentially

remove secondary effluent COD, it appears to be effective for the removal of UV254,

implying that ferrate(VI) s a potential tool for controlling DBP formation in downstream

disinfection, because UV254 is a widely accepted surrogate to measure the DBP

precursors in water (Crittenden et al. 2012).

Page 87: Ferrate as a New Treatment Chemical for Removal of ...

72

Fig. 4-6 UV absorbance of the untreated, Fe(VI)-treated and Fe(III)-treated secondary

effluents ( initial pH = 7.5; chemical dose = 15.0 mg/L)

0

0.2

0.4

0.6

0.8

1

1.2

200 250 300 350 400

UV

25

4(c

m-1

)

Wavelength (nm)

Untreated

Fe(VI) (15.0 mg/L)

Fe(III) (15.0 mg/L)

Page 88: Ferrate as a New Treatment Chemical for Removal of ...

73

Fluorescence EEM spectra of untreated secondary effluent, Fe(VI)-treated

secondary effluent and Fe(III)-treated secondary effluent are presented in Fig. 4-7 (a),

(b), and (c), respectively. Fluorescence EEM spectra is an effective tool to provide

qualitative information on chemical compositions of dissolved organic matter, such as

NOM and EfOM (Chen et al. 2012, Bro 1997, Lee and Hur 2016). Generally, the

fluorescence EEM spectra can be divided into five unique Ex– Em regions that represent

different dissolved organic matter (DOM) types (Chen et al. 2003) : 1) regions I and II,

proteins (e.g. tyrosine) at Ex/Em of < 250 nm/< 380 nm; 2) region III, fulvic acid-like

materials at Ex/Em of < 250 nm/> 380 nm; 3) region IV, soluble microbial byproduct-

like materials at Ex/Em of 250-280 nm/< 380 nm; and 4) region V, humic acid-like

organics at Ex/Em of > 250 nm/> 380 nm.

The fluorescence EEM image of untreated secondary effluent is shown in Fig. 4-

7(a). Obviously, two EEM peaks were observed in the regions III and V, representing the

presence of fulvic acid-like and humic acid-like DOM, respectively. Because the two

DOM types are hydrophobic organic matter, the secondary effluent EfOM was

characterized with hydrophobic organic compounds. Of note, both of the two

hydrophobic compounds are characterized with abundant UV-absorbing moieties

(Tchobanoglous et al. 2003, Zhao et al. 2013). As ferrate(VI) dose was gradually

increased, the EEM spectra in all the five regions were weakened, to the different degrees

(Fig. 4-(b)). In particular, the EEM peak in the region V was almost eliminated at 5.00

mg/L Fe(VI), and the EEM peak in the region V was dramatically reduced at 15.00 mg/L.

In contrast, for the Fe(III) treatment, only the EEM peak in the region III was somewhat

Page 89: Ferrate as a New Treatment Chemical for Removal of ...

74

alleviated as the Fe(III) dose was increased from 0.00 to 15.00 mg/L (Fig.4-(c)).

However, the region V peak was not significantly altered. This finding suggests that the

capability of Fe(III) for the removal of UV-quenching hydrophobic matters was limited.

Fe(III) coagulation only mitigated fulvic acid-like substances, but did not obviously

remove the humic-like organic matter. In contrast, ferrate(VI) treatment was very

effective for the removal of both fulvic-like and humic-like substances because of the

dual mechanisms, i.e. chemical oxidation and coagulation. The finding was in agreement

with the aforementioned observation that ferrate(VI) could better remove UV254

absorbance from secondary effluent, considering that the both hydrophobic organic

matters contributed to UV254 absorbance due to their strong UV-quenching properties.

The finding is also similar to the observation from a recent study (Song et al. 2016) to

investigate ferrate(VI) treatment of NOM in a drinking water source. Song et al. (2016)

also found that ferrate(VI) favorably removed hydrophobic DOM rather than hydrophilic

constituents. They ascribed the unique treatability of ferrate(VI) to the selectivity of

ferrate(VI) oxidation. Namely, ferrate(VI) better reacts with aromatic rings and double

bonds in the hydrophobic groups.

Page 90: Ferrate as a New Treatment Chemical for Removal of ...

75

Fig. 4-7(a) Fluorescence excitation-emission matrix (EEM) images of untreated

secondary effluent (X-axis is the excitation wavelength and Y-axis is the emission

wavelength: regions I and II, proteins at Ex/Em of < 250 nm/< 380 nm; region III, fulvic

acid-like materials at Ex/Em of < 250 nm/> 380 nm; region IV, soluble microbial

byproduct-like materials at Ex/Em of 250-280 nm/< 380 nm; and region V, humic acid-

like organics at Ex/Em of > 250 nm/> 380 nm.

(V) (IV)

(III) (II) (I)

Page 91: Ferrate as a New Treatment Chemical for Removal of ...

76

Fig. 4-7(b) Fluorescence excitation-emission matrix (EEM) images of untreated and

Fe(VI) treated secondary effluent (initial pH = 7.5)

Page 92: Ferrate as a New Treatment Chemical for Removal of ...

77

Fig. 4-7(c) Fluorescence excitation-emission matrix (EEM) images of untreated and

Fe(III) treated secondary effluent (initial pH = 7.5)

4.4 Conclusions

Ferrate(VI) application for wastewater reclamation targets at certain pollutants in

secondary effluent (e.g. emerging micro-pollutants, phosphorus and pathogens).

However, it unavoidably reacts with other water matrix constituents. One such example is

EfOM. The major findings in this study include:

1) Ferrate(VI) decomposition in a secondary effluent matrix follows a 2nd order

reaction with respect to Fe(VI) concentration. The rate constant exponentially

decreases with the increasing ferrate(VI) dose.

Page 93: Ferrate as a New Treatment Chemical for Removal of ...

78

2) Ferrate(VI) is more effective for reduction of UV254 than Fe(III) in secondary

effluent, but not better than Fe(III) in the removal of COD.

3) At a low dose, Fe(VI) preferentially decomposes high (> 10 k Da) and medium

(10-1 k Da) MW molecules into low MW compounds (< 1 k Da) in terms of

UV254, thereby causing an increased low MW fraction. As chemical dose further

increases, the UV254 absorbance of all the MW groups are somewhat decreased.

4) Fluorescence EEM spectra results reveal that ferrate(VI) treatment effectively

removes hydrophobic DOM, including fulvic-like and humic-like substances. In

contrast, Fe(III) coagulation only obviously alleviates fulvic-like dissolved

organic compounds.

5) Although both chemical oxidation and coagulation in ferrate(VI) treatmetn could

potentially contribute to EfOM removal, ferrate(VI)-driven oxidation seemed to

play a more essential role in the EfOM transformation under the tested conditions.

Page 94: Ferrate as a New Treatment Chemical for Removal of ...

79

4.5 References

Bro, R. (1997) PARAFAC. Tutorial and applications. Chemometrics and intelligent

laboratory systems 38(2), 149-171.

Chen, W., Westerhoff, P., Leenheer, J.A. and Booksh, K. (2003) Fluorescence excitation-

emission matrix regional integration to quantify spectra for dissolved organic matter.

Environmental science & technology 37(24), 5701-5710.

Chen, Y.Z., He, Q., Yu, D.N., Li, S. and Tan, X.M. (2012) Research on the Degradation

of BaP with Potassium Ferrate Characterized by Fluorescence. Spectroscopy and Spectral

Analysis 32(7), 1842-1845.

Crittenden, J.C., Trussell, R.R., Hand, D.W., Howe, K.J. and Tchobanoglous, G. (2012)

MWH's Water Treatment: Principles and Design, John Wiley & Sons.

Goodwill, J.E., Jiang, Y., Reckhow, D.A., Gikonyo, J. and Tobiason, J.E. (2015)

Characterization of Particles from Ferrate Preoxidation. Environmental Science &

Technology 49(8), 4955-4962.

Graham, N.J.D., Khoi, T.T. and Jiang, J.Q. (2010) Oxidation and coagulation of humic

substances by potassium ferrate. Water Science and Technology 62(4), 929-936.

Gur-Reznik, S., Katz, I. and Dosoretz, C.G. (2008) Removal of dissolved organic matter

by granular-activated carbon adsorption as a pretreatment to reverse osmosis of

membrane bioreactor effluents. Water Research 42(6), 1595-1605.

Ito, K., Jian, W., Nishijima, W., Baes, A.U., Shoto, E. and Okada, M. (1998) Comparison

of ozonation and AOPs combined with biodegradation for removal of THM precursors in

treated sewage effluents. Water science and technology 38(7), 179-186.

Jiang, Y., Goodwill, J.E., Tobiason, J.E. and Reckhow, D.A. (2015) Effect of different

solutes, natural organic matter, and particulate Fe (III) on ferrate (VI) decomposition in

aqueous solutions. Environmental Science & Technology 49(5), 2841-2848.

Korshin, G.V., Li, C.-W. and Benjamin, M.M. (1997) Monitoring the properties of

natural organic matter through UV spectroscopy: a consistent theory. Water Research

31(7), 1787-1795.

Lee, S. and Hur, J. (2016) Heterogeneous adsorption behavior of landfill leachate on

granular activated carbon revealed by fluorescence excitation emission matrix (EEM)-

parallel factor analysis (PARAFAC). Chemosphere 149, 41-48.

Lee, Y., Cho, M., Kim, J.Y. and Yoon, J. (2004) Chemistry of ferrate (Fe(VI)) in aqueous

solution and its applications as a green chemical. Journal of Industrial and Engineering

Chemistry 10(1), 161-171.

Page 95: Ferrate as a New Treatment Chemical for Removal of ...

80

Lee, Y., Yoon, J. and von Gunten, U. (2005) Spectrophotometric determination of ferrate

(Fe(Vl)) in water by ABTS. Water Research 39(10), 1946-1953.

Lee, Y. and von Gunten, U. (2010) Oxidative transformation of micropollutants during

municipal wastewater treatment: Comparison of kinetic aspects of selective (chlorine,

chlorine dioxide, ferrate(VI), and ozone) and non-selective oxidants (hydroxyl radical).

Water Research 44(2), 555-566.

Liu, H. and Fang, H.H. (2002) Extraction of extracellular polymeric substances (EPS) of

sludges. Journal of Biotechnology 95(3), 249-256.

Michael-Kordatou, I., Michael, C., Duan, X., He, X., Dionysiou, D., Mills, M. and Fatta-

Kassinos, D. (2015) Dissolved effluent organic matter: Characteristics and potential

implications in wastewater treatment and reuse applications. Water Research 77, 213-248.

Shon, H.K., Vigneswaran, S., Ngo, H.H., Kim, D.H., Park, N.E., Jang, N.J. and Kim, I.S.

(2003) Characterization of effluent organic matter (EfOM) of fouled nanofilter (NF)

membranes.

Shon, H., Vigneswaran, S., Kim, I.S., Cho, J. and Ngo, H. (2004) The effect of

pretreatment to ultrafiltration of biologically treated sewage effluent: a detailed effluent

organic matter (EfOM) characterization. Water Research 38(7), 1933-1939.

Shon, H., Vigneswaran, S. and Snyder, S. (2006) Effluent organic matter (EfOM) in

wastewater: constituents, effects, and treatment. Critical reviews in environmental

science and technology 36(4), 327-374.

Song, Y., Deng, Y. and Jung, C. (2016) Mitigation and degradation of natural organic

matters (NOMs) during ferrate (VI) application for drinking water treatment.

Chemosphere 146, 145-153.

Tchobanoglous, G., Franklin, L. and Stensel, H.D. (2003) Wastewater engineering:

treatment and reuse. Metcalf & Eddy, Inc., New York.

Yang, B., Ying, G.G., Zhao, J.L., Liu, S., Zhou, L.J. and Chen, F. (2012) Removal of

selected endocrine disrupting chemicals (EDCs) and pharmaceuticals and personal care

products (PPCPs) during ferrate(VI) treatment of secondary wastewater effluents. Water

Research 46(7), 2194-2204.

Zhao, R., Gupta, A. and Novak, J.T. (2013) Using Fenton's Reagent as a Polishing Step

for Biologically Treated Landfill Leachate. Proceedings of the Water Environment

Federation 2013(19), 386-408.

Page 96: Ferrate as a New Treatment Chemical for Removal of ...

81

CHAPTER 5 FERRATE(VI) OXIDATION OF MEFENAMIC ACID

(MEF) IN WATER: KINETICS AND REACTION MECHANISMS

5.1 Introduction

Emerging contaminants are traceable synthetic or naturally occurring chemicals

or microorganisms that are not commonly monitored in the environment but have the

potential to enter into the environment and cause known or suspected adverse ecological

and/or human health effects. Examples of emerging contaminants recently identified in

untreated and treated wastewater include pharmaceuticals and personal care products

(PPCPs), endocrine disrupting compounds (EDCs), brominated flame retardants,

perfluorinated compounds (PFCs), and new disinfection byproducts (e.g. NDMA). Most

of them are biochemically and chemically persistent, unregulated, and of ecological or

health concern even at traceable levels (Choi et al., 2006; Ying et al., 2003).

Unfortunately, traditional water and wastewater treatment plants were not specially

designed for these emerging contaminants. Many emerging contaminants have been

reported to largely flow through these treatment facilities without a sufficient removal.

Therefore, advanced treatment methods are required to eliminate or alleviate these micro-

pollutants before reclaimed water is safely reused.

Since the mid-1990s PPCPs have attracted public attention in water industries as

an emerging contaminant type due to their potentially adverse environmental and health

impacts. Because pharmaceutically active compounds are commonly used and most of

them are not completely metabolized, many pharmaceuticals end up in natural water

Page 97: Ferrate as a New Treatment Chemical for Removal of ...

82

bodies through sewers and treatment systems. Since the 1990s, a variety of PPCPs at low

concentrations have been identified in surface water, groundwater and drinking water in

the world (Doerr-MacEwen and Haight, 2006).

Ferrate(VI) has proven to be very effective for degradation of certain emerging

PPCPs in a buffered deionized (DI) water matrix, hospital wastewater (Wilde et al., 2013)

and municipal wastewater effluent. In this dissertation study, mefenamic acid (MEF) is

selected as a representative emerging PPCP. Structure and basic physical/chemical

properties of MEF are summarized in Table 5-1. It served as a model micro-pollutants

due to the following reasons. Firstly, MEF has been widely used as a common non-

steroidal anti-inflammatory drug (NSAID) for treatment of pain such as menstrual pain.

Secondly, it has been frequently identified in natural and engineered water systems.

Marketed as Ponstel in the United States and known as Ponstan in UK, MEF has been

detected at trace levels in the effluent of WWTPs in many countries such as Switzerland,

UK and Japan (Tauxe-Wuersch et al. 2005; Roberts and Thomas, 2006; Nakada et al.

2005). Thirdly, it has negative impacts on human and ecological health. Jones et al.

(2002) reported that the ratio of its predicted environmental concentration (PEC) to its

no-effect concentration (PNEC) is greater than one (1.03) based on aquatic toxicity data.

Tauxe-Wuersch et al. (2005) found that the PEC/PNEC of MEF is ten times greater than

that of ibuprofen, another commonly used pain killer, thereby indicating that the aquatic

toxicity of MEF is significant. Finally, very few efforts have been made to remove MEF

from secondary effluent for water reuse (Nakada et al. 2007; Gimeno et al. 2010; Chang

et al. 2012; Khalaf et al. 2013; Colombo et al. 2016; Chen et al. 2016a, 2016b). Yang et

Page 98: Ferrate as a New Treatment Chemical for Removal of ...

83

al. (2012) has examined frerrate(VI)_ removal efficiencies of 68 emerging contaminants

including MEF in water. However, the kinetics information and reaction mechanisms

were not studied. Therefore, there is an urgent research need for a better understanding of

ferrate(VI) degradation of MEF in a secondary effluent matrix.

Page 99: Ferrate as a New Treatment Chemical for Removal of ...

84

Table 5-1. Structure and physiochemical properties of mefenamic acid

Compound

Molecular

weight

Water

solubility

(mg/L)

pKa LogKow

Mefenamic acid

241 15.4 4.2

5.12

(Chang et

al. 2012)

5.2 Materials and methods

MEF (analytical grade, ≥99%), anhydrous, sodium phosphate dibasic (Na2HPO4)

and sodium phosphate monobasic (NaH2PO4) were purchased from Sigma-Aldrich (St.

Louis, Missouri, USA). All the buffer solutions were prepared with 18.2 MΩ purified DI

water (Millipore Milli-Q water purification system). HPLC grade methanol and

acetonitrile were purchased from EMD Merk (Gibbstown, NJ, USA).

The experiments to evaluate the ferrate(VI) capability for removing MEF were

first carried out in 100 mL beaker. The reactors were installed in a shaking bed at a rapid

mixing speed (100 rpm) to ensure a complete mixing state during the reaction. The MEF

solution (1.0 mg/L) was prepared in phosphate buffer solution (10 µM) or a secondary

Page 100: Ferrate as a New Treatment Chemical for Removal of ...

85

effluent. Of note, phosphate was selected only for the purpose of pH buffer. Although

ferrate(VI) might remove part of phosphate from water, the mechanisms are associated

with precipitation due to the formation of Fe(III) from ferrate(VI) and/or adsorption to

the Fe(III) oxide particles. Therefore, the possible reactions did not influence Fe(VI)

oxidation of MEF. Furthermore, P was not a chemical species of interest in this study. In

fact, phosphate buffer were commonly used in the previous studies on ferrate(VI)

oxidation of organic compounds in water. Therefore, phosphate buffer was used in this

task. The secondary effluent was the effluent from a secondary settling tank prior to

disinfection in the WWTP in Verona, NJ. Solution pH was adjusted pH 7.5 with 1.0 mM

NaOH or H2SO4. Ferrate(VI) dose was varied from 0.1 to 2.0 mg/L as Fe. The oxidation

treatment was initiated through the addition of an appropriate weight of K2FeO4.

Reaction proceeded for two hours within which all the ferrate(VI) was depleted through

the visual inspection (purple color vanished) and the ABTS method (Lee et al. 2005b).

Fifty milliliters of Fe(VI)-treated samples were collected and then filtered through 0.45

um membrane filters prior to analysis. The concentration of MEF was determined using

the HPLC method (Werner et al. 2005). A Supelcosil LC-18 column (25cm×4.6mm,

5μm) was used. The mobile phase was a mixed solution composed of acetonitrile and

water (60:40, v/v). Prior to use, pH of the mobile phase solution was adjusted to 3.9

with acetic acid. Flow rate was controlled at 1.2 mL/min. MEF samples were analyzed

with a PDA Plus-300 nm detector (Thermo) with a retention time at 6.1 min. Kinetics

tests of ferrate(VI) degradation of MEF were conducted in 10 µM phosphate buffer at

pH 7.5. The initial concentration of MEF was fixed at 4.3 mg/L (0.0178 M). Four

Page 101: Ferrate as a New Treatment Chemical for Removal of ...

86

different molar ratios of Fe(VI) to MEF were tested, including 3:1, 4:1, 5:1 and 6:1. At

different sampling times, appropriate volumes of samples were collected into 25 ml

centrifuge tube for Fe(VI) analysis. Residual concentration of Fe(VI) was quantified

using ABTS method (Lee et al. 2005). To accurately measure residual MEF, the collected

sample was mixed with 0.5 mL 100 mM sodium thiosulfate solution that could quench

any residual Fe(VI). Results showed that MEF degradation was too rapid to be accurately

measured.

In order to determine MEF degradation pathways, a Thermo-Fisher Scientific

linear quadrupole ion trap (LTQ-XL) mass spectrometer equipped with an electrospray

ionization (ESI) source and HPLC was used to identify chemically transformation

products (OPs) of MEF for better understating of its degradation pathway. The treatment

tests were carried out in 10 µM pH 7.5 phosphate buffer solution using the method as

described previously. Ferrate dose was 10.0 mg/L as Fe and the initial MEF was 10.0

mg/L. Although the initial MEF concentration was over the typical MEF occurrence level

in a treated wastewater, the high level of MEF did not influence Fe(VI) degradation of

MEF molecules in water and could produce a strong signal in HPLC/MS analysis for the

identification of OPs. Once the reaction was completed, 5 mL sample was collected and

then filtered through 0.45 µm membrane filters for analysis. The chemical OPs were

separated by HPLC equipped with a C18 column followed by a multiple-stage tandem

mass spectrometry (MS2, MS3, and MS4) of ions of interest. The structure and/or

functionalities of OPs were proposed based on the obtained fragmentation patterns of

multiple-stage tandem mass spectrometry.

Page 102: Ferrate as a New Treatment Chemical for Removal of ...

87

5.3 Results and Discussion

5.3.1 Ferrate(VI) degradation of MEF in phosphate buffered solution (PBS) and

secondary effluent matrixes

MEF removal efficiencies at different ferrate(VI) doses are summarized in Table

5-2 and shown in Fig. 5-1. Generally, the removal was increased with the increase in

ferrate(VI) dose. The removal efficiency had a positively linear relationship with

ferrate(VI) dose in a phosphate buffered solution (PBS) or a secondary effluent.

MEF removal = 0.332 Fe(VI), R2 = 0.96 (5-1)

MEF removal = 0.217 Fe(VI), R2 = 0.99 (5-2)

Here, MEF removal is MEF removal efficiency (%); and Fe(VI) is the ferrate(VI) mass

concentration. At the highest ferrate(VI) dose (i.e. 2.0 mg/L), the maximum removal

efficiencies of MEF were 61% and 44% at PBS and secondary effluent, respectively. The

discounted removal in secondary effluent was ascribed to the matrix effect of secondary

effluent. Various reducing agents present in secondary effluent potentially competed with

MEF for ferrate(VI). The resulted ferrate(VI) consumption could cause a decreased MEF

degradation.

It is of interest to compare ferrate(VI) oxidation with other oxidative processes for

the degradation of MEF in water, as presented in Table 5-3. Among the different

oxidative processes, O3/UV (an AOP) had the highest removal efficiency for MEF (70%)

in terms of the removal efficiency, followed by ferrate(VI) oxidation (61% removal in

PBS). Ferrate(VI) exhibited better removal than other two AOPs (photo-Fenton process

and H2O2/UV) and ozonation in terms of the MEF degradation efficiency. It is of interest

Page 103: Ferrate as a New Treatment Chemical for Removal of ...

88

to compare ferrate(VI) and ozone for the degradation of MEF, because the both oxidants

share some unique oxidative characteristics. For example, the two oxidants are selective

and preferentially oxidize ERMs. Here the molar ratio of oxidized MEF to an oxidant

dose is used as an indicator. Ferrate(VI) had 0.071 mM MEF removed/1.0 mM Fe(VI),

slightly below 0.088 mM removed/1.0 mM O3. This finding indicates that ferrate(VI)

appeared to be slightly less efficient for the MEF degradation than ozone.

Table 5-2 MEF removal efficiencies at different Fe(VI) doses in PBS and secondary

effluent matrixes

Fe(VI) dose

(mg/L)

Fe(VI) dose

(mmol/L)

MEF removal in

PBS(%)

MEF removal in SE

(%)

0.1 0.002 6% 2%

0.5 0.009 23% 7%

1.0 0.018 40% 22%

2.0 0.036 61% 44%

Table 5-3 Comparison of different oxidative processes for the degradation of MEF

Oxidation

processes

Oxidant

dose

Initial MEF

concentration

pH Time

(min)

Degradation of MEF (%)

photo-Fenton 10 mmol/L - 6.5 60 54.61 (Colombo et al. 2016)

H2O2/UV 10 mmol/L - 2.5 60 41.47 (Colombo et al. 2016)

Ozonation 0.015

mmol/L

(0.7 mg/L)

0.002 mmol/L

(0.54 mg/L)

7.0 40 57.5 (Chang et al. 2012)

O3/UV 0.015

mmol/L

(0.7 mg/L)

0.002 mmol/L

(0.54 mg/L)

7.0 40 ~70 (Chang et al. 2012)

Fe(VI) 0.036

mmol/L

(2.0 mg/L)

0.004 mmol/L

(1.0 mg/L)

7.5 > 60 61 (PBS)

44 (secondary effluent)

(this study)

Page 104: Ferrate as a New Treatment Chemical for Removal of ...

89

Fig 5-1 MEF removal rate under different initial ferrate(VI) dose in buffer and secondary

effluent

MEF removal = 0.332 Fe(VI)

R² = 0.96

MEF removal = 0.217 Fe(VI)

R² = 0.99

0%

10%

20%

30%

40%

50%

60%

70%

0.00 0.50 1.00 1.50 2.00

ME

F r

emoval

Fe(VI) (mg/L)

PBS Secondary effluent

Page 105: Ferrate as a New Treatment Chemical for Removal of ...

90

5.3.2 Kinetics studies of Fe(VI) degradation of MEF

Kinetic tests were carried out at pH 7.5 in 10 µM phosphate buffer solution. The

overall Fe(VI) decomposition is through two pathways: 1) self-decomposition; and 2)

reactions with MEF or its OPs. Therefore, the overall ferrate(VI) decomposition kinetics

equation can be written as:

𝑑[𝐹𝑒(𝑉𝐼)]

𝑑𝑡= −𝑘1[𝐹𝑒(𝑉𝐼)]2 − ∑ 𝑘𝑖[Fe(VI)]𝑎𝑖[𝐶𝑖]

𝑏𝑖 𝑛𝑖=2 (5-3)

Here, −𝑘1[𝐹𝑒(𝑉𝐼)]2 is Fe(VI) self-decomposition (a 2nd order reaction with respect to

[Fe(VI)], in which k1 is a rate constant; − ∑ 𝑘𝑖[Fe(VI)]𝑎𝑖[𝐶𝑖]𝑏𝑖 𝑛𝑖=2 is the rate of Fe(VI)

loss due to the reactions with MEF and its OPs; i=2, MEF; i=3, 4,…n for different

chemical oxidation products of MEF; ai is the reaction order with Fe(VI) for the

ferrate((VI) i; and bi is the reaction order with Fe(VI) for the chemical compound i.

Kinetic data of overall Fe(VI) decomposition and its consumption due to MEF

and its OPs are shown in Fig. 5-2. Kinetic data of the overall Fe(VI) decomposition and

Fe(VI) self-decomposition can be measured in a PBS solution in the presence and

absence of MEF, respectively, corresponding to the red and blue symbols and lines in

Fig. 5-2. At any specific sampling time, the difference between the two residual

ferrate(VI) concentrations in the absence and presence of MEF was the kinetic data of

Fe(VI) reactions with MEF and its OPs (the purple symbols and lines in Fig. 5-2). And

the ferrate(VI) consumption could be obtained from the difference between the added

ferrate(VI) dose and residual ferrate(VI) concentration (the green symbols in Fig. 5-2).

Therefore, the blue and red symbols represent experimentally measured data, while the

purple and green symbols are the processed data from experimental results. As seen, at all

Page 106: Ferrate as a New Treatment Chemical for Removal of ...

91

the different [Fe(VI)]:[MEF] ratios, a majority of ferrate(VI) loss was due to the reactions

with MEF and its OPs, rather than self-decomposition. Within 60 s, the residual Fe(VI)

rapidly dropped to 1.0 mg/L or below. It is found that the second item in Eq. (5-3) can be

empirically expressed as a kinetic reaction with respect to [Fe(VI)], as follows.

𝑑[𝐹𝑒(𝑉𝐼)]

𝑑𝑡│𝑀𝐸𝐹&𝑂𝑃𝑠 = − ∑ 𝑘𝑖[Fe(VI)]𝑎𝑖[𝐶𝑖]

𝑏𝑖 =𝑛

𝑖=2 − 𝑘[Fe(VI)]𝑎 (5-4)

Here, k is the rate constant; and a is the reaction order with respect to [Fe(VI)].

Experimental data of Fe(VI) decomposition due to the reactions with MEF and its OPs

were fitted with the kinetic models of 1st and 2nd order reactions with respect to [Fe(VI)],

separately. Key kinetics parameters are summarized in Table 5-4. Clearly, the ferrate(VI)

decomposition due to MEF and OPs better followed a 2nd order reaction pattern (R2 >

0.93) than a 1st order reaction behavior (R2 = 0.72-0.81). Since there is insufficient MEF

kinetic data from literature, it is not possible to compare the rate constant from our

research to others. However, the pattern of ferrate(VI) decomposition in present of MEF

shows similarity with MEF degradation with ozone (Chang et al. 2012). Furthermore, a

plot of k and [Fe(VI)]:[MEF] is shown in Fig. 5-3. The rate constant k decreased from

0.0209 to 0.0067 mM-1 ∙ s-1 as [Fe(VI)]:[MEF] increased from 3 to 6.The data enable us

to determine the kinetics behaviors for Fe(VI) decomposition in ferrate(VI) treatment of

MEF in water and obtain key kinetics parameters, but cannot provide mechanistic

information on the MEF degradation by Fe(VI).

Page 107: Ferrate as a New Treatment Chemical for Removal of ...

92

(a)

(b)

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

0 20 40 60 80

Fe(

VI)

res

idu

al

(mg/L

)

Time (second)

[Fe(VI)]:[MEF]=3:1

Fe(VI) self decay

Total Fe(VI) decay

Fe(VI) comsumption

Fe(VI) decay due to

MEF and OPs

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

4.50

0 20 40 60 80

Fe(

VI)

res

idu

al

(mg/L

)

Time (second)

[Fe(VI)]:[MEF]=4:1

Fe(VI) self decay

Total Fe(VI) decay

Fe(VI) comsumption

Fe(VI) decay due to

MEF and OPs

Page 108: Ferrate as a New Treatment Chemical for Removal of ...

93

(c)

(d)

Fig. 5-2 Kinetics of overall ferrate(VI) decomposition and ferrate(VI) decomposition due

to self-decay and reactions with MEF and its OPs: (a) [Fe(VI)]:[MEF] = 3:1; (b)

[Fe(VI)]:[MEF] = 4:1; (c) [Fe(VI)]:[MEF] = 5:1; and (d) [Fe(VI)]:[MEF] = 6:1. (pH

=7.5, initial MEF = 0.018 mM, i.e. 4.3 mg/L, and [Fe(VI)] = 0.054, 0.072, 0.090, and

0.108 mM, corresponding to 3.0, 4.0, 5.0 and 6.0 mg/L, respectively) .

0.00

1.00

2.00

3.00

4.00

5.00

6.00

0 20 40 60 80

Fe(

VI)

res

idu

al

(mg

/L)

Time (second)

[Fe(VI)]:[MEF]=5:1

Fe(VI) self decay

Total FeVI) decay

Fe(VI) consumption

Fe(VI) decay due to

MEF and OPs

0.00

1.00

2.00

3.00

4.00

5.00

6.00

7.00

0 20 40 60 80

Fe(

VI)

res

idu

al

(mg/L

)

Time (second)

[Fe(VI)]:[MEF]= 6:1

Fe(VI) self decay

Total FeVI) decay

Fe(VI) consumption

Fe(VI) decay due to

MEF and OPs

Page 109: Ferrate as a New Treatment Chemical for Removal of ...

94

Table 5-4 Rate constant k and R2 in different kinetics models for Fe(VI) decomposition

due to the reactions with MEF and OPs

[Fe(VI)]:[MEF]

1st order 2nd order

k2 (s-1) R2 k2 (mM-1 s-1) R2

3 0.0297 0.81 0.0209 0.98

4 0.0281 0.76 0.0144 0.93

5 0.0222 0.75 0.0076 0.94

6 0.0229 0.72 0.0067 0.92

Fig. 5-3 k vs. [Fe(VI)]:[MEF] (pH =7.5, initial MEF = 0.018 mM)The rate constant k2’

has a negative correlation again the initial Fe (VI):MEF ratio.

0.000

0.005

0.010

0.015

0.020

0.025

2 3 4 5 6 7

k (

mM

-1s-1

)

[Fe(VI)]:[MEF]

Page 110: Ferrate as a New Treatment Chemical for Removal of ...

95

5.3.3 Reaction pathways of ferrate(VI) degradation of MEF

HPLC-LTQ was used to identify chemical OPs produced from ferrate(VI)

decomposition of MEF in water. The advanced analytical technique enabled the

determination of multiple leveled MS fragmentation. Five major chromatogram peaks

were noticed at m/z 138, 256, 258, 290, and 511 under a positive mode, respectively

(Table 5-5). Multiple-stage tandem mass spectrometry of the five fragments are shown in

Fig. 5-5, 5-6, 5-7 and 5-8. Two of the five observed peaks (i.e. m/z = 256 and 258) were

also reported in other studies to investigate photo-degradation of MEF (Chen el al, 2016a

and Chen el al. 2016b). In the photo-degradation of MEF, the principle oxidants were

reactive oxygen species (ROS) such as •OH, O2 and •O2-.

Table 5-5 Fragments (m/z) and chemical formula of OPs identified by multiple-stage

tandem MS

m/z Chemical

formula MS2 MS3 MS4

138.17 C7H7NO2 120 92

256.17 C15H13NO3 239, 238, 223 210, 223 206, 195, 167

258.21 C15H15NO3 240 212,225 197

290.11 C15H14NO5 272 244, 243, 254 215, 228

511.20 C30H25N2O6 493,489, 293,

270

475, 252, 447,226

The MS spectrum of OP-138 indicates the breakage of MEF occurred at the N-C

bond. In MS2 and MS3 fragmentation spectrum of OP-138, major fragments ions with

m/z 120 (-18) and 92 (-28) corresponded to the loss of H2O and C=O carbonyl group,

Page 111: Ferrate as a New Treatment Chemical for Removal of ...

96

respectively (Fig. 5-4). OP-138 had a much smaller m/z than those of all the peaks

observed by Chen et al. (2016 a and b), suggesting that ROS produced from photo-

degradation was not capable of breaking down MEF into very small MW fragments that

ferrate(VI) could.

The OP-256 and OP-258 observed in this study were also reported by Chen et al.

(2016a and b). For OP-256, three fragment ions at m/z 239 (-13), 238 (-18) and 223(-18, -

15) were detected in the MS2 spectrum, which corresponded to loss of –OH, -H2O and –

CH3, respectively. In addition, -CO and –CH3 groups could be deduced according to the

MS3 spectrum of OP-256. The further MS4 spectrum of OP-256 revealed that three more

moieties (one hydroxyl group and two carbonyl groups) were included in OP-256, which

had the m/z of 206, 195 and 167, respectively. These fragments could be formed due to

loss of –OH, -CO and two –CO, respectively (Fig. 5- 6). Similar fragments with m/z 210

and m/z 195 were discovered in MS2 fragments by Chen et al. (2016 a and b). Of note,

the multi-stage tandem MS provided the information regarding the functionalities and

their possible locations of OP-256, but the information was not sufficient to identify the

detailed structure of OP-256. Under our MS2 spectrum, the major fragments were ions

with m/z 238, m/z 239 and m/z 223, corresponding to the loss of –OH, -H2O and –CH3,

respectively. Next, the fragments from m/z 238 in MS3 spectrum included two smaller

pieces, i.e. m/z 223 (-15) and m/z 210 (-28), which indicates that the loss of a –CH3 and a

-CO occurred for OP-256. The further MS4 cleavage of m/z 223 shows that the m/z 195

and 167 were due to the loss of a –CO and two –CO. Based on these finding, it is

Page 112: Ferrate as a New Treatment Chemical for Removal of ...

97

plausible that Fe(VI) oxidation led to the addition of one oxygen atom and a carbon from

one of the methyl group of MEF, producing –C=O to replace -CH3.

Moreover, another common transformation compound OP-258 was identified

from these studies of Chen et al. (2016 a and b), which suggests the similarity of Fe(VI)-

driven oxidation process and UV-photolysis. As shown in Fig. 5-7, the fragment with a

m/z at 258 clearly indicates that one oxygen atom was attached on MEF. This

carbonization could take place at either para- or meta-position of the nitrogen atom (Fig

5-7). As seen in the fragments from the MS3 spectrum, the major MS2 molecule with m/z

240 was further cleaved into smaller fragments with m/z 225 (-15) and m/z 212 (-28),

which indicated the loss of a -CH3 and –CO. The MS4 spectrum shows that the m/z 197

was due to one –CO loss from m/z 225. The proposed OP-258 structure in this study is

similar to that proposed by Chen et al. (2016 a and b), likely because both Fe(VI)

oxidation and UV photo-degradation could result in the carbonization of MEF. In other

studies to degrade microcystin-LR (MC-LR), similar chemical oxidation products were

also reported from Fe(VI) and photocatalytic oxidation processes (Jiang et al. 2014; Zong

et al. 2013; Antoniou et al. 2008).

However, two other major OPs were identified after Fe(VI) degradation of MEF in this

study, i.e. OP-290 and OP-511, which were not reported after photo-degradation. OP-

290 was detected at MS2, whose fragments m/z 244 (-28), m/z 243 (-29) and m/z 254 (-

18) were identified from the MS3 fragment ion with m/z 272. These molecule losses were

probably due to the cleavage of –CO, -COH and –H2O. Further MS4 fragment ions from

m/z 243 indicate that one more –CO and one more –CH3 were included in OP-290. The

Page 113: Ferrate as a New Treatment Chemical for Removal of ...

98

difference between OP-290 and MEF suggests that two additional C=O and one –OH

could be attached on MEF after Fe(VI) oxidation (Fig 5-8). This finding is in agreement

with to the observations in the Fe(VI)-driven oxidation studies from other groups (Yang

et al. 2011; Anquandah et al. 2013; Jiang et al. 2014, Karlesa et al. 2014). Jiang et al.

(2014) proposed the formation of a number of products associated with mono-, di- and

trihydroxylation of aromatic rings from MC-LR in Fe(VI) oxidation of MC-LR.

Anquandah et al. (2013) suggested that Fe(VI) oxidation of propranolol involved the

cleavage of an aromatic ring by two oxygen atoms. However, Karlesa et al. (2014)

proposed the either one or two additional oxygen atoms would attack the thioether moiety

rather than the only one aromatic ring when ferrate(VI) was applied to oxidize beta-

lactam antibiotic, likely because thioether ring was less stable than aromatic rings. In

addition, a similar coupling reaction was proposed because of the existing of OP-511.

OP-511 could be a dimer ion of OP-256 as a further oxidation product. The reaction

mechanism was discussed by Wilde (2013) for Fe(VI) degradation of atenolol in water.

Degradation pathways of MEF during ferrate(VI) oxidation is proposed in this

study (Fig. 5-9). It is known that Fe(VI) favorably attacks ERMs such as secondary

amine or Sulphur group in the molecule (Yang et al. 2011). Therefore, the breakage of N-

C bonding occurs to form OP-138 (Pathway I), which is similar to the ether bond

breaking in the study of Yang et al. (2011) and Wilde et al. (2013). In addition, Raphael

et al. (2000) and Ma et al. (2012) found that the energy of N-C bond was low, making it

more vulnerable. The pathway II proposed involves with the reactions such as

hydroxylation, ketonization, and coupling. According to the MS3 and MS4 fragments, a –

Page 114: Ferrate as a New Treatment Chemical for Removal of ...

99

OH attack could occur on one methyl carbon to produce intermediate OP-256. Further

oxidation could result in the formation of OP-256 isomers. This is in agreement with one

of the MEF pathways during UV photo-degradation of MEF (Chen et al. 2016a and b).

The plausible coupling reaction is proposed based on the by identification of OP-511.

Similar mechanism have been discussed by Yang et al. (2011) and Wilde et al. (2013)

that investigated ferrate(VI) degradation of triclosan and atenolol, respectively.

The electrophilic attack on the aromatic ring makes delocalization of positive

charges from the nitrogen atom to the aromatic ring. This enabled the electron deficiency

on the aromatic ring, thus making the para or meta position was readily attacked by

oxygen atom to produce OP-258 (Pathway III). OP-290 formed in the fourth proposed

pathway gets involved with the trihydroxylation of aromatic rings, which was proposed

by Jiang et al. (2014) that studied ferrate(VI) degradation of MC-LR in water.

To sum up, ferrate(VI) degradation of MEF experienced different reaction

pathways. Partially similar to the UV photolysis, it produced OP-256 through

hydroxylation and ketonization. However, more reaction mechanisms occurred

simultaneously, suggesting that ferrate(VI) oxidation was more powerful than UV

photolysis. For example, ferrate(VI) was able to break the N-C bonding to produce

smaller fragments i.e. OP-138. And mono-, di- or trihydroxylation occurred to form OP-

258 and OP-290 when aromatic rings were attacked by oxygen atoms.

Page 115: Ferrate as a New Treatment Chemical for Removal of ...

100

Fig. 5-4 Structure identification and functional group determination of MEF OPs after ferrate(VI) oxidation

Page 116: Ferrate as a New Treatment Chemical for Removal of ...

101

Fig. 5-5 Multiple-stage tandem mass spectrometry studies of m/z 138 ion in the positive

mode. EIC: extracted ion chromatography of m/z 138; MS1: m/z 138 is the parent ion of

interest; MS2: CID of m/z 138 ion; MS3: CID of m/z 120 ion

Page 117: Ferrate as a New Treatment Chemical for Removal of ...

102

Fig. 5-6 Multiple-stage tandem mass spectrometry studies of m/z 256 ion in the positive

mode. EIC: extracted ion chromatography of m/z 256; MS1: m/z 256 is the parent ion of

interest; MS2: CID of m/z 256 ion; MS3: CID of m/z 238 ion; MS4: CID of m/z 223 ion.

Page 118: Ferrate as a New Treatment Chemical for Removal of ...

103

Fig. 5-7 Multiple-stage tandem mass spectrometry studies of m/z 258 ion in the positive

mode. EIC: extracted ion chromatography of m/z 258; MS1: m/z 258 is the parent ion of

interest; MS2: CID of m/z 258 ion; MS3: CID of m/z 240 ion; MS4: CID of m/z 225 ion.

Page 119: Ferrate as a New Treatment Chemical for Removal of ...

104

Fig. 5-8 Multiple-stage tandem mass spectrometry studies of m/z 290 ion in the positive

mode. EIC: extracted ion chromatography of m/z 290; MS1: m/z 290 is the parent ion of

interest; MS2: CID of m/z 290 ion; MS3: CID of m/z 272 ion; MS4: CID of m/z 243 ion.

Page 120: Ferrate as a New Treatment Chemical for Removal of ...

105

Fig. 5-9 Proposed pathways of MEF degradation with Fe(VI)

Page 121: Ferrate as a New Treatment Chemical for Removal of ...

106

5.4 Conclusions

The major findings in this chapter are summarized as below.

1) The MEF degradation efficiency exhibited a linear correlation with ferrate(VI)

dose in both PBS and secondary effluent. A ferrate(VI) dose of 2.0 mg/L

Fe(VI) could remove 61% and 44% MEF at an initial concentration of 1.0

mg/L in PBS and secondary effluent matrixes, respectively. The lower

removal efficiency observed in the secondary effluent was due to the matrix

effect of secondary effluent. MEF was not mineralized by ferrate(VI)

completely. The reduction of MEF is more due to chemical transformation.

2) Ferrate(VI) decomposition due to the reaction with MEF and its OPs followed

a 2nd order reaction pattern. The rate constant was linearly decreased with the

increasing molar ration of Fe(VI) to MEF (i.e. [Fe(VI)]:[MEF]).

3) HPLC-LTQ technique was used to reveal the reaction mechanisms of

ferrate(VI) degradation of MEF in water. Five major oxidation products (OPs)

and four different reaction pathways are proposed to explain the complex

reaction mechanisms. As a variety of complex products were generated and

their impact to water quality is unknown at this point, a future research

regarding toxicity of the OPs are strong recommended.

Page 122: Ferrate as a New Treatment Chemical for Removal of ...

107

5.5 References

Anquandah, G., Sharmar, V., Panditi, V., Gardinali, P, Kim, H. and Oturan, M. (2013)

Ferrate(VI) oxidtation of propranolol: Kinetics and products, chemosphere 91:105-109

Antoniou, M. G.; Shoemaker, J. A.; de la Cruz, A. A.; Dionysiou, D. D. LC/MS/MS

structure elucidation of reaction intermediates formed during the TiO2 photocatalysis of

microcystin-LR. Toxicon 2008, 51, 1103−1118.

Chang, E.E., Liu, T.Y., Huang, C.P., Liang, C.H., Chiang, P.C (2012) Degradation of

mefenamic acid from aqueous solutions by the ozonation and O3/UV processes,

Separation and Purification Technology 98:123-129

Chen, P. Wang, F., Yao, K., Ma, J., Li, F., Ly, W. and Liu, G. (2016 a) Photodegradation

of Mefenamic Acid in Aqueous Media: Kinetics, Toxicity and Photolysis Products, Bull

Environ Contam Toxicol 96:203-209

Chen, P., Wang, F. and Su, H. (2016 b) Photo-catalytical degradation of mefenamic acid

by TiO2 (P25) in aqueous solution, Environmental Chemistry, 35(8):1627-1635

Choi, K.J., Kim, S.G., Kim, C.W. and Park, J.K (2006) removal efficiencies of endocrine

disrupting chemicals by coagulation/flocculation, ozonation, powdered/granular activated

carbon adsorption, and chlorination, Korean J. Chem. Eng. 23: 399-408

Colombo, R. Ferreira, T. Ferreira, R., Lanza, M. (2016) Removal of Mefenamic acid

from aqueous solutions by oxidative process: Optimization through experimental design

and HPLC/UV analysis, Journal of Environmental Management (167): 206-213

Doerr-MacEwen NA, Haight ME (November 2006). "Expert stakeholders' views on the

management of human pharmaceuticals in the environment". Environ Manage. 38 (5):

853

Gimeno, O. Rivas, J. Encinas, A. and Beltran, F. (2010) Application of advance oxidation

processes to mefenamic acid elimination, World Academy of Science, Engineering and

Technology 42:1090-1092

Hata, T., Kawai, S., Okamura, H., Nishida, T., (2010) Removal of diclofenac and

mefenamic acid by the white rot fungus Phanerochaete sordida YK-624 and identification

of their metabolites after fungal transformation, Biodegradation 21:681-689

Jiang, W., Chen, L., Batchu, S., Gardinali, P., Jasa, L., Marsalek, B., Zboril, R.,

Dionysiou, D. O’Shea, K., and Sharmar, V., (2014) Oxidation of Microcystin-LR by

Ferrate (VI): Kinetics, Degradation Pathways, and Toxicity Assessment, Environmental

Science & Technology 48:1216-12172

Page 123: Ferrate as a New Treatment Chemical for Removal of ...

108

Jones, O.A.H. Voulvoulis, N. and Lester, J.N. (2002) Aquatic environmental assessment

of the top 25 english prescription pharmaceuticals, Water Research 36:5013-5022

Karlesa, A., De Vera, G., Dodd, M., Park, J., Espino, P., and Lee, Y. (2014) Ferrate(VI)

oxidation of beta-lactam Antibiotices: Reaction Kinetics, Antibacterial Activity Changes,

and Transformation Products, Environmental Science and Technology, 48: 10380-10389

Khalaf S, Al-Rimawi F, Khamis M, Nir S, Bufo SA, Scrano L, Mecca G, Karaman R ,

(2013) Efficiency of membrane technology, activated charcoal, and a micelle-clay

complex for removal of the acidic pharmaceutical mefenamic acid, J Environ Sci Health

A Tox Hazard Subst Environ Eng. 2013;48(13):1655-62

Lee, Y., Yoon, J. and von Gunten, U. (2005) Spectrophotometric determination of ferrate

(Fe(VI)) in water by ABTS. Water Research 39(10), 1946-1953.

Li, C., Li, X., Graham, N., and Gao, N., (2008) The aqueous degradation of bisphenol A

and steroid estrogens by ferrate, Water Research, 42:109-120

Ma, Y., Gao, N. and Li, C. (2012) Degradation and Pathway of Tetracycline

Hydrochloride in Aqueous Solution by Potassium Ferrate, Environmental Engineering

Science 29:357-362

Nakada, N. Komori, K. and Suzuki, Y. (2005) Occurrence and fate of anti-inflammatory

drugs in wastewater treatment plants in Japan, Environmental Science 12:359-369

Nakada,N. Shinohara,H., Murata,A., Kiri,K., Managaki,S., Sato,N. Takada,H. (2007)

Removal of selected pharmaceuticals and personal care products (PPCPs) and endocrine-

disrupting chemicals (EDCs) during sand filtration and ozonation at a municipal sewage

treatment plant, Water Research 41:4373-4382

Raphael, D., Maume, D., Le bizec, B., and Pouliquen, H. (2000), Preliminary assays to

elucidate the structure of oxytetracycline’s degradation products in sediments,

determination of natural tetracyclines by high-performance liquid chromatography fast

atom bombardment mass spectrometry, Journal Chromatography B 748(2):369-381

Roberts, P.H. and Thomas, K.V. (2006) The occurrence of selected pharmaceuticals in

wastewater effluent and surface waters of the lower Tyne catchment, Science of the Total

Environment 356:143-153

Tauxe-Wuersch, A. De Alencastro, L.F., Grandjean, D. and Tarradellas, J. (2005)

Occurrence of several acidic drugs in sewage treatment plants in Switzerland and risk

assessment, Water Research 39:1761-1772

Werner, J., McNeill, K., and Arnold, W. (2005) Environmental photodegradation of

mefenamic acid, Chemosphere, 58:1339-1346

Page 124: Ferrate as a New Treatment Chemical for Removal of ...

109

Wilde, M., Mahmoud, W., Kummerer, K., and Martins, A., (2013) Oxidation-coagulation

of β-blockers by K2FeVIO4 in hospital wastewater: Assessment of degradation products

and biodegradability, Science of the Total Environment, 452-453:137-147

Yang, B. Ying, G., Zhao, J., Zhang, L., Fang, Y. and Nghiem, L. (2011) Oxidation of

triclosan by ferrate: Reaction kinetics, products identification and toxicity evaluation,

Journal of Hazardous Materials 186:227-235

Ying, G.G., Kookana, R.S. and Dillon, P (2003) Sorption and degradation of selected five

endocrine disrupting chemicals in aquifer material, Water Research 37:3785-3791

Zong, W.; Sun, F.; Sun, X. Oxidation by-products formation of microcystin-LR exposed

to UV/H2O2: Toward the generative mechanism and biological toxicity. Water Res.

2013, 47, 3211−3219

Page 125: Ferrate as a New Treatment Chemical for Removal of ...

110

CHAPTER 6 PERFORMANCE OF FERRATE(VI) FOR

TREATMENT OF SECONDARY EFFLUENT IN A CONTINUOUS-

FLOW REACTOR

6.1 Introduction

Batch reactors are typically restricted to laboratory-scale investigation or small

flow water treatment. In engineering practice, however, continuous-flow reactors are

used more commonly in full-scale treatment facilities because of the large volumes of

water processed. The reactors are operated on a continuous basis with flow into and out

of the reactor.

Proper reactor types are crucial for treatment because a reactor is where both

mixing and reactions occur. Two major continuous-flow reactors are continuous stirred-

tank reactor (CSTR) and plug-flow reactor (PFR). In CSTR, solution is completely

uniform without concentration gradients, and any chemical added to the solution is

instantly and uniformly distributed throughout the reactor. In contrast, in PFR, solution is

ideally mixed in the lateral direction but not mixed longitudinally. Selection of a proper

reactor type depends heavily upon kinetics. Under an ideal flow condition, PFR is more

(for 1st or 2nd-order reactions) efficient than, or at least equally (for zero-order reaction)

efficient with, CSTR, because the concentration of the chemical added sharply drops due

to dilution in CSTR, but gradually decreases to maintain a relatively high level within

PFR. That is, to achieve the same treatment efficiency, the reactor size required for PFR

is less than, or at least equal to, the volume of a corresponding CSTR. Or PFR can

Page 126: Ferrate as a New Treatment Chemical for Removal of ...

111

accomplish higher removal rates than, or at least equal to, the treatment achieved by

CSTR with identical reactor sizes. Besides, in engineering practice, PFR is a preferred

option for chemical oxidation processes (e.g., UV/H2O2 and O3). However, the

information regarding the performance of ferrate(VI) for removal of wastewater

contaminants in a PFR is highly limited.

Another interesting question on ferrate(VI) application in a PFR is the properties

of the produced iron precipitates accompanied with Fe(VI) reduction. Different from the

oxidants (e.g. ozone) subject to a homogenous reduction, ferrate(VI) is reduced to Fe(III)

at an equimolar amount (Jiang 2007, Sharma et al. 2008, Jiang and Lloyd 2002, Carr et

al. 1985, Carr 2008, Lee et al. 2014). The formation of Fe(III) is a continuous process,

rather than a one-step addition of Fe(III). At a typical wastewater treatment condition, the

produced Fe(III) rapidly precipitates in water. Traditionally, it is believed that the

formation of Fe(III) can initiate a unique in-situ coagulation that is capable of removing

suspended particles in water. The iron precipitates are a complex mixture when

ferrate(VI) is applied for treatment of secondary effluent, at least composed of certain

iron hydroxide species and Fe-P minerals.

Previous studies were performed to characterize these Fe(VI)-induced particles

produced from ferrate(VI) reduction, and compared them with Fe(III) resultant iron oxide

particles in a traditional coagulation process (Graham et al. 2010, Prucek et al. 2013,

Goodwill et al. 2015, Tien and Graham 2011, Prucek et al. 2015). Graham et al. (2010)

applied a photometric dispersion analyzer (PDA) instrument to study the flocs from

Fe(VI) in a humic acid solution. PDA is an instrument to monitor rapidly changing

Page 127: Ferrate as a New Treatment Chemical for Removal of ...

112

particle suspensions with an optical technique that analyses the light transmitted through

a flowing suspension (Graham et al. 2010). Although ferrate achieved the comparable, or

better, floc formation to ferric chloride, a less DOC removal was observed in the

ferrate(VI) treatment. Tien and Graham (2011) evaluated the effect of a disperse phase on

the floc formation through the addition of kaolin (colloidal particles) to a Fe(VI)-humic

acid solution system. They found that the magnitude of floc formation with ferrate was

inferior to that with ferric chloride. Moreover, the iron floc growth rate in the ferrate(VI)

group was slower than that in the control group with ferric chloride. This observation

may be because of a slow Fe(VI) reduction rate that led to a gradual production of

Fe(III). Very recently, Goodwill et al. (2015) compared ferrate(VI) and ferric resultant

particles produced in carbonate or phosphate buffered solutions and a real natural water

(reservoir). The particles from two different iron sources exhibited similar surface

charges, but had different size distributions in the buffered deionized water solutions.

Scanning electron microscope (SEM) revealed that the ferrate(VI) resultant particles

looked smoother and more granular. From X-ray photoelectron spectroscopy (XPS)

analysis Fe2O3 was observed in the ferrate(VI)-induced particles, which was not detected

in ferric chloride resultant particles. In the reservoir water, more nanoparticles with

negative surface charges were produced from the Fe(VI) addition than from ferric

addition. It is surprising that a large number of these particles remained suspended,

similar to stable colloids in natural water. Furthermore, Prucek et al. (2013, 2015) studied

ferrate(VI)-resultant particles in water and found that they were characterized with a

unique core (γ-Fe2O3)-shell (γ-FeOOH) structure.

Page 128: Ferrate as a New Treatment Chemical for Removal of ...

113

A question is raised. Do the Fe(VI)-induced particles sufficiently settle in a PFR?

If the iron precipitates produced from ferrate(VI) coagulation tend to settle in a PFR, the

PRR performs like a horizontal flow sedimentation tank, in which a large of iron

precipitates may settle down at the bottom, thus facilitating the removal of suspended

particles in wastewater and reducing the particulate loadings of downstream solid-liquid

separation processes (e.g. sand filter or membrane filtration units). However, if the

particles have a poor settling property, the Fe(VI)-induced particles increase the original

turbidity through the formation of additional iron particles, which requires an efficient

downstream solid-liquid separation process. Batch reactors cannot provide useful

information to answer this question, because the formation of settleable floc are tightly

associated with some factors that cannot be easily controlled in a batch mode, such as

hydraulic conditions.

Therefore, the objective of this chapter is to determine the treatment performance

of ferrate(VI) in a PFR and characterize the produced iron particles. Particularly,

turbidity, P, organic content in wastewater, and the settling properties of the produced

particles were studied.

6.2 Materials and methods

All the reagents used were at least analytical grade, except as noted. Potassium

ferrate (K2FeO4) (> 96%) was purchased from Sigma−Aldrich. Secondary effluent was

the effluent from a secondary clarifier prior to disinfection at a local municipal

wastewater treatment plant (Verona, New Jersey). The sample was delivered to Montclair

State University’s water treatment laboratory and stored at 4oC in a refrigerator until use.

Page 129: Ferrate as a New Treatment Chemical for Removal of ...

114

A 300 mg/L Fe(VI) stock solution was prepared by dissolving an appropriate mass of

K2FeO4 in deionized water. The stock solution pH was over 9.0 under which Fe(VI) was

relatively stable. The Fe(VI) concentration in the stock solution was confirmed with the

ABTS method (Lee et al. 2005b). The stock solution was freshly prepared every day.

Continuous flow reactor. A customized continuous-flow reactor was designed and built

as shown in Fig. 6-1. As seen, the reactor is composed of a CSTR (inlet zone) and a PFR

(reaction and settling zone, sludge zone and outlet zone). Secondary effluent and Fe(VI)

stock solution were continuously fed into the inlet zone through two peristaltic pumps

from a secondary effluent reservoir and a Fe(VI) stock solution container (Fig. 6-2),

respectively. The inlet is a chamber with a size of 4.78 ×4.45×13.97 cm (L×W×D), which

was installed on a magnetic stirrer. Secondary effluent and Fe(VI) stock solution were

sufficiently and instantly mixed in the inlet zone under a complete mixing (100 rpm). The

inlet served as a rapid mixer in WWTPs to ensure a rapid mixing between Fe(VI) and

secondary effluent. It should be noted that no iron precipitates settled down in the inlet

zone due to the mixing, though the reactions of Fe(VI) with secondary effluent

constituents were initiated. The inlet is separated from the following PFR with a slotted

plate, which ensures that water uniformly enters into the reaction and sludge zones.

The ensuing upper section include the reaction and settling zone as shown in Fig.

6-1, in which the added Fe(VI) continued its reactions with secondary effluent and the

iron precipitates were continuously produced. Ideally, water was under a plug-flow state

here (that is, there was not any horizontal mixing). Therefore, concentrations of the

chemicals (e.g. Fe(VI) and wastewater contaminants) were decreased with the increasing

Page 130: Ferrate as a New Treatment Chemical for Removal of ...

115

distance and contact time. Particles were primarily subject to two movements. Firstly,

they move horizontally with the flowing water. Secondly, they also move down vertically

(i.e. settle) because their gravity was greater than buoyant force. If the settling velocity

was sufficiently large, some particles settled down to the bottom before they flew through

the outlet. It is assumed that the iron particles settling to the bottom did not re-suspend.

Therefore, the zone below the reaction and settling zone is called the sludge zone, where

the iron sludge is stored. A slotted plate separates the reaction/settling and sludge zones

with the following outlet zone as shown in Fig. 6-1. The water entering into the outlet

zone overflew from the reactor through an outlet. As the water flow (arrow lines in Fig.

6-1) showed, the water entering into the outlet zone could be completely withdrawn to

the outlet. Therefore, the slotted plate was a crucial design to avoid the iron particles in

the sludge zone from re-suspending and escaping from the reactor.

To visualize the flow patterns in the continuous flow reactor, a food-grade dye

was added into the feeding water, as shown in Fig. 6-3. A sludge zone (no dye) was

observed in the section 3 cm above the bottom.

Continuous flow treatment experiments. The treatment was initiated in the inlet zone

in which secondary effluent and Fe(VI) stock solution were completely mixed. The total

flow rate (the sum of secondary effluent and Fe(VI) stock feeding rates) was fixed to

ensure a 2-hr retention time in the reactor. The feeding rate of the secondary effluent was

adjusted at 58.65, 58.15 and 57.65 ml/min. And the pumping rate of Fe(VI) stock

solution was varied at 0.5, 1 and 1.5 ml/min to achieve 2.50, 5.00 and 7.50 mg/L Fe(VI),

respectively. During the treatment, solution pH was not controlled. Over the different

Page 131: Ferrate as a New Treatment Chemical for Removal of ...

116

distances, four 5 mL aliquot samples were collected at 1.75~7 cm below the water table

for analysis. The different distances corresponded to different reaction times, as shown in

Eq. 6-1.

t = L/v (6-1)

Here, t is the contact time; L is distance; and v is the water speed.

Analyses. Solution pH was measured by a pH meter (Thermo Scientific Orion 5-Star

Plus). Ferrate(VI) was determined using the ABTS method (Lee et al. 2005b). In order to

determine the settling properties of iron particles, turbidity and particulate iron were

measured. In this study, particulate iron was operationally defined as Fe present in these

iron oxide particles that could not pass through a 0.45 µm membrane filter. Turbidity,

particle size, size distribution, zeta potential, particle count, and total iron were measured

using the unfiltered samples. Turbidity was quantified using a portable turbidity meter

(HACH, 2100Q). Zeta potentials (ZPs) were determined using a Nano Zetasizer

(Malvern, ZEN 3690) without any sample dilution. The measurement ranges were as

follows: Z-average sizes, 0.3 nm–10 microns; and ZPs for 3.8 nm - 100 micron particles.

Particle sizes and size distributions were measured using a Dynamic Imaging Particle

Analyzer (FlowCAM® VS Series). Iron was determined with an inductively coupled

plasma-mass spectroscopy (ICP-MS Thermo X-Series II, XO 472). Total iron was

measured after the unfiltered samples were completely digested (EPA Method 3005A).

Dissolved iron was measured after the samples were filtered through 0.45µm syringe

membrane filters (Thermo Scientific, cellulose acetate (CA), 30 mm diameter).

Page 132: Ferrate as a New Treatment Chemical for Removal of ...

117

Morphology of Fe(VI)-induced particles were determined with a Hitachi H-7500

transmission electron microscope (TEM). All the experiments were run in duplicates.

Fig. 6-1 Scheme of the continuous flow reactor design (CSTR (inlet zone) + PFR

(reaction zone + sludge zone + outlet zone)

Peristaltic pump

Secondary effluent reservoir

Fe(VI) stock solution

Magnetic stirrer

Magnetic stirring bar

Slotted plate

Outlet

Reaction zone

(settling zone)

Outlet zoneInlet zone

Sludge zone

Page 133: Ferrate as a New Treatment Chemical for Removal of ...

118

Fig. 6-2 The secondary effluent reservoir (left) and Fe(VI) stock solution container

(right)

Fig. 6-3 Flow patterns in the continuous flow reactor using a food-grade dye

Page 134: Ferrate as a New Treatment Chemical for Removal of ...

119

6.3 Results and Discussion

6.3.1 Quality of secondary effluent

Basic water quality parameters of the secondary effluent sample are shown in

Table 6-1. Solution was slightly alkaline (pH 7.23). In this study, turbidity (16.7 NTU),

particle size (1,005 nm), particle count (5,762/mL), and zeta potential (-9.37 mV) were

used to characterize particles in the sample. Turbidity (16.7 NTU) was selected as an

aggregate parameter to measure colloidal particles, rather than TSS, because turbidity

was more sensitive at a low amount of particles. Moreover, turbidity is more commonly

used for potable and reclaimed water. The suspended particles had a hydrodynamic size

of 1,005 nm, indicating that their size was in a micro-particle scale, but close to a nano

range. The zeta potential of -8.97 mV suggests that these micro-particles had slightly

negatively charges. Although the zeta potential did not fall within a range (> 20.0 mV or

< -20.0 mV) that extremely stabilizes colloidal particles in water, it appeared to

sufficiently keep these micro-particles suspended in secondary effluent.

The secondary effluent had a typical dissolved organic content (COD = 16.0

mg/L). UV254 at 0.123 cm-1 indicates the abundant presence of aromatic structures and/or

double bonds in EfOM, which preferentially react with ferrate(VI) as demonstrated in

Chapter 4. TP was 3.851 mg/L, 98% of which was reactive P. The level was significantly

below TP present in sewage because New Jersey WWTPs typically remove P from

wastewater using a chemical precipitation method.

Page 135: Ferrate as a New Treatment Chemical for Removal of ...

120

Table 6-1 Basic water quality parameters of secondary effluent

Parameter Value Parameter Value

pH 7.23 Dissolved P (mg/L) 3.762

Turbidity (NTU) 16.7 Particulate P (mg/L) 0.089

Hydrodynamic size (nm) 1,005 COD (mg/L) 21.0

Particle # (particles/mL) 1.32×104 UV254 (cm-1) 0.123

Zeta potential (mV) -9.37

Total Fe (mg/L) 0.270

Total P (mg/L) 3.851

6.3.2 Turbidity, morphology, and zeta potential of Fe(VI)-induced particles

Turbidity of Fe(VI) treated secondary effluent at different contact times with

different Fe(VI) doses is shown in Fig. 6-4. The initial turbidity was 6.87 NTU due to the

presence of TSS in untreated secondary effluent. However, the turbidity was increased to

16.70, 30.07, and 34.23 NTU at 2.50, 5.00, and 7.50 mg/L Fe(VI) at 10 min, respectively.

Thereafter, the turbidity almost stabilized until 120 min, regardless of the Fe(VI) dose. It

should be noted that the detention time in the inlet zone was around 10 min, in which the

solution was under a complete mixing state and the turbidity was constant. The turbidity

in Fig. 6-4 was almost the earliest contact time for the following PF reactor. The above

finding suggests that Fe(VI) addition created a large number of particles, of which most

were non-settable.

Page 136: Ferrate as a New Treatment Chemical for Removal of ...

121

TEM images of the suspended particles in the reaction and settling zone are

shown in Fig. 6-5(a) and (b). It is found that Fe(VI) reduction caused the formation of

numerous nanoparticles. The nanoscale particles had spherical shapes and uniform sizes

approximately ranging within 30-70 nm. Many nanoscale particles tended to aggregate

and produce a few micrometer flocs. As demonstrated from Fig. 6-4, these aggregates

had a poor settling velocity to significantly increase the water turbidity. Zeta potentials of

these suspended Fe(VI)-induced iron nanoparticles in the reaction and settling zone were

measured under different experimental conditions (Fig. 6-6). The zeta potentials were

slightly increased from the initial -9.37 mV to -8.23, -8.45 and -7.88 mV at 2.50, 5.00,

and 7.50 mg/L Fe(VI) at 120 min, respectively. The zeta potentials allowed these

particles under a relatively stable state.

Fig. 6-4 Residual turbidity at different contact times with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

45.0

0 20 40 60 80 100 120 140

Tu

rbid

ity

(N

TU

)

Contact time (min)

2.50 mg/L Fe(VI) 5.00 mg/L Fe(VI) 7.50 mg/L Fe(VI)

Page 137: Ferrate as a New Treatment Chemical for Removal of ...

122

(a)

(b)

Fig. 6-5 TEM images of suspended Fe(VI)-induced particles in the reaction and settling

zone during Fe(VI) treatment of secondary effluent (initial pH = 7.23, Fe(VI) = 5.00

mg/L, and contact time = 120 min)

100 nm

100 nm

Page 138: Ferrate as a New Treatment Chemical for Removal of ...

123

Fig. 6-6 Zeta potentials of suspended particles in the reaction and settling zone during

Fe(VI) treatment of secondary effluent (initial pH = 7.23; relative standard deviations are

below 5%, not shown in the figure)

-20.00

-18.00

-16.00

-14.00

-12.00

-10.00

-8.00

-6.00

-4.00

-2.00

0.00

0 20 40 60 80 100 120 140

Zet

a p

ote

nti

al

(mV

)

Contact time (min)

2.50 mg/L Fe(VI)

5.00 mg/L Fe(VI)

7.50 mg/L Fe(VI)

Page 139: Ferrate as a New Treatment Chemical for Removal of ...

124

6.3.3 Particle counting and size of Fe(VI)-induced particles

To further understand formation and coagulative behaviors of Fe(VI)-induced

particles, the numbers of suspended particles were counted using the Dynamic Imaging

Particle Analysis technique, which captures the images of particles in water using a non-

invasive approach. Numbers of suspended particles in secondary effluent at different

contact times are shown in Fig. 6-7 (5.00 mg/L Fe(VI)). The particle number of untreated

secondary effluent was 1.32×104 particles/mL, which was primarily caused by TSS. Of

interest, the profile of particle number with time was composed of two phases. In the 1st

phase (0-300 min), the number was linearly increased to 1.14 ×105 particles/mL as the

contact time reached 30 min (Particle # = 3210 t + 21830, R2 = 0.95). In the 2nd phase, the

particle number linearly dropped to 7.91 ×104 particles/mL with the increasing contact

time to 120 min (Particle # = -380 t + 123545, R2 = 0.95). The variation of particle

number was primarily due to the competition between the formation and aggregation

rates of new Fe(VI)-induced particles. As demonstrated previously, Fe(VI) gradually

decomposed due to self-decomposition and reactions with certain reducing agents to

continuously produce Fe(III), which immediately formed precipitates (nanoparticles in

Fig. 6-5) to contribute to the increase in the particle number. In contrast, the produced

Fe(VI)-induced nanoparticle tended to aggregate to form larger particles, thereby

reducing the particle number. In the 1st phase, the formation rate of new Fe(VI)-induced

particles appeared to be greater than the aggregation rate, because a majority of Fe(VI)

decayed within the first 30 min as demonstrated before. This led to an increase in the net

number of particles. Afterwards, the formation rate of Fe(III) was substantially reduced

Page 140: Ferrate as a New Treatment Chemical for Removal of ...

125

so that the aggregation predominated in the system, which caused an decreasing net

particle number with time.

Size distributions of suspended particles in secondary effluent during Fe(VI)

treatment are shown in Fig. 6-8. At any specific contact time, the particle size was almost

normally distributed. Although the particle numbers varied with contact time, the size

distribution curves were very similar: the peaks were observed around 0.5 µm; and the

most of particles had a size below 5.0 µm. These findings, in addition to the observation

in Fig. 6-4, indicate that the size distribution and turbidity were almost consistent within

120 min and were not influenced by the varied particle number.

Page 141: Ferrate as a New Treatment Chemical for Removal of ...

126

Fig. 6-7 Numbers of suspended particles in secondary effluent during Fe(VI) treatment of

secondary effluent (initial pH = 7.23; 5.00 mg/L Fe(VI); relative standard deviations are

below 5%, not shown in the figure)

Fig. 6-8 Size distributions of suspended particles in secondary effluent during Fe(VI)

treatment of secondary effluent (initial pH = 7.23; 5.00 mg/L Fe(VI))

Particle # = 3210 t + 21830

R² = 0.95

Particle # = -380 t + 123545

R² = 0.99

0.00E+00

2.00E+04

4.00E+04

6.00E+04

8.00E+04

1.00E+05

1.20E+05

1.40E+05

0 20 40 60 80 100 120 140

Part

icle

cou

n (

part

icle

s/m

L)

Contact time (min)

0

5000

10000

15000

20000

25000

30000

35000

40000

45000

50000

0.00 1.00 2.00 3.00 4.00 5.00

Part

icle

cou

nt

(part

icle

s/m

L)

Size (µm)

0 min10 min30 min60 min90 min120 min

Page 142: Ferrate as a New Treatment Chemical for Removal of ...

127

6.3.4 Settleability of Fe(VI)-induced particles

To better evaluate the settleability properties of Fe(VI)-induced particles under

different experimental conditions, the particulate iron content in treated secondary

effluent was studied. In this study, the concentrations of total and dissolved Fe in

secondary effluent were measured. Because particulate Fe = total Fe – dissolved Fe, the

particulate Fe could be accordingly determined. Concentrations of total and particulate Fe

in treated secondary effluent at different contact times are shown in Fig. 6-9 and 6-10,

respectively. It is found that a majority of total Fe existed in the form of solid, because

Fe(III), once formed from Fe(VI) reduction, tended to precipitate at the study pH range.

As seen in Fig. 6-10, at 2.50 mg/L Fe(VI), the particulate Fe was increased to

2.296 mg/L at 10 min, representing the level of total Fe-containing particles at 10 min.

Thereafter, the particulate Fe gradually decreased to 1.394 mg/L, suggested that 45%

particulate Fe settled down. At 5.00 and 7.50 mg/L Fe(VI), the particulate increased to

4.509 and 4.268 mg/L at 10 min and reached 4.869 and 4.307 mg/L at 120 min,

respectively. The fractions of particulate Fe that was removed through sedimentation at

120 min were 3% and 43% for 5.00 and 7.50 mg/L, respectively.

Page 143: Ferrate as a New Treatment Chemical for Removal of ...

128

Fig. 6-9 Total Fe in the reaction and settling zone with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

Fig. 6-10 Particulate Fe in the reaction and settling zone with different Fe(VI) doses

during ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

0.000

1.000

2.000

3.000

4.000

5.000

6.000

7.000

8.000

0 20 40 60 80 100 120 140

To

tal

Fe

(mg

/L)

Contact time (min)

2.50 mg/L Fe(VI) 5.00 mg/L Fe(VI) 7.50 mg/L Fe(VI)

0.000

1.000

2.000

3.000

4.000

5.000

6.000

7.000

8.000

0 20 40 60 80 100 120 140

Pa

rtic

ula

te F

e (m

g/L

)

Contact time (min)

2.50 mg/L Fe(VI) 5.00 mg/L Fe(VI) 7.50 mg/L Fe(VI)

Page 144: Ferrate as a New Treatment Chemical for Removal of ...

129

6.3.5 P in Fe(VI) treated secondary effluent

Total and dissolved P were also measured during Fe(VI) treatment of secondary

effluent, as shown in Fig. 6-11 and 6-12, respectively. Experimental analysis showed that

the reactive phosphorus (i.e. orthophosphate) accounted for 98% of total P, while the

remaining 2% was nonreactive (poly or organic phosphorus). And almost all the P existed

in a dissolved state. As seen in Fig. 6-11, over 120 min, TP in the secondary effluent in

the reaction and settling zone was not obviously altered at 2.50 mg/L, while TP at 5.00

and 7.50 mg/L Fe(VI) slightly dropped to 3.209 and 3.129 mg/L, respectively, both

corresponding to a 85% TP removal through sedimentation. The finding suggests that a

majority of TP flew through the continuous-flow reactor and existed in the effluent.

At Fig. 6-12, dissolved P dropped to 3.308, 2.330, and 1.917 mg/L at 2.50, 5.00,

and 7.50 mg/L at 10 min, respectively, indicating that 12%, 38%, and 49% dissolved P

were transformed to a particulate state. As the treatment proceeded until 120 min,

dissolved P further decreased to 2.982, 2.222, and 1.57 mg/L, respectively, indicating that

21%, 41%, and 58% of dissolved P became particulate P. Of extreme interest, the ratio of

the transformed dissolved P to Fe(VI) dose was 0.30 mg dissolved P removed /mg Fe (i.e.

0.54 mM dissolved P removed/1.00 mM Fe) , regardless of the ferrate(VI) dose.

Generally speaking, dissolved orthophosphate can be transformed to a particulate state by

Fe(III) through two possible pathways, i.e. the direct precipitation between Fe(III) and

phosphate (Eq.(6-2)) and adsorption of P to iron hydroxides through surface

complexation.

Fe(III) + PO43- = FePO4 ↓ (6-2)

Page 145: Ferrate as a New Treatment Chemical for Removal of ...

130

The first mechanism requires a theoretic molar ratio of dissolved P removed to Fe at 1:1.

Moreover, the second mechanism would provide additional P removal. Therefore, the

overall theoretic molar ratio should be greater than 1:1. However, the observed ratio in

this study was 0.54:1.00, likely because Fe(III) produced from Fe(VI) reduction is also

subject to other competition reactions, such as hydrolysis. It is technically difficult to

separate contributions of the two different mechanisms in this study.

Page 146: Ferrate as a New Treatment Chemical for Removal of ...

131

Fig. 6-11 Total P in the reaction and settling zone with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

Fig. 6-12 Dissolved P in the reaction and settling zone with different Fe(VI) doses during

ferrate(VI) treatment of secondary effluent (initial pH = 7.23)

0.000

0.500

1.000

1.500

2.000

2.500

3.000

3.500

4.000

4.500

5.000

0 20 40 60 80 100 120 140

TP

(m

g/L

)

Contact time (min)

2.50 mg/L Fe(VI) 5.00 mg/L Fe(VI) 7.50 mg/L Fe(VI)

0.000

0.500

1.000

1.500

2.000

2.500

3.000

3.500

4.000

4.500

0 20 40 60 80 100 120 140

DP

(m

g/L

)

Contact time (min)

2.50 mg/L Fe(VI) 5.00 mg/L Fe(VI) 7.50 mg/L Fe(VI)

Page 147: Ferrate as a New Treatment Chemical for Removal of ...

132

6.3.6 EfOM in Fe(VI) treated secondary effluent

Measurement of COD in the presence of ferrate(VI) is technically difficult,

because ferrate(VI) (an oxidant) can underestimate the solution COD after the sample is

added to the HACH COD test kits (Deng 2007). Therefore, only the COD data in effluent

is discussed here. COD removal efficiencies in effluent at different ferrate(VI) doses are

shown in Fig. 6-13. The COD removals were 22%, 40%, and 38% at 2.50, 5.00, and

7.500 mg/L, respectively. It should be noted that the COD removal was dramatically

increased at a low ferrate(VI) range (0.00-5.00 mg/L), but almost stabilized at a high

ferrate(VI) dose (> 5.00 mg/L). The trend was in agreement with the finding from batch

tests in Chapter 4, again validating that ferrate(VI) could effectively alleviate EfOM in

secondary effluent. However, the observed COD removal was greater than the COD

removal achieved in Chapter 4. One or more of the following reasons may contribute to

the disparate observations. Firstly, a lower pH (pH 7.23, the original pH) in this chapter

was applied than that in Chapter 4 (pH 7.50). It is well known that ferrate(VI) is more

reactive at a lower pH. Therefore, it is not surprising that more COD was removed in the

PFR flow reactor. Secondly, the two secondary effluent samples had different initial

COD (21.0 mg/L in this study vs. 31.7-33.1 mg/L in Chapter 4). Thirdly, chemical

compositions might be different between the two batches of secondary effluent samples,

e.g. the nature of EfOM (e.g. MW distribution) and the concentrations of potentially

competing reducing species.

Page 148: Ferrate as a New Treatment Chemical for Removal of ...

133

Fig. 6-13 COD removal with different Fe(VI) doses during ferrate(VI) treatment of

secondary effluent (initial pH = 7.23, COD0 = 21.0 mg/L, and reaction time = 2 hr)

0%

10%

20%

30%

40%

50%

2.50 mg/L 5.00 mg/L 7.50 mg/L

CO

D r

emoval

Fe(VI)

Page 149: Ferrate as a New Treatment Chemical for Removal of ...

134

6.4 Conclusions

The major conclusions of this chapter are described as follows.

1) Ferrate(VI) reduction in secondary effluent produces numerous nanoscale iron

particles in a continuous-flow reactor. These nanoparticles tend to aggregate to

microscale particles.

2) A majority of these micro-particles remained suspended due to negative surface

charge, thereby increasing effluent turbidity.

3) Number of Fe(VI)-induced particles is initially increased because a large number

of Fe(III) is in-situ produced to form new iron oxide particles at the initial phase.

Thereafter, as the Fe(III) production rate decreases and the particle aggregation

rate prevails in the system, the particle number gradually drops in the ensuing

phase. However, the variation in particle number does not significantly alter water

turbidity.

4) Sizes of Fe(VI)-induced particles at different contact times are almost normally

distributed with their size peaks around 0.5 µm.

5) In this study, 0.30 mg/L dissolved P is transformed to a particulate state by 1.00

mg/L Fe(VI), regardless of Fe(VI) dose. The transformation is likely due to

chemical precipitation and/or adsorption to iron precipitates.

6) COD was removed be 5.00 mg/L Fe(VI) by 40% in this study, validating that

EfOM could be largely removed by ferrate(VI).

Page 150: Ferrate as a New Treatment Chemical for Removal of ...

135

6.5 References

Carr, J.D., Kelter, P.B., Tabatabai, A., Splichal, D., Erickson, J. and McLaughlin, C. (1985)

Properties of ferrate (VI) in aqueous solution: an alternate oxidant in wastewater treatment, pp.

1285-1298.

Carr, J.D. (2008) Kinetics and product identification of oxidation by ferrate (VI) of water and

aqueous nitrogen containing solutes, pp. 189-196, Oxford University Press.

Deng, Y. (2007) Physical and oxidative removal of organics during Fenton treatment of mature

municipal landfill leachate. Journal of Hazardous Materials 146(1), 334-340.

Goodwill, J.E., Jiang, Y., Reckhow, D.A., Gikonyo, J. and Tobiason, J.E. (2015) Characterization

of Particles from Ferrate Preoxidation. Environmental Science & Technology 49(8), 4955-4962.

Graham, N.J.D., Khoi, T.T. and Jiang, J.Q. (2010) Oxidation and coagulation of humic

substances by potassium ferrate. Water Science and Technology 62(4), 929-936.

Jiang, J.Q. and Lloyd, B. (2002) Progress in the development and use of ferrate(VI) salt as an

oxidant and coagulant for water and wastewater treatment. Water Research 36(6), 1397-1408.

Jiang, J.Q. (2007) Research progress in the use of ferrate(VI) for the environmental remediation.

Journal of Hazardous Materials 146(3), 617-623.

Lee, Y., Yoon, J. and von Gunten, U. (2005) Spectrophotometric determination of ferrate (Fe(Vl))

in water by ABTS. Water Research 39(10), 1946-1953.

Lee, Y., Kissner, R. and von Gunten, U. (2014) Reaction of ferrate (VI) with ABTS and self-

decay of ferrate (VI): Kinetics and mechanisms. Environmental Science & Technology 48(9),

5154-5162.

Prucek, R., Tucek, J., Kolarik, J., Filip, J., Marusak, Z., Sharma, V.K. and Zboril, R. (2013)

Ferrate(VI)-Induced Arsenite and Arsenate Removal by In Situ Structural Incorporation into

Magnetic Iron(III) Oxide Nanoparticles. Environmental Science & Technology 47(7), 3283-3292.

Prucek, R., Tucek, J., Kolarik, J., Huskova, I., Filip, J., Varma, R.S., Sharma, V.K. and Zboril, R.

(2015) Ferrate (VI)-Prompted Removal of Metals in Aqueous Media: Mechanistic Delineation of

Enhanced Efficiency via Metal Entrenchment in Magnetic Oxides. Environmental Science &

Technology 49(4), 2319-2327.

Sharma, V.K., Yngard, R.A., Cabelli, D.E. and Baum, J.C. (2008) Ferrate(VI) and ferrate(V)

oxidation of cyanide, thiocyanate, and copper(I) cyanide. Radiation Physics and Chemistry 77(6),

761-767.

Page 151: Ferrate as a New Treatment Chemical for Removal of ...

136

Tien, K. and Graham, N. (2011) Evaluating the Coagulation Performance of Ferrate: Coagulation

of Kaolin Suspension, Kumamoto, Japan.

Page 152: Ferrate as a New Treatment Chemical for Removal of ...

137

CHAPTER 7 ENVIRONMENTAL IMPLICATIONS

7.1 Preliminary cost analysis

Cost is a key factor for a technology application. Although a few hundreds of

publications on ferrate(VI) studies for environmental applications have been available,

economic information regarding ferrate(VI) treatment is extremely limited (Waite

2012a). An effort of this dissertation is to preliminarily explore costs associated with

ferrate(VI) treatment of secondary effluent. The expenses associated with capital and

operating/maintenance (O&M) costs need to be considered. In order to have a clear

picture on the costs for ferrate(VI) treatment, the estimated expenses would be compared

with those from existing treatments that provide an equivalent treatment in water reuse.

Ferrate(VI) serves as a multi-treatment agent. The primary function of ferrate(VI) for

water reclamation is chemical oxidation for the degradation of EfOM and emerging

contaminants. Among existing chemical oxidation technologies for water reuse,

ozonation plays a very similar role because:1) ozonation is also a selective oxidant that

favorably reacts with ERMs (Lee and von Gunten 2010, AWWA 1991); 2) ozone and

ferrate(VI) have similar reduction potentials (2.07 V for O3 vs. 2.20 V for ferrate(VI));

and 3) the COD removal efficiencies achieved by ozone are very similar to those by

ferrate(VI) in this study (Tripathi et al. 2011, Domenjoud et al. 2011). Therefore,

ozonation is selected in this dissertation as the existing chemical oxidation process for the

comparison.

The secondary purpose of ferrate(VI) treatment of secondary effluent is the

alleviation of phosphate. In practices, the most commonly used method for the phosphate

Page 153: Ferrate as a New Treatment Chemical for Removal of ...

138

elimination is chemical precipitation with a coagulant (e.g. ferric chloride)

(Tchobanoglous et al. 2003, Wilfert et al. 2015). This technology is particularly

commonly applied at WWTPs in New Jersey, United States. Ferric chloride can

transform dissolve phosphate into a solid phase that can be subsequently removed

through a solid-liquid separation process such as sand filtration. It shares a similar

treatment mechanism with ferrate(VI) for the removal of phosphate from secondary

effluent. Therefore, ferric chloride-driven coagulation is selected as the existing treatment

option for phosphate removal for the comparison purpose.

Cost analysis for the two treatment options (Option1: ferrate(VI) treatment alone;

Option 2: ozonation combined with ferric chloride precipitation) are made below. The

treatment performance data of ferrate(VI) is obtained from the continuous-flow reactor in

Chapter 6. It is assumed that ferrate(VI) can provide an equivalent treatment of ozonation

combined with ferric chloride precipitation. Ferrate(VI) treatment and O3 + FeCl3 are

compared in terms of the expenses (capital and O&M costs) spent for treatment of 1,000

gallons of secondary effluent. Data on chemical price, equipment, and treatment

performance of ozonation and ferric chloride precipitation are obtained from literature.

All the costs are expressed as dollars in 2016. The inflation rate is 1.6% which is an

average one in the United States over 2006-2016.

Cost estimation for Option 1- ferrate(VI) treatment. Our continuous- flow reactor

tests showed that ferrate(VI) at 5.0 mg/L Fe(VI) could reduce COD from 21 mg/L to 12.6

mg/L with a decrease by 8.4 mg/L COD, corresponding to 40% COD removal. EPA has

recently estimated the cost of ferrate(VI) treatment using a set of Ferrator® (on-site

Page 154: Ferrate as a New Treatment Chemical for Removal of ...

139

ferrate(VI) generator) systems (Cashman et al. 2014). The annual cost including

electricity usage, chemical inputs, incidental repairs, and amortized capital investment

(15-yr lifetime) is $833,300/yr (dollars in 2014) for a capacity of 10 MGD at a dose of

5.00 mg/L Fe(VI). The unit cost is $0.23/1000 gal treated water (dollars in 2014). The

corresponding value in 2016 is $0.24/1000 gal, which is calculated as below.

PV = FV / (1 + r)Y (Eq. 7-1)

Where PV is the present value (= starting principal), FV is the future value, r the inflation

rate (1.6%), and Y is the number of years for investment.

Cost estimation for Option 2- ozonation treatment and ferric chloride

precipitation. Ozonation can provide a very similar treatment result for EfOM.

Domenjoud et al. (2011) reported that 5.00 mg/L O3 could reduce a secondary effluent

COD approximately by 14%, corresponding to a COD reduction by 11.0 mg/L (CODo =

78.4 mg/L). Amortized costs for ozonation at 5.00 mg/L O3 in water treatment was

estimated at $3.48/1000 gal (dollars in 1998). The corresponding value in 2016 is

$4.13/1000 gal.

Ferric chloride can be added during primary or secondary wastewater treatment in

engineering practices. The produced precipitates can be removed together with TSS in a

primary or secondary settling tank. Therefore, any special mixing equipment or reactor is

not required to achieve the precipitation. In this dissertation, only chemical cost (the

major O&M cost) is considered.

The average price of ferric chloride (100% grade) was $325/US ton in the United

States chemical market in 2006 (ICIS 2006, Van Savage 2001), corresponding to

Page 155: Ferrate as a New Treatment Chemical for Removal of ...

140

$381/US ton (dollars in 2016). One US ton is 907.19 kg. So the unit price for FeCl3 is

$0.0012/g as Fe. When chemical dose is 5.0 mg/L Fe(III), the treatment cost of ferric

precipitation is $0.02/1000 gal.

Cost comparison. Cost estimates for the two treatment options (i.e. ferrate(VI)

treatment and ozonation combined with ferric chloride precipitation) for water

reclamation are summarized in Table 7-1. Unit costs for ferrate(VI) treatment and O3 +

FeCl3 are $0.25/1000 gal and $4.15/1000 gal, respectively. Therefore, ferrate(VI)

treatment is a much cost-effective option, accounting for approximately 6% of ozonation

+ ferric chloride precipitation. Although costs in practices are largely influenced by

many site-specific factors (e.g. material and transportation costs), the preliminary cost

analyses indicates the economic competitiveness of ferrate(VI) treatment in water

reclamation.

The majority of ferrate(VI) treatment cost is the price of ferrate(VI) generators. A

ferrate(VI) generator unit costs approximately $836,127 (2016 dollar) with a capacity of

10 MGD (Cashman et al. 2014). On the other hand, ozonation contributes to 99.5%

expense of the second option. Ozonation is a costly wastewater treatment technology.

The major costs during ozonation come from ozone generators, oxygen feed gas,

compressors, and energy consumption (EPA 1999). Furthermore, the transfer efficiency

of O3 generation from oxygen gas is a critical economic consideration. This requires a

deep and tightly covered contact chambers for the reactions, which leads to a high capital

costs in design, materials and construction.

Page 156: Ferrate as a New Treatment Chemical for Removal of ...

141

Table 7-1 Cost comparison between Fe(VI) treatment and ozonation + FeCl3 for

treatment of secondary effluent

Treatment option Treatment unit Cost ($/1000 gal)1,2 Total cost ($/1000 gal)

1 Ferrate(VI) 0.24 0.24

2 Ozonation 4.13 4.15

FeCl3 precipitation 0.02

1Cost is the expense spent for treatment of 1000-gallon secondary effluent; 2Price is expressed in dollars in 2016.

Recently, an EPA project (Cashman et al. 2014) compared ferrate(VI) technology

with two other treatment trains, including traditional water treatment processes

(coagulation + sedimentation + filtration + chlorination) plus granular activated carbon

(GAC) adsorption and for drinking water treatment. Cost analysis showed that

ferrate(VI) technology was the least costly among the candidate treatment options, which

is in consistence with the finding in this study. However, it should be noted that

transportation, treatment, and disposal of ferrate(VI)-induced sludge was not considered

in this preliminary cost analysis.

7.2 Implications in Environmental Management

This dissertation research made an effort to explore a new water reclamation

process with an environmentally friendly treatment agent – ferrate(VI). Results from this

study suggests that ferrate(VI) has a potential to revolutionize water reclamation

technologies and bring about profound impacts on environmental management.

Page 157: Ferrate as a New Treatment Chemical for Removal of ...

142

7.2.1 Implication to water reuse industry

Although the demand for safe water is increasing with the increasing population,

growing agricultural and industrial needs, and increasingly serious pollution, there is not

an obvious increase in reliable water supply to match the growing demand. As a result of

overused water resources and changed climate patterns, available fresh water is

decreasing. Undoubtedly, water reuse serves as a sound approach to closing the gap

between the demand and supply and supporting the sustainable development of our

society. It should be noted that the market for water reuse technologies reached $29

billion in 2010 and increased to $57 billion in 2015. Development of new-generation

water reuse technologies is crucial to water reuse industry. Ferrate(VI)-based treatment

represents a new direction due to its technical and economic competitiveness.

Findings from this dissertation provides a scientific basis for industrialization of

ferrate(VI)-based treatment process for water reuse industry. Data acquired from this

study is critically important to engineering design for the ferrate(VI) application at least

in the following aspects.

1) Data from the experiments to evaluate the treatability of ferrate(VI) for

different wastewater contaminants demonstrate the multiple treatment

functions of ferrate(VI). It may be of extreme interest to the industry because:

a) it can achieve different treatment goals (e.g. disinfection, precipitation of

phosphate, elimination of emerging contaminants, and EfOM degradation) in

a single reactor, thereby simplifying the engineering design and operation; b)

it requires a much smaller space to accommodate a reactor, rather than a

Page 158: Ferrate as a New Treatment Chemical for Removal of ...

143

treatment train composed of different treatment units, which typically need a

large space for the reactor installation. This may be of significance to water

reclamation facilities in remote areas where less labor is needed to operate and

maintain the treatment systems, as well as in urban areas where land space has

a high priority in engineering investment.

2) Kinetics data of ferrate(VI) decomposition in a secondary effluent matrix

enables the appropriate sizing of a ferrate(VI) reactor. In this study, ferrate(VI)

was observed to follow a 2nd order reaction pattern, and key kinetics data were

determined. The data can be used to determine ferrate(VI) oxidant exposure

and life time in a reactor, which is important to correctly size a ferrate(VI)

reactor in engineering design.

3) Ferrate(VI) removals of EfOM and P were investigated and the effects of

operating factors were evaluated. The knowledge is useful when ferrate(VI) is

used prior to RO filtration during water reclamation. Membrane filtration is

currently commanding 70% water reuse market share. Both EfOM and P

significantly contribute to RO fouling, which is a key barrier to RO

technologies, reducing permeate flux, increasing energy consumption, and

leading to frequent chemical cleaning and even membrane replacement.

EfOM and P can serve as substrate and nutrient to enhance microbial growth

on membranes, respectively. They also contribute to organic and inorganic

membrane fouling, respectively (Vrouwenvelder et al. 2010). For example,

phosphate can precipitate calcium that deposits on RO membrane to reduce

Page 159: Ferrate as a New Treatment Chemical for Removal of ...

144

water reflux (Tang et al. 2011, Ning and Troyer 2007). Ferrate(VI) application

is capable of alleviating such membrane fouling through mitigation of EfOM

and P in the RO influent.

4) Data on ferrate(VI) decomposition of MEF again validates that ferrate(VI)

oxidation is a powerful tool to screen unregulated emerging contaminants

from treated wastewater. Identification of MEF degradation byproducts also

reveals that ferrate(VI)-driven oxidation has a different degradation pattern

from other chemical oxidation processes. This finding suggests that ferrate(VI)

may be an alternative for some emerging compounds, which are persistent to

traditional chemical oxidants.

5) Cost analysis provides water reuse industry a picture on affordable expenses

for ferrate(VI) applications for water reclamation. The quantitative

information, though preliminary, clearly indicates that ferrate(VI) is more cost

competitive than traditional ozonation and chemical precipitation technologies.

Based on the preliminary cost analyses in this dissertation, ferrate(VI)

treatment saves $3.91/1000 gal than ozonation combined with ferric

precipitation. Therefore, for a 10 MGD water reclamation facility (a small

treatment capacity), ferrate(VI) treatment can reduce capital and O&M costs

by $14.3 million (dollars in 2016) annually. Unquestionably, the economic

advantage allows the new treatment option more competitive in the water

reuse market and encourages more major intended end users, e.g. water

Page 160: Ferrate as a New Treatment Chemical for Removal of ...

145

managers/planners, engineers, and water reclamation facilities, to accept and

apply the affordable treatment.

7.2.2 Environmental Benefits

Ferrate(VI)-based technologies for water reuse are beneficial for environmental

quality from two aspects. Firstly, ferrate(VI) treatment directly removes different

pollutants from reclaimed water and significantly alleviates these pollutant loadings into

the environment (for example, the pollutants in reclaimed are mitigated when it is reused

for agricultural irrigation and wetland restoration), thus safeguarding ecological health.

This dissertation has demonstrated that ferrate(VI) is capable of effectively removing

several pollutants of particular concern, such as phosphorus (a nutrient to cause algal

blooming), pathogens, dissolved organic matter, and emerging contaminants. The

reduction of these unwanted microbes, organic substances and nutrients can protect the

health of ecological systems and alleviate the effects of these stressors on vulnerable

wildlife populations.

Secondly, ferrate(VI) treatment has a potential to bring about less environmental

impact than other technologies that provide similar treatment effects. Life cycle

assessment (LCA) was recently made to compare ferrate(VI) and other oxidation

technologies for water disinfection (Cashman et al. 2014). This study compared

ferrate(VI)-based water treatment technology with conventional water treatment scenario

(coagulation + sedimentation + filtration + chlorination) in addition to granular activated

carbon adsorption in a representative moderate-sized water treatment facility. Results

show that ferrate(VI) reduced global warming potential, smog formation, energy demand,

Page 161: Ferrate as a New Treatment Chemical for Removal of ...

146

fossil depletion, acidification, human health criteria impacts, ozone depletion impacts,

metal depletion results, human health non-cancer results, human health cancer results,

and ecotoxicity results are decreased by 7%, 10%, 8%, 4%, 5%, 9%, 15%, 15%, 36%,

11%, and 12%, respectively (Cashman et al. 2014). It is also expected that ferrate(VI) in

water reuse similarly has less negative influence on the environment and provides a more

environmentally friendly water reuse treatment option.

7.2.3 Social benefits

As a result of the shortage of clean fresh water resources, the public and policy

makers have paid more attention to water reclamation. United States EPA and many state

level agencies (e.g. NJ Department of Environmental Protection) have issued their

guidance to highly encourage water reuse. The efforts from this dissertation research will

have multiple social impacts.

1) Water supply reliability. As a technically reliable and cost competitive water

reclamation treatment option, ferrate(VI) technology enables locally produced

sewage as a water reuse source and thus mitigates the dependence upon water

conveyance from remote sources. This benefit is of significance for the U.S.

arid states such as California, Nevada, Arizona, and Texas.

2) Local economic influence. Safe and reliable water can stimulate the economy

by providing local business with the assurance of water supplies for

agricultural, manufacturing, recreational, or other activities. This will

strongly support the creation of more job opportunities.

Page 162: Ferrate as a New Treatment Chemical for Removal of ...

147

3) Life quality and public health. The aforementioned different environmental

benefits from the environmentally friendly water reuse technology improve

public health and allow for increasing more recreational spaces, thereby

creating more livable and resilient communities.

4) Sustainability awareness. Water reuse represents a sustainable approach to

addressing water shortage issues in arid and water-stressed areas. Success

implementation of new and green water reclamation projects will help expand

the public awareness and understanding of sustainability in their normal living.

5) Aesthetics. Safe recycled water can be used to irrigate parks, golf courses and

other recreational facilities (e.g. fountains and lakes) to create aesthetic values.

7.3 Limitations of Ferrate(VI) Technology

Although ferrate(VI) treatment brings about many benefits, making it

advantageous over many existing treatment options, its application is restricted, more or

less, in the following five aspects. To overcome these issues will enable the emerging

treatment technology to be widely applied in engineering practices. Firstly, commercial

ferrate(VI) suppliers are very limited. To the best of knowledge, there is only one

ferrate(VI) manufacturer, i.e. the Ferrate Treatment Technology (FTT), LLC (Orlando,

FL), in the U.S. market. The on-site ferrate(VI) generator from FTT is called Ferrator,

which produces sodium ferrate solution via the wet chemistry method. Although it can

continuously produce ferrate solution on site, the ferrate(VI) solution is alkaline, under

which ferrate(VI) remains stable. Addition of the ferrate solution may cause a pH

increase in the treated water and wastewater. Secondly, ferrate(VI) treatment produces

Page 163: Ferrate as a New Treatment Chemical for Removal of ...

148

iron sludge as a final product. The sludge should be removed using a solid-liquid

separation method and then appropriately disposed of. The additional treatment and

disposal increases the complexity of ferrate(VI) treatment and the overall costs. Thirdly,

ferrate(VI) reactivity is highly pH sensitive. Ferrate(VI) oxidation is much more reactive

at an acidic condition than at an alkaline condition. Therefore, ferrate(VI) treatment is

very effective for the removal of many organic pollutants at a low pH. However, the

treatment may be substantially discounted at an alkaline environment. In practices, this

may be problematic when water or wastewater has an original pH within an alkaline

range. Fourthly, ferrate(VI) is highly unstable at an acidic condition. Ferrate(VI) can

rapidly self-decompose at an acidic condition, thereby reducing ferrate(VI) oxidant

exposure. The quick self-decomposition can compete with target pollutants for available

ferrate(VI) to reduce the expected treatment efficiency. Fifthly, ferrate(VI) self-

decomposition in water can release OH- and increase pH. The extent of pH increase

depends on the amount of produced OH- and solution chemistry (e.g. acidity). The pH

increase may be substantial at a high ferrate(VI) dose and for a water with a low pH

buffer capacity. If this is the case, pH adjustment has to be used to satisfy the pH

requirement in effluent.

7.4 Overall Conclusions

Ferrate(VI) treatment as a new-generation water reclamation technology has a

potential to revolutionize the water reuse industry. A principal advantage of ferrate(VI)

treatment is to achieve multiple treatment goals with a single reactor and thus bring about

many attractive benefits, such as simple system design and operation, low investment,

Page 164: Ferrate as a New Treatment Chemical for Removal of ...

149

and small physical footprint. This dissertation, in addition to literature, demonstrates the

capability of ferrate(VI) for mitigation and elimination of sewage organic matter, total

suspended particles, phosphorus, pathogenic microorganism, and unregulated micro-

pollutants from treated wastewater. The treatment performance relies heavily upon two

operating factors, i.e. ferrate(VI) dose and solution pH. Their effects on the treatment

performance need to be carefully evaluated when ferrate(VI) treatment is used in full

scale applications.

The major expected mechanism from ferrate(VI) is chemical oxidation. In many

previous studies, ferrate(VI) has been long used as a chemical oxidant alone. It is capable

of chemically transforming EfOM in secondary effluent. Although its capability to

mineralize these organic compounds is limited, Fe(VI) preferentially attacks aromatic

structures and double bonds to significantly lower UV254 . This can reduce the formation

potential of DBPs because strong UV-absorbing EfOM compounds characterized with

high UV254 absorbance generally serve as major DBP precursors. The production and

presence of DBPs in reclaimed water are recognized as a great challenge in water reuse,

because the toxic byproducts pose a threat to public health and reduce the safety of

reclaimed water. The other benefit is to mitigate EfOM-caused fouling for RO membrane

filtration if ferrate(VI) is combined with RO in water reclamation practices (RO treatment

accounts for 70% of the U.S. water reuse market). EfOM can organically foul RO

membrane materials and also enhance the growth of biofilms on RO membrane as a

substrate. On the other hand, ferrate(VI)-driven oxidation can simultaneously oxidize

unregulated wastewater-driven micro-pollutants such as pharmaceutical and personal care

Page 165: Ferrate as a New Treatment Chemical for Removal of ...

150

products. Over the two past decades, hundreds of them have been identified in treated

wastewater due to their persistence and the poor treatment capability of conventional

wastewater treatment. The reactivity of these micro-pollutants toward ferrate(VI) allows

ferrate(VI) treatment to be a powerful tool to address the emerging challenge in water

reuse. However, the degradation products of these compounds should be carefully

identified. Their toxicity should be determined and compared with their parent

compounds to examine whether more harmful compounds are produced after ferrate(VI)

oxidation.

Generally, higher chemical dose and lower pH achieve better degradation of

target compounds. A high dose brings more Fe(VI) into water to react with pollutants. A

threshold likely exists beyond which the pollutant removal increases slightly with the

increasing ferrate(VI) dose. In this dissertation research, such a trend was observed for

ferrate(VI) removal of EfOM in terms of COD or UV254 absorbance. At an acidic

condition, ferrate(VI) is much more reactive but more unstable. Therefore, ferrate(VI) is

competed by the reactions with target pollutants and ferrate(VI) self-decomposition.

Coagulation and precipitation also occur after the in-situ formation of Fe(III) from

Fe(VI) reduction. Of interest, the expected TSS removal is not observed in the results

from batch or continuous-flow experiments. Moreover, the nanoscale iron particles

quickly aggregate to form micro-scale particles that remain suspended, significantly

increasing effluent turbidity. Meanwhile, Fe(III) from Fe(VI) reduction can transform

dissolved phosphorus in secondary effluent into a solid phase. However, a majority of

these Fe-P solids still remain suspended. These findings are different from the widely

Page 166: Ferrate as a New Treatment Chemical for Removal of ...

151

accepted viewpoint – ferrate(VI) also functions as a coagulant. Therefore, the Fe(VI)-

induced particles with a poor settling property require an ensuing solid-liquid separation

process (e.g. sand filtration) to remove TSS and P.

Costs of ferrate(VI) treatment are low when compared with the expenses of other

treatment practices. Ferrate(VI) application was extremely restricted by the ferrate(VI)

manufacture. However, the barrier is being removed with the advances in ferrate(VI)

synthesis. Similar to ozone, ferrate(VI) needs to be generated on site due to its chemical

instability. Recently, such on-site ferrate(VI) generators have been commercially

available to produce sodium ferrate solution through a wet chemistry method. Other

methods such as electrochemical methods are being explored to produce a large amount

of ferrate(VI) to meet full-scale applications. Capital cost on the generators and O&M

costs assoicated with electricity are less than those of ozone or energy intensive advanced

oxidation processes. As a result, the affordable costs enable ferrate(VI) technology more

acceptable and competitive than existing treatment options.

7.5 Future Research

Future research needs in ferrate(VI) treatment for water reuse are recommended

as follows.

1) Effect of water sources. Quality of secondary effluent may vary with different

sources. In this study, the effluent from activated sludge treatment, the most

common secondary treatment option in the United States, was used. It is likely

that the treatment performance of ferrate(VI) is changed with the varying

compositions and nature of pollutants and water matrix constituents in secondary

Page 167: Ferrate as a New Treatment Chemical for Removal of ...

152

effluent. Besides the activated sludge process, many other treatments such as

trickling filters and rotating biological contactors are also broadly applied in the

United States. Nature of EfOM and chemical composition (e.g. the nitrite

concertation) from different treatments are likely different. Effect of the different

secondary biological treatments on the following ferrate(VI) treatment should be

evaluated.

2) Effect of temperature. All the experiments in this study were completed at a room

temperature (25oC). If ferrate(VI) is truly applied in practices, the treatment will

proceed under a widely varied temperature range in different seasons. However,

the information on the effect of temperature on ferrate(VI) oxidation or

coagulation is extremely limited. The impact of temperature should be elucidated

prior to full-scale applications.

3) Identification the toxicity of MEF oxidation products. Oxidation products from

Fe(VI) reactions with MEF are identified and the reaction pathways are proposed.

However, it remains unclear whether these OPs are less or more harmful than the

parent compound. The toxicity information is essential to determine whether

ferrate(VI) is selected for addressing the MEF pollution.

4) Lifecycle assessment. Lifecycle assessment is needed for ferrate(VI) treatment of

secondary effluent to quantitatively understand the impacts of ferrate(VI) in water

reclamation. To the best of knowledge, only one study was performed to conduct

lifecycle assessment of ferrate(VI) for drinking water. But such an analysis for

water reuse is unavailable. Information of lifecycle assessment regarding

Page 168: Ferrate as a New Treatment Chemical for Removal of ...

153

ferrate(VI) application for water treatment is of importance for comprehensive

comparison of ferrate(VI) and other treatment candidates for water reclamation.

5) Solid-liquid separation. A solid-liquid separation process need to be selected and

optimized for removing suspended iron particles after Fe(VI) treatment. In the

current treatment market, different separation processes are available, such as

rapid sand filtration, microfiltration, and centrifugation. Investigations are needed

to compare the treatment performance and costs of different separation processes

and select the optimal candidate.

Page 169: Ferrate as a New Treatment Chemical for Removal of ...

154

7.6 References

AWWA (1991) Ozone in water treatment: application and engineering, CRC press.

Cashman, S., Gaglione, A., Mosley, J., Weiss, L., Hawkins, T., Ashbolt, N., Cashdollar,

J., Xue, X., Ma, C. and Arden, S. (2014) Environmental and Cost Life Cycle Assessment

of Disinfection Options for Municipal Wastewater Treatment. Office of Research and

Development. National Homeland Security Research Center.

EPA (1999) Wastewater Technology Fact Sheet Ozone Disinfection. EPA, U. (ed).

Domenjoud, B., Tatari, C., Esplugas, S. and Baig, S. (2011) Ozone-based processes

applied to municipal secondary effluents. Ozone: Science & Engineering 33(3), 243-249.

ICIS (2006) List of chemical prices, pp. http://www.icis.com/chemicals/channel-info-

chemicals-a-z/.

Lee, Y. and von Gunten, U. (2010) Oxidative transformation of micropollutants during

municipal wastewater treatment: Comparison of kinetic aspects of selective (chlorine,

chlorine dioxide, ferrate(VI), and ozone) and non-selective oxidants (hydroxyl radical).

Water Research 44(2), 555-566.

Ning, R.Y. and Troyer, T.L. (2007) Colloidal fouling of RO membranes following

MF/UF in the reclamation of municipal wastewater. Desalination 208(1), 232-237.

Tang, C.Y., Chong, T. and Fane, A.G. (2011) Colloidal interactions and fouling of NF

and RO membranes: a review. Advances in colloid and interface science 164(1), 126-

143.

Tchobanoglous, G., Franklin, L. and Stensel, H.D. (2003) Wastewater engineering:

treatment and reuse. Metcalf & Eddy, Inc., New York.

Tripathi, S., Pathak, V., Tripathi, D.M. and Tripathi, B. (2011) Application of ozone

based treatments of secondary effluents. Bioresource technology 102(3), 2481-2486.

Van Savage, E. (2001) US FERRIC CHLORIDE PRODUCERS RAISE PRICES IN

HEALTHY MARKET. Chemical Market Reporter 260(8), 19.

Vrouwenvelder, J., Beyer, F., Dahmani, K., Hasan, N., Galjaard, G., Kruithof, J. and Van

Loosdrecht, M. (2010) Phosphate limitation to control biofouling. Water Research

44(11), 3454-3466.

Page 170: Ferrate as a New Treatment Chemical for Removal of ...

155

Waite, T.D. (2012) On-Site Production of Ferrate for Water and Wastewater Purification.

American Laboratory 44(10), 26-28.

Wilfert, P., Suresh Kumar, P., Korving, L., Witkamp, G.-J. and Van Loosdrecht, M.C.

(2015) The relevance of phosphorus and iron chemistry to the recovery of phosphorus

from wastewater: a review. Environmental Science & Technology.

Page 171: Ferrate as a New Treatment Chemical for Removal of ...

156