169
The Pennsylvania State University The Graduate School Department of Civil and Environmental Engineering NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING MICROBIAL ELECTROCHEMICAL TECHNOLOGIES A Dissertation in Environmental Engineering by Roland D. Cusick © 2013 Roland D. Cusick Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2013

NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

The Pennsylvania State University

The Graduate School

Department of Civil and Environmental Engineering

NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

MICROBIAL ELECTROCHEMICAL TECHNOLOGIES

A Dissertation in

Environmental Engineering

by

Roland D. Cusick

© 2013 Roland D. Cusick

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

December 2013

Page 2: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

The dissertation of Roland D. Cusick was reviewed and approved* by the following:

Bruce E. Logan Evan Pugh Professor of Engineering, Kappe Professor of Environmental Engineering Dissertation Advisor Chair of Committee

John M. Regan Professor of Environmental Engineering

Brian A. Dempsey Professor of Environmental Engineering

Michael A. Hickner Associate Professor of Material Science and Engineering

Peggy Johnson Professor of Civil Engineering Head of the Department of Civil and Environmental Engineering

*Signatures are on file in the Graduate School

Page 3: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

iii

ABSTRACT

Microbial Electrochemical Technologies (METs) are an emerging design platform

capable of converting organic matter directly into electrical energy. Two new METs were

developed to enhance the sustainability of wastewater treatment. The first set of systems

utilizes electrical and ionic currents to couple bio-electricity generation from wastewater

organics with the recovery of wastewater nutrients via salt crystallization. The second set

of METs was designed to enable the recovery of waste heat using engineered salinity

gradients.

The first system was an energy efficient method of recovering both energy and

nutrients from wastewater in a microbial electrolysis cell (MEC) by producing hydrogen

to increase solution pH and precipitate struvite (MgNH4PO4·6H2O). The concept was first

proven in a single-chamber MEC in which hydrogen evolution cathodes were either

stainless steel 304 mesh or flat plates. The cathodes accumulated a layer of crystals,

which were verified as struvite using a scanning electron microscope capable of energy

dispersive spectroscopy (SEM-EDS). Phosphate removals were low (20 – 38%) because

the pH increase was localized to the cathode surface. Energy produced as hydrogen was

higher than the energy invested to drive hydrogen evolution, implying that MECs may be

useful both as a method for hydrogen gas production and struvite production.

In order to increase phosphate removal and prevent cathode scale accumulation, a

two-chamber MEC was designed where the cathode operated as a fluidized bed

crystallization reactor. Within the cathode, hydrogen production continuously raised

cathode solution pH, and a fluidized bed of seed particles provided surface area for

Page 4: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

iv

struvite crystal growth while scoring the cathode surface. When MEC electrodes

generated current, the cathode solution pH was maintained between 8.3 – 8.8 under

continuous flow conditions, and soluble phosphorus removal ranged from 70 – 85%, as

compared to 10 – 20% when the fluidized bed was operated under open circuit

conditions. At low current densities (≤ 2 mA/m2), scouring by fluidized particles

prevented cathodic scale formation for the duration of continuous flow experiments. At

an applied voltage of 1.0 V, energy consumption from the power supply and pumping

was 0.2 Wh/L (7.5 Wh/g-P). This energy requirement could be fully offset if future

hydrogen recoveries could exceed 80%. These results indicate that a fluidized bed

cathode MEC is a promising method of sustainable electrochemical nutrient recovery.

The second type of MET was developed for converting waste heat and wastewater

organics into electricity using ammonium bicarbonate (AmB) salt solutions in a microbial

reverse electrodialysis cell (MRC). A MRC couples microbial fuel cell (MFC) electrodes

with a reverse electrodialysis (RED) membrane stack, a technology capable of recovering

energy released when salt and fresh water mix. By feeding the RED stack salt solutions

composed of AmB, which decomposes to ammonia and carbon dioxide gas at low

temperatures (40 – 60°C), MRCs can create electricity from abundant sources of low-

grade thermal energy such as waste heat. The maximum power density using acetate

reached 5.6 ± 0.04 W/m2-cathode surface area, which was five times that produced

without the dialysis stack, and 3.0 ± 0.05 W/m2 with domestic wastewater. The greatest

contribution to MRC power came from enhanced MFC electrode performance, which

increased by more than 300% with acetate and over 700% with domestic wastewater.

Page 5: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

v

MFC electrode power production was further investigated in MRCs having a minimal

number of membranes. It was observed that using only a single RED cell pair (CP),

created by operating the cathode in concentrated AmB, dramatically increased power

production normalized to cathode area from both acetate and wastewater. Galvanostatic

electrochemical impedance spectroscopy indicated that power increases in the 1-CP MRC

were caused by reductions in solution and kinetic resistances at the cathode. The addition

of a second RED cell pair further increased power production and anode biofilm activity.

Power densities achieved here were close to those previously achieved using 11

membranes, indicating near optimal electrode performance with only one or two cell

pairs. When normalized to total membrane area, the minimal cell pair MRC power

densities was 4 – 10 times higher than previous MRCs. The rate of wastewater COD

removal, normalized to reactor volume, was 30 – 50 times higher in 1-CP and 2-CP

MRCs than that in a single-chamber MFC. These findings showed that even a single cell

pair AmB RED stack could significantly enhance electrical power production and

wastewater treatment rates.

Page 6: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

vi

TABLE OF CONTENTS

List of Figures .............................................................................................................. ix

List of Tables ............................................................................................................... xiii

Acknowledgements ...................................................................................................... xiv

Chapter 1 General Introduction ................................................................................... 1

1.1 Dissertation Scope and Outline ................................................................................... 2 1.2 Additional Research Publications ............................................................................... 6 1.3 Literature Cited ........................................................................................................... 8

Chapter 2 Literature Review ........................................................................................ 10

2.1 Microbial Electrochemical Technologies ................................................................... 10 2.1.1 MFC and MEC Basic Principles ...................................................................... 10 2.1.2 Anode Materials ............................................................................................... 12 2.1.3 Cathode Materials ............................................................................................. 13 2.1.4 Membranes ....................................................................................................... 13 2.1.5 Internal Resistance ........................................................................................... 14 2.2 Phosphorus Removal and Recovery from Wastewater ............................................... 15 2.2.1 Introduction ...................................................................................................... 15 2.2.2 Phosphorus Management at Wastewater Treatment Plants ............................. 17 2.2.3 Phosphorus Precipitation from Nutrient Rich Side-streams ............................. 18 2.2.4 Electrochemical Phosphate Salt Precipitation .................................................. 23 2.3 Reverse Electrodialysis ............................................................................................... 24 2.3.1 Introduction ...................................................................................................... 24 2.3.2 Reverse Electrodialysis Power Production ....................................................... 27 2.3.3 Membrane Resistance ....................................................................................... 28 2.3.4 Solution Resistance .......................................................................................... 29 2.3.5 Spacers .............................................................................................................. 30 2.3.6 Redox Electrode Pairs ...................................................................................... 30 2.4 Literature Cited ........................................................................................................... 33

Chapter 3 Energy Efficient Phosphate Recovery as Struvite Within a Single Chamber Microbial Electrolysis Cell ................................................................... 44

3.1 Abstract ....................................................................................................................... 44 3.2 Introduction ................................................................................................................. 45 3.3 Materials and Methods ................................................................................................ 47 3.3.1 Reactor Construction and Operation ................................................................ 47 3.3.2 Analytical Techniques ...................................................................................... 49 3.4 Results and Discussion ............................................................................................... 50 3.4.1 Cathode Crystal Accumulation and Phosphate Removal ................................. 50

Page 7: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

vii

3.4.2 Crystal Analysis ............................................................................................... 51 3.4.3 MEC Performance ............................................................................................ 52 3.4.4 Electrochemical Impedance Analysis of Cathodes .......................................... 55 3.5 Conclusions ................................................................................................................. 56 3.6 Figures ........................................................................................................................ 57 3.7 Literature Cited ........................................................................................................... 62

Chapter 4 Electrochemical Struvite Precipitation from Digester Effluent with a Fluidized Bed Cathode Microbial Electrolysis Cell ............................................. 65

4.1 Abstract ....................................................................................................................... 65 4.2 Introduction ................................................................................................................. 66 4.3 Materials and Methods ................................................................................................ 69 4.3.1 Reactor Construction ........................................................................................ 69 4.3.2 Solutions ........................................................................................................... 70 4.3.3 Reactor Operation ............................................................................................. 70 4.3.4 Solution and Electrochemical Measurements .................................................. 71 4.3.5 Analysis ............................................................................................................ 72 4.4 Results and Discussion ............................................................................................... 73 4.4.1 Phosphorus Removal in the Fluidized Bed Cathode ........................................ 73 4.4.2 Phosphorus Precipitation in the Fluidized Bed Cathode .................................. 74 4.4.3 Cathode Scaling ................................................................................................ 75 4.4.4 Molar Ionic Removal in the Fluidized bed Cathode ........................................ 76 4.4.5 Electrolyte Modeling ........................................................................................ 77 4.4.6 COD Removal in the Anode Chamber ............................................................. 78 4.4.7 Energy Consumption ........................................................................................ 78 4.5 Conclusions ................................................................................................................. 79 4.6 Tables .......................................................................................................................... 81 4.7 Figures ........................................................................................................................ 83 4.8 Literature Cited ........................................................................................................... 92

Chapter 5 Energy Capture from Thermolytic Salt Solutions in Microbial Reverse Electrodialysis Cells ............................................................................................. 97

5.1 Abstract ....................................................................................................................... 97 5.2 Introduction ................................................................................................................. 98 5.3 Materials and Methods ................................................................................................ 100 5.3.1 Reactor Construction ........................................................................................ 100 5.3.2 Solutions ........................................................................................................... 101 5.3.3 Power Measurement ......................................................................................... 102 5.3.4 Analysis ............................................................................................................ 103 5.4 Results and Discussion ............................................................................................... 104 5.4.1 Effect of AmB Concentration on MRC Power Production .............................. 104 5.4.2 Effect of Salinity Ratio and Flow Rate on MRC Power Production ................ 106 5.4.3 Energy Production in Batch Recycle Experiments .......................................... 106 5.4.4 Energy Recovery and Efficiency ...................................................................... 107 5.4.5 Ammonia Transport into the Anode ................................................................. 108 5.4.6 Power Production from Domestic Wastewater ................................................ 108 5.5 Conclusions ................................................................................................................. 109

Page 8: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

viii

5.6 Figures ........................................................................................................................ 110 5.7 Literature Cited ........................................................................................................... 118

Chapter 6 Minimal RED Cell Pairs Markedly Improve Electrode Kinetics and Power Production in Microbial Reverse Electrodialysis Cells ............................. 122

6.1 Abstract ....................................................................................................................... 122 6.2 Introduction ................................................................................................................. 123 6.3 Materials and Methods ................................................................................................ 125 6.3.1 Reactor Construction ........................................................................................ 125 6.3.2 Solutions ........................................................................................................... 127 6.3.3 Electrochemical Measurements and Analysis .................................................. 128 6.4 Results and Discussion ............................................................................................... 130 6.4.1 MRC Power Production ................................................................................... 130 6.4.2 Electrode and RED Stack Potential Losses in MRCs ...................................... 131 6.4.3 Anolyte Composition and Membrane Transport Resistance ............................ 133 6.4.4 Effect of RED Stack Potential on Bio-Anode Charge Transfer Resistance ..... 135 6.4.5 Wastewater Treatment Rates and Energy Recovery ........................................ 135 6.5 Outlook ....................................................................................................................... 137 6.6 Tables .......................................................................................................................... 138 6.7 Figures ........................................................................................................................ 139 6.8 Literature Cited ........................................................................................................... 143 Appendix A Energy Efficient Phosphate Recovery as Struvite Within a Single

Chamber Microbial Electrolysis Cell ....................................................................... 146 Appendix B Energy Production from Thermolytic Solutions in Microbial Reverse

Electrodialysis Cells ................................................................................................. 149 Appendix C Minimal RED Cell Pair Microbial Reverse Electrodialysis Cells ............... 152

Page 9: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

ix

LIST OF FIGURES

Figure 2-1: Diagram of microbial electrochemical technologies. .......................................... 12

Figure 2.2: Examples of scale accumulation in pipes and crystallized phosphate salts recovered from a fluidized bed reactor. ........................................................................... 21

Figure 2.3: Diagram of a commercial scale fluidized bed reactor for struvite precipitation. .................................................................................................................... 22

Figure 2.4: Diagram of a reverse electrodialysis reactor. ....................................................... 26

Figure 3.1: Single chamber MESC reactor schematic. ........................................................... 57

Figure 3.2: Struvite Crystal Growth and phosphate removal in MESC with stainless steel mesh (SSM) and shim (SSF) cathodes. ............................................................................ 57

Figure 3.4: (a) Hydrogen production rate (Q) and (b) current density of single chamber MEC (C) and MESC (S) reactors. ................................................................................... 59

Figure 3.5: (a) Coulombic Efficiency (CE), (b) electrical efficiency (ηe) and (c) overall efficiency (ηe+s) of single chamber MEC (C) and MESC (S) reactors. ........................... 60

Figure 3.6: Energy comsumption and recovery for SSM and SSF in MEC (C) and MESC (S) reactors a) 0.75 b) 0.90 and c) 1.05 V. ....................................................................... 61

Figure 4.1: Process flow of fluidized bed cathode microbial electrolysis cell with close up of electrode interface. .................................................................................................. 83

Figure 4.2: a) Molar ionic removal and b) soluble (sP) and total (tP) phosphorus removal in the fluidized bed cathode. ............................................................................................ 84

Figure 4.3: a) MEC current density (b) cathode pH and (c) effluent soluble P concentrations during eight day continuous flow experiments. ....................................... 85

Figure 4.4: a) Phosphorus precipitation and collection rate in the fluidized bed cathode and (b) energy input normalized to precipitated and collected phosphorus. .................... 86

Figure 4.5: SEM images of stainless steel mesh cathodes operated at a) open circuit, b) 0.8 V, c) 1.0 V and d) 1.4 V. ............................................................................................ 87

Figure 4.6: a) Ionic, (b) ammonium and (c) COD removal in the anode chamber. ................ 88

Figure 4.7: Modeled energy balance and a function of cathodic hydrogen recovery. ............ 89

Figure 4.8: Modeled energy input predictions as a function of a) applied voltage and b) influent phosphorus concentration. .................................................................................. 90

Figure 4.9: Electrolyte modeling of a) equivalent NaOH addition and b) super-saturation of sparingly soluble salts, and c) ionization in the digestate. ........................................... 91

Page 10: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

x

Figure 5.1: (a) Main components of the microbial reverse electrodialysis cell (MRC), showing the membrane stack between the electrodes, the reference electrodes and the circuit containing a load (resistor). (b) Example of how the anion- (AEM) and cation- (CEM) exchange membranes are used to selectively drive the flow of positive ions to the right (towards the cathode) and the negatively-charged ions to the left (towards the anode). (c) Expanded view of the membrane stack showing flow path of the high (HC) and low (LC) concentrate solutions of ammonium bicarbonate. (d) Construction of the gaskets used to direct the flow from one LC chamber to the next LC chamber, avoiding the HC chamber through a short flow path through the membrane and gasket. ........................................................................... 110

Figure 5.2: Peak power density of MRC as well as the contributions to total power from the RED stack, and MFC electrodes at various HC concentrations. The dashed line represents peak power density of the same electrodes in a single chamber MFC. .......... 111

Figure 5.3: Polarization curves of the MRC at various HC concentrations as well and the polarization curves of the electrodes in a single chamber MFC. ..................................... 112

Figure 5.4: a) RED stack potential as well as b) anode (filled symbols) and cathode (open symbols) potentials, measured during polarization curves, in the single chamber MFC (SC) and the three highest HC concentrations. ........................................ 113

Figure 5.5: Salinity ratio vs. peak power density at HC concentration of 0.95 M and flow rate of 1.6 mL/min. Dashed line represented peak power density of single chamber MFC fed identical substrate. ............................................................................................ 114

Figure 5.6: a) Energy recovery, efficiency and b) and energy balance of the MRC vs. HC concentration. ................................................................................................................... 115

Figure 5.7: a) Peak power density and (b) anode and cathode potentials of MRC and single chamber MFC fed domestic wastewater. .............................................................. 116

Figure 5.8: Batch recycle component (MFC, RED and total MRC) power profile of MRC fed domestic wastewater. ....................................................................................... 117

Figure 6.1: Schematics of tested reactor configurations. ........................................................ 141

Figure 6.2: (a,b) Power densities normalized to cathode area, (c,d) anode and cathode potentials (vs. SHE), and (e,f) reverse electrodialysis (RED) stack potentials of MFCs, 1-CP and 2-CP MRCs fed (filled markers) bicarbonate buffered acetate and (open markers) domestic wastewater. .............................................................................. 141

Figure 6.3: Internal resistance of 1-CP reactors components fed raw domestic wastewater (WW), buffered wastewater (WW-HC), and buffered acetate (Ac), determined from (a) direct current (DC) polarization and (b) alternating current galvanostatic electrochemical impedance spectroscopy. ....................................................................... 142

Figure 6.4: Internal resistance of MFC, 1-CP, and 2-CP MRC reactors components fed buffered acetate (Ac) determined from (a) direct current (DC) polarization and (b) alternating current (AC) galvanostatic electrochemical impedance spectroscopy. ......... 142

Page 11: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

xi

Figure 6.5: The relative effects of reduced internal resistance and RED stack potential on power enhancement in 1-CP and 2-CP MRCs fed wastewater (WW) and acetate (Ac). ................................................................................................................................. 143

Figure 6.6: a) Volumetric COD removal rates and (b) energy recovery from wastewater and acetate in batch recycle experiments. ........................................................................ 144

Figure 6.7: Estimation of 1-CP and 2-CP MRC power density with buffered acetate (Ac) and wastewater (WW) normalized to (a) total membrane area and (b) cathode area versus reduction in RED stack resistance. ................................................................ 145

Figure A.1: Concurrent LSV (filled) and pH (open) measurements at the surface of SSM cathode in 3mM PBS and 75 mM HCO3

- electrolyte. ...................................................... 150

Figure A.2: Average batch time and crystal accumulation of MESC reactors with both SSF and SSM cathodes at applied voltages of 0.75, 0.90 and 1.05 V. ............................ 151

Figure A.4: EIS determination of internal resistance for SSM and SSF in MEC (C) and MESC (S) reactors at cathode potentials observed in fed-batch operation at a) 0.75 b) 0.90 and c) 1.05 V. ....................................................................................................... 152

Figure B.1: Relationship between ammonium bicarbonate concentration and solution. Figure A.3: Anode (an) and cathode (cat) potentials measured in fed-batch operation at a) 0.75 b) 0.90 and c) 1.05 V. ...................................................................... 153

Figure B.2: Relationship between total ammonium bicarbonate salt concentration and species concentration at pH =7 and T = 25 °C as estimated with OLI stream analyzer software. ........................................................................................................................... 154

Figure B.3: Relationship between species concentration and species activity at pH =7 and T = 25 °C as estimated with OLI stream analyzer software. ..................................... 154

Figure B.4: Power density curves of the MRC (HC = 0.95 M, SR = 100) at different salt solution flow rates. ........................................................................................................... 155

Figure B.5: Batch recycle component (MFC, RED and total MRC) power profile of MRC fed sodium acetate and operating with an external resistance of 300 Ω. ............... 155

Figure C.1: Internal resistance determined from the slope of linear polarization curves for the MFC, 1-CP and 2-CP MRC fed acetate (Ac) and domestic wastewater (WW). .. 156

Figure C.2: a) Power density, (b) electrode potentials, and (c) stack potentials of 1-CP reactor fed acetate (Ac), high conductivity wastewater (WW-HC) and raw wastewater (WW). ........................................................................................................... 157

Figure C.3: Nyquist plots of GEIS data collected from 1-CP MRCs fed acetate, wastewater, and high conductivity wastewater. Impedance spectra were simultaneously collected for (a) whole cell and (b) anode (c) RED stack and (d) cathode. ............................................................................................................................ 157

Page 12: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

xii

Figure C.4: Nyquist plots of GEIS data collected from electrode and RED of MFC and MRCs fed acetate. Impedance spectra were simultaneously collected for (a) anode (c) RED stack and (d) cathode. ........................................................................................ 158

Figure C.5: Nyquist plots with equivalent circuit models and fits of GEIS data collected from the (a) anode (c) RED stack and (d) cathode of 1-CP MRC fed acetate. ................ 158

Page 13: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

xiii

LIST OF TABLES

Table 2.1: Molecular formulas and formation constants for phosphate and carbonate salts. .................................................................................................................................. 19

Table 4.1: Common name, formula and solubility constant of sparing soluble salts with potential to form in digester effluent. ............................................................................... 81

Table 4.2: Solution characteristics of digestate fed to cathode and anode chambers. ............ 81

Table 4.3: Cathode effluent ionic concentrations. .................................................................. 82

Table 4.4: Molar composition of collected solids as determined by acid dissolution. ........... 82

Table 6.1: Anode substrate characteristics .............................................................................. 138

Page 14: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

xiv

ACKNOWLEDGEMENTS

I would like to thank Dr. Bruce E. Logan, for providing me this opportunity to pursue

a graduate degree at Penn State as well as a myriad of life altering experiences over the

last five years. His encouragement and constructive feedback have pushed me to become

a better researcher, communicator and engineer. I would also like to thank Dr. John M.

Regan, Dr. Michael Hickner and Dr. Brian A. Dempsey for their support and serving on

my committee. I would like to acknowledge my fellow researchers at the H2E center. In

particular I want to thank David Jones, Dr. Younggy Kim, for his research guidance,

Mark Ullery for data collection assistance, as well as Marta Hatzell and Fang Zhang for

their assistance in experimental design and data analysis.

I am eternally grateful to my wife, Elise Belknap, for providing me with the

unconditional love, compassion and support needed to complete my graduate education.

Without the guidance and encouragement of my family, I would have never embarked on

this journey.

Finally, I would like to thank the King Abdullah University of Science and

Technology for financially supporting my research.

Page 15: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

Chapter 1

General Introduction

If properly managed, wastewater can sustainably provide society with three of the

most critical resources of the 21st century: energy, nutrients and water. Each day roughly

126 billion liters of wastewater are treated in the United States [1]. Contained within this

out-flow are roughly 240 mega-watt hours (MWh) of energy in the form of organic

matter, 880 MWh of energy as heat, 5 metric tons of nitrogen, and 1 metric ton of

phosphorus [2]. Although wastewater contains nine times the energy required for

treatment [3], current wastewater treatment practices consumes 3% of the national energy

grid [4].

Microbial electrochemical technologies (METs) have attracted much attention

recently due to the unique ability to convert soluble organic matter directly into electrical

energy. Similar to fuel cells and batteries, METs are composed of an anode, where

electrical current is generated, and a cathode, where current is consumed. At the anode,

dissimilatory metal reducing bacteria (commonly referred to as exoelectrogens [5])

oxidize the organic matter present in wastewater and continuously transfer electrons to

the circuit. At MET cathodes, electrical current is consumed to reduce oxygen to water

(MFC) [6], or protons to hydrogen gas (MEC) [7]. In a MFC, electricity generation is

spontaneous because the oxidation of organic matter occurs at a negative potential (− 0.3

V vs. SHE for acetate) and the potential of oxygen reduction to water is positive

Page 16: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

2

(thermodynamic potential of 0.8 V vs. SHE, in practice 0.2 V vs. SHE) [8]. In order to

generate hydrogen at an anaerobic cathode in a MEC, an additional input of electrical

energy is needed (0.4 to 1.0 V in practice); however, the required input is significantly

lower than that required for traditional electrolytic hydrogen generation technologies [9].

Although using METs to directly generate electrochemical energy from wastewater is

an exciting prospect, low power densities and high material costs have limited

commercialization efforts. In comparison, anaerobic digestion (AD), an established

biotechnology, converts organic matter to energy dense methane at roughly 1% of the

capital cost of a MFC [10]. However, the ability of METs to drive ionic current creates

the opportunity to leverage wastewater organics towards displacing energy and cost

intensive industrial processes such as pH adjustment [11] and ion separations [12]. It is

shown here that METs can be designed to not only generate electrical current from

soluble organic matter but also recover solid-phase nutrients from wastewater and

transform low-grade thermal energy into electricity.

1.1 Dissertation Scope and Outline

To increase the potential sustainability of wastewater treatment METs, two new

technologies were developed that simultaneously convert organic wastewater into energy

and recover resources such as phosphate, nitrogen and waste heat.

In Chapter 3, I examine nutrient recovery as magnesium ammonium phosphate

hexahydrate (struvite) in a single chamber microbial electrolysis cell. Struvite can be

produced from crystallization of inorganics in nutrient rich animal and domestic

wastewaters during wastewater treatment. Controlled struvite recovery from wastewater

Page 17: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

3

can be achieved by raising solution pH via chemical base addition, carbon dioxide

stripping, or electrolysis. In order to make electrochemical struvite precipitation more

energy efficient, I examined production of struvite in a microbial electrolysis cell (MEC).

The objectives of this study were therefore to show that struvite could be precipitated at

the cathode of a MEC, and to quantify the affect applied potential and cathode material

had on system metrics such as struvite precipitation, phosphate removal, hydrogen

production and energy efficiency. The results of this chapter are summarized in a paper

by R. D. Cusick and B. E. Logan entitled “Phosphate recovery as struvite within a single

chamber microbial electrolysis cell” published in the journal Bioresource Technology. I

conducted all experiments and analysis and prepared the first manuscript draft.

In Chapter 4, I addressed two of the deficiencies observed for phosphate recovery in

a single chamber MEC (chapter 3) of low P removal rates and cathode scaling by

constructing a continuous flow, two-chambered MEC in which the cathode

simultaneously operates as a hydrogen evolution chamber and high pH fluidized bed

crystallization chamber. Fluidized bed crystallization reactors (FBRs) are designed to

rapidly recover phosphate salts and reduce loss of colloidal fine particles in solution by

providing abundant surface area for molecular growth and agglomeration. The

hypotheses of this study was that a fluidized bed of seed granules could enhance

phosphate salt removal efficiency and maintain MEC current by scouring crystals from

the cathode surface. The FBR-MEC system effectiveness was determined by comparing

phosphate salt removal rates in open circuit voltage operation (no pH adjustment) and

various applied voltages as well as the sustainability of current generation and energy

demand of phosphorus removal. The results of this chapter will be summarized in a paper

Page 18: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

4

by R.D. Cusick, M. Ullery, B. A. Dempsey and B.E. Logan entitled “Electrochemical

struvite precipitation from digestate with a fluidized bed cathode microbial electrolysis

cell”. I designed, constructed and operated the reactor as well as performed data analysis

and manuscript preparation. M. Ullery assisted in data collection. B.A. Dempsey assisted

in electrolyte modeling. All co-authors will assist in the final preparation of the

manuscript.

In Chapter 5, I examined the potential of converting waste heat and wastewater

organics into electricity by using ammonium bicarbonate salt solutions in a microbial

reverse electrodialysis cell (MRC). A MRC couples the spontaneous electrode kinetics of

MFCs with salinity gradient energy of a reverse electrodialysis (RED) stack. Previously,

RED technology had been proposed to harness energy released by the mixing of seawater

and fresh water. With RED saline solutions were composed of ammonium bicarbonate

(AmB), which decomposes to ammonia and carbon dioxide gas at low temperatures (40 –

60°C) [13], MRCs can create electricity from abundant sources of low grade thermal

energy such as waste heat (204 GW) [14], solar (120,000 TW) and geothermal (12 TW)

energy [15]. Therefore, in this study, MRC power production using ammonium

bicarbonate saline solutions was investigated. The effects of saline solution

concentration, salinity ratio (HC/LC), and flow rate on power density were examined

with both synthetic substrate containing acetate and primary clarifier effluent from a local

domestic wastewater treatment plant. The results of this chapter were published in the

journal Science as a paper by R.D Cusick, Y. Kim and B.E. Logan entitled “Energy

capture from thermolytic solutions in microbial reverse-electrodialysis cells.” Y. Kim

assisted in reactor construction and experimental design. I performed all experiments and

Page 19: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

5

analysis as well as prepared the first draft of the manuscript. All co-authors contributed to

the preparation of the final manuscript.

In Chapter 6, I investigated the source of dramatic MRC electrode power production

increases (reported in chapter 5) by constructing and operating MRCs with a minimal

number of RED stacks. The objectives of this study were to investigate potential catalytic

benefits AmB on oxygen reduction, and determine to what extent MRC performance

could be improved using only a minimal number of RED cell pairs. To study the effects

of the AmB and cell pairs on MRC performance, power production and internal

resistance were examined using three different reactor configurations: (1) a single

chamber MFC with a bicarbonate buffer used as the single electrolyte in contact with

both electrodes; (2) a one cell pair (1-CP) MRC where the anolyte formed an equivalent

low concentration chamber with the bicarbonate buffer, and the catholyte was the

concentrated solution (1 M AmB); and (3) a two cell pair (2-CP) MRC containing an

additional membrane pair containing high and low concentration AmB solutions. Both

sodium acetate (in 50 mM bicarbonate buffer) and domestic wastewater were used as

fuels to investigate the effect of anolyte composition on electrode and RED stack

performance. To quantify the specific effects of AmB catholyte and RED stack potential

on electrode activity, internal resistance was measured using linear polarization and

galvanostatic electrochemical impedance spectroscopy. The results presented in the

chapter will be summarized in a paper by R.D. Cusick, M.C. Hatzell, F. Zhang, and B.E.

Logan entitled “Minimal RED cell pairs markedly improve electrode kinetics and power

production in microbial reverse electrodialysis cells.” I conducted all experiments and

Page 20: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

6

analysis and prepared the initial manuscript draft. M.C. Hatzell and F. Zhang assisted in

experimental design. All co-authors will contribute the to preparation of the manuscript.

1.2 Additional Research Publications

The following list includes published research to which I contributed as a co-investigator

but fall outside the scope of this dissertation:

1. Yates, M.D., R.D. Cusick, I. Ivanov, and B.E. Logan, Exoelectrogenic biofilm as a

template for sustainable formation of a catalytic mesoporous structure. Nano Letters,

2013. Submitted.

2. Hatzell, M.C., R.D. Cusick, and B.E. Logan, Capacitive Mixing Power Production

from Salinity Gradient Energy Enhanced through Exoelectrogen-Generated Ionic

Currents. Energy and Environmental Science, 2013. Submitted.

3. Yates, M.D., R. Cusick, and B.E. Logan, Extracellular Palladium Nanoparticle

Production using Geobacter sulfurreducens. ACS Sustainable Chemistry &

Engineering, 2013. doi: 10.1021/sc4000785

4. Zhu, X., Hatzell, M. C., R.D. Cusick, B.E. Logan, Microbial reverse-electrodialysis

chemical-production cell for acid and alkali production. Electrochemistry

Communications, 2013. 31: p. 52-55.

5. Tenca, A., R.D. Cusick, A. Schievano, R. Oberti, and B.E. Logan, valuation of low

cost cathode materials for treatment of industrial and food processing wastewater

using microbial electrolysis cells. International Journal of Hydrogen Energy, 2013.

38(4): p. 1859-1865.

6. Nam, J.-Y., R.D. Cusick, Y. Kim, and B.E. Logan, Nam, J.-Y., et al., Hydrogen

Page 21: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

7

generation in microbial reverse-electrodialysis electrolysis cells using a heat-

regenerated salt solution. Environmental Science and Technology-Columbus, 2012.

46(9): p. 5240.

Page 22: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

8

1.3 Literature Cited

1. WIN, Water infrastructure now: recommendations for clean and safe water in the 21st

century, 2001.

2. McCarty, P.L., J. Bae, and J. Kim, Domestic Wastewater Treatment as a Net Energy

Producer–Can This be Achieved? Environmental Science & Technology, 2011. 45(17):

p. 7100-7106.

3. Shizas, I. and D.M. Bagley, Experimental determination of energy content of unknown

organics in municipal wastewater streams. J Energy Eng, 2004. 130(2): p. 45-53.

4. EPA, Wastewater Management Fact Sheet, Energy Conservation, 2006. p. 7.

5. Logan, B.E., Exoelectrogenic bacteria that power microbial fuel cells. Nature Reviews

Microbiology, 2009. 7(5): p. 375-381.

6. Liu, H., R. Ramnarayanan, and B.E. Logan, Production of electricity during wastewater

treatment using a single chamber microbial fuel cell. Environmental Science &

Technology, 2004. 38(7): p. 2281-2285.

7. Liu, H., S. Grot, and B.E. Logan, Electrochemically assisted microbial production of

hydrogen from acetate. Environ Sci Technol, 2005. 39(11): p. 4317-4320.

8. Logan, B.E., Microbial fuel cells2008, Hoboken, NJ: John Wiley & Sons, Inc. 300.

9. Logan, B.E., et al., Microbial electrolysis cells for high yield hydrogen gas production

from organic matter. Environmental Science & Technology, 2008. 42(23): p. 8630-8640.

10. Rozendal, R.A., et al., Towards practical implementation of bioelectrochemical

wastewater treatment. Trends in Biotechnology, 2008. 26(8): p. 450-459.

11. Rabaey, K., et al., High Current Generation Coupled to Caustic Production Using a

Lamellar Bioelectrochemical System. Environmental Science & Technology, 2010.

44(11): p. 4315-4321.

Page 23: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

9

12. Mehanna, M., et al., Using microbial desalination cells to reduce water salinity prior to

reverse osmosis. Energy & Environmental Science, 2010. 3(8): p. 1114-1120.

13. McGinnis, R.L., J.R. McCutcheon, and M. Elimelech, A novel ammonia-carbon dioxide

osmotic heat engine for power generation. Journal of Membrane Science, 2007. 305: p.

13-19.

14. EIA, U.S., Annual Energy Review 2010, DOE, Editor 2010.

15. Kamat, P.V., Meeting the clean energy demand: Nanostructure architectures for solar

energy conversion. The Journal of Physical Chemistry C, 2007. 111(7): p. 2834-2860.

Page 24: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

10

Chapter 2

Literature Review

2.1 Microbial Electrochemical Technologies

Electrical energy production in microbial electrochemical technologies (METs)

occurs in a similar fashion to other more established electrochemical cells, such as fuel

cell and batteries, by coupling an oxidation reaction at an anode terminal with a reduction

reaction at a cathode terminal. METs are unique in that microbes growing on the anode

catalyze the oxidation of soluble organic matter in wastewater to generate current [1].

Several METs have been designed by tailoring the cathode (both abiotic and biotic) to

accomplish a specific reaction such as oxygen reduction [2], proton reduction to produce

hydrogen [3], metal reduction [4, 5], water reduction to produce hydroxide [6], pollutant

reduction [7] and carbon dioxide reduction to produce organic chemicals [8, 9]. The

METs described in this dissertation utilize either oxygen reduction to produce electricity

or proton reduction to produce hydrogen so this section of the literature review will focus

on substrate and material characteristics that affect power and hydrogen production.

2.1.1 MFC and MEC Basic Principles

Both MFCs and MECs rely on anaerobic microbial oxidation of organic matter at the

anode. Acetate oxidation at pH 7, standard temperature and pressure, is shown as an

example:

CH3CHOO- + H2O → HCO3- + CO2 + 8e- + 8H+ Eo = −0.32 V (2.1)

Page 25: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

11

Although many microbes have proved capable of generating current at the anode a

MET, the most ubiquitous bacterial strain in mixed culture biofilms fed wastewater

appears to be Geobacter sulfurreducens [10-12]. G. sulfurreducens can generate high

current densities (~10 A/m2) by directly transmitting electrons to the anode surface

through conductive pili [13]. Electrons travel from the anode through an external circuit

and are consumed at the cathode. At the cathode of MFCs, oxygen is reduced to water.

2O2 + 8e- + 8H+ → 4H2O Eo = +0.82 V (2.2)

Electricity production in MFCs is spontaneous because the cathode potential is more

positive than the anode potential, which creates a driving force for current production.

The maximum thermodynamic potential for a MFC fed acetate should be 1.1 V, but due

to significant potential losses at the electrodes. MFC open circuit potentials are typically

between 0.5 – 0.7 V [14]. At the anode, open circuit potential losses (~0.1 V) result from

bacterial energy recovery for reproduction and cell maintenance [15]. The most

significant potential losses (~0.6 V) occur at the cathode and are caused by catalyst

properties, solution pH and conductivity [16].

In MECs, electrons are consumed in the hydrogen evolution reaction at the cathode:

8e- + 8H+ → 4H2 Eo = −0.414 V (2.3)

Since the HER potential is less than the maximum anode potential, supplemental voltage

(−0.114 V) must be applied to the cell in order to drive current. Similar to in MFC, HER

cathode over-potentials significantly increase the magnitude of applied potential (0.3 –

1.0 V in practice) required to produce hydrogen in an MEC [17, 18].

Page 26: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

12

2.1.2 Anode Materials

Graphitic carbon based materials are most commonly used as anodes in both MFCs

and MECs. Graphite is advantageous because it is highly electrically conductive, does not

corrode at the operating potential and pH of the electroacitve biofilms, microbes attach

well to the material, and it can be made from waste biomass. High surface area and

porous materials such as graphite fiber brushes have been used to maximize area (4,600

m2/m3 reactor) for biofilm attachment and minimize substrate transport limitations [19].

Anodes can also be thermally or ammonia treated before inoculation to promote

attachment and decrease enrichment times [20, 21].

Figure 2-1: Diagram of microbial electrochemical technologies.

Organic(Carbon(

CO2(+(4H+(

e1(

e1( e1(

e1(

O2(+(4H+(

2H2O(

O2(+(4H+(

H2O(

4H+(

2H2(

e1( e1(

2H2O(+(O2(

2H2O2(

NO2(

½N2(

Page 27: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

13

2.1.3 Cathode Materials

Oxygen reduction in MFCs occurs at the intersection of three phases (liquid, air and

solid) and must simultaneously prevent water leakage while allowing for the transport of

oxygen to the catalytic interface. MFC air breathing cathodes are constructed on a planar

current collector by applying a catalyst layer to the water facing side, and multiple

hydrophobic oxygen diffusion layers to the air facing side [2]. Cathode material cost and

performance are main limiting factors in MFC scale up and performance. To lower

catalyst costs, alternative non-noble catalysts have been explored with recent results

indicating that activated carbon can replace platinum as an oxygen reduction catalyst

without significant losses in power production [22, 23].

The performance of hydrogen evolution cathodes in MECs is highly dependent on the

activation potential loss (or over-potential) of the inert metal catalyst. When platinum, an

ideal HER catalyst with an over-potential near −0.1 V, is used the applied potential

required to produce hydrogen is as low as 0.3 V and the energy produced as hydrogen

can be nearly three times greater than the electricity invested to apply a cell potential

[17]. Inexpensive metal alloys such as stainless steel 304 can be used in MECs but

significantly higher potentials (0.6 – 1.0 V) are required to overcome the HER over-

potential (-0.7 V) [18, 24].

2.1.4 Membranes

Ion exchange membranes are often used to separate the anode and cathode chambers

of METs. In MFCs, anion, cation, and proton exchange membranes as well as filtration

membranes have been used to prevent oxygen intrusion into the anode [25, 26]. In

Page 28: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

14

general, dual-chambered MFCs produce significantly less power than in single chamber

reactors because at neutral pH, and in low conductivity solutions such as wastewater,

membranes substantially increase internal resistance and lead to acidification of the

anode and alkalization of the cathode. Trans-membrane pH gradients and also incur

potential losses, estimated at 60 mV per pH unit difference across the membrane.

Negative membrane potentials can increase (0.4 – 1.0 V) the applied potential required to

produce hydrogen in MEC [27].

2.1.5 Internal Resistance

MET performance is related to factors that resist current production and ionic

transport. In MFCs power production (P) is a function of both cell voltage (V) and

external resistance (Rext) and maximum power is reached when the external resistance is

equal to the internal resistance of the cell (PMFC = IV = V2/Rint) [1]. When current (I) in

generated, internal resistances lead to cell voltage losses which reduces power production

in MFCs and increases the required applied potential in MECs. Potential losses can occur

in solution (Rsol), across a membrane separator (Rmem) if one is present in the system, and

at the electrodes (Ran, Rcat). Internal resistance of METs is composed of ohmic, kinetic,

and diffusional resistances [28]. Ohmic resistance is a measure of how quickly charge is

transported through solution and it is affected by the conductivity of the anode and

cathode electrolytes, and the distance between electrode. Kinetic resistance, also known

as charge transfer resistance (Rct), is a measure of how favorable the anode and cathode

redox reactions are under a given set of system conditions [29]. Diffusion (Rdiff)

Page 29: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

15

resistances are caused by transport imbalances at the electrode reaction interphase [30]

and across the membrane [31].

Electrochemical methods can be used to characterize bulk resistance of the reactor

components (i.e. Ran, Rmem, and Rcat) and specific ohmic and capacitive resistances (i.e.

Rsol, Rct, and Rdiff). Internal resistance can be measured from the slope of a voltage versus

current plot (linear slope method) [32]. The bulk resistances of each of the reactor

components can be determined measuring electrode or trans-membrane potential against

reference electrodes (typically Ag/AgCl) during a linear polarization experiment.

Electrochemical impedance spectroscopy (EIS) is often used to determine the ohmic and

capacitive resistances of a MFC [33]. In EIS experiments, a sinusoidal wave of potential

is applied to the electrochemical cell over a range of frequencies to induce a phase shift in

current [30]. Using EIS is was determined that solution conductivity of wastewater (~1

mS/cm) dramatically limits power production by increasing both solution and diffusion

resistance at the cathode [28]. To optimize power production within the confines of non-

ideal wastewater characteristics, new reactor designs must be explored to minimize the

effect of low wastewater conductivity on power production.

2.2 Phosphorus Removal and Recovery from Wastewater

2.2.1 Introduction

Phosphorus is a vital and irreplaceable nutrient for all living organisms. Phosphate,

the soluble inorganic form of phosphorus, fulfills a diverse array of functions within

cellular organisms including energy storage and transfer (ATP, NADH), information

transfer (DNA, RNA), and structure (Phospholipid bilayer) [34].

Page 30: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

16

Human activity has dramatically increased the mobilization of phosphorus over the

last half century [35]. Industrial and agricultural activities consume roughly 30 million

metric tonnes (MMt) of phosphorus each year, with the majority of P being utilized to

grow crops for food and biofuel production. Current rates of consumption (~18 MMt)

predominantly rely on mined phosphorus [36] and rapid rise in demand and price over the

last three decades have led some researchers to predict phosphorus reserves will be

depleted in the next century [35]. According to the U.S. Geological Survey, at the current

consumption rate, worldwide proven reserves of sedimentary phosphate rock would last

roughly 1,600 years. While the longevity of our reserves is debatable, the effect of

current phosphorus consumption on natural ecosystems is clearly unsustainable. Each

year, 25.3 MMT of P are lost to landfills and waterways [35], directly contributing to the

exponential rise in coastal eutrophication since 1960 [37]. According to a recent

watershed survey conducted by the U.S. EPA, 53% of rivers and streams, 67% of lakes,

reservoirs and ponds, 67% of bays and estuaries, and 86% of coastal shorelines are

considered impaired by pollution [38]. Nutrients are the third largest contributor to

watershed impairment [38].

Livestock and human waste effluents are major contributors to aquatic eutrophication.

Animal feeding operations in the United States generate 1.27 billion tonnes of animal

waste each year, which is 137 times the amount generated by the human population [39].

These waste streams contain 11 MMt of P, and could be used to offset 55% of global P

demand annually if the phosphorus could sustainably recovered during waste treatment

[35]. Unfortunately, despite efforts by the EPA, a recent court ruling vacated the

requirement that concentrated animal feeding operations (CAFOs) that intend to

Page 31: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

17

discharge to watersheds apply for a national pollution discharge elimination system

(NPDES) permit and meet nutrient discharge limits [40]. If adequate technology is

developed to economically and sustainably treat animal wastes and recover nutrients, it is

possible that this decision could be reversed.

2.2.2 Phosphorus Management at Wastewater Treatment Plants

Municipal wastewater treatment plants offer the opportunity to capture significant

amounts of phosphorus (3 MMt yr-1) within an established infrastructure [41].

Phosphorus concentrations entering wastewater treatment plants are dilute (~9 mg-P/L)

and sustainable treatment systems must simultaneously achieve removal from solution

and concentration of phosphorus in a reusable form [42]. The most commonly employed

method of phosphorus removal from primary wastewater streams that fits both criteria is

biological phosphorus removal.

Biological phosphorus removal can occur from bacterial cell growth and enhanced

uptake for energy storage. Since the first observation of biological phosphorus uptake by

aerated sludge in the 1950s, biological nutrient removal has undergone extensive

microbiological investigation and process engineering, and it is now a widely used and

effective biotechnology for phosphorus removal from municipal wastewater [43]. The

enhanced biological phosphorus removal (EBPR) process relies on a group of bacterial

known as phosphate accumulating organisms (PAO) that are capable of absorbing and

storing P in excess of cell growth requirements [43].

The EBPR process requires a strictly anaerobic zone immediately followed by an

aerobic or anoxic zone. In the anaerobic zone, PAOs use stored polyphosphates to

Page 32: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

18

generate adenosine triphosphate (ATP). ATP and absorbed acetate are consumed to

generate polyhydroxyalkanoates (PHAs) an energy storage molecule. The strictly

anaerobic phase is required to ensure the presence of VFAs from fermentation and

provide PAO an opportunity to outcompete other microbes for VFAs. Under

aerobic/anoxic conditions and in the absence of volatile fatty acids, the PAO utilize

stored PHB as a substrate to produce new biomass and restore pools of polyphosphates

and glycogen. Soluble P is absorbed from solution to restore poly-P reserves. If both

VFAs and an electron donor (O2, NO2/NO3) are present, then PAOs will make PHA and

phosphate will be released. The PAO Accumulibacter phosphatis has been shown to

dominate (~80%) the microbial ecology of EBPR sludge [44].

Clarified EBPR sludge contains 5 - 7% of phosphorus by mass and can be reused as

an agricultural fertilizer if risks associated with transmission of pathogens and heavy

metals are eliminated [45]. Thermophilic anaerobic digestion of EBPR sludge is a

common method of stabilization.

2.2.3 Phosphorus Precipitation from Nutrient Rich Side-streams

Aerobic biological wastewater treatment such as activated sludge and EBPR generate

excessive quantities of sludge. Sludge processing and disposal can account for up to 65%

of plant operating costs [46]. To stabilize and thicken sludge prior to disposal or land

application, solids are commonly anaerobically digested and the produced effluent

centrifuged. The remaining liquid often contains high levels of Ca2+, Mg2+, K+, PO43-, and

NH4+. During anaerobic digestion of EBPR sludge, PAOs release phosphate, and 40 -

80% of the phosphorus that enters the digester as sludge is solubilized and transferred to

Page 33: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

19

the supernatant [47, 48]. The release of nutrients during digestion often leads to the

precipitation of calcium and magnesium phosphate salts within the digester as well as on

equipment surfaces. Scale formation on pumps and within conduit can lead to costly

equipment failure and conduit replacement [49-51]. The annual operational cost of

dealing with scale formation at wastewater treatment plants is a function of treatment

capacity and has been estimated to be around $10,000 per million gallons per day (MGD)

of dry weather capacity [52]. A number of mineral salts can form in nutrient rich digester

supernatant (Table 1) with the magnesium ammonium phosphate hexahydrate (struvite)

and amorphous calcium phosphate being the most prevalent.

Table 2.1: Molecular formulas and formation constants for phosphate and carbonate salts.

Common Name Formula pKso Ref.

Struvite MgNH4PO4·6H2O(s) 13.2 [53]

Newberyite MgHPO4·H2O(s) 18.2 [54]

Bobierrite Mg3(PO4)2·22H2O(s) 25.2 [54]

Calcium Hydrogen phosphate CaHPO4(s) 6.7 [55]

β-Tricalcium Phosphate (TCP) β-Ca3(PO4)2(s) 24.0 [55]

Hydroxyapatite (HAP) Ca5(PO4)3OH(s) 55.9 [55] Calcite CaCO3(s) 8.3 [55]

Heterogeneous solid precipitation occurs when the ionic concentration product

components exceeds the thermodynamic solubility limit for a given solid. The primary

growth potential of a salt has been defined as the ratio of the conditional ionic product

(IP) to the solubility limit. Saturation is reached when the ratio is ≥ 1. For struvite (2) and

calcium phosphate (3), the ratios are [53, 56]:

𝛽 = !"!!"

(2.4)

Page 34: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

20

IPStruvite = (αMgγMgCT,Mg) (αMgγMgCT,Mg) (αPO4γPO4CT,PO4) (2.5)

IPTCP = (αCaγCaCT,Ca)3(αPO4γPO4CT,PO4) 2 (2.6)

IPHAP = (αCaγCaCT,Ca)5(αPO4γPO4CT,PO4) 3(αOHγOHCT,OH) (2.7)

where β is the supersaturation ratio, IP is the ionic product, α is the ionization fraction of

each component ion, γ is the activity coefficient, and CT is the total measure

concentration of the ion in solution.

The ionic product of struvite and calcium phosphate are both strongly affected by pH.

At equimolar concentrations, supersaturation of struvite occurs in the shared availability

regions of ammonium (pKa = 9.3) and ortho-phosphate. The availability of ortho-

phosphate increases as pH approaches 12.3 (pKa3 of phosphate) but minimum solubility

is reached at pH = 9.3 – 10.7 [51, 55]. In carbonate rich solutions (>10mM) such as

digestate, complexation of HCO3– and CO3

2– with calcium and magnesium increases the

solubility struvite and calcium phosphate salts [57, 58]. High concentrations of carbonate

have shown to lower the pKso of hydroxylapatite from 56 at pH = 7 to 44 at pH = 11 [57].

Phosphate salt crystal formation occurs in solution from primary nucleation (the

formation of new crystals) and secondary growth onto existing crystal surfaces [59, 60].

Secondary growth on the surface of granules is governed by mass transfer conditions and

surface reaction kinetics and has been modeled as a two-step process [59]:

!"!"= 𝑘!(𝛽 − 𝛽!) (2.8)

!"!"= 𝑘!𝛽!! (2.9)

where dm/dt is the crystal accumulation rate, m is moles of salt, t is reaction time (s), β is

the overall supersaturation and βi is the supersaturation at the granule surface-solution

Page 35: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

21

interface, kd is the overall diffusion coefficient of the salt components, kr is the reaction

constant of the salt, and x is the surface reaction order.

To preventing equipment failure and recovering nutrients has led to development of

digestate and centrate treatment systems. Both physical and chemical methods have been

explored to induce supersaturation of phosphate salts. Precipitation techniques typically

involve raising solution pH to increase the availability of phosphate ions. Sodium

hydroxide (NaOH) is often used because it is highly soluble and can rapid increase the

solution pH of alkaline wastewater between 8.0 and 10.0 [49, 61-63]. Magnesium and

calcium bases (Mg(OH)2 and MgO) are attractive because, in addition to raising pH,

supplemental soluble cation further increases the IP of phosphate salts [54]. Stripping

carbonate alkalinity by aeration has also been shown to effectively raise pH and reduce

cationic complexation with carbonate [56, 64].

Figure 2.2: Examples of scale accumulation in pipes and crystallized phosphate salts recovered from a fluidized bed reactor.

Page 36: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

22

The two most popular reactor designs employed for phosphate salt crystallization are

aeration basins and fluidized beds. Aeration basin crystallization systems are designed to

simultaneously raise pH by stripping alkalinity, promote secondary crystal growth with

upward turbulent mixing, and continuously suspend growing salt granules. Although

simple in design, aeration basins are energy intensive to operate and provide little control

of particle size. Fluidized bed reactors (FBRs) are tall and narrow columns designed to

generate up-flow velocities strong enough to vertically segregate a bed of granules

according to particle diameter. FBRs are often used for crystallized phosphate recovery

because the design creates an abundance of reaction surface area and solution turbulence

[60]. These conditions promote rapid molecular growth at the surface interface of the

granules essentially eliminating the induction time of nucleation [55]. As a result, 70 −

90% removal of soluble phosphate at low hydraulic retention times (1 − 4 h) are typical

for FBRs [60, 61, 65, 66]. Another major advantage of fluidized beds is that by

Figure 2.3: Diagram of a commercial scale fluidized bed reactor for struvite precipitation.

Page 37: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

23

controlling up-flow velocity within the reactor phosphate crystals can grow to a desired

size (3 − 5 mm at full scale), settle and be harvested without shutting down the reactor

[66]. The settling velocity of a particle of known particle density is reached when the

gravitational force acting on the particle matches the drag force exerted by the viscous

liquid fluidizing the particle within the reactor:

𝜗! =!!(!!! !! )

!𝑔𝑅! (2.10)

where υs is the settling velocity (m/s), ρp is the particle density (kg/m3), R is the particle

radius (m), ρf is the fluid density (kg/m3), µ is the fluid dynamic viscosity (N-s/m2), and g

is gravitational acceleration (m/s2) [67].

2.2.4 Electrochemical Phosphate Salt Precipitation

Recently, electrochemical methods of inducing struvite precipitation have been

explored. The fundamental attraction to electrochemical methods tends to revolve around

replacing pH increases by hydroxide addition with consumption of protons via hydrogen

evolution [68, 69]. The counter electrode oxidation reaction was originally water

oxidation, and more recently metal magnesium passivation [70]. Tested electrochemical

systems have been operated with both electrodes in the same reactor chamber. In water

electrolysis, precipitation at cathode surface can occur due to a localized pH increase

caused by hydrogen evolution. In the Mg dissolution cell, a bulk phase pH increase was

created by the release of Mg2+ into solution from oxidation of Mg(0) at the anode, and the

consumption of H+ at the cathode. Both of these systems suffer from critical flaws: water

electrolysis required >3.0 V of applied voltage, making energy input prohibitively

Page 38: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

24

expensive [68]. The electrolytic magnesium passivation system had a low energy input

(0.1 - 2.2 Wh/g-P), but the cost of magnesium electrode led to a struvite production cost

of $6/kg-struvite [70]. Most critically, the effect of cathode scale formation on the

sustainability of current generation, and thus pH increase, has not been explored with any

of the electrochemical struvite precipitation systems.

2.3 Reverse Electrodialysis

2.3.1 Introduction

When volumes of fresh and salt water mix, energy is released, the magnitude of

which is a function of the molar entropy difference between the concentrated and dilute

saline solutions [71]. For ideal solutions, the Gibb’s free energy of mixing (ΔGmix) can be

described as:

ΔGmix = ΔGM – (ΔGHC + ΔGLC) (2.11)

ΔGmix = -(nHC + nLC)TΔSM + (nHCTΔSHC + nHCTΔSLC) (2.12)

ΔS = -RΣxilnxi (2.13)

where subscript HC represents high concentrate solution, LC represents low concentrate

solution, M is the resulting mixed solution, n is moles, ΔS is the molar entropy, R is the

gas rate constant, T is temperature, and xi is the mole fraction of an ionic substance in

solution [72].

Mixing 1 m3 of seawater (~ 0.5 M NaCl) and 1 m3 of freshwater (~0.01 NaCl)

theoretically releases 1.4 MJ of entropic energy [72, 73]. Based on the global discharge

of river water (1.3 × 106 m3/s) into the ocean, roughly 2.8 TW of mixing energy is

released worldwide [74, 75]. Discharge of treated wastewater into the sea releases

Page 39: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

25

another 18.5 GW of energy [76]. Recent research has also explored using ammonium

bicarbonate (NH4HCO3), a thermolytic salt that decomposes at low temperatures (40 – 60

°C), to generate salinity gradient from thermal energy such as industrial waste heat [77,

78]. Each year in the United States over 204 GW of energy is lost as heat in the industrial

sector [79]. One advantage of engineered salinity gradients is that the salinity ratios are

not predetermined by the natural bodies of water and can be increased up to the solubility

of the mixing salt (2 M for AmB). Mixing 1 m3 of concentrated AmB (2.0 M NaCl) with

1 m3 of freshwater (~0.01 AmB) theoretically releases 5.4 MJ (1.5 kWh) of entropic

energy.

Engineers have designed systems to control salinity gradient mixing and convert free

energy into electrical energy. Research and development efforts have focused on two

membrane-based systems: pressure retarded osmosis (PRO) and reverse electrodialysis

(RED) [76, 80]. PRO makes used of the osmotic pressure difference between saline and

fresh solutions to drive water flux across a thin uncharged membrane. As water transports

across the membrane, from dilute to concentrated, volumetric displacement can perform

work on a turbine and generate electricity [81]. RED systems generate electricity by

separating salt and fresh water streams with ion-selective membranes to control ionic

mixing, and convert ionic flux into electrical current. At the time of this work, AmB

solutions have only been used to generate salinity gradients for use in PRO [82] and have

not been used to generate power with RED based systems. Since one of the key

objectives of this dissertation is to recovery heat using AmB solutions within a microbial

RED cell, the material in this literature review will focus on the fundamentals of RED

power generation.

Page 40: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

26

When an ion-selective barrier separates solutions of different ionic concentration, an

electrochemical junction potential (Δϕ) is observed. The potential produced by a HC/LC

junction, or cell pair, is related to the activity difference between the two solutions and

can be estimated with the Nerst equation,

Δϕ = 2(RgT/F) ln(γi,HCxi,HC/γi,LCxi,LC) = 2(RgT/F) ln(ai,HC/ai,LC) (2.14)

where γi represents the mole fraction based activity coefficient and ai is ion activity. The

theoretical potential for a cell pair of 0.5 M (seawater) and 0.01 NaCl (river water) is 220

mV [83]. The theoretical potential for a cell pair of 2.0 M and 0.01 AmB is 240 mV. A

fraction of the potential based on bulk phase concentration difference is lost to membrane

non-idealities and concentration polarization at the membrane interface, leading to

observed junction potentials that ranged from 0.1 – 0.2 V per cell pair [84, 85].

To produce power using RED, multiple cell pairs are constructed in series to drive

current generation at redox electrodes placed at either end of the membrane stack. At the

anode, an oxidation reaction releases electrons to an external circuit and generates a net

Figure 2.4: Diagram of a reverse electrodialysis reactor.

CEM$ AEM$HC#

Stream#LC#

Stream#

A,#

A,#

A,#

A,#

A,#

C+#

C+#

C+#

C+#

A,#

C+#

A,#

A,#

A,#

A,#

A,#

C+#

C+#

C+#

C+#

A,#

C+#

C+#

C+#

C+#A,#

A,#

A,#

2H2O$

O2$+$4H+$ 4H2O$

4OH+$+$2H2$

e+$ e+$

e+$e+$

Page 41: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

27

positive charge increase in the anode chamber. Electrons flow from anode to cathode

through an external circuit. At the cathode a reduction reaction occurs, resulting in a net

decrease in positive charge in the cathode chamber. Solution charge imbalance allows

ionic flux through the selective membranes of the RED stack. The membranes are

arranged to direct anionic flux toward the anode and cationic flux to the cathode to

maintain electro-neutrality within the system. As current flows through the external

circuit, ionic mixing between the HC and LC chambers increases, decreasing the additive

junction potential of the stack [80, 86].

2.3.2 Reverse Electrodialysis Power Production

The power produced by the RED is dependent on the amount of current passing

between the electrodes and the net cell voltage. Maximum power is produced when the

resistance the external circuit (Rext) is equal to the internal resistance of the system (Rint).

Gross power production can be estimated from the open circuit voltage (OCV) and

internal resistance of an electrochemical cell [84].

W = I2Rload = IV = (Vo)2Rext/(Rext + Rint)2 (2.15)

Wmax = (Vo)2Rext/4Rint (2.16)

The OCV of a stack in dependent on several factors including the salinity ratio the

salt solutions, number of salinity gradient cell pairs, membrane selectivity, fluid dynamic

within the stack, and the redox chemistry of the electrode pair. Internal resistance of the

cell is dependent on membrane properties fluid flow geometry, solution characteristics

and the kinetics of the redox catalysis electrodes (Rint = Rstack + Relectrodes) [87].

Page 42: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

28

2.3.3 Membrane Resistance

In a RED system, electrical resistance arises from membrane properties that contribute to

ionic transport resistance. Ion-exchange membranes are relatively water impermeable

barriers that contain fixed charge groups that allow for the passage or either cations or

anions [88]. Anion exchange membranes typically contain fixed quaternary ammonium

groups that interact with anions in solution. Cation exchange membranes employ fixed

sulfonate groups that attract and transport soluble cations across a potential gradient.

Ionic transport in ion-exchange membranes is a function of material properties such as

conductivity, thickness, and the projected area for ionic transport.

The ion conductivity of a membrane, commonly referred to as the ion-exchange

capacity (IEC), determines how quickly ions diffuse through the interface and is

dependent on the density of fixed charged groups attached to the support polymer matrix

[89]. In theory, membrane conductivity and resistance have an inverse relationship: as the

number of fixed charges increases the faster counter-ions can pass through the interface.

However, membrane conductivity must be optimized to prevent membrane swelling that

occurs when soluble counter-ions and water molecules form hydrations shells around the

fixed charge groups, causing the membrane to swell [88]. Membrane swelling increases

the ion-exchange interface thickness, which can in turn increases ionic flux resistance.

Membrane geometry, specifically thickness and area, also greatly contribute to RED

stack resistance. Reported membrane resistance values are typically normalized to area

(i.e. Ω-cm2) indicating that resistance decreases as the area for ionic transport increases

[90].

Page 43: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

29

2.3.4 Solution Resistance

Solution resistance can be calculated from solution conductivity and flow

chamber geometry (Rsol = K d/A). Because of low conductivity of freshwater (≤1

mS/cm), solution resistance in the low concentrate chambers severely limit power

production. The highest reported power production (2.2 W/m2) was reported in a study

that focused on reducing internal resistance by decreasing the solution flow chamber

thickness [91]. A thickness of 100 µm was determined to be optimal because are lower

thickness (60 µm), pressure drop within the stack increased substantially, which increased

the pumping energy input to the system.

Solution conductivity can also have a dramatic effect of the electrical resistance of

membrane interfaces by increasing the transport boundary layer thickness. As current

passes through an ion-exchange membrane, the majority of the ionic current will be due

to counter-ion transport, creating a localized counter-ion gradient at the membrane

interface that acts in opposition to the bulk concentration salinity gradient in a RED. This

transport phenomenon is known as concentration polarization and the thickness of this

ion-specific gradient is known as the transport boundary layer [92]. It was recently shown

that under stagnant conditions, observed membrane resistance increases from 1 – 3 Ω-

cm2 in 0.5 M NaCl to 10 – 20 Ω-cm2 for AEMs and 20 – 50 Ω-cm2 for CEM in fresh

water (0.017 M NaCl) [93]. The boundary layer thickness, and resistance, in dilute

electrolytes can be significantly reduced with hydrodynamic turbulence such as mixing

[94] or increased solution flow rate at the interface [93].

Page 44: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

30

2.3.5 Spacers

Spacers are used to prevent membrane deformation and create mixing within a flow

path. Spacers add to internal resistance by blocking active membrane surface area, a

phenomenon commonly referred to as spacer shadowing effect. In one study, a spacer

was constructed from perforated anion and cation exchange membranes that were

sandwiched together in the RED solution flow paths. Although this is not an economical

solution to spacer shadowing, power production was shown to increase by ~300% [95]

Profiled membranes that create linear flow paths when positioned side by side have also

been used to eliminate the use of spacers [96]. While this approach eliminated spacer

resistance, the linear flow paths lead to an increase in boundary layer resistance and

minimally affected power density.

2.3.6 Redox Electrode Pairs

Since electrodes covert the ionic current of RED stacks to electrical current, proper

consideration must be paid to the redox couples chosen for current generation. An

effective redox couple must be sustainable, and minimize energy loss due to reaction

kinetics (over-potentials). In the last thirty years, the evolution of RED electrodes has

transitioned from reversible reactive electrodes [71, 97], to inert electrodes with reactive

or capacitive electrolytes [85, 96, 98, 99], to most recently, biologically active electrodes

[100, 101].

Reversible redox electrodes utilize identical reactive metal electrodes and an

electrode rinse. At the anode the current is generated by metal corrosion (i.e. Zn → Zn2+

Page 45: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

31

+ 2 e-). The electrode rinse, containing a background concentration of NaCl) carries the

dissolved metal ions from the anode to cathode, where the metal ion is reduced (Zn2+ + 2

e- → Zn). The benefit of this system is that there is no net electrode resistance and current

could be generated with only a few cell pairs. However, reversible electrode pair design

is not sustainable because of metal loss at the anode and accumulation at the cathode

(approximately 2 mm of electrode thickness per week at 50 A/m2) [71].

To avoid electrode material transport, researchers have explored coupling inert

electrodes (e.g. graphite or Pt/Ir) with homogeneous redox active iron based electrolytes.

The soluble redox couples tested include ferric chloride [99] and hexacyanoferrate

([Fe(CN)6]3-/[Fe(CN)6]4-) [91]. In both systems, ferric iron donates one electron to the

anode and receives one electron at the cathode. The benefit of this system is the same as

the reversible electrode couple in that there is no potential loss as the electrodes. The

drawbacks of this system are iron fouling of the ion-exchange membranes, iron

precipitation at the cathode surface and the possibility of cyanide or chlorine gas

formation at electrodes [99, 100].

Water splitting electrodes has been used to avoid membrane fouling and toxic gas

formation. At the anode, water is oxidized to oxygen or hydroxyl radicals. At the cathode,

water is reduced to form hydrogen gas. The theoretical minimum potential for this

reaction is 1.24 V but due to catalyst inefficiency, the net resistance to current generation,

nearly 2.5 V is lost to electrode activation over-potentials [85].

Recently, RED membrane stacks have been paired with MFC electrodes [100]. MFCs

contain anodes where microbes grow and release electrons, and air cathodes, where

electrons are consumed in the oxygen reduction reaction [102]. By coupling a RED stack

Page 46: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

32

with spontaneous MFC electrode kinetics (i.e., oxidation of organic matter at the anode

and oxygen reduction at the cathode), significant energy losses which occur with water

electrolysis (up to 2.5 V in cell voltage loss in a 25 membrane stack due to electrode

over-potentials) are prevented [85]. System damage associated with the hexacyanoferrate

redox couple ([Fe(CN)6]3-/[Fe(CN)6]4-) can also be avoided [72, 86, 103-105]. The

spontaneous nature of MRC current generation also allows salinity-gradient energy

generation with much fewer membranes (5 membrane pairs compared to 25 – 50) [100]

representing a substantial reduction in the membrane associated capitol costs ($100 –

200/m2) [76].

Page 47: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

33

2.4 Literature Cited

1. Logan, B.E., et al., Microbial fuel cells: methodology and technology. Environmental

Science & Technology, 2006. 40(17): p. 5181-5192.

2. Liu, H. and B.E. Logan, Electricity generation using an air-cathode single chamber

microbial fuel cell in the presence and absence of a proton exchange membrane.

Environmental Science & Technology, 2004. 38(14): p. 4040-4046.

3. Liu, H., S. Grot, and B.E. Logan, Electrochemically assisted microbial production of

hydrogen from acetate. Environ Sci Technol, 2005. 39(11): p. 4317-4320.

4. ter Heijne, A., et al., Ferric iron reduction as an alternative for platinum-based cathodes

in microbial fuel cells. Environmental Science & Technology, 2006. 40: p. 5200-5205.

5. Heijne, A.T., et al., Copper Recovery Combined with Electricity Production in a

Microbial Fuel Cell. Environmental Science & Technology, 2010. 44(11): p. 4376-4381.

6. Rabaey, K., et al., High Current Generation Coupled to Caustic Production Using a

Lamellar Bioelectrochemical System. Environmental Science & Technology, 2010.

44(11): p. 4315-4321.

7. Aulenta, F., et al., Electron transfer from a solid-state electrode assisted by methyl

viologen sustains efficient microbial reductive dechlorination of TCE. Environmental

Science & Technology, 2007. 41(7): p. 2554-2559.

8. Nevin, K.P., et al., Electrosynthesis of organic compounds from carbon dioxide is

catalyzed by a diversity of acetogenic microorganisms. Applied and Environmental

Microbiology, 2011. 77(9): p. 2882-2886.

9. Rabaey, K. and R.A. Rozendal, Microbial electrosynthesis- revisiting the electrical route

for microbial production. Nat Rev Micro, 2010. 8(10): p. 706-716.

Page 48: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

34

10. Cusick, R.D., P.D. Kiely, and B.E. Logan, A monetary comparison of energy recovered

from microbial fuel cells and microbial electrolysis cells fed winery or domestic

wastewaters. International Journal of Hydrogen Energy, 2010. 35(17): p. 8855-8861.

11. Kiely, P.D., et al., Anode microbial communities produced by changing from microbial

fuel cell to microbial electrolysis cell operation using two different wastewaters.

Bioresource Technology, 2011. 102(1): p. 388-394.

12. Call, D.F., R.C. Wagner, and B.E. Logan, Hydrogen production by Geobacter species

and a mixed consortium in a microbial electrolysis cell. Applied and Environmental

Microbiology, 2009. 75(24): p. 7579-7587.

13. Malvankar, N.S., et al., Tunable metallic-like conductivity in microbial nanowire

networks. nature nanotechnology, 2011. 6(9): p. 573-579.

14. Cheng, S., H. Liu, and B.E. Logan, Increased power generation in a continuous flow

MFC with advective flow through the porous anode and reduced electrode spacing.

Environmental Science & Technology, 2006. 40: p. 2426-2432.

15. Logan, B.E., Exoelectrogenic bacteria that power microbial fuel cells. Nature Reviews

Microbiology, 2009. 7(5): p. 375-381.

16. Rismani-Yazdi, H., et al., Cathodic limitations in microbial fuel cells: An overview.

Journal of Power Sources, 2008. 180: p. 683–694.

17. Call, D. and B.E. Logan, Hydrogen production in a single chamber microbial electrolysis

cell (MEC) lacking a membrane. Environmental Science & Technology, 2008. 42(9): p.

3401-3406.

18. Call, D., M.D. Merrill, and B.E. Logan, High surface area stainless steel brushes as

cathodes in microbial electrolysis cells (MECs). Environmental Science & Technology,

2009. 43(6): p. 2179-2183.

Page 49: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

35

19. Logan, B.E., et al., Graphite fiber brush anodes for increased power production in air-

cathode microbial fuel cells. Environmental Science & Technology, 2007. 41(9): p.

3341-3346.

20. Cheng, S. and B.E. Logan, Ammonia treatment of carbon cloth anodes to enhance power

generation of microbial fuel cells. Electrochem Commun, 2007. 9(3): p. 492-496.

21. Wang, X., et al., Use of Carbon Mesh Anodes and the Effect of Different Pretreatment

Methods on Power Production in Microbial Fuel Cells. Environmental Science &

Technology, 2009. 43(17): p. 6870-6874.

22. Watson, V.J., C. Nieto Delgado, and B.E. Logan, Influence of chemical and physical

properties of activated carbon powders on oxygen reduction and microbial fuel cell

performance. Environmental Science & Technology, 2013.

23. Zhang, F., D. Pant, and B.E. Logan, Long-term performance of activated carbon air

cathodes with different diffusion layer porosities in microbial fuel cells. Biosensors and

Bioelectronics, 2011. 30(1): p. 49-55.

24. Zhang, Y., M.D. Merrill, and B.E. Logan, The use and optimization of stainless steel

mesh cathodes in microbial electrolysis cells. International Journal of Hydrogen Energy,

2010. 35(21): p. 12020-12028.

25. Kim, J.R., et al., Power generation using different cation, anion and ultrafiltration

membranes in microbial fuel cells. Environmental Science & Technology, 2007. 41(3): p.

1004-1009.

26. Harnisch, F., U. Schröder, and F. Scholz, The suitability of monopolar and bipolar ion

exchange membranes as separators for biological fuel cells. Environmental Science &

Technology, 2008. 42: p. 1740–1746.

Page 50: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

36

27. Rozendal, R.A., et al., Effect of the type of ion exchange membrane on performance, ion

transport, and pH in biocatalyzed electrolysis of wastewater. Water Science and

Technology, 2008. 57(11): p. 1757-1762.

28. Hutchinson, A.J., J.C. Tokash, and B.E. Logan, Analysis of carbon fiber brush loading in

anodes on startup and performance of microbial fuel cells. Journal of Power Sources,

2011. 196(22): p. 9213-9219.

29. Manohar, A.K. and F. Mansfeld, The internal resistance of a microbial fuel cell and its

dependence on cell design and operating conditions. Electrochimica Acta, 2009. 54(6): p.

1664-1670.

30. Bard, A.J. and L.R. Faulkner, Electrochemical Methods: Fundamentals and Applications.

2nd ed2001, New York: John Wiley & Sons.

31. Ji, E., et al., Interface resistances of anion exchange membranes in microbial fuel cells

with low ionic strength. Biosensors and Bioelectronics, 2011. 26(7): p. 3266-3271.

32. Logan, B.E., Microbial fuel cells2008, Hoboken, NJ: John Wiley & Sons, Inc. 300.

33. He, Z. and F. Mansfeld, Exploring the use of electrochemical impedance spectroscopy

(EIS) in microbial fuel cell studies. Energy & Environmental Science, 2009. 2(2): p. 215-

219.

34. Madigan, M.T., et al., Brock Biology of Microorganisms. Eleventh ed2009, Upper Saddle

River, NJ: Pearson Eductation Inc.

35. Cordell, D., J.O. Drangert, and S. White, The story of phosphorus: Global food security

and food for thought. Global Environmental Change, 2009. 19(2): p. 292-305.

36. Villalba, G., et al., Global Phosphorus Flows in the Industrial Economy From a

Production Perspective. Journal of Industrial Ecology, 2008. 12(4): p. 557-569.

37. Diaz, R.J. and R. Rosenberg, Spreading dead zones and consequences for marine

ecosystems. Science, 2008. 321(5891): p. 926-929.

Page 51: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

37

38. EPA, U.S. Watershed Assesment, Tracking & Environmental Results. 2013 [cited 2013

5/17/2013]; Available from: ofmpub.epa.gov/waters10/attains_nation_cy.control.

39. USDA, Manure Management for Water Quality: Costs to Animal Feeding Operations of

Applying Manure Nutrients to Land, E.R. Service, Editor 2003.

40. EPA, U.S., National Pollutant Discharge Elimination System Permit Regulation for

Concentrated Animal Feeding Operations: Removal of Vacated Elements in Response to

2011 Court Decision, 2012: The Federal Register. p. 44494 -44497.

41. Yuan, Z., S. Pratt, and D.J. Batstone, Phosphorus recovery from wastewater through

microbial processes. Current Opinion in Biotechnology, 2012.

42. Cornel, P. and C. Schaum, Phosphorus recovery from wastewater: needs, technologies

and costs. Water science and technology, 2009. 59(6): p. 1069.

43. van Loosdrecht, M.C.M., et al., Biological phosphate removal processes. Applied

Microbiology and Biotechnology, 1997. 48(3): p. 289-296.

44. Martín, H.G., et al., Metagenomic analysis of two enhanced biological phosphorus

removal (EBPR) sludge communities. Nature biotechnology, 2006. 24(10): p. 1263-1269.

45. Tchobanoglous, G., F.L. Burton, and H.D. Stensel, Wastewater Engineering Treatment

and Reuse (4th edn.). Metcalf & Eddy Inc, 2003, McGraw-Hill, New York.

46. Pérez-Elvira, S., P.N. Diez, and F. Fdz-Polanco, Sludge minimisation technologies.

Reviews in Environmental Science and Bio/Technology, 2006. 5(4): p. 375-398.

47. Jardin, N. and H. Popel, Phosphate release of sludges from enhanced biological P-

removal during digestion. Water science and technology, 1994. 30(6): p. 281-292.

48. Liao, P., D.S. Mavinic, and F.A. Koch, Release of phosphorus from biological nutrient

removal sludges: A study of sludge pretreatment methods to optimize phosphorus release

for subsequent recovery purposes. Journal of Environmental Engineering and Science,

2003. 2(5): p. 369-381.

Page 52: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

38

49. Ohlinger, K.N., T.M. Young, and E.D. Schroeder, Kinetics effects on preferential struvite

accumulation in wastewater. Journal of Environmental Engineering, 1999. 125: p. 730.

50. Ohlinger, K.N., T.M. Young, and E.D. Schroeder, Postdigestion struvite precipitation

using a fluidized bed reactor. Journal of Environmental Engineering, 2000. 126: p. 361.

51. Doyle, J.D. and S.A. Parsons, Struvite formation, control and recovery. Water Research,

2002. 36(16): p. 3925-3940.

52. Neethling, J. and M. Benisch, Struvite control through process and facility design as well

as operation strategy. Water science and technology: a journal of the International

Association on Water Pollution Research, 2004. 49(2): p. 191.

53. Ohlinger, K.N., T.M. Young, and E.D. Schroeder, Predicting struvite formation in

digestion. Water Research, 1998. 32(12): p. 3607-3614.

54. Dempsey, B.A. Removal and reuse of ammonia and phosphate by precipitation of

struvite. in PROCEEDINGS OF THE INDUSTRIAL WASTE CONFERENCE-PURDUE

UNIVERSITY. 1998. MOSBY YEAR BOOK EUROPE LTD.

55. Snoeyink, V.L. and D. Jenkins, Water chemistry1980: John Wiley.

56. Battistoni, P., et al., Phosphorus removal from a real anaerobic supernatant by struvite

crystallization. Water Research, 2001. 35(9): p. 2167-2178.

57. Ferguson, J.F. and P.L. McCarty, Effects of carbonate and magnesium on calcium

phosphate precipitation. Environmental Science & Technology, 1971. 5(6): p. 534-540.

58. Cao, X. and W. Harris, Carbonate and Magnesium Interactive Effect on Calcium

Phosphate Precipitation. Environmental Science & Technology, 2007. 42(2): p. 436-442.

59. Bhuiyan, M.I.H., D. Mavinic, and R. Beckie, Nucleation and growth kinetics of struvite

in a fluidized bed reactor. Journal of Crystal Growth, 2008. 310(6): p. 1187-1194.

Page 53: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

39

60. Seckler, M.M., O.S.L. Bruinsma, and G.M. Van Rosmalen, Phosphate removal in a

fluidized bed--I. Identification of physical processes. Water Research, 1996. 30(7): p.

1585-1588.

61. Bhuiyan, M., D. Mavinic, and F. Koch, Phosphorus recovery from wastewater through

struvite formation in fluidized bed reactors: a sustainable approach. Water Science &

Technology, 2008. 57(2): p. 175-181.

62. Le Corre, K.S., et al., Struvite crystallisation and recovery using a stainless steel

structure as a seed material. Water Research, 2007. 41(11): p. 2449-2456.

63. Le Corre, K.S., et al., Impact of calcium on struvite crystal size, shape and purity. Journal

of Crystal Growth, 2005. 283(3-4): p. 514-522.

64. Moerman, W., et al., Phosphate removal in agro-industry: Pilot- and full-scale

operational considerations of struvite crystallization. Water Research, 2009. 43(7): p.

1887-1892.

65. Seckler, M.M., O.S.L. Bruinsma, and G.M. Van Rosmalen, Calcium phosphate

precipitation in a fluidized bed in relation to process conditions: A black box approach.

Water Research, 1996. 30(7): p. 1677-1685.

66. Forrest, A., et al., Optimizing struvite production for phosphate recovery in WWTP.

Journal of Environmental Engineering, 2008. 134: p. 395.

67. Munson, B.R., et al., Fundamentals of fluid mechanics. Vol. 3. 1998: Wiley New York.

68. Moussa, S.B., et al., Electrochemical precipitation of struvite. Electrochemical and Solid-

State Letters, 2006. 9: p. C97.

69. Wang, C.-C., et al., Formation of pure struvite at neutral pH by electrochemical

deposition. Chemical Engineering Journal, 2010. 159(1): p. 280-283.

70. Hug, A. and K.M. Udert, Struvite precipitation from urine with electrochemical

magnesium dosage. Water Research, 2012.

Page 54: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

40

71. Pattle, R., Production of electric power by mixing fresh and salt water in the

hydroelectric pile. 1954.

72. Post, J.W., H.V.M. Hamelers, and C.J.N. Buisman, Energy Recovery from Controlled

Mixing Salt and Fresh Water with a Reverse Electrodialysis System. Environmental

Science & Technology, 2008. 42(15): p. 5785-5790.

73. Post, J., et al., Towards implementation of reverse electrodialysis for power generation

from salinity gradients. Desalination and Water Treatment, 2010. 16(1-3): p. 182-193.

74. Veerman, J., Reverse Electrodialysis: Design and Optimization by Modeling and

Experimentation2010: University of Groningen.

75. Wick, G.L., Power from salinity gradients. Energy, 1978. 3(1): p. 95-100.

76. Ramon, G.Z., B.J. Feinberg, and E.M.V. Hoek, Membrane-based production of salinity-

gradient power. Energy & Environmental Science, 2011.

77. McCutcheon, J.R., R.L. McGinnis, and M. Elimelech, A novel ammonia-carbon dioxide

forward (direct) osmosis desalination process. Desalination, 2005. 174: p. 1-11.

78. McGinnis, R.L. and M. Elimelech, Energy requirements of ammoina-carbon dioxide

forward osmosis desalination. Desalination, 2007. 207(370-382).

79. DOE, U.S., Waste Heat Recovery: Technology and Opportunities in U.S. Industry, I.T.P.

U.S. Department of Energy, Editor 2008.

80. Post, J.W., et al., Salinity-gradient power: Evaluation of pressure-retarded osmosis and

reverse electrodialysis. Journal of Membrane Science, 2007. 288(1-2): p. 218-230.

81. Lee, K., R. Baker, and H. Lonsdale, Membranes for power generation by pressure-

retarded osmosis. Journal of Membrane Science, 1981. 8(2): p. 141-171.

82. McGinnis, R.L., J.R. McCutcheon, and M. Elimelech, A novel ammonia-carbon dioxide

osmotic heat engine for power generation. Journal of Membrane Science, 2007. 305: p.

13-19.

Page 55: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

41

83. Post, J.W., Blue Energy: electricity production from salinity gradients by reverse

electrodialysis2009: Wageningen Universiteit.

84. Długołȩcki, P., et al., Practical Potential of Reverse Electrodialysis As Process for

Sustainable Energy Generation. Environmental Science & Technology, 2009. 43(17): p.

6888-6894.

85. Veerman, J., et al., Reverse electrodialysis: Comparison of six commercial membrane

pairs on the thermodynamic efficiency and power density. Journal of Membrane Science,

2009. 343(1-2): p. 7-15.

86. Lacey, R., Energy by reverse electrodialysis. Ocean engineering, 1980. 7: p. 1-47.

87. Veerman, J., et al., Reverse electrodialysis: A validated process model for design and

optimization. Chemical Engineering Journal, 2011. 166(1): p. 256-268.

88. Strathmann, H., Ion-exchange membrane separation processes. Vol. 9. 2004: Access

Online via Elsevier.

89. Xu, T., Ion exchange membranes: state of their development and perspective. Journal of

Membrane Science, 2005. 263(1): p. 1-29.

90. Dlugolecki, P., et al., Current status of ion exchange membranes for power generation

from salinity gradients. Journal of Membrane Science, 2008. 319(1-2): p. 214-222.

91. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Doubled power density from salinity

gradients at reduced intermembrane distance. Environmental Science & Technology,

2011. 45(16): p. 7089-7095.

92. Choi, J.-H., J.-S. Park, and S.-H. Moon, Direct measurement of concentration

distribution within the boundary layer of an ion-exchange membrane. Journal of Colloid

and Interface Science, 2002. 251(2): p. 311-317.

Page 56: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

42

93. Długołęcki, P., et al., On the resistances of membrane, diffusion boundary layer and

double layer in ion exchange membrane transport. Journal of Membrane Science, 2010.

349(1): p. 369-379.

94. Park, J.-S., et al., An electrical impedance spectroscopic (EIS) study on transport

characteristics of ion-exchange membrane systems. Journal of Colloid and Interface

Science, 2006. 300(2): p. 655-662.

95. Dlugolecki, P., et al., Ion conductive spacers for increased power generation in reverse

electrodialysis. Journal of Membrane Science, 2010. 347(1-2): p. 101-107.

96. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Power generation using profiled

membranes in reverse electrodialysis. Journal of Membrane Science, 2011. 385–386(0):

p. 234-242.

97. Audinos, R., Electrodialyse inverse. Etude de l'energie electrique obtenue a partir de

deux solutions de salinites differentes. Journal of Power Sources, 1983. 10(3): p. 203-

217.

98. Vermaas, D.A., et al., Clean energy generation using capacitive electrodes in reverse

electrodialysis. Energy & Environmental Science, 2013.

99. Veerman, J., et al., Reverse electrodialysis: evaluation of suitable electrode systems.

Journal of Applied Electrochemistry, 2010. 40(8): p. 1461-1474.

100. Kim, Y. and B.E. Logan, Microbial Reverse Electrodialysis Cells for Synergistically

Enhanced Power Production. Environmental Science & Technology, 2011. 45(13): p.

5834-5839.

101. Kim, Y. and B.E. Logan, Hydrogen production from inexhaustible supplies of fresh and

salt water using microbial reverse-electrodialysis electrolysis cells. Proceedings of the

National Academy of Sciences, 2011.

Page 57: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

43

102. Liu, H., R. Ramnarayanan, and B.E. Logan, Production of electricity during wastewater

treatment using a single chamber microbial fuel cell. Environmental Science &

Technology, 2004. 38(7): p. 2281-2285.

103. Veerman, J., et al., Reverse electrodialysis: Performance of a stack with 50 cells on the

mixing of sea and river water. Journal of Membrane Science, 2009. 327(1-2): p. 136-144.

104. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Doubled power density from salinity

gradients at reduced inter-membrane distance. Environmental Science & Technology,

2011. 45: p. 7089-7095.

105. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Power generation using profiled

membranes in reverse electrodialysis. Journal of Membrane Science, 2011.

Page 58: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

Chapter 3

Energy Efficient Phosphate Recovery as Struvite Within a Single

Chamber Microbial Electrolysis Cell

3.1 Abstract

An energy efficient method of concurrent hydrogen gas and struvite (MgNH4PO4·6H2O)

production was investigated based on bioelectrochemically driven struvite crystallization

at the cathode of a single chamber microbial electrolysis struvite-precipitation cell

(MESC). The MESC cathodes were either stainless steel 304 mesh or flat plates.

Phosphate removal ranged from 20 – 40%, with higher removals obtained using mesh

cathodes than with flat plates. Cathode accumulated crystals were verified as struvite

using a scanning electron microscope capable of energy dispersive spectroscopy (SEM-

EDS). Crystal accumulation did not affect the rate of hydrogen production in struvite

reactors. The rate of struvite crystallization (g/m2-hr) and hydrogen production (m3/m3-d)

were shown to be dependent on applied voltage and cathode material. Overall energy

efficiencies (substrate and electricity) were high (73 ± 4%) and not dependant on applied

voltage. These results show that MESCs may be useful both as a method for hydrogen

gas production and struvite production.

Page 59: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

45

3.2 Introduction

Phosphorous is an essential nutrient for all forms of life, but it exists at low

concentrations in the earth’s crust. Modern agricultural practices primarily rely on

synthetic fertilizers containing mined phosphate rock to increase crop yields. Extensive

mining of phosphorus to meet global demands has led to speculation that known global

phosphate deposits may not last through the 21st century [1], with peak in phosphorus

production projected to occur in 2030 [2]. As worldwide reserves diminish, renewable

methods of phosphate production must be developed to help stabilize global food

production. One renewable source of phosphate is magnesium ammonium phosphate

hexahydrate (MgNH4PO4·6H2O), or struvite, which is known to be an excellent slow

release fertilizer [3, 4].

Struvite can be produced from crystallization of inorganics in nutrient rich animal and

domestic wastewaters. Crystal precipitation occurs when concentrations of Mg2+, NH4+,

and PO43– exceed the solubility limit for struvite formation. Concentrations of the three

struvite components increase during anaerobic digestion of biosolids, specifically cell

lysis (magnesium and phosphate) and protein degradation (ammonia) [5]. The

concentration of the specific chemical species needed to form struvite is highly pH

dependent, resulting in a struvite solubility minimum near pH = 10 [5]. However,

effective struvite crystallization has been observed at pH > 8 [6].

Controlled struvite recovery from wastewater can be achieved by several approaches:

chemical addition, carbon dioxide stripping or electrolysis. Most struvite recovery studies

have focused on increasing solution pH via chemical base addition (NaOH, Mg(OH)2,

Ca(OH)2) and stripping carbon dioxide through aeration [5, 7, 8]. While these approaches

Page 60: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

46

are effective, operational costs associated with continuous chemical addition or blower

operation are high ranging from $140-460/tonne of struvite (as compared to $40-50/tonne

of mined phosphate rock) [6]. Chemical base addition can account for up to 97% of

struvite precipitation costs [9]. If operation costs were reduced, sale of recovered struvite

could replace a monetary sink with a revenue stream.

Struvite can also be crystallized at the cathode of a water electrolysis cell where the

consumption of protons (via hydrogen evolution) results in a localized pH increase [10].

The main disadvantage of this process is the high energy cost for producing the potential

needed to split water at the anode (1.25 V in theory, or more than 1.8 V in practice).

Electrochemical precipitation studies have been conducted in the presence of turbulent

mixing (which in itself can cause precipitation), and have not investigated phosphate

removal, hydrogen production or energy efficiency, or addressed whether precipitation is

possible in a solution with a carbonate concentration similar to anaerobic digestion

effluent.

In order to make electrochemical struvite precipitation more energy efficient, we

examined production of struvite in a microbial electrolysis cell (MEC). In an MEC,

microorganisms convert organic and inorganic matter into electrical current at a

significantly lower potential (minimum of ~0.2 V when bacteria used) than that needed

for splitting water [11]. In addition, no precious metals are needed for the MEC,

compared water electrolyzers, which utilize noble metal catalysts on the anode. Recovery

of the hydrogen gas can also offset energy costs, as electrical energy efficiencies in

MECs have reached 400% (a ratio of the energy in the hydrogen gas produced, to the

electrical energy needed) [11]. Thus, the recovery and utilization of the hydrogen gas for

Page 61: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

47

sustaining the process could possibly reduce energy costs at wastewater treatment plants

with combined heat and power systems. It has also been shown that seawater, in

combination with a freshwater source (such as treated wastewater) can be used to provide

the needed potential using a reverse electrodialysis stack [12] and therefore no electrical-

grid power would be needed. Although MECs have been developed for hydrogen and

methane production [13, 14], and microbial fuel cells have been developed to release

orthophosphate from iron phosphate [15], struvite crystallization in an MEC has not yet

been explored. The objectives of this study were therefore to show that struvite could be

precipitated at the cathode of a MEC in the presence of high carbonate concentrations,

and to quantify the affect applied potential and cathode material had on system metrics

such as struvite precipitation, phosphate removal, hydrogen production and efficiency.

3.3 Materials and Methods

3.3.1 Reactor Construction and Operation

Four single chamber MEC reactors (28 mL empty bed volume) were constructed as

previously described [11]. Heat-treated graphite fiber brushes (Gordon Brush, OD = 2.5

cm, L = 2.5) were used as anodes [16, 17]. Stainless steel 304 mesh #60 (SSM, McMaster

Carr, USA) was selected as a cathode based on previous tests showing this mesh size to

be optimum for hydrogen evolution in MECs. Both SSM and a stainless steel flat plates

(SSF, 0.09 mm thick, Trinity Brand Industries, IL, USA) were tested to determine the

relative effects of surface area on hydrogen production rates, struvite crystal precipitation

and phosphate removal. An area of 7 cm2 was used for both the SSM and SSF cathodes in

all calculations based on projected area. However, the SSF cathodes had an exposed

Page 62: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

48

surface area of 19 cm2 due to the mesh fibers, and thus the SSM had a surface area to

projected area ratio of As/Ap = 2.7 [18]. Cathodes were placed 2-cm from the anode at the

opposite side of the chamber (Figure 3.1).

The reactors were inoculated with effluent from working microbial fuel cells. Two

solutions were used: an ammonia phosphate solution (5.96 g/L NaHCO3 1 g/L

NaC2H3O2, 0.46 g/L NH4H2PO4, trace vitamins and minerals) used during startup; and a

magnesium chloride solution (5.96 g/L NaHCO3, 1 g/L NaC2H3O2, 1.66 g/L MgCl2, trace

vitamins and minerals) used in struvite precipitation tests. The concentrations of

carbonate (75 mM) and phosphate (4 mM) were selected to mimic those commonly

found in anaerobic digestion supernatant at domestic wastewater treatment plants [19].

During start-up, all reactors were fed the ammonia phosphate solution and operated in

fed-batch mode at 30°C and an applied potential of 0.9 V until voltages and gas

production were reproducible over three batch cycles. Once stable, two of the four

reactors were converted to operate as microbial electrochemical struvite cells (MESCs)

by feeding a solution containing 4 mM concentrations of the struvite ionic components

(Mg2+, NH4+, and PO4

3-). The struvite reactor feed solution was mixed before each feed

cycle to prevent crystallization during storage, and was composed by combining equal

aliquots of the ammonia phosphate and magnesium chloride solutions. The four reactors

(2 control, 2 MESC) were run at applied voltages (Eap) of 0.75, 0.90, and 1.05 V for at

least three batch cycles with either SSM or SSF cathodes. The four reactors were also

operated at the open circuit voltage (OCV) for three cycles (each 24 hours) to determine

if struvite precipitated at the cathode with no current.

Page 63: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

49

3.3.2 Analytical Techniques

Voltage was added to the reactor circuits with DC power supplies (Circuit Specialists,

Inc., USA). Current was calculated using Ohm’s law by measuring the voltage drop

across a resistor (10 Ω) using a multimeter (Keithley Instruments, OH) [11]. Cathode

potentials were determined using a reference electrode (Ag/AgCl, +0.197 V). Gas

production, gas composition [11], hydrogen production (Q, m3/m3-d), coulombic

efficiency (CE), electrical efficiency (ηe), and overall system efficiency (ηe+s) were

calculated as previously described [20], where the CE is based on the difference in the

initial and final solution chemical oxygen demand (COD). The electrical energy input per

kg of COD removed (Whin, kWh/kg-COD) was determined for each condition:

𝑊ℎ!" =!!"!"

!!!!∆!"#∙!

(3.1)

where Win is the previously described [20] electrical input and ∆COD is the change in

solution chemical oxygen demand and V is the volume of solution fed to the reactor. The

energy recovered as hydrogen per kg of COD (WhH2, kWh/kg-COD) was determined

with equation 3.2:

𝑊ℎ!! = 𝑌!! ∙ 𝐻𝐻𝑉!! (3.2)

where YH2 (kg H2/kg-COD) is the is the hydrogen yield [21], HHVH2 is the higher heating

value of hydrogen (39.4 kWh/kg-H2) [22].

Crystal accumulation (g/m2-hr) on the cathode surface was measured gravimetrically

(before and after each cycle) using a high precision balance (Mettler Toledo AG104,

USA), with the mass normalized by the fed-batch cycle time and the cathode projected

surface area (7 cm2). Used cathodes were dried at room temperature overnight before

Page 64: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

50

being weighed. Crystals were imaged and characterized using scanning electron

microscopy coupled with energy dispersive spectroscopy (Hitachi S-3500N SEM EDS

BSE, Japan). Crystals were verified to be struvite by comparing energy dispersive

spectrographs (EDS) of a pure struvite standard (99.998%, Alfa Aesar) to spectra

collected from cathode accumulated crystals. Phosphate removal due to struvite

precipitation was determined for filtered (0.2 µm, Restek, USA) samples obtained before

and after each batch cycle, using ion chromatography (Dionex-120, AS-16 anion

exchange column, CA).

Electrochemical impedance spectroscopy (EIS, Biologic Science Intruments, VMP 3,

France) was used to investigate the effects of struvite precipitation on cathode kinetics.

An abiotic three electrode cell (working electrode (WE), cathode; reference electrode

(REF), Ag/AgCl; counter electrode, Pt ring) filled with 75 mM carbonate electrolyte was

used in all EIS experiments. To replicate MEC experimental conditions, the WE

potential was set at the average operating potential (Figure S2) for each cathode (SSM

and SSF) at each applied voltage (0.75, 0.90, and 1.05 V).

3.4 Results and Discussion

3.4.1 Cathode Crystal Accumulation and Phosphate Removal

Phosphate removal was achieved by struvite crystallization in MESCs. In

comparison to reactors operated in open circuit (OCV), phosphate removal was

significantly higher in cells where hydrogen evolution and crystal precipitation occurred

Page 65: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

51

(Figure 3.2). Removal rates increased with applied voltages for both SSM and SSF

cathodes. However, the rate of increase in phosphate removal was much higher for

MESCs with mesh than flat plate cathodes. At Eap = 0.75 V phosphate removal in MESC

cells operated with SSM (18 ± 10%) and SSF (20 ± 12%) cathodes were very similar, but

as the applied voltage increased to 0.90 V and 1.05 V, SSM cathode reactors out-

performed SSF cathode reactors. At 1.05 V, phosphate removal reached 38 ± 9% with

SSM cathode reactors at 1.05 V, in comparison to 26 ± 13% in SSF cathode reactors.

Struvite crystals precipitated on both types of cathodes at all applied cell voltages.

Crystal accumulation rates on SSM cathodes increased linearly with applied voltage from

0.32 ± 0.05 g/m2-hr at 0.75 V, to 0.85 ± 0.09 g/m2-hr at 1.05 V (Figure 3.2). The rate of

crystal accumulation on SSF cathodes increased with applied voltage (from 0.25 ± 0.05

g/m2-hr at 0.75 V, to 0.53 ± 0.09 g/m2-hr at 1.05 V) at significantly lower rates than

observed with SSM cathodes. The difference in accumulation rates was most likely due

to the significantly higher surface area of mesh than flat cathodes. Cathodes did not

increase in mass when MESC reactors were operated in OCV, or in control MECs fed

magnesium deficient solution.

3.4.2 Crystal Analysis

Crystals that accumulated on the cathodes were analyzed by SEM-EDS to examine

the morphology of the crystal as well as elemental composition. Crystal growth did not

result in a uniform scale layer but rather produced blooms growing away from the

cathode surface (Figure 3.3). On the SSM cathodes, crystal blooms were concentrated at

wire junctions in the woven mesh (Figure 3.3a) compared to more sporadic growth on the

Page 66: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

52

SSF cathodes (Figure 3b). Crystals on all cathodes displayed the needle shaped prismatic

morphology typical of struvite [23]. Energy dispersive spectra of crystals on SSM

(Figure 3d) and SSF (Figure 3.3e) cathodes (at all applied voltages) verified the

composition as that of struvite based on comparison with pure struvite standards (Figure

3.3f).

3.4.3 MEC Performance

Hydrogen production (Figure 3.4a) in MESCs with SSM cathodes ranged from Q =

1.0 ± 0.3 m3-H2/m3-d (0.75 V) to 2.3 ± 0.2 m3-H2/m3-d (1.05 V). Hydrogen produced by

MESCs with SSF cathodes increased from 0.7 ± 0.02 m3-H2/m3-d (0.75 V) to 2.0 ± 0.6

m3-H2/m3-d (1.05 V). Hydrogen production rates for both MESCs and control reactors

with SSM and SSF cathodes increased linearly with applied voltage. Production values

for control reactors (fed magnesium deficient solution) with SSM and SSF cathodes were

slightly lower but very similar to reactors with struvite production, with differences

ranging from 3% (1.05 V) to 16% (0.75 V). This indicates that struvite crystal

precipitation in MESCs did not negatively affect hydrogen production rates.

The most significant difference in hydrogen production was observed between

reactors with SSM and SSF cathodes. Reactors with SSM cathodes produced hydrogen at

a greater rate than SSF cathode reactors at all applied voltages because reactors with SSM

cathodes operated at higher densities than reactors with SSF cathodes. The difference in

production decreased from 50% at 0.75 V to 11% at 1.05 V implying that the SSF

cathodes could be limited by surface area at lower applied voltages.

Page 67: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

53

Current density (A/m2-cathode, Figure 3.4b) increased linearly with applied voltage

for all reactors. Reactors with mesh cathodes produced slightly more current than those

with sheet reactors. Struvite precipitating reactors produced more current that control

MECs. This difference is most likely attributed to differences is anode performance

between reactors.

Coulombic efficiencies (CEs) were very near or above 100% for all tested conditions

(Figure 5a). The CE in control MECs decreased linearly with applied voltage from 125 ±

14% at 0.75 V to 114 ± 5% at 1.05 V. Generally, struvite reactors had lower CEs than

control reactors and did not show a linear relationship to applied voltage. MESC reactors

with sheet cathodes had an average CE of 111 ± 7%. MESCs with mesh cathodes had an

average CE of 101 ± 8%. Coulombic efficiencies above 100% indicate electron cycling

via hydrogen gas (current generation due to hydrogen oxidation by anodic bacteria)

occurred within both MEC and MESC reactors. Electron cycling results in excess current

generation and additional input of external energy which negatively effects electrical

efficiency [24].

The electrical efficiency of hydrogen production in MESC (the ratio of energy

recovered as hydrogen to the electrical energy consumed by the power supply) exceeded

100% at all applied voltages (Figure 3.5b). The electrical efficiency of control MEC

reactors with SSM and SSF cathodes decreased linearly with each increase in applied

voltage. Observed values of electrical efficiency were similar (within 10%) for MEC and

MESC reactors with both SSM and SSF cathodes at 0.90 V (165% – 181%) and 1.05 V

(139% – 161%). Even with high CE values, the electrical efficiencies presented in this

study suggest that the energy demands of struvite production with an MEC could be

Page 68: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

54

significantly offset by energy recovered as hydrogen. This could substantially lower the

production cost of struvite precipitation in wastewater treatment systems.

The overall system efficiency (ηe+s), which compares the energy recovered as

hydrogen gas to the energy input as electricity and substrate, ranged from 67 – 78% for

MESCs and 78 – 83% for control MECs (Figure 3.5c). MESC reactors with both mesh

and sheet cathodes peaked in overall efficiency at 0.90 V. At this applied voltage interval,

the overall hydrogen production efficiency of struvite precipitating and control reactors

with both mesh and sheet cathodes were within 6% of each other. MESC reactors with

mesh cathodes were only 4 – 6% less efficient than control reactors at all applied

voltages. This similarity indicates that system efficiency in reactors with mesh cathodes

were not significantly affected by struvite crystallization. There was a significant

disparity in efficiency between control MECs and MESCs with sheet cathodes at 0.75 V

(19%) and 1.05 V (15%). The reason for this difference was that MESCs with sheet

cathodes had lower hydrogen yields than MECs with sheet cathodes indicating that

struvite crystallization negatively affected catalysis at the surface of sheet cathodes.

Energy production normalized to COD removal was higher than electrical

consumption at all tested conditions (Figure 6a-c). Electrical energy input rate increased

with applied voltage and ranged from 3.1 ± 0.3 kWh/kg-COD at 0.75 V to 4.6 ± 0.3

kWh/kg-COD at 1.05 V. Electrical energy consumption values were higher for control

reactors that MESC reactors. The observed rates of energy input in this study are higher

than the estimated for activated sludge treatment of organic wastewater (~1 kWh/kg-

COD) [25]. Electricity consumption was amplified by electron cycling within the reactor

as well as the high applied voltages required to overcome the electrochemical over-

Page 69: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

55

potential of the stainless steel cathodes. Energy production in the form of hydrogen

(which is related to acetate concentration) did not follow an obvious trend and ranged

from 4.2 ± 0.3 kWh/kg-COD at 0.75 V to 5.8 ± 0.3 kWh/kg-COD at 1.05 V. Energy

recoveries were lower in MESC reactors than control MECs. The most significant

difference was observed at 0.75 V where MESC reactors with SSM cathodes were 17%

lower and MESCs with SSF cathodes were 29% lower than control MECs. The net

energy recovered as hydrogen for control MEC reactors decreased as applied voltage

(electricity input) increased ranging from 3.1 ± 0.3 kWh/kg-COD at 0.75 V to 4.6 ± 0.3

kWh/kg-COD at 1.05 V. Energy recovery in MESCs reactors with both SSM and SSF

cathodes peaked at 0.9 V. This peak in energy recovery implies that the presence of

crystals on the MESC cathodes effected hydrogen recovery at 0.75 and 1.05 V.

3.4.4 Electrochemical Impedance Analysis of Cathodes

Used mesh and flat plate cathodes from both MEC and MESCs were analyzed with

electrochemical impedance spectroscopy to quantify the effect of crystal accumulation

based on changes in individual factors that contribute to overall internal resistance.

Charge transfer resistances (Rct) were very similar for all cathodes at all potentials, at 1.7

± 0.4 Ω (range of 1.5 – 2.8 Ω) (Figure A3a-c). As expected, there was little change in

solution resistance (Rs= 21 ± 1.4 Ω). As cathode potential decreased, large changes in

diffusion resistance were observed for the SSM cathodes, with decreases from Rd = 41 ±

1.3 Ω at a cathode potential of –0.95 V (0.75 V) to 21 ± 2.8 Ω at –1.2 V (1.05V).

Changes in the diffusion resistance were similar for SSM cathodes with struvite

precipitation (MESC) compared to the control cathodes without struvite crystals (MEC).

Page 70: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

56

In contrast, as cathode potential decreased there was little change in the diffusion

resistance using both the struvite precipitation (MESC) and control (MEC) SSF cathodes.

The similarities of internal resistance between MEC and MESC reactor cathodes provides

evidence that although energy recovery was affected by crystal precipitation, it did not

affect hydrogen evolution kinetics.

3.5 Conclusions

Phosphorus was successfully precipitated as struvite (0.3 – 0.9 g/m2-hr) in single

chamber MESCs with up to 40% of soluble phosphate removed at high electrical energy

efficiencies (135% – 181%). The hydrogen production rates (0.7 – 2.3 m3-H2/m3-d) and

electrical energy efficiencies obtained in this study suggest the energy demands of

struvite production with an MESC would be significantly offset by energy recovered as

hydrogen. This could substantially lower the operational costs for struvite recovery.

Page 71: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

57

3.6 Figures

Figure 3.1: Single chamber MESC reactor schematic.

Figure 3.2: Struvite Crystal Growth and phosphate removal in MESC with stainless steel mesh (SSM) and shim (SSF) cathodes.

Page 72: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

58

c)

d) e) f)

a) b)

2

Page 73: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

59

Figure 3.3: (a) Hydrogen production rate (Q) and (b) current density of single chamber MEC (C) and MESC (S) reactors.

b)

a)

Page 74: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

60

Figure 3.4: (a) Coulombic Efficiency (CE), (b) electrical efficiency (ηe) and (c) overall efficiency (ηe+s) of single chamber MEC (C) and MESC (S) reactors.

a)

b)

c)

Page 75: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

61

Figure 3.5: Energy comsumption and recovery for SSM and SSF in MEC (C) and MESC (S) reactors a) 0.75 b) 0.90 and c) 1.05 V.

a)

b)

c)

Page 76: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

62

3.7 Literature Cited

1. Gilbert, N., The disappearing nutrient. Nature, 2009. 461(8): p. 716-718.

2. Cordell, D., J.O. Drangert, and S. White, The story of phosphorus: Global food security

and food for thought. Global Environmental Change, 2009. 19(2): p. 292-305.

3. Plaza, C., et al., Greenhouse Evaluation of Struvite and Sludges from Municipal

Wastewater Treatment Works as Phosphorus Sources for Plants. Journal of Agricultural

and Food Chemistry, 2007. 55(20): p. 8206-8212.

4. Massey, M.S., et al., Effectiveness of recovered magnesium phosphates as fertilizers in

neutral and slightly alkaline soils. Agronomy Journal, 2009. 101(2): p. 323-329.

5. Ohlinger, K.N., T.M. Young, and E.D. Schroeder, Predicting struvite formation in

digestion. Water Research, 1998. 32(12): p. 3607-3614.

6. Doyle, J.D. and S.A. Parsons, Struvite formation, control and recovery. Water Research,

2002. 36(16): p. 3925-3940.

7. Suzuki, K., et al., Removal of phosphate, magnesium and calcium from swine wastewater

through crystallization enhanced by aeration. Water Research, 2002. 36(12): p. 2991-

2998.

8. Le Corre, K.S., et al., Struvite crystallisation and recovery using a stainless steel

structure as a seed material. Water Research, 2007. 41(11): p. 2449-2456.

9. Jaffer, Y., et al., Potential phosphorus recovery by struvite formation. Water Research,

2002. 36(7): p. 1834-1842.

10. Moussa, S.B., et al., Electrochemical precipitation of struvite. Electrochemical and Solid-

State Letters, 2006. 9: p. C97.

11. Call, D. and B.E. Logan, Hydrogen production in a single chamber microbial electrolysis

cell (MEC) lacking a membrane. Environmental Science & Technology, 2008. 42(9): p.

3401-3406.

Page 77: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

63

12. Kim, Y. and B.E. Logan, Hydrogen production from inexhaustible supplies of fresh and

salt water using microbial reverse-electrodialysis electrolysis cells. Proceedings of the

National Academy of Sciences, 2011.

13. Cheng, S., et al., Direct biological conversion of electrons into methane by

electromethanogenesis. Environmental Science & Technology, 2009. 43(10): p. 3953-

3958.

14. Clauwaert, P. and W. Verstraete, Methanogenesis in membraneless microbial electrolysis

cells. Applied Microbiology and Biotechnology, 2009. 82(5): p. 829-836.

15. Fischer, F., et al., Microbial fuel cell enables phosphate recovery from digested sewage

sludge as struvite. Bioresource Technology, 2011. 102(10): p. 5824-5830.

16. Wang, X., et al., Use of Carbon Mesh Anodes and the Effect of Different Pretreatment

Methods on Power Production in Microbial Fuel Cells. Environmental Science &

Technology, 2009. 43(17): p. 6870-6874.

17. Logan, B.E., et al., Graphite fiber brush anodes for increased power production in air-

cathode microbial fuel cells. Environmental Science & Technology, 2007. 41(9): p.

3341-3346.

18. Zhang, Y., M.D. Merrill, and B.E. Logan, The use and optimization of stainless steel

mesh cathodes in microbial electrolysis cells. International Journal of Hydrogen Energy,

2010. 35(21): p. 12020-12028.

19. Battistoni, P., et al., Phosphorus removal from a real anaerobic supernatant by struvite

crystallization. Water Research, 2001. 35(9): p. 2167-2178.

20. Logan, B.E., et al., Microbial electrolysis cells for high yield hydrogen gas production

from organic matter. Environmental Science & Technology, 2008. 42(23): p. 8630-8640.

Page 78: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

64

21. Cusick, R.D., P.D. Kiely, and B.E. Logan, A monetary comparison of energy recovered

from microbial fuel cells and microbial electrolysis cells fed winery or domestic

wastewaters. International Journal of Hydrogen Energy, 2010. 35(17): p. 8855-8861.

22. NAE, N.a., The Hydrogen Economy: Opportunities, Costs, Barriers, and R&D

Needs2004, Washington, D. C.: The National Academies Press.

23. Le Corre, K.S., et al., Impact of calcium on struvite crystal size, shape and purity. Journal

of Crystal Growth, 2005. 283(3-4): p. 514-522.

24. Call, D.F., R.C. Wagner, and B.E. Logan, Hydrogen production by Geobacter species

and a mixed consortium in a microbial electrolysis cell. Applied and Environmental

Microbiology, 2009. 75(24): p. 7579-7587.

25. Rabaey, K. and W. Verstraete, Microbial fuel cells: novel biotechnology for energy

generation. Trend Biotechnol, 2005. 23(6): p. 291-298.

Page 79: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

Chapter 4

Electrochemical Struvite Precipitation from Digester Effluent with a

Fluidized Bed Cathode Microbial Electrolysis Cell

4.1 Abstract

Microbial electrolysis cells (MECs) can be used to simultaneously convert wastewater

organics to hydrogen and precipitate struvite, but scale formation at the cathode surface

can block catalytic active sites and limit extended operation. To promote bulk phase

struvite precipitation and minimize cathode scaling, a two-chamber MEC was designed

with a fluidized bed to provide particles with a high surface area for crystal growth and to

enable particle cleaning of the cathode surface. Current generation with this design

elevated cathode pH between 8.3 – 8.7 under continuous flow conditions. Soluble

phosphorus removal using digester effluent ranged from 70 – 85%, compared to 10 –

20% for the control (open circuit conditions). At low current densities (≤ 2 mA/m2),

scouring of the cathode by fluidized particles prevented scale accumulation over a period

of 8 days. There was nearly identical removal of soluble phosphorus and magnesium

from solution, and an equimolar composition in the collected solids, supporting

phosphorus removal by struvite formation. At an applied voltage of 1.0 V, energy

consumption from the power supply and pumping (0.2 Wh/L, 7.5 Wh/g-P) was

significantly less than that needed by other struvite formation methods based on pH

adjustment such as aeration and NaOH addition. Total energy consumption could be

Page 80: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

66

completely offset if hydrogen recoveries >80% could be achieved. In the anode chamber,

current generation led to COD oxidation (1.1 – 2.1 g-COD/L-d) and ammonium removal

(7 – 12 mM) from digestate amended with 1 g/L of sodium acetate. These results indicate

that a fluidized bed cathode MEC is a promising method of sustainable electrochemical

nutrient and energy recovery method for nutrient rich wastewaters.

4.2 Introduction

Wastewater treatment plants use enhanced biological phosphorus removal (EBPR) to

minimize nutrient discharge into waterways. Stabilization of excess activated and EBPR

sludge in an anaerobic digester results in significant releases of nutrients, minerals, and

alkalinity (Ca2+, Mg2+, K+, PO43-, NH4

+, and HCO3–) [1]. During digestion, 40 − 80% of

the phosphorus contained in EBPR sludge is solubilized, leading to nutrient recycling

within the plant and calcium and magnesium phosphate salt precipitation within the

digester and on equipment surfaces [2, 3].

In nutrient rich digester effluent, a number of sparingly soluble salts can form (Table

4.1). The most prevalent are magnesium ammonium phosphate (struvite), calcium

carbonate (calcite) and amorphous calcium phosphate (ACP) [4]. Precipitation occurs

when the ionic product of a salt exceeds the solubility limit (aka super saturation) [5].

Since the ionic product of phosphate salts are pH dependent, precipitation takes place

where carbonate striping raises solution pH, such as pumps and conduits due to the high

turbulence [6]. Scale formation on pumps and within conduit can lead to reduced pipe

capacity, conduit replacement, and costly equipment failure [7-9]. The annual operational

cost of dealing with scale formation at wastewater treatment plants has been estimated at

Page 81: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

67

roughly $10,000 per million gallon per day (MGD) of dry weather capacity [10]. The

potential economic benefit of preventing scale formation while recovering renewable

sources of nitrogen and phosphorus has led to the development of side-stream phosphate

salt crystallization systems [6, 8, 11]. The majority of systems have focused on struvite

crystallization because both nitrogen and phosphorus are removed.

Fluidized bed reactors (FBRs) are most commonly used for crystallized phosphate

recovery because the design creates an abundance of reactive surface area and solution

turbulence [12]. These conditions promote rapid molecular growth at the crystal surface

interface which eliminates the induction time associated with nucleating crystal directly

from solution [13]. Struvite precipitation within crystallization reactors affected by pH

and molar concentration ratios. The minimum solubility of struvite has been reported at a

pH of 10.7, but effective secondary crystal growth has been observed at pH values greater

than 8.0 [9]. When solution pH within an FBR is increased to 8.5 or greater, soluble

phosphate removal often reaches 70 − 90% of at low hydraulic retention times (1 − 4

hours) [11, 12, 14, 15]. Struvite precipitation is adversely affected by the presence of

calcium and carbonate. When solution Ca:P ratios are equal to or greater than Mg:P

ratios, struvite crystal purity, size and phosphorus removal can be compromised by

amorphous calcium phosphate co-precipitation [16-18]. In digester effluent, high

alkalinity can lead to magnesium carbonate complexation, which reduces the amount of

magnesium available for struvite precipitation [19, 20].

Despite high nutrient removal efficiencies, proliferation of phosphate salt

crystallization systems has been limited by the monetary cost of increasing solution pH.

Sodium hydroxide (NaOH) is often used because it is highly soluble and can rapid

Page 82: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

68

increase the solution pH [7, 11, 16, 21]. The costs of chemical addition can drive

phosphorus recovery costs (~$3,500/ton-P) to be significantly higher than the value of

nutrients contained in struvite (~$765/ton-P)[9, 22]. Stripping carbonate alkalinity by

aeration has also been shown to effectively raise pH and reduce cationic complexation

with carbonate [17, 23], but blowers are energy intensive and aeration releases ammonia

into the atmosphere.

Recently, microbial electrochemical technologies (METs), such as microbial fuel

cells (MFCs) and electrolysis cells (MECs) have been explored as a chemical free and

energy efficient method of precipitation struvite [24-26]. METs consist of an anode,

where anaerobic microbes oxidize organic matter and transfer electrons to an external

circuit, and a cathode where the electrons and protons catalytically combine by reducing

oxygen to water (MFC) or producing hydrogen gas (MEC). Proton consumption

increases pH at the cathode surface and induces struvite precipitation [27].

Simultaneously removing organic matter and nutrients from wastewater while generating

renewable energy and fertilizer is in concept a sustainable approach for struvite

production, but initial results with batch-fed, single-chamber reactors have shown low

phosphorus removal efficiencies (20 – 50%) and cathode failure due to scale

accumulation.

In this study, a two-chamber MEC reactor with a fluidized bed cathode chamber was

developed to enhance phosphorus removal and minimize cathode scaling. To induce a

bulk phase pH increase in the cathode chamber, a cation exchange membrane separated

the electrodes. The cathode chamber was operated as a fluidized bed to increase struvite

precipitation through production of high surface area particles. In addition, particles

Page 83: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

69

scoured the cathode and helped to minimize scale accumulation. To investigate the effect

of current density on struvite precipitation rate and cathode scaling, the FBR-MEC was

operated under continuous flow at three applied cell voltages (0.8V, 1.0V, and 1.4V). At

each applied voltage, the influent flow rate was adjusted to maintain a cathode pH near

8.5. The reactor was also operated at open circuit (control) to evaluate phosphorus

removal within the fluidized bed in the absence of current.

4.3 Materials and Methods

4.3.1 Reactor Construction

The MEC was composed of two concentric cylindrical chambers. A tubular cation

exchange membrane separated the inner cathode chamber from the outer anode chamber.

The inner cathode chamber was designed to replicate the dimensions of fluidized bed

columns previously used for struvite precipitation. The fluidization section of the cathode

had a diameter to cross sectional area ratio of 25 (inner diameter of d=1.1 cm, height of h

= 28 cm) [12, 28]. The fluidization chamber consisted of a cylindrical cation exchange

membrane (d = 1.1 cm, h = 15 cm) and sections of clear PVC tubing at each end of the

membrane tube (d = 1.1, h = 15 cm). The cation exchange membrane tube was

constructed by applying a CEM monomer solution to the surface of a high-density

polyethylene filtration membrane (Porex, USA) with 30 µm diameter pores. The CEM

layer created an ionic connection between chambers but prevented convective liquid

flow. Above the fluidization section, an expansion vessel (d = 3.8 cm, h = 7.9 cm) was

constructed to reduce liquid velocity and promote solids retention [23]. To collect biogas

produced at the cathode surface, a cut glass anaerobic tube was glued to a section of the

Page 84: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

70

1.1 cm diameter PVC above the expansion section and liquid exit port [29]. The total

cathode volume was 160 mL. The cathode electrode was a 3 cm × 15 cm piece of

stainless steel 304 mesh (30x30, McMaster Carr, US) rolled into a 1 cm diameter cylinder

and placed within the fluidization column. The outer anode chamber was contained a

carbon mesh anode [30], and had an empty bed volume of 180 mL.

4.3.2 Solutions

Anaerobic digester effluent samples were collected from the top section of the

secondary digester at a local domestic wastewater treatment plant (University Park, PA)

at least once per week and stored at 4°C. The treatment plant is comprised of two

secondary treatment trains (activated sludge and trickling filters) followed by biological

nutrient removal. The cathode chamber was fed raw digestate. The organic matter in the

digestate at the point of collection was not adequate for MEC current generation. To

provide sufficient electron donor to the anode microbial community, 1 g/L of sodium

acetate was added to the digestate fed to the anode chamber and the COD increased from

0.6 g/L to 1.3 g/L (Table 2).

4.3.3 Reactor Operation

Solutions were pumped from reservoirs to the MEC reactor through Viton tubing

(Masterflex L/S size 16, USA) using variable speed peristaltic pumps (Cole Parmer,

USA). Solution was recycled in both the anode and cathode chambers at a flow rate of 20

mL/min. Recirculation in the cathode chamber created an up-flow velocity of 0.3 cm/s,

Page 85: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

71

which was sufficient to fluidize a bed of glass particles (30 – 50 µm diameter,

Polyscience, Inc, USA).

The MEC was operated at applied cell voltages (Eap) of 0.8, 1.0 and 1.4 V using an

external power supply (BK Precision, USA). The reactor was also operated under open

circuit (control) to establish a baseline for ion removal in the absence of electrolytic pH

changes. The digestate influent flow rate for each applied voltage was adjusted to

maintain a cathode pH near 8.5. The influent flow rates for both the anode and cathode

chambers were: 0.4 mL/min (0.8 V), 0.7 mL/min (1.0 V), and 1.0 mL/min (1.4 V). At

each applied cell voltage condition, the MEC was continuously operated for 8 days.

4.3.4 Solution and Electrochemical Measurements

To monitor current generation, the voltage drop across a 10 Ω resistor placed within

the MEC circuit was measured every 15 minutes using a multimeter (Keithley

Instruments, USA) connected to a computer. Catholyte pH was measured every 15

minutes by multi-parameter probe placed in a flow cell within the recirculation flow path.

The pH probe was calibrated before each run. Liquid samples (50 mL) were collected

from the influent reservoir as well as anode and cathode effluent ports each day.

Phosphorus, magnesium, calcium, potassium, and sodium concentrations were

determined by inductively coupled plasma atomic emission spectroscopy (ICP-AES)

(Perkin Elmer Optima 5300 DV, USA). To determine total ion concentrations, samples

(10 mL) were acidified to dissolve any fine particles prior to filtration (0.45 µm pore

diameter). Soluble ion concentrations were acidified after filtration to prevent

precipitation during sample storage. Ammonium concentrations were determined with an

Page 86: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

72

ammonia probe (Thermo Scientific Orion, USA). Carbonate alkalinity was determined by

titration to pH 4.5 with 20 mM H2SO4 according to standard methods [31]. Chemical

oxygen demand was determined with a chromic acid colorimetric method (Hach, CO).

4.3.5 Analysis

Phosphorus conversion efficiency (X) accounts for both fine phosphate salt particles

leaving the reactor as well as solids retained and was calculated as the difference between

reactor influent total P and effluent soluble P as:

𝑋 = 100 !"!"!!"!"#!"!"

(4.1)

where tPin is total influent phosphorus (M), and sPout is the effluent soluble phosphorus

concentration (M). Precipitation efficiency (L), which measures the extent of precipitate

retained within the fluidized bed, was calculated as the difference between influent and

effluent total P concentrations as:

𝐿 = 100 !"!"!!"!"#!"!"

(4.2)

The collection efficiency (η), which is used to determine the retention of converted

particles within the fluidized bed, is the ratio of conversion to precipitation (X/L) [17, 23].

The rate of conversion and precipitation were calculated for each sample interval using:

𝑚! = (!"!"!!"!"#)!!"#

!!"#∙∆! (4.3)

𝑚! = (!"!"!!"!"#)!!"#

!!"!∙∆! (4.4)

where mx is the phosphorus mass conversion rate (g-P/m3cat-d), mL is the

phosphorus precipitation rate (g-P/m3cat-d), Vdig is the volume of digestate treated by the

Page 87: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

73

cathode chamber during the sample interval (m3), Vcat is the cathode volume (m3), and ∆t

is the sample interval (d).

Energy input (Ein) to the MEC, from pumping and applied potential was calculated

for each sample interval (Win = WCell + Wpump). Energy input based on current generation

and applied cell potential (Wcell) was calculated as previously described [29]. Energy

input for the recirculation and influent feed lines were calculated as previously described

based on hydraulic head and flow rates [32]. At the end of each run, solids were collected

from the FBR cathode. A portion of the collected solids from each run was dissolved in

an acidic solution (0.2 M HCl) to determine elemental composition. Solid dissolution

supernatant was filtered and analyzed using ICP-AES. Super saturation of sparingly

soluble salts and ionic complexation within the disgestate were predicted at pH values

ranging from 7 – 11 using an electrolyte modeling software package (OLI, USA). The

same software package was also used to estimate the NaOH dose required to raise the

raw digestate to a pH of 8.5.

4.4 Results and Discussion

4.4.1 Phosphorus Removal in the Fluidized Bed Cathode

Enhanced phosphorus removal and elevated pH were maintained in the fluidized bed

cathode MEC. By generating current from soluble organic matter, solution pH within the

cathode chamber was maintained between 8.3 and 8.7 under continuous flow conditions

(Figure 4.3a). Under elevated pH conditions, phosphorus conversion (X: 70 – 85%) and

precipitation efficiency (L: 66 – 71%) within the FBR cathode chamber were

significantly higher than open circuit conditions (Figure 4.2b). Without current

Page 88: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

74

generation cathode solution pH (7.3 ± 0.1) did not increase as it passed through the

fluidized bed and phosphorus conversion (8 ± 6%) and precipitation within the reactor (8

± 6%) were negligible (Figure 4.2b). The level phosphorus removal achieved by the FBR

MEC was comparable to previous crystallization reactor investigations in which solution

pH was increased by aeration or chemical base addition [8, 11, 23] and is a marked

improvement over previous MET struvite investigations.

4.4.2 Phosphorus Precipitation in the Fluidized Bed Cathode

The rate of solids precipitation within the FBR was dependent on influent flow rate

and current density. Current density increased with applied voltage (Figure 4.3a), which

allowed the cathode influent flow rate to be raised while maintaining elevated solution

pH and phosphorus removal within the FBR. To maintain elevated pH at the lowest

applied potential and current density (0.8 V: 0.8 ± 0.03 A/m2-membrane) solution flow

rate was set to 0.4 mL/min (θh = 11.1 hrs), resulting in a precipitation rate of 67 ± 3

grams of phosphorus per cubic meter of cathode volume per day (Figure 4.4a). At 1.0 V,

the current density (1.7 ± 0.2 A/m2) and flow rate (Qf,c = 0.7 mL/min, θh = 4.8 hrs) were

approximately twice that produced at the lower applied voltage, and the phosphorus

precipitation rate increased to 146 ± 15 g-P/m3cat-day. At the applied voltage (1.4 V, 2.4 ±

0.4 A/m2-membrane) the flow rate was 1.0 mL/min (θh = 2.8 h), producing an average

phosphorus precipitation rate of 210 ± 14 g-P/m3cat-day. The highest rate is within the

range previously obtained in fluidized bed crystallization reactors with chemical base

addition to increase pH [8]. Under open circuit conditions, the phosphate precipitation

rate was 16 ± 15 g-P/m3cat-day at an influent flow rate of 0.7 mL/min.

Page 89: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

75

Solids analysis indicated that struvite was the dominant precipitate in the MEC

cathode. Acidic dissolution analysis of solids recovered from the FBR cathode indicated

that phosphorus to magnesium that were equal to, or slightly greater than 1 for all of the

elevated pH conditions (Table 4.4). Molar Mg:P ratios greater than one at applied

voltages of 1.0 and 1.4 V suggest that Newberyite may have co-precipitated struvite. At

OCV, the Mg:P ratio of recovered solids was 0.8 suggesting solids accumulated near

neutral pH were primarily magnesium phosphate salts. Calcium, potassium and sodium

were present in all acid dissolution samples, but at very low concentrations.

Struvite precipitation may have been promoted by favorable influent ionic

concentration ratios. The Mg:P ratio in digestate fed to the reactor varied slightly (1.3 –

1.5) and were above a common target ratio (1.2) which has led to efficient struvite

precipitation in previous investigations [33]. Ratios of Mg:Ca and Ca:P greater than 2

have previously been shown promote small diameter (<10 um) amorphous calcium

phosphate precipitation [17, 34] which easily leave fluidization bed reactors and decrease

precipitation efficiency. Relatively low influent magnesium to calcium (1.1 – 1.2) and

phosphorus to calcium ratios (1.2 – 1.4) also likely contributed to high collection

efficiency.

4.4.3 Cathode Scaling

Cathode performance was maintained by particle scouring at current densities less

than 2 A/m2-membrane (Figure 3). At 0.8 and 1.0 V, current density and cathode pH

were held stable through the entire continuous flow run. As seem in SEM images, the

active surface area SS mesh cathodes operated at 0.8 and 1.0 V was maintained after 8

Page 90: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

76

days of operation (Figure 5 b and c). Minor scale accumulation was localized to the wire

junctions. At 1.4 V, the initial current density of 2.8 ± 0.2 A/m2 began to decrease after 3

days of operation and reduced to 2 A/m2 by the end of the 8-day run. By the end of the

1.4 V run, scale had completely covered the cathode surface. It is unclear whether scale

accumulation was caused by localized pH increase at higher current density or by higher

precipitation rates. It is possible that larger particles and higher up-flow velocity could

maintain the surface area at higher precipitation rates.

4.4.4 Molar Ionic Removal in the Fluidized bed Cathode

Soluble magnesium and phosphorus removal from digestion effluent fed to the

cathode were nearly equimolar, providing further evidence that struvite precipitation was

the dominant P removal pathway within the fluidization bed. The similarity in molar

phosphorus (0.9 – 1.1 mM) and magnesium (0.9 – 1.0 mM) removal for each applied

potentials interval also implied that hydraulic retention time, which decreased from 11

hours at 0.8 V to 2.8 hours at 1.4 V, did not influence the extent of struvite precipitation

within the cathode chamber. When magnesium flux from the anode chamber is

considered, the molar Mg:P removal ratios for all of the elevated pH conditions are all

near unity (Table 4.3). Co-precipitation of calcium phosphates did not occur in the MEC

cathode and soluble calcium concentrations in cathode solution actually increased due to

transport across the CEM during MEC current generation (Figure 4.2a). Soluble

phosphorus concentration in anode chamber solution did not increase, indicating that P

removal in the cathode was caused by precipitation and that the CEM maintained high

selectivity through out the study (Figure 4.6a).

Page 91: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

77

Sodium acetate amendment to anode digestate produced a concentration gradient

between anode and cathode, which led to increasing sodium flux across the CEM with

HRT (Figure 4.2a). Measurement of ammonium flux into the cathode solution was highly

variable due to volatilization caused by elevated pH and turbulence. Ammonium

transport from anode to cathode was measured with greater accuracy from changes in

anode concentration. Ammonia removal from the anode increased with current density

and flow rate, ranging from 7.0 ± 1.3 mM at 0.8 V to 12 ± 2.3 mM at 1.4 V (Figure 4.6b).

4.4.5 Electrolyte Modeling

The influent digestate was modeled using electrolyte software to gain a better

understanding of why magnesium phosphate, rather than calcium phosphates, formed in

the fluidized bed. Modeling showed at pH ≥ 8, both magnesium and calcium phosphates

became supersaturated. At the operating pH of the FBR cathode, several salts were

supersaturated including calcite, hydroxyapatite (HAP), struvite and amorphous calcium

phosphate (ACP). Although hydroxyapatite was the most saturated of the sparingly

soluble phosphate salt (Figure 4.9b), crystalline HAP does not form directly. The

precursor, ACP, transforms into a homogeneous crystalline structure over a period of

days, which is unlikely to occur within the hydraulic retention time of the FBR [19, 35].

Despite high saturation of HAP, the competition for soluble phosphorus results primarily

in the formation of ACP and struvite. At the operating pH of the FBR cathode, model

results shows that struvite was much more saturated than ACP. Additionally, ACP and

calcite precipitation rates can be severely inhibited by CO32-, Mg2+, soluble organic

ligands and organic particles [19, 36-40].

Page 92: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

78

4.4.6 COD Removal in the Anode Chamber

MEC current was generated by microbial oxidation of organic matter at the anode.

COD removal rate in the anode chamber increased with current density ranging from 1.1

± 0.01 g-COD/L-d at 0.8V to 2.1 ± 0.2 g-COD/L-d at 1.4 V (Figure 4.6c). Coulombic

efficiency ranged from 20 – 30%, indicating that acetoclastic methanogenesis likely

contributed to COD removal.

4.4.7 Energy Consumption

Energy consumption during MEC operation resulted almost entirely from the power

used to set an applied potential, and as affected by current density, precipitation rate, and

influent flow rate. At 0.8 V, the energy consumption normalized to the amount of

phosphorus precipitated was 6.5 ± 0.5 Wh/g-P (0.20 ± 0.4 Wh/L). When the applied

potential was increased to 1.0 V, a substantial increase in both current and precipitation

rate offset the increase energy needed at the higher applied voltage, and therefore the

normalized energy input (6.6 ± 0.4 Wh/g-P, 0.22 ± 0.01 Wh/L) was similar to that

obtained at 0.8 V. At 1.4 V, the increase in current density and precipitation rate were

insufficient to compensate for the increased energy, resulting in a much higher energy

consumption of 10 ± 1.0 Wh/g-P (0.3 ± 0.07 Wh/L).

Recovery of hydrogen produced in the MEC could offset electrical energy demand.

Based on the current produced. Hydrogen produced at the MEC cathode could not be

properly collected during continuous flow experiments. It is possible that produced

hydrogen diffused through the PVC reactor body and fitting. If cathodic hydrogen

Page 93: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

79

recovery from the FBR-MEC could reach 80%, energy in the hydrogen gas produced

(based on the lower heating value) would exceed energy consumption (Figure 4.7). Even

without hydrogen gas recovery, energy consumption by the FBR-MEC remain below that

needed for struvite recovery using other pH adjustment methods. Energy consumption

rates observed here were 85 – 90% lower than a system in which aeration was used to

increase digestate pH to 8.4 (~65 Wh/g-P, 1 Wh/L) [4]. It is estimate, based on

electrolyte modeling of the cathode influent digestate, that 5 mM (200 mg/L) of NaOH

would be required to achieve an equivalent pH increase to 8.5 (Figure 4.9a). Considering

that the embodied electrical energy of caustic soda produced by the Chlo-Alkali industry

(114 Wh/mol-NaOH, [41]), equivalent pH and phosphorus removal (1.0 V case) by

caustic addition would represent 0.57 Wh/L and 17.5 Wh/g-P of energy consumption.

The energy required for operation of the FBR-MEC could be further improved in

several ways (Figure 4.8a). For example, constructing the cathode chamber with a

thinner CEM would significantly reduced ionic transport resistance which has limited

MEC hydrogen production in previous investigations [42]. Selecting a catalyst with less

activation over-potential such as nickel foam [43] or molybdenum disulfide [44] would

also reduce applied potential and energy consumption. With the system used in this study

energy consumption normalized to phosphorus precipitation would be significantly

reduced by treating waste with higher influent P concentrations (Figure 4.8b).

4.5 Conclusions

The results of this study indicate that continuous microbial electrochemical struvite

precipitation from digester supernatant is possible if the cathode is operated as a fluidized

Page 94: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

80

bed and the current if the scouring rate is sufficient to prevent cathode scaling. Mg:P

molar ratios of soluble ionic removal and recover solids indicated that struvite formation

was the primary method of phosphorus removal within the FBR cathode. MEC energy

consumption (0.2 - 0.3 Wh/L) was lower than conventional pH adjustment systems used

to recover struvite, such as aeration and chemical base addition. Reducing internal

resistance and producing hydrogen with a low over-potential catalyst could optimize

energy consumption. Capturing MEC produced biogas at a cathodic efficiency of 80%

could offset energy consumption.

Page 95: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

81

4.6 Tables

Table 4.1: Common name, formula and solubility constant of sparing soluble salts with potential to form in digester effluent.

Table 4.2: Solution characteristics of digestate fed to cathode and anode chambers.

Parameter Cathode Anode

pH 7.2 ± 0.1 7.2 ± 0.1

COD (g/L) 0.61 ± 0.1 1.3 ± 0.1

tC (mg CaCO3/L) 2100 ± 120 2300 ± 130

NH4-N (mg/L) 490 ± 45 510 ± 52

tP (mg/L) 46 ± 3 46 ± 3

sP (mg/L) 45 ± 3 45 ± 3

sMg (mg/L) 51 ± 3 52 ± 4

sCa (mg/L) 76 ± 4 76 ± 3

sK (mg/L) 57 ± 4 60 ± 5

sNa (mg/L) 230 ± 15 476 ± 30

Mg:P 1.4 ± 0.1 1.5 ± 0.1

Ca:P 1.3 ± 0.1 1.3 ± 0.1

Mg:Ca 1.1 ± 0.03 1.1 ± 0.1

Common Name Formula pKso Reference

Struvite MgNH4PO4·6H2O(s) 13.2 [6]

Newberyite MgHPO4·H2O(s) 18.2 [45]

Bobierrite Mg3(PO4)2·22H2O(s) 25.2 [45]

Calcium Hydrogen phosphate CaHPO4(s) 6.7 [5]

β-Amorphous Calcium Phosphate (ACP) β-Ca3(PO4)2(s) 24.0 [5]

Hydroxyapatite (HAP) Ca5(PO4)3OH(s) 55.9 [5]

Calcite CaCO3(s) 8.3 [5]

Page 96: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

82

Table 4.3: Cathode effluent ionic concentrations.

FBR Cathode Effluent Ionic Removal Ratios Applied Voltage

tP (mg/L)

sP (mg/L)

sMg (mg/L)

sCa (mg/L) Mg:P Ca:P

OCV 46 ± 3 46 ± 2 49 ± 2 76 ± 2 1.2 0.6

0.8V 14 ± 1 13 ± 1 34 ± 1 87 ± 2 1.0 0.0

1.0V 13 ± 3 7.9 ± 2 23 ± 2 71 ± 3 0.99 0.0

1.4V 15 ± 1 13 ± 3 31 ± 2 76 ± 3 0.98 0.0

Table 4.4: Molar composition of collected solids as determined by acid dissolution.

Solid Sample Mg:P Ca:P K:P Na:P Glass Beads 0 0 0 0

OCV 0.84 0.00 0.08 0.01 0.8 V 1.00 0.01 0.04 0.02 1.0 V 1.06 0.05 0.01 0.01 1.4 V 1.06 0.07 0.02 0.03

Page 97: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

83

4.7 Figures

Figure 4.1: Process flow of fluidized bed cathode microbial electrolysis cell with close up of electrode interface.

Raw digestate

Anod

e FB

R C

atho

de

Recycle Pump

Cathode Effluent Anode

Effluent

Digestate w/ 1 g/L NaAc

Car

bon

Mes

h An

ode

Tubu

lar C

atio

n Ex

chan

ge M

embr

ane

SS M

esh

Cat

hode

Car

bon

Mes

h An

ode

Tubu

lar C

atio

n Ex

chan

ge M

embr

ane

SS M

esh

Cat

hode

Fluidized Particle Bed

CO2

Power Supply

COD Biofilm

e- e-

Page 98: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

84

Figure 4.2: a) Molar ionic removal and b) soluble (sP) and total (tP) phosphorus removal in the fluidized bed cathode.

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

Ion

Rem

oval

in C

atho

de (m

M) P Mg Ca K Na

0

20

40

60

80

100

OCV 0.8V 1.0V 1.4V

Phos

phor

us R

emov

al (%

) sPtP

b)

a)

Page 99: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

85

0.0

1.0

2.0

3.0

4.0

0 1 2 3 4 5 6 7 8

Cur

rent

Den

sity

(A/m

2 )

Days

0.8V 1.0V 1.4V

7.0

7.5

8.0

8.5

9.0

0 1 2 3 4 5 6 7 8

Cat

hode

pH

Days

0.8V 1.0V1.4V OCV

0

10

20

30

40

50

60

0 1 2 3 4 5 6 7 8

Cat

hode

Effl

uent

sP

(mg/

L)

Days

OCV 0.8V 1.0V 1.4V

Figure 4.3: a) MEC current density (b) cathode pH and (c) effluent soluble P concentrations during eight day continuous flow experiments.

a) b)

c)

Page 100: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

86

Figure 4.4: a) Phosphorus precipitation and collection rate in the fluidized bed cathode and (b) energy input normalized to precipitated and collected phosphorus.

0

50

100

150

200

250

OCV 0.8V 1.0V 1.4V

Rem

oval

Rat

e (g

-P/m

3 -rx

tr-d) Precipitation Rate

Collection Rate

0

2

4

6

8

10

12

14

OCV 0.8V 1.0V 1.4V

Ener

gy In

put (

Wh/

g-P)

PrecipitatedCollected

a)

b)

Page 101: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

87

Figure 4.5: SEM images of stainless steel mesh cathodes operated at a) open circuit, b) 0.8 V, c) 1.0 V and d) 1.4 V.

a) b)

c) d)

Page 102: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

88

Figure 4.6: a) Ionic, (b) ammonium and (c) COD removal in the anode chamber.

0.0

0.5

1.0

1.5

2.0

2.5

OCV 0.8V 1.0V 1.4V

CO

D R

emov

al (g

-CO

D/L

-rxtr-

d)

-10

-5

0

5

10

15

20

NH

4+R

emov

al fr

om A

node

(mM

)-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

2.0

Ion

Rem

oval

in A

node

(mM

)P Mg Ca K Na

a)

b)

c)

Page 103: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

89

0.00

0.05

0.10

0.15

0.20

0.25

0.30

0 20 40 60 80 100

Ener

gy B

alan

ce (W

h/L)

Cathodic Recovery (%)

H2Input

Figure 4.7: Modeled energy balance and a function of cathodic hydrogen recovery.

Page 104: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

90

0123456789

10

0 50 100 150 200

Influent Conc. (mg-P/L)

0

2

4

6

8

10

0.0 0.2 0.4 0.6 0.8 1.0

Ene

rgy

Inpu

t (kW

h/kg

-P)

Applied Voltage (V)

Figure 4.8: Modeled energy input predictions as a function of a) applied voltage and b) influent phosphorus concentration.

a) b)

Page 105: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

91

Figure 4.9: Electrolyte modeling of a) equivalent NaOH addition and b) super-saturation of sparingly soluble salts, and c) ionization in the digestate.

0

10

20

30

40

50

6 7 8 9 10 11E

quiv

alen

t NaO

H A

dditi

on (m

M)

pH

-3

-2

-1

0

1

2

3

6 7 8 9 10 11

log

(IP/K

so)

pH

CaCO3

MgNH4PO4

β-­‐Ca3(PO4)2(s)

0.0

0.2

0.4

0.6

0.8

1.0

7 8 9 10 11

Ioni

zatio

n Fr

actio

n

pH

Ca2+

CaH2PO4+

Mg2+

MgCO3(o)

NH4+

PO43-­‐

CaCO3(o)

HPO42-­‐

H2PO4-­‐

a)

b)

c)

Page 106: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

92

4.8 Literature Cited

1. Pérez-Elvira, S., P.N. Diez, and F. Fdz-Polanco, Sludge minimisation technologies.

Reviews in Environmental Science and Bio/Technology, 2006. 5(4): p. 375-398.

2. Jardin, N. and H. Popel, Phosphate release of sludges from enhanced biological P-

removal during digestion. Water science and technology, 1994. 30(6): p. 281-292.

3. Liao, P., D.S. Mavinic, and F.A. Koch, Release of phosphorus from biological nutrient

removal sludges: A study of sludge pretreatment methods to optimize phosphorus release

for subsequent recovery purposes. Journal of Environmental Engineering and Science,

2003. 2(5): p. 369-381.

4. Battistoni, P., et al., P removal from anaerobic supernatants by struvite crystallization:

long term validation and process modelling. Water Research, 2002. 36(8): p. 1927-1938.

5. Snoeyink, V.L. and D. Jenkins, Water chemistry1980: John Wiley.

6. Ohlinger, K.N., T.M. Young, and E.D. Schroeder, Predicting struvite formation in

digestion. Water Research, 1998. 32(12): p. 3607-3614.

7. Ohlinger, K.N., T.M. Young, and E.D. Schroeder, Kinetics effects on preferential struvite

accumulation in wastewater. Journal of Environmental Engineering, 1999. 125: p. 730.

8. Ohlinger, K.N., T.M. Young, and E.D. Schroeder, Postdigestion struvite precipitation

using a fluidized bed reactor. Journal of Environmental Engineering, 2000. 126: p. 361.

9. Doyle, J.D. and S.A. Parsons, Struvite formation, control and recovery. Water Research,

2002. 36(16): p. 3925-3940.

10. Neethling, J. and M. Benisch, Struvite control through process and facility design as well

as operation strategy. Water science and technology: a journal of the International

Association on Water Pollution Research, 2004. 49(2): p. 191.

Page 107: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

93

11. Bhuiyan, M., D. Mavinic, and F. Koch, Phosphorus recovery from wastewater through

struvite formation in fluidized bed reactors: a sustainable approach. Water Science &

Technology, 2008. 57(2): p. 175-181.

12. Seckler, M.M., O.S.L. Bruinsma, and G.M. Van Rosmalen, Phosphate removal in a

fluidized bed--I. Identification of physical processes. Water Research, 1996. 30(7): p.

1585-1588.

13. Bhuiyan, M.I.H., D. Mavinic, and R. Beckie, Nucleation and growth kinetics of struvite

in a fluidized bed reactor. Journal of Crystal Growth, 2008. 310(6): p. 1187-1194.

14. Seckler, M.M., O.S.L. Bruinsma, and G.M. Van Rosmalen, Calcium phosphate

precipitation in a fluidized bed in relation to process conditions: A black box approach.

Water Research, 1996. 30(7): p. 1677-1685.

15. Forrest, A., et al., Optimizing struvite production for phosphate recovery in WWTP.

Journal of Environmental Engineering, 2008. 134: p. 395.

16. Le Corre, K.S., et al., Impact of calcium on struvite crystal size, shape and purity. Journal

of Crystal Growth, 2005. 283(3-4): p. 514-522.

17. Moerman, W., et al., Phosphate removal in agro-industry: Pilot- and full-scale

operational considerations of struvite crystallization. Water Research, 2009. 43(7): p.

1887-1892.

18. Hao, X.D., et al., Struvite formation, analytical methods and effects of pH and Ca(2 + ).

Water Sci Technol, 2009. 59(6): p. 1077-1083.

19. Cao, X. and W. Harris, Carbonate and Magnesium Interactive Effect on Calcium

Phosphate Precipitation. Environmental Science & Technology, 2007. 42(2): p. 436-442.

20. Mañas, A., et al., Parameters influencing calcium phosphate precipitation in granular

sludge sequencing batch reactor. Chemical Engineering Science, 2012. 77(0): p. 165-

175.

Page 108: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

94

21. Le Corre, K.S., et al., Struvite crystallisation and recovery using a stainless steel

structure as a seed material. Water Research, 2007. 41(11): p. 2449-2456.

22. Dockhorn, T. About the economy of phosphorus recovery. in International conference on

nutrient recovery from wastewater streams. 2009.

23. Battistoni, P., et al., Phosphorus removal from a real anaerobic supernatant by struvite

crystallization. Water Research, 2001. 35(9): p. 2167-2178.

24. Cusick, R.D. and B.E. Logan, Phosphate recovery as struvite within a single chamber

microbial electrolysis cell. Bioresource Technology, 2012. 107(0): p. 110-115.

25. Hirooka, K. and O. Ichihashi, Phosphorus recovery from artificial wastewater by

microbial fuel cell and its effect on power generation. Bioresource Technology, 2013.

137(0): p. 368-375.

26. Ichihashi, O. and K. Hirooka, Removal and recovery of phosphorus as struvite from

swine wastewater using microbial fuel cell. Bioresource Technology, 2012. 114(0): p.

303-307.

27. Moussa, S.B., et al., Electrochemical precipitation of struvite. Electrochemical and Solid-

State Letters, 2006. 9: p. C97.

28. Seckler, M., et al., Phosphate removal in a fluidized bed--II. Process optimization. Water

Research, 1996. 30(7): p. 1589-1596.

29. Call, D. and B.E. Logan, Hydrogen production in a single chamber microbial electrolysis

cell (MEC) lacking a membrane. Environmental Science & Technology, 2008. 42(9): p.

3401-3406.

30. Wang, X., et al., Use of Carbon Mesh Anodes and the Effect of Different Pretreatment

Methods on Power Production in Microbial Fuel Cells. Environmental Science &

Technology, 2009. 43(17): p. 6870-6874.

Page 109: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

95

31. APHA, ed. Standard Methods for the Examination of Water and Wastewater. 20th ed.,

ed. L.S. Clesceri, A.E. Greenberg, and A.D. Eaton1998, American Public Health

Association, American Water Works Association, Water Environment Federation:

Washington DC.

32. Zhang, F., et al., Long-Term Performance of Liter-Scale Microbial Fuel Cells Treating

Primary Effluent Installed in a Municipal Wastewater Treatment Facility. Environmental

Science & Technology, 2013. 47(9): p. 4941-4948.

33. Nelson, N.O., R.L. Mikkelsen, and D.L. Hesterberg, Struvite precipitation in anaerobic

swine lagoon liquid: effect of pH and Mg: P ratio and determination of rate constant.

Bioresource Technology, 2003. 89(3): p. 229-236.

34. Le Corre, K.S., et al., Impact of calcium on struvite crystal size, shape and purity. Journal

of Crystal Growth, 2005. 283(3): p. 514-522.

35. Abbona, F. and A. Baronnet, A XRD and TEM study on the transformation of amorphous

calcium phosphate in the presence of magnesium. Journal of Crystal Growth, 1996.

165(1–2): p. 98-105.

36. Ferguson, J.F., D. Jenkins, and J. Eastman, Calcium Phosphate Precipitation at Slightly

Alkaline pH Values. Journal (Water Pollution Control Federation), 1973. 45(4): p. 620-

631.

37. Ferguson, J.F. and P.L. McCarty, Effects of carbonate and magnesium on calcium

phosphate precipitation. Environmental Science & Technology, 1971. 5(6): p. 534-540.

38. Van der Houwen, J.A., et al., The effect of organic ligands on the crystallinity of calcium

phosphate. Journal of Crystal Growth, 2003. 249(3): p. 572-583.

39. Reddy, M.M., Crystallization of calcium carbonate in the presence of trace

concentrations of phosphorus-containing anions: I. Inhibition by phosphate and

Page 110: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

96

glycerophosphate ions at pH 8.8 and 25 C. Journal of Crystal Growth, 1977. 41(2): p.

287-295.

40. Reddy, M.M. and K.K. Wang, Crystallization of calcium carbonate in the presence of

metal ions: I. Inhibition by magnesium ion at pH 8.8 and 25 C. Journal of Crystal

Growth, 1980. 50(2): p. 470-480.

41. Pellegrino, J.L., Energy and environmental profile of the US chemical industry2000:

Office of Industrial Technologies, Energy Efficiency and Renewable Energy, US

Department of Energy.

42. Rozendal, R.A., et al., Effect of the type of ion exchange membrane on performance, ion

transport, and pH in biocatalyzed electrolysis of wastewater. Water Science and

Technology, 2008. 57(11): p. 1757-1762.

43. Jeremiasse, A.W., et al., Ni foam cathode enables high volumetric H< sub> 2</sub>

production in a microbial electrolysis cell. International Journal of Hydrogen Energy,

2010. 35(23): p. 12716-12723.

44. Tokash, J.C. and B.E. Logan, Electrochemical evaluation of molybdenum disulfide as a

catalyst for hydrogen evolution in microbial electrolysis cells. International Journal of

Hydrogen Energy, 2011. 36(16): p. 9439-9445.

45. Dempsey, B.A. Removal and reuse of ammonia and phosphate by precipitation of

struvite. in PROCEEDINGS OF THE INDUSTRIAL WASTE CONFERENCE-PURDUE

UNIVERSITY. 1998. MOSBY YEAR BOOK EUROPE LTD.

Page 111: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

Chapter 5

Energy Capture from Thermolytic Salt Solutions in Microbial Reverse

Electrodialysis Cells

5.1 Abstract

A system was developed capable of converting wastewater organics and low-grade

thermal energy (such as solar, geothermal, and waste heat) to electricity by using a

thermolytic salt, ammonium bicarbonate, as the saline solutions in a microbial reverse

electrodialysis cell (MRC). To examine the potential of this system, ammonium

bicarbonate solutions of a high concentration (HC) and low concentration (LC) were used

to generate electricity in a lab-scale MRC. The reverse-electrodialysis stack containing 5

pairs of HC and LC cells (HC = 1.1 M; LC = 0.011 M) directly contributed 2.4 ± 0.01

W/m2 (43%) to the maximum power, and also enhanced power recoveries from electrode

reactions. The electrode power density (from acetate oxidation) increased from 1.08 ±

0.03 W/m2 in a single chamber microbial fuel cell (MFC) to 3.3 ± 0.04 W/m2 in the

MRC. Maximum power density of the MRC reached 5.6 W/m2, which was 20% higher

than that reported with artificial seawater and river water. When fed domestic

wastewater, the MRC peak power density was 3.0 ± 0.05 W/m2, with the MFC electrodes

contributing 2.0 ± 0.05 W/m2, the highest reported value of electricity production from

domestic wastewater. The results of this study show the potential of using ammonium

Page 112: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

98

bicarbonate salt solutions to recovery energy from low grade thermal and wastewater

with a MRC.

5.2 Introduction

Worldwide energy demand is expected to increase to 23 TW by 2050 [1]. The

development of 10 TW of renewable energy [1] is needed to sustain a habitable

ecosystem while meeting the energy demand of a human population projected to eclipse

9 billion by mid-century [2]. Organic wastewater and salinity energy are two sources of

renewable energy that remain essentially untapped. Microbial fuel cells (MFCs) have

recently emerged as an efficient method of directly producing electricity from organic

matter present in wastewater (the treatment of which currently accounts for 1% of total

energy use in the U.S.[3])[4, 5]. Reverse electrodialysis (RED) is a salinity-gradient

energy system designed to convert the entropic energy released by mixing solutions of

different ionic-strength into electrical energy within a fuel cell arrangement [6-8]. A

recently proven combination of these technologies, the microbial reverse electrodialysis

cell (MRC), resulted in a syntrophic enhancement of energy recovery from both

wastewater and salinity-gradient energy [9].

MFCs capture electrical current produced by exoelectrogenic microbes that oxidize

soluble organic matter present in wastewater [10, 11]. MFCs contain anodes where

microbes grow and release electrons, and air cathodes, where electrons are consumed in

the oxygen reduction reaction [4]. Spontaneous power production in single chamber air

cathode MFCs (normalized to projected cathode area) has reached 2.7 W/m2 under

idealized conditions [12] but is substantially lower when reactors are fed complex and

Page 113: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

99

poorly conductive substrates such as domestic wastewater (0.26 – 0.45 W/m2) [13-15].

When coupled with a RED stack, the MRC produced 4.3 W/m2 when fed buffered acetate

as well as seawater and river water; however, its potential applications are limited only in

estuary areas or treatment plants that discharge treated effluent into the ocean [9].

In a RED system, a stack of alternating anion- and cation-exchange membranes

separating solutions of high concentration (HC) and low concentration (LC) is

constructed between fuel cell electrodes. The additive electrochemical junction potential

(0.1 – 0.2 V per cell pair for typical seawater and fresh water) [16] of the electrolytic-pile

provides a driving force for current generation at the electrodes. Because of this energy

losses at the electrodes, the net output energy of an RED system is highly dependent on

energy conversion at the electrodes especially under high current conditions [17]. By

coupling a RED stack with spontaneous MFC electrode kinetics (i.e., oxidation of

organic matter at the anode and oxygen reduction at the cathode), significant energy

losses which occur with water electrolysis (up to 2.5 V in cell voltage loss in a 25

membrane stack due to electrode overpotentials) [18] and system damage associated with

the hexacyanoferrate redox couple ([Fe(CN)6]3-/[Fe(CN)6]4-) such as membrane

poisoning, iron precipitation, and HCN gas production at the electrodes can be avoided

[17, 19-22]. The spontaneous nature of MRC current generation also allows salinity-

gradient energy generation with much fewer membranes (5 membrane pairs compared to

25 – 50) [9] representing a substantial reduction in the membrane associated capitol costs

($100 – 200/m2) [8].

At the time of this work, RED technology had only been proposed to harness energy

released by the mixing of seawater and fresh water limiting the application to estuaries

Page 114: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

100

(1.7 TW) [23], or for the case of MRC technology to ocean side wastewater treatment

plants (18.5 GW) [8]. If RED saline solutions were composed of ammonium bicarbonate,

which decomposes to ammonia and carbon dioxide gas at low temperatures (40 – 60°C),

[24] MRCs can create electricity from abundant sources of low grade thermal energy

such as waste heat (204 GW), [25] solar (120,000 TW) and geothermal (12 TW) energy

[1]. In this study, MRC power production using ammonium bicarbonate saline solutions

is investigated. The effects of saline solution concentration, salinity ratio (HC/LC), and

flow rate on power density were examined with both synthetic wastewater containing

acetate and effluent from primary clarifier at a local domestic wastewater treatment plant.

5.3 Materials and Methods

5.3.1 Reactor Construction

The lab scale MRC reactor was constructed as previously described [9] with minor

modifications. The 4-cm cubic anode chamber (Lexan, 30 mL empty bed volume)

contained a graphite brush anode (D = 2.7 cm, L = 2.3 cm, Mill-Rose Labs Inc., OH).

The brush anode was heat treated [26] before it was inoculated with the effluent from an

existing MFC and enriched in a conventional single chamber MFC prior to MRC

operation. The cathode chamber (2-cm cubic chamber, 18 mL empty bed volume)

contained a Pt loaded (0.5 mg Pt/cm2) carbon cloth passive air cathode constructed as

previously described [27] with a Nafion catalyst binder on the water side and four layers

of polytetrafluoroethylene diffusion layers on the air side. The cathode chamber also

Page 115: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

101

served as the first flow channel of the high concentrate salt stream to prevent the pH rise

in the cathode chamber.

The RED stack, assembled between the anode and cathode chambers, consists of with

6 anion- and 5 cation-exchange membranes (Selemion AMV and CMV, Asahi glass,

Japan), creating 5 pairs of alternating HC and LC chambers as previously described[9].

Inter-membrane chambers were sealed and separated by silicon gaskets, each with an 8

cm2 (2 × 4 cm) rectangular cross section cut out. Inter-membrane chamber width (1.3

mm) was maintained with a 2 cm2 (0.5 × 4 cm) strip of polyelthylene mesh. The total

ionic exchange membrane area in the RED stack was 88 cm2. The total MRC empty bed

volume was 58.4 mL (RED stack + Cathode = 28.4 mL; Anode = 30 mL). The HC

solution entered the reactor at the cathode and flowed serially through the 5 HC cells in

the stack, exiting from the cell next to the anode chamber. The LC stream entered the

RED stack near the anode and flowed serially through the 5 LC cells in the stack, exiting

from the cell next to the cathode chamber. A peristaltic pump (Cole Parmer, IL)

continuously fed the HC and LC solutions at a flow rate of 1.6 mL/min, unless specified

otherwise. During power density curve experiments fresh saline solutions were pumped

through the RED stack with the effluent collected in separate reservoirs. In batch recycle

experiments 0.1 L of each solution was recycled through the stack in airtight flow paths.

Before each batch the stack and tubing were flushed with matching solutions.

5.3.2 Solutions

Ammonium bicarbonate HC solution was prepared by dissolving ammonium

bicarbonate salt (Alfa Aesar, MA) into deionized water within an airtight vessel. The

Page 116: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

102

initial HC tested was 1.8, 1.1, 0.95, 0.8, and 0.5 M. The LC solutions were prepared

depending on given salinity ratios of 50, 100, and 200 by diluting an aliquot of the HC

solution. The anode solution contained 1 g/L of sodium acetate, in 50 mM carbonate

buffer (4.2 g/L NaHCO3-) containing 0.231 g/L NH4H2PO4 and trace vitamins and

minerals [28]. Domestic wastewater was collected from the primary clarifier of the Penn

State University wastewater treatment plant.

A second order relationship between ammonium bicarbonate solution concentration

and solution conductivity (determined by conducting a stepwise dilution series, Figure

S4) was used to estimate initial and final concentrations of HC and LC streams.

Conductivity and pH of the HC and LC streams were measured (Mettler-Toledo, OH)

before and after each batch recycle experiment.

5.3.3 Power Measurement

Power production in batch recycle system efficiency experiments was determined by

measuring the potential drop across a fixed external resistance (300 Ω) for both MRC and

single chamber MFC operations. Voltage drop was recorded every 20 minutes by a

digital multimeter (Keithley Instruments, OH). Electrical current (i) was determined by

Ohm’s law. Power was calculated by multiplying the electrical current and total cell

voltage. To determine the maximum MRC power (PMRC) production at each condition the

reactor was held at open circuit voltage for one hour and then the external resistance was

decreased from 1,000 to 50 Ω every 20 minutes with the voltage read at each resistance

interval. Power contribution by the electrode reactions (PMFC) was determined by

measuring the anode potential (Ean) and cathode potential (Ecat) against Ag/AgCl

Page 117: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

103

reference electrodes (BASi, IN): PMFC = (Ecat – Ean)× i. The RED stack power

contribution was calculated by finding stack voltage (Vstk) with two reference electrodes

located on both ends of the stack as: PRED = Vstk × i.

The MRC anode was transferred to a single chamber MFC to determine baseline

power production in batch mode and peak power production from carbonate buffered

acetate and domestic wastewater.

5.3.4 Analysis

Coulombic efficiency was determined as previously described [27]. Energy recovery

(rE) is defined by the ratio of energy produced by the MRC reactor and the energy input

as substrate and salinity gradient as written in Eq. (3). Energy efficiency (ηE) is the ratio

of energy produced over the energy consumed as substrate and salinity gradient, were

determined for batch recycle experiments [19]:

𝑟! =!!"#$

!!,!∙∆!!!∆!!"#,!∙ 100% (5.1)

𝜂! =!!"#$

(!!,!!!!,!)∙∆!!!(∆!!"#,!!∆!!"#,!)∙ 100% (5.2)

where EMRC is the energy produced per batch (kJ), ns is the moles of substrate (acetate)

fed to the anode initially (0) and at the end of the batch cycle (f), ∆Gs is the Gibb’s free

energy of substrate (acetate = –846.6 kJ/mol [29], domestic wastewater = 17.8 kJ/g-COD

[30]), ∆Gmix is the free energy that can be created by mixing of HC and LC solutions until

the two solutions reach equilibrium concentration as:

∆𝐺!"# = 𝑅𝑇 (𝑉!"𝑐!,!"𝑙𝑛!!,!"#!!,!"

+ 𝑉!"𝑐!,!"𝑙𝑛!!,!"#!!,!"

)! (5.3)

Page 118: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

104

where R is the ideal gas constant (8.314 J/mol-K), T is solution temperature, V is the

volume of solution, c is the molar concentration of ionic species i in the solution, and a is

the activity of species i in the solution.

At a neutral pH, concentrated ammonium bicarbonate is dominated by ammonium

(NH4+) and bicarbonate (HCO3

-) ions but significant amounts of carbamate (NH4CO3-)

and carbonate (CO32-) also contribute to ionic strength. Species specific concentrations

and activities were estimated with OLI Stream Analysis software (OLI Systems, Inc.,

Morris Plains, NJ) at a pH of 7 and temperature of 25 °C (Figures S4 and 5).

To determine ammonia transport into the anode, total ammonia nitrogen (TAN, NH3

+ NH4+) concentration in the substrate was determined before and after each fed-batch

cycle (HACH, CO) [31]. Free ammonia concentration (FAN) was determined also

determined for each fed-batch cycle:

𝑁𝐻! = 𝑇𝐴𝑁 1+ !"!!"

!"! !.!"!#$! !"!#.!"

!

!!

(5.4)

5.4 Results and Discussion

5.4.1 Effect of AmB Concentration on MRC Power Production

At a fixed salinity ratio (SR = 100; HC = 1.1 M; LC = 0.011 M) and flow rate (1.6

mL/min), the MRC produced a maximum power density of 5.6 ± 0.04 W/m2 (0.5 ± 0.003

W/m2 of projected membrane area) (Figure 5.2), 20% higher than the power density

achieved with artificial seawater and freshwater [9]. Peak power was reached at a total

cell voltage was 0.77 V and current density of 0.73 mA/cm2. At this salinity ratio and

Page 119: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

105

concentration interval, the RED stack contributed 2.4 ± 0.01 W/m2 (43%) and the

electrode reactions contributed 3.3 ± 0.04 W/m2 (57%) to the total MRC power.

As the molar concentration in the AmB HC solution increased from 0.5 to 1.8 M,

RED stack power increased from 1.2 ± 0.3 to 2.6 ± 0.1 W/m2 (Figure 5.2). By

maintaining salinity ratio (SR 100), the effect of ionic concentration within the stack

could be seen in the polarization curves of the MRC and RED stack at the various HC

concentrations (Figure 5.3, 5.4a). Since the salinity ratio was fixed at 100, the MRC open

circuit voltages of all HCs were identical. However, internal resistance within the MRC

(slope of the cell voltage polarization curves, Figure 5.3) decreased as ammonium

bicarbonate concentration increased, from 170 Ω at 0.5 M to 138 Ω at 1.8 M.

The RED stack directly contributed to MRC power and dramatically enhanced power

production from MFC electrodes. Electrode reactions produced nearly 300% more power

in the MRC than in a conventional single chamber MFC. The power associated with

electrode reactions in the MRC was 3.2 ± 0.2 W/m2 as compared to 1.08 ± 0.03 W/m2

(dashed line of Figure 5.2) from the same electrodes in a single chamber MFC. The stack

enhanced electrode power production by maintaining electrode potentials near open

circuit potentials as current density increased. In the single chamber MFC, the electrode

potentials approached each other (anode increased and cathode decreased) and as current

density increased (Figure 5.4b).

At a HC of 1.8 M maximum power decreased to 4.8 ± 0.1 W/m2. Upon reaching a

current density of 0.6 mA/cm2 (Figure 5.4b) the anode potential rapidly increased

(polarized), significantly reducing the MFC contribution to total MRC power. Anode

polarization was also observed at HC = 1.1 M, but at a much higher current density 0.9

Page 120: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

106

mA/cm2, and did not occur at lower HCs of 0.5, 0.8, 0.95 M. Anode polarization at

higher HCs caused permanent damage to the anode biofilm and reparative enrichment

was required to restore performance.

5.4.2 Effect of Salinity Ratio and Flow Rate on MRC Power Production

By varying salinity ratio at a fixed HC of 0.95 M (highest achieved power density

without anode polarization) it was determined that SR 100 produced the highest peak

MRC power production. Since MFC electrode performance was considerably stable over

a wide range of SRs from 50 to 200 (3.0 ± 0.01 W/m2) the difference in MRC peak power

density can be attributed to stack performance (Figure 5.5). Increasing salinity ratio from

100 to 200 increased the resistance of the LC chambers and resulted in a 34% drop in

stack power. Lowering SR to 50 reduced LC solution resistance but also lowered the

RED potential, resulting in a 20% decrease in stack power. With HC solution flowing

through both RED stack flow channels (SR 1), stack power dropped to zero but the MFC

electrodes still produced more power (1.7 ± 0.05 W/m2) in comparison to the single

chamber reactor (1.08 ± 0.03 W/m2). It is likely that the observed power increase was due

to the presence of HC solution in the cathode chamber (65.5 mS/cm conductivity).

Lowering the flow rate (Figure A2.7) from 1.6 to 0.85 mL/min (4.9 ± 0.1 W/m2) had

nearly the same effect as lowering the SR from 100 to 50 (4.7 ± 0.1 W/m2).

5.4.3 Energy Production in Batch Recycle Experiments

To maximize energy recovery and efficiency, HC and LC salt solutions (0.1 L each)

were recycled in airtight flow paths for the duration of anode feeding cycles. During

Page 121: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

107

recirculation, the salinity gradient decreased, reducing RED stack power contribution to

total MRC power. As a result, total MRC energy production was similar (108 ± 7 J,

Figure 6b) for all tested HC concentration intervals. Although the stack contribution to

total power decreased, the MRC energy recovery was more than 3× higher than the

energy recovery from acetate in a single chamber MFC reactor (32.6 ± 4 J). The salinity-

gradient energy input increased from 98 ± 1 J at 0.5 M (which was less than the MRC

power production at 0.5 M, 101 ± 2 J) to as high as 360 ± 1 J at 1.8 M. Acetate energy

(193 ± 6 J) was the highest input at all HC concentrations except 1.8 M indicating that

energy recovery and efficiency was mostly limited by the rate and efficiency of substrate

oxidation at the anode.

5.4.4 Energy Recovery and Efficiency

Energy recovery and efficiency were significantly higher in the MRC than in the

single chamber MFC. MRC energy recovery ranged from a high of 30 ± 0.5% at 0.5 M,

to 20 ± 0.01% at 1.8 M (Figure 5.6a). There was little variation in MRC energy efficiency

for HCs below 1.8 M (34 ± 0.5%). At 1.8 M, the MRC produced the lowest measured

energy efficiency (25 ± 0.1). In comparison, both energy recovery (14 ± 2%) and

efficiency (16 ± 2%) the single chamber MFC. The coulombic efficiency of acetate

oxidation in the MRC (66 ± 4%) was also markedly higher than the single chamber MFC

(35 ± 4%) because the RED stack prevented oxygen intrusion into the anode chamber of

the MRC. Reduced energy recovery and efficiency with increasing HCs may be

attributed to volatilization of ammonia and carbon dioxide through the air cathode as well

as osmotic water flux from the LC chambers. Also, the volatility of ammonia and

Page 122: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

108

carbonate at high HC concentrations caused observable bubble formation within the

stack, which could have limited the active membrane transport area.

5.4.5 Ammonia Transport into the Anode

Ammonia transport into the anode chamber from the RED stack affected anode

performance for HCs > 1 M. As anodic bacteria produce protons from the oxidation of

organic matter, negative bicarbonate and carbamate ions transported across the terminal

anion exchange membrane to balance charge within the anode chamber. Effluent TAN

concentrations in the anode after batch recycle experiments ranged from 263 ± 32 mg/L

at 0.5 M to 590 ± 36 mg/L at 1.8 M. Effluent FAN, which has been linked to extracellular

polysaccharide destruction in activated sludge [32] and power inhibition in single

chamber MFC [33], ranged from 1.5 ± 0.2 mg/L at 0.5 M to 3.3 ± 0.2 mg/L at 1.8 M

(Figure S8). The effluent TAN concentration for HC of 1.8 M exceeded previously

reported thresholds for power inhibition in MFCs (TAN = 500 mg/L) but inhibition was

observed in power density curves, not in batch recycles experiments.

5.4.6 Power Production from Domestic Wastewater

At a HC concentration of 0.95 M, SR of 100, and 1.6 mL/min flow rate, the peak

power density of the MRC fed domestic wastewater was 2.9 ± 0.05 W/m2 (Figure 5.7a).

The MFC electrode contribution to peak power was 2.0 ± 0.05 W/m2, a 740% increase in

power over when the same electrodes operated in a single chamber MFC fed the same

wastewater (0.27 ± 0.05 W/m2). This MFC electrode power is the highest power ever

Page 123: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

109

recorded with domestic wastewater and exceeds a recently reported value for a domestic

wastewater fed MFC with carbon nanotube coated electrodes by ~50% [34].

In batch recycle experiment with domestic wastewater, power production was

sustained for only two hours. As MFC electrode power dropped below the stack, the total

output power quickly reduced to zero (Figure 5.8). This inflection point corresponded to

when the anode potential quickly polarized from negative to positive. COD removal in

batch recycle experiments was 58 ± 5%. The energy production rate during batch recycle

was 0.94 kWh/kg-COD.

5.5 Conclusions

To examine the potential of a heat recovery microbial reverse electrodialysis cell

(MRC), concentrated (HC) and dilute (LC) ammonium bicarbonate solutions at various

concentrations were used to generate electricity in a lab-scale reactor. At an ammonium

bicarbonate HC concentration of 1.1 M, maximum power density reached 5.6 W/m2,

which was 20% higher than reported from an artificial sea and river water MRC. When

fed domestic wastewater, the MRC achieved a peak power density of 2.9 ± 0.05 W/m2

with the electrodes reactions contributing 2.0 ± 0.05 W/m2. The RED stack directly

contribute to total MRC power (2.4 ± 0.01 W/m2 with acetate, 0.8 ± 0.01 W/m2 with

domestic wastewater) and symbiotically enhanced the power production of MFC

electrodes fed acetate (330% increase) and domestic wastewater (740% increase).

Page 124: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

110

5.6 Figures

Figure 5.1: (a) Main components of the microbial reverse electrodialysis cell (MRC), showing the membrane stack between the electrodes, the reference electrodes and the circuit containing a load (resistor). (b) Example of how the anion- (AEM) and cation- (CEM) exchange membranes are used to selectively drive the flow of positive ions to the right (towards the cathode) and the negatively-charged ions to the left (towards the anode). (c) Expanded view of the membrane stack showing flow path of the high (HC) and low (LC) concentrate solutions of ammonium bicarbonate. (d) Construction of the gaskets used to direct the flow from one LC chamber to the next LC chamber, avoiding the HC chamber through a short flow path through the membrane and gasket.

Page 125: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

111

Figure 5.2: Peak power density of MRC as well as the contributions to total power from the RED stack, and MFC electrodes at various HC concentrations. The dashed line represents peak power density of the same electrodes in a single chamber MFC.

0

1

2

3

4

5

6

0.5 1.0 1.5 2.0

Peak

Pow

er D

ensi

ty (W

/m2 )

HC Concentration (M)

MRC Electrodes RED

MFC

Page 126: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

112

Figure 5.3: Polarization curves of the MRC at various HC concentrations as well and the polarization curves of the electrodes in a single chamber MFC.

0.0

0.5

1.0

1.5

2.0

0.0 0.2 0.4 0.6 0.8 1.0

Cel

l Vol

tage

(V)

Current Density (mA/cm2)

MFC0.5M0.8M0.95M1.1M1.8M

Page 127: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

113

-0.4

-0.2

0.0

0.2

0.4

0.6

0.0 0.2 0.4 0.6 0.8 1.0

Ele

ctro

de P

onte

ntia

l vs.

SH

E (V

)

Current Density (mA/cm2)

1.1M C 1.1M A 1.8M C

1.8M A 0.95M A 0.95M C

SC C SC A

0.0

0.3

0.5

0.8

1.0

RE

D S

tack

Vol

tage

(V) 1.8M

1.1M0.95M

b)

a)

Figure 5.4: a) RED stack potential as well as b) anode (filled symbols) and cathode (open symbols) potentials, measured during polarization curves, in the single chamber MFC (SC) and the three highest HC concentrations.

Page 128: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

114

Figure 5.5: Salinity ratio vs. peak power density at HC concentration of 0.95 M and flow rate of 1.6 mL/min. Dashed line represented peak power density of single chamber MFC fed identical substrate.

0

1

2

3

4

5

6

0 50 100 150 200

Peak

Pow

er D

ensi

ty (W

/m2 )

Salinity Ratio

MRC Electrodes RED

MFCSR1

Page 129: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

115

0

10

20

30

40

Rec

over

y/Ef

ficie

ncy

(%)

nErE

MFC

0

50

100

150

200

250

300

350

0.5 1.0 1.5 2.0

Ene

rgy

per B

atch

(J)

HC concentration (M)

Salinity InAcetate InMRC Out

MFC

Figure 5.6: a) Energy recovery, efficiency and b) and energy balance of the MRC vs. HC concentration.

a)

b)

Page 130: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

116

-0.4-0.3-0.2-0.10.00.10.20.30.4

0.0 0.2 0.4 0.6

Elec

trode

Pon

tent

ial v

s. S

HE

(V)

Current Density (mA/cm2)

MRC WW A MFC WW AMRC WW C MFC WW C

0.0

1.0

2.0

3.0

Pow

er D

ensi

ty (W

/m2 )

MRC WW

MFC WW

Figure 5.7: a) Peak power density and (b) anode and cathode potentials of MRC and single chamber MFC fed domestic wastewater.

a)

b)

Page 131: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

117

Figure 5.8: Batch recycle component (MFC, RED and total MRC) power profile of MRC fed domestic wastewater.

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0 0.1 0.2 0.3 0.4 0.5 0.6

Pow

er (W

/m2 )

Time (days)

MRCREDMFC

Page 132: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

118

5.7 Literature Cited

1. Kamat, P.V., Meeting the clean energy demand: Nanostructure architectures for solar

energy conversion. The Journal of Physical Chemistry C, 2007. 111(7): p. 2834-2860.

2. Division, U.N.P., World Population Prospects: the 2010 Revision, 2010, United Nations.

3. IPCC, Climate Change 2007: Synthesis Report. , in Contribution of Working Groups I, II

and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate

Change2007, IPCC: Geneva.

4. Liu, H., R. Ramnarayanan, and B.E. Logan, Production of electricity during wastewater

treatment using a single chamber microbial fuel cell. Environmental Science &

Technology, 2004. 38(7): p. 2281-2285.

5. Cusick, R.D., P.D. Kiely, and B.E. Logan, A monetary comparison of energy recovered

from microbial fuel cells and microbial electrolysis cells fed winery or domestic

wastewaters. International Journal of Hydrogen Energy, 2010. 35(17): p. 8855-8861.

6. Wick, G.L., Power from salinity gradients. Energy, 1978. 3(1): p. 95-100.

7. Post, J.W., et al., Salinity-gradient power: Evaluation of pressure-retarded osmosis and

reverse electrodialysis. Journal of Membrane Science, 2007. 288(1-2): p. 218-230.

8. Ramon, G.Z., B.J. Feinberg, and E.M.V. Hoek, Membrane-based production of salinity-

gradient power. Energy & Environmental Science, 2011.

9. Kim, Y. and B.E. Logan, Microbial Reverse Electrodialysis Cells for Synergistically

Enhanced Power Production. Environmental Science & Technology, 2011. 45(13): p.

5834-5839.

10. Logan, B.E., Exoelectrogenic bacteria that power microbial fuel cells. Nat Rev

Microbiol, 2009. 7(5): p. 375-81.

11. Logan, B.E., et al., Microbial fuel cells: methodology and technology. Environ Sci

Technol, 2006. 40(17): p. 5181-92.

Page 133: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

119

12. Fan, Y., H. Hu, and H. Liu, Sustainable Power Generation in Microbial Fuel Cells Using

Bicarbonate Buffer and Proton Transfer Mechanisms. Environmental Science &

Technology, 2007. 41(23): p. 8154-8158.

13. Ahn, Y. and B.E. Logan, Effectiveness of domestic wastewater treatment using microbial

fuel cells at ambient and mesophilic temperatures. Bioresource Technology, 2010.

101(2): p. 469-475.

14. Liu, H. and B.E. Logan, Electricity generation using an air-cathode single chamber

microbial fuel cell in the presence and absence of a proton exchange membrane.

Environmental Science & Technology, 2004. 38(14): p. 4040-4046.

15. Cheng, S., H. Liu, and B.E. Logan, Increased power generation in a continuous flow

MFC with advective flow through the porous anode and reduced electrode spacing.

Environmental Science & Technology, 2006. 40: p. 2426-2432.

16. Weinstein, J.N. and F.B. Leitz, Electric power from differences in salinity: the dialytic

battery. Science, 1976. 191(4227): p. 557.

17. Veerman, J., et al., Reverse electrodialysis: Performance of a stack with 50 cells on the

mixing of sea and river water. Journal of Membrane Science, 2009. 327(1-2): p. 136-144.

18. Veerman, J., et al., Reverse electrodialysis: Comparison of six commercial membrane

pairs on the thermodynamic efficiency and power density. Journal of Membrane Science,

2009. 343(1-2): p. 7-15.

19. Post, J.W., H.V.M. Hamelers, and C.J.N. Buisman, Energy Recovery from Controlled

Mixing Salt and Fresh Water with a Reverse Electrodialysis System. Environmental

Science & Technology, 2008. 42(15): p. 5785-5790.

20. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Doubled power density from salinity

gradients at reduced inter-membrane distance. Environmental Science & Technology,

2011. 45: p. 7089-7095.

Page 134: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

120

21. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Power generation using profiled

membranes in reverse electrodialysis. Journal of Membrane Science, 2011.

22. Lacey, R., Energy by reverse electrodialysis. Ocean engineering, 1980. 7: p. 1-47.

23. Kuleszo, J., et al., The potential of blue energy for reducing emissions of CO2 and non-

CO2 greenhouse gases. Journal of Integrative Environmental Sciences, 2010. 7(S1): p.

89-96.

24. McGinnis, R.L., J.R. McCutcheon, and M. Elimelech, A novel ammonia-carbon dioxide

osmotic heat engine for power generation. Journal of Membrane Science, 2007. 305: p.

13-19.

25. EIA, U.S., Annual Energy Review 2010, DOE, Editor 2010.

26. Wang, X., et al., Use of carbon mesh anodes and the effect of different pretreatment

methods on power production in microbial fuel cells. Environmental Science &

Technology, 2009. 43(17): p. 6870-6874.

27. Cheng, S., H. Liu, and B.E. Logan, Increased performance of single-chamber microbial

fuel cells using an improved cathode structure. Electrochemistry Communications, 2006.

8(3): p. 489-494.

28. Ambler, J., Perfomance of stainless steel 304 cathodes and bicarbonate buffer in a

Microbial Electrolysis Cell using a new methode of gas characterization, in Civil and

Environmental Engineering2010, The Pennsylvania State University: University Park.

29. Rozendal, R.A., et al., Towards practical implementation of bioelectrochemical

wastewater treatment. Trends in Biotechnology, 2008. 26(8): p. 450-459.

30. Heidrich, E.S., T.P. Curtis, and J. Dolfing, Determination of the Internal Chemical

Energy of Wastewater. Environmental Science & Technology, 2010. 45(2): p. 827-832.

31. APHA, ed. Standard Methods for the Examination of Water and Wastewater. 20th ed.,

ed. L.S. Clesceri, A.E. Greenberg, and A.D. Eaton1998, American Public Health

Page 135: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

121

Association, American Water Works Association, Water Environment Federation:

Washington DC.

32. Yang, S.-F., J.-H. Tay, and Y. Liu, Inhibition of free ammonia to the formation of aerobic

granules. Biochemical Engineering Journal, 2004. 17(1): p. 41-48.

33. Nam, J.Y., H.W. Kim, and H.S. Shin, Ammonia inhibition of electricity generation in

single-chambered microbial fuel cells. Journal of Power Sources, 2010. 195(19): p. 6428-

6433.

34. Xie, X., et al., Carbon nanotube-coated macroporous sponge for microbial fuel cell

electrodes. Energy & Environmental Science, 2012.

Page 136: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

Chapter 6

Minimal RED Cell Pairs Markedly Improve Electrode Kinetics and

Power Production in Microbial Reverse Electrodialysis Cells

6.1 Abstract

Power production from microbial reverse electrodialysis cell (MRC) electrodes is

substantially improved compared to microbial fuel cells (MFCs) by using ammonium

bicarbonate (AmB) solutions in a multiple-membrane stack and the cathode chamber, but

reducing the number of membranes pairs could help to reduce capital costs. We show

here that using only a single RED cell pair (CP), created by operating the cathode in

concentrated AmB, dramatically increased power production normalized to cathode area

from both acetate (Acetate: from 0.9 to 3.1 W/m2-cat) and wastewater (WW: 0.3 to 1.7

W/m2), by reducing solution and kinetic resistances at the cathode. The addition of a

second RED cell pair further increased power production (Acetate: 4.2 W/m2; WW: 1.9

W/m2) and anode biofilm activity. These power densities are close to those previously

achieved using 11 membranes, indicating near optimal electrode performance with only

one or two cell pairs. When normalized to total membrane area, the 1-CP (Acetate: 3.1

W/m2-mem; WW: 1.7 W/m2) and 2-CP (Acetate: 1.3 W/m2; WW: 0.6 W/m2) power

densities were much higher than previous MRCs (0.3 − 0.5 W/m2-membrane with

acetate). The rate of wastewater COD removal, normalized to reactor volume, was 30 –

Page 137: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

123

50 times higher in 1-CP and 2-CP MRCs than that in a single chamber MFC. It is

estimated that further reductions in RED solution and transport resistance could increase

MRC power production to 6 – 8 W/m2-cat (2 – 3 W/m2-mem). These findings show that

even a single cell pair AmB RED stack can significantly enhance the electrical power

production and wastewater treatment.

6.2 Introduction

Microbial reverse electrodialysis cells (MRCs) are biotechnologies designed to generate

renewable energy from unconventional sources of organic wastewater and salinity gradients. In

the United States alone, ~155 GWh (1.23 kWh/m3) could be generated from organics in

wastewater [1-3]. Ammonium bicarbonate (AmB) is a thermolytic salt that decomposes to

ammonia and carbon dioxide gas at low temperatures (40 – 60°C) [4]. The low decomposition

temperature of AmB could enable generation of salinity gradients from abundant sources of low

grade thermal energy such as waste heat (204 GW) [5], geothermal (12 TW), and solar (120,000

TW) energy [6].

Within a MRC, the same electrodes used in a microbial fuel cell (MFC) are placed on each

side of a reverse electrodialysis (RED) membrane stack [7]. MFCs spontaneously generate

electrical current by pairing a bio-anode, on which exoelectrogenic microbes oxidize soluble

wastewater organics and release electrons [8], and cathodes with oxygen reduction [9]. The

entropic energy released by mixing solutions of different ionic strength is converted in a RED

stack into electrical energy by separating chambers of high concentration (HC) and low

concentration (LC) saline solutions with alternating anion (AEM) and cation (CEM) exchange

membranes [10-12]. The additive electrochemical junction potential (typically 0.1 – 0.2 V per cell

pair) [13] of the electrolytic-pile provides additional driving force for current generation at the

Page 138: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

124

MFC electrodes. Power generation in MRCs is synergistic because the RED stacks not only

directly contribute voltage to power production but also enhance MFC electrode performance [7].

When the electrodes were transferred from a single-chamber MFC to a MRC with an 11

membrane RED stack fed AmB solutions, power production increased by >300% using acetate

and by >700% with domestic wastewater [14].

The improved performance of the MRC relative to a MFC is due to the use of multiple pairs

of membranes in the RED stack, the use of AmB in the stack, and the improved electrode

potentials. Power production with an 11 membrane stack with acetate reached 5.6 W/m2 based on

projected cathode area (or equivalently cross sectional area between the electrodes), compared to

1.1 W/m2 using only the MFC. While it is typical to normalize power density based on projected

cathode area for MFCs, the total membrane area is typically reported in RED studies. When

evaluated on the basis of total membrane area, MRC power densities have ranged from 0.3 to 0.5

W/m2 [7, 14]. The use of these membranes in the MRC adds substantial capital costs for reactor

construction compared to MFCs. Thus, reducing the number of membranes pairs or further

increasing power production could make MRCs a more commercially viable biotechnology.

The effect of AmB on MRC performance relative to that of NaCl has not been well

examined. AmB solutions can be prepared to create larger salinity gradients than those possible

with river water and seawater. However, there are other important differences that have not been

previously addressed. A HC solution with NaCl in the catholyte improves conductivity, but it

does not buffer pH changes. As a result, the cathode pH increases due to the consumption of

protons by the cathode and accumulation of OH–, leading to potential losses of >0.3 V [15-17].

Phosphate or carbonate buffers can be used to mitigate pH changes, but comparisons to high

conductivity NaCl solutions have shown that these negatively-charged buffer species do not

improve performance [17]. The use of AmB as the catholyte may have advantages compared to

these other chemical species because positively charged ammonium ions would be attracted to the

Page 139: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

125

cathode, which could greatly affect oxygen reduction kinetics as well as hydroxide ion gradients

near the electrode. Since AmB has only been used as a catholyte in the presence of an 11

membrane RED stack [14], the contribution of improved cathode kinetics to MRC electrode

performance has not been separately examined from that of the stack performance.

The objectives of this study were to investigate potential catalytic benefits AmB on oxygen

reduction, and determine to what extent MRC performance could be improved using only a

minimal number of RED cell pairs. To study the effects of the AmB and cell pairs on MRC

performance, power production and internal resistance were examined using three different

reactor configurations: (1) a single chamber MFC with a bicarbonate buffer used as the single

electrolyte in contact with both electrodes; (2) a one cell pair (1-CP) MRC where the anolyte

formed an equivalent LC chamber with the bicarbonate buffer, and the catholyte was the HC

solution (1 M AmB); and (3) a two cell pair (2-CP) MRC containing an additional membrane pair

containing AmB LC and HC solutions (Figure 6.1). To investigate the effect of anolyte

composition on electrode and RED stack performance, sodium acetate (in 50 mM bicarbonate

buffer) and domestic wastewater were used as fuels. The effect on anolyte conductivity was

further examined in the 1-CP MRC configuration by amending the domestic wastewater with

bicarbonate buffer (50 mM). To quantify the specific effects of AmB catholyte and RED stack

potential on electrode activity, internal resistance was measured using linear polarization and

galvanostatic electrochemical impedance spectroscopy.

6.3 Materials and Methods

6.3.1 Reactor Construction

MFCs and MRCs were constructed as previously described [7] with minor

modifications. MFCs had a single 4-cm diameter chamber (Lexan, 28 mL empty bed

Page 140: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

126

volume) (Figure 6.1). To maintain the same electrode spacing in both MFC and MRC

tests, the anode and cathode chambers of 1- and 2-CP MRCs were constructed using two,

2-cm long cubic chambers (14 mL empty bed volume). The anodes were graphite fiber

brushes (2.7 cm in diameter, 2.3 cm in length; Mill-Rose Labs Inc., OH) that were heat

treated [18] prior to being inoculated with effluent from an existing MFC, and enriched in

single chamber MFCs. The same anodes were used in all three reactor configurations.

The air-cathodes contained a Pt catalyst (0.5 mg Pt/cm2) on carbon-cloth, and were

constructed as previously described [19] with a Nafion catalyst binder on the water side,

and four layers of polytetrafluoroethylene (PTFE) diffusion layers on the air side.

The 1-CP MRC contained a single anion exchange membrane (1-CP; Selemion

AMV, Asahi Glass, Japan, 7 cm2 active area) separating the anode and cathode chambers

(Figure 6.1). The 2-CP MRC reactor contained an additional HC/LC cell pair between the

anode and cathode chambers, requiring three membranes (1 cation- and 2 anion-

exchange membranes, Selemion CMV and AMV, Asahi glass, Japan). Inter-membrane

chambers were sealed and separated by silicon gaskets (0.5 mm), each with an 8 cm2 (2

cm × 4 cm) rectangular cross section cut out. The total ionic exchange membrane area in

the 2-CP MRC was 24 cm2. The HC solution entered the reactor at the cathode and exited

from the HC cell next to the anode chamber. The LC AmB stream entered the RED stack

near the anode and flowed upward through the LC cell in the stack. Both the HC and LC

AmB solutions were continuously fed using a peristaltic pump (Cole Parmer, IL) at a

flow rate of 1.0 mL/min. During polarization tests fresh saline solutions were pumped

through the RED stack with the effluent collected in separate reservoirs. In batch recycle

experiments 0.1 L of each solution was recycled through the stack in airtight flow paths.

Page 141: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

127

Before each fed-batch experiment the stack and tubing were flushed with matching

solutions.

6.3.2 Solutions

Anodes were fed either a buffered sodium acetate solution, or raw domestic

wastewater except as noted. The bicarbonate buffered acetate solution was designed to

reduce solution and membrane resistances and provide a high concentration of electron

donor for the anodic biofilm. The acetate solution contained 2.0 g/L of sodium acetate in

the 50 mM bicarbonate buffer (4.2 g/L NaHCO3) as well as 0.23 g/L NH4H2PO4, 0.16

g/L KCl, and trace vitamins and minerals [20]. The volume of acetate was 30 mL for

each reactor configuration. In the 1-CP and 2-CP configurations, the reduced anode

volume required recirculation of 16 mL of sodium acetate solution. Domestic wastewater

(WW) samples were collected from the Penn State University wastewater treatment plant

primary clarifier effluent. High conductivity wastewater (WW-HC) was prepared by

dissolving 4.2 g/L of NaHCO3 in the wastewater. To avoid electron donor limitations that

could occur due to low COD, experiments using wastewater were conducted by recycling

250 mL of substrate through the anode chamber of each reactor. The anolyte feed

reservoirs were airtight glass vessels submerged in ice water to minimize external organic

degradation.

The HC ammonium bicarbonate solution used in the 1- and 2-CP MRC reactors was

prepared by dissolving chemical grade ammonium bicarbonate salt (Alfa Aesar, MA) into

deionized water within an airtight vessel. The initial HC (1.0 M) and LC (0.01 M)

concentrations tested were selected based on previous optimization experiments [14].

Page 142: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

128

6.3.3 Electrochemical Measurements and Analysis

Power densities and bulk internal resistances for each reactor configuration were

determined using chronopotentiometry. Whole cell current intervals (20 minutes) were

set with a galvanostat (Biologic VMP-3, USA). Prior to galvanostatic operation, reactors

were held at open circuit for 30 minutes.

Power density was calculated by multiplying the electrical current (i) and total cell

voltage. MFC electrode power contribution (PMFC) was determined by measuring the

anode potential (Van) and cathode potential (Vcat) against Ag/AgCl (+0.210 V vs SHE)

reference electrodes (BASi, IN), and calculated as PMFC = i(Vcat – Van).[14] Power

contributed by the RED stack was calculated as PRED = iVRED based on the stack voltage

(VRED), which was measured using two reference electrodes located on either side of the

stack.

MRC internal resistance (RMRC) is the sum of potential dissipative forces at the anode

(Ran), cathode (Rcat), and within the RED stack (Rstk), or RMRC = Ran + RRED + Rcat. Internal

resistance can be further separated into solution, membrane, kinetic, diffusion and

capacitive resistances. The solution and membrane resistance (Rsol+m) is dependent on

electrode spacing, electrolyte and membrane conductivity, as well as diffusion area.

Kinetic charge transfer occurs during current generation at the anode (Ran,ct) and current

consumption at the cathode (Rcat,ct). Diffusion resistance can arise due to transport of

substrate to the anode (Ran,diff) as well as protons and oxygen to the cathode (Rcat,diff).

Diffusion boundary layer resistance (RRED,dbl), also known as concentration polarization,

is caused by concentration and depletion of ions at the interface of ion-exchange

Page 143: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

129

membranes. The fixed charges of IEMs also create ionic double layer capacitance

(RRED,dl) that impede ion diffusion within the RED stack [21-23].

Bulk internal resistances of the whole cell, anode, cathode and RED membrane

interfaces were determined from the slopes of linear polarization curves constructed for

each reactor component [22, 24]. Galvanostatic EIS (GEIS) was used to characterize the

effect of RED stacks on specific resistances. GEIS was conducted at a set current of 1.5

mA and amplitude of 0.21 mA. GEIS spectra were simultaneously recorded for each

reactor component by using three galvanostat channels operated in unison (one master

channel that applied AC current and two reference channels that collected EIS spectra).

Whole cell impedance was collected from the master channel terminals by setting the

cathode as the working electrode and the anode as both the counter and reference

electrode. The two reference channels were used to obtain anode, RED, and cathode

impedance spectra. Anode and cathode impedance was measured against an Ag/AgCl

reference electrode. RED stack impedance was obtained with reference electrodes

positioned at either side of the membrane stack. All EIS experiments were conducted

within one hour of chronopotentiometry measurements to ensure similar working

conditions. EIS data were analyzed using EC-Lab V10.32 to determine the specific

resistances of each reactor component. Equivalent circuit model (ECM) software was

used to determine specific resistances from the Nyquist plots of each reactor component

(Figure A3.5).

During batch-fed experiments, the voltage drop across an external resistor was

recorded every 10 minutes by a digital multi-meter (Keithley Instruments, OH). To

maximize energy production, the fixed external resistance was set to match the internal

Page 144: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

130

resistance of the cell at each condition as determined from polarization data. All batch

recycle experiments were conducted at 30 °C in a constant temperature room.

Elelctrical energy recovery, in batch experiments with recycled anolyte, was

calculated as the energy produced (Wh) from the substrate based on chemical oxygen

demand removed (g-COD). Power production (W) was calculated for each measured

voltage (V2/Rext), multiplied by the sample interval (h) and summed over the entire fed-

batch cycle. Volumetric COD removal rates were normalized by the duration of the fed-

batch cycle and the anode (an) volume (g-COD/Lan-d) to allow direct comparison to

commercialized technologies. COD removal was determined as previously described

[25].

6.4 Results and Discussion

6.4.1 MRC Power Production

The use of the concentrated AmB solution as the catholyte in the 1-CP MRC (two

chamber) configuration resulted in a >300% increase in power (3.1 ± 0.1 W/m2)

compared to that obtained the single-chamber MFC containing only bicarbonate buffer

(0.9 ± 0.05 W/m2-cat;). The 1-CP MRC also produced higher power with domestic

wastewater (1.8 ± 0.03 W/m2) than the MFC (0.3 ± 0.02 W/m2-cat). Previously, power

densities produced using two-chambered MFCs with concentrated NaCl or a phosphate

buffer as catholytes, have always been lower those obtained with single-chamber MFCs

[15, 17]. The use of concentrated AmB catholyte in the two-chamber configuration (1-

CP) reduced several limitations compared to two-chamber MFCs for bio-electricity

production, that include: transport of HCO3– into the anode chamber instead of Cl–

Page 145: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

131

buffered anode pH and prevented acidification; the presence of NH4+ near the cathode

mitigated accumulation of OH-; a favorable concentration gradient that created a positive

membrane potential and minimized power loss due to membrane resistance [17].

The inclusion of a second RED cell pair between the electrodes (2-CP) further

increased the peak power density to 4.1 ± 0.1 W/m2-cat (1.3 ± 0.1 W/m2-mem, Figure

6.2a). Power density from the 2-CP reactor fed wastewater was only slightly higher (1.9

± 0.05 W/m2-cat, 0.6 ± 0.02 W/m2-mem, Figure 6.2b) than the 1-CP configuration. When

fed the dilute and complex organic matter in wastewater, bio-anodes did not sustain

negative potentials at current densities beyond 6.5 A/m2-cat (Figure 6.2d). At these

higher current densities, rapid anode polarization was observed. The low conductivity of

wastewater also significantly reduced the effectiveness of the RED stack.

6.4.2 Electrode and RED Stack Potential Losses in MRCs

At peak power production in both 1- and 2-CP MRCs, the RED stack potential was

nearly 0 mV or negative. Under these conditions, the major reason for the power increase

relative to MFC conditions was therefore a reduction in electrode potential loss at higher

current densities. Under direct current conditions (DC), electrode potential loss (mV/mA)

can be quantified as polarization resistance (Ω). During operation with the high

concentrate AmB electrolye, cathode potential loss with increased current was uniformly

reduced (relative to the MFC) in both the 1- and 2-CP reactor configurations (Figure

6.2c,d). The cathode linear polarization resistance in the single-chamber MFC was 55 Ω

(or 55 mV/mA) with buffered acetate, and 220 Ω with domestic wastewater. When

operated in HC AmB, the cathode polarization resistance dropped dramatically to 12 - 15

Page 146: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

132

Ω in both the 1-CP and 2-CP MRC reactors, and was independent of anolyte

composition.

Direct current anode potential loss was influenced by the use of AmB catholyte as

well as the magnitude of RED stack potential. In the single chamber MFCs, anode

potential loss was highly dependent on the type of anolyte, with nearly a 90 Ω (or

mV/mA) difference between bicarbonate buffered acetate (33 Ω) and domestic

wastewater (120 Ω). In the 1-CP MRC, anode potential loss decreased significantly for

both substrates (Ac: 21 Ω, WW: 28 Ω). In the 2-CP MRC configuration, additional RED

stack voltage minimized anode DC resistance to the point at which potential loss was not

appreciably affected by substrate composition (10 Ω in WW versus 8 Ω in buffered

acetate, Figure S1). The similarity in 2-CP MRC anode potential loss implies that the

magnitude of RED potential may influence bio-anode activity.

Electrode potentials, and the rate of potential loss, in the 1-CP and 2-CP MRCs were

nearly identical with both organic fuel solutions, but power production from acetate

remained significantly higher than from wastewater. This disparity can be linked to the

effect that substrate composition had on RED stack resistance. With buffered acetate in

the anode chamber, RED resistance was 17 Ω in the 1-CP reactor, and it increased to 36

Ω with the inclusion of an additional cell pair (Figure 4b). When domestic wastewater

filled the anode chamber, the RED stack resistance significantly increased in both the 1-

CP (38 Ω) and 2-CP (97 Ω) MRCs. In order to increase power production from low

conductivity wastewater, transport resistance at the terminal AEM/anolyte interface must

be reduced. Minimizing anode chamber thickness would reduce solution resistance and

Page 147: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

133

increasing solution flow rate at the membrane surface could reduce boundary layer

resistance [23, 26].

6.4.3 Anolyte Composition and Membrane Transport Resistance

The two anode substrates evaluated in this study had different characteristics in terms

of electrode donor concentration and composition, buffer capacity, and specific

conductance. In order to understand how buffer concentration influenced the specific

resistances limiting MRC power production, experiments using GEIS were conducted on

the individual reactor components (anode, RED interface, and cathode) with the same set

current (1.5 mA) and amplitude (0.21 mA), using three different anode substrates:

buffered acetate; raw domestic wastewater (WW); and wastewater amended with 4.2 g/L

NaHCO3 (WW-HC) that matched the conductivity of the buffered acetate.

The differences between membrane interface impedance values in the 1-CP MRC

were primarily due to anolyte conductivity. Membrane resistance determined by

alternating current (AC) impedance spectroscopy showed ohmic resistance (from anode

and cathode solutions as well as the AEM, Rsol+m) and two overlapping semicircles

representing double layer capacitance (Rdl) and diffusion boundary layer (Rdbl) resistance

(Figure A3.3c). All three types of membrane impedance were highest when the MRC was

fed raw domestic wastewater (Rsol+m: 8 ± 0.1 Ω, Rdl: 15 ± 1.3 Ω, Rdbl: 17 ± 1.1 Ω, Figure

6.3b). Raising the conductivity of the wastewater decreased both double layer and

diffusion boundary layer resistance (Rsol+m: 6.0 ± 0.1 Ω, Rdl: 6.0 ± 0.4 Ω, Rdbl: 12 ± 0.2 Ω,

Figure 6.3b). By raising anolyte conductivity and reducing membrane transport

resistance, power density of the 1-CP MRC fed high conductivity wastewater (2.1 ± 0.05

Page 148: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

134

W/m2, Figure S2a) exceeded that of the 2-CP MRC with raw domestic wastewater.

However, raising wastewater conductivity did not reduce membrane impedance to the

levels of buffered acetate (Rsol/m: 5.0 ± 0.3 Ω, Rdl: 4.0 ± 0.1 Ω, Rdbl: 7.0 ± 0.4 Ω). This

implies that some other distinguishing characteristic of the wastewater, such as presence

of particulate organic matter or other ions, may have adversely affect membrane

capacitance and boundary layer resistance. The summation of individual membrane

resistances determined by GEIS closely corresponded with linear polarization resistance

determined under direct current (DC) conditions (differences ranged from 0.5 − 2.8 Ω).

Electrode resistance values determined by GEIS were nearly identical for each

substrate condition. Alternating current (AC) cathode impedance included 3.7 ± 0.4 Ω of

solution resistance and 12 ± 0.5 Ω of charge transfer resistance for all three conditions

(Figure 3b, S3c). The total cathode impedance values, based on the sum of individual AC

resistances, were all quite similar (within 2 Ω) to cathode resistance values determined by

the linear polarization slope method (DC resistance). Anode impedance spectra for all

three conditions were comprised of a single charge transfer semicircle (Figure S3a).

Anode charge transfer resistance values ranged from 13 − 15 Ω (Figure 3b, S3a) and

were lower than DC resistances for all three substrates (Figure 3a,b). The discrepancy

between DC (linear polarization) and AC (GEIS) anode resistance (WW: 15 Ω, WW-HC:

13 Ω, Ac: 7 Ω) decreased with the magnitude of RED stack resistance (WW: 40 Ω, WW-

HC: 24 Ω, Ac: 16 Ω) implying that 1-CP anode potential, under direct current conditions,

may have been adversely affected by the negative potential of the membrane interface.

Page 149: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

135

6.4.4 Effect of RED Stack Potential on Bio-Anode Charge Transfer Resistance

Anode resistances, determined from linear polarization experiments (DC resistance),

sharply decreased as internal resistance decreased (1-CP) and RED stack potential

increased (2-CP). GEIS was used to track anode biofilm catalytic activity, represented by

charge transfer resistance, in the single chamber MFC, 1-CP and 2-CP MRCs. In each

configuration, the anode was fed buffered acetate and GEIS was conducted as the same

set current (1.5 mA) and amplitude (0.21 mA).

Nyquist plots of anode impedance in the single chamber MFC showed ohmic solution

resistance (Rsol: 12 ± 0.1 Ω) between anode and reference as well as charge transfer and

diffusion resistance (Rct: 16 ± 0.01 Ω, Rdiff: 11 ± 0.7 Ω, Figure 6.4b, S4a). Anode

impedance in the 1-CP configuration was represented by a single charge transfer

semicircle that was similar, based on radius (Rct: 14 ± 0.2 Ω, Figure 6.4b, S4a), to the

anode kinetic resistance in the single chamber MFC. This similarity in charge transfer

implies that eliminating solution and diffusion resistance in the 1-CP reactor led to the

observed reduction in anode potential loss. In the 2-CP reactor, anode charge transfer

resistance decreased to 4.2 ± 0.8 Ω. At the GEIS set current, the 1-CP reactor RED

potential was 0.005 mV and 2-CP RED potential was 150 mV, implying that positive

RED potential not only adds to MRC power, but also enhances the catalytic activity of

anodic biofilms.

6.4.5 Wastewater Treatment Rates and Energy Recovery

To investigate how increased MRC power would translate to COD removal and

energy recovery, additional tests were conducted with buffered acetate or domestic

Page 150: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

136

wastewater recycled through the anode chamber to maintain a similar amount of total

COD in the samples. The rate of COD removal with the 1-CP system with acetate was

1.5 ± 0.3 g-COD/Lan-d, nearly twice that of the MFC (Figure 6.6a). In the 2-CP MRC,

COD removal further increased to 1.9 ± 0.1 g-COD/Lan-d COD. Overall COD removal

(77 – 91%) was very high in both the 1-CP and 2-CP MRCs fed acetate. Energy recovery

increased from 0.4 ± 0.01 Wh/g-COD for the MFC, to 0.7 ± 0.01 Wh/g-COD for the 1-

CP and 0.9 ± 0.1 for the 2-CP MRCs (Figure 6.6b).

Volumetric COD removal rates and energy recoveries were also greatly increased

with domestic wastewater. The COD removal rate with the 1-CP MRC (1.2 ± 0.1 g-

COD/Lrxtr-d) was 20 times larger than that of the single chamber MFC (0.05 ± 0.007 g-

COD/Lrxtr-d, Figure 6.6a). Using the 2-CP MRC, the COD removal rate of 2.3 ± 0.2 g-

COD/Lrxtr-d was nearly twice that of the 1-CP, and it was actually larger than the acetate

removal rate in this system. This COD removal rate for wastewater is within the range

reported for activated sludge systems [27]. The overall COD removal with the MRCs (1-

CP, 40 ± 6%, 2-CP, 49 ± 1%) was lower than that obtained with the single chamber MFC

(68 ± 2%). Energy recovery with wastewater followed the same trend as acetate, but

values were lower due COD removal from processes other than oxidation by anode

bacteria (Figure 6.6b). Coulombic efficiencies, which measure the percentage of COD

removed by current generation, were significantly lower with wastewater (35 – 50%)

than with acetate (72 – 81%).

Page 151: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

137

6.5 Outlook

By using AmB as catholyte and minimizing the number of RED cell pairs, it was

shown that electrode kinetics could be dramatically enhanced with only a single

membrane. Although adding additional RED cell pairs could further increase power

densities, calculated on the basis of projected cathode area, power does not proportionally

increase on the basis of total membrane area. The inclusion of additional membranes

lowered the membrane area normalized power density from 3.1 ± 0.1 W/m2 in the 1-CP

configuration to 1.4 ± 0.1 W/m2. These values are significantly greater than membrane

area normalized power densities of 5-CP MRCs (0.3 — 0.5 W/m2) and quite comparable

to those reported for optimized abiotic RED system.[7, 14, 26, 28]

In order to increase the MRC performance on the basis of total membrane area, the

resistance per RED cell pair must be decreased. The electrode chamber and RED flow

channel thickness used in this study promoted high solution and transport boundary layer

resistance, leading to 17 Ω per CP with acetate and 48 Ω per CP with wastewater.

Previous studies have shown that optimization of RED flow channel thickness and

hydrodynamics can reduce RED stack resistance to less than 1 Ω per CP.[26, 29] If MRC

stack resistance can attain similar optimization, 2-CP power densities would approach 10

W/m2-cathode (3 W/m2-membrane, Figure 6.8a, b). With stack optimization, 2-CP MRCs

could exceed the maximum membrane normalized power densities reported for abiotic

RED systems (2.2 W/m2) [26] showing that increasing MFC electrode power with

minimal RED stacks makes efficient use of membrane area and could enable high power

densities even using domestic wastewater.

Page 152: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

138

6.6 Tables

Table 6.1: Anode substrate characteristics

Anolyte tCOD (g/L) sCOD (g/L) κ (mS/cm) pH

Ac 1.8 ± 0.05 1.8 ± 0.05 5.5 ± 0.1 8.1 ± 0.1

WW 0.38 ± 0.04 0.20 ± 0.02 1.5 ± 0.3 7.6 ± 0.2

WW-HC 0.38 ± 0.04 0.20 ± 0.02 5.1 ± 0.1 8.1 ± 0.1

Page 153: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

139

6.7 Figures

Figure 6.1: Schematics of tested reactor configurations.

Single Chamber MFC

Ref.

2-CP MRC(Additional HC/LC)

AEM

CEMRef. Ref.

1-CP MRC(HC Cathode)

Ref. Ref.

AEM

-0.2

-0.1

0.0

0.1

0.2

0.3

0 2 4 6 8 10 12

RED

Sta

ck P

oten

tial (

V)

Current Density (A/m2)

2-CP Ac

1-CP Ac

-0.2

-0.1

0.0

0.1

0.2

0.3

0 1 2 3 4 5 6 7

RE

D S

tack

Pot

entia

l (V

)

Current Density (A/m2 )

1-CP WW

2-CP WW

Figure 6.2: (a,b) Power densities normalized to cathode area, (c,d) anode and cathode potentials (vs. SHE), and (e,f) reverse electrodialysis (RED) stack potentials of MFCs, 1-CP and 2-CP MRCs fed (filled markers) bicarbonate buffered acetate and (open markers) domestic wastewater.

-0.4

-0.2

0.0

0.2

0.4

0.6

0 2 4 6 8 10 12

Ele

ctro

de P

oten

tial (

V)

Current Density (A/m2)

2-CP Ac C 2-CP Ac A1-CP Ac C 1-CP Ac AMFC Ac C MFC Ac A

-0.4

-0.2

0.0

0.2

0.4

0.6

0 1 2 3 4 5 6 7

Ele

ctro

de P

oten

tial (

V)

Current Density (A/m2)

2-CP WW C 2-CP WW A1-CP WW C 1-CP WW AMFC WW C MFC WW A

c)

d)

e)

f)

0.0

1.0

2.0

3.0

4.0

5.0

0 2 4 6 8 10 12

Pow

er D

ensi

ty (W

/m2 -

cat)

Current Density (A/m2)

MFC Ac1-CP Ac2-CP Ac

0.0

0.5

1.0

1.5

2.0

2.5

0 1 2 3 4 5 6 7

Pow

er D

ensi

ty (W

/m2 -

cat)

Current Density (A/m2)

MFC WW1-CP WW2-CP WW

b)

a)

Page 154: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

140

0

20

40

60

80

100

1-CP WW 1-CP HKWW

1-CP Ac

DC

Inte

rnal

Res

ista

nce

(Ω) Cathode

REDAnode

0

20

40

60

80

100

1-CP WW 1-CP HKWW

1-CP Ac

AC

Inte

rnal

Res

ista

nce

(Ω)

Cat_Rct Cat_RsolRED_Rdbl RED_RdlRED_Rs/m An_Rct

Figure 6.3: Internal resistance of 1-CP reactors components fed raw domestic wastewater (WW), buffered wastewater (WW-HC), and buffered acetate (Ac), determined from (a) direct current (DC) polarization and (b) alternating current galvanostatic electrochemical impedance spectroscopy.

a)

0

20

40

60

80

100

MFC Ac 1-CP Ac 2-CP Ac

AC

Inte

rnal

Res

ista

nce

(Ω) Cat_Rct Cat_Rsol

RED_Rdbl RED_RdlRED_Rs/m An_RsolAn_Rdiff An_Rct

0

20

40

60

80

100

MFC Ac 1-CP Ac 2-CP Ac

DC

Inte

rnal

Res

ista

nce

(Ω) Cathode

REDAnode

b)

b) a)

Figure 6.4: Internal resistance of MFC, 1-CP, and 2-CP MRC reactors components fed buffered acetate (Ac) determined from (a) direct current (DC) polarization and (b) alternating current (AC) galvanostatic electrochemical impedance spectroscopy.

Page 155: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

141

Figure 6.5: a) Volumetric COD removal rates and (b) energy recovery from wastewater and acetate in batch recycle experiments.

b)

b)

a)

Page 156: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

142

0

1

2

3

4

5

6

0 25 50 75 100(W

/m2 -

mem

bran

e)

0

2

4

6

8

10

12

0 25 50 75 100

Pow

er D

ensi

ty (W

/m2 -

cat)

Reduction in RED Resistance (%)

2-CP Ac2-CP WW1-CP Ac1-CP WW

Figure 6.6: Estimation of 1-CP and 2-CP MRC power density with buffered acetate (Ac) and wastewater (WW) normalized to (a) total membrane area and (b) cathode area versus reduction in RED stack resistance.

b)

Page 157: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

143

6.8 Literature Cited

1. WIN, Clean & Safe Water for the 21st Century: A Renewed National Commitment to

Water and Wastewater Infrastructure, 2000.

2. IPCC, Climate Change 2007: Synthesis Report. , in Contribution of Working Groups I, II

and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate

Change2007, IPCC: Geneva.

3. McCarty, P.L., J. Bae, and J. Kim, Domestic Wastewater Treatment as a Net Energy

Producer–Can This be Achieved? Environmental Science & Technology, 2011. 45(17):

p. 7100-7106.

4. McGinnis, R.L., J.R. McCutcheon, and M. Elimelech, A novel ammonia-carbon dioxide

osmotic heat engine for power generation. Journal of Membrane Science, 2007. 305: p.

13-19.

5. EIA, U.S., Annual Energy Review 2010, DOE, Editor 2010.

6. Kamat, P.V., Meeting the clean energy demand: Nanostructure architectures for solar

energy conversion. The Journal of Physical Chemistry C, 2007. 111(7): p. 2834-2860.

7. Kim, Y. and B.E. Logan, Microbial Reverse Electrodialysis Cells for Synergistically

Enhanced Power Production. Environmental Science & Technology, 2011. 45(13): p.

5834-5839.

8. Logan, B.E., Exoelectrogenic bacteria that power microbial fuel cells. Nature Reviews

Microbiology, 2009. 7(5): p. 375-381.

9. Liu, H., R. Ramnarayanan, and B.E. Logan, Production of electricity during wastewater

treatment using a single chamber microbial fuel cell. Environmental Science &

Technology, 2004. 38(7): p. 2281-2285.

10. Wick, G.L., Power from salinity gradients. Energy, 1978. 3(1): p. 95-100.

Page 158: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

144

11. Post, J.W., et al., Salinity-gradient power: Evaluation of pressure-retarded osmosis and

reverse electrodialysis. Journal of Membrane Science, 2007. 288(1-2): p. 218-230.

12. Ramon, G.Z., B.J. Feinberg, and E.M.V. Hoek, Membrane-based production of salinity-

gradient power. Energy & Environmental Science, 2011.

13. Weinstein, J.N. and F.B. Leitz, Electric power from differences in salinity: the dialytic

battery. Science, 1976. 191(4227): p. 557.

14. Cusick, R.D., Y. Kim, and B.E. Logan, Energy Capture from Thermolytic Solutions in

Microbial Reverse-Electrodialysis Cells. Science, 2012. 335(6075): p. 1474-1477.

15. Kim, J.R., et al., Power generation using different cation, anion and ultrafiltration

membranes in microbial fuel cells. Environmental Science & Technology, 2007. 41(3): p.

1004-1009.

16. Popat, S.C., et al., Importance of OH− Transport from Cathodes in Microbial Fuel Cells.

ChemSusChem, 2012. 5(6): p. 1071-1079.

17. Ahn, Y. and B.E. Logan, Saline catholytes as alternatives to phosphate buffers in

microbial fuel cells. Bioresource Technology, 2013. 132(0): p. 436-439.

18. Wang, X., et al., Use of carbon mesh anodes and the effect of different pretreatment

methods on power production in microbial fuel cells. Environmental Science &

Technology, 2009. 43(17): p. 6870-6874.

19. Cheng, S., H. Liu, and B.E. Logan, Increased performance of single-chamber microbial

fuel cells using an improved cathode structure. Electrochemistry Communications, 2006.

8(3): p. 489-494.

20. Ambler, J.R. and B.E. Logan, Evaluation of stainless steel cathodes and a bicarbonate

buffer for hydrogen production in microbial electrolysis cells using a new method for

measuring gas production. International Journal of Hydrogen Energy, 2011. 36(1): p.

160-166.

Page 159: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

145

21. Zhang, F., et al., Mesh optimization for microbial fuel cell cathodes constructed around

stainless steel mesh current collectors. Journal of Power Sources, 2011. 196(3): p. 1097-

1102.

22. Hutchinson, A.J., J.C. Tokash, and B.E. Logan, Analysis of carbon fiber brush loading in

anodes on startup and performance of microbial fuel cells. Journal of Power Sources,

2011. 196(22): p. 9213-9219.

23. Długołęcki, P., et al., On the resistances of membrane, diffusion boundary layer and

double layer in ion exchange membrane transport. Journal of Membrane Science, 2010.

349(1): p. 369-379.

24. Logan, B.E., Microbial fuel cells2008, Hoboken, NJ: John Wiley & Sons, Inc. 300.

25. Liu, H., S. Cheng, and B.E. Logan, Power generation in fed-batch microbial fuel cells as

a function of ionic strength, temperature, and reactor configuration. Environmental

Science & Technology, 2005. 39(14): p. 5488-5493.

26. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Doubled power density from salinity

gradients at reduced intermembrane distance. Environmental Science & Technology,

2011. 45(16): p. 7089-7095.

27. Rozendal, R.A., et al., Towards practical implementation of bioelectrochemical

wastewater treatment. Trends in Biotechnology, 2008. 26(8): p. 450-459.

28. Vermaas, D.A., M. Saakes, and K. Nijmeijer, Power generation using profiled

membranes in reverse electrodialysis. Journal of Membrane Science, 2011. 385–386(0):

p. 234-242.

29. Długołȩcki, P., et al., Practical Potential of Reverse Electrodialysis As Process for

Sustainable Energy Generation. Environmental Science & Technology, 2009. 43(17): p.

6888-6894.

Page 160: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

146

Appendix A

Energy Efficient Phosphate Recovery as Struvite Within a Single Chamber Microbial Electrolysis Cell

Figure A.1: Concurrent LSV (filled) and pH (open) measurements at the surface of SSM cathode in 3mM PBS and 75 mM HCO3

- electrolyte.

-30

-25

-20

-15

-10

-5

0

6.5

7.0

7.5

8.0

8.5

9.0

-0.1 -0.4 -0.7 -1.0

I (m

A)pH

E (V) vs. SHE

pH, 3mM PBS

pH, 75 mM HCO3

I (mA), 3 mM PBS

I (mA), 75 mM HCO3

Page 161: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

147

Figure A.2: Average batch time and crystal accumulation of MESC reactors with both SSF and SSM cathodes at applied voltages of 0.75, 0.90 and 1.05 V.

10

20

30

40

0.0

2.0

4.0

6.0

8.0

10.0

0.7 0.8 0.9 1.0 1.1

Bat

ch D

urat

ion

(hr)

Cry

stal

l Gro

wth

(mg/

batc

h)

Applied Voltage (V)

SSMSSFBatch

2

A.3: Anode (an) and cathode (cat) potentials measured in fed-batch operation at a) 0.75 b) 0.90 and c) 1.05 V.

-1.2

-1.0

-0.8

-0.6

-0.4

-0.2

0.00.7 0.8 0.9 1.0 1.1

Ele

ctro

de P

ot.

(V) v

s. S

HE

Applied Voltage (V)

SSM-C-cat

SSM-S-cat

SSF-C-cat

SSF-S-cat

SSM-C-an

SSM-S-an

SSF-C-an

SSF-S-an

Page 162: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

148

Figure A.4: EIS determination of internal resistance for SSM and SSF in MEC (C) and MESC (S) reactors at cathode potentials observed in fed-batch operation at a) 0.75 b) 0.90 and c) 1.05 V.

a)

b)

c)

Page 163: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

Appendix B

Energy Production from Thermolytic Solutions in Microbial Reverse Electrodialysis Cells

1

Figure B.1: Relationship between ammonium bicarbonate concentration and solution. Figure

y = -16.505x2 + 91.209xR² = 1

0

20

40

60

80

100

120

140

0.0 0.5 1.0 1.5 2.0

Con

duct

ivity

(mS

/cm

)

Concentration (M)

Page 164: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

150

Figure B.2: Relationship between total ammonium bicarbonate salt concentration and species concentration at pH =7 and T = 25 °C as estimated with OLI stream analyzer software.

0.0

0.5

1.0

1.5

2.0

0.0 0.5 1.0 1.5 2.0

Spe

cies

Con

cent

ratio

n (M

)

Total Concentration (M)

NH4+HCO3-NH4CO3-CO32-NH3

Figure B.3: Relationship between species concentration and species activity at pH =7 and T = 25 °C as estimated with OLI stream analyzer software.

0.0

0.3

0.5

0.8

1.0

0.0 0.5 1.0 1.5 2.0

Spe

cies

act

vity

Species Concentration (M)

NH4+HCO3-NH4CO3-CO32-

Page 165: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

151

0.0

1.0

2.0

3.0

4.0

5.0

6.0

0.0 0.2 0.4 0.6 0.8 1.0P

ower

Den

sity

(W/m

2 )

Current Density (mA/cm2)

1.6 mL/min0.85 mL/min

Figure B.4: Power density curves of the MRC (HC = 0.95 M, SR = 100) at different salt solution flow rates.

1

Figure B.5: Batch recycle component (MFC, RED and total MRC) power profile of MRC fed sodium acetate and operating with an external resistance of 300 Ω.

0.0

1.0

2.0

3.0

4.0

0.0 0.2 0.4 0.6 0.8

Pow

er (W

/m2 )

Time (days)

MRCMFCRED

Page 166: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

152

Appendix C

Minimal RED Cell Pair Microbial Reverse Electrodialysis Cells

Figure C.1: Internal resistance determined from the slope of linear polarization curves for the MFC, 1-CP and 2-CP MRC fed acetate (Ac) and domestic wastewater (WW).

050

100150200250300350

MFCWW

MFCAc

1-CPWW

1-CPAc

2-CPWW

2-CPAc

Inte

rnal

Res

ista

nce

(Ω)

CathodeREDAnode

Page 167: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

153

a) b)

Figure C.2: a) Power density, (b) electrode potentials, and (c) stack potentials of 1-CP reactor fed acetate (Ac), high conductivity wastewater (WW-HC) and raw wastewater (WW).

) ) )

Figure C.3: Nyquist plots of GEIS data collected from 1-CP MRCs fed acetate, wastewater, and high conductivity wastewater. Impedance spectra were simultaneously collected for (a) whole cell and (b) anode (c) RED stack and (d) cathode.

b) d)

a) c)

b) d)

Page 168: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

154

0

10

20

30

40

0 10 20 30 40

-Img

(Z)/

Ohm

Real(Z)/Ohm

Cathode 2-CP Ac

Cathode 1-CP Ac

Cathode MFC

0

10

20

30

40

0 10 20 30 40

-Img

(Z)/

Ohm

Real(Z)/Ohm

RED 2-CP AcRED 1-CP Ac

05

10152025303540

0 10 20 30 40

-Img

(Z)/

Ohm

Real(Z)/Ohm

Anode MFC

Anode 1-CP Ac

Anode 2-CP Ac

Figure C.4: Nyquist plots of GEIS data collected from electrode and RED of MFC and MRCs fed acetate. Impedance spectra were simultaneously collected for (a) anode (c) RED stack and (d) cathode.

a) b) c))

0

5

10

15

20

0 5 10 15 20

-Img

(Z)/

Ohm

Real(Z)/Ohm

Anode 1-CP Ac

Anode model fit

0

5

10

15

20

0 5 10 15 20

-Img

(Z)/

Ohm

Real(Z)/Ohm

RED 1-CP Ac

Stack model fit

0

5

10

15

20

0 5 10 15 20

-Img

(Z)/

Ohm

Real(Z)/Ohm

1-CP Ac Cat

Cathode model fit

Figure C.5: Nyquist plots with equivalent circuit models and fits of GEIS data collected from the (a) anode (c) RED stack and (d) cathode of 1-CP MRC fed acetate.

a) b) c))

v

Rdl

>>Qa

Rsol+m v

Rdbl

>>Qa

v

Rct

>>Qa

Rsolv

Rct

>>Qa

Rsol

Page 169: NUTRIENT AND HEAT RECOVERY FROM WASTE STREAMS USING

VITA

Roland D. Cusick

EDUCATION

Doctor of Philosophy in Environmental Engineering, Department of Civil and Environmental Engineering, The Pennsylvania State University, University Park, PA, 08/2010 – 12/2013 (expected). Advisor: Bruce E. Logan

Master of Science in Environmental Engineering, Department of Civil and Environmental Engineering, The Pennsylvania State University, University Park, PA, 08/2008 – 08/2010. Advisor: Bruce E. Logan

Bachelor of Science in Environmental Engineering, Department of Chemical and Environmental Engineering, The University of California, Riverside, CA, 09/2001–12/2005. AWARDS

• W. Wesley Eckenfelder Graduate Research Award, American Academy of Environmental Engineers and Scientists, 2013

• Alumni Association Dissertation Award, Penn State University, 2013

• DOW Sustainability Innovation Student Challenge Award, Grand Prize, Penn State University, 2012

• George W. Johnstone Graduate Fellowship, Department of Civil and Environmental Engineering, Penn State University, 2012

• Cecil Pepperman Memorial Fellowship, Department of Civil and Environmental Engineering, Penn State University, 2010

INVENTION DISCLOSURES AND PATENT APPLICATIONS

• M. Yates, Cusick, R. D., and B. E. Logan (2013). Formation of a Nanoporous Structure using an Active Exoelectrogenic Biofilm as a Template.

• Y. Kim, Cusick, R. D., and B. E. Logan (2012). Reverse Electrodialysis Supported Microbial Fuel Cells and Microbial Electrolysis Cells.